1:
2:
3:
4: %\documentclass{svjour2} % onecolumn
5: %\documentclass[twocolumn,fleqn]{svjour2} % twocolumn
6: \documentclass[onecollarge]{svjour2} % onecolumn "king-size"
7: %
8: \smartqed % flush right qed marks, e.g. at end of proof
9:
10: \usepackage{graphicx,epsfig}
11: \usepackage{natbib}
12:
13: %
14: % \usepackage{mathptmx} % use Times fonts if available on your TeX system
15: %
16: % insert here the call for the packages your document requires
17: %\usepackage{latexsym}
18: % etc.
19: %
20:
21:
22: %%% Example macros (some are not used in this sample file) %%%
23:
24: \newcommand{\beq}{\begin{equation}}
25: \newcommand{\eeq}{\end{equation}}
26: \newcommand{\beqa}{\begin{eqnarray}}
27: \newcommand{\eeqa}{\end{eqnarray}}
28: \newcommand{\bea}{\begin{array}}
29: \newcommand{\ea}{\end{array}}
30:
31: \newcommand{\dd}{{\rm d}}
32: \newcommand{\pl}{\partial}
33: \newcommand{\inta}{\int_{-i\infty}^{+i\infty}}
34: \newcommand{\lag}{\langle}
35: \newcommand{\rag}{\rangle}
36: \newcommand{\law}{\stackrel{\rm law}{=}}
37: \newcommand{\ii}{{\rm i}}
38: \newcommand{\pv}{{\rm p.v.}}
39:
40: \newcommand{\cP}{{\cal P}}
41: \newcommand{\cK}{{\cal K}}
42: \newcommand{\tG}{\tilde{G}}
43: \newcommand{\Hi}{H_{\infty}}
44: \newcommand{\tH}{\tilde{H}}
45: \newcommand{\tD}{\tilde{\Delta}}
46: \newcommand{\tR}{\tilde{R}}
47: \newcommand{\gam}{\gamma}
48: \newcommand{\cI}{{\cal I}}
49: \newcommand{\cC}{{\cal C}}
50: \newcommand{\hu}{{\hat u}}
51: \newcommand{\hf}{{\hat f}}
52: \newcommand{\hF}{{\hat F}}
53: \newcommand{\hp}{{\hat p}}
54: \newcommand{\hP}{{\hat P}}
55: \newcommand{\bP}{{\overline P}}
56: \newcommand{\PPsi}{\chi}
57: \newcommand{\cQ}{{\cal Q}}
58: \newcommand{\cX}{{\cal X}}
59: \newcommand{\bp}{{\overline p}}
60: \newcommand{\Pk}{{\cal P}}
61: \newcommand{\tN}{\tilde{N}}
62: \newcommand{\cG}{{\cal G}}
63: \newcommand{\cF}{{\cal F}}
64:
65: \newcommand{\Ai}{\mbox{Ai}}
66: \newcommand{\Bi}{\mbox{Bi}}
67: \newcommand{\hphi}{\hat{\phi}}
68: \newcommand{\Arg}{\mbox{Arg}}
69: \newcommand{\erfc}{\mbox{erfc}}
70: \newcommand{\sinc}{\mbox{sinc}}
71:
72:
73: \journalname{Journal of Statistical Physics}
74:
75: \begin{document}
76:
77: \title{Statistical properties of the Burgers equation with Brownian initial velocity}
78: %\subtitle{} %Do you have a subtitle?\\ If so, write it here
79: \titlerunning{Burgers equation with Brownian initial velocity} % if too long for running head
80:
81: \author{Patrick Valageas}
82:
83: \institute{Institut de Physique Th{\'e}orique, CEA Saclay,
84: 91191 Gif-sur-Yvette, France\\
85: \email{valag@spht.saclay.cea.fr}}
86:
87: \date{Received: date / Accepted: date}
88:
89:
90: \maketitle
91:
92: \begin{abstract}
93: We study the one-dimensional Burgers equation in the inviscid limit for Brownian
94: initial velocity (i.e. the initial velocity is a two-sided Brownian motion that
95: starts from the origin $x=0$). We obtain the one-point distribution of the velocity
96: field in closed analytical form.
97: In the limit where we are far from the origin, we also
98: obtain the two-point and higher-order distributions. We show how they factorize
99: and recover the statistical invariance through translations for the distributions
100: of velocity increments and Lagrangian increments. We also derive the velocity
101: structure functions and we recover the bifractality of the inverse Lagrangian map.
102: Then, for the case where the initial density
103: is uniform, we obtain the distribution of the density field and its $n$-point
104: correlations. In the same limit, we derive the $n-$point distributions of the
105: Lagrangian displacement field and the properties of shocks.
106: We note that both the stable-clustering ansatz and the Press-Schechter
107: mass function, that are widely used in the cosmological context, happen to
108: be exact for this one-dimensional version of the adhesion model.
109: \keywords{Inviscid Burgers equation \and Turbulence \and Cosmology: large-scale structure of the universe}
110: \end{abstract}
111:
112:
113:
114: \section{Introduction}
115: \label{sec:intro}
116:
117:
118: The Burgers equation \cite{Burgersbook} is a very popular nonlinear evolution
119: equation that appears in many physical problems, see \cite{Bec2007}
120: for a recent review. It was first introduced
121: as a simplified model of fluid turbulence, as it shares the same hydrodynamical
122: (advective) nonlinearity and several conservation laws with the Navier-Stokes
123: equation. Even though it was shown later on by \cite{Hopf1950} and \cite{Cole1951}
124: that it can be explicitly integrated and lacks the chaotic character associated
125: with actual turbulence, it still retains much interest for hydrodynamical studies.
126: In particular, it can serve as a useful benchmark to test various approximation
127: schemes devised for turbulence studies, since the nonlinearity is the same
128: for both dynamics \cite{Fournier1983}. On the other hand, it has appeared in other
129: physical situations, such as the propagation of nonlinear acoustic waves in
130: non-dispersive media \cite{Gurbatov1991}, the study of disordered systems and
131: pinned manifolds \cite{LeDoussal2008}, or the formation of large-scale
132: structures in cosmology \cite{Gurbatove1989,Vergassola1994}. There,
133: in the limit of vanishing viscosity, it is known as the ``adhesion model'' and it
134: provides a good description of the large-scale filamentary structure of the
135: cosmic web \cite{Melott1994}. In this context, one is interested in the
136: statistical properties of the dynamics, starting with random Gaussian initial
137: conditions \cite{Kida1979,Gurbatov1997} (i.e. ``decaying Burgers turbulence'' in
138: the hydrodynamical context).
139: Moreover, in addition to the velocity field, one is also interested in the
140: properties of the density field generated by this dynamics, starting with
141: an initial uniform density.
142:
143: This problem has led to many studies, focusing
144: on power-law initial spectra (fractional Brownian motion), especially for the
145: two peculiar cases of white-noise initial velocity
146: \cite{Burgersbook,Kida1979,She1992,Frachebourg2000} or Brownian motion initial
147: velocity \cite{She1992,Sinai1992,Bertoin1998}. The initial velocity fluctuations
148: are dominated by short wavelengths in the former case and by large wavelengths
149: in the latter case. In the present Universe, where the power spectrum is not a power
150: law and converges at both ends, the velocity fluctuations are governed by scales
151: that are somewhat larger than those where structures have already formed (thus
152: the variance of the velocity field is still set by the linear theory) and this
153: scale ratio was larger in the past (as the size of nonlinear structures was
154: smaller). In this sense the case of Brownian initial conditions is closer to the
155: cosmological scenario. From the viewpoint of hydrodynamics, this is also
156: an interesting configuration since in many hydrodynamical systems the power is
157: generated by the larger scales. For instance, the Kolmogorov spectrum of turbulence,
158: $E(k)\propto k^{-5/3}$, displays such an infrared divergence. Thus, the case of
159: Brownian initial velocity was recently used in \cite{Frisch2005} to address
160: the issue of local homogeneity
161:
162: In this article, we revisit the one-dimensional Burgers dynamics with two-sided
163: Brownian initial velocity. In the spirit of the approach of \cite{Frachebourg2000},
164: using analysis methods (Laplace transforms)
165: we obtain closed analytical results for $n$-point distributions
166: (mostly in the limit where we are far from the origin of the initial Brownian
167: motion if $n\geq 2$). We check that our results agree with already known
168: properties. In particular, we recover the property, derived by \cite{Bertoin1998}
169: through probabilistic tools for the one-sided Brownian initial velocity, that
170: increments of the inverse Lagrangian map are independent and homogeneous.
171: In our case this only holds for particles that are on the same side of the
172: origin. We pay attention to issues that arise in the hydrodynamical
173: context (e.g., velocity structure functions, Lagrangian displacement field)
174: as well as the cosmological context (e.g., statistics of the density field,
175: mass function of the collapsed structures associated with shocks).
176: In particular, we compare our exact results with phenomenological models
177: that are often used to describe the cosmological dynamics.
178:
179: We first describe in section~\ref{sec:Initial} the initial Brownian conditions
180: and the standard geometrical interpretation in terms of parabolas of the
181: Hopf-Cole solution of the dynamics \cite{Burgersbook}.
182: Adapting to our case the method presented in
183: \cite{Frachebourg2000}, this will allow us to express all statistical
184: properties in terms of the transition kernel associated with Brownian particles
185: moving above parabolic absorbing barriers. We present this propagator in
186: sect.~\ref{sec:transition}, decomposed over a continuous set of eigenfunctions
187: built from the Airy function (whereas the white-noise case leads to a discrete
188: spectrum, that is also built from Airy functions). Then, we derive closed
189: analytical expressions for the one-point velocity distribution
190: $p_x(v)$ in sect.~\ref{sec:velocity1}, as well as the distribution, $p_x(q)$, of the
191: initial Lagrangian position $q$ of the particle that is located at the position
192: $x$ at time $t$. Next, we study the two-point and higher-order distributions
193: in sect.~\ref{sec:velocity2}, and we obtain simple analytical results in the
194: limit where all particles are far from the origin. This allows us to derive the
195: distribution of the density field in sect.~\ref{sec:Density}, for the case of
196: a uniform initial density. Next, we consider the statistics of the Lagrangian
197: displacement field in sect.~\ref{sec:Displacement}. In the same limit where the
198: particles are far from the origin, we obtain the $n-$point distributions,
199: $p_{q_i}(x_i)$, of the positions $x_i$ at time $t$ of the particles that were
200: initially at positions $q_i$. We also derive the probability $\bp_q^{\rm shock}$
201: that two particles initially separated by a distance $q$ have coalesced into
202: a single shock by time $t$. Finally, we obtain in sect.~\ref{sec:shock} the
203: mass function of shocks and their spatial distribution.
204:
205: The reader who is not interested in the technical details of our derivations
206: may directly go to section~\ref{sec:velocity2} to survey most of our practical
207: results.
208:
209:
210:
211:
212: \section{Initial conditions and geometrical solution}
213: \label{sec:Initial}
214:
215:
216: We consider the one-dimensional Burgers equation for the velocity field $v(x,t)$
217: in the limit of zero viscosity,
218: \beq
219: \frac{\pl v}{\pl t} + v \frac{\pl v}{\pl x} = \nu \frac{\pl^2 v}{\pl x^2}
220: \hspace{1cm} \mbox{with} \hspace{1cm} \nu \rightarrow 0^+ .
221: \label{Burg}
222: \eeq
223: As is well-known \cite{Hopf1950,Cole1951}, introducing the velocity potential
224: $\psi(x,t)$ and making the change of variable $\psi(x,t)=-2\nu\ln\theta(x,t)$
225: transforms the nonlinear Burgers equation into the linear heat equation.
226: This gives the explicit solution
227: \beq
228: v(x,t) = \frac{\pl\psi}{\pl x} \hspace{0.5cm} \mbox{with} \hspace{0.5cm}
229: \psi(x,t)= -2\nu \ln \int_{-\infty}^{\infty} \frac{\dd q}{\sqrt{4\pi\nu t}} \;
230: \exp\left[-\frac{(x-q)^2}{4\nu t}-\frac{\psi_0(q)}{2\nu}\right] ,
231: \label{psinu}
232: \eeq
233: where we introduced the initial condition $\psi_0(q)=\psi(q,t=0)$.
234: Then, in the limit $\nu \rightarrow 0^+$ the steepest-descent method
235: gives
236: \beq
237: \psi(x,t) = \min_q \left[ \psi_0(q) + \frac{(x-q)^2}{2t} \right]
238: \hspace{0.5cm} \mbox{and} \hspace{0.5cm} v(x,t) = \frac{x-q(x,t)}{t} ,
239: \label{psinu0}
240: \eeq
241: where we introduced the Lagrangian coordinate $q(x,t)$ defined by
242: \beq
243: \psi_0(q) + \frac{(x-q)^2}{2t} \hspace{0.5cm} \mbox{is minimum at the point}
244: \hspace{0.5cm} q = q(x,t) .
245: \label{qmin}
246: \eeq
247: The Eulerian locations $x$ where there are two solutions $q_-<q_+$ to the
248: minimization problem (\ref{qmin}) correspond to shocks (and all the matter
249: initially between $q_-$ and $q_+$ is gathered at $x$). The application
250: $q \mapsto x(q,t)$ is usually called the Lagrangian map, and
251: $x \mapsto q(x,t)$ the inverse Lagrangian map (which is discontinuous at
252: shock locations). For the case of Brownian initial velocity that we consider
253: in this paper, it is known that the set of regular Lagrangian points has
254: a Hausdorff dimension of $1/2$ \cite{Sinai1992}, whereas shock locations
255: are dense in Eulerian space \cite{Sinai1992,She1992}.
256:
257: In this article, we take for the initial velocity field $v_0(q)$ a bilateral
258: Brownian motion starting from the origin $v_0(0)=0$, and we also normalize
259: the potential $\psi_0$ by $\psi_0(0)=0$. Thus, introducing a Gaussian
260: white noise $\xi(q)$, we can express the initial conditions by
261: \beq
262: v_0(q)= \int_0^q\dd q' \, \xi(q') , \;\;\;\;
263: \psi_0(q)= \int_0^q\dd q' \int_0^{q'}\dd q'' \, \xi(q'') .
264: \label{xidef}
265: \eeq
266: All initial fields are Gaussian and fully determined by their two-point
267: correlation, which we normalize by
268: \beq
269: \lag\xi(q)\rag = 0, \;\;\;\; \lag\xi(q)\xi(q')\rag = D \, \delta(q-q') ,
270: \label{Ddef}
271: \eeq
272: where $\lag .. \rag$ is the average over all realizations of $\xi$.
273: This gives for instance
274: \beq
275: \lag v_0(q_1) v_0(q_2)\rag = D \, q_1 , \;\;\;
276: \lag \psi_0(q_1) \psi_0(q_2)\rag = \frac{D}{2}
277: \left[q_1^2 q_2 -\frac{q_1^3}{3}\right] ,
278: \;\;\; \mbox{for} \;\;\; 0 \leq q_1 \leq q_2 ,
279: \label{variance0}
280: \eeq
281: and for the initial velocity distribution at location $q$,
282: \beq
283: t=0 : \;\;\;\; p_{q}(v) = \frac{1}{\sqrt{2\pi}\sigma_{v_0}} \,
284: e^{-v^2/(2\sigma_{v_0}^2)} \;\;\;\; \mbox{with} \;\;\;\;
285: \sigma_{v_0}^2(q) = D q .
286: \label{Gaussian_t0}
287: \eeq
288: Note that the initial fields over the two sides $q<0$ and $q>0$ are independent.
289: The initial velocity $v_0(q)$ is not homogeneous, since the origin $q=0$
290: clearly plays a special role, but it has homogeneous increments, as seen
291: from the equality,
292: \beq
293: \mbox{for any} \; q_1,q_2 : \;\;\;
294: v_0(q_2)-v_0(q_1) = \int_{q_1}^{q_2}\dd q \, \xi(q) ,
295: \;\;\; \lag [v_0(q_2)-v_0(q_1)]^2 \rag = D |q_2-q_1| .
296: \label{Dv0}
297: \eeq
298: Then, the energy spectrum $E_0(k)$ of the initial velocity field is
299: \beq
300: E_0(k)=\frac{D}{2\pi} k^{-2} , \;\;\; \mbox{with} \;\;\;
301: \lag [v_0(q_2)-v_0(q_1)]^2 \rag = 2 \int_{-\infty}^{\infty} \dd k \,
302: (1-e^{\ii k (q_2-q_1)}) \, E_0(k) .
303: \label{E0}
304: \eeq
305:
306: Thanks to the scale invariance of the
307: Brownian motion, the scaled initial potential $\psi_0(\lambda q)$ has the same
308: probability distribution as $\lambda^{3/2} \psi_0(q)$, for any $\lambda>0$.
309: Then, using the explicit solution (\ref{psinu0}) we obtain the scaling laws
310: \beq
311: \psi(x,t) \law t^3 \psi(x/t^2,1) , \;\;\; v(x,t) \law t v(x/t^2,1) , \;\;\;
312: q(x,t) \law t^2 q(x/t^2,1) ,
313: \label{scalings}
314: \eeq
315: where $\law$ means that both sides have the same probability distribution.
316: Thus, any equal-time statistics at a given time $t>0$ can be expressed in terms
317: of the same quantity at the time $t=1$ through appropriate rescalings.
318: In this article we only investigate equal-time statistics, so that $t$ can
319: be seen as a mere parameter in the explicit solution (\ref{psinu})
320: from which we derive our results.
321:
322: In the cosmological context, the time $t$ in the Burgers equation (\ref{Burg})
323: actually stands for the linear growing mode $D_+(t)$ of the density fluctuations,
324: the spatial coordinate $x$ is a comoving coordinate (that follows the
325: uniform Hubble expansion) and, up to a time-dependent factor, the velocity $v$ is
326: the peculiar velocity (where the Hubble expansion has been subtracted),
327: see \cite{Gurbatove1989,Vergassola1994}.
328: In these coordinates, the evolution of the density
329: field is still given by the continuity equation (\ref{continuity}) below,
330: where the density $\rho$ is the comoving density. If we take $\nu=0$,
331: that is we remove the right hand side in Eq.(\ref{Burg}), this is the well-known
332: Zeldovich approximation \cite{Zeldovich1970,Valageas2007}, where particles
333: always keep their
334: initial velocity and merely follow straight trajectories. The diffusive term
335: of (\ref{Burg}) is then added as a phenomenological device to prevent particles from
336: escaping to infinity after crossing each other and to mimic the gravitational
337: trapping of particles within the potential wells formed by the overdensities
338: \cite{Gurbatove1989}.
339: Of course, this cannot describe the inner structure of collapsed objects
340: (e.g., galaxies) but it provides a good description of the large-scale structure
341: of the cosmic web \cite{Melott1994}.
342:
343: As is well-known \cite{Burgersbook}, the minimization problem (\ref{qmin})
344: has a nice geometrical solution. Indeed, let us consider the
345: downward\footnote{In the literature one usually defines the velocity potential
346: as $v=-\pl_x\psi$, which leads to upward parabolas. Here we prefer to define
347: $v=\pl_x\psi$ to simplify the interpretation of the process $(q,\psi_0,v_0)$
348: in terms of the dynamics of a Brownian particle.}
349: parabola $\cP_{x,c}(q)$ centered at $x$ and of maximum $c$, i.e. of vertex
350: $(x,c)$, of equation
351: \beq
352: \cP_{x,c}(q) = - \frac{(q-x)^2}{2 t} + c .
353: \label{paraboladef}
354: \eeq
355: Then, starting from below with a large negative value of $c$, such that the
356: parabola is everywhere well below $\psi_0(q)$ (this is possible thanks to the
357: scaling $\psi_0(\lambda q) \law \lambda^{3/2} \psi_0(q)$ which shows
358: that $\psi_0(q)$ only grows as $|q|^{3/2}$
359: at large $|q|$), we increase $c$ until the two curves touch one another.
360: Then, the abscissa of the point of contact is the Lagrangian coordinate
361: $q(x,t)$ and the potential is given by $\psi(x,t)=c$.
362: (We show below in Fig.~\ref{figP1} the case where the Lagrangian coordinate
363: $q'(x,t)$ is somewhere in the range $0\leq q'\leq q$.)
364:
365:
366: \section{Transition kernel with parabolic absorbing barrier}
367: \label{sec:transition}
368:
369:
370: For the Brownian initial conditions (\ref{xidef}), the process
371: $q\mapsto\{\psi_0,v_0\}$ is Markovian, going from $q=0$ towards positive or
372: negative values. Then, following the approach of \cite{Frachebourg2000}
373: (where it was applied to white-noise initial velocity),
374: from the geometrical construction (\ref{paraboladef})
375: we can see that a key quantity is the conditional probability density
376: $K_{x,c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2)$ for the Markov process
377: $\{\psi_0(q),v_0(q)\}$, starting from $\{\psi_1,v_1\}$ at $q_1 \geq 0$, to end at
378: $\{\psi_2,v_2\}$ at $q_2 \geq q_1 \geq 0$, while staying above the parabolic barrier,
379: $\psi_0(q)>\cP_{x,c}(q)$, for $q_1\leq q\leq q_2$. It obeys the
380: advective-diffusion equation
381: \beq
382: \left[ \frac{\pl}{\pl q_2} + v_2 \frac{\pl}{\pl \psi_2} \right]
383: K_{x,c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2) = \frac{D}{2} \frac{\pl^2}{\pl v_2^2}
384: K_{x,c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2)
385: \label{diffK}
386: \eeq
387: over the domain $\psi \geq \cP_{x,c}(q)$, with the initial condition at $q_2=q_1$
388: \beq
389: K_{x,c}(q_1,\psi_1,v_1;q_1,\psi_2,v_2) = \delta(\psi_2-\psi_1) \delta(v_2-v_1) ,
390: \label{initial}
391: \eeq
392: and the boundary condition
393: \beq
394: K_{x,c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2) = 0
395: \;\;\; \mbox{at} \;\;\; \psi_2=\cP_{x,c}(q_2)
396: \;\;\; \mbox{for} \;\;\; v_2 \geq \frac{\dd \cP_{x,c}}{\dd q}(q_2) .
397: \label{Kboundary}
398: \eeq
399: Equation (\ref{diffK}) is also the Klein-Kramers equation for the distribution
400: function $P(x,v;t)$ of Brownian particles, in the limit of zero external force
401: and zero friction coefficient but finite diffusion coefficient, where we identify
402: the position, velocity and time coordinates as $\{x,v;t\}=\{\psi_2,v_2;q_2\}$.
403: The boundary condition (\ref{Kboundary}) simply means that particles cannot
404: come back from the absorbing region (i.e. curves that cross the parabola
405: are ``lost'' and do not contribute to the probability density $K_{x,c}$).
406:
407: In the case of white-noise initial velocity studied in \cite{Frachebourg2000},
408: the velocity potential $\psi$ itself is a Brownian motion so that the
409: relevant propagator only involves one dependent variable, $\psi$,
410: as $K_{x,c}^{\rm w.n.}(q_1,\psi_1;q_2,\psi_2)$. In our case, since $\psi$ is now
411: the integral of the Brownian motion $v$, the propagator $K_{x,c}$ introduced
412: in (\ref{diffK}) involves the two dependent variables $v$ and $\psi$.
413: Thus, we have a diffusion in a two-dimensional $\{\psi,v\}-$space rather than
414: the one-dimensional $\psi-$space as in \cite{Frachebourg2000}.
415: As we shall see below, the propagator $K_{x,c}$ involves an expansion
416: over a continuous spectrum of eigenfunctions that are built from the Airy function,
417: whereas the white-noise case leads to a different expansion over eigenfunctions that
418: are still built from the Airy function but form a discrete spectrum,
419: see \cite{Frachebourg2000}.
420:
421:
422:
423: The conditional probability density $K_{x,c}$ associated with the left-handed
424: Brownian motion $q_2\leq q_1 \leq 0$ can be obtained from the symmetry
425: $q\rightarrow-q$ as:
426: \beq
427: 0 \leq q_1 \leq q_2 : \;\; K_{x,c}(-q_1,\psi_1,v_1;-q_2,\psi_2,v_2)
428: = K_{-x,c}(q_1,\psi_1,-v_1;q_2,\psi_2,-v_2) ,
429: \label{Kmq}
430: \eeq
431: hence we only need consider Eq.(\ref{diffK}) for $0\leq q_1 \leq q_2$.
432: To solve this equation it is convenient to make the change of variables
433: \beq
434: K_{x,c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2) = \cK(q_1,y_1,w_1;q_2,y_2,w_2) ,
435: \label{cKdef}
436: \eeq
437: \beq
438: \mbox{with} \;\;\;\;\;\; y=\psi-\cP_{x,c}(q) =\psi+\frac{(q-x)^2}{2 t}-c ,
439: \;\;\;\;\; w= v - \frac{\dd \cP_{x,c}}{\dd q}(q) = v + \frac{q-x}{t} ,
440: \label{yw}
441: \eeq
442: to obtain a simpler boundary at the fixed vertical half-line $(y=0,w\geq 0)$
443: in the $(y,w)$ half-plane for $\cK$, $y\geq 0$ and $-\infty<w<\infty$, instead of
444: the parabolic boundary for $K$.
445: From Eq.(\ref{diffK}) the kernel $\cK$ satisfies the equation with
446: constant external force
447: \beq
448: \left[ \frac{\pl}{\pl q_2} + w_2 \frac{\pl}{\pl y_2} + \frac{1}{t}
449: \frac{\pl}{\pl w_2} \right] \cK = \frac{D}{2} \frac{\pl^2}{\pl w_2^2} \cK .
450: \label{cKdiff}
451: \eeq
452: Then, making the transformation
453: \beq
454: \cK(q_1,y_1,w_1;q_2,y_2,w_2) = \frac{2}{D} \, G(\tau;r_1,u_1;r_2,u_2) \,
455: \exp\left[\frac{w_2-w_1}{Dt}-\frac{q_2-q_1}{2Dt^2}\right] ,
456: \label{Gdef}
457: \eeq
458: \beq
459: \mbox{with} \;\;\;\;\;\; \tau=q_2-q_1 , \;\;\;\;\; r=\sqrt{\frac{2}{D}} \, y ,
460: \;\;\;\;\; u=\sqrt{\frac{2}{D}} \, w ,
461: \label{xu}
462: \eeq
463: we obtain the simpler advective-diffusion equation for $\tau \geq 0$ and
464: $x\geq 0$,
465: \beq
466: \frac{\pl G}{\pl\tau} + u_2 \frac{\pl G}{\pl r_2} = \frac{\pl^2 G}{\pl u_2^2} ,
467: \label{Gdiff}
468: \eeq
469: with the initial and boundary conditions
470: \beq
471: G(0;r_1,u_1;r_2,u_2) = \delta(r_2-r_1) \delta(u_2-u_1) ,
472: \;\;\; G(\tau;r_1,u_1;0,u_2) = 0 \;\;\; \mbox{for} \;\;\; u_2 \geq 0 .
473: \label{Gboundary}
474: \eeq
475: Thus, $G(\tau;r_1,u_1;r_2,u_2)$ is the conditional probability density
476: of Brownian particles with unit diffusion coefficient and absorbing barrier
477: at $r=0$. This quantity was obtained in \cite{Burkhardt1993} and we briefly
478: recall below his procedure using our notations.
479: We first take the Laplace transform of $G$ as
480: \beq
481: \tG(s;r_1,u_1;r_2,u_2) = \int_0^{\infty} \dd\tau \, e^{-s\tau}
482: G(\tau;r_1,u_1;r_2,u_2) ,
483: \label{tGdef}
484: \eeq
485: hence Eq.(\ref{Gdiff}) gives
486: \beq
487: \left(s+u_2 \frac{\pl}{\pl r_2}-\frac{\pl^2}{\pl u_2^2}\right)
488: \tG(s;r_1,u_1;r_2,u_2) = \delta(r_2-r_1) \delta(u_2-u_1) .
489: \label{tGdiff}
490: \eeq
491: Next, to obtain an ordinary differential equation, it is convenient to expand
492: over the eigenfunctions $e^{-\nu^3 r_2} g_{s,\nu}(u_2)$ associated with
493: Schr\"{o}dinger's equation
494: \beq
495: \left(s-\nu^3 u-\frac{\dd^2}{\dd u^2}\right) g_{s,\nu}(u) = 0 ,
496: \;\;\; \mbox{whence} \;\;\;
497: g_{s,\nu}(u) = \Ai\left[-\nu u +\frac{s}{\nu^2}\right] ,
498: \label{gsnudiff}
499: \eeq
500: using the fact that the standard Airy function $\Ai(x)$ is the only solution
501: of $\Ai''(x)=x \Ai(x)$ that vanishes at both ends $x \rightarrow \pm\infty$
502: \cite{Abramowitz}. We recall in Appendix~\ref{Airy} some useful properties
503: of this entire function. Using the integral representation (\ref{Ai1int}),
504: we obtain the orthogonality property
505: \beq
506: \int_{-\infty}^{\infty} \dd u \, u \, \Ai\left[-\nu u +\frac{s}{\nu^2}\right]
507: \Ai\left[-\nu' u +\frac{s}{\nu'^2}\right] = \frac{1}{3\nu} \, \delta(\nu-\nu') ,
508: \label{orthogonality}
509: \eeq
510: and the closure relation
511: \beq
512: \int_{-\infty}^{\infty} \dd\nu \, 3\nu \, \Ai\left[-\nu u +\frac{s}{\nu^2}\right]
513: \Ai\left[-\nu u' +\frac{s}{\nu^2}\right] = \frac{1}{u} \, \delta(u-u') .
514: \label{closure}
515: \eeq
516: Therefore, we can see from Eqs.(\ref{gsnudiff})-(\ref{closure}) that
517: Eq.(\ref{tGdiff}) has the particular solution
518: \beqa
519: \tG_0(s;r_1,u_1;r_2,u_2) & = & \int_{-\infty}^{\infty} \dd\nu \,
520: e^{-\nu^3(r_2-r_1)} \, 3\nu \, \Ai\left[-\nu u_1 +\frac{s}{\nu^2}\right]
521: \Ai\left[-\nu u_2 +\frac{s}{\nu^2}\right] \nonumber \\
522: && \times \left[ - \theta(-\nu) \theta(r_1-r_2)
523: + \theta(\nu) \theta(r_2-r_1) \right]
524: \label{tG0}
525: \eeqa
526: where $\theta$ is the Heaviside function.
527: We can check that $\tG_0$ vanishes for $|r|\rightarrow\infty$ and for
528: $|u|\rightarrow\infty$. Then, since we have not taken into account the
529: boundary condition at $r_2=0$ of (\ref{Gboundary}) yet, $\tG_0$ is the Laplace
530: transform of the probability density of Brownian particles over the unbounded
531: plane $(r,u)$ (thus $\tG_0$ only depends on the length $|r_2-r_1|$).
532: Note that the solution to this unbounded problem is well known to be
533: the Gaussian \cite{Burkhardt1993}
534: \beq
535: G_0(\tau;r_1,u_1;r_2,u_2) = \frac{\sqrt{3}}{2\pi\tau^2} \,
536: e^{-\frac{3}{\tau^3}(r_2-r_1-u_1\tau)^2
537: +\frac{3}{\tau^2}(r_2-r_1-u_1\tau)(u_2-u_1)-\frac{1}{\tau}(u_2-u_1)^2} ,
538: \label{G0tau}
539: \eeq
540: as can be checked by substitution into Eq.(\ref{Gdiff}). Therefore,
541: Eq.(\ref{G0tau}) is the inverse Laplace transform of Eq.(\ref{tG0}).
542:
543: Next, in order to satisfy the second constraint (\ref{Gboundary}), we must
544: subtract to $\tG_0$ an appropriate solution $\tG_1$ of the homogeneous form of
545: Eq.(\ref{tGdiff}). From Eq.(\ref{gsnudiff}), we can see that $\tG_1$
546: can be written as a combination of eigenfunctions
547: $e^{-\mu^3 r_2} g_{s,\mu}(u_2)$, that must be restricted
548: to $\mu>0$ to ensure that $\tG$ vanishes for $r_2\rightarrow +\infty$.
549: Moreover, for $r_2=0$ only the first part $\theta(-\nu) \theta(r_1-r_2)$
550: contributes to $\tG_0$ in Eq.(\ref{tG0}). Therefore, to compensate for this
551: term at $r_2=0$ for $u_2\geq 0$, we must look for a function $\tG_1$
552: of the form
553: \beq
554: \tG_1(s;r_1,u_1;r_2,u_2) = \int_0^{\infty} \dd\nu \, e^{-\nu^3 r_1} \, 3\nu \,
555: \Ai\left[\nu u_1+\frac{s}{\nu^2}\right] \phi_{s,\nu}(r_2,u_2) ,
556: \label{tG1def}
557: \eeq
558: where the function $\phi_{s,\nu}(r,u)$ can be written as
559: \beq
560: \phi_{s,\nu}(r,u) = \int_0^{\infty} \dd\mu \, W_{s,\nu}(\mu) \, e^{-\mu^3 r}
561: \, \Ai\left[-\mu u+\frac{s}{\mu^2}\right] ,
562: \label{phiW}
563: \eeq
564: with some weight $W_{s,\nu}(\mu)$, and satisfies the constraint
565: \beq
566: \phi_{s,\nu}(r=0,u) = \Ai\left[\nu u+\frac{s}{\nu^2}\right] \;\;\;\;\;
567: \mbox{for} \;\;\;\;\; u\geq 0.
568: \label{phix0}
569: \eeq
570: This is a half-range problem as we must decompose a given function (here
571: $\Ai[\nu u+s/\nu^2]$) over half the domain ($u\geq 0$) using only half of the
572: eigenfunctions $g_{s,\mu}(u)$.
573: Using the results of \cite{Marshall1985}, who studied the Klein-Kramers equation,
574: and taking the limit of zero friction but non-zero diffusion, \cite{Burkhardt1993}
575: obtained:
576: \beq
577: \nu>0 : \;\;\; \phi_{s,\nu}(r,u) = \int_0^{\infty} \frac{\dd\mu}{2\pi} \,
578: \frac{3\nu^{1/2}\mu^{3/2}}{\nu^3+\mu^3} \,
579: e^{-\frac{2}{3} s^{3/2} (\nu^{-3}+\mu^{-3})} \, e^{-\mu^3 r} \,
580: \Ai\left[-\mu u+\frac{s}{\mu^2}\right] .
581: \label{phiint}
582: \eeq
583: Substituting into Eq.(\ref{tG1def}), we obtain for the solution $\tG$
584: of Eq.(\ref{tGdiff}), with the boundary conditions (\ref{Gboundary}),
585: \beq
586: \tG=\tG_0-\tG_1 , \;\;\; \mbox{with}
587: \eeq
588: \beq
589: \tG_1 = \int_0^{\infty} \frac{\dd\nu \dd\mu}{2\pi} \,
590: \frac{9\nu^{3/2}\mu^{3/2}}{\nu^3+\mu^3}
591: \, e^{-\frac{2}{3} s^{3/2} (\nu^{-3}+\mu^{-3})} \, e^{-\nu^3 r_1-\mu^3 r_2}
592: \, \Ai\left[\nu u_1+\frac{s}{\nu^2}\right]
593: \Ai\left[-\mu u_2+\frac{s}{\mu^2}\right] .
594: \label{tG1}
595: \eeq
596: We describe in Appendix~\ref{Half-range} how the
597: solution (\ref{phiint}) can be directly obtained for the half-range expansion
598: problem (\ref{phiW})-(\ref{phix0}), associated with the Brownian dynamics
599: (\ref{Gdiff}), rather than first solving the problem associated with the
600: Klein-Kramers dynamics and next taking the limit of zero friction, see
601: Eq.(\ref{phi0intk1l0}).
602: This also allows us to derive the more general identities (\ref{phi0int}),
603: (\ref{Expintkl}), that we need in the following sections.
604:
605: We can see from the explicit expressions (\ref{tG0}), (\ref{tG1}),
606: that the kernel $G$ also satisfies the backward evolution equations
607: (compare with Eqs.(\ref{Gdiff}), (\ref{tGdiff}))
608: \beqa
609: \left(\frac{\pl}{\pl\tau} - u_1 \frac{\pl}{\pl r_1} - \frac{\pl^2}{\pl u_1^2} \right)
610: G(\tau;r_1,u_1;r_2,u_2) & = & 0 , \label{Gbackward} \\
611: \left(s-u_1 \frac{\pl}{\pl r_1}-\frac{\pl^2}{\pl u_1^2}\right)
612: \tG(s;r_1,u_1;r_2,u_2) & = & \delta(r_2-r_1) \delta(u_2-u_1) ,
613: \label{tGbackward}
614: \eeqa
615: as well as the boundary condition (compare with (\ref{Gboundary}))
616: \beq
617: G(\tau;0,u_1;r_2,u_2) = 0 \;\;\; \mbox{for} \;\;\; u_1 \leq 0 .
618: \label{Gboundary1}
619: \eeq
620: Equation (\ref{Gboundary1}) merely states that the trajectory $r(\tau)$
621: starting on the absorbing barrier at $r_1=0$ must start in the upward
622: direction $u_1>0$ not to be immediately absorbed.
623:
624: For later calculations we also need two kernels $\Delta$ and $H$ that are derived
625: from $G$. Thus, we define the propagator $\Delta$, that will be associated with
626: Brownian particles that come within a small distance $\epsilon$ from the
627: parabolic absorbing barrier, by
628: \beq
629: \Delta(\tau;r_1,u_1;r_2,u_2) = \lim_{\epsilon\rightarrow 0} \frac{1}{\epsilon}
630: [ G(\tau;r_1+\epsilon,u_1;r_2+\epsilon,u_2) - G(\tau;r_1,u_1;r_2,u_2) ] .
631: \label{Deltadef}
632: \eeq
633: From Eqs.(\ref{tG0}) and (\ref{tG1}) we have for its Laplace transform $\tD$
634: \beqa
635: \tD(s;r_1,u_1;r_2,u_2) & = & \int_0^{\infty} \frac{\dd\nu\dd\mu}{2\pi} \,
636: 9\nu^{3/2}\mu^{3/2} \, e^{-\frac{2}{3} s^{3/2} (\nu^{-3}+\mu^{-3})}
637: \, e^{-\nu^3 r_1-\mu^3 r_2} \nonumber \\
638: && \times \Ai\left[\nu u_1+\frac{s}{\nu^2}\right]
639: \Ai\left[-\mu u_2+\frac{s}{\mu^2}\right] .
640: \label{Ds}
641: \eeqa
642: Next, we define the kernel $\Hi(r_1,u_1)$, associated with Brownian particles that
643: stay forever above the parabolic absorbing barrier, by
644: \beq
645: \Hi(r_1,u_1) = \lim_{\tau\rightarrow+\infty} e^{-\tau/\gam^2} H(\tau;r_1,u_1) ,
646: \label{Hidef}
647: \eeq
648: \beq
649: \mbox{with} \;\;\;\;
650: H(\tau;r_1,u_1) = \int_0^{\infty} \dd r_2 \int_{-\infty}^{\infty} \dd u_2
651: \, e^{u_2/\gam} \, G(\tau;r_1,u_1;r_2,u_2) .
652: \label{Hdef}
653: \eeq
654: Using Eqs.(\ref{tG0}) and (\ref{tG1}), and the property (\ref{IntExpAi}),
655: we obtain after integration over $r_2$ and $u_2$ for the Laplace transform $\tH$,
656: \beqa
657: \tH(s;r_1,u_1) & = & \int_{-\infty}^{\infty} \dd\nu \frac{3}{\nu^3}
658: \Ai\left[-\nu u_1+\frac{s}{\nu^2}\right]
659: e^{(\frac{s}{\gam}-\frac{1}{3\gam^3})/\nu^3}
660: \left[ \theta(\nu) - \theta(-\nu) (1-e^{\nu^3 r_1}) \right] \nonumber \\
661: && - \int_0^{\infty} \frac{\dd\nu\dd\mu}{2\pi} \,
662: \frac{9\nu^{3/2}\mu^{-5/2}}{\nu^3+\mu^3} \,
663: e^{-\frac{2}{3} s^{3/2} (\nu^{-3}+\mu^{-3})} \, e^{-\nu^3 r_1}
664: \Ai\left[\nu u_1+\frac{s}{\nu^2}\right]
665: e^{(\frac{s}{\gam}-\frac{1}{3\gam^3})/\mu^3} .
666: \label{Hs}
667: \eeqa
668: The behavior for $\tau\rightarrow\infty$ of $H(\tau;r_1,u_1)$ is determined by
669: the rightmost singularity of $\tH$, which is located at $s=1/\gam^2$.
670: At this point, the first integral in Eq.(\ref{Hs}) diverges for
671: $\nu\rightarrow 0^+$ whereas the second integral diverges for
672: $\mu\rightarrow 0^+$. Therefore, the singularity is governed by the behavior
673: of the integrand for $\nu\rightarrow 0^+$ and $\mu\rightarrow 0^+$, so that we
674: can expand the first Airy function and the ratio $1/(\nu^3+\mu^3)$, which yields
675: \beq
676: s\rightarrow\gam^{-2} : \;\; \tH \sim \frac{1}{s-\gam^{-2}} \left\{ e^{u_1/\gam}
677: - \int_0^{\infty} \frac{\dd\nu}{\sqrt{\pi}} 3 \nu^{-3/2}
678: e^{-\frac{2}{3}\nu^{-3}-\nu^3 r_1/\gam^3}
679: \Ai\left[\nu \frac{u_1}{\gam}+\frac{1}{\nu^2}\right] \right\} .
680: \label{Hsasymp}
681: \eeq
682: This gives for the function $\Hi(r_1,u_1)$:
683: \beq
684: \Hi(r_1,u_1) = e^{u_1/\gam} - \int_0^{\infty} \frac{\dd\nu}{\sqrt{\pi}}
685: 3 \nu^{-3/2} e^{-\frac{2}{3}\nu^{-3}-\nu^3 r_1/\gam^3}
686: \Ai\left[\nu \frac{u_1}{\gam}+\frac{1}{\nu^2}\right] .
687: \label{Hieq}
688: \eeq
689:
690: Finally, using the transformations (\ref{cKdef}) and (\ref{Gdef}), we obtain
691: in terms of the original variables
692: \beq
693: K_{x,c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2) \, \dd \psi_2 \dd v_2 =
694: e^{-\tau/\gamma^2+(u_2-u_1)/\gamma} \, G(\tau;r_1,u_1;r_2,u_2) \, \dd r_2 \dd u_2 ,
695: \label{KG1}
696: \eeq
697: with
698: \beq
699: \tau=\gamma^2 (Q_2-Q_1) , \;\;\;
700: r_i=2\gamma^3\left[ \Psi_i+\frac{(Q_i-X)^2}{2}-C \right],
701: \;\;\; u_i= 2\gamma (V_i+Q_i-X) .
702: \label{riui}
703: \eeq
704: Here we introduced the dimensionless spatial coordinates
705: (which we shall note by capital letters in this article)
706: \beq
707: Q=\frac{q}{\gam^2} = \frac{q}{2Dt^2} , \hspace{0.5cm}
708: X= \frac{x}{\gam^2} = \frac{x}{2Dt^2} , \hspace{0.5cm} \mbox{with}\hspace{0.5cm}
709: \gam = \sqrt{2D} \, t ,
710: \label{QXdef}
711: \eeq
712: and the dimensionless velocity
713: \beq
714: V=\frac{tv}{\gam^2}=\frac{v}{2Dt} , \;\;\; \mbox{whence} \;\;\; X=Q+V \;\;\;
715: \mbox{for regular points} .
716: \label{Vdef}
717: \eeq
718: In a similar fashion, the dimensionless velocity potential coordinates
719: in (\ref{riui}) are
720: \beq
721: \Psi = \frac{t\psi}{\gamma^4} \;\;\;\; \mbox{and} \;\;\;\; C = \frac{t c}{\gamma^4} .
722: \label{PsiCdef}
723: \eeq
724: Next, from Eq.(\ref{Hidef}) the kernel associated with Brownian particles that
725: remain forever above the parabola $\cP_{x,c}$ reads as
726: \beq
727: \lim_{q_2\rightarrow+\infty} \int \dd\psi_2\dd v_2 \,
728: K_{x,c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2) = e^{-u_1/\gamma} \Hi(r_1,u_1) ,
729: \label{KcHi}
730: \eeq
731: whereas the propagator associated with Brownian particles that come within a small
732: distance $\delta c$ from the parabolic absorbing barrier is from Eq.(\ref{Deltadef})
733: \beqa
734: \lim_{\delta c \rightarrow 0} \, \frac{1}{\delta c}
735: [K_{x,c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2)
736: - K_{x,c+\delta c}(q_1,\psi_1,v_1;q_2,\psi_2,v_2)] \, \dd\psi_2\dd v_2 & = &
737: \nonumber \\
738: && \hspace{-5cm} 2\frac{t}{\gamma} \, e^{-\tau/\gamma^2+(u_2-u_1)/\gamma}
739: \, \Delta(\tau;r_1,u_1;r_2,u_2) \, \dd r_2 \dd u_2 .
740: \label{KDelta}
741: \eeqa
742:
743:
744: \section{One-point distributions}
745: \label{sec:velocity1}
746:
747:
748:
749: \subsection{Results for arbitrary Eulerian location $x$}
750: \label{subsec:arbitrary_Eulerian_location_x}
751:
752:
753:
754: In this section we consider the one-point velocity distribution $p_x(v)$
755: at the Eulerian location $x$. From the explicit solution (\ref{psinu0}),
756: it can be derived from the probability distribution $p_x(q)$ of the Lagrangian
757: coordinate $q(x,t)$. Thus, we have from Eq.(\ref{psinu0})
758: \beq
759: p_x(v) = t \, p_x(q) \hspace{0.8cm} \mbox{and} \hspace{0.8cm} q = x - v t ,
760: \label{pvpq}
761: \eeq
762: where we note $p_x(v)$ and $p_x(q)$ the probability distributions of the velocity
763: $v$ and of the Lagrangian coordinate $q$, at the Eulerian location $x$ and time
764: $t$.
765: Here we used the property that $q(x,t)$ is well defined for any $x$ except
766: over a set of zero measure in Eulerian space associated with shocks
767: \cite{She1992}.
768:
769: Then, from the geometrical construction (\ref{paraboladef}),
770: we are led to consider the bivariate probability distribution,
771: $p_x(0\leq q'\leq q,c)\dd c$, that the first contact point of the potential
772: $\psi_0(q')$
773: with the family of downward parabolas $\cP_{x,c}(q')$, with $c$ increasing from
774: $-\infty$, occurs at an abscissa $q'$ in the range $0\leq q'\leq q$, with a
775: parabola of height between $c$ and $c+\dd c$.
776: This will give us in turn the cumulative distribution $p_x(0\leq q' \leq q)$
777: by integrating over $c$.
778: Then, for $q \geq 0$, we can write this probability distribution as
779: \beqa
780: p_x(0\leq q'\leq q,c)\dd c & = & \lim_{q_{\pm}\rightarrow\pm\infty}
781: \int \dd\psi_-\dd v_- \dd\psi\dd v \dd\psi_+\dd v_+
782: \, K_{x,c}(0,0,0;q_-,\psi_-,v_-) \nonumber \\
783: && \hspace{-0.5cm} \times \, [ K_{x,c}(0,0,0;q,\psi,v) - K_{x,c+\dd c}(0,0,0;q,\psi,v) ]
784: \, K_{x,c}(q,\psi,v;q_+,\psi_+,v_+) ,
785: \label{pxcq1}
786: \eeqa
787: where we used the Markovian character of the process $q\mapsto\{\psi,v\}$.
788: Thus, we could factorize in Eq.(\ref{pxcq1}) the probability
789: $p_x(0\leq q'\leq q,c)\dd c$ into three terms, which correspond to the
790: probabilities that i) $\psi_0(q')$ stays above $\cP_{x,c}$ for $q'<0$,
791: ii) $\psi_0(q')$ stays above $\cP_{x,c}$, but does not everywhere remain above
792: $\cP_{x,c+\dd c}$, over the range $0\leq q'\leq q$, while reaching an arbitrary value
793: $\{\psi,v\}$ at $q$, over which we will integrate, and iii) $\psi_0(q')$
794: stays above $\cP_{x,c}$ for $q '> q$. We show in Fig.~\ref{figP1} the geometrical
795: interpretation of Eq.(\ref{pxcq1}) (where we did not try to draw an actual Brownian
796: curve $\psi_0(q)$ which has no finite second-derivative).
797:
798:
799:
800: \begin{figure}
801: \begin{center}
802: \epsfxsize=9.5 cm \epsfysize=6. cm {\epsfbox{P1.eps}}
803: \end{center}
804: \caption{(color online) Geometrical interpretation of the initial conditions $\psi_0(q'')$
805: associated with the probability $p_x(0\leq q'\leq q,c)\dd c$. The Brownian curve
806: $\psi_0(q'')$ is everywhere above the parabola $\cP_{x,c}$ and goes below
807: $\cP_{x,c+\dd c}$ somewhere in the range $0\leq q''\leq q$. From the constraints
808: $\psi_0(0)=0$ and $\psi_0'(0)=0$, see Eqs.(\ref{xidef}), it goes through the
809: origin with an horizontal tangent. To obtain the cumulative probability,
810: $p_x(0\leq q'\leq q)$, we must then integrate over the height $c$ of the parabola.}
811: \label{figP1}
812: \end{figure}
813:
814:
815:
816: We can easily check that in the limit $x\rightarrow +\infty$ and
817: $q\rightarrow +\infty$ with $q\gg x$, the integral over $c$ of Eq.(\ref{pxcq1})
818: gives unity as it should. It is convenient to first compute cumulative
819: probabilities as in (\ref{pxcq1}) and to take the derivatives afterwards to
820: derive the probability densities. This ensures that probabilities are well
821: normalized and it avoids coming across ill-defined expressions. Indeed,
822: since the curve $\psi_0(q)$ has a continuous derivative, it is tangent to
823: the parabola $\cP_{x,c}$ at the first contact point. Then, this point corresponds to
824: $r=0$ and $u=0$ in terms of the reduced variables (\ref{xu}), where the Brownian
825: kernels are singular. For instance, the expression (\ref{tG0}) is not well
826: defined if we naively put $r_1=r_2=0$. Other ambiguities or seemingly divergent
827: quantities appear if we try to directly compute probability densities by using
828: Taylor expansions.
829:
830: Then, using the relations (\ref{KG1})-(\ref{KDelta}), we obtain
831: \beq
832: p_x(0\leq q'\leq q,r_0)\dd r_0 = e^{-q/\gam^2} \dd r_0 \int \dd r \dd u
833: \, \Hi(r_0,\hu) \Delta(q;r_0,-\hu;r,u) \Hi(r,u) ,
834: \label{pHHD}
835: \eeq
836: where we defined
837: \beq
838: \hu= \sqrt{\frac{2}{D}} \frac{x}{t}
839: = \frac{2x}{\gamma} .
840: \label{hudef}
841: \eeq
842: Using the results of section~\ref{sec:transition}, the integration over
843: $r$ and $r_0$ gives
844: \beq
845: p_x(0\leq q'\leq q) = \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q} \, I(s) J(s,2X) ,
846: \label{pxQX1}
847: \eeq
848: where we introduced the dimensionless variables $Q$ and $X$ as in (\ref{QXdef})
849: and we defined the functions
850: \beq
851: I(s) = \int_{-\infty}^{\infty} \dd z \, J(s,2z) ,
852: \label{Is1}
853: \eeq
854: and
855: \beqa
856: J(s,y) & = & e^y \int_0^{\infty} \frac{\dd\nu}{\sqrt{\pi}} \, 3\nu^{-3/2}
857: \, e^{-\frac{2}{3} s^{3/2} \nu^{-3}}
858: \Ai\left[-\nu y+\frac{s}{\nu^2}\right] \nonumber \\
859: && - \int_0^{\infty} \frac{\dd\nu\dd\mu}{\pi} \,
860: \frac{9\nu^{-3/2}\mu^{3/2}}{\nu^3+\mu^3} \,
861: e^{-\frac{2}{3} (\nu^{-3}+s^{3/2}\mu^{-3})}
862: \Ai\left[\nu y+\frac{1}{\nu^2}\right] \Ai\left[-\mu y+\frac{s}{\mu^2}\right] .
863: \label{JsX1}
864: \eeqa
865: For $y\geq 0$, we obtain using Eqs.(\ref{phi0int})-(\ref{Expu_int}),
866: \beq
867: y\geq 0 : \;\;\; J(s,y) = s^{-1/4} e^{(1-\sqrt{s})y} - 6 \int_0^{\infty}
868: \frac{\dd\nu}{\nu^2} \, e^{\frac{2}{3} (s^{3/2}-1) \nu^{-3}}
869: \Ai\left[\nu y+\frac{1}{\nu^2}\right] \Ai\left[\nu y+\frac{s}{\nu^2}\right] .
870: \label{JsX2}
871: \eeq
872: For $y\leq 0$, using Eq.(\ref{Expintkl}) in the second term of Eq.(\ref{JsX1}), we
873: obtain
874: \beq
875: y\leq 0 : \;\;\; J(s,y) = 6 \int_0^{\infty}
876: \frac{\dd\mu}{\mu^2} \, e^{-\frac{2}{3} (s^{3/2}-1) \mu^{-3}}
877: \Ai\left[-\mu y+\frac{1}{\mu^2}\right] \Ai\left[-\mu y+\frac{s}{\mu^2}\right] .
878: \label{JsX3}
879: \eeq
880: Therefore, since we have the primitive
881: \beqa
882: \int \dd u \, \Ai\left[\nu u+\frac{s_1}{\nu^2}\right]
883: \Ai\left[\nu u+\frac{s_2}{\nu^2}\right] & = & \frac{\nu}{s_1-s_2}
884: \nonumber \\
885: && \hspace{-2cm} \times \left\{ \Ai\,'\left[\nu u+\frac{s_1}{\nu^2}\right]
886: \Ai\left[\nu u+\frac{s_2}{\nu^2}\right] - \Ai\left[\nu u+\frac{s_1}{\nu^2}\right]
887: \Ai\,'\left[\nu u+\frac{s_2}{\nu^2}\right] \right\} ,
888: \label{prim1}
889: \eeqa
890: the integral (\ref{Is1}) reads as
891: \beq
892: I(s) = \frac{s^{-1/4}}{2(\sqrt{s}-1)} + \frac{3}{s-1} \int_{-\infty}^{\infty} \!
893: \frac{\dd\nu}{\nu} \, e^{\frac{2}{3}(s^{3/2}-1)\nu^{-3}} \left[
894: \Ai\,'\left(\frac{s}{\nu^2}\right) \Ai\left(\frac{1}{\nu^2}\right) \!
895: - \! \Ai\left(\frac{s}{\nu^2}\right) \Ai\,'\left(\frac{1}{\nu^2}\right) \right] .
896: \label{Is2}
897: \eeq
898: Using Eq.(\ref{intmu1AipAiui02}) this yields the simple result
899: \beq
900: I(s) = \frac{1}{s-1} .
901: \label{Isresult}
902: \eeq
903: Therefore, in terms of dimensionless variables, the cumulative probability
904: (\ref{pxQX1}) reads as
905: \beq
906: P_X(0\leq Q'\leq Q) = \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q} \,
907: \frac{J(s,2X)}{s-1} .
908: \label{PXQ2}
909: \eeq
910: On the other hand, since the system is statistically invariant through reflection
911: about the origin, we have the symmetry
912: $p_x(0\leq q'\leq q) = p_{-x}(-q\leq q'\leq 0)$. This implies that the cumulative
913: probability distribution associated with a Lagrangian coordinate $q'$ on the
914: negative real axis reads as
915: \beq
916: P_X(-Q\leq Q'\leq 0) = \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q} \,
917: \frac{J(s,-2X)}{s-1} .
918: \label{PXQm1}
919: \eeq
920: We can also check Eq.(\ref{PXQm1}) through an explicit calculation similar to
921: (\ref{pxcq1}).
922:
923: From Eq.(\ref{pvpq}) and Eqs.(\ref{PXQ2})-(\ref{PXQm1}), the cumulative velocity
924: distribution is given by
925: \beqa
926: v\leq \frac{x}{t} & : & \;\;\; p_x(v\leq v'\leq x/t) = p_x(0\leq q'\leq x-v t)
927: = P_X(0\leq Q' \leq X-V) ,
928: \label{pxv1} \\
929: v\geq \frac{x}{t} & : & \;\;\; p_x(x/t\leq v'\leq v) = p_x(x-v t\leq q'\leq 0)
930: = P_{-X}(0\leq Q'\leq V-X) ,
931: \label{pxv2}
932: \eeqa
933: where we introduced the dimensionless velocity $V$ defined as in (\ref{Vdef}).
934: Of course, Eqs.(\ref{PXQ2})-(\ref{pxv2}) agree with the
935: scalings (\ref{scalings}).
936:
937:
938: Letting $|Q|\rightarrow\infty$ in Eqs.(\ref{PXQ2})-(\ref{PXQm1}), or
939: $|V|\rightarrow\infty$ in Eqs.(\ref{pxv1})-(\ref{pxv2}), we obtain the
940: probabilities that the Lagrangian coordinate $q$, associated with the Eulerian
941: coordinate $x$, is located on either side of the origin (or that the velocity
942: $v$ is smaller or greater than $x/t$):
943: \beqa
944: p_x(q\geq 0) = p_x(v\leq x/t) & = & J(1,2X) ,
945: \label{pqp1} \\
946: p_x(q\leq 0) = p_x(v\geq x/t) & = & J(1,-2X) .
947: \label{pqm1}
948: \eeqa
949: Here we used the fact that the large-$Q$ behavior of Eqs.(\ref{PXQ2})-(\ref{PXQm1})
950: is set by the rightmost singularity of the ratio $J(s,\pm 2X)/(s-1)$, which is
951: the simple pole at $s=1$. From Eqs.(\ref{JsX2})-(\ref{JsX3}) we obtain for $x\geq 0$:
952: \beq
953: x\geq 0 : \;\;\; p_x(q\leq 0) = \int_0^{\infty} \dd\nu \, \frac{6}{\nu^2} \,
954: \Ai\left[\nu 2X+\frac{1}{\nu^2}\right]^2 ,
955: \;\;\;\;\;\; p_x(q\geq 0) = 1 - p_x(q\leq 0) .
956: \label{pqpm2}
957: \eeq
958: We can check that the sum of these two probabilities is equal to unity.
959: As expected, Eq.(\ref{pqpm2}) shows that $p_x(q\leq 0)$ decreases as $x$ gets larger
960: and it goes to zero for $x\rightarrow+\infty$. For $x=0$, both quantities are equal
961: to $J(1,0)=1/2$, as can be checked from the explicit computation of the integral
962: in Eq.(\ref{pqpm2}).
963:
964: Finally, from Eqs.(\ref{PXQ2})-(\ref{PXQm1}) the probability densities are given
965: by:
966: \beq
967: Q\geq 0: \;\; P_X(Q) = \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q} J(s,2X) ,
968: \;\;\;\;\; P_X(-Q) = \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q} J(s,-2X) .
969: \label{PXQdens}
970: \eeq
971: This also gives the velocity distributions through the relation (\ref{Vdef}).
972:
973:
974:
975: \subsection{Velocity distribution at the origin $x=0$}
976: \label{subsec:origin_x=0}
977:
978:
979:
980:
981:
982: \begin{figure}
983: \begin{center}
984: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{P0v.ps}}
985: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{lP0v.ps}}
986: \end{center}
987: \caption{(color online) {\it Left panel:} The probability distribution $P_0(V)$ of the
988: reduced velocity $V=v/(2Dt)$ at the origin $x=0$, from Eq.(\ref{p0V}). The
989: dashed lines show the asymptotic behaviors (\ref{p0V0}) and (\ref{p0Vinfty}).
990: {\it Right panel:} Same as left panel but on a logarithmic scale.}
991: \label{figP0v}
992: \end{figure}
993:
994:
995:
996: We consider here the one-point distribution at the origin $x=0$.
997: From Eq.(\ref{PXQdens}) we can check that the distribution
998: is even (we have $Q=-V$ at $X=0$). For $V\geq 0$ it is given by
999: \beq
1000: V\geq 0 : \;\;\; P_0(V) = \inta\frac{\dd s}{2\pi\ii}\, e^{(s-1)V} \, J(s,0) ,
1001: \;\;\;\; P_0(-V) = P_0(V) .
1002: \label{p0V}
1003: \eeq
1004: The behavior for $V\rightarrow 0^+$ is determined by the behavior at
1005: $s\rightarrow+\infty$ of $J(s,0)$. From Eq.(\ref{JsX3}) and using
1006: Eq.(\ref{IntAiExp1}) we obtain
1007: \beq
1008: s\rightarrow+\infty: \;\;\; J(s,0) \sim \frac{\sqrt{3}}{\pi} \, s^{-1/2} ,
1009: \label{Js0infty}
1010: \eeq
1011: which leads to
1012: \beq
1013: V \rightarrow 0^+ : \;\;\; P_0(V) \sim \frac{1}{\pi} \sqrt{\frac{3}{\pi V}} .
1014: \label{p0V0}
1015: \eeq
1016: Thus, we obtain an inverse square-root divergence for $P_0(V)$ at
1017: $V \rightarrow 0$.
1018:
1019: The behavior of $P_0(V)$ for large $V$ is governed by the
1020: rightmost singularity of $J(s,0)$, located at $s=0$ (associated with the
1021: branch cut along the negative real axis). There, $J(s,0)$ behaves as
1022: $J(s,0) \sim s^{-1/4}$, because of the first term in Eq.(\ref{JsX2}). This yields
1023: \beq
1024: V \rightarrow +\infty : \;\;\; P_0(V) \sim \frac{1}{\Gamma[1/4]} \, V^{-3/4}
1025: \, e^{-V} .
1026: \label{p0Vinfty}
1027: \eeq
1028: Note that initially, at time $t=0$, the velocity at the origin is not random
1029: as it is equal to zero, see (\ref{xidef}). Then, for $t>0$ the nonlinear
1030: evolution of the velocity field $v(x,t)$ broadens this initial Dirac peak and
1031: gives rise to the exponential tail (\ref{p0Vinfty}) at large velocities and to
1032: the power-law peak (\ref{p0V0}) at low velocities.
1033: We show in Fig.~\ref{figP0v} the velocity distribution $P_0(V)$, as well as
1034: the asymptotic behaviors (\ref{p0V0}) and (\ref{p0Vinfty}), that happen to
1035: describe very well most of the distribution.
1036:
1037: We can note that since all quantities can be expressed in terms of the scaling
1038: variables (\ref{QXdef})-(\ref{Vdef}), the exponential tail (\ref{p0Vinfty})
1039: can be understood from simple scaling arguments applied to the initial velocity
1040: field. Thus, for a particle of initial Lagrangian position $q>0$ to reach the
1041: Eulerian position $x=0$ at time $t$, we can expect its initial velocity to be of
1042: order $v_0\sim -q/t$. From Eq.(\ref{Gaussian_t0}) this corresponds to a probability
1043: of order $e^{-v_0^2/(2\sigma_{v_0}^2(q))} \sim e^{-q/t^2} \sim e^{-Q}$,
1044: where we did not write factors of order unity in the exponent, which cannot be
1045: obtained by such arguments. Thus, we recover the exponential tail (\ref{p0Vinfty})
1046: (at $X=0$ we have $V=-Q$).
1047:
1048:
1049:
1050:
1051: \subsection{Velocity distribution for $|x|\rightarrow\infty$}
1052: \label{subsec:far_away_x}
1053:
1054:
1055:
1056: Finally, we consider the one-point velocity distribution at large $|x|$.
1057: By symmetry, we only need consider $x\rightarrow+\infty$.
1058: Using the relation $X=Q+V$ and Eq.(\ref{PXQdens}), we can write the velocity
1059: distribution in terms of the reduced variables $X$ and $V$ as
1060: \beq
1061: V\leq X : \;\;\; P_X(V) = \inta\frac{\dd s}{2\pi\ii}\, e^{(s-1)(X-V)} \, J(s,2X) .
1062: \label{pXinfV1}
1063: \eeq
1064: We now consider the limit $X\rightarrow+\infty$ at fixed $V$. Then, making the
1065: change of variable $s=1+\ii k$, we obtain at leading order ($k$ being of order
1066: $X^{-1/2}$)
1067: \beq
1068: P_{X}(V) \sim \int_{-\infty}^{\infty} \frac{\dd k}{2\pi}\, e^{\ii k(X-V)}
1069: \, e^{(1-\sqrt{1+\ii k})2X} \sim
1070: \int_{-\infty}^{\infty} \frac{\dd k}{2\pi}\, e^{-\ii k V - X k^2/4}
1071: = \frac{e^{-V^2/X}}{\sqrt{\pi X}} .
1072: \label{pXinfV2}
1073: \eeq
1074: Therefore, in terms of the variable $v$, we recover as expected the initial Gaussian
1075: (\ref{Gaussian_t0}).
1076: This can be understood as follows.
1077: A remote region $[x-L/2,x+L/2]$, with $L\ll x$, has a mean initial velocity
1078: $v_0 \sim \sqrt{Dx}$ that is much larger than its initial velocity dispersion
1079: $\Delta v_0 \sim \sqrt{DL}$, see Eq.(\ref{Dv0}). Then, this domain remains
1080: well-defined and not strongly disturbed by neighboring regions until times
1081: of order $t_*$ with $\Delta v_0 t_*=L$, that is $D t_*^2=L$. Conversely,
1082: at any time $t$, for $x\gg D t^2$ (i.e. $X\gg 1$) it is possible to make
1083: such a separation of scales and to identify a region of size $L$ around $x$,
1084: with $D t^2\ll L \ll x$, that moves in a collective fashion with a mean velocity
1085: $\simeq v_0(x)$ that is set by the initial velocity. Therefore, we recover at
1086: leading order the initial Gaussian velocity distribution, of variance
1087: $\sigma_v = \sqrt{D x}$ (i.e. $\sigma_V= \sqrt{X/2} \gg 1$), and the nonlinear
1088: evolution only modifies the velocity distribution by changes of order
1089: $\Delta v \sim D t$ (i.e. $\Delta V \sim 1$).
1090: The result (\ref{pXinfV2}) confirms this simple scaling argument.
1091: This is an illustration of the ``principle of permanence of large eddies''
1092: \cite{Gurbatov1997}, that holds for more general energy spectra,
1093: $E_0(k) \propto k^n$, with $n<1$.
1094: As suggested by this discussion, and as checked in numerical simulations
1095: \cite{Aurell1993,Gurbatov1999}, the stability of large-scale structures is not
1096: only a statistical property but actually holds on an individual basis, that is
1097: for each random realization of the velocity field.
1098:
1099:
1100:
1101:
1102:
1103: \begin{figure}
1104: \begin{center}
1105: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{Pqnegative.ps}}
1106: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{lPqnegative.ps}}
1107: \end{center}
1108: \caption{(color online) {\it Left panel:} The probability $p_x(q\leq 0)$ (equal to the reduced
1109: cumulative probability $P_X(Q\leq 0)$), that a particle, located
1110: at a position $x>0$ at time $t$, was initially located on the negative real axis,
1111: $q<0$, from Eq.(\ref{pqpm2}). The dashed line is the asymptotic behavior
1112: (\ref{pqmxinf}).
1113: {\it Right panel:} Same as left panel but on a logarithmic scale.}
1114: \label{figPqnegative}
1115: \end{figure}
1116:
1117:
1118:
1119: Of course, this reasoning does not apply to rare events, such as those where the
1120: displacement $x-q$ remains of order $x$. In particular, from Eq.(\ref{pqpm2}),
1121: we obtain for the cumulative probability to have a negative Lagrangian coordinate
1122: $q$ the asymptotic behavior
1123: \beq
1124: x \rightarrow+\infty : \;\;\; p_x(q\leq 0) = p_x(v\geq x/t) \sim
1125: \left(8\pi\sqrt{3}X\right)^{-1/2} \, e^{-4\sqrt{3}X} .
1126: \label{pqmxinf}
1127: \eeq
1128: Thus, we obtain an exponential tail for these very rare events. It can again
1129: be understood from simple scaling arguments, as for the exponential tail
1130: (\ref{p0Vinfty}). Thus, for a particle with Lagrangian coordinate $q<0$ to reach
1131: the position $x \gg 2Dt^2$, we can associate the initial velocity $v_0=(x-q)/t$
1132: and the probability $e^{-(x-q)^2/(t^2|q|)}$, using Eq.(\ref{Gaussian_t0}) without
1133: writing numerical factors. Then, the maximum over $q<0$ of this exponential weight
1134: is reached for $q=-x$, which gives a weight $\sim e^{-x/t^2} \sim e^{-X}$
1135: that agrees with Eq.(\ref{pqmxinf}). We show in Fig.~\ref{figPqnegative}
1136: the probability $p_x(q\leq 0)$, as well as the asymptotic decay (\ref{pqmxinf}).
1137:
1138:
1139:
1140:
1141: \section{Two-point and higher-order distributions}
1142: \label{sec:velocity2}
1143:
1144:
1145: \subsection{General results for $x_1 < x_2$ and $0<q_1<q_2$}
1146: \label{subsec:general}
1147:
1148:
1149:
1150: We now study the two-point Eulerian velocity distribution $p_{x_1,x_2}(v_1,v_2)$,
1151: with $x_1 < x_2$. As in section~\ref{sec:velocity1}, we first consider the
1152: distribution $p_{x_1,x_2}(q_1,q_2)$ of the Lagrangian coordinates $q_1,q_2$,
1153: associated with the Eulerian positions $x_1,x_2$.
1154: For the Brownian initial conditions
1155: (\ref{xidef})-(\ref{Ddef}) shocks are dense \cite{Sinai1992,She1992}.
1156: Therefore, for $x_1<x_2$ there is almost surely
1157: a shock between $x_1$ and $x_2$ and these two Eulerian points are associated
1158: with two different Lagrangian coordinates $q_1 \neq q_2$.
1159: This can also be understood from the fact that at the contact point $q_1$
1160: (resp. $q_2$) the curve $\psi_0(q)$ is tangent to a parabola $\cP_{x_1,c_1}(q)$
1161: (resp.$\cP_{x_2,c_2}(q)$), from the geometric construction recalled in
1162: (\ref{paraboladef}). Then, since two parabolas $\cP_{x_1,c_1}$ and $\cP_{x_2,c_2}$
1163: with $x_1\neq x_2$ have different tangents at any point $q$ (indeed
1164: $\dd \cP_{x,c}/\dd q = - (q-x)/t$), the curve $\psi_0(q)$ cannot be tangent to
1165: both parabolas at a common point $q_1=q_2$ (in both steps we used the property
1166: that the derivative $\psi_0'(q)$ is continuous, being a Brownian motion). Therefore,
1167: we almost surely have $q_1\neq q_2$. Then, since particles do not cross each other
1168: we have $q_1 < q_2$ for $x_1 < x_2$.
1169:
1170:
1171:
1172: \begin{figure}
1173: \begin{center}
1174: \epsfxsize=9.5 cm \epsfysize=5.5 cm {\epsfbox{P2s.eps}}
1175: \end{center}
1176: \caption{(color online) Geometrical interpretation of the initial conditions $\psi_0(q)$
1177: associated with the contribution
1178: $p_{x_1,x_2}^>(0\leq q_1'\leq q_1,c_1;q_2'\geq q_2)\dd c_1$. The Brownian curve
1179: $\psi_0(q)$ is everywhere above the parabola $\cP_{x_1,c_1}$, it goes below
1180: $\cP_{x_1,c_1+\dd c_1}$ somewhere in the range $0\leq q_1'\leq q_1$, and it goes
1181: below the parabola $\cP_{x_2,c_*}$, of center $x_2$, that intersects $\cP_{x_1,c_1}$
1182: at $q_*=q_2$. This counts all paths with a first-contact parabola $\cP_{x_2,c_2}$
1183: such that $c_2\leq c_*$ and $q_*\geq q_2$ (which implies $q_2'\geq q_2$).}
1184: \label{figP2s}
1185: \end{figure}
1186:
1187:
1188:
1189: As in section~\ref{subsec:arbitrary_Eulerian_location_x}, we first consider the
1190: cumulative probability distribution, $p_{x_1,x_2}(0\leq q_1'\leq q_1;q_2'\geq q_2)$,
1191: that the Lagrangian coordinates $q_1',q_2'$, associated with the Eulerian positions
1192: $x_1,x_2$, are within the ranges $0\leq q_1'\leq q_1$ and $q_2\leq q_2'< +\infty$.
1193: Let us consider this probability in two steps. First, as for Eq.(\ref{pxcq1}),
1194: we consider the initial conditions such that $\psi_0(q)$ stays everywhere above
1195: a parabola $\cP_{x_1,c_1}$ but goes below $\cP_{x_1,c_1+\dd c_1}$ somewhere in
1196: the range $0\leq q_1'\leq q_1$. Integrating over the height $c_1$ this will take
1197: care of the first constraint $0\leq q_1'\leq q_1$ for the Lagrangian coordinate
1198: associated with $x_1$. Second, we must only count among those initial conditions
1199: the ones that also satisfy $q_2'\geq q_2$. We split them into two contributions as
1200: follows. Let us note $q_*$ the unique abscissa where the two contact parabolas
1201: $\cP_{x_1,c_1}$ and $\cP_{x_2,c_2}$ intersect. From Eq.(\ref{paraboladef})
1202: it is given by
1203: \beq
1204: q_* = \frac{x_1+x_2}{2} - \frac{c_2-c_1}{x_2-x_1} t .
1205: \label{q*def}
1206: \eeq
1207: Then, we note $p^>$ the first contribution, associated with initial conditions
1208: such that $q_*>q_2$ (which implies $q_2'>q_*>q_2$).
1209: Clearly, this actually corresponds to curves $\psi_0(q)$
1210: that at some point go below the parabola $\cP_{x_2,c_*}$ where $c_*$ is such
1211: that $q_*=q_2$ (i.e. the second parabola intersects $\cP_{x_1,c_1}$ at $q_2$).
1212: We note $p^<$ the second contribution, associated with $q_1<q_*<q_2$
1213: (since afterwards we shall consider the probability density
1214: $p_{x_1,x_2}(q_1;q_2'\geq q_2)$ we only need to include the cases with $q_*>q_1$).
1215: We show in Figs.~\ref{figP2s} and \ref{figP2m} the geometrical interpretation
1216: of these two contributions $p^>$ and $p^<$.
1217:
1218:
1219:
1220: \begin{figure}
1221: \begin{center}
1222: \epsfxsize=9.5 cm \epsfysize=5.5 cm {\epsfbox{P2m.eps}}
1223: \end{center}
1224: \caption{(color online) Geometrical interpretation of the initial conditions $\psi_0(q)$
1225: associated with the contribution
1226: $p_{x_1,x_2}^<(0\leq q_1'\leq q_1,c_1;q_2'\geq q_2,c_2)\dd c_1\dd c_2$.
1227: The Brownian curve $\psi_0(q)$ is everywhere above the parabolas $\cP_{x_1,c_1}$
1228: and $\cP_{x_2,c_2}$, it goes below $\cP_{x_1,c_1+\dd c_1}$ somewhere in the range
1229: $0\leq q_1'\leq q_1$, and below the parabola $\cP_{x_2,c_2+\dd c_2}$
1230: somewhere in the semi-infinite range $q_2'\geq q_2$. The height $c_2$ of the second
1231: parabola is such that both parabolas intersect at $q_*$ in the range
1232: $q_1\leq q_*\leq q_2$.}
1233: \label{figP2m}
1234: \end{figure}
1235:
1236:
1237:
1238: We describe in appendix~\ref{Computation-of-two-point} the computation of the
1239: two-point distribution $p_{x_1,x_2}(q_1,q_2)$ from these two contributions $p^>$
1240: and $p^<$. As for the one-point distribution computed in
1241: section~\ref{subsec:arbitrary_Eulerian_location_x}, we first express the kernels
1242: $K$ in terms of the Brownian propagator $G$ obtained in section~\ref{sec:transition}
1243: and we use various properties of the Airy functions, described in
1244: Appendices~\ref{Airy} and \ref{Half-range}, to simplify the integrals.
1245: We finally obtain for the sum of both contributions the probability density
1246: \beq
1247: P_{X_1,X_2}(Q_1,Q_2) = \inta \frac{\dd s_1 \dd s_2}{(2\pi\ii)^2} \,
1248: e^{(s_1-1) Q_1+(s_2-1)Q_{21}} J(s_1,2X_1) \, e^{-(\sqrt{s_2}-1)2 X_{21}} .
1249: \label{PX1X2}
1250: \eeq
1251: Comparing with the one-point probability density (\ref{PXQdens}), we find that
1252: for $0\leq Q_1\leq Q_2$ the two-point probability density factorizes as
1253: \beq
1254: X_1\leq X_2, \;\; 0\leq Q_1\leq Q_2 : \;\;\; P_{X_1,X_2}(Q_1,Q_2) = P_{X_1}(Q_1)
1255: \bP_{X_{21}}(Q_{21}) ,
1256: \label{PX1X2fact}
1257: \eeq
1258: where we introduced
1259: \beq
1260: \bP_X(Q) = \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q} \, e^{-(\sqrt{s}-1)2 X} .
1261: \label{pBdef}
1262: \eeq
1263: Therefore, the conditional probability $P(X_2,Q_2|X_1,Q_1)$ obeys the property
1264: \beq
1265: X_1\leq X_2, \;\; 0\leq Q_1\leq Q_2 : \;\;\; P(X_2,Q_2|X_1,Q_1) =
1266: \frac{P_{X_1,X_2}(Q_1,Q_2)}{P_{X_1}(Q_1)} = \bP_{X_{21}}(Q_{21}) ,
1267: \label{Pcond}
1268: \eeq
1269: that is, it only depends on the relative distances $X_{21}$ and $Q_{21}$,
1270: and no longer on $Q_1$ and $X_1$, over the range $Q_1\geq 0$.
1271: Thus, the system is statistically homogeneous with respect
1272: to Lagrangian and velocity increments as long as we remain on either side
1273: of the origin $Q=0$. Indeed, by symmetry through reflection about the origin
1274: we also have
1275: \beqa
1276: X_1\leq X_2, \;\; Q_1\leq Q_2\leq 0 : \;\;\; P_{X_1,X_2}(Q_1,Q_2)
1277: & = & P_{-X_2,-X_1}(-Q_2,-Q_1) \nonumber \\
1278: && \hspace{-0.5cm} = P_{-X_2}(-Q_2) \bP_{X_{21}}(Q_{21})
1279: = P_{X_2}(Q_2) \bP_{X_{21}}(Q_{21}) .
1280: \label{PX1X2mfact}
1281: \eeqa
1282: Then, in the limits $X_1\rightarrow+\infty$, or $X_2\rightarrow-\infty$, where the
1283: weight of configurations such that $Q_1$ and $Q_2$ have different signs vanishes,
1284: we recover the invariance through translation of the probability distributions
1285: of relative displacements and velocity increments. Of course, this is related to the
1286: fact that the initial conditions have homogeneous velocity increments,
1287: see Eq.(\ref{Dv0}). However, the invariance through translation is only recovered
1288: in the exact nonlinear velocity distribution if we go infinitely far from the
1289: origin $Q=0$. As we get closer to the origin it is broken by the increasing weight
1290: of configurations, such that $Q_1$ and $Q_2$ are on different sides of the origin,
1291: which do not satisfy the factorizations (\ref{PX1X2fact}) or (\ref{PX1X2mfact}).
1292: Indeed, note that Eq.(\ref{PX1X2mfact}) shows that the factorization
1293: (\ref{PX1X2fact}) cannot be extended to $Q_1<0$, as for $Q_1<0$ and $Q_2<0$
1294: we must reach the other regime (\ref{PX1X2mfact}) that cannot hold simultaneously.
1295:
1296: Thus, at finite distance from the origin the invariance through translation is
1297: always partly broken for the distribution $P_{x_1,x_2}(q_1,q_2)$ considered over
1298: the full range $-\infty<q_1\leq q_2<\infty$. Nevertheless, the invariance
1299: is exactly recovered over either the partial range $0\leq q_1\leq q_2<\infty$,
1300: or $-\infty<q_1\leq q_2\leq 0$. This can be understood as follows, focussing
1301: on the case $q_1>0$ with again $x_1<x_2$. The probability density $p_{x_1}(q_1)$
1302: counts the configurations $\psi_0(q)$ that are tangent at $q_1$ with the highest
1303: first contact parabola $\cP_{x_1,c_1}$, from the geometric construction
1304: described below (\ref{paraboladef}). Then, the conditional probability density
1305: $p(x_{21},q_{21}|x_1,q_1)$ only counts among those the configurations that are also
1306: tangent at $q_2$ with the highest first contact parabola $\cP_{x_2,c_2}$.
1307: Independently of the behavior of the curve $\psi_0$ on either side of $q_1$,
1308: the first-contact height parameter $c_2$ must be smaller than the value $c_*$
1309: such that $\cP_{x_2,c_*}$ runs through the point $\{q_1,\psi_0(q_1)\}$.
1310: Then, for any $c_2<c_*$, we clearly have $\cP_{x_2,c_2}(q)<\cP_{x_1,c_1}(q)$
1311: for all $q<q_1$ (using $x_1<x_2$), whence $\cP_{x_2,c_2}(q)<\psi_0(q)$ for all
1312: $q<q_1$ since we have already selected those configurations
1313: associated with $p_{x_1}(q_1)$ that are above $\cP_{x_1,c_1}$ (and make contact
1314: at $q_1$). (In other words, the additional requirement $q(x_2)=q_2$ does not
1315: bring any additional constraint on $\psi_0(q)$ over $q<q_1$.)
1316: Therefore, we are only sensitive to the behavior of $\psi_0$ to
1317: the right of $q_1$.
1318: For the Brownian initial conditions (\ref{xidef}), the latter
1319: is fully determined by $\{\psi_0(q_1), v_0(q_1)\}$ and the white noise $\xi(q)$
1320: at $q\geq q_1$ (which is statistically homogeneous).
1321: Next, $p(x_{21},q_{21}|x_1,q_1)$ does not depend on $\psi_0(q_1)$ since
1322: a vertical translation of the curves $\psi_0$ and $\cP_{x_1,c_1}$ is fully
1323: absorbed by the same vertical translation of the parabola $\cP_{x_2,c_2}$,
1324: without affecting spatial coordinates $q$ and $x$.
1325: On the other hand, through Galilean invariance the relative
1326: displacements of the particles only depend on their relative velocities, hence
1327: $p(x_{21},q_{21}|x_1,q_1)$ only depends on the relative velocity field
1328: $v_0(q)-v_0(q_1)$
1329: over $q\geq q_1$. For the Brownian initial conditions (\ref{xidef}), with
1330: homogeneous velocity increments, the statistical properties of this relative
1331: velocity field $v_0(q)-v_0(q_1)$ do not depend on $v_0(q_1)$, but only on the
1332: distance $q-q_1$, see (\ref{Dv0}). Therefore, the distributions
1333: $p(x_{21},q_{21}|x_1,q_1)$, of the Lagrangian position increment $q_{21}$, and
1334: $p(x_{21},v_{21}|x_1,v_1)$, of the velocity increment $v_{21}$, only depend
1335: on $x_{21}$, as in (\ref{Pcond}) and in (\ref{pBVnboth}) below.
1336:
1337: In agreement with (\ref{PX1X2mfact}), we can check that this argument fails for
1338: $q_1<0$. Indeed, again we are only sensitive to the behavior of $\psi_0(q)$ to
1339: the right of $q_1$, but this range now includes the special point $q=0$ with
1340: the constraints $\psi_0(0)=0$ and $v_0(0)=0$ that prevent us from absorbing
1341: $\psi_0(q_1)$ and $v_0(q_1)$. For instance, we now have the new constraint
1342: that the first-contact parabola $\cP_{x_2,c_2}$ cannot go upward of the
1343: point $\{0,\psi_0(0)=0\}$ (which was irrelevant in the previous case $q_1>0$,
1344: since we already had $\cP_{x_2,c_2}<\cP_{x_1,c_1}$ over $q<q_1$ and
1345: $\cP_{x_1,c_1}(0)\leq 0$ by construction, being everywhere below $\psi_0$).
1346:
1347: The property that the increments of the inverse Lagrangian map, $q(x_2)-q(x_1)$,
1348: are independent and homogeneous, as in Eq.(\ref{Pcond}), and the probability
1349: distribution (\ref{pBdef}), were already obtained by
1350: \cite{Carraro1998} for intrinsic statistical solutions, and
1351: by \cite{Bertoin1998} through probabilistic tools for $x\geq 0$ in the case of
1352: one-sided Brownian initial conditions (i.e. $v_0(q)=0$ for $q\leq 0$).
1353: The latter work involves
1354: a similar reasoning to the one described above, using the property
1355: that the distribution of a Markov process after last passage at a given point does
1356: not depend on its previous path, but this mathematical proof uses the convex hull
1357: of the Lagrangian potential rather than the parabolas construction used here.
1358: For one-sided
1359: Brownian initial conditions, it is clear that if we consider Eulerian locations
1360: at $x\geq 0$, the particles can only come from the right side $q\geq 0$ so that
1361: we recover the configuration analyzed above for particles that are all located
1362: on the same side of the origin. The agreement with the results of \cite{Bertoin1998}
1363: provides a nice check of our calculations. The probabilistic proof is remarkably
1364: concise, as it first shows that the increments $q_{21}$ are independent and
1365: homogeneous and next derives their distribution. However, the analysis method
1366: presented in the present work has the advantage of a large range of applicability.
1367: Thus, it allowed us to obtain the one-point distributions in closed form in
1368: section~\ref{sec:velocity1} and it could also be applied to different-time
1369: statistics, where the parabolas would have different curvatures.
1370: Another application of the method described in this paper is presented in
1371: \cite{Valageas2008}, where we study ballistic aggregation for one-sided Brownian
1372: initial velocity.
1373:
1374: The previous discussion can be extended to $n-$point distributions, which
1375: thus factorize as
1376: \beqa
1377: X_1\leq .. \leq X_n, \;\; 0 \leq Q_1 \leq .. \leq Q_n : \;\;\;
1378: P_{X_1,..,X_n}(Q_1,..,Q_n) & = & P_{X_1}(Q_1) \bP_{X_2-X_1}(Q_2-Q_1) \nonumber \\
1379: && \hspace{-3cm} \times \bP_{X_3-X_2}(Q_3-Q_2) ... \bP_{X_n-X_{n-1}}(Q_n-Q_{n-1}) .
1380: \label{pBn}
1381: \eeqa
1382: We obtain a similar identity for $Q_1 \leq .. \leq Q_n\leq 0$ by reflection
1383: through the origin, as for Eq.(\ref{PX1X2mfact}).
1384: This also extends to the general case where the Lagrangian coordinates are
1385: located on both sides of the origin as
1386: \beqa
1387: X_m'\leq .. \leq X_1' \leq X_1 \leq .. \leq X_n, \;\;\;
1388: Q_m'\leq .. \leq Q_1' \leq 0 \leq Q_1 \leq .. \leq Q_n & : & \nonumber \\
1389: && \hspace{-11cm} P_{X_i';X_j}(Q_i';Q_j)
1390: = P_{X_1',X_1}(Q_1',Q_1) \prod_{i=2}^m \bP_{X_{i-1,i}'}(Q_{i-1,i}')
1391: \prod_{j=2}^{n} \bP_{X_{j,j-1}}(Q_{j,j-1}) ,
1392: \label{pBnboth}
1393: \eeqa
1394: where we defined relative distances such as $X_{j,j-1}=X_j-X_{j-1}$.
1395: However, it appears that the probability distribution $P_{X_1',X_1}(Q_1',Q_1)$,
1396: with $Q_1' \leq 0 \leq Q_1$, does not greatly simplify and is given by
1397: intricate multiple integrals. Therefore, we shall not consider it further in
1398: this article. Note that for practical purposes one is mostly interested in
1399: the behavior far from the origin, where the invariance through translations
1400: is fully restored.
1401:
1402: In terms of velocities, using the relation (\ref{Vdef}) for the dimensionless
1403: velocities $V_i$, we obtain from the previous results the factorization
1404: \beqa
1405: X_m'\leq .. \leq X_1' \leq X_1 \leq .. \leq X_n,
1406: \;\; V_1' \geq X_1' , \;\; V_{i,i-1}' \geq X_{i,i-1}' ,
1407: \;\; V_1 \leq X_1, \;\; V_{j,j-1} \leq X_{j,j-1} & : & \nonumber \\
1408: && \hspace{-13cm} P_{X_i';X_j}(V_i';V_j)
1409: = P_{X_1',X_1}(V_1',V_1) \prod_{i=2}^m \bP_{X_{i-1,i}'}(V_{i-1,i}')
1410: \prod_{j=2}^{n} \bP_{X_{j,j-1}}(V_{j,j-1}) ,
1411: \label{pBVnboth}
1412: \eeqa
1413: where the various factors are the velocity probabilities that may be obtained
1414: from the Lagrangian $Q-$probability densities through (\ref{Vdef}).
1415: As noticed above, the factorizations (\ref{pBnboth})-(\ref{pBVnboth}) also
1416: follow from the analysis of \cite{Bertoin1998}. However, although this provides
1417: the conditional distribution $\bP_{X_2-X_1}(Q_2-Q_1)$ of the Lagrangian increment
1418: it does not give the distributions $P_{X_i';X_j}(Q_i';Q_j)$ or
1419: $P_{X_1',X_1}(V_1',V_1)$ that appear in these $n$-point distributions.
1420: Nevertheless, in the limit where we are far from the origin, we only need the
1421: one-point distribution $P_{X_1}(Q_1)$, which goes to the Gaussian (\ref{pXinfV2}),
1422: as would also be the case for one-sided initial conditions, besides in that limit
1423: we are mostly interested in the distributions of relative increments.
1424:
1425: We can note that the the Burgers equation with Brownian initial velocity which
1426: we study in this paper was also used in a recent article \cite{Frisch2005}
1427: to discuss the concept of local homogeneity that is used in turbulence studies.
1428: Indeed, for systems which are not strictly homogeneous (the energy shows an
1429: infrared divergence) it is customary to assume incremental homogeneity so that
1430: the physical quantities of interest (e.g. velocity increments) remain homogeneous.
1431: However, as noticed in \cite{Frisch2005} this is not fully consistent because
1432: initial incremental homogeneity is destroyed at later times by the nonlinearity
1433: of the equations of hydrodynamics (the quadratic advective term).
1434: Then, they used numerical simulations and perturbative analysis of the $1$-D
1435: Burgers dynamics with two-sided Brownian initial velocity to illustrate this point
1436: and to note that local homogeneity is only asymptotically recovered far from the
1437: reference point.
1438: The results (\ref{pBnboth}) and (\ref{pBVnboth}) above explicitly show how
1439: the incremental homogeneity is indeed destroyed at finite distance from the
1440: origin but asymptotically recovered at large distances. A peculiarity of this system
1441: is that at finite distance it is already exactly recovered over a partial range
1442: of velocities.
1443: In fact, for the case of one-sided initial conditions ($v_0(q)=0$ for $q\leq 0$)
1444: the system is exactly homogeneous over $x>0$, as shown by the previous discussion
1445: and \cite{Bertoin1998}.
1446:
1447: The factorizations (\ref{pBnboth}) and (\ref{pBVnboth}) also show that small scales
1448: are largely decoupled from long-wavelength modes. Note that this key property is
1449: usually assumed in hydrodynamical systems (so that one can ignore the details of
1450: the large-scale boundary conditions) but is often difficult to prove in a precise
1451: manner.
1452:
1453:
1454:
1455: \subsection{Distribution of Lagrangian increments (i.e. of relative initial
1456: Lagrangian distance)}
1457: \label{relative}
1458:
1459:
1460:
1461: We now study in more details the probability distribution, $\bP_X(Q)$, of
1462: the relative Lagrangian positions (i.e. relative initial distance
1463: $q$ between particles that are separated by distance $x$ at time $t$),
1464: that is, of the increments of the inverse Lagrangian map
1465: (here we omit the subscripts $''21''$ to simplify the notations).
1466: The following results apply far from the origin, or at any location on the
1467: right side of the origin if we have one-sided initial conditions ($v_0(q)=0$ at
1468: $q\leq 0$).
1469:
1470: We can check from the integral representation (\ref{pBdef}) that for
1471: $X\rightarrow 0$ we obtain as expected the Dirac distribution
1472: $\bP_X(Q)\rightarrow\delta(Q)$ (whence $Q_2\rightarrow Q_1$ for
1473: $X_2\rightarrow X_1$). In fact, Eq.(\ref{pBdef}) is a well-known inverse
1474: Laplace transform \cite{Abramowitz} which gives the explicit expression
1475: \beq
1476: X\geq 0, \;\; Q \geq 0 : \;\;\; \bP_X(Q) = \frac{X}{\sqrt{\pi}} \, Q^{-3/2}
1477: \, e^{2X-Q-X^2/Q} =
1478: \frac{X}{\sqrt{\pi}} \, Q^{-3/2} \, e^{-(\sqrt{Q}-X/\sqrt{Q})^2} .
1479: \label{pB1}
1480: \eeq
1481: Therefore, we obtain an exponential tail at large $Q$, as $\sim e^{-Q}$,
1482: and a strong falloff at small $Q$, as $\sim e^{-X^2/Q}$.
1483: For large relative distance $X$ this gives the Gaussian
1484: \beq
1485: X\rightarrow +\infty , \;\;\; |Q-X| \ll X : \;\; \bP_X(Q) \sim
1486: \frac{1}{\sqrt{\pi X}} \, e^{-(Q-X)^2/X} .
1487: \label{PXQinf}
1488: \eeq
1489: This agrees with the expectation that over large distances particles are still
1490: governed by the initial velocity field, as discussed in
1491: section~\ref{subsec:far_away_x} for Eq.(\ref{pXinfV2}).
1492: This is again an illustration of the ``principle of permanence of large eddies''
1493: \cite{Gurbatov1997}, see the discussion below Eq.(\ref{pXinfV2}).
1494:
1495:
1496: We show the probability density $\bP_X(Q)$ obtained for three relative distances
1497: $X$ in Fig.~\ref{figPQ}. We clearly see that for large $X$, which corresponds to
1498: large scales or small times, we recover a Gaussian centered on $X$, whereas for
1499: small $X$ we obtain a skewed distribution with an intermediate power-law regime
1500: $Q^{-3/2}$.
1501:
1502:
1503:
1504: \begin{figure}
1505: \begin{center}
1506: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{PQ.ps}}
1507: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{lPQ.ps}}
1508: \end{center}
1509: \caption{(color online) {\it Left panel:} The probability density $\bp_x(q)$ that two particles,
1510: separated by the distance $x>0$ at time $t$, were initially separated by a
1511: distance $q$ (in the limit where the particles are far from the origin,
1512: or anywhere on the right side for one-sided initial conditions).
1513: This is the distribution of the increments of the inverse Lagrangian map,
1514: $x\mapsto q$.
1515: We show the reduced probabilities, $\bP_X(Q)$, in terms of the dimensionless
1516: variables $X=x/(2Dt^2)$ and $Q=q/(2Dt^2)$, for three values of $X$,
1517: from Eq.(\ref{pB1}). The probability is zero for $Q<0$.
1518: For large relative distance $X$ we recover a Gaussian of center $X$ and variance
1519: $\lag (Q-X)^2\rag=X/2$.
1520: {\it Right panel:} The probability density $\bP_X(Q)$ on a logarithmic scale,
1521: for three values of $X$.}
1522: \label{figPQ}
1523: \end{figure}
1524:
1525:
1526:
1527:
1528: From Eq.(\ref{pB1}) we obtain the moments of the Lagrangian increments
1529: $Q$ as \cite{Gradshteyn}
1530: \beq
1531: \lag Q^n \rag = \frac{2}{\sqrt{\pi}} \, X^{n+1/2} \, e^{2X} \, K_{n-1/2}(2X)
1532: = X^n \sum_{k=0}^{n-1} \frac{(n-1+k)!}{k!(n-1-k)!(4X)^k} ,
1533: \label{Qn}
1534: \eeq
1535: where the last equality only holds for $n\geq 1$, and $K_{\nu}$ is the
1536: modified Bessel function of the second kind.
1537: This gives for the first few moments
1538: \beq
1539: \lag Q \rag = X, \;\;\; \lag Q^2 \rag = X^2+\frac{X}{2} ,
1540: \;\;\; \lag Q^3 \rag = X^3+\frac{3X^2}{2}+\frac{3X}{4} .
1541: \label{Q123}
1542: \eeq
1543: We can note that the mean of the relative displacement, $\PPsi=X-Q$, is zero:
1544: the mean distance between particles does not change (far from the origin).
1545: On the other hand, if we define the usual moment-generating function $\Psi(y)$
1546: by
1547: \beq
1548: \Psi(y)= \sum_{n=0}^{\infty} \frac{(-y)^n}{n!} \lag Q^n\rag
1549: = \int_0^{\infty} \! \dd Q \, e^{-y Q} \, \bP_X(Q) ,
1550: \;\;\; \bP_X(Q)= \inta \frac{\dd y}{2\pi\ii} \, e^{Qy} \, \Psi(y) ,
1551: \label{Psimomentsdef}
1552: \eeq
1553: we obtain from Eq.(\ref{pBdef}), making the change of variable $s=1+y$,
1554: \beq
1555: \Psi(y)= e^{-(\sqrt{1+y}-1)2X} .
1556: \label{Psimoments1}
1557: \eeq
1558: Therefore, the cumulant-generating function $\Phi(y)$, which satisfies
1559: the standard relation
1560: \beq
1561: \Phi(y) = \sum_{n=1}^{\infty} \frac{(-y)^n}{n!} \lag Q^n\rag_c = \ln[\Psi(y)] ,
1562: \label{Phicumdef}
1563: \eeq
1564: is given by
1565: \beq
1566: \Phi(y) = -(\sqrt{1+y}-1)2X = - X y + 2X \sum_{n=2}^{\infty}
1567: \frac{(2n-3)!!}{2^n \, n!} \, (-y)^n .
1568: \label{Phicum1}
1569: \eeq
1570: This yields the simple results
1571: \beq
1572: \lag Q \rag_c= X , \;\;\; \mbox{and for} \;\; n \geq 2 : \;\;\;
1573: \lag Q^n \rag_c= \frac{(2n-3)!!}{2^{n-1}} \, X .
1574: \label{Qcum1}
1575: \eeq
1576:
1577: We can note that the first equality in (\ref{Qn})
1578: also holds for non-integer $n$, and we obtain for small Eulerian distance,
1579: $(x_2-x_1)\rightarrow 0^+$,
1580: \beqa
1581: \nu > \frac{1}{2} & : & \;\; \lag(q_2-q_1)^{\nu}\rag \sim (2Dt^2)^{(\nu-1)}
1582: \, \frac{\Gamma[\nu-\frac{1}{2}]}{\sqrt{\pi}} \, (x_2-x_1) , \label{Qnuscalp} \\
1583: \nu < \frac{1}{2} & : & \;\; \lag(q_2-q_1)^{\nu} \rag \sim (2Dt^2)^{-\nu}
1584: \, \frac{\Gamma[-\nu+\frac{1}{2}]}{\sqrt{\pi}} \, (x_2-x_1)^{2\nu} .
1585: \label{Qnuscalm}
1586: \eeqa
1587: Note that the second scaling also holds for any negative $\nu$. Indeed, the
1588: strong cutoff, $e^{-X^2/Q}$, of the probability distribution (\ref{pB1}),
1589: ensures that all negative moments are finite. Equations
1590: (\ref{Qnuscalp})-(\ref{Qnuscalm})
1591: show that we recover the bifractality of the inverse Lagrangian map,
1592: that was already derived in \cite{Aurell1997} for $\nu\geq 0$.
1593: As is well-known \cite{Frisch2001}, the scaling (\ref{Qnuscalp}) is universal as
1594: it is due to shocks. Indeed, if we have a shock of finite Lagrangian increment
1595: $\delta q$ at position $x$, it gives a contribution
1596: $[q(x+\ell/2)-q(x-\ell/2)]^n \sim (\delta q)^n$ which remains of order unity for
1597: $\ell \rightarrow 0^+$ for any $n$. Next, the probability to have a shock of a
1598: given finite strength $\delta q$ in a small Eulerian interval $\ell$ scales as
1599: $\ell$ at small distances, which gives rise to the factor $(x_2-x_1)$ in
1600: Eq.(\ref{Qnuscalp}). Note that in our case, the total number of shocks per
1601: unit length is actually infinite \cite{She1992,Sinai1992}, see
1602: sect.~\ref{subsec:shockmass} below, as the shock mass function (\ref{Nshock2})
1603: leads to a divergence at low mass, but the number of shocks above a finite mass
1604: threshold is finite and this is sufficient to make the scaling (\ref{Qnuscalp})
1605: universal.
1606: However, the behavior observed at $\nu<1$ (the critical value $\nu_c=1/2$ and the
1607: exponent $2\nu$ observed below $\nu_c$ in Eq.(\ref{Qnuscalm})) depends on the
1608: initial energy spectrum, through the low-mass tail of the shock mass function, see
1609: also \cite{Aurell1997} for more detailed discussions.
1610:
1611:
1612: \subsection{Distribution of Eulerian velocity increments}
1613: \label{relative_velocities}
1614:
1615:
1616: We now consider the probability distribution, $\bP_X(V)$, of the relative Eulerian
1617: velocities, that is of the velocity increments $V(X_2)-V(X_1)$.
1618: From Eq.(\ref{pB1}) we obtain
1619: \beq
1620: X\geq 0, \;\; V\leq X : \;\;\; \bP_X(V) = \frac{X}{\sqrt{\pi}} \, (X-V)^{-3/2}
1621: \, e^{-(\sqrt{X-V}-X/\sqrt{X-V})^2} .
1622: \label{pBV1}
1623: \eeq
1624: In the limit of large relative Eulerian distance $X\rightarrow \infty$,
1625: at fixed $V$, the distribution (\ref{pBV1}) can be expanded around the
1626: maximum of the exponent at $V=0$ (corresponding to $Q=X$) and we again recover
1627: the initial Gaussian
1628: \beq
1629: |V|\ll X : \;\; \bP_X(V) \sim \frac{1}{\sqrt{\pi X}} \, e^{-V^2/X} ,
1630: \label{PVinf}
1631: \eeq
1632: in agreement with the fact that over large distances particles are still
1633: governed by the initial velocity field (see also section~\ref{subsec:far_away_x}).
1634: We show in Fig.~\ref{figPV} the velocity distribution $\bP_X(V)$ for three values
1635: of $X$. We can again check that we recover a Gaussian for large $X$ (i.e. large
1636: scales or small times), whereas for smaller $X$ the upper bound $V\leq X$ is
1637: increasingly apparent while a power law develops at intermediate negative velocities.
1638:
1639:
1640:
1641: \begin{figure}
1642: \begin{center}
1643: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{PV.ps}}
1644: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{lPV.ps}}
1645: \end{center}
1646: \caption{(color online) {\it Left panel:} The probability density $\bp_x(v)$ of the
1647: velocity increment $v=v(x_2)-v(x_1)$ for two positions separated by the distance
1648: $x=x_2-x_1$ (in the limit where we are far from the origin,
1649: or anywhere on the right side for one-sided initial conditions).
1650: We show the reduced probabilities, $\bP_X(V)$ in terms of the dimensionless variables
1651: $X=x/(2Dt^2)$ and $V=v/(2Dt)$, for three values of $X$, from Eq.(\ref{pBV1}).
1652: The probability is zero for $V>X$.
1653: At large relative distance $X$ we recover a symmetric Gaussian of variance
1654: $\lag V^2\rag=X/2$.
1655: {\it Right panel:} The probability density $\bP_X(V)$ on a semi-logarithmic scale,
1656: for three values of $X$.}
1657: \label{figPV}
1658: \end{figure}
1659:
1660:
1661:
1662:
1663: From Eq.(\ref{Qcum1}) the velocity cumulants are given by
1664: \beq
1665: n \geq 2 : \;\; \lag V^n\rag_c = (-1)^n \frac{(2n-3)!!}{2^{n-1}} \, X ,
1666: \;\; \mbox{whence} \;\;
1667: \lag V\rag=0 , \;\; \lag V^2\rag= \frac{X}{2} , \;\;
1668: \lag V^3\rag = - \frac{3X}{4} .
1669: \label{V123}
1670: \eeq
1671: We can note that the first moment exactly vanishes whereas the variance
1672: $\lag V^2\rag$ remains equal to that of the initial Gaussian field,
1673: see (\ref{Dv0}), even though $\bP_X(V)$ is no longer Gaussian. Thus, in terms
1674: of the dimensional variables the velocity energy spectrum remains equal to the
1675: initial one,
1676: \beq
1677: \lag [v(x_2,t)-v(x_1,t)]^2 \rag = D |x_2-x_1| , \;\;\;\;
1678: E(k,t)= E_0(k)= \frac{D}{2\pi} k^{-2} .
1679: \label{Ekt}
1680: \eeq
1681: Finally, in the limit of small separations we obtain from Eq.(\ref{V123})
1682: \beq
1683: n\geq 2 , \;\; (x_2-x_1)\rightarrow 0^+ : \;\;\; \lag (v_2-v_1)^n\rag \sim \frac{(2Dt^2)^{n-1}}{t^n} \, (-1)^n \frac{(2n-3)!!}{2^{n-1}} (x_2-x_1) .
1684: \label{VnX0}
1685: \eeq
1686: Therefore, we recover the universal scaling at small distances
1687: of the structure functions \cite{Frisch2001},
1688: $\lag [v(x+\ell)-v(x)]^n\rag \propto \ell$, that was also observed in the numerical
1689: simulations of \cite{She1992}. This is due to the contribution from shocks, as
1690: discussed below Eqs.(\ref{Qnuscalp})-(\ref{Qnuscalm}).
1691: Thus, if we have a shock of finite velocity jump $-\delta v=\delta q/t$ at location
1692: $x$, then $[v(x+\ell/2)-v(x-\ell/2)]^n \sim (-\delta v)^n$ for $\ell \rightarrow 0^+$.
1693: Note that $\delta v$ is positive, since a shock is associated with particles from
1694: the left overtaking particles on the right, so that $v(x^-)>v(x^+)$, which agrees
1695: with the factor $(-1)^n$ in Eq.(\ref{VnX0}). Again, the factor $(x_2-x_1)$ in
1696: Eq.(\ref{VnX0}) comes from the probability to encounter a shock of strength larger
1697: than some finite threshold $\delta q$ in a small Eulerian interval $[x_1,x_2]$.
1698:
1699:
1700:
1701: \section{Density field}
1702: \label{sec:Density}
1703:
1704:
1705: \subsection{Overdensity within finite size domains}
1706: \label{Overdensity}
1707:
1708:
1709: We consider here the evolution of a density field $\rho(x,t)$ that evolves through
1710: the usual continuity equation,
1711: \beq
1712: \frac{\pl \rho}{\pl t} + \frac{\pl}{\pl x} (\rho v) = 0 ,
1713: \label{continuity}
1714: \eeq
1715: whereas the velocity field $v(x,t)$ evolves through the Burgers equation
1716: (\ref{Burg}). The initial conditions for the velocity are set by Eqs.(\ref{xidef})
1717: as in previous sections, whereas the initial density is a constant $\rho_0$.
1718: Thus, the mass $m$ between particles $q_1$ and $q_2$, with $q_1<q_2$, is
1719: $m=\rho_0(q_2-q_1)$. This quantity is conserved by the dynamics since particles
1720: do not cross each other (though it is ambiguous at shock locations, but the latter
1721: have zero measure in Eulerian space).
1722: Then, the overall overdensity, $\eta=m/(\rho_0 x)$, over the length $x=x_2-x_1$,
1723: is $\eta=(q_2-q_1)/(x_2-x_1)$ by conservation of matter, where $q_i$ is the
1724: initial Lagrangian position of the particle that is located at $x_i$ at time $t$.
1725: In terms of dimensionless variables this reads as the ratio of relative distances
1726: $\eta=Q/X$. Therefore, far from the origin ($|x_1| \rightarrow \infty$),
1727: or on the right side of the origin for one-sided initial conditions,
1728: we obtain from Eq.(\ref{pB1}) the probability distribution of the overdensity at
1729: scale $X$ as
1730: \beq
1731: \eta = \frac{m}{\rho_0 x} , \;\; \eta \geq 0 : \;\;\;
1732: P_X(\eta) = \sqrt{\frac{X}{\pi}} \, \eta^{-3/2} \,
1733: e^{-X(\sqrt{\eta}-1/\sqrt{\eta})^2} = \sqrt{\frac{X}{\pi}} \, e^{2X} \,
1734: \eta^{-3/2} \, e^{-X(\eta+1/\eta)} .
1735: \label{Peta}
1736: \eeq
1737: Over large scales we recover a Gaussian distribution, as for the variables $Q$
1738: and $V$, while on small scales, $X\rightarrow 0$, we obtain the power law
1739: $\eta^{-3/2}$ between the low and high density cutoffs at
1740: $\eta_- \sim x/(2Dt^2)$ and $\eta_+ \sim (2Dt^2)/x$, as we can see in
1741: Fig.~\ref{figPeta} where we show the overdensity distribution $P_X(\eta)$
1742: for three values of $X$. We can note that this is very similar to the behavior
1743: that is observed in cosmological numerical simulations for gravitational clustering
1744: \cite{Balian1989,Colombi1994,ValMun2004}.
1745:
1746:
1747:
1748: \begin{figure}
1749: \begin{center}
1750: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{Peta.ps}}
1751: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{lPeta.ps}}
1752: \end{center}
1753: \caption{(color online) {\it Left panel:} The probability distribution $p_x(\eta)$ of the
1754: overdensity, $\eta=m/(\rho_0 x)$, over a region of length $x$.
1755: We show $P_X(\eta)$ for three values of the reduced length $X=x/(2Dt^2)$, from
1756: Eq.(\ref{Peta}).
1757: Thus larger $X$ corresponds to larger scale or smaller time. For large
1758: $X$ we recover a Gaussian of mean $1$ and variance $\lag(\eta-1)^2\rag=1/(2X)$.
1759: For small $X$ the distribution becomes skewed and an intermediate power-law
1760: region develops.
1761: {\it Right panel:} The probability density $P_x(\eta)$ on a logarithmic scale,
1762: for three values of $X$.}
1763: \label{figPeta}
1764: \end{figure}
1765:
1766:
1767:
1768: From Eqs.(\ref{Qn}) and (\ref{Qcum1}) we obtain for the moments and the cumulants
1769: of the overdensity at scale $X$:
1770: \beq
1771: n\geq 1 : \;\; \lag \eta^n \rag = \sum_{k=0}^{n-1}
1772: \frac{(n-1+k)!}{k!(n-1-k)!(4X)^k} , \;\;\; \mbox{and for} \; n\geq 2 : \;\;
1773: \lag \eta^n \rag_c= \frac{(2n-3)!!}{(2X)^{n-1}} ,
1774: \label{etan}
1775: \eeq
1776: whence for the lowest orders
1777: \beq
1778: \lag \eta \rag=1, \;\;\; \lag \eta^2 \rag_c= \frac{1}{2X} , \;\;\;
1779: \lag \eta^3 \rag_c= \frac{3}{4X^2} = 3 \lag \eta^2 \rag_c^2 .
1780: \label{eta123}
1781: \eeq
1782: We can note that the second result (\ref{etan}) gives the cumulant hierarchy
1783: \beq
1784: S_n = \frac{\lag \eta^n \rag_c}{\lag \eta^2 \rag_c^{n-1}}
1785: = (2n-3)!! \;\;\; \mbox{and} \;\;\;
1786: \varphi(y) = \sum_{n=1}^{\infty} (-1)^{n-1} \, S_n \,\frac{y^n}{n!} = \sqrt{1+2y}-1 ,
1787: \label{Sndef}
1788: \eeq
1789: which shows that the ratios $S_n$ are constants that do not depend on time nor
1790: scale. We can note that in the cosmological context, associated with
1791: a gravitational dynamics in a 3-dimensional expanding Universe, for the case of
1792: an initial power-law density power spectrum, the coefficients $S_n$, still
1793: defined as in (\ref{Sndef}), only show a weak dependence on scale in the
1794: highly nonlinear regime, and they also asymptotically reach (different) finite
1795: values at large scales in the quasi-linear regime
1796: \cite{Colombi1994,Bernardeau2002}. Then, it has been proposed
1797: to use the approximation of constant $S_n$ to describe the highly nonlinear
1798: regime \cite{Peebles1980}. Moreover, the form (\ref{Sndef}) of the
1799: reduced cumulant generating function $\varphi(y)$ is one of the possibilities
1800: that have been studied in this context \cite{Balian1989}.
1801: This phenomenological ansatz is known
1802: as the ``stable clustering model'' as it was derived by assuming that on small
1803: physical scales, after nonlinear collapse and gravitational relaxation,
1804: overdensities decouple from the Hubble expansion and keep a constant physical size
1805: \cite{Davis1977}. In the present case, collapsed objects are actually Dirac
1806: peaks (shocks) of vanishing size. Then, it is easy to see from a multifractal
1807: analysis that shocks lead to finite ratios $S_n$ in the small-scale limit
1808: \cite{Balian1989b,Valageas1999}, so that the hierarchy (\ref{Sndef}) is universal,
1809: in the sense that the generating function $\varphi(y)$ has a finite limit at small
1810: scale, $x\rightarrow 0$. However, this non-trivial
1811: limit depends on the initial energy spectrum.
1812: A specific property of the Brownian initial
1813: conditions studied in this paper is that the ratios $S_n$ are actually constant
1814: over all scales, from the linear to highly nonlinear scales.
1815: Thus, it is interesting to note that the $1$-D Burgers
1816: equation with Brownian initial velocity provides an exact physical realization
1817: of this ansatz.
1818:
1819:
1820: \subsection{Density correlations}
1821: \label{Density-correlations}
1822:
1823:
1824: We now consider the unsmoothed density field $\rho(x)$ itself (again in the limit
1825: where we are far from the origin so that the previous results apply).
1826: It is related to the smoothed overdensity $\eta$ over scale $x$ introduced
1827: above through
1828: \beq
1829: \eta = \int_{x_1}^{x_1+x} \frac{\dd x'}{x} \; \frac{\rho(x')}{\rho_0} .
1830: \label{etax}
1831: \eeq
1832: Then, introducing the density power spectrum, $\Pk(k)$, by going to Fourier space as
1833: \beq
1834: \rho(x) - \rho_0 = \int_{-\infty}^{\infty} \dd k \, e^{\ii kx} \, \rho(k) ,
1835: \;\;\; \lag \rho(k_1) \rho(k_2) \rag = \delta(k_1+k_2) \, \rho_0^2 \, \Pk(k_1) ,
1836: \label{Pkdef}
1837: \eeq
1838: we obtain using the second Eq.(\ref{eta123}) and Eqs.(\ref{etax})-(\ref{Pkdef})
1839: \beq
1840: \frac{Dt^2}{x} = \lag\eta^2\rag_c = \int_{-\infty}^{\infty} \dd k \,
1841: \sinc^2(\frac{kx}{2}) \,\Pk(k) , \;\;\;\; \mbox{and} \;\;\;\;
1842: \Pk(k,t) = \frac{Dt^2}{2\pi} ,
1843: \label{Pkt}
1844: \eeq
1845: where $\sinc(x)=\sin(x)/x$ is the cardinal sine. Thus, we obtain a white-noise
1846: density power spectrum, with an amplitude that grows as $t^2$. This yields the
1847: connected density two-point correlation
1848: \beq
1849: \lag \rho(x_1,t) \rho(x_2,t) \rag_c = \rho_0^2 \, C_2(x_1,x_2) , \;\;\;\;
1850: \mbox{with} \;\;\;\; C_2(x_1,x_2) = Dt^2 \, \delta(x_2-x_1) ,
1851: \label{Ceta}
1852: \eeq
1853: which remains a Dirac function at all times.
1854:
1855: In fact, the factorization (\ref{pBn}) implies a similar factorization for the
1856: multivariate distributions of the smoothed density field, far from the origin
1857: ($X_1\rightarrow +\infty$),
1858: \beq
1859: X_1\leq .. \leq X_n \;\; : \;\;
1860: P_{X_{2,1};..;X_{n,n-1}}(\eta_{2,1};..;\eta_{n,n-1})
1861: = P_{X_{2,1}}(\eta_{2,1}) .. P_{X_{n,n-1}}(\eta_{n,n-1}) ,
1862: \label{Peta1n}
1863: \eeq
1864: where $\eta_{i,i-1}$ is the mean overdensity over the interval $[X_{i-1},X_i]$.
1865: Thus, the densities within non-overlapping domains are completely independent
1866: random variables. This agrees with the Dirac obtained in Eq.(\ref{Ceta})
1867: for the connected density two-point correlation.
1868: Moreover, this can be extended to all higher orders.
1869: Indeed, let us consider the density $n-$point connected correlation, defined as
1870: \beq
1871: \lag \rho(x_1) .. \rho(x_n) \rag_c = \rho_0^n \, C_n(x_1,..,x_n) .
1872: \label{Cndef}
1873: \eeq
1874: If there exists a position $x_i$ that is different from all other positions
1875: $x_j$, with $j\neq i$, then we can build a small region
1876: $[x_i-\epsilon,x_i+\epsilon]$ where the density is independent from the density
1877: at all other points $x_j$, using the property (\ref{Peta1n}) and
1878: $\epsilon\rightarrow 0^+$. Therefore, by definition of connected correlations,
1879: $C_n$ must vanish. Then, the $n-$point connected correlation
1880: can be written as the product of $n-1$ Dirac factors
1881: \beqa
1882: C_n(x_1,..,x_n) & = & (2n-3)!! \, (Dt^2)^{n-1} \,
1883: \delta(x_2-x_1) \delta(x_3-x_1) ... \delta(x_n-x_1) \label{CnDirac} \\
1884: & = & (2n-3)!! \, C_2(x_1,x_2) C_2(x_1,x_3) ... C_2(x_1,x_n) ,
1885: \label{Cnfact}
1886: \eeqa
1887: where the amplitude is obtained from Eq.(\ref{etan}), as well as Eq.(\ref{etax})
1888: which implies the relation
1889: $\lag\eta^n\rag_c = x^{-n} \int_0^x \dd x_1 .. \dd x_n C_n(x_1,..,x_n)$.
1890:
1891:
1892:
1893: \begin{figure}
1894: \begin{center}
1895: \epsfxsize=13 cm \epsfysize=2.5 cm {\epsfbox{tree1.eps}}
1896: \end{center}
1897: \caption{The 15 heap ordered trees that can be associated with the $4-$point
1898: correlation $C_4(x_1,..,x_4)$. The labels refer to the positions $x_i$, which are
1899: ordered as $x_1\leq x_2 \leq x_3 \leq x_4$. We only show the 5 tree structures,
1900: as the additional terms can be obtained from the previous diagram by permutations
1901: over the labels that satisfy the ordering constraint that each path from the root
1902: has increasing labels as we proceed down to the leaves. Each link between nodes
1903: $i$ and $j$ yields a contribution $C_2(x_i,x_j)$ and the contribution of a tree
1904: is the product of the 3 factors $C_2$ associated with its 3 links.
1905: This gives $C_4$ as the sum over all these tree contributions.}
1906: \label{figtree}
1907: \end{figure}
1908:
1909:
1910:
1911: We can note that $(2n-3)!!$ also counts the number of heap ordered trees with
1912: $n$ nodes, that is, rooted trees where the $n$ nodes are labelled as
1913: $\{1,2,..,n\}$ and each path from the root has increasing labels
1914: \cite{Prodinger1983}. In our case, we can therefore construct the following
1915: combinatorial interpretation of Eq.(\ref{CnDirac}). First, the points $x_1,..,x_n$,
1916: are ordered such that $x_1\leq x_2 \leq .. \leq x_n$ (we choose
1917: one among several possibilities if several positions are equal). Then,
1918: the $n-$point connected correlation (\ref{CnDirac}) is obtained as the sum of the
1919: contributions of all heap ordered trees, where the contribution of each tree
1920: is simply the product of the $n-1$ factors $C_2(x_i,x_j)$ associated with the
1921: $n-1$ links between nodes $x_i$ and $x_j$.
1922:
1923: Of course, we may also write (\ref{CnDirac}) as the sum over
1924: the products of $C_2(x_i,x_j)$ associated with any other class of $N$ trees,
1925: multiplied by a weight $(2n-3)!!/N$.
1926: However, this no longer recovers the amplitude $(2n-3)!!$ of Eq.(\ref{CnDirac})
1927: in a natural manner. We note that in the cosmological context,
1928: within the stable-clustering ansatz discussed above, it has been proposed to
1929: use as a phenomenological model a diagrammatic description such as
1930: Fig.~\ref{figtree}, where the connected $n-$point density correlation is written
1931: as the sum over trees of each product of $(n-1)$ factors $C_2(x_i,x_j)$
1932: associated with the internal links \cite{Fry1984,Schaeffer1985}.
1933: However, the tree diagrams used in this context are usually
1934: not ordered, and each topology may have an additional multiplicative weight.
1935: In the present case of the $1$-dimensional Burgers dynamics, we can note that
1936: the concept of ordering naturally arises since particles do not cross each other
1937: and one can order both Lagrangian and Eulerian positions on the line
1938: (this would no longer be the case for higher dimensions).
1939:
1940: Thus, the $1$-D Burgers dynamics with Brownian initial velocity provides
1941: a physical realization of the hierarchical structure such as (\ref{Cnfact}) for
1942: the many-body correlation functions. It is an interesting question to ask whether
1943: other real dynamical systems can be built that display the same factorization
1944: property (possibly over other classes of trees) with other two-point correlations
1945: $C_2$.\footnote{In fact, as for the weaker property of constant ratios $S_n$,
1946: the author is not aware of other dynamical systems that
1947: exactly obey such a factorization property. In view of the many phenomenological
1948: works that have used such a diagrammatic construction for many-body correlations,
1949: it is satisfying to find that it is at least obeyed by one truly dynamical system,
1950: even though the expression in terms of Dirac factors and the lack of large
1951: distance correlation make this a very simple and specific case.}
1952: On the other hand, one may wonder whether a factorization such as (\ref{Cnfact}),
1953: and a diagrammatic construction such as Fig.~\ref{figtree}, could be generalized,
1954: as an exact asymptotic solution or as a useful phenomenological model,
1955: to the $1$-dimensional Burgers dynamics with other initial conditions, where
1956: $C_2$ would no longer be the simple Dirac function (\ref{Ceta}).\footnote{The results
1957: of \cite{Bertoin1998} show that the Brownian case can be generalized to
1958: Levy processes with no positive jumps, where the increments of the inverse Lagrangian
1959: map again remain homogeneous and independent at all times. However, this does not
1960: provide another hierarchy for the many-body correlations as they remain of the
1961: Dirac type.}
1962:
1963:
1964: \subsection{Comparison with a perturbative approach}
1965: \label{perturbative-approach}
1966:
1967:
1968: We can note that the exact (far from the origin) nonlinear results
1969: (\ref{Pkt})-(\ref{Ceta}) are identical to the perturbative predictions
1970: that would be obtained at linear order from Eq.(\ref{continuity}).
1971: Indeed, if we linearize the continuity and inviscid Burgers equations, we obtain
1972: at lowest order for the density field $\pl\rho_L/\pl t= - \rho_0\pl v_L/\pl x
1973: = - \rho_0 \pl v_0/\pl x = - \rho_0 \xi$, where $\xi$ is the initial white-noise
1974: of Eq.(\ref{xidef}).
1975: This gives $\rho_L(x,t)=\rho_0 (1-t\xi(x))$, which leads in turn to
1976: Eqs.(\ref{Pkt})-(\ref{Ceta}). The fact that for these Brownian initial
1977: conditions the nonlinear Burgers
1978: dynamics preserves the linear density power spectrum is reminiscent of the
1979: invariance of the energy velocity spectrum (\ref{Ekt}). In both cases, one needs to
1980: consider higher-order correlations (or the full distribution) to measure the
1981: effects of the nonlinearities.
1982:
1983: In fact, the agreement of the exact density two-point function with the linear theory
1984: actually extends to all order cumulants $\lag\eta^n\rag_c$, computed at leading order
1985: from quasi-linear theory. Indeed, at tree-order in perturbation theory, in the
1986: inviscid limit, it can be shown that the cumulant-generating function $\varphi(y)$,
1987: defined as in Eq.(\ref{Sndef}), is given by the implicit system
1988: \beq
1989: \left\{ \bea{l} \tau= - y \, \cG'(\tau) \\ \\
1990: \varphi(y) = y \, \cG(\tau)+\frac{\tau^2}{2} \ea \right. \;\;\;
1991: \mbox{with} \;\;\; \cG(\tau) = \cF\left[-\tau \frac{\sigma(\cG \, x)}{\sigma(x)}
1992: \right] = \cF\left[-\tau \, \cG^{-1/2} \right] ,
1993: \label{varphiyGtau}
1994: \eeq
1995: where $\sigma(x)^2=\lag \delta_L^2\rag=Dt^2/x$ is the variance of the linear density
1996: contrast $\delta_L$ at scale $x$, and the function $\cF(\delta_L)$ describes
1997: the evolution of spherical (here symmetric) density fluctuations
1998: (see \cite{Bernardeau1994,Bernardeau2002,Valageas2002}
1999: for the similar case of the cosmological gravitational dynamics).
2000: The system $\{\tau,\cG\}\leftrightarrow\{y,\varphi\}$ in (\ref{varphiyGtau}) is
2001: actually a Legendre transform and it arises from a saddle-point approximation.
2002: Indeed, in the quasi-linear limit (i.e. $\sigma\rightarrow 0$, which also corresponds
2003: to $t\rightarrow 0$ or $x\rightarrow\infty$)
2004: the cumulant ratios $S_n$ are governed by the tails of the density distribution
2005: and the generating function $\varphi(y)$ can be obtained from a steepest-descent
2006: method\footnote{In fact, the steepest-descent method described in \cite{Valageas2002}
2007: is a non-perturbative approach, which can also be applied to the other limit of rare
2008: events at finite $\sigma$, where it allows to go beyond perturbative methods
2009: \cite{Valageas2002IV}. However, in the quasi-linear limit $\sigma\rightarrow 0$
2010: it gives the same results for $\varphi(y)$ as the usual perturbative expansion
2011: over powers of the linear growing mode of the density field (provided the latter
2012: gives finite results in this limit).}
2013: \cite{Valageas2002}. (In a somewhat similar fashion, the minimization
2014: problem (\ref{psinu0}), that also arises from a saddle-point method, can be written
2015: in terms of a Legendre transform of the Lagrangian potential, see \cite{Bec2007}.)
2016: As compared with Eq.(69) of \cite{Valageas2002} we made the change
2017: $\sigma[(1+\cG)^{1/3}x] \rightarrow \sigma(\cG \, x)$ by taking $1+\cG\rightarrow\cG$
2018: and the exponent $1/3$ is changed to unity as we go from $3$-D to $1$-D.
2019: For the present $1$-D Burgers dynamics, we have:
2020: \beq
2021: \eta=\frac{q}{x} = \frac{q}{q+tv} , \;\;\;\; \mbox{whence at linear order} \;\;\;\;
2022: \eta_L=1-\frac{t v}{q} \;\;\; \mbox{and} \;\;\; \delta_L = - t \frac{v}{q} .
2023: \label{etaL}
2024: \eeq
2025: This yields
2026: \beq
2027: \cF(\delta_L) = \frac{1}{1-\delta_L} , \;\;\; \mbox{whence} \;\;\;
2028: \cG(\tau)= \frac{(-\tau+\sqrt{\tau^2+4})^2}{4} \;\;\; \mbox{and} \;\;\;
2029: \varphi(y) = \sqrt{1+2y} -1 .
2030: \label{FGvarphi}
2031: \eeq
2032: Thus, we recover at tree-order the exact result (\ref{Sndef}). Therefore,
2033: for Brownian initial velocity the Burgers dynamics happens to preserve the
2034: density cumulant-generating function $\varphi(y)$ that would be obtained at leading
2035: order (tree-order) from a perturbative approach, which does not take into account
2036: collisions between particles (shell-crossings). This is also why the ratios $S_n$
2037: are constants that apply to all scales, from the quasi-linear to the highly nonlinear
2038: scales. Note that the perturbative approach breaks down beyond leading order
2039: as next-to-leading corrections actually involve divergent integrals (which means
2040: that one can no longer discard shocks, which requires non-perturbative methods).
2041: For other initial conditions the coefficients $S_n$ would no longer remain equal
2042: to their tree-order values. However, they still asymptote to finite values
2043: in the highly nonlinear regime, because of the contribution from shocks, just
2044: as the Lagrangian and velocity increments scale linearly with $\ell$ for small
2045: distances $\ell \rightarrow 0$, as discussed below
2046: Eqs.(\ref{Qnuscalp})-(\ref{Qnuscalm}) and Eq.(\ref{VnX0}).
2047:
2048:
2049:
2050: \section{Lagrangian displacement field}
2051: \label{sec:Displacement}
2052:
2053:
2054: \subsection{One-point distributions}
2055: \label{subsec:One-point}
2056:
2057:
2058:
2059: We now consider the dynamics associated with the Burgers equation (\ref{Burg})
2060: from a Lagrangian point of view. That is, labelling particles by their initial
2061: position $q$ at time $t=0$, we follow their trajectory $x(q,t)$ and we note
2062: $\PPsi(q,t)=x(q,t)-q$ their displacement with respect to their initial location.
2063: Note that for regular points, which have kept their initial velocity, we have
2064: $\PPsi= t v$, see Eq.(\ref{psinu}). Since particles do not cross each other
2065: they remain well-ordered. Then, it is clear that the probability, $p_q(x'\geq x)$,
2066: for the particle $q$ to be to the right of the Eulerian position $x$, at time $t$,
2067: is equal to the probability, $p_x(q'\leq q)$, for the Eulerian location $x$ to
2068: be ``occupied'' by particles that were initially to the left of particle $q$.
2069: (Since shocks have zero measure in Eulerian space there are no ambiguities.)
2070: Therefore, we obtain in terms of dimensionless variables, for the case $q\geq 0$,
2071: \beqa
2072: Q\geq 0 : \;\;\; P_Q(X'\geq X) & = & P_X(Q'\leq Q) =
2073: P_X(Q'\leq 0)+P_X(0\leq Q'\leq Q) \nonumber \\
2074: & = & J(1,-2X) + \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q} \,
2075: \frac{J(s,2X)}{s-1} ,
2076: \label{PQX1}
2077: \eeqa
2078: where we used the results of section~\ref{subsec:arbitrary_Eulerian_location_x}
2079: and the integration contour runs to the right of the pole $s=1$.
2080: Therefore, the probability density of the Eulerian position $X$ of particle $Q$
2081: reads as
2082: \beqa
2083: Q\geq 0 : \;\;\; P_Q(X) & = & - \frac{\pl}{\pl X} P_Q(X'\geq X) \nonumber \\
2084: && = - \frac{\pl}{\pl X} \left[ J(1,-2X) + \inta \frac{\dd s}{2\pi\ii} \,
2085: e^{(s-1)Q} \, \frac{J(s,2X)}{s-1} \right] .
2086: \label{PQX2}
2087: \eeqa
2088: The case $Q<0$ can be obtained from Eq(\ref{PQX2}) through a reflection about the
2089: origin.
2090:
2091:
2092:
2093: \begin{figure}
2094: \begin{center}
2095: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{P0x.ps}}
2096: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{lP0x.ps}}
2097: \end{center}
2098: \caption{(color online) {\it Left panel:} The probability distribution, $P_0(X)$, of the
2099: reduced position $X=x/(2Dt)$ of the particle that was initially at the origin,
2100: $q=0$, from Eq.(\ref{PQ0X1}). The dashed lines show the asymptotic behaviors
2101: (\ref{PQ0X0}) and (\ref{PQ0Xinf}).
2102: {\it Right panel:} Same as left panel but on a logarithmic scale.}
2103: \label{figP0x}
2104: \end{figure}
2105:
2106:
2107:
2108: Let us consider more precisely the case of the particle that was initially at
2109: rest at the origin, $q=0$. Following the previous discussion, we have
2110: $p_0(x'\geq x)=p_x(q'\leq 0)$, hence we can directly use
2111: Eqs.(\ref{pqp1})-(\ref{pqpm2}) which give
2112: \beq
2113: X \geq 0 : \;\; P_0(X) = - \frac{\pl}{\pl X} \int_0^{\infty} \dd\nu \,
2114: \frac{6}{\nu^2} \, \Ai\left[\nu 2X+\frac{1}{\nu^2}\right]^2 ,
2115: \;\;\; P_0(-X) = P_0(X) .
2116: \label{PQ0X1}
2117: \eeq
2118: At small $X$, we obtain from Eq.(\ref{PQ0X1}) the asymptotic
2119: \beq
2120: X \rightarrow 0^+ : \;\; P_0(X) \sim - \frac{4\sqrt{3}}{\pi} \ln X ,
2121: \label{PQ0X0}
2122: \eeq
2123: whereas the behavior for large displacement $X$ is set by the asymptotic
2124: (\ref{pqmxinf}), which yields
2125: \beq
2126: X \rightarrow + \infty : \;\; P_0(X) \sim
2127: \left(\pi X /(2\sqrt{3})\right)^{-1/2} \, e^{-4\sqrt{3}X} .
2128: \label{PQ0Xinf}
2129: \eeq
2130: Thus, the central particle $q=0$, that was initially at rest ($v_0(0)=0$),
2131: has moved by time $t$ by a distance $\PPsi=X$ whose distribution shows an
2132: exponential tail at large $|X|$ and a logarithmic peak at low $|X|$.
2133: We can note that both the large-$X$ tail and the low-$X$ divergence
2134: are different from the asymptotics of the distribution of the
2135: Lagrangian coordinate, $Q=-V$, of the particle located at the Eulerian location
2136: $X=0$ at the same time, see Eqs.(\ref{p0V0})-(\ref{p0Vinfty}).
2137: We show in Fig.~\ref{figP0x} the distribution $P_0(X)$ as well as the asymptotic
2138: behaviors (\ref{PQ0X0}) and (\ref{PQ0Xinf}). It appears that the logarithmic
2139: asymptote is only reached at very low $X$.
2140:
2141:
2142:
2143: Finally, far away from the origin, in the limit $Q\rightarrow\infty$ at
2144: fixed $\PPsi=X-Q$, we obtain from Eq.(\ref{PQX2}) the asymptotic behavior
2145: (making the change of variable $s=1+\ii k$ as for Eq.(\ref{pXinfV2}))
2146: \beq
2147: Q\rightarrow +\infty, \;\; |\PPsi|\ll Q : \;\; P_Q(\PPsi) \sim
2148: \frac{e^{-\PPsi^2/Q}}{\sqrt{\pi Q}} .
2149: \label{pQinPsi1}
2150: \eeq
2151: Therefore, we recover the property that far from the origin the displacement
2152: of the particles is governed at leading order by the Gaussian distribution of the
2153: initial velocity field $v_0(q)$. This agrees with the discussion presented
2154: below Eq.(\ref{pXinfV2}) in section~\ref{subsec:far_away_x}.
2155: Again, this expresses the fact that at very large distances, where the initial
2156: velocity becomes increasingly large as $\sqrt{|q|}$, the motion with respect to
2157: the origin is dominated by the ``large-scale flow'' and the local fluctuations
2158: of the initial velocity field have only produced local subdominant shifts,
2159: see also \cite{Gurbatov1999,Aurell1993}.
2160:
2161:
2162:
2163: \subsection{Two-point distributions}
2164: \label{subsec:Two-point}
2165:
2166:
2167: We now investigate the two-point probability distribution of the Lagrangian
2168: displacement field. As for the one-point distribution (\ref{PQX1}), it is
2169: related to its Eulerian counterpart through
2170: \beq
2171: P_{Q_1,Q_2}(X_1'\geq X_1,X_2'\geq X_2) = P_{X_1,X_2}(Q_1'\leq Q_1,Q_2'\leq Q_2)
2172: \label{PQ1Q2PX1X2}
2173: \eeq
2174: Using Eqs.(\ref{pXinfV2}), (\ref{pBdef}), we obtain far away from the origin,
2175: in the limit $Q_1\rightarrow+\infty$ at fixed $Q_{21}=Q_2-Q_1>0$,
2176: \beqa
2177: Q_1\rightarrow+\infty , \; Q_{21}>0 : \;\;\;
2178: P_{Q_1,Q_2}(X_1'\geq X_1,X_2'\geq X_2) & \sim & \int_0^{Q_1} \dd Q_1'
2179: \int_{Q_1'}^{Q_2} \dd Q_2' \nonumber \\
2180: && \hspace{-5cm} \times \int_{-\infty}^{\infty} \frac{\dd k}{2\pi}
2181: e^{-\ii k(X_1-Q_1')-X_1 k^2/4} \inta \frac{\dd s}{2\pi\ii} \,
2182: e^{(s-1)(Q_2'-Q_1')} \, e^{-(\sqrt{s}-1)2 X_{21}} .
2183: \label{PQ1Q2_1}
2184: \eeqa
2185: Integrating over $Q_1'$ and $Q_2'$ gives the cumulative distribution
2186: \beqa
2187: Q_1\rightarrow+\infty , \; Q_{21}>0 : \;\;\;
2188: P_{Q_1,Q_2}(X_1'\geq X_1,X_2'\geq X_2) & \sim & \int \frac{\dd k}{2\pi}
2189: \inta \frac{\dd s}{2\pi\ii} \nonumber \\
2190: && \hspace{-3.5cm} \times \frac{e^{-\ii k X_1-X_1 k^2/4-(\sqrt{s}-1)2 X_{21}}}
2191: {(s-1)(\ii k-s+1)} \left[ e^{\ii k Q_1+(s-1)Q_{21}} - e^{(s-1)Q_2} \right] .
2192: \label{PQ1Q2_2}
2193: \eeqa
2194: Then, taking the derivatives with respect to $X_1$ and $X_2$, and integrating
2195: over $k$, gives in the limit $Q_1\rightarrow+\infty$, at finite $\PPsi_1=X_1-Q_1$,
2196: the probability density
2197: \beq
2198: Q_1\rightarrow+\infty , \; Q_{21}>0 : \;\;\;
2199: P_{Q_1,Q_2}(X_1,X_2) \sim \frac{e^{-\PPsi_1^2/Q_1}}{\sqrt{\pi Q_1}}
2200: \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q_{21}-(\sqrt{s}-1)2X_{21}} \,
2201: \frac{4}{(\sqrt{s}+1)^2} .
2202: \label{PQ1Q2_3}
2203: \eeq
2204: Thus, the comparison with Eq.(\ref{pQinPsi1}) shows that we obtain as expected
2205: a factorization of the form
2206: \beq
2207: Q_1\rightarrow+\infty , \; Q_{21}>0 : \;\; P_{Q_1,Q_2}(X_1,X_2) \sim P_{Q_1}(X_1)
2208: \bP_{Q_{21}}(X_{21}) ,
2209: \label{PQ1PQ21}
2210: \eeq
2211: where the distribution, $\bP_{Q_{21}}(X_{21})$, of the relative Eulerian distance
2212: $X_{21}$ of the particles that were initially separated by the distance $Q_{21}$
2213: reads as
2214: \beq
2215: Q \geq 0 , \;\; X > 0 : \;\;\; \bP_Q(X) = \inta \frac{\dd s}{2\pi\ii}
2216: \, e^{(s-1)Q-(\sqrt{s}-1)2X} \, \frac{4}{(\sqrt{s}+1)^2} .
2217: \label{PQX}
2218: \eeq
2219: Therefore, we recover a factorization of the form (\ref{PX1X2fact}) that was
2220: obtained in the Eulerian framework. However, it is no longer exact at a finite
2221: distance from the origin and only applies in the limit $Q_1\rightarrow\infty$
2222: (again, by symmetry we have a similar result for $Q_2\rightarrow-\infty$).
2223:
2224: Multiplying Eq.(\ref{PQX}) by $e^{-4X}$ and taking the derivative with respect
2225: to $X$ we obtain a standard inverse Laplace transform \cite{Abramowitz}
2226: \beqa
2227: \frac{\dd}{\dd X} \left[e^{-4X}\bP_Q(X)\right] & = & - \inta \frac{\dd s}{2\pi\ii}
2228: \, e^{(s-1)Q} \, \frac{8 \, e^{-(\sqrt{s}+1)2X}}{\sqrt{s}+1} \nonumber\\
2229: & = & - 8 \left[ \frac{1}{\sqrt{\pi Q}} e^{-(\frac{X}{\sqrt{Q}}+\sqrt{Q})^2} -
2230: \erfc\left(\frac{X}{\sqrt{Q}}+\sqrt{Q}\right) \right] ,
2231: \label{dPQX2}
2232: \eeqa
2233: where $\erfc(z)$ is the complementary error function. This can be integrated to
2234: give
2235: \beq
2236: \bP_Q(X) = 8 \, \bigl(X+Q+\frac{1}{2}\bigr) \, e^{4X} \,
2237: \erfc\left(\frac{X}{\sqrt{Q}}+\sqrt{Q}\right) - 8 \, \sqrt{\frac{Q}{\pi}}
2238: \, e^{-(\frac{X}{\sqrt{Q}}-\sqrt{Q})^2} .
2239: \label{PQX3}
2240: \eeq
2241: Using the asymptotic expansion of the complementary error function
2242: \cite{Abramowitz} we obtain for large Lagrangian separation, $Q$, and fixed
2243: relative displacement, $\PPsi=X-Q$,
2244: \beq
2245: Q\rightarrow+\infty : \;\;\; \bP_Q(\PPsi) \sim \frac{1}{\sqrt{\pi Q}} \,
2246: e^{-\PPsi^2/Q} .
2247: \label{PQPsi1}
2248: \eeq
2249: As for Eqs.(\ref{pXinfV2}), (\ref{PVinf}), (\ref{pQinPsi1}), we recover
2250: as expected the property that over large distances particles are still
2251: governed at leading order by the initial Gaussian velocity field.
2252: Next, equation (\ref{PQX3}) yields for the asymptotic behavior at large $X$
2253: for finite $Q$,
2254: \beq
2255: X\rightarrow+\infty : \;\;\; \bP_Q(X) \sim \sqrt{\frac{Q}{\pi}} \, \frac{4}{X} \,
2256: e^{-(\frac{X}{\sqrt{Q}}-\sqrt{Q})^2}
2257: = 4 \sqrt{\frac{Q}{\pi}} \, e^{-Q} \, X^{-1} \, e^{2X-X^2/Q} ,
2258: \label{PQXinf}
2259: \eeq
2260: whereas $\bP_Q(X)$ remains finite for $X\rightarrow 0^+$. We show our results
2261: for three values of $Q$ in Fig.~\ref{figPX}, that clearly illustrate the evolution
2262: of $\bP_Q(X)$ with scale or time (smaller $Q$ corresponds to smaller scale or larger
2263: time). As for the Eulerian probability distribution, the Gaussian tail
2264: (\ref{PQXinf}) can be understood from a simple scaling argument applied to the
2265: initial velocity field. Thus, the expansion of the initial Lagrangian interval
2266: $q$ up to a very large size $x$ at time $t$ requires an initial velocity increment
2267: of order $v_0 \sim x/t$ (since $x\gg q$) which gives rise to a probability of order
2268: $e^{-(x/t)^2/q} \sim e^{-X^2/Q}$, using Eq.(\ref{Gaussian_t0}), which agrees with
2269: the large-$X$ tail (\ref{PQXinf}).
2270:
2271:
2272:
2273: \begin{figure}
2274: \begin{center}
2275: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{PX.ps}}
2276: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{lPX.ps}}
2277: \end{center}
2278: \caption{(color online) {\it Left panel:} The probability density $\bp_q(x)$ that two particles,
2279: that were initially separated by a distance $q$, are separated by the distance
2280: $x>0$ at time $t$ (in the limit where the particles are far from the origin).
2281: We show the reduced probabilities, $\bP_Q(X)$, in terms of the dimensionless variables
2282: $Q=q/(2Dt^2)$ and $X=x/(2Dt^2)$, for three values of $Q$, from Eq.(\ref{PQX3}).
2283: The probability is zero for $X<0$.
2284: For large initial relative distance $Q$ we recover a Gaussian of center $Q$ and
2285: variance $\lag (X-Q)^2\rag=Q/2$.
2286: {\it Right panel:} The probability density $\bP_Q(X)$ on a logarithmic scale,
2287: for three values of $Q$.}
2288: \label{figPX}
2289: \end{figure}
2290:
2291:
2292:
2293:
2294:
2295:
2296: \begin{figure}
2297: \begin{center}
2298: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{PQshock.ps}}
2299: \epsfxsize=6.3 cm \epsfysize=5 cm {\epsfbox{lPQshock.ps}}
2300: \end{center}
2301: \caption{(color online) {\it Left panel:} The probability, $\bp_q^{\rm shock}$, that two
2302: particles, that were initially separated by a distance $q$, have coalesced within
2303: a single shock by time $t$ (in the limit where the particles are far from the
2304: origin). We show the reduced probability, $\bP_Q^{\rm shock}$, from
2305: Eq.(\ref{PQshock1}). The dashed lines are the asymptotic behaviors
2306: (\ref{PshockQinf}) and (\ref{PshockQ0}).
2307: {\it Right panel:} The probability $\bP_Q^{\rm shock}$ on a logarithmic scale.}
2308: \label{figPQshock}
2309: \end{figure}
2310:
2311:
2312:
2313:
2314:
2315: From Eq.(\ref{PQX3})
2316: we also obtain for the asymptotic behaviors of $\bP_Q(0^+)$ with respect to
2317: $Q$
2318: \beq
2319: Q\rightarrow 0 : \;\;\; \bP_Q(0^+) \rightarrow 4 , \;\;\;\;\;\;\;\;
2320: Q\rightarrow+\infty : \;\;
2321: \bP_Q(0^+) \sim \frac{4}{\sqrt{\pi}} \, Q^{-3/2} \, e^{-Q} .
2322: \label{PQX0}
2323: \eeq
2324: However, note that the limits $X\rightarrow 0$ and $Q\rightarrow 0$ do not commute.
2325: Indeed, it is clear from Eq.(\ref{PQX3}) that for any $X>0$ we have
2326: $\bP_Q(X)\rightarrow 0$ for $Q\rightarrow 0$. As we shall see below, this is
2327: the signature of the contribution due to shocks.
2328: Thus, from Eq.(\ref{PQX}) we obtain the cumulative distribution as
2329: \beq
2330: Q \geq 0 , \;\; X > 0 : \;\;\; \bP_Q(X'\geq X) = \inta \frac{\dd s}{2\pi\ii}
2331: \, e^{(s-1)Q} \, \frac{2 \, e^{-(\sqrt{s}-1)2X}}{(s-1)(\sqrt{s}+1)} , \;\;
2332: \Re(s) > 1 ,
2333: \label{PQ>X}
2334: \eeq
2335: where the integration path is located to the right of the pole at $s=1$,
2336: so that $\bP_Q(X'\geq X)\rightarrow 0$ for $X\rightarrow+\infty$.
2337: However, we note that it does not reach unity in the limit $X\rightarrow 0^+$,
2338: since we have
2339: \beq
2340: Q \geq 0 : \;\; \lim_{X\rightarrow 0^+} \bP_Q(X'\geq X) = 1 +
2341: \inta \frac{\dd s}{2\pi\ii} \, e^{(s-1)Q} \, \frac{2}{(s-1)(\sqrt{s}+1)} ,
2342: \;\; 0 < \Re(s) < 1 ,
2343: \label{PQ>X0}
2344: \eeq
2345: where the integration path crosses the real axis in the range $0<s<1$.
2346: This means that there is a non-zero contribution due to shocks, where particles
2347: that were initially located at different positions, $q_1\neq q_2$, have collided
2348: by time $t$ and are now located at the same Eulerian position, $x_1=x_2$, in the
2349: same massive shock. Therefore, to the contribution (\ref{PQX}) we must add the
2350: contribution from shocks, that reads as
2351: \beqa
2352: Q \geq 0 & : & \;\;\; \bP_Q^{\rm shock}(X) = \bP_Q^{\rm shock} \, \delta(X-0^+) ,
2353: \;\;\;\; \mbox{with the amplitude} \nonumber \\
2354: && \bP_Q^{\rm shock} = \inta \frac{\dd s}{2\pi\ii} \,
2355: e^{(s-1)Q} \, \frac{2}{(1-s)(\sqrt{s}+1)} ,
2356: \;\;\; 0 < \Re(s) < 1 ,
2357: \label{PQshock}
2358: \eeqa
2359: so that the full probability is normalized to unity, see Eq.(\ref{PQ>X0}).
2360: Thus, the amplitude $\bP_Q^{\rm shock}$ is the probability that two particles,
2361: that were initially separated by the (dimensionless) distance $Q$, are both located
2362: in the same shock at time $t$ (in the limit where the particles are far from
2363: the origin, or anywhere on the right side for one-sided initial conditions).
2364: At large initial Lagrangian separation $Q$, we obtain from Eq.(\ref{PQshock})
2365: the exponential decay (that again can be understood from simple scaling arguments)
2366: \beq
2367: Q\rightarrow+\infty : \;\;\; \bP_Q^{\rm shock} \sim \frac{1}{\sqrt{\pi}}
2368: \, Q^{-3/2} \, e^{-Q} ,
2369: \label{PshockQinf}
2370: \eeq
2371: whereas for small initial distance $Q$ we have
2372: \beq
2373: Q\rightarrow 0 : \;\;\; \bP_Q^{\rm shock} \sim 1 - 4 \sqrt{\frac{Q}{\pi}} .
2374: \label{PshockQ0}
2375: \eeq
2376: Therefore, in the limit $Q\rightarrow 0$ the probability that both particles are
2377: within the same shock reaches unity whereas the weight associated with the
2378: ``regular'' contribution (\ref{PQX3}) vanishes (while its cutoff decreases as $Q$).
2379: This agrees with the well-known result that the set of regular Lagrangian points
2380: has a Hausdorff dimension equal to $1/2$ \cite{She1992,Sinai1992},
2381: so that with probability $1$ a random Lagrangian point $q$ belongs to a shock
2382: at any given time $t>0$. This clearly implies that $\bP_Q^{\rm shock}\rightarrow 1$
2383: for $Q\rightarrow 0$, as in Eq.(\ref{PshockQ0}). Moreover, the behavior
2384: $1-\bP_Q^{\rm shock} \propto Q^{1/2}$ also shows that the set of regular Lagrangian
2385: points has a box-counting dimension equal to $1/2$, in agreement with these works.
2386:
2387: Taking the derivative of Eq.(\ref{PQshock}) yields again a standard inverse Laplace
2388: transform \cite{Abramowitz} that provides a convenient integral expression for
2389: $\bP_Q^{\rm shock}$,
2390: \beq
2391: \frac{\dd \bP_Q^{\rm shock}}{\dd Q} = 2 \left[ \erfc(\sqrt{Q}) -
2392: \frac{e^{-Q}}{\sqrt{\pi Q}} \right] , \;\;\; \mbox{hence} \;\;\;
2393: \bP_Q^{\rm shock} = 2 \int_Q^{\infty} \dd Q' \, \left( \frac{e^{-Q'}}{\sqrt{\pi Q'}}
2394: - \erfc(\sqrt{Q'}) \right) .
2395: \label{PQshock1}
2396: \eeq
2397: We show in Fig.~\ref{figPQshock} the same-shock probability $\bP_Q^{\rm shock}$
2398: as a function of $Q$, as well as the asymptotic behaviors (\ref{PshockQinf})
2399: and (\ref{PshockQ0}).
2400:
2401:
2402:
2403: \subsection{Higher-order distributions}
2404: \label{subsec:n-point}
2405:
2406:
2407: We can obtain the higher-order $n-$point distributions $p_{q_1,..,q_n}(x_1,..,x_n)$
2408: by the same method which we applied in the previous section for the two-point
2409: distribution. Thus, as in Eq.(\ref{PQ1Q2PX1X2}), we can related the Lagrangian
2410: and Eulerian cumulative probabilities by
2411: \beq
2412: P_{Q_1,..,Q_n}(X_1'\geq X_1,..,X_n'\geq X_n) =
2413: P_{X_1,..,X_n}(Q_1'\leq Q_1,..,Q_n'\leq Q_n) ,
2414: \label{PQ1QnPX1Xn}
2415: \eeq
2416: with $Q_1 < Q_2 <..<Q_n$ and $X_1 < X_2 <..<X_n$.
2417: Then, in the limit $Q_1\rightarrow+\infty$, using again the factorization
2418: (\ref{pBn}) and the expressions (\ref{pXinfV2}) and (\ref{pBdef}), we can
2419: integrate over $Q_1',..,Q_n'$. Differentiating with respect to $X_1,..,X_n$, gives
2420: the $n-$point probability density (compare with Eq.(\ref{PQ1Q2_3}))
2421: \beqa
2422: Q_1\rightarrow\infty : \;\; P_{Q_1,..,Q_n}(X_1,..,X_n) & \sim &
2423: \frac{e^{-\PPsi_1^2/Q_1}}{\sqrt{\pi Q_1}}
2424: \inta \! \frac{\dd s_2 .. \dd s_n}{(2\pi\ii)^{(n-1)}} \,
2425: e^{(s_2-1)Q_{2,1}+..+(s_n-1)Q_{n,n-1}} \nonumber \\
2426: && \hspace{-0cm} \times
2427: \frac{2^n \, e^{-(\sqrt{s_2}-1)2X_{2,1}-..-(\sqrt{s_n}-1)2X_{n,n-1}}}
2428: {(1+\sqrt{s_2})(\sqrt{s_2}+\sqrt{s_3})..(\sqrt{s_{n-1}}+\sqrt{s_n})(\sqrt{s_n}+1)} ,
2429: \label{PQ1Qn_1}
2430: \eeqa
2431: with $X_{i,i-1}=X_i-X_{i-1}$, $Q_{i,i-1}=Q_i-Q_{i-1}$, $\Psi_1=X_1-Q_1$.
2432: Note that this $n-$point distribution does not factorize.
2433:
2434: From Eq.(\ref{PQ1Qn_1}) we can obtain the contributions associated with shocks
2435: in the same manner as in section~\ref{subsec:Two-point}. For instance, far from the
2436: origin ($Q_1\gg 1$), the probability density,
2437: $\bP_{Q_{21},Q_{32},Q_{43}}^{\rm shock}(X_{32})$, that each pair $\{Q_1,Q_2\}$,
2438: and $\{Q_3,Q_4\}$, has coalesced within two shocks that are separated by a distance
2439: in the range $[X_{32},X_{32}+\dd X_{32}]$, reads as
2440: \beqa
2441: \bP_{Q_{21},Q_{32},Q_{43}}^{\rm shock}(X_{32}) & = & \inta
2442: \frac{\dd s_2 \dd s_3 \dd s_4}{(2\pi\ii)^3} \,
2443: e^{(s_2-1)Q_{21}+(s_3-1)Q_{32}+(s_4-1)Q_{43}} \nonumber \\
2444: && \hspace{-0cm} \times \, \frac{4 \, e^{-(\sqrt{s_3}-1)2X_{32}}}
2445: {(s_2-1)(\sqrt{s_2}+\sqrt{s_3})(\sqrt{s_3}+\sqrt{s_4})(s_4-1)} ,
2446: \label{PX32shock}
2447: \eeqa
2448: where the integration contour is such that $\Re(s_2)<1$, $\Re(s_3)>1$ and
2449: $\Re(s_4)<1$.
2450:
2451:
2452:
2453: \subsection{Computing the density power spectrum from the Lagrangian statistics}
2454: \label{subsec:PkLag}
2455:
2456:
2457: Finally, it is interesting to note that the statistics of the Lagrangian
2458: displacement field also allow us to compute the Eulerian density power spectrum
2459: and to recover the result (\ref{Pkt}). Indeed, as is well-known the conservation
2460: of matter implies that the density $\rho(x)$ may be written as
2461: \beq
2462: \rho(x) = \rho_0 \left(\frac{\pl x}{\pl q}\right)^{-1} = \rho_0 \int \dd q \,
2463: \delta(x-q-\PPsi(q)) ,
2464: \label{rhoPsi1}
2465: \eeq
2466: where $\PPsi(q)=x(q)-q$ is the Lagrangian displacement of particle $q$.
2467: Note that the last expression is still valid when there are shocks, as may be
2468: seen by computing the mass within some interval $[x_1,x_2]$. Going to Fourier
2469: space as in (\ref{Pkdef}), we can write
2470: \beq
2471: \lag \rho(k_1)\rho(k_2)\rag = \rho_0^2 \int\frac{\dd q_1 \dd q_2}{(2\pi\ii)^2} \,
2472: e^{-\ii(k_1q_1+k_2q_2)} \lag e^{-i(k_1\PPsi_1+k_2\PPsi_2)}\rag .
2473: \eeq
2474: Then, in the regime where the invariance through translations is recovered (i.e.
2475: far from the origin), making the changes of variables $q_2=q_1+q$ and
2476: $\PPsi_2=\PPsi_1+\PPsi$, we obtain
2477: \beqa
2478: \cP(k) & = & \int_0^{\infty} \frac{\dd q}{\pi} \int_0^{\infty} \dd x \, \bp_q(x)
2479: \cos(kx) \\
2480: & = & \gamma^2 \int_0^{\infty} \frac{\dd Q}{\pi} \int_{0^+}^{\infty}
2481: \dd X \, \bP_Q(X) \cos(\gamma^2kX) + \gamma^2 \int_0^{\infty} \frac{\dd Q}{\pi}
2482: \, \bP_Q^{\rm shock} .
2483: \eeqa
2484: In the second line, written in terms of dimensionless variables, we separated
2485: the two contributions associated with the regular part (\ref{PQX3}) and with
2486: the singular part (\ref{PQshock}) of the distribution of the relative Eulerian
2487: distance $X$. Using Eq.(\ref{PQX}), we can check that the first contribution
2488: actually vanishes (as expected since all the mass is enclosed within shocks,
2489: see Eq.(\ref{Nshocknorm}) below and \cite{She1992,Sinai1992}), whereas the
2490: shock contribution is obviously independent of $k$ and using Eq.(\ref{PQshock})
2491: we recover the amplitude (\ref{Pkt}).
2492:
2493:
2494:
2495:
2496: \section{Properties of shocks}
2497: \label{sec:shock}
2498:
2499:
2500:
2501: \subsection{Shock mass function}
2502: \label{subsec:shockmass}
2503:
2504:
2505: From the probability $\bp_q^{\rm shock}$ that two particles, initially separated
2506: by a distance $q$, are located in the same shock at a time $t>0$,
2507: we can now derive the mass function of shocks. First, we note that if a shock
2508: has Lagrangian end-points $q_-$ and $q_+$, its mass is simply $m=\rho_0 (q_+-q_-)$,
2509: where $\rho_0$ is the uniform initial density, as discussed in
2510: section~\ref{sec:Density}. Then, within an interval of length $\cQ$ in
2511: Lagrangian space,
2512: we now count the number, $\cQ n_{\cQ}(m) \dd m$, of shock-intervals of length in
2513: the range $[q,q+\dd q]$, whence of mass in $[m,m+\dd m]$, with $m=\rho_0 q$.
2514: The limit $\cQ\rightarrow \infty$ gives the probability density $n(m)$
2515: for a Lagrangian point to belong to a shock of mass $m$. Since on large scales
2516: particles are still governed by the initial Gaussian velocity field and have
2517: only moved by a relative distance $\PPsi \sim \sqrt{\cQ}$, see for instance
2518: Eq.(\ref{PQPsi1}), the corresponding Eulerian relative distance, $\cX$, obeys
2519: $\cX/\cQ\rightarrow 1$ for $\cQ\rightarrow\infty$. Therefore, $n(m)\dd m$ is
2520: also the mean number of shocks, per unit Eulerian length, with a mass in the range
2521: $[m,m+\dd m]$.
2522:
2523: Let us now relate the mass function $n(m)$ to the shock probability
2524: $\bp_q^{\rm shock}$ that we obtained in Eq.(\ref{PQshock}). The latter is the
2525: probability that a Lagrangian interval, $q=q_2-q_1$, chosen at random, has
2526: coalesced by time $t$ within a single shock. This clearly means that this shock
2527: has a length $q^{\rm s}$ larger than $q$ and that $q_1$ is located within
2528: a distance smaller than $q^{\rm s}-q$ from its left boundary. Therefore, in terms
2529: of dimensionless variables, we have the relation
2530: \beq
2531: \bP_Q^{\rm shock} = \int_Q^{\infty} \dd M \, N(M) \, (M-Q) ,
2532: \;\;\; \mbox{with} \;\;\; M = \frac{m}{\rho_0 \gam^2} .
2533: \label{Nshock1}
2534: \eeq
2535: Here $N(M)$ is the dimensionless mass function. From Eq.(\ref{Nshock1}) and
2536: Eq.(\ref{PshockQ0}), taking $Q=0$, we obtain at once the normalization
2537: \beq
2538: \int_0^{\infty} \dd M \, M \, N(M) = 1 ,
2539: \label{Nshocknorm}
2540: \eeq
2541: which means that all the mass is included within shocks, at any time $t>0$.
2542: This agrees with previous results discussed below Eq.(\ref{PshockQ0}),
2543: see \cite{She1992} and \cite{Sinai1992}.
2544: Differentiating twice Eq.(\ref{Nshock1}) with respect to $Q$, and using the first
2545: Eq.(\ref{PQshock1}), gives the simple expression
2546: \beq
2547: N(M)= \left. \frac{\dd^2 \bP_Q^{\rm shock}}{\dd Q^2} \right|_{Q=M} , \;\;\;
2548: \mbox{whence} \;\;\; N(M) = \frac{1}{\sqrt{\pi}} \, M^{-3/2} \ e^{-M} .
2549: \label{Nshock2}
2550: \eeq
2551: Thus, we recover the low-mass power law $M^{-3/2}$ that was already obtained
2552: in \cite{She1992,Sinai1992}. At large masses we obtain the exponential falloff
2553: that was heuristically derived in \cite{Vergassola1994}, following the same scaling
2554: arguments as those described in previous sections for the tails of the Eulerian or
2555: Lagrangian distributions. The full mass function
2556: (\ref{Nshock2}) was also obtained in \cite{Bertoin1998} for the one-sided
2557: Brownian initial velocity. Indeed, in that case the Lagrangian increments $q_{21}$
2558: have the same distribution for $x\geq 0$, as discussed in
2559: section~\ref{subsec:general}, which clearly leads to identical shock properties.
2560:
2561: It is interesting to compare the exact result (\ref{Nshock2}) with the
2562: Press-Schechter ansatz that is widely used in the cosmological context to
2563: count the number of collapsed objects \cite{Press1974}. This model attempts
2564: to identify such objects from the properties of the linear fields, obtained from
2565: the linearization of the equations of motion. For our case,
2566: this heuristic approach would state that the fraction of matter, $F(>m)$,
2567: that is enclosed within collapsed objects (here infinitesimally thin shocks,
2568: as we consider the Burgers equation in the inviscid limit) of mass larger than
2569: $m$, with $m=\rho_0 q$, is given by the probability that, choosing a Lagrangian
2570: point $q_c$ at random, the linear-theory Eulerian relative distance $x_L$ at
2571: time $t$ between the particles $q_c+q/2$ and $q_c-q/2$ vanishes.
2572: (For a 3-dimensional Universe one considers the probability the a sphere of
2573: mass $m$ centered on $q_c$ has collapsed to a point.) In terms of dimensionless
2574: variables this reads as
2575: \beq
2576: F^{\rm PS}(\geq M) = P_Q^L(X_L \leq 0) \;\; \mbox{at} \;\; M=Q ,
2577: \;\; \mbox{with} \;\; P_Q^L(X_L)= \frac{e^{-(X_L-Q)^2/Q}}{\sqrt{\pi Q}} ,
2578: \label{FPS1}
2579: \eeq
2580: where $P^L$ refers to the distribution obtained by linear theory, where particles
2581: always keep their initial velocity and shocks are discarded.
2582: This gives
2583: \beq
2584: F^{\rm PS}(\geq M) = \int_{\sqrt{M}}^{\infty} \dd y \,
2585: \frac{e^{-y^2}}{\sqrt{\pi}} .
2586: \label{FPS2}
2587: \eeq
2588: As usual, Eq.(\ref{FPS2}) implies $F^{\rm PS}(\geq 0)=1/2$, which means that only
2589: half of the mass would be within collapsed structures. Therefore, it is customary
2590: to multiply this by a somewhat ad-hoc factor $2$ \cite{Press1974}.
2591: Thus, differentiating with respect to $M$, the standard Press-Schechter recipe gives
2592: in our case the mass function
2593: \beq
2594: 2F^{\rm PS}(\geq M) = \int_M^{\infty} \dd M \, M \, N^{\rm PS}(M) , \;\;
2595: \mbox{whence} \;\; N^{\rm PS}(M) = \frac{1}{\sqrt{\pi}} \, M^{-3/2} \ e^{-M} .
2596: \label{NPS}
2597: \eeq
2598: Therefore, we find that for the $1$-D Burgers dynamics with Brownian initial
2599: velocity the Press-Schechter ansatz happens to give the exact mass function
2600: (\ref{Nshock2}). The agreement of the Press-Schechter mass function at both small
2601: and large masses for the one-dimensional case was already noticed in
2602: \cite{Vergassola1994}, for more general power-law initial velocity energy
2603: spectra (although there were no exact results available at large masses but
2604: heuristic predictions). This can be somewhat surprising in view of the many
2605: effects that could have made the Press-Schechter ansatz fail (especially at the
2606: low-mass tail), such as the so-called ``cloud-in-cloud problem'', associated
2607: here with the fact that, even though particles $q_c\pm q/2$, evolved according to
2608: linear theory, may have not collided yet, on a larger scale $\ell >q$ it may happen
2609: that particles $q_c\pm \ell/2$ had very large inward velocities and have formed
2610: by time $t$ a massive shock that includes the smaller scale $q$.
2611: Nevertheless, in the cosmological context, numerical simulations have shown that,
2612: even though the Press-Schechter mass function is not exact, is usually gives
2613: reasonably good estimates (e.g., \cite{Sheth1999}), so that it is still widely
2614: used today. It is satisfying to find out that in a related dynamical system,
2615: it actually happens to
2616: coincide with the exact result. This makes the reasonable agreement observed in
2617: other cases somewhat less surprising than would be expected at first sight.
2618: As found in \cite{Vergassola1994}, this also suggests that it could provide
2619: a reasonable estimate for the Burgers dynamics itself with more general initial
2620: conditions.
2621:
2622:
2623:
2624:
2625:
2626: \subsection{Spatial distribution of shocks}
2627: \label{subsec:shockcorrelation}
2628:
2629:
2630: Finally, from the $n-$point distributions (\ref{PQ1Qn_1}) we can derive the
2631: many-body distributions of shocks, far from the origin.
2632: Thus, in a fashion similar to Eq.(\ref{Nshock1}), we can relate the trivariate mass
2633: function, $N(M_1,M,M_2;X)\dd M_1 \dd M \dd M_2 \dd X_1 \dd X$, that counts the
2634: probability to have a shock of mass $M_1$ within $[X_1,X_1+\dd X_1]$, another
2635: shock of mass $M_2$ at distance $[X,X+\dd X]$, and a mass $M$
2636: in-between both shocks, to the three-point conditional shock probability,
2637: $\bP_{Q_{21},Q_{32},Q_{43}}$, of Eq.(\ref{PX32shock}). This reads as
2638: \beqa
2639: \bP_{Q_{21},Q_{32},Q_{43}}(X) & = & \int_0^{Q_{32}} \dd M
2640: \int_{Q_{21}}^{\infty} \dd M_1 \int_{Q_{43}}^{\infty} \dd M_2 \, N(M_1,M,M_2;X)
2641: \int_0^{M_1-Q_{21}} \dd Q_1 \nonumber \\
2642: && \hspace{0cm} \times \, \theta(Q_1\!+\!Q_{21}\!+\!Q_{32}\!-\!M_1\!-\!M) \,
2643: \theta(M_1\!+\!M\!+\!M_2\!-\!Q_1\!-\!Q_{21}\!-\!Q_{32}\!-\!Q_{43}) ,
2644: \label{N12shocks1}
2645: \eeqa
2646: where the two Heaviside factors ensure that $Q_3$ and $Q_4$ are within the second
2647: shock of mass $M_2$. Then, differentiating with respect to $Q_{21}$ and $Q_{43}$
2648: yields
2649: \beq
2650: \frac{\pl^2 \bP_{Q_{21},Q_{32},Q_{43}}}{\pl Q_{21}\pl Q_{43}} =
2651: \int_0^{Q_{32}} \! \dd M \int_{Q_{43}}^{Q_{43}+Q_{32}-M} \! \dd M_2 \,
2652: N(Q_{21}\!+\!Q_{32}\!+\!Q_{43}-\!M\!-\!M_2,M,M_2;X) .
2653: \label{N12shocks2}
2654: \eeq
2655: Taking the shifted Laplace transform of both quantities, in the form
2656: \beq
2657: N(M_1,M,M_2;X) = \inta \frac{\dd s_1 \dd s \dd s_2}{(2\pi\ii)^3}
2658: e^{(s_1-1) M_1+(s-1)M+(s_2-1)M_2} \, \tN(s_1,s,s_2;X) ,
2659: \label{tNs1}
2660: \eeq
2661: and using Eq.(\ref{PX32shock}), we obtain
2662: \beq
2663: \tN(s_1,s,s_2;X) = \frac{4 \, (s-s_1)(s-s_2) \, e^{-(\sqrt{s}-1)2X}}
2664: {(\sqrt{s_1}+\sqrt{s})(\sqrt{s}+\sqrt{s_2})} .
2665: \label{tNs2}
2666: \eeq
2667:
2668: If we now consider the multiplicity of shocks at positions $X_1$ and $X_2$,
2669: independently of the mass $M$ in-between, we integrate over $M$ and $s$,
2670: which gives the bivariate mass function of shocks separated by a distance $X$:
2671: \beq
2672: N(M_1,M_2;X) = \inta \frac{\dd s_1 \dd s_2}{(2\pi\ii)^2}
2673: e^{(s_1-1) M_1+(s_2-1)M_2} \, \frac{4(1-s_1)(1-s_2)}
2674: {(\sqrt{s_1}+1)(1+\sqrt{s_2})} = N(M_1)\, N(M_2) ,
2675: \label{N12shocks3}
2676: \eeq
2677: where we used Eqs.(\ref{Nshock2}) and (\ref{PQshock}) to recognize the product
2678: $N(M_1) N(M_2)$. Therefore, we find that the bivariate mass function $n(m_1,m_2;x)$
2679: does not depend on the inter-shock distance $x$ and merely factorizes as
2680: $n(m_1) \times n(m_2)$ (far from the origin $x_1=0$).
2681:
2682: Thus, shocks are not correlated and there is no bias: knowing that there is a shock
2683: of mass $m_1$ at position $x_1$ does not bias in any way the shock multiplicity
2684: at position $x_2=x_1+x$. We can note that this is consistent with the fact that
2685: the densities within separate regions are uncorrelated, as seen in
2686: Eqs.(\ref{Ceta}) and (\ref{Peta1n}). In fact, from Eq.(\ref{Peta1n}), which shows
2687: that the density fields over separate regions are completely independent,
2688: we can see that this must extend to all $n-$point shock mass functions, thus
2689: shocks at different positions are uncorrelated. The fact that shocks form
2690: a Poisson point process was also obtained in \cite{Bertoin1998} for the case of
2691: one-sided Brownian velocity.
2692:
2693:
2694:
2695: \section{Conclusion}
2696: \label{sec:conclusion}
2697:
2698:
2699:
2700: We have shown in this paper how to derive the equal-time statistical properties of
2701: the solution of the Burgers equation with Brownian initial velocity, using the
2702: transition kernel associated with Brownian particles over a parabolic absorbing
2703: barrier. This initial velocity field is not homogeneous, as the initial velocity
2704: is Gaussian with a variance
2705: $\lag v_0^2\rag \propto |x|$ at distance $|x|$. However, it has homogeneous
2706: increments. Then, although the one-point distributions,
2707: $p_x(v)$ and $p_x(q)$, of the velocity $v$ and initial Lagrangian position $q$,
2708: depend on the position $x$, the two-point distributions exactly factorize provided
2709: that all spatial coordinates remain on the same side of $x=0$, such as
2710: $p_{x_1,x_2}(v_1,v_2)=p_{x_1}(v_1) \bp_{x_{21}}(v_{21})$ and
2711: $p_{x_1,x_2}(q_1,q_2)=p_{x_1}(q_1) \bp_{x_{21}}(q_{21})$ for
2712: $q_i>0$. A similar factorization holds for higher-order distributions.
2713: This agrees with the results that were obtained for the one-sided Brownian
2714: initial velocity by \cite{Bertoin1998}.
2715: In the limit where we are far from the origin, this implies that we recover the
2716: invariance through translations for the distributions of velocity and Lagrangian
2717: increments. Then, we have focussed on the properties of the system
2718: in this limit, where many simple explicit results can be derived.
2719:
2720: As expected, we have found that on large scales, or at early times, all statistical
2721: properties converge to the Gaussian distributions set by the initial conditions
2722: (the nonlinear evolution being subdominant). On small scales, or at late times,
2723: the distributions of the velocity increment, Lagrangian increment,
2724: or mean overdensity $\eta$ within a region of length $x$, increasingly depart
2725: from the Gaussian. They exhibit widely separated exponential cutoffs, of the form
2726: $e^{-\eta}$ and $e^{-1/\eta}$, while a power law $\eta^{-3/2}$ develops in the
2727: intermediate range. However, we find that the variance of these distributions
2728: remains unchanged by the nonlinear dynamics, that is, it is equal to the value
2729: that would be obtained by discarding collisions and shocks, and letting particles
2730: cross each other and always keep their initial velocity. In particular,
2731: the second-order velocity structure function and its energy spectrum do not evolve
2732: with time, while the density correlation remains a Dirac function with an amplitude
2733: that grows as $t^2$.
2734:
2735: In fact, the densities within non-overlapping regions are uncorrelated
2736: and, at any order, the $n-$point connected density correlation can be written
2737: as a product of $n-1$ Dirac factors: it is non-vanishing only when all points
2738: coincide. Then, it can also be written as a product of $n-1$ two-point correlations,
2739: that connect the $n$ points, with a constant amplitude that happens to be the number
2740: of heap ordered trees. This allows a combinatorial interpretation that is similar
2741: to the hierarchical tree models that were devised as a phenomenological
2742: tool in the cosmological context, with the difference that in our case we must
2743: consider ordered trees (note that in the present $1$-D system,
2744: where particles do not cross, it is meaningful to order particles by their
2745: positions so that the concept of ordering appears rather natural).
2746: Then, the cumulants of the overdensity exactly scale as
2747: $\lag\eta^n\rag_c \propto \lag\eta^2\rag_c^{(n-1)}$, with an amplitude that is
2748: independent of time and scale. Thus, they happen to exactly satisfy the
2749: so-called ``stable-clustering ansatz''. In fact, the density cumulant-generating
2750: function remains exactly equal to the one obtained at tree-order from a perturbative
2751: approach (which breaks down beyond leading order as next-to-leading corrections
2752: actually are divergent, which signals the need for a non-perturbative method that
2753: takes into account shocks).
2754:
2755: We have also studied the Lagrangian displacement field, associated with a Lagrangian
2756: description of the dynamics. In the limit where we are far from the origin, we
2757: find that it satisfies a similar factorization and recovers the invariance
2758: through translations for the distributions of relative displacements.
2759: This also allows us to derive the properties of shocks. Thus, in agreement with
2760: previous works, we find that all of the mass is enclosed within shocks, at any time
2761: $t>0$, and that the shock mass function has the very simple expression
2762: (\ref{Nshock2}). It agrees with the asymptotic behaviors that were already
2763: obtained through analytical means or numerical simulations in
2764: \cite{She1992,Sinai1992} or \cite{Vergassola1994} and the exact result of
2765: \cite{Bertoin1998}. Finally, shocks are not correlated as the bivariate
2766: multiplicity function
2767: $n(m_1,m_2;x)$, that counts shocks of mass $m_1$ and $m_2$ separated by the
2768: distance $x$, factorizes as $n(m_1) \times n(m_2)$, in agreement with the same
2769: lack of correlation obtained for the density field.
2770:
2771: Thus, the equal-time statistical properties of the Burgers dynamics with Brownian
2772: initial velocity are remarkably simple. It appears that the nonlinear dynamics
2773: preserves some properties of the initial fields (e.g. the second-order structure
2774: functions, the independence and homogeneity of velocity increments and of the
2775: densities in separate domains) and that simple
2776: explicit expressions can be derived in the limit where we are far from the origin
2777: (or on the right side for one-sided initial conditions).
2778:
2779: At finite distance from the origin, in addition
2780: to the quantities given here we need the two-point distribution associated with
2781: the case where the two particles are on different sides of the origin $x=0$.
2782: Although we can obtain explicit expressions by the method presented in this article,
2783: this leads to multidimensional integrals that do not seem to greatly simplify
2784: (although we obtained simple expressions for the one-point distributions).
2785: However, for practical purposes, one is mostly interested in the limit
2786: where we are far from the origin and homogeneity is recovered.
2787: From a physical point of view, the initial conditions (\ref{xidef}) are meant
2788: to represent a system with homogeneous velocity increments, which scale as
2789: $\lag(\Delta v_0)^2\rag \propto |\Delta q|$ as in(\ref{Dv0}) over a finite
2790: range, and an energy spectrum $E_0(k)\propto k^{-2}$ as in (\ref{E0}) over the
2791: range of interest. Thus, in practice there would be an infrared cutoff, $\Lambda$,
2792: below which $E_0(k)$ would grow more slowly than $1/k$ so that the velocity
2793: field is actually homogeneous (in an experimental setup there would actually be
2794: a finite lower wavenumber, set by the size of the box, and homogeneity would only
2795: apply far from the boundaries). Then, the initial conditions (\ref{xidef})
2796: studied in this article can be viewed as a convenient mathematical device
2797: to represent such a system, with the understanding that the special role
2798: played by the origin is a mathematical artifact and that only the properties
2799: far from the origin are meaningful in the physical sense described above.
2800: Note that this identification is possible because small scales are not strongly
2801: coupled to large scales, in agreement with the fact that over large scales
2802: we recover the initial fields and no strong correlations develop.
2803:
2804: To put this study in a broader context, it may be useful to recall here the
2805: main properties of ``decaying Burgers turbulence'' for more general Gaussian
2806: initial conditions. It is customary to study the Burgers dynamics (\ref{Burg})
2807: for power-law energy spectra, $E_0(k) \propto k^n$ (here we focussed on the
2808: case $n=-2$, see Eq.(\ref{E0})). Indeed, at late times the asymptotic statistical
2809: properties of the velocity field no longer depend on the details of the high-$k$
2810: spectrum (assuming a strong enough falloff) nor on the precise value of the
2811: viscosity $\nu$, as a self-similar evolution develops
2812: \cite{Gurbatov1991,Molchanov1995}.
2813: Then, depending on the exponent $n$, the integral scale of turbulence, $L(t)$,
2814: which measures the typical distance between shocks and the correlation length,
2815: and the shock and velocity probability distributions show the following behaviors.
2816:
2817: For $-3<n<-1$ (which includes the case $n=-2$ studied in this article, associated
2818: with a Brownian initial velocity), the initial velocity is not homogeneous but it
2819: has homogeneous increments, while for $-1<n<1$ (which includes the case $n=0$
2820: associated with a white-noise initial velocity), the initial velocity itself is
2821: homogeneous. In both cases, the integral scale grows as $L(t) \sim t^{2/(n+3)}$,
2822: and the tails of the cumulative shock distribution and velocity distribution
2823: satisfy $\ln[n(>m)] \sim -m^{n+3}$, $\ln[n(>|v|)] \sim -|v|^{n+3}$, for $m\rightarrow
2824: \infty,|v|\rightarrow\infty$, see \cite{She1992,Molchan1997}.
2825: At low wavenumbers, below $1/L(t)$, the energy spectrum keeps its initial form,
2826: $E(k,t) \propto k^n$, whereas at high wavenumbers it shows the universal law,
2827: $E(k,t) \propto k^{-2}$, due to shocks \cite{Gurbatov1997,Noullez2005}.
2828: The preservation of the large-scale part, $E(k,t) \propto k^n$, is associated with
2829: the ``principle of permanence of large eddies'' \cite{Gurbatov1997}. Physically,
2830: this means that, at a given time $t$, structures of size larger than $L(t)$ have not
2831: had time to be strongly distorted by the dynamics (in agreement with the simple
2832: scaling argument $t\sigma_{v_0}(x) \ll x$ which gives $x\gg L(t)$). In particular,
2833: not only statistical properties but each random realization is stable against
2834: small-scale perturbations \cite{Aurell1993,Gurbatov1999}.
2835: Then, the tails of the shock and velocity distributions can be understood from the
2836: initial velocity field. Thus, the velocity difference between the left and right
2837: boundaries of a shock of mass $m=\rho_0 q$ is $q/t$, which leads to a probability
2838: $\sim e^{-(q/t)^2/\sigma_{v_0}(q)^2} \sim e^{-m^{n+3}/t^2}$ (where we did not write
2839: constants in the exponent) \cite{Vergassola1994}.
2840: We can check that these properties agree with the results derived for $n=-2$
2841: in this paper.
2842:
2843: For $1<n<2$ the system shows a more complex behavior, since there are three scaling
2844: regions for the energy spectrum: first a $k^n$ region at very low wavenumbers,
2845: below $k_s(t) \sim t^{-1/2(2-n)}$, next a $k^2$ region between $k_s(t)$ and
2846: $k_L(t) \sim t^{-1/2}$, and finally the standard $k^{-2}$ region above $k_L(t)$
2847: \cite{Gurbatov1997}. Therefore the evolution is no longer self-similar.
2848: For $n>2$ the $k^n$ region disappears (it gives subdominant corrections) and the
2849: leading-order evolution is again self-similar but independent of $n$
2850: \cite{Gurbatov1997,Gurbatove1981}.
2851:
2852: We can hope that the exact results presented in this article for the case
2853: of Brownian initial velocity could serve as a useful benchmark to test
2854: approximation schemes which could be devised to handle other initial
2855: conditions where no exact results are available.
2856: In particular, in the cosmological context, the Zeldovich approximation, which
2857: corresponds to removing the diffusive term altogether, has already been used to
2858: test for instance field-theoretic methods that attempt to resum perturbative series
2859: \cite{Valageas2007}. The Burgers equation in the inviscid limit, which corresponds
2860: to the more efficient adhesion model, might also be used for such purposes.
2861: In a similar fashion, the general properties of the Burgers dynamics (associated
2862: with shocks) have already been used to test approximation schemes devised for
2863: the study of turbulence \cite{Fournier1983}.
2864:
2865: Finally, we note that the method described in this article could also be applied
2866: to different-time statistics, where the parabolas used in the geometrical
2867: interpretations would now have different curvatures. However, we leave such
2868: studies for future works.
2869:
2870:
2871:
2872:
2873:
2874:
2875: \appendix
2876:
2877:
2878: \section{Some properties of the Airy functions}
2879: \label{Airy}
2880:
2881:
2882: We recall here some properties of the Airy functions $\Ai(x)$ and $\Bi(x)$ that
2883: are used repeatedly in the calculations presented in this article. These two
2884: Airy functions are two linearly independent solutions to the second-order
2885: differential equation $y''(x)-x y(x)=0$. The first one, $\Ai(x)$, is the only
2886: solution that vanishes at both ends, $x \rightarrow \pm\infty$,
2887: whereas $\Bi(x)$ grows to infinity at $x\rightarrow +\infty$ \cite{Abramowitz}.
2888: Both are entire functions and they are related through
2889: \beq
2890: \Bi(x) = e^{\ii \pi/6} \Ai(e^{\ii 2\pi/3} x)
2891: + e^{-\ii \pi/6} \Ai(e^{-\ii 2\pi/3} x) ,
2892: \label{BiAi}
2893: \eeq
2894: while their Wronskian is constant and given by
2895: \beq
2896: \Ai(x) \, \Bi\,'(x) - \Ai\,'(x) \, \Bi(x) = \frac{1}{\pi} .
2897: \label{Wronskian}
2898: \eeq
2899: We also have the integral representation \cite{Abramowitz}
2900: \beq
2901: \Ai(x) = \int_{-\infty}^{\infty} \frac{\dd t}{2\pi}
2902: \, e^{\ii (\frac{t^3}{3}+x t)} .
2903: \label{Ai1int}
2904: \eeq
2905: At $x=0$ we have
2906: \beq
2907: \frac{\Bi(0)}{\sqrt{3}} = \Ai(0)= \frac{1}{3^{2/3} \Gamma[2/3]} , \;\;\;\;
2908: \frac{-\Bi\,'(0)}{\sqrt{3}} = \Ai\,'(0)= \frac{-1}{3^{1/3} \Gamma[1/3]} ,
2909: \label{Aix0}
2910: \eeq
2911: and for $|x|\rightarrow\infty$:
2912: \beq
2913: |\Arg(x)| < \pi : \;\;\; \Ai(x) \sim \frac{1}{2\sqrt{\pi}} \, x^{-1/4} \,
2914: e^{-\frac{2}{3}x^{3/2}} ,
2915: \label{Aiinf}
2916: \eeq
2917: \beq
2918: |\Arg(x)| < \frac{2\pi}{3} : \;\;\; \Ai(-x) \sim \frac{1}{\sqrt{\pi}} \,
2919: x^{-1/4} \, \sin\left[\frac{2}{3}x^{3/2}+\frac{\pi}{4}\right] ,
2920: \label{Aipinf}
2921: \eeq
2922: \beq
2923: |\Arg(x)| < \frac{\pi}{3} : \;\;\; \Bi(x) \sim \frac{1}{\sqrt{\pi}} \, x^{-1/4} \,
2924: e^{\frac{2}{3}x^{3/2}} ,
2925: \label{Biinf}
2926: \eeq
2927: \beq
2928: |\Arg(x)| < \frac{2\pi}{3} : \;\;\; \Bi(-x) \sim \frac{1}{\sqrt{\pi}} \,
2929: x^{-1/4} \, \cos\left[\frac{2}{3}x^{3/2}+\frac{\pi}{4}\right] .
2930: \label{Bipinf}
2931: \eeq
2932: For $|\Arg(x)|<2\pi/3$, the Airy function can also be expressed in terms of the
2933: modified Bessel function of the second kind $K_{\nu}$ as
2934: \beq
2935: |\Arg(x)|<\frac{2\pi}{3} : \;\; \Ai(x)= \frac{1}{\pi} \sqrt{\frac{x}{3}}
2936: K_{1/3}\left(\frac{2}{3}x^{3/2}\right) ,
2937: \;\;\; \Ai\,'(x)= \frac{-x}{\pi\sqrt{3}} K_{2/3}\left(\frac{2}{3}x^{3/2}\right) .
2938: \label{AiKnu}
2939: \eeq
2940: Four useful integrals, that may be obtained from the integral representation
2941: (\ref{Ai1int}), are \cite{ValleeSoares}
2942: \beq
2943: \int_{-\infty}^{\infty} \dd x \, e^{\alpha x} \Ai(x) = e^{\alpha^3/3} ,
2944: \;\;\; \mbox{whence} \;\;\;
2945: \int_{-\infty}^{\infty} \dd x \, e^{\alpha x} x \Ai(x) = \alpha^2 e^{\alpha^3/3} ,
2946: \label{IntExpAi}
2947: \eeq
2948: \beq
2949: \nu_1 \neq \nu_2 : \;\;\; \int_{-\infty}^{\infty} \dd u \,
2950: \Ai\left[\nu_1 u+\frac{s_1}{\nu_1^2}\right]
2951: \Ai\left[\nu_2 u+\frac{s_2}{\nu_2^2}\right] =
2952: \frac{1}{|\nu_1^3-\nu_2^3|^{1/3}} \Ai \left[ (\nu_1^3-\nu_2^3)^{-1/3}
2953: \left( \frac{\nu_1 s_2}{\nu_2^2} - \frac{\nu_2 s_1}{\nu_1^2} \right) \right] ,
2954: \label{IntAi3}
2955: \eeq
2956: and
2957: \beq
2958: \nu_1 \neq \nu_2 : \;\;\; \int_{-\infty}^{\infty} \dd u \, u \,
2959: \Ai\left[\nu_1 u+\frac{s_1}{\nu_1^2}\right]
2960: \Ai\left[\nu_2 u+\frac{s_2}{\nu_2^2}\right] =
2961: \frac{s_2-s_1}{(\nu_1^3-\nu_2^3)^{4/3}} \Ai \left[ (\nu_1^3-\nu_2^3)^{-1/3}
2962: \left( \frac{\nu_1 s_2}{\nu_2^2} - \frac{\nu_2 s_1}{\nu_1^2} \right) \right] ,
2963: \label{IntAi4}
2964: \eeq
2965: with the conventions:
2966: $(\nu_1^3-\nu_2^3)^{-1/3} \rightarrow - (\nu_2^3-\nu_1^3)^{-1/3}$ and
2967: $(\nu_1^3-\nu_2^3)^{4/3} \rightarrow - (\nu_2^3-\nu_1^3)^{4/3}$
2968: if $\nu_1<\nu_2$. This also implies the relation
2969: \beq
2970: \nu_1 \neq \nu_2 : \;\;\; \int_{-\infty}^{\infty} \dd u \, u \,
2971: \Ai\left[\nu_1 u+\frac{s_1}{\nu_1^2}\right]
2972: \Ai\left[\nu_2 u+\frac{s_2}{\nu_2^2}\right] =
2973: \frac{s_2-s_1}{\nu_1^3-\nu_2^3} \int_{-\infty}^{\infty} \dd u \,
2974: \Ai\left[\nu_1 u+\frac{s_1}{\nu_1^2}\right]
2975: \Ai\left[\nu_2 u+\frac{s_2}{\nu_2^2}\right] ,
2976: \label{IntAi5}
2977: \eeq
2978: that could be obtained from the property $\Ai''(x)=x\Ai(x)$.
2979:
2980: Finally, using the property \cite{Gradshteyn}
2981: \beqa
2982: \mu>\nu , \;\; \alpha+\beta>0 : \;\;\; \int_0^{\infty} \dd x \, x^{\mu-1}
2983: e^{-\alpha x} K_{\nu}(\beta x) & = & \frac{\sqrt{\pi} (2\beta)^{\nu}}
2984: { (\alpha+\beta)^{\mu+\nu}} \frac{\Gamma(\mu+\nu) \Gamma(\mu-\nu)}
2985: {\Gamma(\mu+\frac{1}{2})} \nonumber \\
2986: && \hspace{-2cm} \times \; _2F_1\left(\mu+\nu,\nu+\frac{1}{2};\mu+\frac{1}{2};
2987: \frac{\alpha-\beta}{\alpha+\beta}\right) ,
2988: \eeqa
2989: and the relations (\ref{AiKnu}), we obtain
2990: \beqa
2991: \nu>0 , \;\; \alpha+\beta>0 : \;\; \int_0^{\infty} \dd x \, x^{\nu-1}
2992: e^{-\alpha x} \Ai\left[\left(\frac{3\beta x}{2}\right)^{2/3}\right] & = &
2993: \frac{1}{\sqrt{\pi}} \, 3^{-1/6} \beta^{2/3} (\alpha+\beta)^{-\nu-\frac{2}{3}}
2994: \nonumber \\
2995: && \hspace{-4cm} \times \; \frac{\Gamma(\nu+\frac{2}{3}) \Gamma(\nu)}
2996: {\Gamma(\nu+\frac{5}{6})} \; _2F_1\left(\nu+\frac{2}{3},\frac{5}{6};
2997: \nu+\frac{5}{6};\frac{\alpha-\beta}{\alpha+\beta}\right) ,
2998: \label{IntAiExp1}
2999: \eeqa
3000: \beqa
3001: \nu>0 , \;\; \alpha+\beta>0 : \;\; \int_0^{\infty} \dd x \, x^{\nu-1}
3002: e^{-\alpha x} \Ai\,'\left[\left(\frac{3\beta x}{2}\right)^{2/3}\right] & = &
3003: \frac{-1}{\sqrt{\pi}} \, 3^{1/6} \beta^{4/3} (\alpha+\beta)^{-\nu-\frac{4}{3}}
3004: \nonumber \\
3005: && \hspace{-4.5cm} \times \; \frac{\Gamma(\nu+\frac{4}{3}) \Gamma(\nu)}
3006: {\Gamma(\nu+\frac{7}{6})} \; _2F_1\left(\nu+\frac{4}{3},\frac{7}{6};
3007: \nu+\frac{7}{6};\frac{\alpha-\beta}{\alpha+\beta}\right) .
3008: \label{IntAipExp1}
3009: \eeqa
3010:
3011:
3012:
3013:
3014: %\beq
3015: %|\Arg(x)|<\frac{2\pi}{3} : \;\;\; \Ai(x) = \frac{1}{2\sqrt{\pi}} \, x^{-1/4}
3016: %\, e^{-\frac{2}{3}x^{3/2}}
3017: %\int_0^{\infty} \dd z \, \frac{\rho(z)}{1+\frac{3}{2}x^{-3/2} z} ,
3018: %\label{Ai_int}
3019: %\eeq
3020: %\beq
3021: %\mbox{with} \;\;\; \rho(z)= \frac{1}{\sqrt{\pi}} \, \left(\frac{3}{2}\right)^{1/6}
3022: %\, z^{-5/6} \, e^{-z} \, \Ai\left[\left(\frac{3z}{2}\right)^{2/3}\right] ,
3023: %\label{rhoz}
3024: %\eeq
3025:
3026:
3027:
3028:
3029:
3030:
3031:
3032:
3033:
3034: \section{Half-range expansion and useful integrals}
3035: \label{Half-range}
3036:
3037:
3038: We show in this appendix how to obtain the solution (\ref{phiint}) to the
3039: half-range expansion problem (\ref{phiW})-(\ref{phix0}). The same method also
3040: allows us to derive other useful identities that we need to perform the
3041: calculations presented in this article. Thus, we consider
3042: the function $f(p)$ of the complex variable $p$ defined by
3043: \beq
3044: s \geq 0 , \;\;\; u \geq 0 , \;\;\; |\Arg(p)|<\pi : \;\;\;
3045: f(p)= p^{-1/6} \, e^{\frac{2}{3}s^{3/2}/p} \,
3046: \Ai\left[p^{1/3} u +\frac{s}{p^{2/3}}\right] .
3047: \label{fdef}
3048: \eeq
3049: This function is regular over the complex plane except for a branch cut along
3050: the negative real axis. Moreover, from the asymptotic behavior (\ref{Aiinf})
3051: of the Airy function we obtain
3052: \beq
3053: u>0 : \;\; f(p) \sim \frac{u^{-1/4}}{2\sqrt{\pi}} \, p^{-1/4} \,
3054: e^{-\frac{2}{3} u^{3/2} p^{1/2}} \;\; \mbox{as} \;\;
3055: |p|\rightarrow\infty \;\; \mbox{with} \;\; |\Arg(p)|<\pi ,
3056: \label{fpasymp1}
3057: \eeq
3058: and
3059: \beq
3060: u=0 : \;\; f(p) \sim \Ai(0) \, p^{-1/6} \;\; \mbox{as} \;\;
3061: |p|\rightarrow\infty \;\; \mbox{with} \;\; |\Arg(p)|<\pi .
3062: \label{fpasymp2}
3063: \eeq
3064: Next, we introduce the general integral
3065: $F_{k,\ell}(\nu_1,..,\nu_k;\lambda_1,..,\lambda_{\ell})$ defined by
3066: \beq
3067: F_{k,\ell}(\nu_i;\lambda_j) = \int_{c-\ii\infty}^{c+\ii\infty}
3068: \frac{\dd p}{2\pi\ii} \, \frac{f(p)}{\prod_{i=1}^k (p-\nu_i^3) \,
3069: \prod_{j=1}^{\ell} (p+\lambda_j^3)} \;\;\;\; \mbox{with} \;\;\;\;
3070: c>\max_i\{\nu_i^3\} ,
3071: \label{fIdef}
3072: \eeq
3073: with the conditions
3074: \beq
3075: \nu_i>0, \;\; \nu_i\neq\nu_{i'} \;\; \mbox{for} \;\;i\neq i' ; \;\;\;
3076: \lambda_j>0 , \;\; \lambda_j\neq\lambda_{j'} \;\; \mbox{for} \;\; j\neq j' ;
3077: \;\;\; k+\ell \geq 1 .
3078: \eeq
3079: If $k=0$ or $\ell=0$ one of the products in Eq.(\ref{fIdef}) is removed and
3080: replaced by a factor $1$. If $k=0$ the contour in Eq.(\ref{fIdef}) again runs
3081: to the right of all singularities, that is $c>0$.
3082:
3083: Then, from the asymptotics (\ref{fpasymp1})-(\ref{fpasymp2}), we can see that
3084: we can push the contour in Eq.(\ref{fIdef}) to the right, as
3085: $c\rightarrow+\infty$, which shows that $F_{k,\ell}=0$ (using $k+\ell \geq 1$).
3086: On the other hand, by pushing the contour to the
3087: left, using again the asymptotics (\ref{fpasymp1})-(\ref{fpasymp2}), we can see
3088: that $F_{k,\ell}$ is the sum of the $k$ residues at $p=\nu_i^3$ and of the
3089: contribution associated with the branch cut along the negative real axis.
3090: This yields
3091: \beq
3092: 0 = \sum_{i=1}^k \frac{\nu_i^{-\frac{1}{2}} \, e^{\frac{2}{3}s^{3/2}\nu_i^{-3}}
3093: \, \Ai\left[\nu_i u +\frac{s}{\nu_i^2}\right]}{\prod_{j\neq i} (\nu_i^3-\nu_j^3)
3094: \, \prod_j (\nu_i^3+\lambda_j^3)} + \int_{\cal C} \frac{\dd p}{2\pi\ii}
3095: \, \frac{p^{-\frac{1}{6}} \, e^{\frac{2}{3}s^{3/2}/p} \,
3096: \Ai\left[p^{1/3} u +\frac{s}{p^{2/3}}\right] }{\prod_j (p-\nu_j^3) \,
3097: \prod_j (p+\lambda_j^3)} ,
3098: \eeq
3099: where ${\cal C}$ is the anticlockwise Hankel contour that bends around the
3100: negative real axis. Then, pushing the contour towards both sides of the negative
3101: real axis, making the change of variable $p=-\mu^3 \pm \ii \epsilon$ with
3102: $\mu>0$ and $\epsilon\rightarrow 0^+$, and using
3103: \cite{Abramowitz}
3104: \beq
3105: \Ai\left[ e^{\pm \ii 2\pi/3} x\right] = \frac{1}{2} \,
3106: e^{\pm \ii \pi/3} \left[ \Ai(x) \mp \ii \,\Bi(x) \right] ,
3107: \label{AiBi1}
3108: \eeq
3109: as well as the Sokhatsky-Weierstrass theorem, written here in concise form as
3110: \beq
3111: \lim_{\epsilon\rightarrow 0^+} \frac{1}{\mu^3-\nu^3+\ii\epsilon}
3112: = \pv \frac{1}{\mu^3-\nu^3} - \frac{\ii\pi}{3\mu^2} \delta(\mu-\nu) ,
3113: \label{Weierstrass}
3114: \eeq
3115: we obtain
3116: \beqa
3117: \!u\geq 0 : \;\; \pv \!\int_0^{\infty} \frac{\dd\mu}{2\pi} \, \frac{3\mu^{3/2} \,
3118: e^{-\frac{2}{3} s^{3/2} \mu^{-3}} \, \Ai\left[-\mu u+\frac{s}{\mu^2}\right]}
3119: {\prod_{j=1}^k (\mu^3+\nu_j^3) \, \prod_{j=1}^{\ell} (\mu^3-\lambda_j^3)}
3120: & = &
3121: \sum_{i=1}^k \frac{\nu_i^{-\frac{1}{2}} \, e^{\frac{2}{3}s^{3/2}\nu_i^{-3}}
3122: \, \Ai\left[\nu_i u +\frac{s}{\nu_i^2}\right]}{\prod_{j\neq i} (-\nu_i^3+\nu_j^3)
3123: \, \prod_j (-\nu_i^3-\lambda_j^3)}
3124: \nonumber \\
3125: && \hspace{-3cm} + \frac{1}{2} \sum_{i=1}^{\ell}
3126: \frac{\lambda_i^{-\frac{1}{2}} \, e^{-\frac{2}{3}s^{3/2}\lambda_i^{-3}}
3127: \, \Bi\left[-\lambda_i u +\frac{s}{\lambda_i^2}\right]}
3128: {\prod_j (\lambda_i^3+\nu_j^3) \, \prod_{j\neq i} (\lambda_i^3-\lambda_j^3)} .
3129: \label{phi0int}
3130: \eeqa
3131: Here, the symbol $\pv$ stands for the Cauchy principal value and must be understood
3132: with respect to $\mu^3$ (rather than $\mu$). That is, the integrals are regularized
3133: by cutting around each pole $\lambda_j$ the interval $[\mu_-,\mu_+]$, with
3134: $\mu^3_{\pm}=\lambda_j^3\pm\epsilon$, which is symmetric in terms of $\mu^3$
3135: around $\lambda_j^3$, and taking the limit $\epsilon\rightarrow 0^+$. If $\ell=0$
3136: the integral is regular and there is no need to introduce the principal value.
3137: In particular, the case $k=1$ and $\ell=0$ yields, with $\nu>0$,
3138: \beq
3139: u\geq 0 : \;\; \int_0^{\infty} \frac{\dd\mu}{2\pi} \,
3140: \frac{3\mu^{3/2}}{\mu^3+\nu^3} \, e^{-\frac{2}{3} s^{3/2} \mu^{-3}}
3141: \, \Ai\left[-\mu u+\frac{s}{\mu^2}\right]
3142: = \nu^{-1/2} \, e^{\frac{2}{3}s^{3/2}\nu^{-3}} \,
3143: \Ai\left[\nu u +\frac{s}{\nu^2}\right] .
3144: \label{phi0intk1l0}
3145: \eeq
3146: This implies that $\phi_{s,\nu}(r,u)$ defined by Eq.(\ref{phiint}) is
3147: the solution of the form (\ref{phiW}) that satisfies the constraint
3148: (\ref{phix0}). Note that the restriction to $u\geq 0$ is essential.
3149: For instance, as $\Arg(u)$ grows from $0$, the function $f(p)$ displays
3150: an exponential growth for $\pi-3\Arg(u)<\Arg(p)<\pi$ and we can no longer
3151: bend the integration contour onto the negative real axis.
3152:
3153: Alternatively, as in \cite{Burkhardt1993}, we can obtain Eq.(\ref{phi0intk1l0})
3154: from the analysis of \cite{Marshall1985} and \cite{Marshall1987},
3155: or of \cite{Hagan1989}, who studied several problems associated with the
3156: Klein-Kramers equation, by taking the limit of zero friction. However, these
3157: problems lead to discrete spectra and require a sophisticated analysis
3158: that involves infinite products to handle the poles associated with all
3159: eigenvalues.
3160:
3161: Finally, making the change $k\rightarrow k+1$ and taking the limit
3162: $\nu_{k+1}\rightarrow 0^+$ in Eq.(\ref{phi0int}) gives the useful identity
3163: \beqa
3164: \hspace{0cm} u\geq 0 : \;\; \pv \int_0^{\infty} \frac{\dd\mu}{2\pi} \,
3165: \frac{3\mu^{-3/2} \,
3166: e^{-\frac{2}{3} s^{3/2} \mu^{-3}} \, \Ai\left[-\mu u+\frac{s}{\mu^2}\right]}
3167: {\prod_{j=1}^k (\mu^3+\nu_j^3) \, \prod_{j=1}^{\ell} (\mu^3-\lambda_j^3)}
3168: & = & \frac{s^{-\frac{1}{4}} \, e^{-\sqrt{s}u}}{2\sqrt{\pi} \,
3169: \prod_j \nu_j^3 \, \prod_j (-\lambda_j^3)}
3170: \nonumber \\
3171: && \hspace{-6.3cm} - \sum_{i=1}^k \frac{\nu_i^{-\frac{7}{2}} \,
3172: e^{\frac{2}{3}s^{3/2}\nu_i^{-3}} \, \Ai\left[\nu_i u +\frac{s}{\nu_i^2}\right]}
3173: {\prod_{j\neq i} (-\nu_i^3+\nu_j^3) \, \prod_j (-\nu_i^3-\lambda_j^3)}
3174: + \frac{1}{2} \sum_{i=1}^{\ell}
3175: \frac{\lambda_i^{-\frac{7}{2}} \, e^{-\frac{2}{3}s^{3/2}\lambda_i^{-3}}
3176: \, \Bi\left[-\lambda_i u +\frac{s}{\lambda_i^2}\right]}
3177: {\prod_j (\lambda_i^3+\nu_j^3) \, \prod_{j\neq i} (\lambda_i^3-\lambda_j^3)} .
3178: \label{Expintkl}
3179: \eeqa
3180: Equation~(\ref{Expintkl}) now applies to any $k\geq 0, \ell\geq 0$. In particular,
3181: the case $k=\ell=0$ yields
3182: \beq
3183: u\geq 0 : \;\;\; e^{-\sqrt{s} u} = s^{1/4} \int_0^{\infty}
3184: \frac{\dd\mu}{\sqrt{\pi}} \, 3\mu^{-3/2} \, e^{-\frac{2}{3} s^{3/2} \mu^{-3}}
3185: \, \Ai\left[-\mu u+\frac{s}{\mu^2}\right] .
3186: \label{Expu_int}
3187: \eeq
3188: Note that $s$ can be absorbed in Eq.(\ref{Expu_int}) through the change of
3189: variables $v=\sqrt{s} u$ and $\nu=\mu/\sqrt{s}$.
3190: By letting $m$ parameters $\nu_i$ going to zero in a sequential manner, we could
3191: derive a series of similar identities for integrals of the form of
3192: (\ref{Expintkl}) with a prefactor $\mu^{3/2-3m}$ for any $m\geq 0$. However,
3193: in this article we do not need to go beyond $m=1$ as in
3194: Eqs.(\ref{Expintkl})-(\ref{Expu_int}).
3195:
3196: In a similar fashion, we now consider the function $\hf(p)$ of the complex
3197: variable $p$ defined for $s_i \geq 0, u_i\geq 0$, by
3198: \beq
3199: \hf(p)= e^{\frac{2}{3}(s_1^{3/2}-s_2^{3/2})/p} \,
3200: \Ai\,'\left[p^{1/3} u_1 +\frac{s_1}{p^{2/3}}\right] \,
3201: \Ai\left[e^{-\ii \pi/3} p^{1/3} u_2
3202: + e^{\ii 2\pi/3}\frac{s_2}{p^{2/3}}\right] .
3203: \label{hfdef}
3204: \eeq
3205: It has a branch cut along the negative real axis and it grows as most as
3206: $p^{1/12}$ for $|p|\rightarrow\infty$ and $0< \Arg(p)<\pi$.
3207: Then, in a manner similar to Eq.(\ref{fIdef}), we introduce the integral
3208: $\hF_{k,\ell}(\nu_1,..,\nu_k;\lambda_1,..,\lambda_{\ell})$ defined by
3209: \beq
3210: \hF_{k,\ell}(\nu_i;\lambda_j) = \int_{-\infty+\ii c}^{\infty+\ii c}
3211: \frac{\dd p}{2\pi} \, \frac{\hf(p)}{\prod_{i=1}^k (p-\nu_i^3) \,
3212: \prod_{j=1}^{\ell} (p+\lambda_j^3)} \;\;\;\;
3213: \mbox{with} \;\;\;\; c > 0 ,
3214: \label{hfIdef}
3215: \eeq
3216: with the conditions
3217: \beq
3218: \nu_i>0, \;\; \nu_i\neq\nu_{i'} \;\; \mbox{for} \;\;i\neq i' ; \;\;\;
3219: \lambda_j>0 , \;\; \lambda_j\neq\lambda_{j'} \;\; \mbox{for} \;\; j\neq j' ;
3220: \;\;\; k+\ell \geq 2 .
3221: \eeq
3222: Note that the integration contour is now parallel to the real axis, in the upper
3223: half-plane, and that the asymptotic behavior of $\hf(p)$ for large
3224: $|p|$, with $0< \Arg(p)<\pi$, now requires $k+\ell\geq 2$. Pushing the contour
3225: upward, as $c\rightarrow+\infty$, we can see that $\hF_{k,\ell}=0$.
3226: Next, pushing the contour towards the real axis, by making the change of variable
3227: $p=\pm\mu^3 + \ii \epsilon$ with $\mu>0$ and $\epsilon\rightarrow 0^+$,
3228: and using Eq.(\ref{AiBi1}), that also yields
3229: \beq
3230: \Ai\,'\left[ e^{\pm \ii 2\pi/3} x\right] = \frac{1}{2} \,
3231: e^{\mp \ii \pi/3} \left[ \Ai\,'(x) \mp \ii \,\Bi\,'(x) \right] ,
3232: \label{AipBip1}
3233: \eeq
3234: as well as Eq.(\ref{Weierstrass}), we obtain after taking the real part
3235: \beqa
3236: \hspace{0cm}u_i\geq 0 : \;\; \pv \int_{-\infty}^{\infty} \frac{\dd\mu}{2\pi} \,
3237: \frac{3\mu^2 \, e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\mu^{-3}} \,
3238: \Ai\,'\left[-\mu u_1+\frac{s_1}{\mu^2}\right]
3239: \Ai\left[\mu u_2+\frac{s_2}{\mu^2}\right]}
3240: {\prod_{j=1}^k (\mu^3+\nu_j^3) \, \prod_{j=1}^{\ell} (\mu^3-\lambda_j^3)}
3241: & = & \nonumber \\
3242: && \hspace{-6cm} - \frac{1}{2} \sum_{i=1}^k
3243: \frac{e^{\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\nu_i^{-3}} \,
3244: \Ai\,'\left[\nu_i u_1 +\frac{s_1}{\nu_i^2}\right]
3245: \Bi\left[-\nu_i u_2 +\frac{s_2}{\nu_i^2}\right]}
3246: {\prod_{j\neq i} (-\nu_i^3+\nu_j^3) \, \prod_j (-\nu_i^3-\lambda_j^3)} \nonumber \\
3247: && \hspace{-6cm} + \frac{1}{2} \sum_{i=1}^{\ell}
3248: \frac{e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\lambda_i^{-3}} \,
3249: \Bi\,'\left[-\lambda_i u_1 +\frac{s_1}{\lambda_i^2}\right]
3250: \Ai\left[\lambda_i u_2 +\frac{s_2}{\lambda_i^2}\right]}
3251: {\prod_j (\lambda_i^3+\nu_j^3) \, \prod_{j\neq i} (\lambda_i^3-\lambda_j^3)} .
3252: \label{intAipAi1}
3253: \eeqa
3254: Taking the imaginary part gives another identity, that involves the integral
3255: over $\mu$ of products $\Bi\,'\Ai$ and $\Ai\,'\Bi$, which we do not need for
3256: the present calculations. Again, in Eq.(\ref{intAipAi1}) the Cauchy principal
3257: value is understood with respect to $\mu^3$.
3258: Exchanging the derivative in expression
3259: (\ref{hfdef}), we obtain an identity similar to Eq.(\ref{intAipAi1}) where
3260: the derivatives are exchanged:
3261: \beqa
3262: \hspace{0cm}u_i\geq 0 : \;\; \pv \int_{-\infty}^{\infty} \frac{\dd\mu}{2\pi} \,
3263: \frac{3\mu^2 \, e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\mu^{-3}} \,
3264: \Ai\left[-\mu u_1+\frac{s_1}{\mu^2}\right]
3265: \Ai\,'\left[\mu u_2+\frac{s_2}{\mu^2}\right]}
3266: {\prod_{j=1}^k (\mu^3+\nu_j^3) \, \prod_{j=1}^{\ell} (\mu^3-\lambda_j^3)}
3267: & = & \nonumber \\
3268: && \hspace{-6cm} - \frac{1}{2} \sum_{i=1}^k
3269: \frac{e^{\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\nu_i^{-3}} \,
3270: \Ai\left[\nu_i u_1 +\frac{s_1}{\nu_i^2}\right]
3271: \Bi\,'\left[-\nu_i u_2 +\frac{s_2}{\nu_i^2}\right]}
3272: {\prod_{j\neq i} (-\nu_i^3+\nu_j^3) \, \prod_j (-\nu_i^3-\lambda_j^3)}
3273: \nonumber \\
3274: && \hspace{-6cm} + \frac{1}{2} \sum_{i=1}^{\ell}
3275: \frac{e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\lambda_i^{-3}} \,
3276: \Bi\left[-\lambda_i u_1 +\frac{s_1}{\lambda_i^2}\right]
3277: \Ai\,'\left[\lambda_i u_2 +\frac{s_2}{\lambda_i^2}\right]}
3278: {\prod_j (\lambda_i^3+\nu_j^3) \, \prod_{j\neq i} (\lambda_i^3-\lambda_j^3)} .
3279: \label{intAiAip1}
3280: \eeqa
3281: Next, making the change $k\rightarrow k+1$ and taking the limit
3282: $\nu_{k+1}\rightarrow 0$ gives, for $k+\ell\geq 1$,
3283: \beqa
3284: \hspace{0cm}u_i\geq 0 : \;\; \pv \! \int_{-\infty}^{\infty} \frac{\dd\mu}{2\pi} \,
3285: \frac{3\mu^{-1} \, e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\mu^{-3}}}
3286: {\prod_j (\mu^3\!+\!\nu_j^3) \, \prod_j (\mu^3\!-\!\lambda_j^3)}
3287: \left[ \Ai\,'\bigl(-\mu u_1\!+\!\frac{s_1}{\mu^2}\bigr)
3288: \Ai\bigl(\mu u_2\!+\!\frac{s_2}{\mu^2}\bigr) - \Ai \, \Ai\,' \right]
3289: & = & \nonumber \\
3290: && \hspace{-11cm} \frac{1}{4\pi} \left[ \bigl(\frac{s_1}{s_2}\bigr)^{1/4}
3291: +\bigl(\frac{s_1}{s_2}\bigr)^{-1/4} \right]
3292: \frac{e^{-\sqrt{s_1}u_1-\sqrt{s_2}u_2}}{\prod_j \nu_j^3 \prod_j (-\lambda_j^3)}
3293: \nonumber \\
3294: && \hspace{-11cm} + \frac{1}{2} \sum_{i=1}^k
3295: \frac{\nu_i^{-3}\, e^{\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\nu_i^{-3}}}
3296: {\prod_{j\neq i} (-\nu_i^3\!+\!\nu_j^3) \, \prod_j (-\nu_i^3\!-\!\lambda_j^3)}
3297: \left[ \Ai\,'\bigl(\nu_i u_1\!+\!\frac{s_1}{\nu_i^2}\bigr)
3298: \Bi\bigl(-\nu_i u_2\!+\!\frac{s_2}{\nu_i^2}\bigr)
3299: - \Ai \, \Bi\,' \right]
3300: \nonumber \\
3301: && \hspace{-11cm} + \frac{1}{2} \sum_{i=1}^{\ell}
3302: \frac{\lambda_i^{-3} \, e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\lambda_i^{-3}}}
3303: {\prod_j (\lambda_i^3\!+\!\nu_j^3) \, \prod_{j\neq i} (\lambda_i^3\!-\!\lambda_j^3)}
3304: \left[ \Bi\,'\bigl(-\lambda_i u_1\!+\!\frac{s_1}{\lambda_i^2}\bigr)
3305: \Ai\bigl(\lambda_i u_2\!+\!\frac{s_2}{\lambda_i^2}\bigr) - \Bi \, \Ai\,' \right] .
3306: \label{intmu1AipAi1}
3307: \eeqa
3308: Here we combined both Equations (\ref{intAipAi1})-(\ref{intAiAip1}), and in
3309: each bracket the second product, such as $\Ai \, \Ai\,'$, is equal to the first
3310: product where we exchange the derivative. Next, for the case $u_1=u_2=0$,
3311: we again make the change $k\rightarrow k+1$ and take the limit
3312: $\nu_{k+1}\rightarrow \infty$. This yields for any $k\geq 0$, $\ell\geq 0$,
3313: \beqa
3314: \pv \int_{-\infty}^{\infty} \frac{\dd\mu}{2\pi} \, \frac{3\mu^{-1} \,
3315: e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\mu^{-3}}}
3316: {\prod_j (\mu^3\!+\!\nu_j^3) \, \prod_j (\mu^3\!-\!\lambda_j^3)}
3317: \left[ \Ai\,'\bigl(\frac{s_1}{\mu^2}\bigr) \Ai\bigl(\frac{s_2}{\mu^2}\bigr)
3318: - \Ai \, \Ai\,' \right] & = & - \frac{\delta_{k+\ell,0}}{2\pi} \nonumber \\
3319: && \hspace{-8cm} + \frac{\bigl(\frac{s_1}{s_2}\bigr)^{1/4}
3320: +\bigl(\frac{s_1}{s_2}\bigr)^{-1/4}}{4\pi \prod_j \nu_j^3 \prod_j (-\lambda_j^3)}
3321: + \frac{1}{2} \sum_{i=1}^k
3322: \frac{\nu_i^{-3}\, e^{\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\nu_i^{-3}}}
3323: {\prod_{j\neq i} (-\nu_i^3\!+\!\nu_j^3) \, \prod_j (-\nu_i^3\!-\!\lambda_j^3)}
3324: \left[ \Ai\,'\bigl(\frac{s_1}{\nu_i^2}\bigr) \Bi\bigl(\frac{s_2}{\nu_i^2}\bigr)
3325: - \Ai \, \Bi\,' \right] \nonumber \\
3326: && \hspace{-8cm} + \frac{1}{2} \sum_{i=1}^{\ell}
3327: \frac{\lambda_i^{-3} \, e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\lambda_i^{-3}}}
3328: {\prod_j (\lambda_i^3\!+\!\nu_j^3) \, \prod_{j\neq i} (\lambda_i^3\!-\!\lambda_j^3)}
3329: \left[ \Bi\,'\bigl(\frac{s_1}{\lambda_i^2}\bigr)
3330: \Ai\bigl(\frac{s_2}{\lambda_i^2}\bigr) - \Bi \, \Ai\,' \right] ,
3331: \label{intmu1AipAiui01}
3332: \eeqa
3333: where we used the Wronskian property (\ref{Wronskian})
3334: and $\delta_{k+\ell,0}$ is the Kronecker symbol. In particular, for $k=\ell=0$
3335: we obtain
3336: \beq
3337: \int_{-\infty}^{\infty} \frac{\dd\mu}{\mu} \,
3338: e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})\mu^{-3}}
3339: \left[ \Ai\,'\bigl(\frac{s_1}{\mu^2}\bigr) \Ai\bigl(\frac{s_2}{\mu^2}\bigr)
3340: - \Ai \, \Ai\,' \right] = \frac{1}{6} \left[ \bigl(\frac{s_1}{s_2}\bigr)^{1/4}
3341: + \bigl(\frac{s_1}{s_2}\bigr)^{-1/4} - 2 \right] .
3342: \label{intmu1AipAiui02}
3343: \eeq
3344: Since the integral is convergent at $\mu=0$ there is no need to
3345: use the principal value.
3346:
3347:
3348:
3349: \section{Computation of the two-point distribution}
3350: \label{Computation-of-two-point}
3351:
3352: We present here the computation of the two-point distribution (\ref{PX1X2})
3353: from the two contributions $p^>$ and $p^<$ described in
3354: Figs.~\ref{figP2s} and \ref{figP2m}.
3355:
3356: Let us first consider the contribution $p^>$.
3357: In a fashion similar to Eq.(\ref{pxcq1}), using the Markovian character of the
3358: process $q\mapsto\{\psi,v\}$, it reads as
3359: \beqa
3360: p_{x_1,x_2}^>(0 \leq q_1' \leq q_1,c_1; q_2' \geq q_2) \dd c_1 & = &
3361: \lim_{q_{\pm}\rightarrow\pm\infty} \int \dd\psi_-\dd v_-
3362: K_{x_1,c_1}(0,0,0;q_-,\psi_-,v_-) \nonumber \\
3363: && \hspace{-5cm} \times \int \dd\psi_1\dd v_1
3364: [ K_{x_1,c_1}(0,0,0;q_1,\psi_1,v_1) - K_{x_1,c_1+\dd c_1}(0,0,0;q_1,\psi_1,v_1) ]
3365: \nonumber \\
3366: && \hspace{-5cm} \times \left\{ \int \dd\psi_+\dd v_+
3367: K_{x_1,c_1}(q_1,\psi_1,v_1;q_+,\psi_+,v_+)
3368: - \int \dd\psi_2\dd v_2 K_{x_1,c_1}(q_1,\psi_1,v_1;q_2,\psi_2,v_2) \right.
3369: \nonumber \\
3370: && \hspace{-4cm} \times \left. \int \dd\psi_+\dd v_+
3371: K_{x_2,c_2}(q_2,\psi_2,v_2;q_+,\psi_+,v_+) \right\} .
3372: \label{px1x2cq1}
3373: \eeqa
3374: Then, we recognize $p_{x_1}(0\leq q_1'\leq q_1)$ in the contribution associated
3375: with the first term in the bracket, and we can write
3376: \beq
3377: p_{x_1,x_2}^>(0 \leq q_1' \leq q_1; q_2' \geq q_2) = p_{x_1}(0\leq q_1'\leq q_1)
3378: - \hp_{x_1,x_2}^>(0 \leq q_1' \leq q_1; q_2' \geq q_2) ,
3379: \label{hp>def}
3380: \eeq
3381: where we introduced the remaining part
3382: \beqa
3383: \hp_{x_1,x_2}^>(0 \leq q_1' \leq q_1,c_1; q_2' \geq q_2) \dd c_1 & = &
3384: \lim_{q_{\pm}\rightarrow\pm\infty} \int \dd\psi_-\dd v_- \dd\psi_1\dd v_1
3385: \dd\psi_2\dd v_2 \dd\psi_+\dd v_+ \nonumber \\
3386: && \hspace{-5cm} \times K_{x_1,c_1}(0,0,0;q_-,\psi_-,v_-)
3387: [ K_{x_1,c_1}(0,0,0;q_1,\psi_1,v_1) - K_{x_1,c_1+\dd c_1}(0,0,0;q_1,\psi_1,v_1) ]
3388: \nonumber \\
3389: && \hspace{-5cm} \times K_{x_1,c_1}(q_1,\psi_1,v_1;q_2,\psi_2,v_2)
3390: K_{x_2,c_2}(q_2,\psi_2,v_2;q_+,\psi_+,v_+) .
3391: \label{hp>1}
3392: \eeqa
3393: In Eq.(\ref{hp>def}) we have also integrated $\hp^>$ over $c_1$.
3394:
3395: We can note that $p_{x_1,x_2}^>$ and $\hp_{x_1,x_2}^>$ satisfy the following
3396: boundary conditions. First, taking the derivative with respect to $q_1$
3397: to obtain the probability density $p_{x_1,x_2}^>(q_1;q_2'\geq q_2)$, we have
3398: \beq
3399: \lim_{q_2\rightarrow q_1^+} p_{x_1,x_2}^>(q_1;q_2'\geq q_2) \rightarrow
3400: p_{x_1}(q_1) , \;\;\; \mbox{whence} \;\;\; \lim_{q_2\rightarrow q_1^+}
3401: \hp_{x_1,x_2}^>(q_1;q_2'\geq q_2) = 0 .
3402: \label{limp>q2q1}
3403: \eeq
3404: Indeed, following the discussion at the beginning of section~\ref{subsec:general},
3405: if $\psi_0(q)$ is tangent to
3406: $\cP_{x_1,c_1}$ at $q_1$, the second parabola $\cP_{x_2,c_2}$ can only cross
3407: the first one at a point $q_*>q_1$ (otherwise, $\psi_0'$ being continuous,
3408: if we had $q_*=q_1$ the curve $\psi_0(q)$ would go below $\cP_{x_2,c_2}$ just beyond
3409: $q_1$). Then, all curves tangent to $\cP_{x_1,c_1}$ at $q_1$ satisfy both
3410: properties $q_*>q_1$ and $q_2>q_1$, so that they are all included in the contribution
3411: $p_{x_1,x_2}^>$ as we take the limit $q_2\rightarrow q_1^+$ and we must recover
3412: $p_{x_1}(q_1)$, as stated in (\ref{limp>q2q1}).
3413: Second, for large $q_2$ we obviously have the asymptotics
3414: \beq
3415: \lim_{q_2\rightarrow +\infty} p_{x_1,x_2}^>(0 \leq q_1' \leq q_1; q_2' \geq q_2)
3416: = 0 , \;\;\; \lim_{q_2\rightarrow +\infty}
3417: \hp_{x_1,x_2}^>(0 \leq q_1' \leq q_1; q_2' \geq q_2) = p_{x_1}(q_1) .
3418: \label{limp>q2inf}
3419: \eeq
3420: This latter constraint can be directly checked on Eq.(\ref{hp>1}).
3421:
3422: Using the transformations (\ref{cKdef}) and (\ref{Gdef}), we obtain
3423: \beqa
3424: \hp_{x_1,x_2}^>(0 \leq q_1' \leq q_1; q_2' \geq q_2) & = &
3425: e^{\frac{\hu_{21}}{\gam}-\frac{q_2}{\gam^2}} \int \dd r_1 \dd u_1
3426: \dd r_2 \dd u_2 \dd r_3 \, \Hi(r_3,\hu_1) \nonumber \\
3427: && \hspace{-3cm} \times \Delta(q_1;r_3,-\hu_1;r_1,u_1)
3428: G(q_2-q_1;r_1,u_1;r_2,u_2+\hu_{21}) \Hi(r_2,u_2) ,
3429: \label{hp>3}
3430: \eeqa
3431: where we introduced as in (\ref{hudef}) the quantities
3432: \beq
3433: \hu_i = \sqrt{\frac{2}{D}} \frac{x_i}{t} , \;\;\;
3434: \hu_{21} = \hu_2 - \hu_1 .
3435: \label{hudef21}
3436: \eeq
3437: Next, taking the derivative with respect to $q_1$, using the backward
3438: equation (\ref{Gbackward}) for $G$ and the forward equation (\ref{Gdiff})
3439: that is also satisfied by $\Delta$, we obtain a total differential over $r_1$,
3440: which only leaves the boundary term at $r_1=0$:
3441: \beqa
3442: \hp_{x_1,x_2}^>(q_1; q_2' \geq q_2) & = &
3443: e^{\frac{\hu_{21}}{\gam}-\frac{q_2}{\gam^2}} \int \dd u_1
3444: \dd r_2 \dd u_2 \dd r_3 \, \Hi(r_3,\hu_1) u_1 \Delta(q_1;r_3,-\hu_1;0,u_1)
3445: \nonumber \\
3446: && \hspace{2cm} \times G(q_2-q_1;0,u_1;r_2,u_2+\hu_{21}) \Hi(r_2,u_2) .
3447: \label{hp>4}
3448: \eeqa
3449: We can check from the explicit expressions of $\Delta$ and $G$ that the
3450: integrations by parts leading to Eq.(\ref{hp>4}) are valid. The expression
3451: (\ref{hp>4}) clearly satisfies the property (\ref{limp>q2q1}).
3452: Indeed, for $q_2\rightarrow q_1^+$ the factor $G$ implies $r_2\rightarrow 0$,
3453: using the first boundary condition (\ref{Gboundary}), which in turns
3454: leads to $u_2+\hu_{21}\leq 0$ because of the second boundary condition in
3455: (\ref{Gboundary}). However, the last factor $\Hi$ also implies $u_2\geq 0$,
3456: using the boundary condition (\ref{Gboundary1}) applied to $\Hi$.
3457: Since $\hu_{21}>0$ both constraints on $u_2$ cannot be simultaneously satisfied
3458: which leads to (\ref{limp>q2q1}). This can also be checked on Eq.(\ref{hp>4})
3459: using the explicit expressions of $\Delta$ and $\Hi$.
3460:
3461: Then, using the explicit expressions of $G,\Delta$, and $\Hi$, and the results
3462: of Appendices \ref{Airy} and \ref{Half-range}, it is possible to greatly simplify
3463: Eq.(\ref{hp>4}).
3464: Indeed, the integrals over $r_i$ are immediate (they only involve factors of the
3465: form $e^{-\nu^3 r}$) whereas the integral over $u_1$ can be transformed using
3466: Eq.(\ref{IntAi5}). Next, integrals over $u_i$ are typically split over
3467: $u_i\leq 0$ and $u_i\geq 0$, and each factor of the form $\Ai(-\mu u_i + s/\mu^2)$
3468: with $u_i\geq 0$, or $\Ai(\mu u_i + s/\mu^2)$ with $u_i\leq 0$, can be integrated
3469: over $\mu$ using the results (\ref{phi0int})-(\ref{Expu_int}) of
3470: Appendix~\ref{Half-range}. This leads to products of the form
3471: $\Ai(\mu u_i + s_1/\mu^2) \Ai(\mu u_i + s_2/\mu^2)$ that can be integrated
3472: over $u_i$ using the primitive (\ref{prim1}), that also extends to the second
3473: Airy function $\Bi$. Then, these terms can be further simplified using
3474: the Wronskian (\ref{Wronskian}) and the results
3475: (\ref{intAipAi1})-(\ref{intmu1AipAiui02}) of Appendix~\ref{Half-range}.
3476: We eventually obtain in terms of dimensionless variables
3477: \beq
3478: \hP_{X_1,X_2}^>(Q_1,Q_2'\geq Q_2) = \inta \frac{\dd s_1 \dd s_2}{(2\pi\ii)^2} \,
3479: e^{(s_1-1)Q_1+(s_2-1)Q_{21}} J(s_1,2X_1) I(s_2,2 X_{21}) ,
3480: \label{hP>5}
3481: \eeq
3482: with
3483: \beq
3484: Q_1 \geq 0, \;\; Q_{21}=Q_2-Q_1\geq 0 , \;\; \mbox{and} \;\;
3485: I(s,2X) = \frac{1}{s-1} \, e^{-(\frac{2}{3}s^{3/2}-s+\frac{1}{3})2X/(s-1)} .
3486: \label{Is2Xdef}
3487: \eeq
3488: We can check that Eq.(\ref{hP>5}) agrees with both constraints
3489: (\ref{limp>q2q1})-(\ref{limp>q2inf}). Then, taking the derivative with respect
3490: to $Q_2$ we obtain the full probability density associated with $q_*>q_2$ as
3491: \beq
3492: P_{X_1,X_2}^>(Q_1,Q_2) = \inta
3493: \frac{\dd s_1 \dd s_2}{(2\pi\ii)^2} \, e^{(s_1-1)Q_1+(s_2-1)Q_{21}} J(s_1,2X_1)
3494: (s_2-1) I(s_2,2 X_{21}) ,
3495: \label{P>6}
3496: \eeq
3497: since the derivative with respect to $Q_2$ of the first term in the right hand
3498: side of (\ref{hp>def}) vanishes.
3499:
3500: We now consider the second contribution, $p^<$, associated with the intersection
3501: $q_*$ between both parabolas, $\cP_{x_1,c_1}$ and $\cP_{x_2,c_2}$, being in the
3502: range $q_1<q_*<q_2$. Proceeding as for (\ref{pxcq1}) and (\ref{px1x2cq1}) it
3503: reads as
3504: \beqa
3505: p_{x_1,x_2}^<(0 \leq q_1' \leq q_1,c_1; q_2' \geq q_2,c_2) \dd c_1 \dd c_2 & = &
3506: \lim_{q_{\pm}\rightarrow\pm\infty} \int \dd\psi_-\dd v_-
3507: K_{x_1,c_1}(0,0,0;q_-,\psi_-,v_-) \nonumber \\
3508: && \hspace{-6cm} \times \int \dd\psi_1\dd v_1
3509: [ K_{x_1,c_1}(0,0,0;q_1,\psi_1,v_1) - K_{x_1,c_1+\dd c_1}(0,0,0;q_1,\psi_1,v_1) ]
3510: \nonumber \\
3511: && \hspace{-6cm} \times \int \dd\psi_*\dd v_*
3512: K_{x_1,c_1}(q_1,\psi_1,v_1;q_*,\psi_*,v_*) \int \dd\psi_2\dd v_2
3513: K_{x_2,c_2}(q_*,\psi_*,v_*;q_2,\psi_2,v_2)
3514: \nonumber \\
3515: && \hspace{-6cm} \times \int \dd\psi_+\dd v_+
3516: [ K_{x_2,c_2}(q_2,\psi_2,v_2;q_+,\psi_+,v_+) -
3517: K_{x_2,c_2+\dd c_2}(q_2,\psi_2,v_2;q_+,\psi_+,v_+) ] ,
3518: \label{px1x2<1}
3519: \eeqa
3520: which we must integrate over both $c_1$ and $c_2$. We can note that it satisfies
3521: the boundary conditions
3522: \beq
3523: \lim_{q_2\rightarrow q_1^+} p_{x_1,x_2}^<(q_1;q_2'\geq q_2) \rightarrow 0 ,
3524: \;\;\; \mbox{and} \;\;\; \lim_{q_2\rightarrow +\infty}
3525: p_{x_1,x_2}^<(0 \leq q_1' \leq q_1;q_2'\geq q_2) = 0 .
3526: \label{limp<q2q1}
3527: \eeq
3528: Using again Eqs.(\ref{cKdef}), (\ref{Gdef}), we obtain
3529: \beqa
3530: p_{x_1,x_2}^<(0 \leq q_1' \leq q_1; q_2' \geq q_2) & = &
3531: e^{\frac{\hu_{21}}{\gam}-\frac{q_2}{\gam^2}} \int \dd r_1 \dd u_1 \dd r_* \dd u_*
3532: \dd r_2 \dd u_2 \dd r_3 \dd r_4 \, \Hi(r_3,\hu_1) \nonumber \\
3533: && \hspace{-3cm} \times \Delta(q_1;r_3,-\hu_1;r_1,u_1)
3534: G(q_*-q_1;r_1,u_1;r_*,u_*) G(q_2-q_*;r_*,u_*-\hu_{21};r_2,u_2)
3535: \frac{\pl\Hi}{\pl r_2}(r_2,u_2) .
3536: \label{p<2}
3537: \eeqa
3538: Next, taking again the derivative with respect to $q_1$ and using the forward
3539: and backward equations (\ref{Gdiff}), (\ref{Gbackward}), gives
3540: \beqa
3541: p_{x_1,x_2}^<(q_1; q_2' \geq q_2) & = &
3542: e^{\frac{\hu_{21}}{\gam}-\frac{q_2}{\gam^2}} \int \dd u_1 \dd r_* \dd u_*
3543: \dd r_2 \dd u_2 \dd r_3 \dd r_4 \, \Hi(r_3,\hu_1) u_1 \Delta(q_1;r_3,-\hu_1;0,u_1)
3544: \nonumber \\
3545: && \hspace{0cm} \times G(q_*-q_1;0,u_1;r_*,u_*)
3546: G(q_2-q_*;r_*,u_*-\hu_{21};r_2,u_2) \frac{\pl\Hi}{\pl r_2}(r_2,u_2) .
3547: \label{p<3}
3548: \eeqa
3549: As for the derivation of Eq.(\ref{hP>5}), using the explicit expressions of
3550: $G,\Delta$, and $\Hi$, and the results of Appendices \ref{Airy} and
3551: \ref{Half-range}, as well as the property (\ref{closure}), we obtain
3552: \beqa
3553: \! P_{X_1,X_2}^<(Q_1,Q_2'\geq Q_2) & = & 2 X_{21} \, e^{2X_{21}-Q_2} \!
3554: \int_{Q_1}^{Q_2} \!\! \dd Q_* \inta \! \frac{\dd s_1 \dd s \dd s_2}{(2\pi\ii)^3} \,
3555: e^{s_1 Q_1+s(Q_*-Q_1)+s_2(Q_2-Q_*)} \nonumber \\
3556: && \hspace{0cm} \times J(s_1,2X_1) \frac{1}{s_2-1}
3557: [ L(1,s;2 X_{21}) - L(s_2,s;2 X_{21}) ] ,
3558: \label{P<4}
3559: \eeqa
3560: with
3561: \beq
3562: L(s_1,s_2;2X) = \frac{1}{2X} \, e^{-\frac{2}{3}(s_1^{3/2}-s_2^{3/2})2X/(s_1-s_2)} ,
3563: \; \mbox{whence} \;\;\; L(s,s;2X) = \frac{1}{2X} \, e^{-\sqrt{s}2X} .
3564: \label{Ls1s22Xdef}
3565: \eeq
3566: We now have three inverse Laplace transforms because of the three terms
3567: $\Delta G G$ in Eq.(\ref{p<3}). The integration over $Q_*$ is associated
3568: with $r_4$ in Eq.(\ref{p<3}) and $c_2$ in Eq.(\ref{px1x2<1}) ($q_*$ being
3569: related to $c_2$ through Eq.(\ref{q*def})).
3570: It gives a factor $[e^{s Q_{21}}-e^{s_2 Q_{21}}]/(s-s_2)$.
3571: Then, choosing for instance a contour such that $\Re(s_2)>\Re(s)>1$,
3572: we can integrate the first term over $s_2$, which gives zero by pushing the
3573: contour to the right, $\Re(s_2)\rightarrow+\infty$, and the second term over $s$,
3574: which gives the contribution associated with the pole at $s=s_2$. This yields
3575: \beqa
3576: P_{X_1,X_2}^<(Q_1,Q_2'\geq Q_2) & = & 2 X_{21} \, e^{2X_{21}}
3577: \inta \frac{\dd s_1 \dd s_2}{(2\pi\ii)^2} \,
3578: e^{(s_1-1) Q_1+(s_2-1)Q_{21}} \nonumber \\
3579: && \hspace{-1cm} \times J(s_1,2X_1) \frac{1}{s_2-1}
3580: [ L(1,s_2;2 X_{21}) - L(s_2,s_2;2 X_{21}) ] ,
3581: \label{P<5}
3582: \eeqa
3583: We can check that Eq.(\ref{P<5}) agrees with the constraints (\ref{limp<q2q1}).
3584: Then, taking the derivative with respect to $Q_2$ we obtain the probability
3585: density
3586: \beqa
3587: P_{X_1,X_2}^<(Q_1,Q_2) & = & 2 X_{21} \, e^{2X_{21}}
3588: \inta \frac{\dd s_1 \dd s_2}{(2\pi\ii)^2} \,
3589: e^{(s_1-1) Q_1+(s_2-1)Q_{21}} \nonumber \\
3590: && \hspace{0cm} \times J(s_1,2X_1)
3591: [ L(s_2,s_2;2 X_{21}) - L(1,s_2;2 X_{21}) ] .
3592: \label{P<6}
3593: \eeqa
3594: Finally, combining Eqs.(\ref{P>6}) and (\ref{P<6}) we find that two terms
3595: cancel out and we are left with the total probability density (\ref{PX1X2}).
3596:
3597:
3598:
3599: % BibTeX users please use one of
3600: %\bibliographystyle{spbasic} % basic style, author-year citations
3601: %\bibliographystyle{spmpsci} % mathematics and physical sciences
3602: %\bibliographystyle{spphys} % APS-like style for physics
3603: \bibliographystyle{plain}
3604:
3605: \bibliography{ref} % name your BibTeX data base
3606:
3607:
3608: \end{document}
3609: