ARTICLE IN PRESS Physica A 350 (2005) 245–262 www.elsevier.com/locate/physa Advection and dispersion in time and space B. Baeumera,1, D.A. Bensonb,2, M.M. Meerschaertc,,3 a Department of Mathematics and Statistics, University of Otago, Dunedin, New Zealand b Desert Research Institute, 2215 Raggio Parkway, Reno NV 89512, USA c Department of Physics, University of Nevada, Reno NV 89557, USA Received 19 July 2004; received in revised form 20 October 2004 Available online 24 November 2004 Abstract Previous work showed how moving particles that rest along their trajectory lead to timenonlocal advection–dispersion equations. If the waiting times have infinite mean, the model equation contains a fractional time derivative of order between 0 and 1. In this article, we develop a new advection–dispersion equation with an additional fractional time derivative of order between 1 and 2. Solutions to the equation are obtained by subordination. The form of the time derivative is related to the probability distribution of particle waiting times and the subordinator is given as the first passage time density of the waiting time process which is computed explicitly. r 2004 Elsevier B.V. All rights reserved. PACS: 05.40.Fb; 02.50.Cw; 05.60.k; 47.55.Mh Keywords: Anomalous diffusion; Continuous time random walks; First passage time; Fractional calculus; Subdiffusion; Power laws Corresponding author. Department of Mathematics and Statistics, Graduate Program in Hydrologic Sciences, Unviersity of Nevada, Reno, NV 89557-0045, USA. E-mail addresses: bbaeumer@maths.otago.ac.nz (B. Baeumer), dbenson@dri.edu (D.A. Benson), mcubed@unr.edu (M.M. Meerschaert). 1 B. Baeumer was partially supported by the Marsden Fund administered by the Royal Society of New Zealand. 2 D.A. Benson was supported by DOE-BES grant DE-FG03-98ER14885 and NSF grants DES-9980489 and DMS-0139943. 3 M.M. Meerschaert was partially supported by NSF grants DMS-0139927 and DMS-0417869, and the Marsden Fund administered by the Royal Society of New Zealand. 0378-4371/$ - see front matter r 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.physa.2004.11.008 ARTICLE IN PRESS 246 B. Baeumer et al. / Physica A 350 (2005) 245–262 1. Introduction Continuous time random walks (CTRW) can be used to derive governing equations for anomalous diffusion [1–4]. The CTRW is a stochastic process model for the movement of an individual particle [5,6]. In the long-time limit, the process converges to a simpler form whose probability densities solve the governing equation, leading to a useful model for anomalous diffusion. For a simple random walk with zero-mean, finite-variance particle jumps, the limit process is a Brownian motion AðtÞ governed by the classical diffusion equation qp=qt ¼ q2 p=qx2 where pðx; tÞ is the probability density of the random variable AðtÞ: For symmetric infinitevariance jumps (i.e., those with probability density function tails that fall off like jxj1a [7] with some index 0oao2), the limit process AðtÞ is an a-stable Lévy motion, and the governing equation becomes qp=qt ¼ qa p=qjxja [8]. When waiting times between the jumps are introduced, the limiting process is altered via subordination [9–11]. In this study, we examine the case where the waiting times are independent of jump size, also called an ‘‘uncoupled’’ CTRW. For infinite mean waiting times (whose probability distribution is assumed to decay algebraically with some index 0ogo1) the limit process is AðEðtÞÞ where EðtÞ is the inverse or first passage time process for the g-stable subordinator. By virtue of its construction, the process EðtÞ counts the number of particle jumps by time tX0; accounting for the waiting time between particle jumps. In the scaling limit, EðtÞ keeps track of the possibly nonlinear link between real time and the operational time that a particle actually spends in motion. The governing equation becomes qg p=qtg ¼ qa p=qjxja [2,3]. Some applications [12] seem to indicate a time derivative of order 1ogp2: In this paper, we develop one such equation by extending the CTRW approach to processes with finite-mean waiting times, and compute the distribution of the relevant first passage time process. 2. The model In the usual CTRW formalism, the long-time limit for the waiting time process is a g-stable subordinator DðtÞ [2]. Then the inverse Lévy process EðtÞ ¼ inffx : DðxÞ4tg counts the number of particle jumps by time tX0; reflecting the fact that the time T n of the nth particle jump and the number N t ¼ maxfn : T n ptg of jumps by time t are also inverse processes. When the waiting times between particle jumps have heavy tails with 0ogo1 (i.e., infinite mean), subordination of the particle location process AðtÞ via the inverse Lévy process EðtÞ is necessary in the long-time limit to account for the amount of time that a particle is not participating in the motion process. The subordination leads to a time derivative of order g in the governing equation of motion [3]. When waiting times have heavy tails of order 1ogp2; meaning that the probability of waiting longer than t falls off like tg ; a different model is needed [13]. In this case, convergence of the waiting time process requires centering to the mean waiting time w, which is not necessary when 0ogo1: Accounting for this leads to a waiting time process W ðtÞ ¼ DðtÞ þ wt where DðtÞ is a completely positively ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 247 skewed4 stable Lévy process with index g; so that W ðtÞ is a Lévy process with drift. The drift ensures that W ðtÞ ! 1 with probability one as t ! 1: Since g41; the process W ðtÞ is not strictly increasing, so it is proper to use MaxðtÞ ¼ supfW ðuÞ : 0puptg to represent the particle jump times. This substitution is explained as a particle that makes more than one jump without intervening rest periods [13]. Then the inverse or first passage time process HðtÞ ¼ inffx : MaxðxÞXtg counts the number of particle jumps. The first passage time process HðtÞ serves as the subordinator in place of (and when go1 is identical to) EðtÞ: We show this in subsequent sections. 3. First passage time density According to the previous section we model the waiting time process as a Lévy process with drift; i.e., W ðtÞ has characteristic function g E½eixW ðtÞ ¼ etðiwxþaðixÞ Þ ; where w40 is the mean waiting time and a is a shape variable akin to the variance. For the remainder of this paper we make the assumption w ¼ 1; which entails no loss of generality, since any mean time w can be recovered by a simple rescaling in time. In order to compute the density of the process H we use the fact that PfHðTÞ4sg ¼ PfMaxðsÞoTg for all s; TX0: Theorem 1 in Ref. [14] shows that the distribution of the maximum process Mðs; TÞ ¼ PfMaxðsÞoTg satisfies Z 1Z 1 u euslT d T ðMðs; TÞÞ ds 0 0 Z 1Z 1 1 l ix þ aðixÞg ¼ exp ð1Þ g dx dx 2p u 1 xðx ilÞ xðx ðix þ aðixÞ ÞÞ for all l; u40: The integrand has two poles in the upper complex halfplane, at x ¼ il and whenever x ¼ ix þ aðixÞg : That the second equality holds only once, follows from the following Lemma (for a proof see Appendix A) investigating the inverse function of azg z: This involves the following region. Let a40; 0oapp=g and sinðyÞ sinðaÞ org1 o OðaÞ:¼ reiy : aoyoa and ; a sinðgyÞ a sinðgaÞ where we take sinðaÞ=a sinðgaÞ ¼ 1 when a ¼ p=g: Lemma 1. Let a40; 1ogp2: There exists a unique holomorphic function q : Cnð1; aðg 1ÞðagÞg=ð1gÞ ! Oðp=gÞ such that aqðzÞg qðzÞ ¼ z : Furthermore, there exists an analytic F with supt40 t11=g ezt jF ðtÞjo1 for all R 1function g=ð1gÞ zt 0ozoaðg 1ÞðagÞ such that 0 e F ðtÞ dt ¼ 1=qðzÞ for z40: 4 The skew is irrelevant in the normal case g ¼ 2: ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 248 Using the Lemma, we can simplify (1) to (see Appendix B) Z 1Z 1 1 l=qðuÞ euslT d T ðMðs; TÞÞ ds ¼ : u þ l alg 0 0 Let Hðs; TÞ ¼ PfHðTÞpsg and recall that Mðs; TÞ ¼ PfMaxðsÞoTg: Then Hðs; TÞ ¼ PfHðTÞpsg ¼ PfMaxðsÞXTg ¼ 1 Mðs; tÞ; and hence we can integrate by parts to get Z 1Z 1 Z 1Z 1 euslT d T ðMðs; TÞÞ ds ¼ l euslT Mðs; TÞ dT ds 0 0 0 0 Z 1Z 1 ¼l euslT ð1 Hðs; TÞÞ ds dT 0 0 Z 1Z 1 ¼ 1=u l euslT Hðs; TÞ ds dT : 0 0 In other words, the Laplace–Laplace transform of the distribution function of the first passage time process is given by Z 1Z 1 1 1 l=qðuÞ euslT Hðs; TÞ ds dT ¼ ul lðu þ l alg Þ 0 0 ¼ 1 alg1 þ u=qðuÞ : uðu þ l alg Þ ð2Þ In the following theorem, we invert this Laplace–Laplace transform to obtain an expression for the cumulative distribution function of the first passage time process (corresponding in our model to the ‘‘number of jumps by time t’’). ~ Theorem 2. Let 1ogp2; let m be the function with Laplace transform mðuÞ ¼ 1=qðuÞ given by Lemma 1 and gg be the maximally skewed, standard g-stable density; i.e., the g Fourier transform of gg is g^ g ðkÞ ¼ eðikÞ : Then the cumulative distribution function of the first passage time of a Lévy-stable waiting time process with drift is given by ! Z s Z 1 mðs uÞ tu gg ðuÞ du þ gg Hðs; tÞ ¼ du : (3) 1=g ðauÞ1=g ðtsÞ=ðasÞ1=g 0 ðauÞ Remark 3. For tb0; Z 1 Hðs; tÞ ðtsÞ=ðasÞ1=g gg ðuÞ du ¼ PðW ðsÞXtÞ (4) in view of the fact that W ðsÞ is identically distributed with ðasÞ1=g W g þ s where W g is the stable random variable with density gg : If W ðsÞ were an increasing process, the left-hand and right-hand expressions in (4) would be equal. Hence the second term in (3) accounts for the fact that W ðsÞ is not monotone. ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 249 4. The differential equations In this section, we show that the density of the uncoupled CTRW limit process solves a fractional partial differential equation with an extra time derivative term of order 1ogp2 on the right-hand side whose coefficient is negative. This is in contrast to Mainardi and colleagues’ [15–18] investigation of the transition via fractional derivatives from the second order hyperbolic equation to the first order parabolic case. Our equations stay parabolic in nature (given an elliptic operator in space). We begin with a related result concerning the first passage time density. We say that a mild solution to a fractional partial differential equation is a function whose Laplace transform solves the equivalent algebraic equation in Laplace–Laplace space. The following theorem employs the Caputo derivative ðd=dtÞg ; which can be defined for 1ogp2 by requiring that ðd=dtÞg F ðtÞ has Laplace transform lg F~ ðlÞ lg1 F ð0Þ lg2 F 0 ð0Þ where F~ ðlÞ is the Laplace transform of F ðtÞ; see for example Refs. [19,20]. Theorem 4. There exists a unique distribution f such that the density uðt; sÞ ¼ dHðs; tÞ=ds of the first passage time distribution Hðs; tÞ in (3) is the unique mild solution to g d d d a (5) uðt; sÞ þ uðt; sÞ ¼ uðt; sÞ þ f ðsÞdðtÞ dt dt ds with conditions: uð0; sÞ ¼ dðsÞ; uðt; 0Þ ¼ ut ð0; sÞ ¼ 0 8s; t40; and s7!uðt; sÞ is a probability density for all t40: For any other distribution f Eq. (5) has no solution. We now extend this result to incorporate more general spatial derivative operators. When the CTRW particle jumps are in the generalized domain of attraction of an operator stable limit with probability distribution n; the long-time limiting particle location process AðtÞ is operator stable with distribution nt [21]. For a simple random walk, the density pðx; tÞ of AðtÞ defines R a strongly continuous semigroup ðTðtÞÞtX0 on L1 ðRd Þ via the formula TðtÞf ðxÞ ¼ f ðx yÞpðy; tÞ dy: If L is the generator of this semigroup, then Cðx; tÞ ¼ TðtÞf ðxÞ also solves the abstract Cauchy problem dC=dt ¼ LC; Cð0Þ ¼ f : For an uncoupled CTRW with infinite mean waiting times, the limiting particle location is given by the subordinated process AðEðtÞÞ: The density hðx; tÞ of this process solves a fractional Cauchy problem ðd=dtÞg h ¼ Lh where 0ogo1 [22]. The next theorem extends this result to 1ogp2: Theorem 4.1 in Ref. [13] shows that the long-time CTRW limit process in this caseR is AðHðtÞÞ; and Corollary 4.2 in Ref. [13] shows that this limit process has 1 density 0 pðx; sÞd s ðHðt; sÞÞ: The next theorem shows that this density is the Green’s function solution to a fractional partial differential equation (6) with time derivative of order 1ogp2: This result extends Theorem 3.1 in Ref. [22] showing that these CTRW limits provide stochastic solutions of fractional Cauchy problems with time derivative of order 1ogp2: Theorem 5. Let L be the generator of an operator stable semigroup ðTðtÞÞtX0 on ^ L1 ðRd Þ such that for f 2 L1 ðRd Þ; Lf ¼ F1 ðLðkÞf ðkÞÞ: Let a40; 1ogp2: Then there ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 250 exists a unique distribution g such that g d d CðtÞ þ CðtÞ ¼ LCðtÞ þ dðtÞg a dt dt (6) with conditions C 0 ð0Þ ¼ 0; Cð0Þ ¼ f ; CðtÞ 2 L1 ðRd Þ for all tX0 has a mild solution and this unique solution is given by Z 1 Cðt; xÞ ¼ TðsÞf ðxÞd s ðHðt; sÞÞ : 0 For any other distribution g Eq. (6) has no solution. 5. Examples In this section we give two examples showing how the more differentiated scaling in time influences the resulting model. Example 6. Consider a random walk composed of zero-mean, finite-variance Gaussian jumps with intervening waiting times that are in the domain of attraction of a Gaussian with unit mean and variance. In this case, the first passage time density of the Gaussian waiting times (with drift) is well known; see, for example, Ref. [23]. The governing equation becomes a d2 d d2 Cðt; xÞ þ Cðt; xÞ ¼ D 2 Cðt; xÞ þ dðtÞg 2 dt dt dx with Cð0; xÞ ¼ f ðxÞ and qC=qt ¼ 0 when t ¼ 0: The solution is given by Z 1 t 1 ðt sÞ2 1 x2 pffiffiffiffiffiffiffiffiffiffi exp pffiffiffiffiffiffiffiffiffiffiffi exp Cðt; xÞ ¼ ds%f ; s 4pas 4as 4Ds 4pDs 0 with ‘%’ being the convolution operator in x. Example 7. To generalize the previous example, let the waiting times be regularly varying with unit mean and infinite variance, so the limit of waiting times DðtÞ converges to a Lévy motion with index 1ogo2: Furthermore, let the jump sizes be symmetric and heavy tailed with index 1oap2: Then the particle position density Cðx; tÞ; assuming an initial particle distribution Cð0; xÞ ¼ f ðxÞ and qC=qt ¼ 0 when t ¼ 0; follows g d d d qa Cðt; xÞ þ a Cðt; xÞ þ Cðt; xÞ ¼ v Cðt; xÞ þ dðtÞg : (7) dt dt dx qjxja ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 251 10 t = 0.1; a = 0.1 t = 1; a = 0.1 t = 2; a = 0.1 t = 0.1; a = 1 t = 1; a = 1 t = 2; a = 1 1 0.1 0.01 0.001 0.0001 -10 -5 0 5 10 Fig. 1. Solutions to Example 6 with pulse initial condition. Solid lines are with a ¼ 0:1 and dashed lines are for a ¼ 1; shown for various t and with diffusion parameter D set to 1. The mild solution is given by Z 1 1 x vs Cðt; xÞ ¼ g %f d s ðHðt; sÞÞ : s1=a a s1=a 0 These two examples show that the particle motion has classical advective (drift) motion with additional dispersion in both space and time. In the first example, the effect of the second time derivative vanishes for large time. Note, however, that when the coefficient on the second time derivative a is larger, the density is more peaked at the origin and has more weight in the tails (Fig. 1). The main effect of the parameter g; when less than two, is to cause particles to spend heavy-tailed amounts of time in an immobile state that decays very slowly (Fig. 2). The density of the first passage time, which provides a map between the number of jumps (a particle’s operational time) and the clock time, shows fewer jumps at any time for lower values of g: For smaller values of g; the time dispersion is noticeable for prolonged time periods (Fig. 2), hence the effect of the g-order derivative is important for longer periods. The solution to Example 7 is obtained by subordinating the shifted Lévy motion against these densities. 6. Conclusions The study of classical random walks and, more recently, CTRW, has focused on the spatial dispersion of the particles. Temporal dispersion in the limit process is ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 252 10 γ = 1.1 γ = 1.5 Density h(s,t) γ = 1.9 t=1 t = 10 t = 100 t = 1,000 t = 10,000 1 0.1 0.01 0.001 0.0001 0.00001 0.1 1 10 100 1,000 10,000 0.1 1 10 100 1,000 10,000 0.1 1 10 100 1,000 10,000 Number of jumps (s) at time t Fig. 2. First passage time densities d s ðHðs; tÞÞ corresponding to g-stable waiting time processes with a ¼ 0:1 and unit drift. In the model, the first passage time tracks the random number of jumps at time t. commonly assumed to be restricted to the infinite-mean waiting time CTRW [6,9,1,4]. However, by viewing the time process in a similar mathematical light as the space process, one sees that the time ‘‘drift’’ follows a different scaling procedure than the time dispersion. The time drift measures the linear portion of the map between clock time and the number of jumps, resulting in the familiar first time derivative in the governing equation of motion. If deviations in the waiting times have a power law distribution with finite mean but infinite variance, then the effects of the deviations do not disappear in the limit, and the governing equation has an additional time operator of order 1ogo2: If deviations in the waiting times have finite mean and variance, this procedure leads to an additional second-order time derivative. The resulting model gives a sharper description of space-time diffusive processes. Appendix A. Proof of Lemma 1 First we show uniqueness. Let 0oaop=g; ( ) sinðyÞ 1=ðg1Þ iy G ¼ e : 0pypa a sinðgyÞ and ( Gr ¼ sinðaÞ a sinðgaÞ 1=ðg1Þ ) iy e : apypa : Then q OðaÞ ¼ G:¼G þ Gr Gþ is clearly a simply closed path around OðaÞ (see for example Ref. [24, Theorem 10.40]). Let pðzÞ ¼ azg z : ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 253 Then ! sinðyÞ 1=ðg1Þ iy sinðyÞ g=ðg1Þ igy sinðyÞ 1=ðg1Þ iy ¼a p e e e a sinðgyÞ a sinðgyÞ a sinðgyÞ sinðyÞ g=ðg1Þ sinðyÞ 1=ðg1Þ ¼a cosðgyÞ cosðyÞ a sinðgyÞ a sinðgyÞ ! sinðyÞ g=ðg1Þ sinðyÞ 1=ðg1Þ sinðgyÞ sinðyÞ i a a sinðgyÞ a sinðgyÞ sinðyÞ 1=ðg1Þ sinðyÞ cosðgyÞ cosðyÞ ¼ a sinðgyÞ sinðgyÞ 1=ðg1Þ sinðyÞ sinððg 1ÞyÞ ¼ : a sinðgyÞ sinðgyÞ A quick calculation shows that for 0oyop=g; y7! sinðyÞ= sinðgyÞ is an increasing function which implies that y7!ðsinðyÞ=a sinðgyÞÞ1=ðg1Þ sinððg 1ÞyÞ= sinð½g=ðg 1Þðg 1ÞyÞ is also increasing. Since sinðyÞ 1=ðg1Þ sinððg 1ÞyÞ ¼ aðg 1ÞðagÞg=ð1gÞ lim y!0 a sinðgyÞ sinðgyÞ we have that the image of G under p is a path on the negative real axis, " # sinðaÞ 1=ðg1Þ sinððg 1ÞaÞ ; aðg 1ÞðagÞg=ð1gÞ : pðG Þ ¼ a sinðgaÞ sinðgaÞ Investigating pðGr Þ we see that for z 2 Gr ; sinðaÞ g=ðg1Þ igy sinðaÞ 1=ðg1Þ iy pðzÞ ¼ a e e a sinðgaÞ a sinðgaÞ sinðaÞ 1=ðg1Þ sinðaÞ ðcosðgyÞ þ i sinðgyÞÞ cosðyÞ i sinðyÞ ¼ a sinðgaÞ sinðgaÞ sinðaÞ 1=ðg1Þ sinðaÞ cosðgyÞ sinðaÞ sinðgyÞ ¼ cosðyÞ þ i sinðyÞ : a sinðgaÞ sinðgaÞ sinðgaÞ Using again that y7! sinðyÞ= sinðgyÞ is increasing for y40; the imaginary part is positive iff y is positive. Furthermore, sinðaÞ 1=ðg1Þ sinðaÞ jpðzÞjX 1 a sinðgaÞ sinðgaÞ for all z 2 Gr ; and this lower bound tends to infinity as a ! p=g: Hence we obtain for a large enough that pðGr Þ is a closed contour going once counterclockwise around the origin, which implies that pðGÞ is a closed contour going once counterclockwise around the origin. ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 254 Fix w 2 Cnð1; aðg 1ÞðagÞg=ð1gÞ ; and then choose aop=g (a depends on w) such that pðGÞ goes around w. Using the counting formula for zeros and poles, we see that there is exactly one z 2 OðaÞ (see for example Ref. [25, Theorem 13.2.2]) such that pðzÞ ¼ w: Furthermore (see for example Ref. [26, p. 153]), Z 1 p0 ðzÞ dz qðwÞ:¼p1 ðwÞ ¼ z 2pi G pðzÞ w is a holomorphic function on Cnð1; aðg 1ÞðagÞg=ð1gÞ : Next, we show that 1=q is the Laplace transform of an analytic function f on ð0; 1Þ with t11=g eot f ðtÞ bounded for o4 aðg 1ÞðagÞg=ð1gÞ : Since aqðzÞg qðzÞ ¼ z we have that for jzj large enough, ða þ 1ÞjqðzÞjg 4jaqðzÞg qðzÞj ¼ jzj: Furthermore, qðzÞ 2 Oðp=gÞ and thus qðzÞ is bounded away from zero. Hence there exists M40 such that MjqðzÞjg Xjzj (A.1) for all zeð1; aðg 1ÞðagÞ rðzÞ:¼ g=ð1gÞ : Let d 1 1 ¼ qðzÞ2 dz qðzÞ agqðzÞg1 1 using the fact that dq=dz ¼ 1=ðdp=dzÞ: The function z7!1=agqðzÞg1 1 has a single pole at z ¼ aðg 1ÞðagÞg=ð1gÞ and is bounded off a neighborhood of that pole. Choose oo0 such that o4 aðg 1ÞðagÞg=ð1gÞ : Then for Sa :¼freid : r40; aodoag we have that z o 1 o1 sup jðz oÞrðzÞj ¼ sup 2 g1 1 z2oþSdþp=2 z2oþSdþp=2 qðzÞ agqðzÞ for all 0odop=2: Let Cþ ¼ fz 2 C : RðzÞ40g: By the analytic representation theorem for Laplace transforms [27, Theorem 2.6.1] there exists a holomorphic function f : Cþ ! C such that supz2Cþ jeoz f ðzÞjo1 and Z 1 rðzÞ ¼ ezt f ðtÞ dt : 0 Furthermore, sup jzz RðzÞ40 1=g 2=g z zðg1Þ=g rðzÞj ¼ o1 : qðzÞ2 agqðzÞg1 1 Using the complex representation theorem ([27, Theorem 2.5.1] with b ¼ 1=g and qðzÞ ¼ z1=g rðzÞ) there exists g 2 C½0; 1Þ with supt40 t1=g jgðtÞjo1 such that Z 1 z1=g rðzÞ ¼ z1=g ezt gðtÞ dt : 0 By the uniqueness of the Laplace transform f ¼ g and hence sup jt1=g eot f ðtÞjo1 : tX0 ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 255 Clearly, F ðtÞ ¼ f ðtÞ=t 2 L1 ð0; 1Þ and hence, using Fubini, Z 1 Z 1Z 1 Z 1 zt st e F ðtÞ dt ¼ e f ðtÞ dt ds ¼ rðsÞ ds ¼ 1=qðzÞ z 0 0 z for z40: Thus the Laplace transform of F ðtÞ is 1=q and the growth conditions follow with z ¼ o: & Appendix B. Simplifying the Laplace–Laplace transform of the maximum process Clearly, z40 implies qðzÞ40: Thus the second pole of the integrand of (1), l ix þ aðixÞg ; xðx ilÞ xðx ðix þ aðixÞg ÞÞ is at x ¼ iqðxÞ: For Gn ¼ fneiy : 0pyppg there exists M40 such that Z 1 l ix þ aðixÞg lim g dx n!1 2p G xðx ilÞ xðx ðix þ aðixÞ ÞÞ n Z 1 M p lim dx ¼ lim M=2n ¼ 0 : n!1 2p G n2 n!1 n Thus, we can apply the residue theorem and obtain Z 1 1 l ix þ aðixÞg dx 2p 1 xðx ilÞ xðx ðix þ aðixÞg ÞÞ Z 1 l 1 þ aðixÞg1 ¼ lim g dx n!1 2pi ½n;n[G ðx ilÞ xðx ðix þ aðixÞ ÞÞ n 1 þ alg1 l 1 þ aðqðxÞÞg1 g þ xðx ðl þ al ÞÞ iqðxÞ il xði þ igaqðxÞg1 Þ l þ alg l 1 þ aðqðxÞÞg1 ¼ þ : xðx þ l alg Þ xðqðxÞ lÞ ð1 gaqðxÞg1 Þ ¼l Since q is an inverse function we can compute its derivative d 1 qðxÞ ¼ dx agqðxÞg1 1 as in the proof of Lemma 1. Thus for u4alg l; Z 1 l þ alg l 1 þ aðqðxÞÞg1 dx þ g xðqðxÞ lÞ ð1 gaqðxÞg1 Þ xðx þ l al Þ u Z 1 1 þ axg1 ¼ ½lnð1 þ ðl alg Þ=xÞ1 dx l u g qðuÞ ðax xÞðx lÞ ARTICLE IN PRESS 256 B. Baeumer et al. / Physica A 350 (2005) 245–262 ¼ lnð1 þ ðl alg Þ=uÞ ½lnð1 l=xÞ1 qðuÞ 1 l=qðuÞ u ul=qðuÞ ¼ ln : ¼ ln 1 þ ðl alg Þ=u u þ l alg Using L’Hopital’s rule, we see that for alg l40; limu!alg l lnððu ul=qðuÞÞ=ðu þ l alg ÞÞo1; since the expression inside the logarithm tends to a finite constant as u ! alg l: Then it is not hard to show that the above equality holds for all u40 (integrate from u to alg l e and alg l þ e to infinity and then let e ! 0). Thus, using (1), Z 1Z 1 1 l=qðuÞ euslT d T ðMðs; TÞÞ ds ¼ : u þ l alg 0 0 Appendix C. Proof of Theorem 2 Clearly, the inverse in u of (2) is given by5 Z s g1 ~ Hðs; lÞ ¼ ð1 al Þ expðrðl alg ÞÞ dr 0 Z s þ mðs rÞ expðrðl alg ÞÞ dr 0 Z s 1 expðsðl alg ÞÞ þ mðs rÞ expðrðl alg ÞÞ dr : ¼ l 0 Inverting with respect to l is a bit more delicate. Using the complex inversion formula (see for example Ref. [28, Theorem 7.6]) we obtain Z t Z 1 elt 1 expðsðl alg ÞÞ Hðs; rÞ dr ¼ 2pi cþiR l l 0 Z s þ mðs rÞ expðrðl alg ÞÞ dr dl 0 Z 1 elt 1 expðsðl alg ÞÞ ¼ dl 2pi cþiR l l Z Z s 1 elt expðrðl alg ÞÞ dl dr mðs rÞ þ 2pi l 0 cþiR the second equality holding due to Fubini since Z s Z jelt j j expðrðl alg ÞÞj dl dr jmðs rÞj 0 cþiR jlj 5 As the Laplace transform in l has to stay bounded as l ! 1; we see that convolution with the function m has to have the following effect for all s40: m% expðsðl alg ÞÞ expðsðl alg ÞÞ=l : ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 Z 257 Z j expðrðl alg ÞÞj dl dr jlj 0 cþiR Z s Z j expðrðl alg ÞÞj dl dr þ ect jmðs rÞj jlj s=2 cþiR Z s=2 Z j expðrðl alg ÞÞj dr dl pect sup jmðrÞj jlj s=2oros cþiR 0 Z j expðrðl alg ÞÞj þ ect kmk1 sup dl jlj s=2oros cþiR ¼ ect s=2 jmðs rÞj o1 : Making a change of variables we see that on the right-hand side we have an expression akin to the formula for the inversion of the Fourier transform. Z 1 elt 1 expðsðl alg ÞÞ dl 2pi cþiR l l Z Z s 1 elt expðrðl alg ÞÞ dl dr mðs rÞ þ 2pi cþiR l 0 Z ect 1 iut 1 expðsðiu þ c aðiu þ cÞg ÞÞ e ¼ du 2p 1 ðiu þ cÞ2 Z Z s ect 1 iut expðrðiu þ c aðiu þ cÞg ÞÞ du dr : ðC:1Þ mðs rÞ e þ iu þ c 2p 1 0 Now the Fourier transform of a shifted g-stable distribution is ! Z 1 g 1 xr ikx e g dx ¼ eikrþraðikÞ : 1=g g 1=g ðarÞ ðarÞ 1 Since ecx gg ðxÞ is bounded for all x 2 R for some cX0 (e.g., Ref. [29, Theorem 4.7.1]) we obtain that ! Z 1 cx g e x r eiux g dx ¼ erðiuþcþaðiuþcÞ Þ : 1=g g 1=g ðarÞ ðarÞ 1 Hence, the expressions in (C.1) are indeed inverse Fourier transforms and Z ect 1 iut 1 expðsðiu þ c aðiu þ cÞg ÞÞ e du 2p 1 ðiu þ cÞ2 Z Z s ect 1 iut expðrðiu þ c aðiu þ cÞg ÞÞ du dr mðs rÞ e þ iu þ c 2p 1 0 ! Z t Z w 1 xs ¼t g dx dw 1=g g ðasÞ1=g 1 1 ðasÞ ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 258 Z Z s t 1 1 ðarÞ1=g mðs rÞ þ 0 gg xr ! dx dr : ðarÞ1=g Therefore, Z t 1 ! Z s mðs rÞ tr ! gg dx þ dr 1=g ðasÞ1=g ðarÞ1=g 0 ðarÞ ! ! Z s Z 1 1 xs mðs rÞ tr gg gg dx þ dr ¼ 1=g ðasÞ1=g ðasÞ1=g ðarÞ1=g t 0 ðarÞ ! Z s Z 1 mðs uÞ tu gg ðuÞ du þ gg du : & ¼ 1=g ðauÞ1=g ðtsÞ=ðasÞ1=g 0 ðauÞ Hðs; tÞ ¼ 1 1 ðasÞ1=g gg xs ðC:2Þ Appendix D. Proof of Theorem 4 Assume there exists a solution to (5). Using the Caputo derivative and the fact that ~~ rÞ lg1 : Then uð0; sÞ ¼ dðsÞ; the Laplace–Laplace transform of ðd=dtÞg uðt; sÞ is lg uðl; it follows easily that g1 ~ ~~ rÞ ¼ 1 al þ fg ðrÞ : uðl; r þ l al ~ rÞjp1: Hence, the Laplace–Laplace Since s7!uðt; sÞ is a probability density, juðt; transform for each r is analytic in the right halfplane in l: Thus 1 alg1 þ f~ðrÞ ¼ 0 if r þ l alg ¼ 0 or equivalently, l ¼ qðrÞ; q given by Lemma 1. Hence g l þ al ¼ r=qðrÞ : f~ðrÞ ¼ l Since the range of l7!alg l contains the right halfplane, f~ is uniquely determined, and so is f. The first passage time distribution has Laplace–Laplace transform Z 1Z 1 1 alg1 þ r=qðrÞ erslT Hðs; TÞ ds dT ¼ rðr þ l alg Þ 0 0 by (2). Since Hð0; TÞ ¼ 0 for all T40 we have that Z 1Z 1 1 alg1 þ r=qðrÞ erslT d s ðHðs; TÞÞ dT ¼ r þ l alg 0 0 (D.1) and hence the density of the first passage time distribution has the same Laplace–Laplace transform as the solution to the differential equation and it is therefore a mild solution. & ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 259 Appendix E. Proof of Theorem 5 Taking the Fourier transform of (6) we obtain g d ^ kÞ ¼ LðkÞ ^ kÞ þ dðtÞgðkÞ ^ kÞ þ d Cðt; ^ Cðt; ^ a : Cðt; dt dt Taking the Laplace transform in t yields ~^ ~^ ~^ ^ Cðl; ^ alg Cðl; kÞ f^ðkÞ ¼ LðkÞ kÞ þ alg1 f^ðkÞ þ lCðl; kÞ þ gðkÞ : Thus ^ f^ðkÞ alg1 f^ðkÞ þ gðkÞ ~^ : Cðl; kÞ ¼ g ^ al þ l LðkÞ ~ Since CðlÞ 2 L1 ðRd Þ we know that its Fourier transform has to be bounded for all RðlÞ40: This implies that the numerator of the above equation has to be zero ^ whenever the denominator is equal to zero, or whenever qðLðkÞÞ ¼ l: Hence g1 ^ ^ f^ðkÞ=qðLðkÞÞ ^ ^ ^ Þf ðkÞ ¼ LðkÞ gðkÞ ¼ ð1 þ aqðLðkÞÞ is uniquely determined. Thus, ^ ^ 1 alg1 LðkÞ=qð LðkÞÞ ~^ f^ðkÞ : Cðl; kÞ ¼ g ^ al þ l LðkÞ R1 To see that CðtÞ ¼ 0 TðsÞf d s ðHðt; sÞÞ has the same Fourier–Laplace transform, take the Fourier transform and observe that Z 1 ^ ^ kÞ ¼ esLðkÞ d s ðHðt; sÞÞf^ðkÞ : Cðt; 0 But this is the Laplace transform in s of the first passage time density evaluated at ^ LðkÞ times f^ðkÞ; i.e., taking the Laplace transform in t, using (D.1) we see that Z 1Z 1 ^ ~^ eltsðLðkÞÞ d s ðHðt; sÞÞ dt f^ðkÞ Cðl; kÞ ¼ 0 ¼ 0 g1 ^ ^ LðkÞ=qð LðkÞÞ f^ðkÞ ^ al þ l LðkÞ 1 al g and therefore Cðt; xÞ is indeed the mild solution of (6). Finally we prove that g is a distribution. The proof depends on the fact that for some positive real constant C we have ^ jLðkÞjpC maxfkkk; kkk2 g for all k 2 Rd : To see this, use the Lévy Representation [7] to write Z 1 ikx ^ LðkÞ ¼ ika kAk þ eikx 1 fðdxÞ ; 2 1 þ kxk2 xa0 (E.1) ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 260 where a 2 Rd ; A is a nonnegative definite matrix, and f is a s-finite Borel measure on Rd nf0g such that Z minf1; kxk2 gfðdxÞo1 : (E.2) xa0 The integral term I in the Lévy Representation satisfies jIjpjI 1 j þ jI 2 j with Z ikx I1 ¼ eikx 1 fðdxÞ 1 þ kxk2 0okxko1 ¼ I 11 þ I 12 ; where Z ikx jI 11 j ¼ ðe 1 ikxÞfðdxÞ 0okxko1 Z 1 p kkk2 kxk2 fðdxÞ 0okxko1 2 pC 1 kkk2 and Z jI 12 j ¼ ik x fðdxÞ 2 1 þ kxk 0okxko1 Z kxk3 pkkk fðdxÞ 2 0okxko1 1 þ kxk x pC 2 kkk ; while eikx 1 Z jI 2 j ¼ kxkX1 ikx fðdxÞ 1 þ kxk2 pD þ D þ Dkkk ; where D ¼ ffx : kxkX1go1 using the fact that jeikx j ¼ 1 and kxk=ð1 þ kxk2 Þp1 for kxkX1: Then (E.1) holds. ^ f^ðkÞ=qðLðkÞÞ ^ ^ Since gðkÞ ¼ LðkÞ and jqðzÞjXM 0 jzj1=g for almost all z 2 Rd by (A.1) we have LðkÞ ^ 1 ^ jLðkÞj11=g p ^ qðLðkÞÞ M0 and note that 1 1=g40: Using (E.1) we obtain 11=g ^ ^ ; kkk22=g g jgðkÞjpM 1 jf ðkÞj maxfkkk ARTICLE IN PRESS B. Baeumer et al. / Physica A 350 (2005) 245–262 261 and then it follows easily that for some D40 we have Z ^ dko1 : ð1 þ kkk2 ÞD gðkÞ Now Example 7.12 (b) on p. 191 of Ref. [30] shows that g is a tempered distribution. & References [1] A. Compte, Stochastic foundations of fractional dynamics, Phys. Rev. E 53 (1996) 4191–4193. [2] M.M.Meerschaert,H.P. Scheffler, Limit theorems for continuous time random walks with infinite mean waiting times, J. Appl. Probab. 41 (3) (2004) 623–638. [3] M.M. Meerschaert, D.A. Benson, H.P. Scheffler, B. Baeumer, Stochastic solution of space-time fractional diffusion equations, Phys. Rev. E 65 (2002) 1103–1106. [4] R. Metzler, J. Klafter, The random walk’s guide to anomalous diffusion: a fractional dynamics approach, Phys. Rep. 339 (2000) 1–77. [5] E.W. Montroll, G.H. Weiss, Random walks on lattices, II, J. Math. Phys. 6 (1965) 167–181. [6] H. Scher, M. Lax, Stochastic transport in a disordered solid. I. Theory, Phys. Rev. B 7 (10) (1973) 4491–4502. [7] M.M. Meerschaert, H.P. Scheffler, Limit Distributions for Sums of Independent Random Vectors, Wiley, New York, 2001. [8] M.M. Meerschaert, D.A. Benson, B. Baeumer, Multidimensional advection and fractional dispersion, Phys. Rev. E 59 (1999) 5026–5028. [9] M. Shlesinger, J. Klafter, Y.M. Wong, Random walks with infinite spatial and temporal moments, J. Stat. Phys. 27 (1982) 499–512. [10] M.M. Meerschaert, D.A. Benson, P. Becker-Kern, H.-P. Scheffler, Governing equations and solutions of anomalous random walk limits, Phys. Rev. E 66 (2003) 060102(R). [11] E. Scalas, R. Gorenflo, F. Mainardi, Uncoupled continuous-time random walks: solution and limiting behavior of the master equation, Phys. Rev. E 69 (2004) 011107. [12] R. Haggerty, S.A. McKenna, L.C. Meigs, On the late-time behavior of tracer test breakthrough curves, Water Resour. Res. 36 (2000) 3467–3479. [13] P. Becker-Kern, M.M. Meerschaert, H.P. Scheffler, Limit theorem for continuous time random walks with two time scales, J. Applied Probab. 41 (2) (2004) 455–466. [14] G. Baxter, M.D. Donsker, On the distribution of the supremum functional for processes with stationary independent increments, Trans. Am. Math. Soc. 85 (1957) 73–87. [15] F. Mainardi, Fractional relaxation-oscillation and fractional diffusion wave phenomena, Chaos Soliton Fract. 7 (1996) 1461–1477. [16] F. Mainardi, The fundamental solutions for the fractional diffusion-wave equation, Appl. Math. Lett. 9 (1996). [17] F. Mainardi, Yu. Luchko, G. Pagnini, The fundamental solution of the space-time fractional diffusion equation, Frac. Calc. Appl. Anal. 4 (2001) 153–192. [18] F. Mainardi, G. Pagnini, R. Gorenflo, Mellin transform and subordination laws in fractional diffusion processes, Frac. Calc. Appl. Anal. 6 (2003) 441–459. [19] M. Caputo, Linear models of dissipation whose Q is almost frequency independent, Part II, Geophys. J. R. Astron. Soc. 13 (1967) 529–539. [20] I. Podlubny, Fractional Differential Equations, Academic Press, San Diego, 1999. [21] M.M. Meerschaert, D.A. Benson, B. Baeumer, Operator Lévy motion and multiscaling anomalous diffusion, Phys. Rev. E 63 (2001) 1112–1117. [22] B. Baeumer, M. Meerschaert, Stochastic solutions for fractional Cauchy problems, Frac. Calc. Appl. Anal. 4 (2001) 481–500. ARTICLE IN PRESS 262 B. Baeumer et al. / Physica A 350 (2005) 245–262 [23] M.C.K. Tweedie, Statistical properties of inverse gaussian distributions, I, Ann. Math. Stat. 28 (2) (1957) 362–377. [24] W. Rudin, Real and Complex Analysis, third ed., McGraw-Hill, New York, 1987. [25] R. Remmert, Theory of Complex Functions, Graduate Texts in Mathematics, Springer, Berlin, New York, vol. 122, 1991. [26] L. Ahlfors, An introduction to the theory of analytic functions of one complex variable, Complex Analysis, third ed., McGraw-Hill, New York, 1978. [27] W. Arendt, C.J.K. Batty, M. Hieber, F. Neubrander, Vector-Valued Laplace Transforms and Cauchy Problems, Birkhauser, Verlag, Basel, Boston, Berlin, 2001. [28] D.V. Widder, The Laplace Transform, Princeton University Press, Princeton, NJ, 1946. [29] V.V. Uchaikin, V.M. Zolotarev, Chance and Stability, Stable Distributions and Their Applications, VSP, Utrecht, 1999. [30] W. Rudin, Functional Analysis, second ed., McGraw-Hill, New York, 1991.