<<

PHYSICAL REVIEW E 71, 016124 ͑2005͒

Theory of relativistic Brownian : The „1+1…-dimensional case Jörn Dunkel* Institut für Physik, Humboldt-Universität zu Berlin, Newtonstraße 15, D-12489 Berlin, Germany

Peter Hänggi Institut für Physik, Universität Augsburg, Theoretische Physik I, Universitätstraße 1, D-86135 Augsburg, Germany ͑Received 23 July 2004; published 18 January 2005͒

We construct a theory for the ͑1+1͒-dimensional Brownian motion in a viscous medium, which is ͑i͒ consistent with Einstein’s theory of and ͑ii͒ reduces to the standard Brownian motion in the Newtonian limit case. In the first part of this work the classical Langevin equations of motion, governing the nonrelativistic dynamics of a free Brownian in the presence of a heat bath ͑white noise͒, are general- ized in the framework of special relativity. Subsequently, the corresponding relativistic Langevin equations are discussed in the context of the generalized Ito ͑prepoint discretization rule͒ versus the Stratonovich ͑midpoint discretization rule͒ dilemma: It is found that the relativistic in the Hänggi-Klimontovich interpretation ͑with the postpoint discretization rule͒ is the only one that yields agreement with the relativistic Maxwell distribution. Numerical results for the relativistic Langevin equation of a free Brownian particle are presented.

DOI: 10.1103/PhysRevE.71.016124 PACS number͑s͒: 02.50.Ey, 05.40.Jc, 47.75.ϩf

I. INTRODUCTION With regard to special relativity, standard Brownian mo- tion faces the problem that it permits velocity jumps ⌬v, that For almost 100 years, Einstein’s theory of special relativ- exceed the speed of light c ͑see also Schay ͓3͔͒. This is due ity ͓1,2͔ is serving as the foundation of our most successful to the fact that in the nonrelativistic theory the velocity in- physical standard models ͑apart from gravity͒. The most crements ⌬v have a Gaussian distribution, which always as- prominent and, probably, also the most important feature of signs a nonvanishing ͑though small͒ probability to events this theory is the absolute character of the speed of light c, ⌬vϾc. This problem is also reflected by the Maxwell distri- representing an unsurmountable barrier for the velocity of bution, which represents the stationary velocity distribution any ͑macroscopic͒ physical process. Due to the great experi- for an ensemble of free Brownian and permits ab- mental success of the original theory, almost all other physi- solute velocity values vϾc ͓20͔. cal theories have successfully been adapted to the framework The first relativistically consistent generalization of Max- of special relativity over the past decades. Surprisingly, how- well’s velocity distribution was introduced by Jüttner ͓24͔ in ever, the scientific literature provides relatively few publica- 1911. Starting from an extremum principle for the , tions on the subject of relativistic Brownian ͑classi- he obtained the probability distribution function of the rela- ͓ ͑ ͒ ͔ cal references are ͓3–5͔ and more recent contributions tivistic ideal Boltzmann see Eq. 67 below . In prin- include ͓6–13͔͒. ciple, however, Jüttner’s approach made no contact with the theory of Brownian motion. Fifty years after Jüttner’s work, Brownian particles are physical objects ͑e.g., grains͒ Schay ͓3͔ performed the first comprehensive mathematical that move randomly through a surrounding medium ͑heat ͒ investigation of relativistic processes based on bath . Their motions are caused by permanent col- Lorentz-invariant transition probabilities. On the mathemati- lisions with much lighter constituents of the heat bath ͑e.g., ͓ ͔ ͒ cal side, Schay’s analysis was complemented by Hakim 5 of a . The classical theory of Brownian mo- and Dudley ͓4͔, who studied in detail the properties of tion or nonrelativistic diffusion theory, respectively, was de- Lorentz-invariant Markov processes in relativistic phase veloped by Einstein ͓14͔ and Einstein and von Smolu- ͓ ͔ ͓ ͔ space. After 40 more years, Franchi and Le Jan 13 have chowski 15 . Since the beginning of the last century, when presented an extension of Dudley’s work to general relativ- their seminal papers were published, the classical theory has ity. In particular, these authors discuss relativistic been investigated and generalized by a large number of ͓ ͔ ͓ ͔ ͓ ͔ in the presence of a Schwarzschild metric 25 . Hence, over physicists 16–20 and mathematicians 21–23 . The intense the past 100 years there has been steady ͑though relatively research led, among others, to different mathematical repre- ͒ ͓ slow progress in the mathematical analysis of relativistic sentations of the Brownian motion dynamics Langevin diffusion processes. equations, Fokker-Planck equations ͑FPE͒, etc.͔͓18–20͔,to ͓ ͔ By contrast, one finds in the physical literature only very the notion of the Wiener processes 21 , and to new tech- few publications that directly address the topic of the relativ- niques for solving partial differential equations ͑Feynman- ͑ ͓ ͔͒ istic Brownian motion despite the fact that relativistic ki- Kac formula, etc. 22,23 . netic theory has been fairly well established for more than 30 years ͓26–29͔͒. Among the few exceptions are the papers by Boyer ͓8,9͔ and Ben-Ya’acov ͓6͔, who have studied the *Electronic address: [email protected] interaction between two energy-level particles and electro-

1539-3755/2005/71͑1͒/016124͑12͒/$23.00016124-1 ©2005 The American Physical Society J. DUNKEL AND P. HÄNGGI PHYSICAL REVIEW E 71, 016124 ͑2005͒ magnetic radiation in thermal equilibrium, the latter acting as bath ͑e.g., small liquid particles͒. In the Langevin approach a heat bath. In contrast to their specific microscopic model, the nonrelativistic dynamics of the Brownian particle is de- we shall adopt a more coarse-grained point of view here by scribed by the stochastic dynamical equations ͑see, e.g., ͓20͔ assuming that the heat bath is sufficiently well described by Chap. IX͒ macroscopic and diffusion coefficients. dx͑t͒ Generally, the objective of the present paper can be sum- = v͑t͒, ͑1a͒ marized as follows: We would like to discuss how one can dt construct, in a physically straightforward manner, a relativ- istic theory of Brownian motion for particles moving in a dv͑t͒ homogeneous, viscous medium. For this purpose it is suffi- m =−␯mv͑t͒ + L͑t͒, ͑1b͒ dt cient to concentrate on the case of 1+1 dimensions ͑gener- alizations to the 1+3 dimensions are straightforward and will where ␯ is the viscous friction coefficient. The Langevin be discussed separately in a forthcoming contribution͒.Asa force L͑t͒ is characterized by starting point we choose the nonrelativistic Langevin equa- tions of the free Brownian particle. In Sec. II these equations ͗L͑t͒͘ =0, ͗L͑t͒L͑s͒͘ =2D␦͑t − s͒, ͑2͒ will be generalized such that they comply with special rela- with all higher cumulants being zero ͑Gaussian white noise͒, tivity. As we shall see in Sec. III due to multiplicative noise and D being constant. More general models may include for the degree of freedom, the resulting relativ- velocity-dependent parameters ␯ and D ͑see, e.g., istic Langevin equations are not sufficient in order to ͓19,30–32͔͒, but we shall restrict ourselves to the simplest uniquely determine the corresponding Fokker-Planck equa- ͑ ͒ case here. It is worthwhile to summarize the physical as- tion generalized Ito-Stratonovich dilemma . Furthermore, it sumptions, implicitly underlying Eqs. ͑1͒ as follows: it is shown that the stationary of a particular form ͑i͒ The heat bath is homogeneous. for the relativistic Fokker-Planck equation coincides with ͑ ͒ ͑ ͒ ii Stochastic impacts between the Brownian particle and Jüttner’s relativistic Maxwell distribution Sec. III B 3 . Fi- the constituents of the heat bath occur virtually uncorrelated. nally, we also discuss numerical results for the mean-square ͑iii͒ On the macroscopic level, the interaction between displacement in Sec. IV. Brownian particle and heat bath is sufficiently well described It might be worthwhile to emphasize that the systematic by the constant viscous friction coefficient ␯ and the white Langevin approach pursued below is methodically different noise force L. from those in Refs. ͓3–13͔ and also from the kinetic theory ͑ ͒ ͑ ͒ ⌺ ͓ ͔ iv Equations 1 hold in the rest frame 0 of the heat approach 26–29 . It is therefore satisfactory that our find- bath ͑corresponding to the specific inertial system, in which ings are apparently consistent with rigorous mathematical the average velocity of the heat bath vanishes for all times t͒. results, obtained by Schay ͓3͔ and Dudley ͓4͔ for the case of ⌺ In the following 0 will also be referred to as laboratory free relativistic diffusion. Moreover, it will become clear in frame. Sec. IV that numerical simulations of the relativistic Lange- In the mathematical literature, Eq. ͑1b͒ is usually written vin equations constitute a very useful tool for the numerical as investigation of relativistic diffusion processes, provided that the discretization rule is carefully chosen. d͓mv͑t͔͒ =−␯mv͑t͒dt +dW͑t͒, ͑3a͒ II. where W͑t͒ is a one-dimensional ͓19,22,23͔, First the main properties of the nonrelativistic Langevin i.e., the density of the increments equations for free Brownian particles are briefly summarized w͑t͒ϵdW͑t͒ϵW͑t +dt͒ − W͑t͒͑3b͒ ͑Sec. II A͒. Subsequently, we construct generalized Lorentz- covariant Langevin equations ͑Sec. II B͒. Finally, the cova- is given by riant Langevin equations will be rewritten in laboratory co- ͑ ͒ 1 w͑t͒2 ordinates Sec. II C . P1͓w͑t͔͒ = expͫ− ͬ. ͑3c͒ The following notations will be used throughout the pa- ͱ4␲D dt 4D dt per. Since we confine ourselves to the ͑1+1͒-dimensional ϵ case, upper and lower Greek indices ␣,␤,... can take values Here the abbreviation w dW has been introduced to sim- ͑ ͒ 0, 1, where 0 refers to the time component. The plify the notation in subsequent formulas. From Eq. 3c one ͑ ͒ ͑1+1͒-dimensional Minkowski metric tensor with respect to finds in agreement with 2 Cartesian coordinates is taken as 0, t Þ s ͗ ͑ ͒͘ ͗ ͑ ͒ ͑ ͒͘ ͭ ͮ ͑ ͒ ␣␤ w t =0, w t w s = 4 ͑␩␣␤͒ = ͑␩ ͒ = diag͑−1,1͒. 2D dt, t = s. Moreover, Einstein’s summation convention is invoked Depending on which notation is more convenient for the throughout. current purpose, we shall use below either the physical for- mulation ͑1͒ or the mathematical formulation ͑3͒. The two A. Physical foundations formulations can be connected by ͑formally͒ setting Consider the nonrelativistic one-dimensional motion of a Brownian particle with mass m that is surrounded by a heat w͑t͒ =dW͑t͒ = L͑t͒dt. ͑5͒

016124-2 THEORY OF RELATIVISTIC BROWNIAN MOTION: … PHYSICAL REVIEW E 71, 016124 ͑2005͒

B. Relativistic generalization trast, in the relativistic theory these equations are exact at ⌺ It is well known that in inertial coordinate systems, which time t only if * is comoving at time t. In the latter case, we ͑ ͒ are comoving with a particle at a given t, the rela- can use Eqs. 10 to construct relativistically covariant equa- ␶ tivistic equations must reduce to the nonrelativistic Newton- tions of motion. Introducing, as usual, the proper time by ian equations ͑see, e.g., ͓25͔ Chap. 2.3͒. Therefore, our strat- the definition egy is as follows. Starting from the Langevin equations ͑1͒ v2 v2 or ͑3a͒, respectively, we construct in the first step the non- ␶ ϵ ͱ ͱ * ͑ ͒ d dt 1− 2 =dt* 1− 2 , 11 relativistic equations of motion with respect to a coordinate c c ⌺ frame *, comoving with the Brownian particle at a given and combining momentum p* =mv* and energy into a moment t. In the second step, the general form of the cova- ͑ ͒ ͑ ␣͒ ͑ 0 ͒ ͑ ͒ 1+1 -vector p* = p ,p* = E* /c,p* , we can rewrite Eqs. riant relativistic equation motions are found by applying a ͑10͒ in the covariant form Lorentz transformation to the nonrelativistic equations that ⌺ ␣ have been obtained for *. dp ␣ ␣ * = f , ͑f ͒ =−m␯͑0,v − V ͒. ͑12͒ It is useful to begin by considering the deterministic d␶ * * * * ͑noise-free͒ limit case, corresponding to a pure damping of ͑ ␣͒ ͑ ␣͒ ͑ ͒ the particle’s motion. This will be done Sec. II B 1. Subse- Let u* and U* denote the 1+1 -velocity components of quently, the stochastic force is separately treated in Sec. Brownian particle and heat bath, respectively. Now it is im- IIB2. portant to realize that the covariant force vector f␣ cannot be simply proportional to the ͑1+1͒-velocity difference, 1. Viscous friction ␣ Þ ␯͑ ␣ ␣͒ ͑ ͒ Setting the stochastic force term to zero ͑corresponding to f* − m u* − U* , 13 ͒ a vanishing of the heat bath , the nonrelativistic because, in general, at time t in ⌺ Eq. ͑1b͒ simplifies to * c c ͑9͒ c dv͑t͒ u0 − U0 = − = c − Þ 0. m =−␯mv͑t͒. ͑6͒ * * ͱ 2 2 ͱ 2 2 ͱ 2 2 dt 1−v*/c 1−V*/c 1−V*/c ͑14͒ The energy of the Brownian particle is purely kinetic, However, we can write f␣ in a manifestly covariant form, if mv͑t͒2 * E͑t͒ = , ͑7͒ we introduce the friction tensor 2 00 ͑␯ ␣ ͒ ͩ ͪ ͑ ͒ and, by virtue of ͑6͒, its time derivative is given by ␤ = , 15 * 0 ␯ dE dv = mv =−␯mv2. ͑8͒ which allows us to rewrite ͑12͒ as dt dt ␣ dp* ␣ ␤ ␤ As stated above, in the nonrelativistic theory the last three =−m␯ ␤͑u − U ͒. ͑16͒ ⌺ d␶ * * * equations are assumed to hold in the rest frame 0 of the heat ⌺ bath. Now consider another inertial coordinate system *,in This equation is manifestly Lorentz-invariant, and we drop which the Brownian particle is temporarily at rest at time t or the asterisk from now on, while keeping in mind that the ͑ ͒ ⌺ t* =t* t , respectively, where t* denotes the *-time coordi- diagonal form of the friction tensor ͑15͒ is linked to the rest nate. That is, in ⌺ we have at time t ⌺ * frame * of the Brownian particle. In this respect the friction ͑ ͒ϵ ͑ ͒ ͑ ͒ tensor is very similar to the tensor, as known from v* t v* t* t =0. 9 „ … the relativistic hydrodynamics of perfect fluids ͑see, e.g., ͓ ͑ ͒ Conventionally, we use throughout the lax notation g* t ͓25͔ Chap. 2.10͒. This analogy yields immediately the fol- ϵ ͑ ͒ ͔ g*(t* t ), where g* is originally a function of t*. With lowing representation: ⌺ respect to the comoving frame *, the heat bath will, in ␣ ͑ ͒ u u␤ general, have a nonvanishing average velocity V*. Then, ␯␣ ␯ͩ␩␣ ͪ ͑ ͒ ͑ ͒ ⌺ ␤ = ␤ + 2 . 17 using a Galilean transformation we find that Eq. 6 in * c coordinates at time t reads as follows: It is now interesting to consider Eq. ͑16͒ in the laboratory ͑ ͒ dv 9 ⌺ * ͑ ͒ ␯ ͑ ͒ ␯ ͑ ͒ frame 0, defined above as the rest frame of the heat bath. m t =− m„v* t − V*…= mV*. 10a dt* There we have Similarly, in ⌺ coordinates Eq. ͑8͒ is given by dt 1 * ͑U␤͒ = ͑c,0͒, ͑u␤͒ = ͑␥c,␥v͒,d␶ = , ␥ ϵ . ͑ ͒ ␥ ͱ 2 2 dE 9 1−v /c * ͑ ͒ ␯ ͑ ͒ ͑ ͒ ͑ ͒ t =− mv* t „v* t − V*…= 0. 10b ͑ ͒ dt* 18 Note that in the nonrelativistic ͑Newtonian͒ theory the left Combining ͑16͒–͑18͒ we find that the relativistic equations ⌺ equalities in Eqs. ͑10͒ are valid for arbitrary time t. By con- of motion in 0 are given by

016124-3 J. DUNKEL AND P. HÄNGGI PHYSICAL REVIEW E 71, 016124 ͑2005͒

relativistic velocity curves, given by Eq. ͑24͒, exhibit essen- tial deviations from the purely exponential decay, predicted by the Newtonian theory.

2. Stochastic force We now construct a relativistic generalization of the sto- chastic force. To this end, we consider Eq. ͑11͒ as an opera- tional definition for the proper time parameter ␶. The gener- alization procedure will be based on the standard assumption ͑ ͒ ⌺ postulate that, in temporarily comoving inertial frames *, the relativistic equations of motions must reduce to the New- tonian equations of motions. According to this assumption, FIG. 1. Velocity curves v͑t͒, corresponding to Eq. ͑24͒, for the ⌺ for frames *, comoving with the particle at laboratory time purely damped motion of a relativistic particle in the rest frame of t, the relativistic stochastic differential equation must reduce ͑ ͒ ͉ ͉ the heat bath laboratory frame . Especially at high velocities v to Շc, the relativistic velocity curves deviate from the exponential decay, predicted by the Newtonian theory. ͑ ͒ ␯ ͑ ͒ ͑ ͒ ͑ ͒ dp* t =− „p* t − mV*…dt* + w* t , 25 ϵ dp mv where the momentum increments w* dW* represent a =−␯ ͑19a͒ Wiener-process with parameter D, i.e., the increments w ͑t͒ ͱ 2 2 * dt 1−v /c have a Gaussian distribution

dE mv2 1 w ͑t͒2 =−␯ . ͑19b͒ P1͓w ͑t͔͒ = expͩ− * ͪ. ͑26͒ ͱ 2 2 * * ͱ ␲ dt 1−v /c 4 D dt* 4D dt* On comparing ͑19a͒ with ͑6͒ and ͑19b͒ with ͑8͒, one readily Note that also in the relativistic theory the momentum incre- observes that the relativistic equations ͑19͒ do indeed reduce ments w͑t͒=dW͑t͒ may tend to infinity, as long as the related to the known Newtonian laws in the limit case v2 /c2 Ӷ1. velocity increments remain bounded. In other words, in the Using the relativistic definitions relativistic theory one must carefully distinguish between stochastic momentum and velocity increments ͑this is not E = ␥mc2, p = ␥m , ͑20͒ v necessary in the nonrelativistic theory, because Newtonian Eqs. ͑19͒ can also be rewritten as momenta are simply proportional to their velocities͒. The next step is now to define the increment ͑1+1͒-vector dp =−␯p, ͑21a͒ by dt ͑ ␣͒ ͑ ͒ ͑ ͒ w* = 0,w* . 27 dE v2 =−␯pv =−␯E . ͑21b͒ This definition is in agreement with the requirement that in a dt c2 ⌺ comoving inertial system * the 0-component of the ͑ ͒ ͑ ͓ ͔ In fact, only one of the two Eqs. ͑19͒ or ͑21͒, respectively, 1+1 -force vector must vanish see, e.g., 25 Chap. 2.3, ͑ ͒ ͑ ͒ ͑ ͒ must be solved due to the fixed relation between relativistic and also compare Eqs. 12 , 31 , and 32 of the present ͒ ⌺ energy and momentum: paper . Moreover, if the Lorentz frame * is comoving with the Brownian particle at given time t, then the ͑equal-time͒ 2 ␣ 2 2 2 2 2 mc white-noise relations ͑4͒ generalize to p␣p =−E /c + p =−m c ⇒ E͑t͒ = . ͱ1−v2/c2 0, ␣ = 0 and/or ␤ =0, ͑ ͒ ͗ ␣͑ ͒͘ ͗ ␣͑ ͒ ␤͑ ͒͘ ͭ ͕ 22 w* t =0, w* t w* t = 2D dt*, otherwise. ͑ ͒ The solution of 21a reads ͑28͒ p͑t͒ = p exp͑− ␯t͒, p͑0͒ = p , ͑23͒ 0 0 The rhs. of the second equation in ͑28͒ makes it plausible to and, by using ͑20͒, one thus obtains for the velocity of the introduce a correlation tensor by ⌺ ͑ particle in the laboratory frame 0 rest frame of the heat ͒ 00 bath ͑ ͒ ͩ ͪ ͑ ͒ D*␣␤ = , 29a v2 v2 −1/2 0 2D dt* ͑ ͒ ͫͩ 0 ͪ 2␯t 0 ͬ ͑ ͒ v t = v0 1− e + . 24 c2 c2 thus, Figure 1 depicts a semi-logarithmic representation of the ve- ͗ ␣͑ ͒ ␤͑ ͒͘ ␣␤ ͑ ͒ ͑ ͒ w* t w* t = D* . 29b locity v t for different values of the initial velocity v0.As one can see in the diagram, at high velocities ͉v͉Շc the Additionally defining an “inverse” correlation tensor by

016124-4 THEORY OF RELATIVISTIC BROWNIAN MOTION: … PHYSICAL REVIEW E 71, 016124 ͑2005͒

␣ 00 from ͑35b͒ and then used that u␣w =0, see Eq. ͑32͒. ͑ ˆ ͒ ΀ ΁ ͑ ͒ D ␣␤ = , 29c * 0 ͑2D dt ͒−1 By virtue of the above results, we are now in the position * to write down the covariant Langevin equations with respect allows us to generalize the distribution of the increments to an arbitrary inertial system: If a Brownian particle with from Eq. ͑26͒ as follows: rest mass m, proper time ␶ and ͑1+1͒-velocity u␤ is sur- rounded by an isotropic, homogeneous heat bath with con- 1 1 P1+1 ␣͑ ͒ ͫ ˆ ␣͑ ͒ ␤͑ ͒ͬ␦ 0͑ ͒ stant 1+1 velocity U␤, then the relativistic Langevin equa- * „w* t … = ͱ exp − D*␣␤w* t w* t „w* t …. 4␲D dt* 2 tions of motions read

͑30͒ ␣ ␣ p ͑␶͒ ␦ dx ͑␶͒ = d␶ ͑36a͒ Here, the Dirac -function on the right-hand side accounts m for the fact that the 0 component of the stochastic force must vanish in every inertial frame, comoving with the Brownian ␣͑␶͒ ␯␣ ␤͑␶͒ ␤ ␶ ␣͑␶͒ ͑ ͒ particle at time t; compare Eq. ͑27͒. This also follows more dp =− ␤„p − mU …d + w , 36b generally from the identity where, according to Eq. ͑17͒, the friction tensor is given by

d 2 d ␣ ␣ ␣ 0 ϵ ͑− mc ͒ = m ͑u␣u ͒ =2u␣f , ͑31͒ u u␤ ␶ ␶ ␯␣ ␯ͩ␩␣ ͪ ͑ ͒ d d ␤ = ␤ + , 36c c2 which, in the case of the stochastic force, translates to ␯ ␣ with denoting the viscous friction coefficient measured in 0=u␣w . ͑32͒ the rest frame of the particle. This is a first main result of this ␣͑␶͒ϵ ␣͑␶͒ Hence, we can rewrite the probability distribution ͑30͒ as work. The stochastic increments w dW are distrib- uted according to ͑35c͒ and, therefore, characterized by ␣ c 1 ␣ ␤ P1+1 ͑ ͒ ͩ ˆ ͑ ͒ ͑ ͒ͪ ␣ w t = exp − D ␣␤w t w t * „ * … ͱ ␲ * * * ͗w ͑␶͒͘ =0, ͑36d͒ 4 D dt* 2 ␣ ϫ␦ u ␣w ͑t͒ , ͑33͒ „ * * … 0, ␶ Þ ␶Ј; ͗ ␣͑␶͒ ␤͑␶Ј͒͘ ͭ ͕ ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ w w = ␣␤ 36e where u*␣ = −c,0 is the covariant 1+1 -velocity of the D , ␶ = ␶Ј, particle a the comoving rest frame. It should be stressed that, ␣␤ because of the constraint ͑32͒, only one of the two incre- with D given by ͑35a͒. Note that in each comoving Lor- ments w␣ ϵdW␣ is to be regarded as “independent,” which is entz frame, in which, at a given moment t, the particle is at reflected by the appearance of the ␦ function in ͑33͒. Also rest, the marginal distribution of the spatial momentum in- note that, due to the prefactor c, the normalization condition crements, defined by takes the simple form ϱ ␣ 1 ϱ P1 w͑t͒ = ͵ d w0͑t͒ P1+1 w ͑t͒ , ͑37͒ ␣ ␣ „ … „ … „ … ͭ͟ ͵ ͑ ͒ ͮP1+1 ͑ ͒ ͑ ͒ −ϱ 1= d„w* t … * „w* t …. 34 ␣=0 −ϱ reduces to a Gaussian. In the Newtonian limit case, corre- Furthermore, analogous to ͑17͒, we have the following more sponding to v2 Ӷc2, one thus recovers from Eqs. ͑35͒ and general representation of the correlation tensors: ͑36͒ the usual nonrelativistic Brownian motion.

u␣u␤ ␶ͩ␩ ͪ ͑ ͒ D␣␤ =2D d ␣␤ + 35a c2 C. Langevin dynamics in the laboratory frame ⌺ A laboratory frame 0 is, by definition, an inertial system, ␤ 1 u␣u␤ ⌺ ͑ ͒ ˆ ͩ␩ ͪ ͑ ͒ in which the heat bath is at rest, i.e., in 0 we have U D␣␤ = ␣␤ + 2 . 35b ͑ ͒ ⌺ 2D d␶ c = c,0 for all times t. Hence, with respect to 0 coordinates, the two stochastic differential Eqs. ͑36b͒ assume with ͑36c͒ ͑ ͒ Then, in an arbitrary Lorentz frame, the density 33 can be the form written as ␯ ͑ ͒ ͑ ͒ c 1 dp =− pdt + w t , 38a P1+1 ␣͑␶͒ ͫ ˆ ␣͑␶͒ ␤͑␶͒ͬ w = exp − D␣␤w w „ … ͱ4␲D d␶ 2 dE =−␯pvdt + cw0͑t͒. ͑38b͒ ϫ␦ ␣͑␶͒ „u␣w … ␣ Here it is important to note that the stochastic increments c w␣͑␶͒w ͑␶͒ ␣ ␦ ␣͑␶͒ w ͑t͒, appearing on the right-hand side. of ͑38͒, are not of = ͱ expͫ− ͬ u␣w . 4␲D d␶ 4D d␶ „ … simple Gaussian type anymore. Instead, their distribution ͑35c͒ now also depends on the particle velocity v. This becomes immediately evident, when we rewrite the increment density ˆ ͑ ͒ ⌺ To obtain the last line from the first, we have inserted D␣␤ 35c in terms of 0 coordinates. Using

016124-5 J. DUNKEL AND P. HÄNGGI PHYSICAL REVIEW E 71, 016124 ͑2005͒

2 III, it will become clear that, for example, choosing p=p͑t͒ −1 v ␣ 0 ͑u␣͒ = ͑− ␥c,␥v͒, ␥ = ͱ1− , ͑w ͒ = ͑w ,w͒, in Eqs. ͑46͒ would be consistent with an Ito-interpretation c2 ͓20,33,34͔ of the stochastic differential equation ͑38a͒. How- ͑39͒ ever, we will also see that alternative interpretations lead to we find reasonable results as well. ␥ 1/2 ͑ ͒2 0͑ ͒2 III. DERIVATION OF CORRESPONDING 1+1 ␣ w t − w t P w ͑t͒ = cͩ ͪ expͩ− ͪ FOKKER-PLANCK EQUATIONS „ … 4␲D dt 4D dt/␥ ϫ␦ c␥w0͑t͒ − ␥vw͑t͒ . ͑40͒ The objective in this part is to derive relativistic Fokker- „ … Planck equations ͑FPE͒ for the momentum density f͑t,p͒ of ␦ ͑ ͒ ⌺ As we already pointed out earlier, the function in 40 a free particle in the laboratory frame 0. Before we deal reflects the fact that the energy increment w0 is coupled to with this problem in Sec. III B, it is useful to briefly recall the spatial ͑momentum͒ increment w via the nonrelativistic case.

␣ 0 0 vw A. Nonrelativistic case 0=u␣w =−c␥w + ␥vw ⇒ w = . ͑41͒ c Consider the nonrelativistic Langevin equation ͑1b͒ Hence, w0 can be eliminated from the Langevin equations dp ͑38b͒, yielding =−␯p + L͑t͒, ͑47a͒ dt dE =−␯pvdt + vw͑t͒ = vdp. ͑42͒ where p͑t͒=mv͑t͒ denotes the nonrelativistic momentum, Using the identity and, in agreement with ͑3c͒, the Langevin force L͑t͒ is dis- cp tributed according to v = , ͑43͒ ͱm2c2 + p2 dt 1/2 dt P L͑t͒ = ͩ ͪ expͩ− L͑t͒2ͪ. ͑47b͒ „ … 4␲D 4D we can further rewrite ͑42͒ as As is well known ͓20,35͔, the related momentum probability cp dE = dp ⇒ E͑t͒ = ͱm2c4 + p͑t͒2c2. density f͑t,p͒ is governed by the Fokker-Planck equation ͱm2c2 + p2 ץ ץ ץ ͑44͒ f = ͩ␯pf + D fͪ, ͑48͒ pץ pץ tץ ⌺ Thus, in the laboratory frame 0 the relativistic Brownian motion is completely described by the Langevin equation whose stationary solution is the Maxwell distribution ͑ ͒ 38a already. If we assume that the Brownian particle has ␯ 1/2 ␯p2 ͑ ͒ ͑ ͒ f͑p͒ = ͩ ͪ expͩ− ͪ. ͑49͒ fixed initial momentum p 0 =p0 or initial velocity v 0 =v0, ␲ respectively, then the formal solution of ͑38a͒ reads ͓͑20͔ 2 D 2D Chap. IX.1͒ B. Relativistic case t ͑ ͒ −␯t −␯t͵ ␯s ͑ ͒ ͑ ͒ p t = p0e + e e w s . 45 We next discuss three different relativistic Fokker-Planck 0 equations for the momentum density f͑t,p͒, related to the ͑ ͒ ͑ ͒ The ͑45͒ is determined by the marginal stochastic processes defined by 38a and 46 . distribution P1(w͑t͒), defined in Eq. ͑37͒. Performing the Our starting point is the relativistic Langevin equation ͑ ͒ ⌺ ͑ integration over the ␦ function in ͑40͒,wefind 38a , which holds in the laboratory frame 0 i.e., in the rest frame of the heat bath͒. Next we define a stochastic process 1 1/2 w͑t͒2 by P1 w͑t͒ = ͩ ͪ expͩ− ͪ, ͑46a͒ „ … ␲ ␥ ␥ 4 D dt 4D dt w͑t͒ y͑t͒ = , ͑50͒ where ͱ␥ v2 −1/2 p2 1/2 and using ͑46b͒, we can rewrite ͑38a͒ as ␥ = ͫ1− ͬ = ͫ1+ ͬ . ͑46b͒ c2 m2c2 dp =−␯pdt + ͱ␥y͑t͒, ͑51a͒ On the basis of Eqs. ͑38a͒ and ͑46͒ one can immediately where y͑t͒ is distributed according to the momentum- perform computer simulations, provided one still specifies the rules of , i.e., which value of p is to be independent density taken to determine ␥ in ͑46͒. In Sec. IV several numerical 1/2 ͑ ͒2 1 1 y t results are presented. Before, it is useful to consider in more P ͓y͑t͔͒ = ͩ ͪ expͩ− ͪ. ͑51b͒ y 4␲D dt 4D dt detail the Fokker-Planck equations of the relativistic Brown- ⌺ ian motion in the laboratory frame 0. By doing so in Sec. Thus, instead of the increments w͑t͒, which implicitly de-

016124-6 THEORY OF RELATIVISTIC BROWNIAN MOTION: … PHYSICAL REVIEW E 71, 016124 ͑2005͒ pend on the stochastic process p via Eqs. ͑46͒, we consider p p2 ordinary p-independent y͑t͒, determined by ͵ dp = c2m2ͱ1+ , ͑56͒ ␥͑p͒ c2m2 ͑51b͒, from now on. Due to the multiplicative coupling of y͑t͒ in ͑51a͒, we must next specify rules for the “multiplica- we find the following explicit representation of ͑55͒: tion with white noise” ͓note that, on viewing Eqs. ͑11͒, ͑35͒, ͑ ͒ p2 −1/2 p2 and 36 as postulates of the relativistic Brownian motion, all f ͑p͒ = C ͩ1+ ͪ expͩ− ␤ͱ1+ ͪ, ͑57͒ above considerations remain valid, independent of this speci- I I m2c2 c2m2 fication͔. where In Secs. III B 1, III B 2, and III B 3, we shall discuss three popular multiplication rules, which go back to proposals ␯m2c2 ͓ ͔ ͓ ͔ ␤ = . ͑58͒ made by Hänggi and Thomas 19,42 , Van Kampen 20 , Ito D ͓33,34͔, Stratonovich ͓36,37͔, Fisk ͓38,39͔, Hänggi ͓40,41͔, and Klimontovich ͓31͔. As it is well-known from ͓19,20,30͔, The dimensionless parameter ␤ can be used to define the these different interpretations of the stochastic process ͑51͒ scalar temperature T of the heat bath via the Einstein relation result in different Fokker-Planck equations, i.e., the Langevin mc2 D equation ͑51͒ per se does not uniquely determine the corre- ϵ ͑ ͒ kBT = , 59 sponding Fokker-Planck equation; it is the stochastic inter- ␤ m␯ pretation of the multiplicative noise that matters from a with kB denoting the . Put differently, the physical point of view. parameter ␤=mc2 /͑k T͒ measures the ratio between rest Nevertheless, the three approaches discussed below have B mass and thermal energy of the Brownian particle. in common that, formally, the related Fokker-Planck equa- tion can be written as a ͓42͔ 2. Stratonovich approach ץ ץ f͑t,p͒ + j͑t,p͒ =0, ͑52͒ According to Stratonovich, the coefficient before y͑t͒ in ,p ͑51a͒ is to be evaluated with the midpoint discretization ruleץ tץ but with different expressions for the probability current i.e., ͑ ͒ j t,p . It is worthwhile to anticipate that only for the Hänggi- p͑t͒ + p͑t +dt͒ Klimontovich approach ͑see Sec. III B 3͒ the current j͑t,p͒ ␥ = ␥ͩ ͪ. ͑60͒ 2 takes such a form that the stationary distribution of ͑52͒ can be identified with Jüttner’s relativistic Maxwell distri- This choice leads to a different expression for the current bution ͓24͔. ͓19,36–38͔, namely, ץ Ito approach .1 j ͑p,t͒ =−ͫ␯pf + Dͱ␥͑p͒ ͱ␥͑p͒fͬ. ͑61͒ pץ According to Ito’s interpretation of the Langevin equation S ͑51a͒, the coefficient before y͑t͒ is to be evaluated at the This Stratonovich-Fisk current j vanishes identically for lower boundary of the interval ͓t,t+dt͔, i.e., we use the pre- S point discretization rule C ␯ p ͑ ͒ S ͫ ͵ ͬ ͑ ͒ fS p = exp − dp , 62 ␥ ␥ ͑ ͒ ͑ ͒ ͱ␥͑p͒ D ␥͑p͒ = „p t …, 53 where as before and, by virtue of ͑56͒, the explicit stationary solution of Stra- tonovich’s Fokker-Planck equation reads p2 1/2 ␥͑ ͒ ͩ ͪ p = 1+ 2 2 . p2 −1/4 p2 m c f ͑p͒ = C ͩ1+ ͪ expͩ− ␤ͱ1+ ͪ. ͑63͒ S S m2c2 c2m2 Ito’s choice leads to the following expression for the current ͓19,20,33,34͔: Hänggi-Klimontovich approach .3 ץ j ͑p,t͒ =−ͫ␯pf + D ␥͑p͒fͬ. ͑54͒ -p Now let us still consider the Hänggi-Klimontovich stoץ I chastic interpretation, sometimes referred to as the The related relativistic Fokker-Planck equation is obtained transport form ͓40–42͔ or also as the kinetic form ͓31͔. Ac- by inserting this current into the conservation law ͑52͒. The cording to this interpretation, the coefficient in front of y͑t͒ ͑ ͒ current 54 vanishes identically for in ͑51a͒ is to be evaluated at the upper boundary value of the ͓ ͔ C ␯ p interval t,t+dt ; i.e., within the postpoint discretization we f ͑p͒ = I expͫ− ͵ dp ͬ, ͑55͒ set I ␥͑p͒ D ␥͑p͒ ␥ ␥ ͑ ͒ ͑ ͒ ͑ ͒ = p t +dt . 64 where CI is the normalization constant. Consequently, fI p „ … is a stationary solution of the Fokker-Planck equation. In This choice leads to the following expression for the current view of the fact that ͓31,41,42͔:

016124-7 J. DUNKEL AND P. HÄNGGI PHYSICAL REVIEW E 71, 016124 ͑2005͒

ץ j ͑p,t͒ =−ͫ␯pf + D␥͑p͒ fͬ. ͑65͒ pץ HK

The current jHK vanishes identically for ␯ p f ͑p͒ = C expͫ− ͵ dp ͬ, ͑66͒ HK HK D ␥͑p͒ and, by virtue of ͑56͒, the stationary solution explicitly reads p2 f ͑p͒ = C expͩ− ␤ͱ1+ ͪ. ͑67a͒ HK HK c2m2 Using the temperature definition in ͑59͒ and the relativistic formula E=ͱm2c4 +p2c2, one can further re- write ͑67a͒ in a more concise form as E ͑ ͒ ͩ ͪ ͑ ͒ fHK p = CHK exp − . 67b kBT The distribution function ͑67͒ is known as the relativistic Maxwell distribution. It was first obtained by Jüttner ͓24͔ back in 1911. Pursuing a completely different line of reason- ing, he found that ͑67͒ describes the velocity distribution of the noninteracting relativistic gas ͑see also ͓43͔͒. In contrast to our approach, which started out with constructing the rela- tivistic generalization of the Langevin equations, Jüttner’s derivation started from a maximum-entropy-principle for the gas. By comparing ͑55͒, ͑63͒, and ͑67a͒ one readily observes that the stationary fI/S differ from the Jüttner func- tion fHK through additional p-dependent prefactors. In order to illustrate the differences between the different stationary solutions, it useful to consider the related velocity probability ␾ ͑ ͒ density functions I/S/HK v , which can be obtained by ap- plying the general transformation law pץ ␾͑v͒ϵf p͑v͒ ͯ ͯ ͑68͒ vץ … „ ␾ ͑ ͒ FIG. 2. Stationary solutions I/S/HK v of the relativistic Fokker- in combination with Planck equations, according to Ito ͑I:solid line͒, Stratonovich ͑S:dotted͒, and Hänggi-Klimontovich ͑HK:dashed-dotted͒. For low mv p = . , i.e., for ␤ӷ1, a Gaussian shape is approached ͓see ͱ 2 2 1−v /c ͑a͔͒. On the other hand, for very high-temperature values, corre- ␤ഛ sponding to 1, the distributions exhibit a bistable shape, and the ͒ ͑ ͉ ץ ץ͉ The determinant factor p/ v in 68 is responsible for the ␾ ͑ ͒ ␾ ͑ ͒ quantitative deviations between I/S/HK v increase significantly as fact that the velocity density functions I/S/HK v are, in fact, ␤→0. zero if v2 Ͼc2. In Fig. 2 we have plotted the probability density functions symmetric around v=0. Instead, they will be centered around ␾ ͑ ͒ ␤ I/S/HK v for different values of the parameter . The nor- the nonvanishing ⌺Ј velocity VЈ of the heat bath. malization constants were determined by numerically inte- An obvious question then arises, which of the above ap- grating ␾͑v͒ over the interval ͓−c,c͔. As one can observe in proaches ͑Ito, Stratonovich, or Hänggi-Klimontovich͒ is the Fig. 2͑a͒, for large values of ␤, corresponding to low- physically correct one. We believe that, at this level of analy- Ӷ 2 temperature values kBT mc , the density functions sis, it is impossible to provide a definite answer to this ques- ␾ ͑ ͒ I/S/HK v approach a common Gaussian shape. On the other tion. Most likely, the answer to this problem requires addi- ജ 2 hand, for high-temperature values kBT mc the deviations tional information about the microscopic structure of the heat from the Gaussian shape become essential. The reason is bath ͑see, e.g., the discussion of Ito-Stratonovich dilemma in that, for a ͑virtual͒ Brownian ensemble in the high- the context of “internal and external” noise as given in Chap. temperature regime, the majority of particles assumes veloci- IX.5 of van Kampen’s textbook ͓20͔͒. At this point, it might ties that are close to the speed of light. It is also clear that in be worthwhile to mention that the relativistic Maxwell dis- other Lorentz frames ⌺Ј, which are not rest frames of the tribution ͑67͒ is also obtained via the transfer probability heat bath, the stationary distributions will no longer stay method used by Schay, see Eq. ͑3.63͒ and ͑3.64͒ in Ref. ͓3͔,

016124-8 THEORY OF RELATIVISTIC BROWNIAN MOTION: … PHYSICAL REVIEW E 71, 016124 ͑2005͒ and that this distribution also results in the relativistic kinetic theory ͓29͔. By contrast, the recent work of Franchi and Le Jan ͓13͔ is based on the Stratonovich approach. From physi- cal insight, however, it is the transport form interpretation of Hänggi and Klimontovich that is expected to provide the physically correct description.

IV. NUMERICAL INVESTIGATIONS The numerical results presented in this section were ob- tained on the basis of the relativistic Langevin equation ͑51͒, ⌺ which holds in the laboratory frame 0. For simplicity, we confined ourselves here to considering the Ito-discretization scheme with fixed time step dt ͑see Sec. III B 1͒. In all simu- lations we have used an ensemble size of N=10 000 par- ticles. Moreover, a characteristic unit system was fixed by setting m=c=␯=1. Formally, this corresponds to using res- caled dimensionless quantities, such as ˜p=p/mc, ˜x=x␯/c, ˜t =t␯, ˜v=v/c, etc. The simulation time-step was always cho- sen as dt=0.001␯−1, and the Gaussian random variables y͑t͒ were generated by using a standard random number genera- tor.

A. Distribution functions In our simulations we have numerically measured the cu- mulative velocity distribution function F͑t,v͒ in the labora- ⌺ ␾͑ ͒ tory frame 0. Given the probability density t,v , the cu- mulative velocity distribution function is defined by

v F͑t,v͒ = ͵ du␾͑t,u͒. ͑69͒ −c In order to obtain F͑t,v͒ from numerical simulations, one simply measures the relative fraction of particles with veloci- ties in the interval ͓−c,v͔. Figure 3 shows the numerically ͑ ͒ determined stationary distribution functions squares , taken FIG. 3. These diagrams show a comparison between numerical at time t=100 ␯−1 and also the corresponding analytical ͑ ͒ and analytical results for the stationary cumulative distribution curves FI/S/HK v . The latter were obtained by numerically ͑ ͒ ⌺ ͑ ͒ function F v in the laboratory frame 0. a In the nonrelativistic integrating Eq. ͑69͒ using the three different stationary den- limit ␤ӷ1 the stationary solutions of the three different FPE are ␾ ͑ ͒ sity functions I/S/HK v from Sec. III. nearly indistinguishable. ͑b͒–͑c͒ In the relativistic limit case ␤ഛ1, As one can see in Fig. 3͑a͒, for low-temperature values however, the stationary solutions exhibit deviations from each corresponding to ␤ӷ1, the three stationary distribution func- other. Because our simulations are based on an Ito-discretization tions are nearly indistinguishable. For high temperatures cor- scheme, the numerical values ͑squares͒ are best fitted by the Ito responding to ␤ഛ1, the stationary solutions exhibit signifi- solution ͑solid line͒. cant quantitative differences, ͓see Figs. 3͑b͒ and 3͑c͔͒. Because our simulations are based on an Ito-discretization ͑ ͒ scheme the numerical values squares are best fitted by the B. Mean-square displacement Ito solution ͑solid line͒. Also note that the quality of the fit is very good for the parameters chosen in the simulations, and In this section we consider the spatial mean-square dis- that this property is conserved over several magnitudes of ␤. This suggests that numerical simulations of the Langevin placement of the free relativistic Brownian motion. Because equations provide a very useful tool if one wishes to study this quantity is easily accessible in experiments, it has played relativistic Brownian motions in more complicated settings an important role in the verification of the nonrelativistic ͑e.g., in higher dimensions or in the presence of additional theory. external fields and interactions͒. In this context, it should As before, we consider an ensemble of N-independent ͑ ͒ ⌺ again be stressed that the appropriate choice of the discreti- Brownian particles with coordinates x͑i͒ t in 0 and initial ͑ ͒ ͑ ͒ zation rule is especially important in applications to realistic conditions x͑i͒ 0 =0,v͑i͒ 0 =0 for i=1,2,...,N. The posi- systems. tion mean value is defined as

016124-9 J. DUNKEL AND P. HÄNGGI PHYSICAL REVIEW E 71, 016124 ͑2005͒

N 1 ͑ ͒ϵ ͑ ͒ ͑ ͒ ¯x t ͚ x͑i͒ t , 70 N i=1 and the related second moment is given by

N 1 2͑ ͒ϵ ͓ ͑ ͔͒2 ͑ ͒ x t ͚ x͑i͒ t . 71 N i=1

The empirical mean-square displacement can then be defined as follows:

␴2͑ ͒ϵ 2͑ ͒ ¯͑ ͒ 2 ͑ ͒ t x t − „x t … . 72 Cornerstone results in the nonrelativistic theory of the one- dimensional Brownian motion are

lim ¯x͑t͒ → 0, ͑73a͒ t→+ϱ

␴2͑t͒ lim → 2Dx, ͑73b͒ t→+ϱ t where the constant FIG. 4. ͑a͒ Mean-square displacement divided by time t as nu- ␤ kBT D merically calculated for different -values in the laboratory frame Dx = = ͑74͒ ⌺ ͑ ͒ m␯ m2␯2 0 rest frame of the heat bath . As evident from this figure, for the relativistic Brownian motion the related asymptotic mean-square ͑ ͒ is the nonrelativistic coefficient of diffusion in coordinate displacement grows linearly with t. b The coordinate space diffu- x ͑␤͒ space ͑not to be confused with noise parameter D͒. sion constant D100 was numerically determined at time t ␯−1 It is therefore interesting to consider the asymptotic be- =100 . The dashed line corresponds to the empirical fitting for- mula Dx͑␤͒=c2␯−1͑␤+2͒−1, which reduces to the classical nonrel- havior of the quantity ␴2͑t͒/t for relativistic Brownian mo- ativistic result Dx Ӎc2 /͑␯␤͒=kT/͑m␯͒ for ␤ӷ2. tions, using again the Ito-relativistic Langevin dynamics from Sec. III B 1. In Fig. 4͑a͒ one can see the corresponding ץ ץ cp ץ ␤ numerical results for different values of . As one can ob- ͑ ͒ ͑ ͒ ͑ ͒ f t,p,x + f t,p,x =− jI/S/HK t,p,x , pץ xץ t ͱm2c2 + p2ץ serve in this diagram, for each value of ␤, the quantity ␴2͑ ͒ t /t converges to a constant value. This means that, at ͑ ͒ ⌺ 76 least in the laboratory frame 0, the asymptotic mean-square displacement of the free relativistic Brownian motions in- which might serve as a suitable starting point for such an creases linearly with t. For completeness, we mention that analysis. Compared to the relativistic Fokker-Planck equa- according to our simulations the asymptotic relation ͑73a͒ tions from Sec. IV, the second term on the left-hand side of holds in the relativistic case, too. ͑76͒ is new. In particular, we recover the relativistic Fokker- In spite of these similarities between nonrelativistic and Planck equations for the marginal density f͑t,p͒, see Sec. III relativistic theory, an essential difference consists of the ex- by integrating Eq. ͑76͒ over a spatial volume with appropri- plicit temperature dependence of the limit value 2Dx. As il- ate boundary conditions. Finally, we mention once again that ͑ ͒ x lustrated in Fig. 4 b , the numerical limit values 2D100, mea- also ͑76͒, as well as all the other results that have been pre- sured at time t=100␯−1, are well fitted by the empirical sented in this section, exclusively refer to the laboratory ⌺ formula frame 0.

c2 Dx = , ͑75͒ V. CONCLUSION ␯͑␤ +2͒ Concentrating on the simplest case of 1+1 dimensions, which reduces to the nonrelativistic result ͑74͒ in the limit we have put forward the Langevin dynamics for the stochas- case ␤ӷ2. tic motion of free relativistic Brownian particles in a viscous We will leave it as an open problem here, to find an ana- medium ͑heat bath͒. Analogous to the nonrelativistic lytical justification for the empirically determined formula Ornstein-Uhlenbeck theory of Brownian motion ͑75͒. Instead we merely mention that, on noting ͑43͒, the ͓17,19,20,44͔, it was assumed that the heat bath can, in good relativistic Fokker-Planck equations for the full-phase space approximation, be regarded as homogenous. Based on this density reads assumption, a covariant generalization of the Langevin equa-

016124-10 THEORY OF RELATIVISTIC BROWNIAN MOTION: … PHYSICAL REVIEW E 71, 016124 ͑2005͒ tions has been constructed in Sec. II. According to these the laboratory coordinate time; the temperature dependence generalized stochastic differential equations, the viscous fric- of the related spatial diffusion constant, however, becomes tion between Brownian particle and heat bath is modeled by more intricate. In principle, the numerical results suggest that a friction tensor ␯␣␤. For a homogeneous heat bath this fric- simulations of the Langevin equations may provide a very tion tensor has the same structure as the pressure tensor of a useful tool for studying the dynamics of relativistic Brown- perfect fluid ͓25͔. In particular, it is uniquely determined by ian particles. In this context it has to be stressed that an ͑ ͒ ␯ the value of the scalar viscous friction coefficient , mea- appropriate choice of the discretization rule is especially im- ͑ sured in the instantaneous rest frame of the particle Sec. portant in applications to realistic physical systems. If, for ͒ IIB1. Similarly, the amplitude of the stochastic force is example, agreement with the kinetic theory ͓29͔ is desirable, also governed by a single parameter D, specifying the Gauss- then a postpoint discretization rule should be used. ian fluctuations of the heat bath, as seen in the instantaneous From the methodical point of view, the systematic relativ- rest frame of the particle ͑Sec. II B 2͒. istic Langevin approach of the present paper differs from In Sec. II C the relativistic Langevin equations have been Schay’s transition probability approach ͓3͔ and also from the derived in special laboratory coordinates, corresponding to a ͓ ͔ specific class of Lorentz frames, in which the heat bath is techniques applied by other authors 4,6,7 . As we shall dis- assumed to be at rest ͑at all times͒. One finds that the corre- cuss in a forthcoming contribution, the above approach can sponding relativistic distribution of the momentum incre- easily be generalized to settings that are more relevant with ͓ ͑ ͒ ments now also depends on the momentum coordinate. This regard to experiments such as the 1+3 -dimensional case, fact is in contrast with the properties of ordinary Wiener the presence of additional external force fields, etc.͔. processes ͓21,23͔, underlying nonrelativistic standard With regard to future work, several challenges remain to Brownian motions with “additive” Gaussian white noise. be solved. For example, one should try to derive an analytic However, as shown in Sec. III it is possible to find an equiva- expression for the temperature dependence of the spatial dif- lent Langevin equation, containing “multiplicative” Gaussian fusion constant. A suitable starting point for such studies white noise. might be the FPE for the full-phase space density given in In order to achieve a more complete picture of the rela- Eq. ͑76͒. Another possible task consists of finding explicit tivistic Brownian motion, the corresponding relativistic exact or at least approximate time-dependent solutions of the Fokker-Planck equations ͑FPE͒ have been discussed in Sec. relativistic FPE. Furthermore, it seems also interesting to III ͑again with respect to the laboratory coordinates with the consider extensions to , as, to some extent, heat bath at rest͒. Analogous to nonrelativistic processes with recently discussed in the mathematical literature ͓13͔. In this multiplicative noise, one can opt for different interpretations context, the physical consequences of the different interpre- of the stochastic differential equation, which result in differ- tations ͑Ito versus Stratonovich versus Hänggi- ent FPE. In this paper, we concentrated on the three most Klimontovich͒ become particularly interesting. popular cases, namely, the Ito, the Stratonovich-Fisk, and the Note added in proof. Recently, we have been informed by Hänggi-Klimontovich interpretations. We discussed and F. Debbasch about two interesting recent papers ͓45,46͔ on a compared the corresponding stationary solutions for a free relativistic generalization of the Ornstein-Uhlenbeck process. Brownian particle. It could be established that only the These two items are related in spirit to the present work: The Hänggi-Klimontovich interpretation is consistent with the authors of those references have postulated a relativistic relativistic Maxwell distribution. This very distribution was Langevin equation with additive noise and a drift term that derived by Jüttner ͓24͔ as the equilibrium velocity distribu- differs from ours; but which also yields the correct relativis- tion of the relativistic ideal gas. Later on, it was also dis- tic Jüttner distribution. Thus, our HK-approach and theirs cussed by Schay in the context of relativistic diffusions ͓3͔ possess the same stationary solution, but notably do exhibit a and by de Groot et al. in the framework of the relativistic different relaxation dynamics. kinetic theory ͓29͔. In Sec. IV we presented numerical results, obtained on the ACKNOWLEDGMENTS basis of an Ito prepoint discretization rule. The simulations indicate that—analogous to the nonrelativistic case—the J. D. would like to thank S. Hilbert, L. Schimansky-Geier, relativistic mean-square displacement grows linearly with and S. A. Trigger for helpful discussions.

͓1͔ A. Einstein, Ann. Phys. ͑Leipzig͒ 17, 891 ͑1905͒. ͓7͔ O. Oron and L. P. Horwitz ͑2003͒, e-print math-ph/0312003. ͓2͔ A. Einstein, Ann. Phys. ͑Leipzig͒ 18, 639 ͑1905͒. ͓8͔ T. H. Boyer, Phys. Rev. D 19, 1112 ͑1979͒. ͓3͔ G. Schay, Ph.D. thesis, Princeton University ͑1961͒, available ͓9͔ T. H. Boyer, Phys. Rev. D 19, 3635 ͑1979͒. through University Microfilms, Ann Arbor, Michigan, https:// ͓10͔ F. Guerra and P. Ruggiero, Lett. Nuovo Cimento Soc. Ital. Fis. wwwlib.umi.com 23, 529 ͑1978͒. ͓4͔ R. M. Dudley, Arkiv Feur Matematik 6, 241 ͑1965͒. ͓11͔ L. M. Morato and L. Viola, J. Math. Phys. 36, 4691 ͑1995͒; ͓5͔ R. Hakim, J. Math. Phys. 6, 1482 ͑1965͒. 37, 4769͑E͒͑1996͒. ͓6͔ U. Ben-Ya’acov, Phys. Rev. D 23, 1441 ͑1981͒. ͓12͔ A. Posilicano, Lett. Math. Phys. 42,85͑1997͒.

016124-11 J. DUNKEL AND P. HÄNGGI PHYSICAL REVIEW E 71, 016124 ͑2005͒

͓13͔ J. Franchi and Y. Le Jan ͑2004͒, e-print math.PR/0403499. NJ, 1990͒. ͓14͔ A. Einstein, Ann. Phys. ͑Leipzig͒ 17, 549 ͑1905͒. ͓29͔ S. R. de Groot, W. van Leeuwen, and C. G. van Weert, Rela- ͓15͔ A. Einstein and M. von Smoluchowski, Untersuchungen über tivistic Kinetic Theory: Principles and Applications ͑North- die Theorie der Brownschen Bewegung/Abhandlungen über Holland, Amsterdam, 1980͒. die Brownsche Bewegung und verwandte Erscheinungen, 3rd ͓30͔ S. A. Trigger, Phys. Rev. E 67, 046403 ͑2003͒. ed. ͑Harri Deutsch, Frankfurt, 1999͒, Vol. 199. ͓31͔ Y. L. Klimontovich, Phys. Usp. 37, 737 ͑1994͒. ͓16͔ M. C. Wang and G. E. Uhlenbeck, Rev. Mod. Phys. 17, 323 ͓32͔ U. Erdmann, W. Ebeling, L. Schimansky-Geier, and F. Sch- ͑1945͒. weitzer, Eur. Phys. J. B 15, 105 ͑2000͒. ͓17͔ G. E. Uhlenbeck and L. S. Ornstein, Phys. Rev. 36, 823 ͓33͔ K. Ito, Proc. Imp. Acad. Tokyo 20, 519 ͑1944͒. ͑1930͒. ͓34͔ K. Ito, Mem. Am. Math. Soc. 4,51͑1951͒. ͓18͔ S. Chandrasekhar, Rev. Mod. Phys. 15,1͑1943͒. ͓35͔ F. Schwabl, Statistische Mechanik ͑Springer, Berlin, 2000͒. ͓19͔ P. Hänggi and H. Thomas, Phys. Rep. 88, 207 ͑1982͒. ͓36͔ R. L. Stratonovich, Vestn. Mosk. Univ., Ser. 1: Mat., Mekh. 1, ͓20͔ N. G. Van Kampen, Stochastic Processes in and 3 ͑1964͒. Chemistry ͑North-Holland, Amsterdam, 2003͒. ͓37͔ R. L. Stratonovich, SIAM J. Control Optim. 4, 362 ͑1966͒. ͓21͔ N. Wiener, J. Math. Phys. 58, 131 ͑1923͒. ͓38͔ D. Fisk, Ph.D. thesis, Michigan State University, Dept. of Sta- ͓22͔ M. Grigoriu, Stochastic Calculus: Applications in and tistics ͑1963͒. Engineering ͑Birkhäuser, Boston, 2002͒. ͓39͔ D. Fisk, Trans. Am. Math. Soc. 120, 369 ͑1965͒. ͓23͔ I. Karatzas and S. E. Shreve, Brownian Motion and Stochastic ͓40͔ P. Hänggi, Helv. Phys. Acta 51, 183 ͑1978͒. Calculus, 2nd ed. Graduate Texts in Vol. 113 ͓41͔ P. Hänggi, Helv. Phys. Acta 53, 491 ͑1980͒. ͑Springer-Verlag, New York, 1991͒. ͓42͔ P. Hänggi and H. Thomas, Phys. Rep. 88, 207 ͑1982͒, see pp. ͓24͔ F. Jüttner, Ann. Phys. ͑Leipzig͒ 34, 856 ͑1911͒. 292–294. ͓25͔ S. Weinberg, Gravitation and Cosmology ͑Wiley, 1972͒. ͓43͔ J. L. Synge, The Relativistic Gas ͑North-Holland, Amsterdam, ͓26͔ R. W. Lindquist, Ann. Phys. ͑N.Y.͒ 37, 487 ͑1966͒. 1957͒. ͓27͔ J. M. Stewart, Non-Equilibrium Relativistic Kinetic Theory, ͓44͔ P. Hänggi and P. Jung, Adv. Chem. Phys. 89, 239 ͑1995͒. Lecture Notes in Physics Vol. 10 ͑Springer-Verlag, Berlin, ͓45͔ F. Debbasch, K. Mallick, and J. P. Rivet, J. Stat. Phys. 88, 945 1971͒. ͑1997͒. ͓28͔ R. L. Liboff, Kinetic Theory ͑Prentice Hall, Englewood Cliffs, ͓46͔ F. Debbasch, J. Math. Phys. 45, 2744 ͑2004͒.

016124-12