Skip to main content

Fundamentals of Fractional Transport

  • Chapter
  • First Online:
A Primer on Complex Systems

Part of the book series: Lecture Notes in Physics ((LNP,volume 943))

  • 1220 Accesses

Abstract

In many physical systems, transport plays a central role. By transport we mean any macroscopic process that moves some physical quantity of interest across the system. For instance, transport processes are responsible for transporting mass and heat, including pollutants, throughout the atmosphere. Or for transporting water and energy, but also debris, across the ocean. Or for transporting plasma density and energy, but also impurities, out of a tokamak. Or for transporting angular momentum out of an accretion disk, but also mass and energy into the black hole at its center.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 59.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 79.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Note also that, in all of these examples, transport is mainly carried out by turbulence!

  2. 2.

    Other famous examples that share the same spirit that Fick’s law are Fourier’s law, that establishes the flux of heat to be proportional to the temperature gradient, q = −κ∂T/∂x, where κ is known as the thermal conductivity; and Newton’s law, that relates the kinematic stress and the velocity shear in a perpendicular direction, τ x  = ν∂u x /∂y. ν is known as the kinematic viscosity.

  3. 3.

    The continuity equation simply expresses mathematically the fact that the number of particles must be conserved.

  4. 4.

    This relation could be made even more general since, as it stands, it assumes translational invariance in time and space. Otherwise, the kernel would depend on either time or space, either explicitly or through other fields. In this chapter, however, we will always assume, for simplicity’s sake, that spatial and temporal translational invariance holds.

  5. 5.

    This last statement establishes a direct connection between transport and the concept of memory that was discussed in Chap. 4.

  6. 6.

    The quotes are used to stress the fact that these two models are just that: mathematical models that provide with a naive, although rich, idealization of the processes that might be taking place at the microscopic level.

  7. 7.

    It might appear to some readers that we discuss them in much deeper detail than what would be needed for the purposes of this subsection. The reasons for doing this will become clearer soon, since these models will provide the basis of the generalizations leading to fractional transport.

  8. 8.

    It is possible to define more general CTRWs. For instance, one could define non-separable CTRWs by specifying instead a joint probability ξ(Δx, Δt) of taking a jump of size Δx after having waited for a lapse of time Δt. It is also possible to define CTRWs that are not invariant under spatial and/or temporal translations, and that depend on time, space or both either explicitly [9, 10] or through dependencies on other fields [11, 12].

  9. 9.

    More precisely, the initial condition distributes the N walkers uniformly in the interval (−dx/2, dx/2), where dx is the spacing used to discretize the problem.

  10. 10.

    From our discussions in Chap. 2 about the central limit theorem, the fact that Gaussians seem to be related to diffusion also suggests that the underlying physical processes behind diffusive transport are of an additive, and not multiplicative nature. This idea will also play a role when extending the description of transport beyond diffusion, as will be seen soon.

  11. 11.

    The impatient reader may ignore this relatively lengthy calculation and jump directly to the final result (Eq. 5.18), and continue reading without any loss of coherence.

  12. 12.

    A primer on the Laplace transform has been included in Appendix 1 for those readers unfamiliar with it. The Laplace transform is very useful to simplify the solution of initial-value problems. When used with partial differential equations, they are often combined with Fourier transforms, that were already introduced in Appendix 1 of Chap. 2.

  13. 13.

    We will use the notation \(\bar f(k,s)\) to denote the double Laplace-Fourier transform. It will be remembered that \(\skew 2\hat f(k,t)\) is used for the Fourier transform, whilst \(\skew 2\tilde f(x,s)\) is used for the Laplace transform.

  14. 14.

    Since the CTRW, as it has been defined here, is translationally invariant in both time and space, its propagator can only depend on the difference x − x 0 and not separately on x and x 0.

  15. 15.

    The spatial and temporal variables associated to the Fourier (k) and Laplace (s) variables are, respectively, Δx = x − x 0 and τ.

  16. 16.

    We will discuss several examples of systems and regimes in which this is indeed the case, in the second part of this book. Particularly, in Chaps. 6 and 9.

  17. 17.

    At that time, we referred to it as the only uncorrelated member of the fractional Brownian motion (fBm) family (see Eq. 3.64).

  18. 18.

    The additive nature of the physical processes behind diffusion are also apparent in the formulation of Langevin equation, where the noise acts as a surrogate for the instantaneous displacement.

  19. 19.

    In the sake of simplicity we have assumed symmetry for the underlying microscopic motion. Naturally, asymmetric CTRWs could also be considered and might be important in some contexts (see, for instance Sect. 5.5).

  20. 20.

    We will refer to the tail-index of these extremal Lévy pdfs as β (instead of α) in order to avoid any confusion with the index of the step-size pdf. We will also use τ (instead of σ) for the scale parameter.

  21. 21.

    Or, superdiffusively, as this behaviour is often referred to.

  22. 22.

    It should be remarked that the same fluid limit would be obtained if one defines the CTRW in terms of non-Lévy pdfs, as long as they have the same asymptotic behaviours. This is completely analogous to what happened in the case of the CTRW with jumps following the Gaussian p(Δx) and waiting times distributed according to the exponential ψ(Δt).

  23. 23.

    In fact, note that the classical diffusion equation is recovered if β → 1 and α → 2.

  24. 24.

    The date in which fractional derivatives were born is actually known exactly. In a letter to L’Hôpital in 1695, Leibniz wrote: Can the meaning of derivatives with integer order be generalized to derivatives with non-integer orders?” Apparently, L’Hôpital was curious about this, and replied himself writing back to Leibniz: What if the order will be 1/2? Leibniz replied that—with a letter dated September 30, 1695—(The idea) will lead to a paradox, from which one day useful consequences will be drawn. Appendix 2 contains a crash course on fractional derivatives, that might be worth reading before the reader continues to read this section.

  25. 25.

    More specifically, \(\mbox{ }_{-\infty }D^{a}_{x}\) is a left-sided Riemann-Liouville fractional derivative of order r, with starting point at − (see Appendix 1).

  26. 26.

    \(\mbox{ }_0D^{q}_{t}\) is in fact, a left-sided Riemann-Liouville fractional derivative of order q ∈ (0,  1), with starting point at t = 0 (see Appendix 1).

  27. 27.

    For the sake of conciseness, we will adopt the convention that L [2,0,0,σ](x) becomes the Gaussian law with zero mean when α = 2. This decision can be justified because the Fourier transform of the symmetric Lévy pdf, \(\hat L_{[\alpha ,0,0,\sigma ]}(k) = \exp (-\sigma ^\alpha |k|{ }^\alpha )\) becomes the Fourier transform of the Gaussian, \(\hat N_{[0,2\sigma ^2]}(k) = \exp (-\sigma ^2k^2)\), for α = 2.

  28. 28.

    Clearly, \({\mathrm {E}}_1(z) = \exp (z)\).

  29. 29.

    This expression of the fTe propagator can be useful to avoid the always delicate evaluation of Laplace transforms at small values of s; instead, we just need to sum the Mittag-LeffLer function up to the value of the index n that ensures the desired level of convergence. However, the inverse of the Fourier transform of Eq. 5.71 must be performed numerically.

  30. 30.

    Due to the assumed symmetry of the underlying motion, all odd moments vanish.

  31. 31.

    This exponent, H, corresponds in fact to the self-similarity exponent of both fBm and fLm (see Chap. 3). It is also the exponent used in the definition of the fractional Langevin equation.

  32. 32.

    It is not known how to do this in most other cases. This is in fact an active area of current research that needs to be developed greatly in the next years.

  33. 33.

    The method is trivially extended to accommodate external sources, if they are present and information about them is available. However, the method is no good if the sources are unknown. One has to use then other approaches, such as the propagator method that is discussed next.

  34. 34.

    For instance, one could use local methods such as the Levenberg-Marquardt algorithm, or global methods, such as genetic algorithms [30].

  35. 35.

    Propagator methods require, in many cases, the use of some kind of tracer field or tracer particles, easily distinguishable from the system background but with similar dynamical behaviour. These tracers must be initialized in a very localized setup, and then followed in time without any further addition of tracers. In this way, one could get an estimate of the system transport exponents without having to consider external sources. We will say more about tracers soon, when discussing Lagrangian methods.

  36. 36.

    That is, if the flow velocity is defined by the field, v(r, t), its Lagrangian trajectory that passes through r 0 at time t 0 is the solution of the differential equation \({\mathbf {\dot r}} = {\mathbf {v}}({\mathbf {r}}, t)\), with r(t 0) = r 0 .

  37. 37.

    Eulerian methods, on the other hand, often use tracer fields, such as oil or dye, in order to be able to follow the evolution in time of a perturbation advected by a system on a Eulerian grid. The use of the tracer field permits to ignore the presence of any external drive when needed to sustain the system and that might be unknown. Or to be able to taylor the initial perturbation at will, as it is required in order to estimate a propagator.

  38. 38.

    Think, for instance, of radioactive isotopes or the polyethylene particles or oil droplets used in particle image velocimetry (PIV) studies of turbulence.

  39. 39.

    In fluids, one might want to adjust their mass or buoyancy. In plasmas, on the other hand, one might also want to use chargeless tracer particles in order to avoid magnetic drifts.

  40. 40.

    Since the fTe is one-dimensional, we will assume either that the displacements occur along one direction of interest, or that we are focusing on one particular component of the motion only.

  41. 41.

    It should be remembered that u is a random number uniformly distributed in [0,  1].

  42. 42.

    In fact, they are not horizontal since particles only advance one position per iteration. However, the scale of the temporal axis used in the figure makes them look so.

  43. 43.

    We have implicitly assumed that the range of cells were the marked particles were initially dropped have similar dynamics, which is the case for the running sandpile.

  44. 44.

    The consequences of boundedness have also been addressed within the CTRW and fTe frameworks by considering truncated Lévy distributions [19, 40,41,42,43]. It is however much more complicated to establish a connection with fractional transport equations that in the unbounded case, since one can no longer rely on the advantages of the Fourier representation to take the long-time, large-distance limit.

  45. 45.

    In fact, no rigorous derivation of any fTe exists, to the best of our knowledge, that obtains the values of the fractional exponents from the physical equations of motion based on Newton’s law or Hamiltonian dynamics.

  46. 46.

    This is, in fact, why the marked particle jumps appeared as (almost) horizontal lines in Fig. 5.8!

  47. 47.

    We will always use the tilde (i.e., \(\skew 2\tilde f\)) to represent the Laplace transform of a function (i.e., f(t)) throughout this book. Also, the Laplace variable will always be represented by the letter s.

  48. 48.

    Note that it is not even required that f(t) be continuous for it to have a Laplace transform.

  49. 49.

    Right-sided RL fractional integrals can also be defined:

    $$\displaystyle \begin{aligned} ^b D^{-p}_{ t} f(t) \equiv \frac{1}{\varGamma(p)}\int_t^b (\tau - t)^{p-1} f(\tau) d\tau, \end{aligned} $$
    (5.122)

    b is known as the ending point.

  50. 50.

    An integral of order n is equivalent to carrying out n consecutive integrals on f:

    $$\displaystyle \begin{aligned} {\mathrm{I}}^n(f) = \int_0^t dt_1 \int_0^{t_1} dt_2 \cdots \int_0^{t_n} f(t_n)dt_n. \end{aligned} $$
    (5.123)
  51. 51.

    In the case of the right-sided RL fractional integral, the Fourier transform, for ending point b = +, is given by:

    $$\displaystyle \begin{aligned} {\mathrm{F}}\left[{}^{\infty} D^{-p}_{t}\cdot f(t) \right] = (-i\omega)^{-p} \skew2\hat f(\omega). \end{aligned} $$
    (5.126)
  52. 52.

    Again, right-sided RL fractional derivatives can also be defined:

    $$\displaystyle \begin{aligned} ^b D^{\,p}_{ t} f(t) \equiv \frac{1}{\varGamma(k-p)} \frac{d^k}{dt^k}\int_t^b (\tau-t)^{k-p-1} f(\tau) d\tau, \end{aligned} $$
    (5.129)

    Their properties are analogous to the left-sided counterpart.

  53. 53.

    For the right side RL derivatives, this property becomes:

    $$\displaystyle \begin{aligned} (-1)^m \frac{d^m }{dt^m} \,{\cdot}\, ^b D^{\,p}_{ t } f(t) =_a D^{\,p+m}_{t} f(t). \end{aligned} $$
    (5.134)
  54. 54.

    For the right-sided fractional integral with ending point b = , the Fourier transform is given by:

    $$\displaystyle \begin{aligned} {\mathrm{F}}\left[{}^{\infty} D^{\,p}_{ t} \cdot f(t) \right] = (-i\omega)^{p} \skew2\hat f(\omega). \end{aligned} $$
    (5.138)
  55. 55.

    The attentive reader will note that we have changed the name of the independent variable (now x) to represent that the range is no (−, ) instead of the range (0, ) used when defining the RL fractional derivatives. Similarly, we will referred to the Fourier variable as k, instead of ω.

  56. 56.

    For the right-sided derivatives, the divergence of the RL derivative happens at the ending point, t = b. The Caputo fractional derivative is defined very similarly:

    $$\displaystyle \begin{aligned} \mbox{}_{\mathrm{C}}^bD_t^p f := \mbox{}^bD_x^p \left[\,f - \sum_{i=0}^{{\mathrm{int}}(p)} f^{(i)}(b)(b-t)^{i} \right]. \end{aligned} $$
    (5.149)
  57. 57.

    Note that we have defined t −1 as the first point int he series. This is done on purpose, since we will need it to discretize the integral that starts at t 0.

References

  1. Fick, A.: Über Diffusion. Annalen der Physik 94, 59 (1855)

    Article  ADS  Google Scholar 

  2. Balescu, R.: Aspects of Anomalous Transport in Plasmas. Institute of Physics, Bristol (2005)

    Book  Google Scholar 

  3. Sanchez, R., Newman, D.E.: Topical Review: Self-organized-criticality and the Dynamics of Near-marginal Turbulent Transport in Magnetically Confined Fusion Plasmas. Plasma Phys. Controlled Fusion 57, 123002 (2015)

    Article  ADS  Google Scholar 

  4. Field, S., Witt, J., Nori, F., Ling, X.: Superconducting Vortex Avalanches. Phys. Rev. Lett. 74, 1206 (1995)

    Article  Google Scholar 

  5. Isichenko, M.: Percolation, Statistical Topography and Transport in Random Media. Rev. Mod. Phys. 64, 961 (1992)

    Article  ADS  MathSciNet  Google Scholar 

  6. Bazant, Z.P.: Scaling of Quasibrittle Fracture: Hypotheses of Invasive and Lacunar Fractality, Their Critique and Weibull Connection. Int. J. Fract. 83, 41 (1997)

    Article  Google Scholar 

  7. Montroll, E.W., Weiss, G.H.: Random Walks on Laticess II. J. Math. Phys. 6, 167 (1965)

    Article  ADS  MATH  Google Scholar 

  8. Einstein, A.: On the Movement of Small Particles Suspended in Stationary Liquids Required by the Molecular-Kinetic Theory of Heat. Annalen der Physik 17, 549 (1905)

    Article  ADS  Google Scholar 

  9. Shugard, W., Reiss, H.: Derivation of the Continuous-Time Random Walk Equation in Non-homogeneous Lattices. J. Chem. Phys. 65, 2827 (1976)

    Article  ADS  Google Scholar 

  10. Barkai, E., Metzler, R., Klafter, J.: From Continuous Time Random Walks to the Fractional Fokker-Plack Equation. Phys. Rev. E 61, 132 (2000)

    Article  ADS  MathSciNet  Google Scholar 

  11. van Milligen, B.P., Sanchez, R., Carreras, B.A.: Probabilistic Finite-Size Transport Models for Fusion: Anomalous Transport and Scaling Laws. Phys. Plasmas 11, 2272 (2004)

    Article  ADS  Google Scholar 

  12. Sanchez, R., Carreras, B.A., van Milligen, B.P.: Fluid Limit of Nonintegrable Continuous-Time Random Walks in Terms of Fractional Differential Equations. Phys. Rev. E 71, 011111 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  13. Morse, P.M., Feshbach, H.: Methods of Mathematical Physics. McGraw Hill, New York (1953)

    MATH  Google Scholar 

  14. Langevin, P.: Sur la theorie du mouvement brownien. C.R. Acad. Sci. (Paris) 146, 530 (1908)

    Google Scholar 

  15. Gardiner, C.W.: Handbook of Stochastic Methods. Springer, New York (1997)

    MATH  Google Scholar 

  16. Compte, A.: Statistical Foundations of Fractional Dynamics. Phys. Rev. E 53, 4191 (1996)

    Article  ADS  Google Scholar 

  17. Metzler, R., Klafter, J.: The Random Walk’s Guide to Anomalous Diffusion: A Fractional Dynamics Approach. Phys. Rep. 339, 1 (2000)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. Zaslavsky, G.M.: Chaos, Fractional Kinetics, and Anomalous Transport. Phys. Rep. 371, 461 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Metzler, R., Klafter, J.: The Restaurant at the End of the Random Walk: Recent Developments in the Description of Anomalous Transport by Fractional Dynamics. J. Phys. A 37, R161 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Samorodnitsky, G., Taqqu, M.S.: Stable Non-Gaussian Processes. Chapman & Hall, New York (1994)

    MATH  Google Scholar 

  21. Podlubny, I.: Fractional Differential Equations. Academic Press, New York (1998)

    MATH  Google Scholar 

  22. Laskin, N., Lambadaris, I., Harmantzis, F.C., Devetsikiotis, M.: Fractional Lévy Motion and Its Application to Network Traffic Modelling. Comput. Netw. 40, 363 (2002)

    Article  Google Scholar 

  23. Mandelbrot, B.B., van Ness, J.W.: Fractional Brownian Motions, Fractional Noises and Applications. SIAM Rev. 10, 422 (1968)

    MATH  Google Scholar 

  24. Calvo, I., Sanchez, R.: The Path Integral Formulation of Fractional Brownian Motion for the General Hurst Exponent. J. Phys. A 41, 282002 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  25. Calvo, I., Sanchez, R.: Fractional Lévy Motion Through Path Integrals. J. Phys. A 42, 055003 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Weiss, G.H., Dishon, M., Long, A.M., Bendler, J.T., Jones, A.A., Inglefield, P.T., Bandis, A.: Improved Computational Methods for the Calculation of Kohlrausch-Williams/Watts Decay Functions. Polymers 35, 1880 (1994)

    Article  Google Scholar 

  27. Medina, J.S., Prosmiti, R., Villarreal, P., Delgado-Barrio, G., Aleman, J.V.: Frequency Domain Description of Kohlrausch Response Through a Pair of Havriliak-Negami-Type Functions: An Analysis of Functional Proximity. Phys. Rev. E 84, 066703 (2011)

    Article  ADS  Google Scholar 

  28. Mainardi, F., Luchko, Y., Pagnini, G.: The Fundamental Solutions for the Fractional Diffusion-Wave Equation. Appl. Math. Lett. 9, 23 (1996)

    Article  MathSciNet  Google Scholar 

  29. Uchaikin, V.V.: Montroll-Weiss Problem, Fractional Equations and Stable Distributions. Int. J. Theor. Phys. 39, 2087 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  30. Nocedal, J., Wright, S.J.: Numerical Optimization. Springer, Heidelberg (2006)

    MATH  Google Scholar 

  31. Press, W.H., Teukolsky, S.A., Vetterling, W.T., Flannery, B.P.: Numerical Recipes. Cambridge University Press, Cambridge (2007)

    MATH  Google Scholar 

  32. Saichev, A., Zaslavsky, G.M.: Fractional Kinetic Equations: Solutions and Applications. Chaos 7, 753 (1997)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  33. del-Castillo-Negrete, D., Carreras, B.A., Lynch, V.E.: Fractional Diffusion in Plasma Turbulence. Phys. Plasmas 11, 3854 (2004)

    Google Scholar 

  34. Dif-Pradalier, G., Diamond, P.H., Grandgirard, V., Sarazin, Y., Abiteboul, J., Garbet, X., Ghendrih, Ph., Strugarek, A., Ku, S., Chang, C.S.: On the Validity of the Local Diffusive Paradigm in Turbulent Plasma Transport. Phys. Rev. E 82, 025401 (2010)

    Article  ADS  Google Scholar 

  35. Balescu, R.: Anomalous Transport in Turbulent Plasmas and Continuous-Time Random Walks. Phys. Rev. E 51, 4807 (1995)

    Article  ADS  Google Scholar 

  36. Carreras, B.A., Lynch, V.E., Zaslavsky, G.M.: Anomalous Diffusion and Exit Time Distribution of Particle Tracers in Plasma Turbulence Model. Phys. Plasmas 8, 5096 (2001)

    Article  ADS  Google Scholar 

  37. Mier, J.A., Sanchez, R., Newman, D.E., Garcia, L., Carreras, B.A.: On the Nature of Transport in Near-critical Dissipative-Trapped-Electron-Mode Turbulence: Effect of a Subdominant Diffusive Channel. Phys. Plasmas 15, 112301 (2008)

    Article  ADS  Google Scholar 

  38. Mier, J.A., Sanchez, R., Newman, D.E., Garcia, L., Carreras, B.A.: Characterization of Nondiffusive Transport in Plasma Turbulence via a Novel Lagrangian Method. Phys. Rev. Lett. 101, 165001 (2008)

    Article  ADS  Google Scholar 

  39. Carreras, B.A., Lynch, V.E., Newman, D.E., Zaslavsky, G.M.: Anomalous Diffusion in a Running Sandpile Model. Phys. Rev. E 60, 4770 (1999)

    Article  ADS  Google Scholar 

  40. Mantegna, R., Stanley, H.E.: Stochastic Process with Ultraslow Convergence to a Gaussian: The Truncated Lévy Flight. Phys. Rev. Lett. 73, 2946 (1994)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  41. Chechkin, A., Gonchar, V., Klafter, J., Metzler, R.: Fundamentals of Lévy Flight Processes. Adv. Chem. Phys. 133B, 439 (2006)

    Google Scholar 

  42. Cartea, A., del-Castillo-Negrete, D.: Fluid Limit of the Continuous-Time Random Walk with General Lévy Jump Distribution Functions. Phys. Rev. E 76, 041105 (2007)

    Google Scholar 

  43. van Milligen, B.P., Calvo, I., Sanchez, R.: Continuous Time Random Walks in Finite Domains and General Boundary Conditions: Some Formal Considerations. J. Phys. A 41, 215004 (2008)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Oldham, K.B., Spanier, J.: The Fractional Calculus. Academic Press, New York (1974)

    MATH  Google Scholar 

  45. Miller, K.S., Ross, B.: An Introduction to the Fractional Calculus and Fractional Differential Equations. Wiley, New York (1993)

    MATH  Google Scholar 

  46. Schiff, J.L.: The Laplace Transform: Theory and Applications. Springer, Heidelberg (1999)

    Book  MATH  Google Scholar 

  47. Samko, S.G., Kilbas, A.A., Marichev, O.I.: Fractional Integrals and Derivatives: Theory and Applications. Gordon and Breach, New York (1993)

    MATH  Google Scholar 

  48. Feller, W.: An Introduction to Probability Theory and Its Applications. Wiley, New York (1968)

    MATH  Google Scholar 

  49. Gorenflo, R., Mainardi, F., Moretti, D., Pagnini, G., Paradisi, P.: Discrete Random Walk Models for Space-Time Fractional Diffusion. Chem. Phys. 284 521 (2002)

    Article  ADS  MATH  Google Scholar 

  50. Carpinteri, A., Mainardi, F.: Fractals and Fractional Calculus in Continuum Mechanics. Springer, New York (1997)

    Book  MATH  Google Scholar 

  51. Lynch, V., Carreras, B.A., del-Castillo-Negrete, D., Ferreira-Mejias, K.M., Hicks, H.R.: Numerical Methods for the Solution of Partial Differential Equations of Fractional Order. J. Comput. Phys. 192, 406 (2003)

    Google Scholar 

  52. del-Castillo-Negrete, D.: Fractional Diffusion Models of Nonlocal Transport. Phys. Plasmas 13, 082308 (2006)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendices

Appendix 1: The Laplace Transform

The Laplace transform Footnote 47 of a function f(t), t > 0, is defined as [46]:

$$\displaystyle \begin{aligned} {\mathrm{L}}[\,f(t)] := \skew2\tilde f(s) \equiv \int_{0}^\infty f(t)e^{-st} dt. \end{aligned} $$
(5.112)

The main requirement for f(t) to have a Laplace transform is that it cannot grow as t → as fast as an exponential. Otherwise, the Laplace integral would not converge.Footnote 48 A list of common Laplace transforms is given in Table 5.1.

Table 5.1 Some useful Laplace transforms

The Laplace transform is a linear operation that has interesting properties. A first one has to do with the translation of the independent variable in Laplace space. That is,

$$\displaystyle \begin{aligned} \skew2\tilde f(s - a) = {\mathrm{L}}\left[ \exp(at) f(t) \right]. \end{aligned} $$
(5.113)

Another one, with the scaling of the independent variable,

$$\displaystyle \begin{aligned} \skew2\tilde f(s/a) = a {\mathrm{L}}\left[\,f(a t) \right]. \end{aligned} $$
(5.114)

The Laplace transform of the n-th order derivative of a function satisfies:

$$\displaystyle \begin{aligned} {\mathrm{L}}\left[ \frac{d^n f}{dt^n} \right] = s^n\skew2\tilde f(s) - s^{n-1}f(0) - s^{n-2}f'(0) -\cdots - f^{(n-1)}(0), \end{aligned} $$
(5.115)

whilst the Laplace transform of the integral of a function,

$$\displaystyle \begin{aligned} {\mathrm{L}}\left[ \int_0^t f(t')dt' \right] = \frac{\skew2\tilde f(s)}{s}. \end{aligned} $$
(5.116)

These properties are very useful to solve linear systems of ordinary differential equations [13]. Another interesting result is known as the convolution theorem. It refers to the temporal convolution of two functions [46],

$$\displaystyle \begin{aligned} h(t) = \int_0^t f(t') g(t - t') dt', \end{aligned} $$
(5.117)

whose Laplace transform is equal to the product of the Laplace transforms of the two functions,

$$\displaystyle \begin{aligned} \tilde h(s) = \skew2\tilde f(s)\cdot \tilde g(s). \end{aligned} $$
(5.118)

We will also discuss a property particularly useful in the context of self-similar functions. The Laplace transform of a power law, f(t) = t a, t ≥ 0 is given by \(\skew 2\tilde f(s) = \varGamma (a+1)/s^{1+a}\), for a > −1, where Γ(x) is Euler’s gamma function [46]. This result, can also be extended to any function that asymptotically scales as (a > −1),

$$\displaystyle \begin{aligned} f(t) \sim t^a,~~~t\rightarrow \infty,~~~\Longleftrightarrow ~~~\skew2\tilde f(s) \sim s^{-(1+a)},~~s\rightarrow 0. \end{aligned} $$
(5.119)

Finally, we will mention that the inverse Laplace transform is given by [46]:

$$\displaystyle \begin{aligned} f(t) = \frac{1}{2\pi i} \int_{c -i\infty}^{c+i\infty} \skew2\tilde f(s) e^{st} ds, \end{aligned} $$
(5.120)

where s varies along an imaginary line, since c is real and must be chosen so that the integral converges.

Appendix 2: Riemann-Liouville Fractional Derivatives and Integrals

Fractional integrals and derivatives were introduced as interpolants between integrals and derivatives of integer order [21, 44, 45]. In this book, we will always use the Riemann-Lioville (RL) definition of fractional integrals and derivatives.

Riemann-Liouville Fractional Integrals

The left-sided RL fractional integral of order p > 0 of a function f(t) is defined as [21]:

$$\displaystyle \begin{aligned} _{a} D^{-p}_{t} f(t) \equiv \frac{1}{\varGamma(p)}\int_a^t (t-\tau)^{p-1} f(\tau) d\tau, \end{aligned} $$
(5.121)

A negative superscript, − p, is used to reveal that we are dealing with a RL fractional integral. a is known as the starting point of the integral.Footnote 49 As advertised, it can be shown that Eq. 5.121 reduces, for p = n > 0, to the usual integral of order n.Footnote 50 But they have some other interesting properties. For instance, they satisfy a commutative composition rule [21]. Indeed, one has that for p, q > 0:

(5.124)

The Laplace transform of the RL fractional integral, if the starting point is a = 0, is particularly simple [21]. It is given by:

$$\displaystyle \begin{aligned} L\left[{}_{0} D^{-p}_{t}\cdot f(t) \right] = s^{-p} \skew2\tilde f(s). \end{aligned} $$
(5.125)

This result naturally follows from the fact that, for a = 0, the RL fractional integral becomes a temporal convolution of f(t) with a power law, t p/Γ(p). Thus, Eq. 5.125 follows from applying the convolution theorem (Eq. 5.118).

A simple relation also provides the Fourier transform (see Appendix 1 of Chap. 2) of the RL fractional integral if the starting point is a = −. It is given byFootnote 51:

$$\displaystyle \begin{aligned} {\mathrm{F}}\left[{}_{-\infty} D^{-p}_{t}\cdot f(t) \right] = (-i\omega)^{-p} \skew2\hat f(\omega). \end{aligned} $$
(5.127)

Riemann-Liouville Fractional Derivatives

The left-sided RL fractional derivative of order p > 0 of a function f(t) is defined as [21]:

$$\displaystyle \begin{aligned} {}_{a} D^{\,p}_{t} f(t) \equiv \frac{1}{\varGamma(k-p)} \frac{d^k}{dt^k}\int_a^t (t-\tau)^{k-p-1} f(\tau) d\tau, \end{aligned} $$
(5.128)

where the integer k satisfies that k − 1 ≤ p < k. Note that they are simply a combination of normal derivatives and RL fractional integrals: \(_{a} D^{\,p}_{t} = (d/dt)^k \cdot _a D^{-(k - p)}_{t} \). Again, it turns out that, for p = n, the RL fractional derivative reduces to the standard derivative of order n.Footnote 52

RL fractional derivatives have interesting, but somewhat not intuitive properties. The most striking property is probably that the fractional derivative of a constant function is not zero. Indeed, using the fact that the derivative of a power law can be calculated to be [21]:

$$\displaystyle \begin{aligned} _{a} D^{\,p}_{t} \cdot (t-a)^\nu = \frac{\varGamma(1+\nu)}{\varGamma(1+\nu -p)} (t-a)^{\nu - p},~~~ p>0,~\nu > -1,~t>0, \end{aligned} $$
(5.130)

it is clear that choosing ν = 0 does not yield a constant, but (ta)p/Γ(1 − p).

RL fractional derivatives can be combined with other derivatives (fractional or integer) and derivatives. But the combinations are not always simple. One of the simplest cases is when a fractional derivative of order p > 0 acts on a fractional integral of order q > 0:

$$\displaystyle \begin{aligned} _a D^{\,p}_{t} \left({}_{a} D^{-q}_{t} f(t)\right) = _a D^{\,p-q}_{t} f(t). \end{aligned} $$
(5.131)

However, the action of a RL fractional integral of order q > 0 on a RL fractional derivative of order p > 0 is given by a much more complicated expression [21].

If one sets p = q in Eq. 5.131, one finds that the inverse (from the left) of a RL fractional derivative of order p > 0 is the fractional integral of order p > 0:

$$\displaystyle \begin{aligned} _a D^{\,p}_{t} \left({}_a D^{-p}_{t} f(t)\right) = f(t). \end{aligned} $$
(5.132)

However, note that the RL fractional derivative of order p > 0 is not the inverse from the left of the RL fractional integral of order p. Instead, one has that [21]

$$\displaystyle \begin{aligned} _a D^{-p}_{t} \left({}_a D^{\,p}_{t} f(t)\right) = f(t) - \sum_{j=1}^k \left[-a D^{\,p-j}_{t} f(t)\right]_{t=a} \frac{(t-a)^{p-j}}{\varGamma(p-j+1)}. \end{aligned} $$
(5.133)

The action of normal derivatives on RL fractional derivatives is also simple to expressFootnote 53:

$$\displaystyle \begin{aligned} \frac{d^m }{dt^m} \cdot~ _a D^{\,p}_{t} f(t) = _a D^{\,p+m}_{t} f(t). \end{aligned} $$
(5.135)

But again, the action of the RL fractional derivative on a normal derivative is much more complicated, and given by the expression:

$$\displaystyle \begin{aligned} _a D^{\,p}_{t}\cdot \frac{d^m }{dt^m} f(t) =_a D^{\,p+m}_{t} f(t) - \sum_{j=0}^{m-1} \frac{f^{(j)}(a)(t-a)^{\,j-p-m}} {\varGamma(1+ j -p-m)}. \end{aligned} $$
(5.136)

Finally, the composition of RL fractional derivatives of different orders is given by a rather complex expression, that is never equal to a fractional derivative of a higher order, except in very special cases [21],

Relatively simple expressions also exist for the Laplace transform of the left-sided RL fractional derivative of order p if the starting point is a = 0:

$$\displaystyle \begin{aligned} L\left[{}_0 D^{\,p}_{ t}\cdot f(t) \right] = s^{p} \skew2\tilde f(s) - \sum_{j=0}^{k-1} s^{\,j} \left[{}_0 D^{\,p-j-1}_{t} \cdot f(t)\right]_{t=0}. \end{aligned} $$
(5.137)

This expression is very reminiscent of the one obtained for normal derivatives (Eq. 5.115). Similarly, the Fourier transform of the left-sided RL fractional derivative satisfies a very simple relation,Footnote 54 but only for starting point a = −:

$$\displaystyle \begin{aligned} {\mathrm{F}}\left[{}_{-\infty} D^{\,p}_{t} \cdot f(t) \right] = (i\omega)^{p} \skew2\hat f(\omega). \end{aligned} $$
(5.139)

Appendix 3: The Riesz-Feller Fractional Derivative

The Riesz fractional derivative of order α is defined by the integralFootnote 55:

$$\displaystyle \begin{aligned} \frac{\partial^\alpha f}{\partial |x|{}^\alpha}:= -\frac{1}{2\varGamma(\alpha)\cos{}(\alpha\pi/2)}\int_{-\infty}^\infty dx' \frac{f(x')}{|x-x'|{}^{\alpha + 1}} dx'. \end{aligned} $$
(5.140)

The most remarkable property of this derivative, and one of the reasons why it appears so often in the context of transport (see Sect. 5.3.1), has to do with its Fourier transform, which is given by [47]:

$$\displaystyle \begin{aligned} {\mathrm{F}}\left[\frac{\partial^\alpha f}{\partial |x|{}^\alpha}\right] = -|k|{}^\alpha \skew2\hat f(k). \end{aligned} $$
(5.141)

Using now the complex identity (\(\i = \sqrt {-1}\)),

$$\displaystyle \begin{aligned} (\i k)^\alpha + (-\i k)^\alpha = 2 \cos\left(\frac{\pi\alpha}{2}\right) |k|{}^\alpha, \end{aligned} $$
(5.142)

it is very easy to prove that the Riesz derivative can also be expressed as a symmetrized sum of two (one left-sided, one right-sided) RL fractional derivatives of order α [21],

$$\displaystyle \begin{aligned} \frac{\partial^\alpha f}{\partial |x|{}^\alpha} = -\frac{1}{2\varGamma(\alpha)\cos{}(\alpha\pi/2)} \left( \mbox{}_{-\infty}D^{\alpha}_{x} +\mbox{}^\infty D^{\alpha}_{x} \right) \end{aligned} $$
(5.143)

It is also possible to define an asymmetric version of the Riesz-Feller derivative, often known as the Riesz-Feller fractional derivative of order α and asymmetry parameter λ [48]. This can be done through its Fourier transform, that is given by:

$$\displaystyle \begin{aligned} {\mathrm{F}}\left[\frac{\partial^{\alpha,\lambda} f}{\partial |x|{}^{\alpha,\lambda}}\right] = -|k|{}^\alpha\left[1- \i \lambda {\mathrm{sgn}}(k) \tan{}(\pi\alpha/2) \right] \skew2\hat f(k). \end{aligned} $$
(5.144)

The indices are restricted, in this case, to α ∈ (0, 2) and λ ∈ [−1, 1]. For λ = 0, the standard symmetric Riesz derivative is recovered. It can also be shown that the Riesz-Feller derivative can also be expressed as an asymmetric sum of the same two RL fractional derivatives of order α [49]:

$$\displaystyle \begin{aligned} \frac{\partial^{\alpha,\lambda} f}{\partial |x|{}^{\alpha,\lambda}} = -\frac{1}{2\varGamma(\alpha)\cos{}(\alpha\pi/2)} \left( c_-(\alpha,\lambda) \mbox{}_{-\infty}D^{\alpha}_{x} + c_+(\alpha,\lambda) \mbox{}^\infty D^{\alpha}_{x} \right) \end{aligned} $$
(5.145)

with the coefficients defined as:

$$\displaystyle \begin{aligned} c_{\pm}(\alpha,\lambda) := \frac{1 \mp \lambda}{1 + \lambda\tan{}(\pi\alpha/2)} \end{aligned} $$
(5.146)

Thus, in the limit of λ = 1, only the left-sided α-RL derivative \( \mbox{ }_{-\infty } D^{\alpha }_{x}\) remains, whilst for λ = −1, only the right-side one \( \mbox{ }^\infty D^{\alpha }_{x}\) does.

Appendix 4: Discrete Approximations for Fractional Derivatives

In order to be useful for actual applications, discrete representations of fractional derivatives are needed that could be easily implemented in a computer [44, 50,51,52]. We discuss here one possible way to do this [51]. The main difficulty has to do with the singularities of the Riemann-Lioville definition, since it turns out that it becomes singular at the starting point, t = a, whenever a is finite. This will certainly be the case in practical cases, since a computer must always deal with finite intervals.

Case 1 < p < 2

One can made these singularities explicit by rewriting the fractional derivative \(\mbox{ }_aD_t^p f\), 1 < p < 2, as the infinite series [52]:

$$\displaystyle \begin{aligned} \frac{1}{\varGamma(1-p)}\frac{f(a)}{(t-a)^p} + \sum_{k=1}^\infty \frac{f^{(k)}(a) (t-a)^{k-p}}{\varGamma(k+1-p)} \end{aligned} $$
(5.147)

simply by Taylor expanding f(t) around t = a and using Eq. 5.130 to derive the different powers of x. Clearly, the first two terms of Eq. 5.147 are singular. At least, unless one sets f(a) = 0, and f′(a) = 0 as well for 1 < p < 2, which puts too much of a restriction in some cases.

One way to circumvent this problem is to replace in practice the Riemann-Liouville derivative by the so-called Caputo fractional derivative [21]:

$$\displaystyle \begin{aligned} \mbox{}^{\mathrm{C}}_aD_t^p f := \mbox{}_aD_t^p \left[\,f - \sum_{i=0}^{{\mathrm{int}}(p)} f^{(i)}(a)(t-a)^{i} \right], \end{aligned} $$
(5.148)

where int(p) is the integer part of p.Footnote 56 The Caputo definition removes the singularities so that the expression is now well defined and can be easily discretized. To do it, we give its analytical expression [21],

$$\displaystyle \begin{aligned} \mbox{}_a^{\mathrm{C}} D^{\,p}_t f(t) \equiv \frac{1}{\varGamma(k-p)} \int_a^t (t-\tau)^{k-p-1} f^{(k)}(\tau) d\tau, \end{aligned} $$
(5.150)

with k − 1 ≤ p < k, which is very similar to the RL definition (Eq. 5.128) but with the integer derivatives acting inside of the integral, instead of outside. It can be shown that RL and Caputo derivatives are identical, for most functions for a →− and t →. However, they are different for finite starting points. For starters, it should be noted that the Caputo derivative of a constant is now zero!

We proceed now to find discrete expressions for the Caputo derivative for 1 < p < 2 on the discrete regular mesh, t i  = (i − 1)Δt, with i = 0, 2, ⋯ , N t . The integral from t = t 1 up to t = t i is then discretizedFootnote 57 following the scheme [44, 51],

$$\displaystyle \begin{aligned} \begin{array}{rcl} \mbox{}^{\mathrm{C}}_aD_t^p f &\displaystyle = &\displaystyle \frac{1}{\varGamma(2-p)} \int_a^t (t-\tau)^{1-p} f^{(2)}(\tau) d\tau \\ &\displaystyle \simeq &\displaystyle ~\frac{1}{\varGamma(2-p)} \sum_{j=0}^{i-1}\int_{t_j}^{t_{j+1}}\frac{f''(t-s)}{s^{p-1}}ds \\ &\displaystyle \simeq &\displaystyle \frac{1}{\varGamma(2-p)}\sum_{j=0}^{i-1}\left[ \frac{f(t_i - t_{j+1}) + f(t_i-t_{j-1}) - 2f(t_i-t_j)}{(\varDelta t)^2} \int_{t_j}^{t_{j+1}}\frac{ds}{s^{p-1}} \right] \\ &\displaystyle \simeq &\displaystyle \sum_{j=0}^{i-1 }\left[ \frac{(f(t_i - t_{j+1}) + f(t_i-t_{j-1}) - 2f(t_i-t_j))\left[ (j+1)^{2-p} - j^{2-p}\right]}{\varGamma(3-p)(\varDelta t)^p} \right]. \end{array} \end{aligned} $$
(5.151)

This expression is more conveniently expressed as,

$$\displaystyle \begin{aligned} \mbox{}^{\mathrm{C}}_aD_t^p f = \sum_{j=-1}^i \frac{W_j^{p>1}}{\varGamma(3-p)(\varDelta t)^p}f(t_i - t_j), \end{aligned} $$
(5.152)

in terms of the weights, \(W_j^{p>1}\),

$$\displaystyle \begin{aligned} W_j^{p>1} = \left\lbrace \begin{array}{ll} \displaystyle 1, & j = -1\\ \displaystyle 2^{2-p}-3, & j = 0\\ \displaystyle (j+2)^{2-p} - 3(j+1)^{2-p} + 3j^{2-p} - (j-1)^{2-p},~~~ & 0 < j < i - i\\ \displaystyle -2i^{2-p} + 3(i-1)^{2-p}- (i-2)^{2-p}, & j = i-1 \\ \displaystyle i^{2-p} - (i-1)^{2-p}, & j = i \end{array} \right. \end{aligned} $$
(5.153)

Case 0 < α < 1

When p < 1 there is only one singularity at the starting point a of the fractional derivative. We can make it explicit by expressing again the fractional derivative \(\mbox{ }_aD_t^p f\), as the infinite series:

$$\displaystyle \begin{aligned} \frac{1}{\varGamma(1-p)}\frac{f(a)}{(t-a)^p} + \sum_{k=1}^\infty \frac{f^{(k)}(a) (t-a)^{k-p}}{\varGamma(k+1-p)}, \end{aligned} $$
(5.154)

done by Taylor expanding f(t) around t = a and using Eq. 5.130 to derive the different powers of x. Clearly, the first term of Eq. 5.154 is singular. At least, unless one sets f(a) = 0.

The solution is again to use the Caputo derivative (Eq. 5.150) instead. Now, int(p) = 0 and the integral to discretize is:

$$\displaystyle \begin{aligned} \mbox{}_a^{\mathrm{C}} D^{\,p}_t f(t) \equiv \frac{1}{\varGamma(1-p)} \int_a^t (t-\tau)^{-p} f^{(k)}(\tau) d\tau, \end{aligned} $$
(5.155)

For 0 < p < 1, on the other hand, we discretize the integral from t = t 0 up to t = t i as,

$$\displaystyle \begin{aligned} \begin{array}{rcl} \mbox{}^{\mathrm{C}}_aD_t^p f &\displaystyle = &\displaystyle \frac{1}{\varGamma(1-p)} \int_a^t (t-\tau)^{-p} f^{(1)}(\tau) d\tau \\ &\displaystyle \simeq &\displaystyle ~\frac{1}{\varGamma(1-p)} \sum_{j=0}^{i-1}\int_{t_j}^{t_{j+1}}\frac{f'(t-s)}{s^{p}}ds \\ &\displaystyle \simeq &\displaystyle \frac{1}{\varGamma(1-p)}\sum_{j=0}^{i-1}\left[ \frac{f(t_i - t_{j+1}) - f(t_i-t_{j-1})}{2\varDelta t} \int_{t_j}^{t_{j+1}}\frac{ds}{s^{p}} \right] \\ &\displaystyle \simeq &\displaystyle \sum_{j=0}^{i-1 }\left[ \frac{(f(t_i - t_{j+1}) - f(t_i-t_{j-1}))\left[ (j+1)^{1-p} - j^{1-p}\right]}{2\varGamma(2-p)(\varDelta t)^p} \right]. \end{array} \end{aligned} $$
(5.156)

Again, we can express this formula more conveniently by introducing weights, \(W_j^{p<1}\),

$$\displaystyle \begin{aligned} \mbox{}^{\mathrm{C}}_aD_t^p f = \sum_{j=-1}^i \frac{W_j^{p<1}}{2\varGamma(2-p)(\varDelta t)^p}f(t_i - t_j), \end{aligned} $$
(5.157)

which are now given by the expressions,

$$\displaystyle \begin{aligned} W_j^{p<1} = \left\lbrace \begin{array}{ll} \displaystyle -1, & j = -1\\ \displaystyle -(2^{1-p}-1), & j = 0\\ \displaystyle j^{1-p} -(j-1)^{1-p} - (j+2)^{1-p} +(j+1)^{1-p},~~~ & 0 < j < i - i\\ \displaystyle (i-1)^{1-p}- (i-2)^{1-p}, & j = i-1 \\ \displaystyle i^{1-p} - (i-1)^{1-p}, & j = i \end{array} \right. \end{aligned} $$
(5.158)

Equations 5.152 and 5.157 are not however, the only possible discrete representation of the Caputo fractional derivative of order p. These formulas are based on central differencing. But other discretizations also exist, that may sometimes offer either better accuracy over some ranges of p or higher orders of discretization [44, 50,51,52]. Equations 5.152 and 5.157 are however sufficient for the purposes of this book.

Problems

5.1

CTRW: Fluid Limit

Write a code that evolves in time a one-dimensional CTRW that has a Gaussian jump size distribution, \(p(\varDelta x) = N_{[0,\sigma ^2]}(\varDelta x)\), and an exponential waiting time pdf, \(\psi (\varDelta t) = E_{\tau _0}(\varDelta t)\). In order to generate values for the jumps and waiting times at run-time, use the algorithms described in Appendix 1 of Chap. 2. Use the code to calculate numerically the CTRW propagator for σ 2 = 2 and τ 0 = 1, using N = 100, 000 particles. Compare the results with its fluid limit: G(x         −     x 0, t)                         =                         N [0,2t] (x − x 0).

5.2

Langevin Equation: Propagator

Write a code that evolves in time a collection of N particles according to the Langevin equation (Eq. 5.26). Use a uniform noise with autocorrelation given by Eq. 5.27 with D = 2. Use N = 100, 000 particles to calculate the propagator of the Langevin equation and compare it with its analytical solution: G(x − x 0, t) = N [0,2t](x − x 0).

5.3

Langevin Equation: Moments of the Propagator

Compute all moments of the propagator of the Langevin equation and show that all odd moments vanish and all even moments n > 2 are given by m n  = (Dt)n(n − 1)!!.

5.4

Fractional Transport Equation: Propagator for α < 2, β = 1

Calculate the propagator of the fractional transport equation (Eq. 5.66) in the case in which β = 1.

5.5

Scale-Invariant Generalization of Fick’s Law

Find the expression that gives the local particle flux for the fractional transport equation. To do it, recast Eq. 5.66 into the form ∂n/∂t + ∇⋅ Γ = S, using the properties of fractional derivatives discussed in Appendices 2 and 3.

5.6

Advanced Problem: Propagator of the Running Sandpile

Write a code that uses the rules discussed in Sect. 5.5 to advances an arbitrary number of tracers on the height profile evolution calculated by the sandpile code previously developed (see Prob 1.5). Use the code to estimate the numerical propagator of the running sandpile for a sandpile with L = 2000, Z c  = 200, N f  = 20, N b  = 10 and p 0 = 10−4.

5.7

Advanced Problem: Numerical Integration of the fTe

Write a code that integrates the fractional transport equation (Eq. 5.66 ) for β = 1 and arbitrary α, arbitrary initial condition, n 0(x), and external source, S(x, t). Use, for that purpose, the discrete expressions of the Caputo fractional derivative (Eqs. 5.152 and 5.157) given in Appendix 4.

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer Science+Business Media B.V.

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Sánchez, R., Newman, D. (2018). Fundamentals of Fractional Transport. In: A Primer on Complex Systems. Lecture Notes in Physics, vol 943. Springer, Dordrecht. https://doi.org/10.1007/978-94-024-1229-1_5

Download citation

Publish with us

Policies and ethics