[22.1.1] The brief historical introduction has shown that fractional derivatives may be defined in numerous ways. [22.1.2] A natural and frequently used approach starts from repeated integration and extends it to fractional integrals. [22.1.3] Fractional derivatives are then defined either by continuation of fractional integrals to negative order (following Leibniz’ ideas [73]), or by integer order derivatives of fractional integrals (as suggested by Riemann [96]).
[22.2.1] Consider a locally integrable1 (This is a footnote:) 1 A function is called locally integrable if it is integrable on all compact subsets (see eq.(B.9)). real valued function whose domain of definition is an interval with . [22.2.2] Integrating
[page 23, §0] times gives the fundamental formula
(2.27) |
where and . [23.0.1] This formula may be proved by induction. [23.0.2] It reduces -fold integration to a single convolution integral (Faltung). [23.0.3] The subscript indicates that the integration has as its lower limit. [23.0.4] An analogous formula holds with lower limit and upper limit . [23.0.5] In that case the subscript will be used.
[23.1.1] Equation (2.27) for -fold integration can be generalized to noninteger values of using the relation where
(2.28) |
is Euler’s -function defined for all .
[23.2.1] Let . [23.2.2] The Riemann-Liouville fractional integral of order with lower limit is defined for locally integrable functions as
(2.29a) | ||
for . [23.2.3] The Riemann-Liouville fractional integral of order with upper limit is defined as | ||
(2.29b) |
for . [23.2.4] For
(2.30) |
completes the definition. [23.2.5] The definition may be generalized to with .
[page 24, §1] [24.1.1] Formula (2.29a) appears in [96, p.363] with and in [76, p.8] with . [24.1.2] The notation is not standardized. [24.1.3] Leibniz, Lagrange and Liouville used the symbol [73, 22, 76], Grünwald wrote , while Riemann used [96] and Most wrote [89]. [24.1.4] The notation in (2.29) is that of [99, 98, 54, 52]. [24.1.5] Modern authors also use [37], [97], [94], [23], [85, 102, 91], or [92] instead of 2 (This is a footnote:) 2 Some authors [97, 26, 92, 23, 85, 91] employ the derivative symbol also for integrals, resp. for derivatives, to emphasize the similarity between fractional integration and differentiation. If this is done, the choice of Riesz and Feller, namely , seems superior in the sense that fractional derivatives, similar to integrals, are nonlocal operators, while integer derivatives are local operators..
[24.2.1] The fractional integral operators are commonly called Riemann-Liouville fractional integrals [99, 98, 94] although sometimes this name is reserved for the case [85]. [24.2.2] Their domain of definition is typically chosen as or [99, 98, 94]. [24.2.3] For the definition of Lebesgue spaces see the Appendix B. [24.2.4] If then and is finite for almost all . [24.2.5] If with and then is finite for all . [24.2.6] Analogous statements hold for [98].
[24.4.1] Examples (2.5) and (2.6) or (A.3) and (A.5) show that Definition 2.1 is well suited for fractional integration of power series, but not for functions defined by Fourier series. [24.4.2] In fact, if is a periodic function with period , and3 (This is a footnote:) 3The notation indicates that the sum does not need to converge, and, if it converges, does not need to converge to .
(2.31) |
then the Riemann-Liouville fractional will in general not be periodic. [24.4.3] For this reason an alternative definition of fractional integrals was investigated by Weyl [124].
[24.5.1] Functions on the unit circle correspond to -periodic functions on the real line. [24.5.2] Let be periodic with period and such that the integral of over the interval vanishes, so that in eq. (2.31). [24.5.3] Then the integral of is itself a periodic function, and the constant of integration can be chosen such that the integral over vanishes again. [24.5.4] Repeating the integration times one finds using (2.6) and the integral representation
[page 25, §0] of Fourier coefficients
(2.32) |
with . [25.0.1] Recall the convolution formula [132, p.36]
(2.33) |
for two periodic functions and . [25.0.2] Using eq. (2.33) and generalizing (2.32) to noninteger suggests the following definition. [99, 94].
[25.1.1] Let be periodic with period and such that its integral over a period vanishes. [25.1.2] The Weyl fractional integral of order is defined as
(2.34) |
where
(2.35) |
for .
[25.2.1] It can be shown that the series for converges and that the Weyl definition coincides with the Riemann-Liouville definition [133]
(2.36a) | ||
respectively | ||
(2.36b) |
for periodic functions whose integral over a period vanishes. [25.2.2] This is eq. (2.29) with resp. . [25.2.3] For this reason the Riemann-Liouville
[page 26, §0] fractional integrals with limits , and , are often called Weyl fractional integrals [24, 99, 85, 94].
[26.1.1] The Weyl fractional integral may be rewritten as a convolution
(2.37) |
where the convolution product for functions on is defined as4 (This is a footnote:) 4 If then exists for almost all and . [26.1.2] If , with and then , the space of continuous functions vanishing at infinity.
(2.38) |
and the convolution kernels are defined as
(2.39) |
for . [26.1.3] Here
(2.40) |
is the Heaviside unit step function, and with the convention that is real for . [26.1.4] For the kernel
(2.41) |
is the Dirac -function defined in (C.2) in Appendix C. [26.1.5] Note that for .
[26.2.1] Riemann-Liouville and Weyl fractional integrals have upper or lower limits of integration, and are sometimes called left-sided resp. right-sided integrals. [26.2.2] A more symmetric definition was advanced in [97].
[26.3.1] Let be locally integrable. [26.3.2] The Riesz fractional integral or Riesz potential of order is defined as the linear combination [99]
(2.42) |
[27.1.1] Riesz fractional integration may be written as a convolution
(2.45a) | |||
(2.45b) |
with the (one-dimensional) Riesz kernels
(2.46) |
for , and
(2.47) |
for . [27.1.2] Subsequently, Feller introduced the generalized Riesz-Feller kernels [26]
(2.48) |
with parameter . [27.1.3] The corresponding generalized Riesz-Feller fractional integral of order and type is defined as
(2.49) |
[27.1.4] This formula interpolates continuously from the Weyl integral for through the Riesz integral for to the Weyl integral for . [27.1.5] Due to their symmetry Riesz-Feller fractional integrals are readily generalized to higher dimensions.
[page 28, §1]
[28.1.1] Fractional integration can be extended to distributions using the convolution formula (2.37) above. [28.1.2] Distributions are generalized functions [105, 31]. [28.1.3] They are defined as linear functionals on a space of conveniently chosen ‘‘test functions’’. [28.1.4] For every locally integrable function there exists a distribution defined by
(2.50) |
where is test function from a suitable space of test functions. [28.1.5] By abuse of notation one often writes for the associated distribution . [28.1.6] Distributions that correspond to functions via (2.50) are called regular distributions. [28.1.7] Examples for regular distributions are the convolution kernels defined in (2.39). [28.1.8] They are locally integrable functions on when . [28.1.9] Distributions that are not regular are sometimes called singular. [28.1.10] An important example for a singular distribution is the Dirac -function. [28.1.11] It is defined as
(2.51) |
for every test function .
[28.1.12] The test function space is usually chosen as a
subspace of , the space of
infinitely differentiable functions.
[28.1.13] A brief introduction to distributions is given in
Appendix C.
[28.2.1] In order to generalize (2.37) to distributions one must define the convolution of two distributions. [28.2.2] To do so one multiplies eq. (2.38) on both sides with a smooth test function of compact support. [28.2.3] Integrating gives
(2.52) |
where the notation means that the functional is applied to the function for fixed . [28.2.4] Explicitly, for fixed
(2.53) |
[page 29, §0] where . [29.0.1] Equation (2.52) can be used as a definition for the convolution of distributions provided that the right hand side has meaning. [29.0.2] This is not always the case as the counterexample shows. [29.0.3] In general the convolution product is not associative (see eq. (2.113)). [29.0.4] However, associative and commutative convolution algebras exist [21]. [29.0.5] Equation (2.52) is always meaningful when or is compact [63]. [29.0.6] Another case is when and have support in . [29.0.7] This will be assumed in the following.
[29.1.1] Let be a distribution with . [29.1.2] Then its fractional integral is the distribution defined as
(2.54) |
for . [29.1.3] It has support in .
[29.2.1] If with then also with .
[29.3.1] The Fourier transformation is defined as
(2.55) |
for functions . [29.3.2] Then
(2.56) |
holds for by virtue of the convolution theorem. [29.3.3] The equation cannot be extended directly to because the Fourier integral on the left hand side may not exist. [29.3.4] Consider e.g. and . [29.3.5] Then const as and does not exist [94]. [29.3.6] Equation (2.56) can be extended to all with for functions in the so called Lizorkin space [99, p.148] defined as the space of functions such that for all .
[29.4.1] For the Riesz potentials one has
(2.57a) | |||
(2.57b) |
for functions in Lizorkin space.
[page 30, §1] [30.1.1] The Laplace transform is defined as
(2.58) |
for locally integrable functions . [30.1.2] Now
(2.59) |
by the convolution theorem for Laplace transforms. [30.1.3] The Laplace transform of leads to a more complicated operator.
[30.2.1] If with , and , for then the formula
(2.60) |
holds. [30.2.2] The formula is known as fractional integration by parts [99]. [30.2.3] For with and the analogous formula
(2.61) |
holds for Weyl fractional integrals.
[30.3.1] These formulae provide a second method of generalizing fractional integration to distributions. [30.3.2] Equation (2.60) may be read as
(2.62) |
for a distribution and a test function . [30.3.3] It shows that right- and left-sided fractional integrals are adjoint operators. [30.3.4] The formula may be viewed as a definition of the fractional integral of a distribution provided that the operator maps the test function space into itself.
[30.4.1] The mapping properties of convolutions can be studied with the help of Youngs inequality. Let obey and . [30.4.2] If and then and Youngs inequality holds. [30.4.3] It follows that if
[page 31, §0] and with . [31.0.1] The Hardy-Littlewood theorem states that these estimates remain valid for although these kernels do not belong to any -space [37, 38]. [31.0.2] The theorem was generalized to higher dimensions by Sobolev in 1938, and is also known as the Hardy-Littlewood-Sobolev inequality (see [37, 38, 113, 63]).
[31.1.1] Let , , . [31.1.2] Then are bounded linear operators from to with ,i.e. there exists a constant independent of such that .
[31.2.1] The basic composition law for fractional integrals follows from
(2.63) |
where Euler’s Beta-function
(2.64) |
was used. [31.2.2] This implies the semigroup law for exponents
(2.65) |
also called additivity law. [31.2.3] It holds for Riemann-Liouville, Weyl and Riesz-Feller fractional integrals of functions.
[31.3.1] Riemann [96, p.341] suggested to define fractional derivatives as integer order derivatives of fractional integrals.
[31.4.1] Let . [31.4.2] The Riemann-Liouville fractional derivative of order with lower limit (resp. upper limit ) is defined for
[page 32, §0] functions such that and as
(2.66) |
and for . [32.0.1] For the definition is extended for functions with as
(2.67) |
where5 (This is a footnote:) 5 is the largest integer smaller than . is smallest integer larger than .
[32.1.1] Here denotes a Sobolev space defined in (B.17). [32.1.2] For the space coincides with the space of absolutely continuous functions.
[32.2.1] The notation for fractional derivatives is not standardized6 (This is a footnote:) 6see footnote 2.1. [32.2.2] Leibniz and Euler used [73, 72, 25] Riemann wrote [96], Liouville preferred [76], Grünwald used or [34], Marchaud wrote , and Hardy-Littlewood used an index [37]. [32.2.3] The notation in (2.67) follows [99, 98, 54, 52]. Modern authors also use [97], [23], [94, 85, 102], [129, 102], [92] instead of .
[32.3.1] Let be absolutely continuous on the finite interval . [32.3.2] Then, its derivative exists almost everywhere on with , and the function can be written as
(2.68) |
Substituting this into gives
(2.69) |
where commutativity of and was used. [32.3.3] It follows that
(2.70) |
for . [32.3.4] Above, the notations
(2.71) |
were used for the first order derivative.
[page 33, §1] [33.1.1] This observation suggests to introduce a modified Riemann-Liouville fractional derivative through
(2.72) |
where . [33.1.2] Note, that must be at least -times differentiable. [33.1.3] Formula (2.72) is due to Liouville [76, p.10] (see eq. (2.18) above), but nowadays sometimes named after Caputo [17].
[33.2.2] For with the Riemann-Liouville fractional derivative exists almost everywhere for . [33.2.3] It can be written as
(2.73) |
in terms of the Liouville(-Caputo) derivative defined in (2.72).
[33.3.1] The Riemann-Liouville fractional derivative is the left inverse of Riemann-Liouville fractional integrals. [33.3.2] More specifically, [99, p.44]
[33.3.3] Let . [33.3.4] Then
(2.74) |
holds for all with .
[33.4.1] For the right inverses of fractional integrals one finds
[33.4.2] Let and . [33.4.3] If in addition where then
(2.75) |
holds. [33.4.4] For this becomes
(2.76) |
[33.5.1] The last theorem implies that for and with the equality
(2.77) |
[page 34, §0] holds only if
(2.78a) | ||
and | ||
(2.78b) |
for all . [34.0.1] Note that the existence of in eq. (2.77) does not imply that can be written as for some integrable function [99]. [34.0.2] This holds only if both conditions (2.78) are satisfied. [34.0.3] As an example where one of them fails, consider the function for . [34.0.4] Then exists. [34.0.5] Now so that (2.78b) fails. [34.0.6] There does not exist an integrable such that . [34.0.7] In fact, corresponds to the -distribution .
[34.1.1] Riemann-Liouville fractional derivatives have been generalized in [52, p.433] to fractional derivatives of different types.
[34.2.1] The generalized Riemann-Liouville fractional derivative of order and type with lower (resp. upper) limit is defined as
(2.79) |
for functions such that the expression on the right hand side exists.
[34.3.1] The type of a fractional derivative allows to interpolate continuously from to . [34.3.2] A relation between fractional derivatives of the same order but different types was given in [52, p.434].
[34.4.1] Marchaud’s approach [78] is based on Hadamards finite parts of divergent integrals [36]. [34.4.2] The strategy is to define fractional derivatives as analytic continuation of fractional integrals to negative orders. [see [99, p.225]]
[34.5.1] Let and . [34.5.2] The Marchaud fractional derivative of order with lower limit is defined as
(2.80) |
[page 35, §0] and the Marchaud fractional derivative of order with upper limit is defined as
(2.81) |
[35.0.1] For (resp. ) the definition is
(2.82) |
[35.0.2] The definition is completed with for all variants.
[35.1.1] The idea of Marchaud’s method is to extend the Riemann-Liouville integral from to , and to define
(2.83) |
where . [35.1.2] However, this is not possible because the integral in (2.83) diverges. [35.1.3] The idea is to subtract the divergent part of the integral,
(2.84) |
obtained by setting for . [35.1.4] Subtracting (2.83) from (2.84) for suggests the definition
(2.85) |
[35.1.5] Formal integration by parts leads to , showing that this definition contains the Riemann-Liouville definition.
[35.2.1] The definition may be extended to in two ways. [35.2.2] The first consists in applying (2.85) to the -th derivative for . [35.2.3] The second possibility is to regard as a first order difference, and to generalize to -th order differences. [35.2.4] The -th order difference is
(2.86) |
where is the identity operator and
(2.87) |
[page 36, §0] is the translation operator. [36.0.1] The Marchaud fractional derivative can then be extended to through [94, 98]
(2.88) |
where
(2.89) |
where the limit may be taken in the sense of pointwise or norm convergence.
[36.1.1] The Marchaud derivatives are defined for a wider class of functions than Weyl derivatives . [36.1.2] As an example consider the function const.
[36.2.1] Let be such that there exists a function with . [36.2.2] Then the Riemann-Liouville derivative and the Marchaud derivative coincide almost everywhere, i.e. for almost all [99, p.228].
[36.3.1] There are two kinds of Weyl fractional derivatives for periodic functions. [36.3.2] The Weyl-Liouville fractional derivative is defined as [99, p.351],[94]
(2.90) |
for where the Weyl integral was defined in (2.34). [36.3.3] The Weyl-Marchaud fractional derivative is defined as [99, p.352],[94]
(2.91) |
for where is defined in eq. (2.35). [36.3.4] The Weyl derivatives are defined for periodic functions of with zero mean in where . [36.3.5] In this space , i.e. the Weyl-Liouville and Weyl-Marchaud form coincide [99]. [36.3.6] As for fractional integrals, it can be shown that the Weyl-Liouville derivative
(2.92) |
coincides with the Riemann-Liouville derivative with lower limit . [36.3.7] In addition one has the equivalence with the Marchaud-Hadamard fractional derivative in a suitable sense [99, p.357].
[page 37, §1]
[37.1.1] To define the Riesz fractional derivative as integer derivatives of Riesz potentials consider the Fourier transforms
(2.93) |
(2.94) |
for . [37.1.2] Comparing this to eq. (2.57) suggests to consider
(2.95) |
as a candidate for the Riesz fractional derivative.
[37.2.1] Following [94] the strong Riesz fractional derivative of order of a function , , is defined through the limit
(2.96) |
whenever it exists. [37.2.2] The convolution kernel defined as
(2.97) |
is obtained from eq. (2.95). [37.2.3] Indeed, this definition is equivalent to eq. (2.94). [37.2.4] A function where has a strong Riesz derivative of order if and only if there exsists a function such that . [37.2.5] Then .
[37.3.1] The basic idea of the Grünwald approach is to generalize finite difference quotients to noninteger order, and then take the limit to obtain a differential quotient. [37.3.2] The first order derivative is the limit
(2.98) |
of a difference quotient. [37.3.3] In the last equality is the identity operator, and
(2.99) |
is the translation operator. [37.3.4] Repeated application of gives
(2.100) |
[page 38, §0] where . [38.0.1] The second order derivative can then be written as
(2.101) |
and the -th derivative
(2.102) |
which exhibits the similarity with the binomial formula. [38.0.2] The generalization to noninteger gives rise to fractional difference quotients defined through
(2.103) |
for . [38.0.3] These are generally divergent for . [38.0.4] For example, if , then
(2.104) |
diverges as if . [38.0.5] Fractional difference quotients were studied in [68]. Note that fractional differences obey [99]
(2.105) |
[38.1.1] The Grünwald-Letnikov fractional derivative of order is defined as the limit
(2.106) |
of fractional difference quotients whenever the limit exists. [38.1.2] The Grünwald Letnikov fractional derivative is called pointwise or strong depending on whether the limit is taken pointwise or in the norm of a suitable Banach space.
[38.2.1] For a definition of Banach spaces and their norms see e.g. [128].
[38.3.1] The Grünwald-Letnikov fractional derivative has been studied for periodic functions in with in [99, 94]. [38.3.2] It has the following properties.
[page 39, §1]
[39.1.1] Let , and . [39.1.2] Then the following statements are equivalent:
[39.1.3] There exists a function such that
where .
[39.1.4] There exists a function such that
holds for almost all .
[39.2.1] Let , and . [39.2.2] Then:
implies for every .
[39.3.1] The basic idea for defining fractional differentiation of distributions is to extend the definition of fractional integration (2.54) to negative . [39.3.2] However, for the distribution becomes singular because is not locally integrable in this case. [39.3.3] The extension of to requires regularization [31, 128, 63]. [39.3.4] It turns out that the regularization exists and is essentially unique as long as .
[39.4.1] Let be a distribution with . [39.4.2] Then the fractional derivative of order with lower limit is the distribution defined as
(2.107) |
where and
(2.108) |
is the kernel distribution. [39.4.3] For one finds and as the identity operator. [39.4.4] For the one finds
(2.109) |
where is the -th derivative of the distribution.
[page 40, §1] [40.1.1] The kernel distribution in (2.108) is
(2.110) |
for . [40.1.2] Its regularized action is
(2.111a) | |||
(2.111b) | |||
(2.111c) | |||
(2.111d) |
where was assumed in the last step and the arbitrary constant was chosen as . [40.1.3] This choice regularizes the divergent first term in (2.111c). [40.1.4] If this rule is used for the distributional convolution
(2.112) |
then the Marchaud-Hadamard form is recovered with .
[40.2.1] It is now possible to show that the convolution of distributions is in general not associative. [40.2.2] A counterexample is
(2.113) |
where is the Heaviside step function.
[40.3.1] has support in . [40.3.2] The distributions in with form a convolution algebra [21] and one finds [31, 99]
[40.3.3] If with then also with . [40.3.4] Moreover, for all
(2.114) |
with for . [40.3.5] For each with there exists a unique distribution with such that .
[page 41, §1] [41.1.1] Note that
(2.115) |
for all . [41.2.1] Also, the differentiation rule
(2.116) |
holds for all . [41.2.2] It contains
(2.117) |
for all as a special case.
[41.3.1] All fractional derivatives defined above are nonlocal operators. [41.3.2] A local fractional derivative operator was introduced in [40, 41, 52].
[41.4.1] For the Riemann-Liouville fractional derivative of order at the lower limit is defined by
(2.118) |
whenever the two limits exist and are equal. [41.4.2] If exists the function is called fractionally differentiable at the limit .
[41.5.1] These operators are useful for the analysis of singularities. [41.5.2] They were applied in [40, 41, 42, 44, 52] to the analysis of singularities in the theory of critical phenomena and to the generalization of Ehrenfests classification of phase transitions. [41.5.3] There is a close relationship to the theory of regularly varying functions [107] as evidenced by the following result [52].
[41.6.1] Let the function be monotonously increasing with and , and such that with and is also monotonously increasing on a neighbourhood for small . [41.6.2] Let , let be a constant and a slowly varying function for . [41.6.3] Then
(2.119) |
holds if and only if
(2.120) |
holds.
[page 42, §1]
[42.1.1] A function is called slowly varying at infinity if for all . [42.1.2] A function is called slowly varying at if is slowly varying at infinity.
[42.2.1] The spectral decomposition of selfadjoint operators is a familiar mathematical tool from quantum mechanics [116]. [42.2.2] Let denote a selfadjoint operator with domain and spectral family on a Hilbert space with scalar product . [42.2.3] Then
(2.121) |
holds for all . [42.2.4] Here is the spectrum of . [42.2.5] It is then straightforward to define the fractional power by
(2.122) |
on the domain
(2.123) |
[42.2.6] Similarly, for any measurable function the operator is defined with an integrand in eq. (2.122). [42.2.7] This yields an operator calculus that allows to perform calculations with functions instead of operators.
[42.3.1] Fractional powers of the Laplacian as the generator of the diffusion semigroup were introduced by Bochner [13] and Feller [26] based on Riesz’ fractional potentials. [42.3.2] The fractional diffusion equation
(2.124) |
was related by Feller to the Levy stable laws [74] using one dimensional fractional integrals of order and type [26]7 (This is a footnote:) 7Fellers motivation to introduce the type was this relation.. [42.3.3] For eq. (2.124) reduces to the diffusion equation. [42.3.4] This type of fractional diffusion will be referred to as fractional diffusion of Bochner-Levy type (see Section 2.3.4 for more discussion). [42.3.5] Later, these ideas were extended to fractional powers of closed8 (This is a footnote:) 8 An operator on a Banach space is called closed if the set of pairs with is closed in . semigroup generators [4, 5, 69, 70]. [42.3.6] If is the infinitesimal generator of a
[page 43, §0] semigroup (see Section 2.3.3.2 for definitions of and ) on a Banach space then its fractional power is defined as
(2.125) |
for every for which the limit exists in the norm of [120, 121, 93, 123]. [43.0.1] This aproach is clearly inspired by the Marchaud form (2.82). Alternatively, one may use the Grünwald approach to define fractional powers of semigroup generators [122, 99].
[43.1.1] The calculus of pseudodifferential operators represents another generalization of the operator calculus in Hilbert spaces. [43.1.2] It has its roots in Hadamard’s ideas [36], Riesz potentials [97], Feller’s suggestion [26] and Calderon-Zygmund singular integrals [16]. [43.1.3] Later it was generalized and became a tool for treating elliptic partial differential operators with nonconstant coefficients.
[43.2.1] A (Kohn-Nirenberg) pseudodifferential operator of order is defined as
(2.126) |
and the function is called its symbol. [43.2.2] The symbol is in the Kohn-Nirenberg symbol class if it is in , and there exists a compact set such that , and for any pair of multiindices there is a constant such that
(2.127) |
[43.2.3] The Hörmander symbol class is obtained by replacing the exponent on the right hand side with where .
[43.3.1] Pseudodifferential operators provide a unified approach to differential and integral or convolution operators that are ‘‘nearly’’ translation invariant. [43.3.2] They have a close relation with Weyl quantization in physics [116, 28]. However, they will not be discussed further because the traditional symbol classes do not contain the usual fractional derivative operators. [43.3.3] Fractional Riesz derivatives are not pseudodifferential operators in the sense above. [43.3.4] Their symbols do not fall into any of the standard Kohn-Nirenberg or Hörmander symbol classes due to lack of differentiability at the origin.
[page 44, §1]
[44.1.1] The eigenfunctions of Riemann-Liouville fractional derivatives are defined as the solutions of the fractional differential equation
(2.128) |
where is the eigenvalue. [44.1.2] They are readily identifed using eq. (A.11) as
(2.129) |
where
(2.130) |
is the generalized Mittag-Leffler function [125, 126]. [44.1.3] More generally the eigenvalue equation for fractional derivatives of order and type reads
(2.131) |
and it is solved by [54, eq.124]
(2.132) |
[page 45, §0] where the case corresponds to (2.128). [45.0.1] A second important special case is the equation
(2.133) |
with . [45.0.2] In this case the eigenfunction
(2.134) |
where is the Mittag-Leffler function [86]. [45.0.3] The Mittag-Leffler function plays a central role in fractional calculus. [45.0.4] It has only recently been calculated numerically in the full complex plane [108, 62]. [45.0.5] Figure 2.1 and 2.2 illustrate for a rectangular region in the complex plane (see [108]). [45.1.1] The solid line in Figure 2.1 is the line , in Figure 2.2 it is .
[45.2.1] Note, that some authors are avoiding the operator in fractional differential equations (see e.g. [112, 101, 84, 111, 7, 82] or chapters in this volume). [45.2.2] In their notation the eigenvalue equation (2.133) becomes (c.f.[112, eq.(22)])
(2.135) |
containing two derivative operators instead of one.