Time-Fractional Differential Equations A Theoretical Introduction (Adam Kubica, Katarzyna Ryszewska Etc.)
Time-Fractional Differential Equations A Theoretical Introduction (Adam Kubica, Katarzyna Ryszewska Etc.)
Adam Kubica
Katarzyna Ryszewska
Masahiro Yamamoto
Time-Fractional
Differential Equations
A Theoretical Introduction
123
SpringerBriefs in Mathematics
Series Editors
Nicola Bellomo, Torino, Italy
Michele Benzi, Pisa, Italy
Palle Jorgensen, Iowa, USA
Tatsien Li, Shanghai, China
Roderick Melnik, Waterloo, Canada
Otmar Scherzer, Linz, Austria
Benjamin Steinberg, New York, USA
Lothar Reichel, Kent, USA
Yuri Tschinkel, New York, USA
George Yin, Detroit, USA
Ping Zhang, Kalamazoo, USA
SpringerBriefs in Mathematics showcases expositions in all areas of mathematics
and applied mathematics. Manuscripts presenting new results or a single new result
in a classical field, new field, or an emerging topic, applications, or bridges between
new results and already published works, are encouraged. The series is intended for
mathematicians and applied mathematicians. All works are peer-reviewed to meet
the highest standards of scientific literature.
Titles from this series are indexed by Scopus, Web of Science, Mathematical
Reviews, and zbMATH.
Time-Fractional Differential
Equations
A Theoretical Introduction
Adam Kubica Katarzyna Ryszewska
Warsaw University of Technology Warsaw University of Technology
Warszawa, Poland Warszawa, Poland
Masahiro Yamamoto
Graduate School of Mathematical Sciences
The University of Tokyo
Tokyo, Japan
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Preface
Recently, fractional differential equations have attracted great attention and many
studies have been performed. However, there are not many works that cover the
theory of partial differential equations, so that unnecessarily lengthy arguments and
also arguments that lack rigor are sometimes presented, and it may be difficult to
gain unified views of the related fields.
This concise book provides rigorous treatments for time-fractional derivatives
in Sobolev spaces and solutions to initial boundary value problems for time-
fractional partial differential equations and establishes the foundation of the theory
for fractional differential equations. The results here should be fundamental also
for discussing other important topics such as nonlinear dynamical systems, optimal
control, and inverse problems for fractional differential equations. Although our
approach can work for more general fractional differential equations, focusing on a
differential equation with a single time-fractional derivative, we describe the theory.
The parts of the book have been presented as graduate courses at Sapienza
University of Rome and The University of Tokyo. The authors thank Professor
Bangti Jin (University College London) for valuable comments.
The first and second authors are supported partly by the National Science
Centre, Poland through 2017/26/M/ST1/00700 Grant, and the third author is partly
supported by Grant-in-Aid for Scientific Research (S) 15H05740 of Japan Society
for the Promotion of Science and by the National Natural Science Foundation of
China (nos. 11771270 and 91730303) and the “RUDN University Program 5–100.”
v
Remarks for Readers
vii
Contents
ix
x Contents
Derivatives of orders of natural numbers have been widely known and commonly
used since the origin of the calculus in the seventeenth century. On the other hand,
derivatives whose orders are not necessarily natural numbers seem not to be well-
known but have been considered since Leibniz (e.g., Ross [25]). Moreover, we can
refer to a historical paper [1] by N.H. Abel. He discussed the following problem:
Given a function f (t), find a function v(t) such that
t
1
√ (t − s)−1/2 v(s)ds = f (t), 0 ≤ t ≤ T. (1.1)
π 0
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 1
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5_1
2 1 Basics on Fractional Differentiation and Integration
t
integrating v(t). Thus we can interpret √1 d
π dt 0 (t − s)−1/2 f (s)ds as a function
which is differentiated 12 = − 12 + 1-times. That is, setting n = − 12 formally in
t √
(n+1) 0 (t − s) v(s)ds and noting that π = (1/2), we see that the right-hand
1 n
side of (1.2) means the half-times derivative of f . Thus we can interpret that (1.2)
means that the half-order derivative of f is v when f is the half-times integration of
v, and establishes the inversion of the half-order integral in terms of the half-order
derivative.
As for other applications of fractional derivatives, we refer to Gorenflo and
Vessella [8].
Although it is later seen that the rigorous treatments are necessary, in this chapter
we intuitively discuss fractional calculus as an introduction.
Throughout the book we assume
0 < α < 1.
and
t
1 du
dtα u(t) := (t − s)−α (s)ds. (1.4)
(1 − α) 0 ds
We call Dtα and dtα the Riemann–Liouville derivative and the Caputo derivative
respectively. The notations of the derivatives are different from e.g., Kilbas et
al. [15], Podlubny [24], and in Chap. 2, the fractional derivative ∂tα in fractional
Sobolev spaces is defined and plays an essential role rather than the classical
Riemann–Liouville derivative and the Caputo derivative.
Throughout this book, we interpret equalities in variables t, x, etc., as that they
hold for almost all t, x in domains under consideration, and we do not specify if
there is no fear of confusion.
In this book, we mainly consider the case 0 < α < 1, but for general α > 0, ∈ N,
choosing m ∈ N satisfying m − 1 < α < m, for u ∈ C m [0, T ], we can define dtα
and Dtα by
t
1 d mu
dtα u(t) = (t − s)m−α−1 (s)ds
(m − α) 0 ds m
and
t
1 dm
Dtα u(t) = (t − s)m−α−1 u(s)ds.
(m − α) dt m 0
1 Basics on Fractional Differentiation and Integration 3
u(0) −α
dtα u(t) + t = Dtα u(t), 0 ≤ t ≤ T. (1.8)
(1 − α)
t −α
Dtα 1 = .
(1 − α)
and
u(0) = 0. (1.11)
Moreover,
t s
For the last equality, we used the exchange of orders of integrals: 0 0 · · · dξ ds =
t t t du
0 ξ · · · ds dξ . By (1.7) we conclude that the right-hand side is 0 dξ (ξ )dξ =
u(t) − u(0), which completes the proof of (1.9). The proof of (1.12) is similar and
so is omitted.
For the Caputo derivative, the usual formulae for derivatives of orders of natural
numbers do not hold. For example, for u, v ∈ C 1 [0, T ], we have
but
and
β α+β
(dtα (dt u))(t) = dt u(t). (1.14)
The violation of the Leibniz rules (see (1.13)) implies the lack of integration by
parts, which causes many difficulties in considering fractional differential equations.
As for (1.14), we consider u(t) = t α and β = α ∈ (0, 1). Then (1.7) yields
dt t α = (1+α)
2α −α , but d α (d α t α ) = d α ((α + 1)) = 0, which means that d 2α t α =
(1−α) t t t t t
α α α
dt (dt t ).
As for the successive derivatives, we can directly prove the following. Let 0 <
α, β < 1 and α + β < 1. Then
β α+β
(dtα (dt u))(t) = dt u(t), 0≤t ≤T
β
for u ∈ C 1 [0, T ] satisfying u(0) = (dt u)(0) = 0, and
β α+β
(Dtα (Dt u))(t) = (Dt u)(t)
for u ∈ C 1 [0, T ].
du
It is important that the derivative characterizes the behavior, that is, increase
dt
or decrease, of u(t). We can accept that such a property holds in a limited sense
for the fractional derivatives, because the fractional derivative at t involves the
integral over (0, t). On the other hand, Luchko [20] proves the following extremum
principle.
Lemma 1.2 Let u ∈ C[0, T ] ∩ C 1 (0, T ] satisfy
T du
(t) dt < ∞. (1.15)
dt
0
We assume that u(t) attains the minimum at some t1 satisfying 0 < t1 ≤ T : u(t1 ) =
min0≤t ≤T u(t). Then
dtα u(t1 ) ≤ 0.
In defining dtα y(t1 ), we divide the integration interval into (0, δ) and (δ, t1 ), so that
for already chosen δ, we have
δ t1
1 dy 1 dy
dtα y(t1 ) = (t1 − s)−α (s)ds + (t1 − s)−α (s)ds =: I1 + I2 .
(1 − α) 0 ds (1 − α) δ ds
(1.16)
Then |I1 | < ε. Moreover, for δ > 0, since y ∈ C 1 (0, T ] and y(t1 ) = 0, we can
choose a constant Cδ > 0 such that |y(s)| ≤ Cδ |t1 − s|, δ ≤ s ≤ T . Therefore
integrating by parts, we have
(t1 − δ)−α y(δ) α t1
I2 = − − (t1 − s)−α−1 y(s)ds.
(1 − α) (1 − α) δ
We note that u ∈ W 1,1 (0, T ) if and only if there exists an absolutely continuous
function
u on [0, T ] such that u(t) = u(t) for almost all t ∈ (0, T ). Here and
henceforth we identify
u with u for u ∈ W 1,1 (0, T ).
We close this chapter with the following lemma which generalizes Lemma 1.1 to
u ∈ W 1,1 (0, T ),
Lemma 1.3
(i) For u ∈ W 1,1 (0, T ), we have J 1−α u ∈ W 1,1 (0, T ) and
d 1−α du u(0) −α
J u(t) = J 1−α (t) + t , 0 < t < T.
dt dt (1 − α)
1 Basics on Fractional Differentiation and Integration 7
(ii) Let un , u ∈ W 1,1 (0, T ) and un −→ u in W 1,1 (0, T ). Then dtα un −→ dtα u in
L1 (0, T ).
(iii) dtα u = Dtα u for u ∈ W 1,1 (0, T ) satisfying u(0) = 0.
(iv) Let α, β > 0. Then
J α (J β u) = J α+β u, u ∈ L2 (0, T ).
dt and 0 · · · ds is possible in
d
t
(t − s)−α u(s)ds.
0
Hence
t s
1 du
(J 1−α u)(t) = (t − s)−α (ξ )dξ + u(0) ds
(1 − α) 0 0 dξ
t s t
1 du 1
= (t − s)−α (ξ )dξ ds + (t − s)−α ds u(0)
(1 − α) 0 0 dξ (1 − α) 0
t t
1 du 1 t 1−α
= (t − s)−α ds (ξ )dξ + u(0).
(1 − α) 0 ξ dξ (1 − α) 1 − α
d 1−α du u(0) −α
J u(t) = J 1−α (t) + t . (1.18)
dt dt (1 − α)
8 1 Basics on Fractional Differentiation and Integration
Here we use Lemma A.1 in the Appendix, which is called the Young inequality for
the convolution:
1−α du 1 −α du
J (t) = t ∗ (t)
dt (1 − α) dt
and
1−α du 1 du
J
≤ −α
t < ∞.
dt L1 (0,T ) (1 − α)
L1 (0,T ) dt L1 (0,T )
Again the Young inequality Lemma A.1 for the convolution yields
1 d
dtα un − dtα u L1 (0,T ) ≤ t −α (un − u)
L1 (0,T ) 1
(1 − α) dt L (0,T )
1 T 1−α
≤ un − u W 1,1 (0,T ) .
(1 − α) 1 − α
Chapter 2
Definition of Fractional Derivatives
in Sobolev Spaces and Properties
2.1 Motivations
and find u(t) satisfying (2.1) with given f ∈ L2 (0, T ) and a ∈ R. Needless to say,
for α = 1, we have
t
u(t) = f (s)ds + a, 0 < t < T.
0
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 9
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5_2
10 2 Definition of Fractional Derivatives in Sobolev Spaces and Properties
is convenient, but the justification of the initial value is indispensable in order to use
the Laplace transform. Because by some repetition of calculations, the formula
is known (e.g., p. 106 in [24]) and we have to justify the sense of u(0), which
requires a certain smoothness of u at t = 0. Such regularity at t = 0 is not well
established for f ∈ L2 (0, T ). Moreover, for such f, we have to justify the solution
formula for (2.1) (e.g., Kilbas et al. [15, p. 141]).
We illustrate with one simple example. Let 0 < α < 12 and
1
f (t) = t δ− 2 , 0 < t < T,
Since δ − 12 > −1, we can formally apply the solution formula (e.g., [15, p. 141])
and obtain
t
1 1 1
u(t) = a + (t − s)α−1 s δ− 2 ds = a + C0 t α+δ− 2 , (2.3)
(α) 0
where we set
δ+ 1
2
C0 = .
α+δ+ 1
2
Moreover, u(t) given by (2.3) cannot satisfy (2.2) if 0 < α < 12 and δ > 0 is
small such that α + δ − 12 < 0. Indeed, limt ↓0 u(t) = ∞, and so the initial condition
1
does not follow common sense. Furthermore we formally calculate dtα t α+δ− 2 :
1 t d α+δ− 1 α + δ − 12 t
(t − s)−α (t − s)−α s α+δ− 2 ds.
1 3
dtα t α+δ− 2 = (s 2 )ds =
(1 − α) 0 ds (1 − α) 0
However, since α +δ − 32 < −1, the integral does not exist. This means that formula
(2.3) does not hold for f ∈ L2 (0, T ) in general, although (2.3) is convergent for
e.g., f ∈ L∞ (0, T ) as a solution formula to (2.1). We can verify that (2.3) gives
dt ∈ L (0, T ), because we can calculate dt u pointwise. In
a solution to (2.1) if du 2 α
terms of an extended Caputo derivative, we will further discuss this later (see, e.g.,
(2.36)).
In some references (e.g., Podlubny [24, pp. 78–79]), it is emphasized that the
Caputo derivative dtα can admit the initial value problem (2.1), unlike the Riemann–
1
Liouville derivative Dtα . However, this simple example f (t) = t δ− 2 shows that the
2.2 Preliminaries: Operational Structure of J α 11
pointwise Caputo derivative dtα cannot make the initial value problem (2.1) well-
posed for general f ∈ L2 (0, T ).
Moreover, for guaranteeing the uniqueness of u satisfying dtα u = f ∈ L2 (0, T ),
we certainly need some extra conditions. In fact, since dtα 1 = 0 by the definition, if
u satisfies dtα u = f in (0, T ), then u + C1 satisfies the same equation with arbitrary
constant C1 . These discussions suggest the necessity for reconsidering the pointwise
Caputo derivative dtα and redefining the Caputo derivative within the framework of
L2 (0, T ).
The function space L2 (0, T ) is reasonable and convenient as data space. Hence
it is natural to formulate the initial value problem and define dtα for f ∈ L2 (0, T ) in
order to establish a more unified theory for fractional differential equations. Thus we
construct the theory where the fractional derivatives should be included in L2 (0, T ).
This is our main motivation in this book, and we construct a seemingly different
fractional derivative ∂tα , but we will prove that it is essentially same as the closure
operator of the Caputo derivative in 0 C 1 [0, T ] (see Sect. 2.3).
First in Sects. 2.2 and 2.3, we define the generalized fractional derivative ∂tα in
Sobolev spaces Hα (0, T ). Then in Chaps. 3 and 4, in terms of such ∂tα , we formulate
an initial value problem and prove the well-posedness.
Our formulation is similar to Zacher [30] (see also Kubica and Yamamoto [17]),
but more essentially relies on the property of the generalized fractional derivative
∂tα defined later in some Sobolev space. These properties are feasible for the
applications such as the clarification of the Sobolev regularity of solutions to initial
boundary value problems.
By the direct observation of the definition (1.4) of the pointwise Caputo derivative,
we need the first derivative du α
ds (s) in order to define the derivative dt u of order
α < 1. In order to define such an adequate fractional derivative, which is denoted
by ∂tα , we should fulfill the following:
1. ∂tα should be well-defined in a subspace of the Sobolev space of order α;
2. the norm equivalence between ∂tα u L2 (0,T ) and some conventional norm of u
such as the norm in a Sobolev space.
For them,
• We will interpret J α as the fractional power of the operator defined by
t
J u(t) = u(s)ds
0
The arguments in this section and a part of Sect. 2.3 are based on Gorenflo and
Yamamoto [9], Gorenflo et al. [11].
By L2 (0, T ) and H α (0, T ) we mean the usual L2 -space and the fractional
Sobolev space on the interval (0, T ) (see e.g., [2, Chapter VII]), respectively, and
we define the norm in H α (0, T ) by
12
2
T T |u(t) − u(s)|2
u H α (0,T ) := u L2 (0,T )
+ dtds .
0 0 |t − s|1+2α
The L2 -norm and the scalar product in L2 are denoted by · L2 (0,T ) and (·, ·) =
(·, ·)L2 (0,T ) , respectively. By ∼ we denote a norm equivalence. Since J α is injective
in L2 (0, T ), by J −α we denote the algebraic inverse to J α .
We set
1
0H
α
(0, T ) = {u ∈ H α (0, T ); u(0) = 0} if < α ≤ 1.
2
We further define the Banach spaces
⎧ 1
⎪ α
⎨ 0H (0, T1 ), 2 < α < 1,
T |v(t )|2
Hα (0, T ) := v ∈ H (0, T ); 0
2 dt < ∞ , α = 12 ,
⎪
⎩ α
t
1
H (0, T ), 0 < α < 2
For later arguments, we need some convenient and not too narrow subspace of
Hα (0, T ). In particular, for H 1 (0, T ), such a subspace is more delicate.
2
We recall
0W
1,1
(0, T ) = {u ∈ W 1,1 (0, T ); u(0) = 0}
du
(t) ≤ Cu t α−1 almost all t, u(0) = 0 . (2.5)
dt
Wα (0, T ) ⊂ Hα (0, T ).
Proof of (2.6) The conclusion (2.6) is trivial for γ1 ≥ or γ2 ≥ 0. Therefore we can
assume that −1 < γ1 , γ2 < 0. We note
T T T t
(s γ1 + t γ1 )|t − s|γ2 dsdt = 2 (s γ1 + t γ1 )|t − s|γ2 ds dt. (2.7)
0 0 0 0
Indeed,
T T T t T
(s γ1 + t γ1 )|t − s|γ2 dsdt = + (s γ1 + t γ1 )|t − s|γ2 ds dt.
0 0 0 0 t
t
Now we complete the proof of Lemma 2.1. Indeed, u(t) = du
0 ds (s) for u ∈
Wα (0, T ). Hence
t
t du du
|u(t) − u(s)| = (ξ )dξ ≤ (ξ ) dξ
s dξ s dξ
1
Let 2 < α < 1. Then
|u(t) − u(s)|2
≤ C(t 2α−2 + s 2α−2 )|t − s|1−2α .
|t − s|1+2α
Since 2α − 2 > −1 by 1
2 < α < 1 and 1 − 2α > −1 by 0 < α < 1, the inequality
(2.6) implies
T T |u(t) − u(s)|2
dsdt < ∞.
0 0 |t − s|1+2α
Since
du
(ξ ) ≤ Cu ξ α−1 ≤ Cu s α−1
dξ
for α = 12 , we have
t
t du
|u(t)| = (s)ds ≤
1 1
Cu s − 2 ds = 2Cu t 2 ,
0 ds 0
T
and so 0 |u(tt )| dt < ∞, which implies that H 1 (0, T ). Thus the proof of
2
2
Lemma 2.1 is completed.
In fact, the space Wα (0, T ) is a convenient subspace of Hα (0, T ).
Remark 2.1 For H 1 (0, T ), Lions and Magenes [19] use a different notation
2
1
2
0 H0(0, T ) (Remark 11.5 (p. 68) in [19] vol. I). However, we use H 1 (0, T ) as well
2
as Hα (0, T ), 0 < α < 1 throughout this book.
Henceforth we set
1
0C [0, T ] = {ϕ ∈ C 1 [0, T ]; ϕ(0) = 0}
which is the interpolation space. Applying Proposition 6.1 (p. 28) in [19], for 12 <
γ < 1, we see that [0 H 1 (0, T ), L2 (0, T )]1−γ is dense in [0 H 1 (0, T ), L2 (0, T )]1−α
by 1 − γ ≤ 1 − α. Therefore
Hα (0,T )
Hγ (0, T ) = Hα (0, T )
1
for 2 < γ < 1. As is already proved, we see that
H (0,T )
1 [0, T ] γ = Hγ (0, T ).
0C
16 2 Definition of Fractional Derivatives in Sobolev Spaces and Properties
Hα (0,T )
1 [0, T ] = Hα (0, T )
0C
H α (0,T ) 1
1 [0, T ] = H α (0, T ), 0≤α≤ (2.8)
0C .
2
1 [0, T ]
H α (0,T ) H α (0, T ), 0 < α ≤ 12 ,
0C = (2.9)
Hα (0, T ), 12 < α < 1.
J α L2 (0, T ) ⊂ H1 (0, T ), α ≥ 1.
However, in this book we limit the range of α to 0 < α ≤ 1 and we omit further
characterization of J α L2 (0, T ).
In terms of Theorem 2.1, (J α )−1 exists and J −α = (J α )−1 . Then:
Theorem 2.2 There exists a constant C > 0 such that
J −α J α u = u, u ∈ L2 (0, T )
and
J α J −α u = u, u ∈ Hα (0, T ).
The first equality is directly seen by the definition, while the second equality
is verified as follows. For u ∈ Hα (0, T ), Theorem 2.1 (i) yields the existence of
w ∈ L2 (0, T ) satisfying u = J α w. Therefore J α J −α u = J α J −α (J α w) = J α w.
Hence J α J −α u = u for u ∈ Hα (0, T ).
Henceforth we write for example (2.10)
J −α v L2 (0,T ) ∼ v Hα (0,T ) ,
for u ∈ L2 (0, T ). We will prove (2.12). We set v(t) = (λI + J )−1 u(t), that is,
t
λv(t) + v(s)ds = u(t), 0 < t < T. (2.13)
0
1
v(0) = u(0). (2.14)
λ
Differentiating (2.13) with respect to t, we have
dv 1 1 du
(t) + v(t) = (t).
dt λ λ dt
With (2.14), we obtain
t
1 −1t 1 du
v(t) = e λ eλs (s)ds + u(0) .
λ 0 dt
Next let u ∈ L2 (0, T ). Then the right-hand side of (2.15) defines a function v ∈
t s t t
L2 (0, T ). Exchanging the orders of the integrals: 0 0 dξ ds = 0 ξ ds dξ ,
we calculate
t t s
1 t 1 s−ξ
v(s)ds = u(s)ds − 2 e− λ u(ξ )dξ ds
0 λ 0 λ 0 0
t t t
1 1 ξ
e− λ ds e λ u(ξ )dξ
s
= u(s)ds − 2
λ 0 λ 0 ξ
t t
1 1 s s=t ξ
= u(s)ds + e− λ e λ u(ξ )dξ
λ 0 λ 0 s=ξ
t t
1 1 −t ξ 1 t
= u(s)ds + e λ e u(ξ )dξ −
λ u(ξ )dξ
λ 0 λ 0 λ 0
1 t t s
= e− λ e λ u(s)ds.
λ 0
t
This and (2.15) yield λv(t) + 0 v(s)ds = u(t), 0 < t < T for each u ∈ L2 (0, T ),
which is (2.12).
2.2 Preliminaries: Operational Structure of J α 19
t −s
and by change of variables η = λ , we obtain
∞
sin πα
λα−1 (λI + J )−1 J u(t) dλ
π 0
∞ t
sin πα −(t −s)/λ
= λα−2
e u(s)ds dλ
π 0 0
t ∞
sin πα
= u(s) λα−2 e−(t −s)/λdλ ds
π 0 0
t ∞
sin πα
= u(s) η−α e−η dη (t − s)α−1 ds
π 0 0
(1 − α) sin πα t
= u(s)(t − s)α−1 ds.
π 0
α
According to [9], the domain D(S 2 ) can be described as follows:
α
D(S 2 ) = Hα (0, T ). (2.20)
The relation (2.20) holds not only algebraically but also topologically:
α α
S2v L2 (0,T ) ∼ v Hα (0,T ) , 0 ≤ α ≤ 1, v ∈ D(S 2 ). (2.21)
α
In particular, the inclusion D(S 2 ) ⊂ H α (0, T ) holds true.
Now we are ready to prove Theorem 2.1.
Proof of Theorem 2.1 We first state the Heinz–Kato inequality (e.g., Theo-
rem 2.3.4 in Tanabe [27]): let X be a Hilbert space and linear operators A, B
in X satisfy D(A) = D(B), Re (Au, u) ≥ 0, Re (Bu, u) ≥ 0 for u ∈ D(A) and
R(I + A) = R(I + B) = X. We assume that there exists a constant C > 0 such that
Bu ≤ C Au , u ∈ D(A).
The proof of the theorem is done by using the norm equivalence between
· α and Hα (0, T ), which is justified by the Heinz–Kato inequality.
D(S )
2
First of all, it can be directly verified that D(J −1 ) = J (L2 (0, T )) = 0 H 1 (0, T ),
dw(t)
(J −1 w)(t) = , and J −1 v L2 (0,T ) = v H 1 (0,T ) for v ∈ 0 H 1 (0, T ).
dt
Therefore by (2.21) we obtain the norm equivalence
J −α v
α α
D(J −α ) = D(S 2 ).
α
L2 (0,T ) ∼ S2v L2 (0,T ) , v ∈ D(S 2 ), (2.22)
α
By (2.20) and (2.22), we have Hα (0, T ) = D(S 2 ) = D(J −α ) = R(J α ),
so that Theorem 2.1 (i) follows. By (2.21) and (2.22), the norm equivalence
J −α v L2 (0,T ) ∼ v Hα (0,T ) holds true for v ∈ D(J −α ) = R(J α ). Next, setting
v = J α u ∈ D(J −α ) with any u ∈ L2 (0, T ), by (2.21), we obtain the following
norm equivalence:
α
u L2 (0,T ) ∼ S 2 (J α u) L2 (0,T ) ∼ J αu H α (0,T ) , u ∈ L2 (0, T ).
Remark 2.3 Since J α is defined by the fractional power of the operator J and J −α
by J −α = (J α )−1 for 0 ≤ α ≤ 1, the general theory (e.g., Tanabe [27]) implies
J α (J β u) = J α+β u, u ∈ L2 (0, T ),
L2 (0, T ), γ ≤ 0,
D(J −γ ) =
Hγ (0, T ), γ ≥ 0.
Proof
(i) We note that
First let u ∈ J α (Hβ (0, T )), that is, u = J α ϕ with some ϕ ∈ Hβ (0, T ). By
Theorem 2.1 (i), there exists ψ ∈ L2 (0, T ) such that ϕ = J β ψ. Then, noting
that J α is defined as fractional power of J , we see that u = J α ϕ = J α (J β ψ) =
J α+β ψ, that is, u ∈ Hα+β (0, T ) by Theorem 2.1 (i). Hence J α (Hβ (0, T )) ⊂
Hα+β (0, T ). Next let u ∈ Hα+β (0, T ). Then Theorem 2.1 yields the existence
of ϕ ∈ L2 (0, T ) such that u = J α+β ϕ. Since J α+β = J α J β , we obtain
u = J α (J β ϕ) ∈ J α (Hβ (0, T )). Hence J α : Hβ (0, T ) −→ Hα+β (0, T ) is
surjective.
The norm equivalence (2.23) is verified as follows:
J −β J α u = J −β J α (J β−α v) = J −β J β v.
∂tα u := J −α u, u ∈ Hα (0, T )
with
d m−1 u d mu
D(∂tα ) = u ∈ H (0, T ); u(0) = · · · = m−1 (0) = 0, m ∈ Hγ (0, T )
m
dt dt
(cf. (3.39)), and we can argue the isomorphism and fractional differential equations
in the same way as the later parts of this book, but we omit the details and postpone
them to forthcoming publications.
By Theorem 2.1, we note that Hα (0, T ) = J α L2 (0, T ). Therefore ∂tα in
Hα (0, T ) is well-defined and ∂tα u ∈ L2 (0, T ) for u ∈ Hα (0, T ). Moreover,
∂tα : Hα (0, T ) −→ L2 (0, T ) is surjective. Indeed, let v ∈ L2 (0, T ) be arbitrarily
given. By Theorem 2.1, we have ϕ := J α v ∈ Hα (0, T ) and so ∂tα ϕ = v by the
definition, which means that ∂tα : Hα (0, T ) −→ L2 (0, T ) is surjective.
On the other hand, we can prove
du
J −1 u = , u ∈ H1 (0, T ).
dt
J −α = J (1−α)−1 = J −1 J 1−α ,
that is, J −α = d
dt (J
1−α ).
24 2 Definition of Fractional Derivatives in Sobolev Spaces and Properties
Moreover,
d 1−α
∂tα u = J −α u = (J u) = Dtα u, u ∈ Hα (0, T ) (2.26)
dt
and
Indeed, (2.27) follows from (2.26) and Lemma 1.3. Thus we can calculate ∂tα u
by means of Dtα u for u ∈ Hα (0, T ).
We are tempted to assert (2.27) for u ∈ Hα (0, T ). However, dtα u does not directly
make sense for u ∈ Hα (0, T ) and with suitable extension of dtα , we can transfer
(2.27) to Hα (0, T ) (see Theorem 2.5).
Formula w = ∂tα u = dt d
(J 1−α u) in (2.26) can correspond to the classical
inversion for finding w solving J α w = u (e.g., Gorenflo and Vessella [8]) for u ∈
1,1 (0, T ), but our construction for ∂ α guarantees the formula for u ∈ H (0, T ),
0W t α
which is a wider space than the set of all absolutely continuous functions on [0, T ].
Moreover,
Proposition 2.1
du
J 1−α (t) = dtα u(t) = Dtα u(t), u ∈ 0 W 1,1 (0, T ).
dt
Therefore
du
J α J 1−α (t) = J α dtα u(t) = J α Dtα u(t), u ∈ 0 W 1,1 (0, T ).
dt
2.3 Definition of Generalized ∂tα in Hα (0, T ) 25
By J α J 1−α = J we have
t
α 1−α du du du
J J (t) = J (t) = (s)ds = u(t)
dt dt 0 ds
This theorem means that our definition of ∂tα is consistent with the classical Caputo
derivative by considering the closure of the operator.
Proof of Theorem 2.5 We first prove that the operator dtα is closable.
Indeed, let un ∈ 0 C 1 [0, T ] and un −→ 0 in L2 (0, T ) and dtα un −→ v in L2 (0, T )
with some v ∈ L2 (0, T ). Then by Lemma 1.3 and Theorem 2.4, we have dtα un =
∂tα un , n = 1, 2, 3, . . . Therefore ∂tα un −→ v in L2 (0, T ). Theorem 2.4 yields
that un ∈ 0 C 1 [0, T ] ⊂ Hα (0, T ) is a Cauchy sequence in Hα (0, T ). Therefore
there exists some u ∈ Hα (0, T ) such that un −→ u in Hα (0, T ). Since un −→ 0
in L2 (0, T ), we see that u = 0, that is, un −→ 0 in Hα (0, T ). Hence again by
Theorem 2.4, we obtain ∂tα un = dtα un −→ ∂tα 0 = 0, that is, v = 0. Thus dtα is
closable.
Now we return to the proof of Theorem 2.5. We recall that u ∈ D(dtα ) if and only
if there exist un ∈ 0 C 1 [0, T ] such that un −→ u in L2 (0, T ) and dtα un is a Cauchy
sequence in L2 (0, T ). Since dtα un = ∂tα un for un ∈ 0 C 1 [0, T ] by Lemma 1.3 and
Theorem 2.4, we see that un is a Cauchy sequence in Hα (0, T ). Therefore there
exists u ∈ Hα (0, T ) such that un −→ u in Hα (0, T ) by Theorem 2.4. Since un −→
u in L2 (0, T ), we obtain u = u and un −→ u in Hα (0, T ) and dtα un −→ ∂tα u in
26 2 Definition of Fractional Derivatives in Sobolev Spaces and Properties
Hence we proved that D(dtα ) ⊂ Hα (0, T ) and dtα u = ∂tα u for u ∈ D(dtα ).
H (0,T )
Conversely assume that u ∈ Hα (0, T ). Since 0 C 1 [0, T ] α = Hα (0, T ) by
Lemma 2.2, there exist un ∈ 0 C 1 [0, T ] such that un −→ u in Hα (0, T ). Then
∂tα un = dtα un −→ ∂tα u in L2 (0, T ) by Theorem 2.4. That is, un ∈ 0 C 1 [0, T ] is dtα -
convergent to u and so u ∈ D(dtα ) and dtα u = ∂tα u. Thus the proof of Theorem 2.5
is complete.
Remark 2.5 In Sect. 2.5, we discuss the case where we start the operator dtα with
D(dtα ) = C 1 [0, T ], not 0 C 1 [0, T ].
In this section, we show that the function t γ and some functions defined by the
Mittag-Leffler functions are in Hα (0, T ). Some of these inclusions are used in later
sections.
For α, β > 0, we define
∞ zk
Eα,β (z) = k=0 (αk+β) ,
∞ zk
(2.28)
Eα,1 (z) = k=0 (αk+1) , z ∈ C.
These functions are called the Mittag-Leffler functions and play important roles in
the fractional calculus (e.g., [15, 24]).
It is known that Eα,β (z) is an entire function in z with α, β > 0. The Mittag-
Leffler functions have been well studied and here we describe only a few of the
important properties used later.
Lemma 2.5
(i) Let 0 < α < 2 and β > 0. We assume that πα 2 < μ < min{π, πα}. Then there
exists a constant C = C(α, β, μ) > 0 such that
C
|Eα,β (z)| ≤ , μ ≤ | arg z| ≤ π.
1 + |z|
dm
Eα,1 (−λt α ) = −λt α−m Eα,α−m+1 (−λt α ), t > 0.
dt m
2.4 Some Functions in Hα (0, T ) 27
The proof of (i) can be found e.g., in [24]. The proof of (ii) is seen directly
because Eα,1 (z) is entire in z and we can differentiate Eα,1 (−λt α ) = ∞ (−λt α )k
k=0 (αk+1)
termwise.
Proposition 2.2 Let 0 < α < 1 and λ ∈ R. Then
d
Eα,1 (−λt α ) = −λt α−1 Eα,α (−λt α )
dt
and
d
Eα,1 (−λt α ) ≤ |λ|t α−1 |Eα,α (−λt α )| ≤ C|λ|t α−1 , t > 0.
dt
By Eα,1 (0) − 1 = 1
(1) − 1 = 0, the definition of Wα (0, T ) implies the
conclusion.
Proposition 2.3 Let 0 < α < 1. Then
t
(t − s)α−1 Eα,α (−λ(t − s)α )f (s)ds ∈ Hα (0, T )
0
and
t
(t − s)α−1 Eα,α (−λ(t − s)α )f (s)ds ≤C f
L2 (0,T )
0 Hα (0,T )
Proof We set
∞
(−λ)k s αk+α−1
gk (s) := , k = 0, 1, 2, 3, .., R(s) = gk (s).
(α(k + 1))
k=N
Then
∞
N −1
(−λ)k (t − s)αk+α−1
(t − s)α−1 Eα,α (−λ(t − s)α ) = = gk (t − s) + R(t − s).
(α(k + 1))
k=0 k=0
(2.29)
αk + α − 1 ≥ αN + α − 1 > 0 if k ≥ N,
ds ∈ L (0, T ). Therefore
Since αN + α − 2 > −1 by (2.30), we see that dR 1
t t
d dR
R(t − s)f (s)ds = (t − s)f (s)ds,
dt 0 0 dt
so that the Young inequality for the convolution (Lemma A.1 in the Appendix)
yields
t
d
R(t − s)f (s)ds ≤C f
dt 2 L2 (0,T ) ,
0 L (0,T )
2.4 Some Functions in Hα (0, T ) 29
t
which means that 0 R(t − s)f (s)ds ∈ W 1,1 (0, T ). Moreover,
t t
d
R(t − s)f (s)ds ≤C (t − s)αN+α−2 |f (s)|ds.
dt
0 0
Since αN > 32 − α by (2.30), we have 2αN + 2α − 4 > −1. Therefore the Cauchy–
Schwarz inequality and (2.30) yield
t t 12
d
R(t − s)f (s)ds ≤ C (t − s) 2αN+2α−4
ds f
dt L2 (0,T )
0 0
C 3
= 1
t αN+α− 2 f L2 (0,T ) ≤ C1 f L2 (0,T ) .
(2αN + 2α − 3) 2
t
Again by (2.30), we see that αN + α − 1 > 0, so that 0 R(t − s)f (s)ds vanishes
at t = 0. Hence we proved
t
R(t − s)f (s)ds ∈ 0 W 1,1 (0, T ) ∩ Wα (0, T ) ⊂ 0 W 1,1 (0, T ) ∩ Hα (0, T ).
0
Hence
t t s
α
−α d
∂
R(t − s)f (s)ds ≤ C (t − s) R(s − ξ )f (ξ )dξ ds
t ds 0
0 0
t
≤C (t − s)−α ds × C1 f L2 (0,T ) ≤ CT 1−α f L2 (0,T ) .
0
Therefore
t 2
α T
∂ R(t − s)f (s)ds ≤ C 2 T 2−2α f 2
dt,
t L2 (0,T )
0 L2 (0,T ) 0
N−1 t
= (−λ)k (Jαk+α f )(t) + R(t − s)f (s)ds.
k=0 0
and so
N−1
N−1
(−λ)k Jαk+α f ∈ Hαk+α (0, T ) ∩ H1 (0, T ) = Hα (0, T ).
k=0 k=0
1
t α+δ− 2 ∈ Hα (0, T ) = H α (0, T ).
where
Γ α+δ+ 1
2 1
w0 (t) := t δ− 2 ∈ L2 (0, T ),
Γ δ+ 1
2
1
θ = −α − δ + .
2
Using
T T T s
· · · ds dt = · · · dt ds,
0 t 0 0
We can prove
T t
s (t − s) t ds dt < ∞
γ1 γ2 γ3
(2.33)
0 0
|t θ − s θ | ≤ |t − s|θ . Hence
|t θ − s θ |2μ ≤ |t − s|2μθ
for 0 ≤ s ≤ t. On the other hand, the mean value theorem and θ < 1 yield
for 0 ≤ s ≤ t. Therefore
and so
Hence
T t |t θ − s θ |2 t −2θ s −2θ T t
dsdt ≤ C s 2(1−μ)(θ −1)−2θ t −2θ |t − s|1−2α−2μ+2μθ dsdt.
0 0 |t − s|1+2α 0 0
and
so that (2.33) yields (2.34). Thus the proof of Proposition 2.4 is complete.
Remark 2.6 It seems that we cannot directly prove the lemma without introducing
the parameter μ ∈ (0, 1).
1 δ+ 12
Since Proposition 2.4 proves C0 t α+δ− 2 ∈ Hα (0, T ), where C0 = , we
α+δ+ 21
can apply Theorem 2.5 to calculate
1 1 1
∂tα (C0 t α+δ− 2 ) = dtα (C0 t α+δ− 2 ) = Dtα (C0 t α+δ− 2 )
t
C0 d
(t − s)−α s α+δ− 2 ds
1
=
(1 − α) dt 0
C0 (1 − α) α + δ + 1
2 d δ+ 1 1
= t 2 = t δ− 2 .
(1 − α) δ+ 3 dt
2
Here we used δ + 32 = δ + 12 δ + 12 .
Therefore with the closure of dtα defined in Sect. 2.3 and ∂tα , we can justify
1 1
dtα (C0 t α+δ− 2 ) = t δ− 2 (2.36)
We have defined ∂tα u for u ∈ Hα (0, T ). Next, it is natural to define the fractional
derivative in H α (0, T ) which is wider than Hα (0, T ) for 0 < α < 1, although we
work mainly within Hα (0, T ) for discussing fractional differential equations.
Since Hα (0, T ) is a set of functions in H α (0, T ) which “vanish” in some sense
at t = 0, the following definition is naive but may be reasonable for 0 < α < 1 with
α = 12 : we define ∂tα v for v ∈ H α (0, T ) by
Then ∂tα u is well-defined because Hα (0, T ) = H α (0, T ) for 0 < α < 12 , and
H α (0, T ) ⊂ C[0, T ] for 12 < α < 1 which enables us to define v(0).
For 0 < α < 1 and α = 12 , we note that this definition gives the same result for
v ∈ Hα (0, T ), and so we can extend the domain of ∂tα from Hα (0, T ) to H α (0, T ).
1
On the other hand, the case of α = 1
2 is delicate, and we do not define ∂t2 u for
1 1
u ∈ H 2 (0, T ) H 1 (0, T ). For example, we note that 1 ∈ H 2 (0, T ) but 1 ∈
2
1 1
H 1 (0, T ). We do not define ∂t2 1 in L2 (0, T ), although we can calculate dt2 1 = 0
2
1
1 − 12
and Dt2 1 = t pointwise.
( 21 )
By the definition we note that
(ii)
for u ∈ C 1 [0, T ].
34 2 Definition of Fractional Derivatives in Sobolev Spaces and Properties
(iii) Let 12 < α < 1 and u ∈ H α (0, T ). Then there exists a sequence ϕk ∈ C 1 [0, T ],
k ∈ N such that ϕk −→ u in H α (0, T ) and ∂tα u = limk→∞ ∂tα ϕk in L2 (0, T ).
(iv) Let 0 < α < 12 . Then there exists a constant C > 0 such that
in L2 (0, T ).
2.5 Definition of ∂tα in H α (0, T ) 35
(iv) For 0 < α < 12 , we note that Hα (0, T ) = H α (0, T ), and so u ∈ H α (0, T )
implies u − a ∈ Hα (0, T ). Therefore, by Theorem 2.2, we have
Therefore
and
operator dtα with the domain D(dtα ) = 0 C 1 [0, T ] and we establish Theorem 2.5.
Our discussion here is similar. We consider the classical Caputo derivative
t
1 du
α
0 dt u(t) := (t − s)−α (s)ds, D(0 dtα ) = C 1 [0, T ].
(1 − α) 0 ds
To avoid confusion, by 0 dtα we denote the classical Caputo fractional derivative with
the domain C 1 [0, T ], and we consider 0 dtα as an operator from C 1 [0, T ] ⊂ L2 (0, T )
to L2 (0, T ).
First we prove:
Lemma 2.6 For 1
2 ≤ α < 1, the operator 0 dtα is closable.
Proof We have to prove that w = 0 if un ∈ C 1 [0, T ] = D(0 dtα ), n = 1, 2, 3, . . .,
un −→ 0 in L2 (0, T ) and 0 dtα un converge to w in L2 (0, T ). By un ∈ C 1 [0, T ],
n = 1, 2, 3, . . ., we see that un − un (0) ∈ 0 C 1 [0, T ] ⊂ Hα (0, T ). Therefore (2.27)
implies
α
0 d t un = 0 dtα (un − un (0)) = dtα (un − un (0)) = ∂tα (un − un (0)).
Hence ∂tα (un − un (0)) −→ w in L2 (0, T ), and so Theorem 2.4 implies un − un (0)
is a Cauchy sequence in Hα (0, T ), and there exists v ∈ Hα (0, T ) such that
Finally, we prove:
Theorem 2.6
(i)
1
D(d
t ) = H (0, T )
α α
if <α<1
2
2.5 Definition of ∂tα in H α (0, T ) 37
and
H 1 (0, T ) ⊂ D(d
1
t ) ⊂ H 2 (0, T ).
α
2
(ii) Let 0 < α < 12 , and let ∂tα be defined by (2.37) for all u ∈ H α (0, T ). Then
d
t = ∂t
α α
on H α (0, T ). (2.42)
Proof
First Step We prove
1
D(d
t ) ⊂ H (0, T ),
α α
≤ α < 1. (2.43)
2
Proof of (2.43) Let u ∈ D(d α α
t ). The definition of dt implies that there exists a
sequence un ∈ C [0, T ], n = 1, 2, 3, . . . such that
1
un −→ u, α
0 d t un −→ dα
t u in L2 (0, T ). (2.44)
un − un (0) −→
u in Hα (0, T ) ⊂ H α (0, T ). (2.45)
Hα (0, T ) ⊂ D(dα
t ) (2.46)
and
d
t u = ∂t u,
α α
u ∈ Hα (0, T ). (2.47)
38 2 Definition of Fractional Derivatives in Sobolev Spaces and Properties
Proof of (2.46) and (2.47) By Lemma 2.2, for any u ∈ Hα (0, T ), we can choose
a sequence un ∈ 0 C 1 [0, T ] ⊂ Hα (0, T ), n = 1, 2, 3, . . ., such that un −→ u
in Hα (0, T ). Theorem 2.4 implies that ∂tα un −→ ∂tα u in L2 (0, T ). On the other
hand, Lemma 1.3 and (2.27) yield that 0 dtα un = ∂tα un , and so 0 dtα un −→ ∂tα u in
L2 (0, T ). Since un ∈ C 1 [0, T ], by the definition of
dtα , we obtain u ∈ D(dα
t ) and
(2.47).
Third Step The inclusions (2.43) and (2.46) imply
Hα (0, T ) ⊂ D(d
t ) ⊂ H (0, T ).
α α
(2.48)
1
To complete the proof of Theorem 2.6, for 2 < α < 1, we have to prove
H α (0, T ) ⊂ D(dα
t ),
dtα u = ∂tα (u − u(0)) for u ∈ H α (0, T ).
On the other hand, in terms of (2.49) we see from Theorem 2.4 and (2.27) that
α
0 dt (un + u(0)) = 0 dtα un = ∂tα un −→ ∂tα (u − u(0))
complete.
In Sects. 2.3 and 2.5, we discussed the two extensions of the classical Caputo
derivative by taking the closures. As another method for the extension, we here
argue the adjoint operator and show the adjoint equality.
The results in this section are not used in the later parts of the book.
2.6 Adjoint of the Fractional Derivative in Hα (0, T ) 39
We introduce:
Definition 2.2 The formal adjoint operator (dtα )∗ to dtα is defined by
T
1 dϕ
(dtα u, ϕ) = (u, (dtα )∗ ϕ) + t 1−α (t)dt u(0) (2.51)
(1 − α)(1 − α) 0 dt
Proof Let u ∈ C 1 [0, T ]. Then, exchanging the orders of the integrals, we have
t
T
1 −α du
=
(dtα u, ϕ) (t − s) (s)ds ϕ(t)dt
0 (1 − α) 0 ds
T T
1 du
= (t − s)−α ϕ(t)dt (s)ds.
(1 − α) 0 s ds
T
1 d 1−α dϕ
+ u, (t − s) (t)dt
(1 − α)(1 − α) ds s dt
T T
−1 dϕ 1 dϕ
= u, (t − s)−α (t)dt + t 1−α (t)dt u(0)
(1 − α) s dt (1 − α)(1 − α) 0 dt
for u ∈ C 1 [0, T ] and ϕ ∈ C 1 [0, T ] satisfying ϕ(T ) = 0. The proof of Lemma 2.7
is completed.
Next:
Definition 2.3 For u ∈ Hα (0, T ), we define dtα u = f if there exists f ∈ L2 (0, T )
such that
1
(dtα 1)(t) = t −α ,
(1 − α)
and repeating the computations in (2.52), we see that (dtα u, ϕ) = (u, (dtα )∗ ϕ) =
(f, ϕ) for all ϕ ∈ C01 (0, T ). Thus the proof is complete.
In fact, we can further prove:
Proposition 2.6
Hα (0, T ) ⊂ D(dtα )
and
In the above example, we see that 1 ∈ D(dtα ), but 1 ∈ Hα (0, T ) for 12 ≤ α < 1.
Therefore Hα (0, T ) D(dtα ) for 12 ≤ α < 1.
Proposition 2.6 generalizes (2.27) with the extension dtα and we notice that dtα
cannot be calculated in general for u ∈ Hα (0, T ).
δ+ 12
Example 2.2 Let 0 < α < 12 , δ > 0 and C0 = . Proposition 2.4 yields
α+δ+ 12
1
t α+δ− 2 ∈ Hα (0, T ) and Proposition 2.6 implies
1 1 1 1
dtα (C0 t α+δ− 2 ) = ∂tα (C0 t α+δ− 2 ) = Dtα (C0 t α+δ− 2 ) = t δ− 2 .
Moreover, we can directly verify the above result by the definition of dtα . For this, it
suffices to prove
1 1
(C0 t α+δ− 2 , (dtα )∗ ϕ) = (t δ− 2 , ϕ), ϕ ∈ C01 (0, T ).
T T
Indeed, using Lemma 2.7, integration by parts and 0 t · · · dη dt
T η
= 0 0 · · · dt dη, we obtain
1
T 1 −1 T dϕ
(C0 t α+δ− 2 , (dtα )∗ ϕ) = C0 t α+δ− 2 (η − t)−α (η)dη dt
0 (1 − α) t dη
42 2 Definition of Fractional Derivatives in Sobolev Spaces and Properties
−C0 T dϕ η
α+δ− 12 −α 1 T 1 dϕ
= (η) t (η − t) dt dη = − ηδ+ 2 (η)dη
(1 − α) 0 dη 0 δ+ 1
2 0 dη
T
1 1 η=T 1 1
=− [ϕ(η)ηδ+ 2 ]η=0 + ηδ− 2 ϕ(η)dη = (t δ− 2 , ϕ).
δ+ 1
2 0
Proof of Proposition 2.6 Let u ∈ Hα (0, T ) be arbitrary. By Lemma 2.2, there exist
un ∈ 0 C 1 [0, T ], n = 1, 2, 3, . . . , such that un −→ u in Hα (0, T ). By Theorem 2.4,
we note that ∂tα un = dtα un , n = 1, 2, 3, . . .. By Theorem 2.2, we have dtα un =
∂tα un −→ ∂tα u in L2 (0, T ). By using un (0) = 0, Lemma 2.7, un ∈ 0 C 1 [0, T ] and
Definition 2.3 imply
Letting n → ∞, we obtain
Since ∂tα u ∈ L2 (0, T ) for u ∈ Hα (0, T ), Definition 2.3 yields that u ∈ D(dtα ) and
dtα u = ∂tα u. Combining (2.26), we complete the proof of Proposition 2.6.
The adjoint equality (2.51) for u ∈ 0 C 1 [0, T ] is generalized as follows.
Proposition 2.7
For u belonging to a certain class, we can define the Laplace transform (Lu)(p) by
∞
(Lu)(p) := e−pt u(t)dt
0
for Re p > p0 : some constant. The formula (2.54) is convenient for solving
fractional differential equations. However, formula (2.54) requires some regularity
for u in order that u(0) is well-defined. Formula (2.54) does not make sense for all
u ∈ H α (0, T ) with 0 < α < 12 .
Moreover, such needed regularity should be consistent with the regularity which
we can prove for solutions to fractional differential equations. In particular, the
regularity for the formula concerning the Laplace transform should be not very
strong. Thus on the regularity assumption for the formula like (2.54), we have to
make adequate assumptions for u.
In this section, we state the formula of the Laplace transform for the fractional
derivative ∂tα in Hα (0, T ).
We set
there exists a constant C = Cu > 0 such that |u(t)| ≤ CeCt for t ≥ 0}.
and
Proof First for u ∈ Hα (0, T ), by Theorem 2.3 (i), we can see that
and so
d 1−α
Dtα u = J u ∈ L2 (0, T ).
dt
d 1−α
Theorem 2.4 yields ∂tα u = J u for u ∈ Hα (0, T ). Let T > 0 be arbitrarily
dt
fixed. Then, in terms of (2.56), we integrate by parts to obtain
T T
d 1−α
e−pt ∂tα u(t)dt = (J u)(t)e−pt dt
0 dt 0
t=T T
= J 1−α u(t)e−pt +p e−pt J 1−α u(t)dt
t=0 0
e−pT T
−α p T
−pt
t
−α
= (T − s) u(s)ds + e (t − s) u(s)ds dt
(1 − α) 0 (1 − α) 0 0
=: I1 + I2 .
T T
p
I2 = e−pt (t − s)−α dt u(s)ds
(1 − α) 0 s
T T −s
p −pη −α
= e η dη e−ps u(s)ds.
(1 − α) 0 0
2.7 Laplace Transform of ∂tα 45
for all s > 0 and T > 0, and the Lebesgue convergence theorem yields
∞ ∞
p −pη −α
lim I2 = e η dη e−ps u(s)ds
T →∞ (1 − α) 0 0
p (1 − α) ∞ −ps
= e u(s)ds = pα (Lu)(p)
(1 − α) p1−α 0
3.1 Examples
and
It is proved that there exists a unique solution to (3.3) and (3.4) respectively, and the
solutions are given by the following formulae (Kilbas et al. [15], Podlubny [24]).
For (3.3)
t
u(t) = at α−1
Eα,α (−λt ) +
α
(t − s)α−1 Eα,α (−λ(t − s)α )f (s)ds, 0 < t < T.
0
(3.5)
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 47
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5_3
48 3 Fractional Ordinary Differential Equations
For (3.4)
t
u(t) = aEα,1 (−λt ) + α
(t − s)α−1 Eα,α (−λ(t − s)α )f (s)ds, 0 < t < T.
0
(3.6)
Here we recall that Eα,β (z) with α, β > 0 are defined by (2.28).
Here we should expect that u(t) given by (3.5) and (3.6) are verified to
pointwisely satisfy (3.3) and (3.4) respectively. Moreover, in general, for f ∈
L2 (0, T ), we do not know whether the system
for u ∈ C 1 [0, T ]. An inequality of the type (3.7) is useful for proving the uniqueness
of solutions. Thanks to (3.7), we can easily verify that if u ∈ H 1 (0, T ) satisfies
du
(t) = λu, 0 < t < T, u(0) = 0,
dt
3.2 Fundamental Inequalities: Coercivity 49
then u = 0 in (0, T ). Indeed, multiplying this equation by u and applying (3.7) and
the Gronwall inequality, we obtain u(t) = 0 for 0 < t < T .
In this section, we discuss inequalities corresponding to (3.7) within the Sobolev
space Hα (0, T ).
We prove:
Theorem 3.1
(i)
T 1 1
u(t)dtα u(t)dt ≥ T −α u 2
− T 1−α |u(0)|2
0 2(1 − α) L2 (0,T ) 2(2 − α)
(3.8)
for all u ∈ W 1,1 (0, T ).
(ii)
T 1
u(t)∂tα u(t)dt ≥ T −α u 2
(3.9)
0 2(1 − α) L2 (0,T )
W 1,1 (0,T )
C 1 [0, T ] = W 1,1 (0, T ). (3.12)
Although this can be proved by a standard method based on the mollifier, we will
give the proof at the end of Theorem 3.2.
50 3 Fractional Ordinary Differential Equations
T −α T 1−α
≥ un 2
− |un (0)|2 , n ∈ N. (3.13)
2(1 − α) L2 (0,T ) 2(2 − α)
By Lemma 1.3 (ii) we have dtα un −→ dtα u in L1 (0, T ). By the Sobolev embedding
W 1,1 (0, T ) ⊂ C[0, T ], the convergence un −→ u in W 1,1 (0, T ) implies un −→ u
in C[0, T ] and un (0) −→ u(0). Therefore, letting n → ∞ in (3.13), we obtain (3.8)
for each u ∈ W 1,1 (0, T ).
Now we prove (3.8) for u ∈ C 1 [0, T ]. Henceforth for simplicity, we set
1
g(t) = t −α , 0 < α < 1, t > 0.
(1 − α)
Then
t du
dtα u(t) = g(t − s) (s)ds.
0 ds
We have
t du
u(t)dtα u(t) = u(t) g(t − s) (s)ds (3.14)
0 ds
t t
du du
=− g(t − s) (s)(u(s) − u(t))ds + g(t − s)u(s) (s)ds
0 ds 0 ds
=: S1 + S2 .
dg
By integration by parts and ds (s) ≤ 0 for 0 < s < T , we see
t
d
S1 = − g(t − s) (u(s) − u(t)) (u(s) − u(t))ds (3.15)
0 ds
t
1 d
=− |u(s) − u(t)|2 ds
g(t − s)
2
0 ds
1 s=t 1 t dg
=− g(t − s)|u(s) − u(t)| 2
− (t − s)|u(s) − u(t)|2 ds
2 s=0 2 0 dξ
1 s=t 1
≥− g(t − s)|u(s) − u(t)|2 = g(t)|u(t) − u(0)|2 ≥ 0.
2 s=0 2
3.2 Fundamental Inequalities: Coercivity 51
T −α
Hence, since g(T − s) ≥ (1−α) for 0 < s < T , we obtain
T T T T
u(t)dtα u(t)dt = S1 dt + S2 dt ≥ S2 (t)dt
0 0 0 0
T T
1 1
≥ g(s)|u(T − s)| ds − 2
g(t)dt |u(0)|2
2 0 2 0
1 T 1 T 1−α |u(0)|2
= g(T − s)|u(s)|2 ds −
2 0 2 (1 − α)(1 − α)
1 T −α 2 1 T 1−α |u(0)|2
≥ u −
2 (1 − α) L2 (0,T ) 2 (2 − α)
for u ∈ C 1 [0, T ]. Thus the proof of (3.8) for u ∈ C 1 [0, T ] is completed, and the
proof of (i) is completed.
Hα (0,T )
Proof of Theorem 3.1 (ii) Since 0 C 1 [0, T ] = Hα (0, T ) by Lemma 2.2, it
suffices to prove (3.9) for u ∈ 0 C 1 [0, T ]. Indeed, for each u ∈ Hα (0, T ), there
exists a sequence un ∈ 0 C 1 [0, T ], n ∈ N such that un −→ u in Hα (0, T ). Then
∂tα un −→ ∂tα u in L2 (0, T ) by Theorem 2.2. Therefore by the passage to the limit,
we obtain (3.9) for each u ∈ Hα (0, T ).
The proof of (3.9) for u ∈ 0 C 1 [0, T ] is done as follows. For u ∈ 0 C 1 [0, T ],
equalities (2.27) and Lemma 1.3 (i) yield
∂tα u = dtα u.
Thus we can repeat the proof of (i) to prove (3.9) for u ∈ 0 C 1 [0, T ]. Thus the proof
of Theorem 3.1 is completed.
52 3 Fractional Ordinary Differential Equations
Proof of Theorem 3.2 (i) First let u ∈ C 1 [0, T ]. Then (3.14) and (3.15) imply
t
1 1 d
u(s)dsα u(s) := S1 + S2 ≥ S2 = (t − s)−α (|u(s)|2 )ds
2 (1 − α) 0 ds
1
= dsα (|u(s)|2 ), 0 < s < T.
2
Hence, since dsα c = 0 for a constant c, we see that
1 α 1
u(s)dsα u(s) ≥ d (|u(s)|2 ) = dsα (|u(s)|2 − |u(0)|2), 0<s<T
2 s 2
d 1−α
dsα (|u(s)|2 − |u(0)|2 ) = J (|u(s)|2 − |u(0)|2 ) = J −α (|u(s)|2 − |u(0)|2 ).
dt
Therefore
1 −α
u(s)dsα u(s) ≥ J (|u(s)|2 − |u(0)|2 ), 0 < s < T , u ∈ C 1 [0, T ]. (3.17)
2
Since
J αu ≥ J αv in (0, T ) if u ≥ v in (0, T ),
which can be verified by (t − s)−α > 0 for 0 < s < t, by (3.17) we have
1 α −α
J α (udtα u)(t) ≥ J (J (|u|2 − |u(0)|2))(t)
2
1
= (|u(t)|2 − |u(0)|2), 0 < t < T,
2
and so
t
1 1
(t − s)α−1 (dsα u(s))u(s)ds ≥ (|u(t)|2 − |u(0)|2 ), 0<t <T
(α) 0 2
for u ∈ C 1 [0, T ].
Next, let u ∈ W 1,1 (0, T ). Then by (3.12) we can choose uk ∈ C 1 [0, T ], k ∈ N
such that uk −→ u in W 1,1 (0, T ). Then dtα uk −→ dtα u in L1 (0, T ) by Lemma 1.3.
Moreover,
t
1 1
(t − s)α−1 uk (s)dsα uk (s)ds ≥ (|uk (t)|2 − |uk (0)|2 ), k ∈ N. (3.18)
(α) 0 2
3.2 Fundamental Inequalities: Coercivity 53
for u ∈ L1loc (R). Then defining vε by (3.19) for v ∈ W 1,1 (R), we see (e.g., Adams
[2]) that vε ∈ C0∞ (R) and vε −→ v in W 1,1 (R). We set
∞
uε (t) = χε (t − s)
u(s)ds.
−∞
54 3 Fractional Ordinary Differential Equations
H −1 () = (H10 ()) . Then H01 () ⊂ L2 () ⊂ H −1 () algebraically and
topologically (e.g., Brezis [5]). By H −1 () ·, ·H 1 () we denote the duality pairing.
0
Theorem 3.3
(i)
T T −α T 1−α
(dtα u(·, t ), u(·, t ))L2() dt ≥ u 2
L2 (0,T ;L2 ())
− u(·, 0) 2
0 2(1 − α) 2(2 − α) L2 ()
In Chap. 2, we define ∂tα in Hα (0, T ). For example, the domain D(∂tα ) requires that
the element vanishes at t = 0 if 12 < α < 1. Thus in (3.4), we have to understand
that u ∈ D(∂tα ), and u(0) = 0 in the sense of the trace if 12 < α < 1. In order to
attach non-zero value to u(0) in some sense not only for 12 < α < 1 but also for
0 < α ≤ 12 , we interpret u(0) = a as u − a ∈ Hα (0, T ). Thus in this section, we
discuss an initial value problem for a linear fractional ordinary differential equation:
We note that if 12 < α < 1, then u − a ∈ Hα (0, T ) yields u(0) = a, which can
justify the initial condition in the pointwise sense. In view of Proposition 2.5 (i), we
can rewrite (3.20) as in the following lemma.
Lemma 3.1 In the sense of (2.37), the system (3.20) is equivalent to
−α
∂tα u = −λu + (1−α)
t
a + f (t), 0 < t < T,
(3.21)
u ∈ H (0, T ), if 0 < α < 12 ,
α
and
Proof Let 0 < α < 12 . Then H α (0, T ) = Hα (0, T ) and u, a ∈ Hα (0, T ). Therefore
t −α
∂tα (u − a) = ∂tα u − ∂tα a = ∂tα u − a.
(1 − α)
Hence the equivalence between (3.20) and (3.21) is verified. Next, let 12 < α < 1.
Assume that u satisfies (3.20). By the Sobolev embedding, we see that u ∈ C[0, T ],
and u(0) = a. Consequently,
by the definition (2.37) of ∂tα , which verifies that a solution to (3.20) satisfies (3.22).
On the other hand, assume that u satisfies (3.22). We see that u(0) = a implies that
u−a ∈ Hα (0, T ). Therefore the definition of ∂tα again means that ∂tα u = ∂tα (u−a).
Thus the proof of the lemma is completed.
For α = 12 , we cannot have the corresponding problem like (3.21) and the
formulation (3.21) is not convenient. Thus we mainly consider (3.20) as the initial
value problem.
We prove:
Theorem 3.5 Let λ, a ∈ R be given. There exists a unique solution u to (3.20).
Moreover, for 0 < α < 1, we can choose a constant C > 0 such that
and
In the pointwise sense, the unique existence of solutions to initial value problems
for fractional ordinary differential equations with Dtα and ∂tα has been well studied
(e.g., Kilbas et al. [15], Podlubny [24]), but such pointwise formulations meet
difficulty in several cases such as f ∈ L∞ (0, T ). Thus we need to formulate the
initial condition by (3.20).
Proof of Theorem 3.5 By Theorem 2.4, we can rewrite (3.20) as
J −α (u − a) = λu + f (t), u − a ∈ Hα (0, T ),
which is equivalent to
Then
t
|u(t)| ≤ C (t − s)α−1 |u(s)|ds, 0 < t < T.
0
We set
t
R(t) = |a| + C (t − s)α−1 |f (s)|ds.
0
for 0 ≤ t ≤ T . We take the norms in L2 (0, T ). The Young inequality Lemma A.1
in the Appendix yields
t
T 1−α
(t − s)α−1 |f (s)|ds ≤ t −α f ≤ f
L1 (0,T ) L2 (0,T ) L2 (0,T ) .
0 L2 (0,T ) 1−α
58 3 Fractional Ordinary Differential Equations
Moreover,
t s t t
(s − ξ )α−1 |f (ξ )|dξ ds = (s − ξ )α−1 ds |f (ξ )|dξ
0 0 0 ξ
t (t − ξ )α
= |f (ξ )|dξ,
0 α
and so
t s α
t
(s − ξ ) α−1
|f (ξ )|dξ ds ≤ f
α 1 L2 (0,T )
0 0 L2 (0,T ) L (0,T )
again by the Young inequality. By exchanging the orders of the integrals, we can
similarly obtain
t s
(t − s) α−1
(s − ξ ) α−1
|f (ξ )|dξ ds
0 0
t t (α)2 t
= |f (ξ )| (t − s) α−1
(s − ξ ) α−1
ds dξ = (t − ξ )2α−1 |f (ξ )|dξ,
0 ξ (2α) 0
and
t s
(t − s)α−1 (s − ξ ) |f (ξ )|dξ ds
α−1
0 0 L2 (0,T )
t
(α)2
≤ (t − ξ ) 2α−1
|f (ξ )|dξ
≤C f L2 (0,T ) .
(2α) 0 L2 (0,T )
We recall (1.17):
du
W (0, T ) := u ∈ L (0, T );
1,1 1
∈ L1 (0, T )
dt
with u W 1,1 (0,T ) = u L1 (0,T ) + du
dt L1 (0,T ) .
Here and henceforth we assume p ∈ L∞ (0, T ) and f ∈ L2 (0, T ).
In (3.28), if u ∈ W 1,1 (0, T ), then the initial condition u(0) = a can be
immediately justified by W 1,1 (0, T ) ⊂ C[0, T ]. Moreover, by u ∈ W 1,1 (0, T ),
we can prove that dtα u ∈ L1 (0, T ) exists. In many monographs (e.g., Diethelm [6],
Kilbas et al. [15], Podlubny [24]), the initial value problem is formulated by (3.28).
Now we consider the formulation:
∂tα (u − a) = p(t)u + f (t), 0 < t < T,
(3.29)
u − a ∈ Hα (0, T ),
Therefore (3.29) implies dtα u(t) = p(t)u + f (t) for 0 < t < T . Since u ∈
W 1,1 (0, T ) ⊂ C[0, T ], we have the initial condition u(0) = a in the sense of
limt →0 u(t) = a. Thus u satisfies (3.28) and the proof of (ii) is complete.
Now we discuss a simple case (3.4) and clarify in which sense the solution
formula (3.6) should be understood.
In fact, we prove:
Proposition 3.2 Let f ∈ L2 (0, T ). Then u given by (3.6) satisfies
The existing references [15, 24] give a representation formula (3.6) for the
solution to (3.4), but it can be justified pointwise only if f has certain regularity
such as f ∈ W 1,1 (0, T ). In other words, for f ∈ L2 (0, T ) it is more consistent
for us to interpret (3.6) as solution formula for the initial value problem (3.31), not
(3.4).
Proof We set
t
u1 (t) = aEα,1 (−λt ), α
u2 (t) = u2 (f )(t) = (t − s)α−1 Eα,α (−λ(t − s)α )f (s)ds.
0
Hence
t C α
|u2 (f )(t)| ≤ C s α−1 ds f C[0,T ] = t f C[0,T ] .
0 α
Therefore
t
du2 (f ) C α
(t) ≤C s α−1 ds f = t f
dt C 1 [0,T ] C 1 [0,T ] ,
0 α
On the other hand, using the power series of Eα,α (−ληα ) for η ≥ 0, we can directly
verify
t
1
(t − η)−α ηα−1 Eα,α (−ληα )dη = Eα,1 (−λt α ), t > 0,
(1 − α) 0
62 3 Fractional Ordinary Differential Equations
which means
t
1
(t − s)−α (s − ξ )α−1 Eα,α (−λ(s − ξ )α )ds
(1 − α) ξ
t −ξ
1
= ((t − ξ ) − η)−α ηα−1 Eα,α (−ληα )dη = Eα,1 (−λ(t − ξ )α ).
(1 − α) 0
Consequently,
t
d
∂tα u2 (f )(t) = Eα,1 (−λ(t − ξ )α )f (ξ )dξ.
dt 0
Since
d
Eα,1 (−λ(t − ξ )α ) = −λ(t − ξ )α−1 Eα,α (−λ(t − ξ )α )
dt
that is,
for f ∈ C0∞ (0, T ). Finally let f ∈ L2 (0, T ) be arbitrary. We choose fn ∈ C0∞ (0, T )
such that fn −→ f in L2 (0, T ) as n → ∞. Then we already proved that
∂tα u2 (fn ) = −λu2 (fn ) + fn (t) for 0 < t < T . Again Proposition 2.3 implies that
u2 (fn ) −→ u2 (f ) in Hα (0, T ) and ∂tα u2 (fn ) −→ ∂tα u2 (f ) in L2 (0, T ). Hence,
letting n → ∞, we obtain ∂tα u2 (f ) = −λu2 (f ) + f (t) also for f ∈ L2 (0, T ).
Thus the proof of Proposition 3.2 is completed.
uN (t) fN (t) aN
and let P (t) = (pij (t))1≤i,j ≤N with pij ∈ L∞ (0, T ) for 1 ≤ i, j ≤ N be an N × N
matrix.
3.5 Systems of Linear Fractional Ordinary Differential Equations 63
(iii) We further assume that F ∈ (W 1,1 (0, T ))N and P ∈ (C 2 [0, T ])N×N . Then
u ∈ (W 1,1 (0, T ))N and u(0) = a.
(iv) Let F = 0 and P ∈ (C 2 [0, T ])N×N . Then u − a ∈ (Wα (0, T ))N .
(v) For F ∈ (C 1 [0, T ])N and P ∈ (C 2 [0, T ])N×N , we have u−a ∈ (Wα (0, T ))N .
The proof of (v) is the same as (iv) and we omit.
Theorem 3.6, in particular parts (iii) and (v) will play an important role in proving
the unique existence of solutions to initial boundary value problems in Chap. 4.
When we can assume stronger regularity than P ∈ (C 2 [0, T ])N×N , we can simplify
the proof but we omit details.
Moreover, we recall that
dv
Wα (0, T ) = v ∈ W 1,1 (0, T ); t 1−α ∈ L∞ (0, T ), v(0) = 0 .
dt
J −α (u − a) = P (t)u(t) + F (t),
(3.34)
u − a ∈ (Hα (0, T ))N ,
and so
By Theorem 2.1 and P ∈ (L∞ (0, T ))N×N , using the compact embedding
(Hα (0, T ))N −→ (L2 (0, T ))N , we verify that u −→ J α (P u) is a compact
operator from (L2 (0, T ))N to (L2 (0, T ))N . Therefore in view of the Fredholm
64 3 Fractional Ordinary Differential Equations
t
|u(t)| ≤ C (t − s)−α P (L∞ (0,T ))N×N |u(s)|ds, 0 < t < T.
0
where we set
t
R(t) = |a| + C (t − s)α−1 |F (s)|ds.
0
Thus, arguing in the same way as in Theorem 3.5, we can complete the proof of
Theorem 3.6 (ii).
Proof of (iii) We set
Here H m+γ (0, T ) is a fractional Sobolev space (e.g., Adams [2]), and the norm in
H m+γ (0, T ) is defined by
m 2
d v
v H m+γ (0,T ) = v 2
L2 (0,T )
+
dt m 2
L (0,T )
m 1
T T d v d m v 2 1 2
+ (t) − (s) dtds . (3.39)
dt m dt m |t − s|1+2γ
0 0
Furthermore,
⎧
⎪
⎨ v H m+γ (0,T ) , 0 < γ < 1, γ = 2 ,
1
1
v Hm+γ (0,T ) = T 1 d m v 2 2 (3.40)
⎪
⎩ v 1 +m + 0 t dt m (t) dt , γ = 12 .
2 H (0,T )
J α (W 1,1 (0, T ))N ⊂ (W 1,1 (0, T ))N and Lα (W 1,1 (0, T ))N ⊂ (W 1,1 (0, T ))N
u − a = Lα (u − a) + G in (0, T ). (3.41)
By Lemma 1.3 (i) and P a, F ∈ (W 1,1 (0, T ))N , we can easily verify
As is already proved in parts (i) and (ii), we see that u − a ∈ (Hα (0, T ))N .
Therefore, setting v1 = Lα (u − a) and w1 = G, Eq. (3.41) and Lemma 3.3 imply
u − a = v1 + w1 with v1 ∈ (H2α (0, T ))N and w1 ∈ (W 1,1 (0, T ))N . Substituting
this into (3.41), we obtain
u − a = Lα v1 + (Lα w1 + G).
By Lemmata 3.2 and 3.3, we see that Lα v1 ∈ (H3α (0, T ))N and Lα w1 ∈
(W 1,1 (0, T ))N . Setting v2 = Lα v1 and w2 = Lα w1 + G, we reach u − a = v2 + w2
with v2 ∈ (H3α (0, T ))N and w2 ∈ (W 1,1 (0, T ))N .
66 3 Fractional Ordinary Differential Equations
u−a =
v + v ∈ (Hk0 α (0, T ))N ⊂ (W 1,1 (0, T ))N and w
w with ∈ (W 1,1 (0, T ))N .
and
t
1
|La (u − a)(t)| = (t − s) (P (s)(u(s) − a))ds
α−1
(α) 0
in view of u − a ∈ (W 1,1 (0, T ))N ⊂ (L∞ (0, T ))N by the Sobolev embedding. We
can directly verify that |G(t)| ≤ Ct α . Therefore limt →0 |u(t) − a| = 0 by (3.43).
Thus the proof of (iii) is complete.
Proof of (iv) First we prove:
Lemma 3.4 Lα h ∈ (Wα (0, T ))N for h ∈ (Wα (0, T ))N .
Proof of Lemma 3.4 Since h ∈ (Wα (0, T ))N , we have h(0) = 0, h ∈
(L∞ (0, T ))N and Lα h(0) = 0. Therefore Lemma 1.3 (i) yields
d d d
(Lα h)(t) = J α (P (t)h(t)) = J α (P (t)h(t))
dt dt dt
dP (t) dh
=Jα h(t) + J α P (t) (t) .
dt dt
and so
t
α dh dh
J
P (t) (t) ≤ C (t − s) α−1
dt P (t) dt (s) ds
0
t
≤ CCh (t − s)α−1 s α−1 ds ≤ Ch t 2α−1 ≤ C1 t α−1 .
0
Therefore
d
(Lα h(t)) ≤ C2 t α−1 ,
dt
which means that Lα h ∈ (Wα (0, T ))N . Thus the proof of Lemma 3.4 is
complete.
By F = 0 we have G(t) = J α (P (t)a), 0 < t < T . In terms of P ∈
(C 1 [0, T ])N×N
and Lemma 1.3 (i), we see that G(0) = 0 and
d d dP (t) P (0)a α−1
G(t) = J α (P (t)a) = J α a + t .
dt dt dt (α)
By (3.44), we obtain J α dP
dt a ∈ (L∞ (0, T ))N and so dG 1−α
dt t ∈ (L∞ (0, T ))N .
Consequently,
3 3
k1 α > , (k1 − 1)α ≤ . (3.47)
2 2
Similarly to the proof of (iii), in terms of (3.41), we will improve the regularity of
u − a. First by Lemma 3.3, u − a ∈ (Hα (0, T ))N yields Lα (u − a) ∈ (H2α (0, T ))N .
Consequently, setting v1 = Lα (u − a) and w1 = G, by (3.41) we see u − a =
v1 +w1 with v1 ∈ (H2α (0, T ))N and w1 ∈ (Wα (0, T ))N . Therefore, continuing this
argument, in view of Lemmata 3.3 and 3.4, we obtain u − a = vk + wk for k ∈ N,
68 3 Fractional Ordinary Differential Equations
where vk ∈ (H(k+1)α (0, T ))N and wk ∈ (Wα (0, T ))N provided that (k − 1)α ≤ 2.
Hence
+W
u−a =V , ∈ (Hk1 α (0, T ))N ,
V ∈ (Wα (0, T ))N .
W
dV
∈ (Hk1 α−1 (0, T ))N ⊂ (L∞ (0, T ))N ,
dt
∈ (Wα (0, T ))N . Thus u − a ∈ (Wα (0, T ))N , and the proof of
which means that V
Theorem 3.6 is complete.
Now we prove Lemmata 3.2 and 3.3.
Proof of Lemma 3.2 The first inclusion was proved as Lemma 1.3 (i).
By P ∈ (C 1 [0, T ])N×N , we see that P (W 1,1 (0, T ))N ) ⊂ (W 1,1 (0, T ))N and so
Thus Lα (W 1,1 (0, T ))N ⊂ (W 1,1 (0, T ))N follows and the proof of Lemma 3.2 is
complete.
Proof of Lemma 3.3 In terms of (2.4), (3.39) and (3.40), using P ∈ (C 2 [0, T ])N×N ,
we obtain
d α dv
J v = Jα
dt dt
3.6 Linear Fractional Ordinary Differential Equations with Multi-Term. . . 69
d α
J v ∈ J α (Hδ (0, T )).
dt
If α + δ ≤ 1, then Theorem 2.3 (i) yields J α (Hδ (0, T )) ⊂ Hα+δ (0, T ). By the
definition of Hα+δ+1 (0, T ), we see
Thus (3.50) holds for all (α, β) satisfying 0 < β ≤ 2 and 0 < α < 1, and so the
proof of (3.49) is complete. Thus the proof of Lemma 3.3 is finished.
We have discussed the unique existence of the solutions to initial value problems for
fractional ordinary differential equations on the basis of ∂tα in the Sobolev spaces
Hα (0, T ), and the method is widely applicable for example, to nonlinear equations.
Rather than comprehensive discussions, here we are restricted to linear fractional
ordinary differential equations with multi-term time fractional derivatives:
m
α
rj (t)∂t j (u − a)(t) = p(t)u(t) + f (t), 0<t <T (3.51)
j =1
with
Here
and
and
1
u H α1 (0,T ) ≤ C(|a| + f L2 (0,T ) ), if 0 < α < 1 and α = . (3.57)
2
Proof We can rewrite (3.51) as
m
J −α1 (u − a)(t) + (t)u(t) + f(t),
rj (t)J −αj (u − a)(t) = p 0 < t < T.
j =2
(3.58)
Here and henceforth we set
rj (t) p(t) f (t)
rj (t) = , j = 2, . . . , m, (t) =
p , f(t) = , 0 ≤ t ≤ T.
r1 (t) r1 (t) r1 (t)
m
w(t) = − rj (t)J α1 −αj w(t)+ p (t)a+f(t),
p J α1 w(t)+ 0<t <T (3.59)
j =2
and
w ∈ L2 (0, T ). (3.60)
m
v(t) = − rj (t)J α1 −αj v(t) + p
J α1 v(t), 0 < t < T, (3.61)
j =2
implies v = 0 in (0, T ), then the Fredholm alternative yields the unique existence
of w satisfying (3.59) for any a ∈ R and f ∈ L2 (0, T ). Let (3.61) hold. Then
m
t t
|v(t)| ≤ C (t − s)α1 −αj −1 |v(s)|ds + C (t − s)α1 −1 |v(s)|ds
j =2 0 0
(t − s)α1 −αj −1 = (t − s)α1 −α2 +(α2 −αj )−1 ≤ T α2 −αj (t − s)(α1 −α2 )−1
and
and so
t
|v(t)| ≤ C (t − s)α1 −α2 −1 |v(s)|ds, 0 < t < T.
0
Noting that α1 −α2 > 0, we apply the generalized Gronwall inequality (Lemma A.2)
to verify that v = 0 in (0, T ). Thus the proof of the unique existence is complete.
The estimates (3.55)–(3.57) are proved similarly to Theorem 3.5.
Chapter 4
Initial Boundary Value Problems for
Time-Fractional Diffusion Equations
∂ ∂2 ∂
∂i = , ∂i2 = , i = 1, . . . , n ∇ = (∂1 , . . . , ∂n ), ∂s = .
∂xi ∂xi2 ∂s
Let −A(t) be a uniform elliptic differential operator of the second order with
(x, t)-dependent coefficients:
n
n
(−A(t)u)(x, t) = ∂i (aij (x, t)∂j u(x, t)) + bj (x, t)∂j u(x, t) + c(x, t)u(x, t),
i,j =1 j =1
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 73
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5_4
74 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
for x ∈ and t > 0, where aij = aj i , bj , c ∈ C 2 ([0, T ]; C 1 ()), and there exists
a constant μ0 > 0 such that
n
n
aij (x, t)ξi ξj ≥ μ0 ξj2 , x ∈ , 0 ≤ t < ∞, (ξ1 , . . . , ξn ) ∈ Rn .
i,j =1 j =1
(4.2)
Remark 4.1 We can relax the regularity conditions of the coefficients similarly to
[17, 30], by approximating them by smooth functions. However, we omit the details
for simplicity.
The main purpose of this chapter is to formulate the initial boundary value
problem for t-dependent A(t) with initial value a and non-homogeneous term F in
L2 -spaces. We refer to Kubica and Yamamoto [17], Zacher [30] for such treatments,
and here we describe a unified approach within the framework by means of ∂tα in
the Sobolev space Hα (0, T ).
In particular, for more regular F , for example F ∈ L∞ (0, T ; L2 ()), there are
several works on the well-posedness for fractional partial differential equations and
we can refer to Gorenflo et al. [11], Li et al. [18], Luchko [21], Sakamoto and
Yamamoto [26]. Here we do not intend any complete list of the references.
We mainly assume that a and F are in some L2 -spaces. Then in general we
cannot prove that u ∈ C([0, T ]; L2 ()). Therefore, similarly to the case of
fractional ordinary differential equations, we must be careful in interpreting the
initial condition u(·, 0) = a in (4.1), which cannot imply that u(·, t) −→ a in
L2 () as t → 0.
Henceforth for H01 () ⊂ L2 (), identifying the dual of L2 () of L2 ()
with itself, we define H −1 () = (H10 ()) . Then H01 () ⊂ L2 () ⊂ H −1 ()
algebraically and topologically (e.g., Brezis [5]).
In terms of the definition ∂tα in Hα (0, T ) defined in Chap. 2, we formulate an
initial boundary value problem as follows:
Remark 4.2 Moreover, in view of (2.38), interpreting ∂tα u by (2.37), we note that
(4.3) is equivalent to
⎧
t −α
⎪
⎪ ∂t u + A(t)u = F +
⎪
α
(1−α) a, in H −1 (), 0 < t < T ,
⎨
if 0 < α < 12 ,
(4.6)
⎪
⎪ ∂tα u + A(t)u = F, −1
in H (), 0 < t < T ,
⎪
⎩
if 12 < α < 1.
In Zacher [30], and Kubica and Yamamoto [17] (Theorem 1.1), with the same
regularity conditions on a and F , the unique existence of the solution u to (4.3)–
(4.5) is proved with the regularity
if and only if
Indeed, the estimate (4.8) is directly seen by (4.7), because Hα (0, T ) ⊂ H α (0, T ) ⊂
C[0, T ] if 12 < α < 1 by the Sobolev embedding.
Remark 4.3 By (2.40), we can rewrite (4.7) as
1
≤ C( a L2 () + F L2 (0,T ;H −1 ()) ), α = . (4.9)
2
Here ∂tα is defined by (2.37).
Proof of (4.9) For 0 < α < 12 , by (2.38) we have
t −α
∂tα (u − a) = ∂tα u − ∂tα a = ∂tα u − a.
(1 − α)
Therefore, since t −α L2 (0,T ) < ∞ for 0 < α < 12 , inequality (4.7) yields
t −α
∂tα u ≤ ∂tα (u − a) +
a
L2 (0,T ;H −1 ()) L2 (0,T ;H −1 ())
(1 − α) L2 (0,T ;H −1 ())
≤ C( F L2 (0,T ;H −1 ()) + a L2 (0,T ;H −1 ()) ).
Next, let 1
2 < α < 1. Then (4.9) is seen by the definition (2.37).
Remark 4.4 In Sakamoto and Yamamoto [26], in a special case where A(t) is
symmetric (i.e., bj = c = 0, 1 ≤ j ≤ n) and the coefficients of A(t) are
independent of t and F = 0, then for (4.1) it is proved that
Here
pα − 2
κ= > 0.
2p
4.1 Main Results 77
In particular,
The proposition means that with more regular a and F , we can prove the
continuity of u(·, t) at t = 0 in L2 (), and so the initial condition (4.5) holds in
the usual sense.
Remark 4.5 For dtα u = div (p(x, t)∇u(x, t)) + F (x, t) in (4.1) with a = 0,
Jin, Li and Zhou (Theorem 2.1 in [12]) proves that u ∈ C([0, T ]; L2 ()) ∩
Lp (0, T ; H 2 ()) and dtα u ∈ Lp (0, T ; L2 ()) if F ∈ Lp (0, T ; L2 ()) with
p > α1 . In fact, we can interpret that p > α1 is a critical condition for the continuity
of u in t. Proposition 4.1 requires a stronger assumption p > α2 for p, but the weaker
spatial regularity is needed. The continuity of u at t = 0 should be exploited more
but here we omit details.
For more regular a and F , we can improve the regularity of u.
Theorem 4.2 We assume regularity aij ∈ C 2 ([0, T ]; C 1 ()), bj , c ∈
C 2 ([0, T ]; C 1 ()), 1 ≤ i, j ≤ n. For F ∈ L2 (0, T ; L2 ()) and a ∈ H01 (),
there exists a unique solution u ∈ L2 (0, T ; H 2() ∩ H01 ()) satisfying
u − a ∈ Hα (0, T ; L2 ()) to (4.3)–(4.5). Moreover, there exists a constant C > 0
such that
u−a Hα (0,T ;L2 ()) + u L2 (0,T ;H 2 ()) ≤ C( F L2 (0,T ;L2 ()) + a H01 () ).
Theorems 4.1 and 4.2 are corresponding results to the classical results for the
parabolic equation (i.e., α = 1) for which we refer to Evans [7], Lions and Magenes
[19], for example.
In the case of F = 0 in Theorem 4.2, we can further prove:
Proposition 4.2 We assume all the conditions in Theorem 4.2 and F = 0 and
Our classes of solutions in Theorems 4.1 and 4.2 are flexible for interpolated
regularity properties. Although we can choose a general uniform elliptic operator,
we introduce the Laplacian with the homogeneous Dirichlet boundary condition:
n
−A0 u(x) = ∂k2 u(x), D(A0 ) = H 2 () ∩ H01 ().
k=1
1
−1
Then it is known that D(A02 ) = H01 () and H −1 () = (H01 ()) = D(A0 2 ).
Moreover, we can verify that for 0 ≤ θ ≤ 1, the interpolation spaces are given by:
n
n
(Av)(x) = − ∂i (aij (x)∂j v(x)) − bj (x)∂j v(x) − c(x)v(x), x ∈ ,
i,j =1 j =1
where aij , bj , c ∈ C 2 () and (4.2) is satisfied. A convenient way for constructing
a solution to (4.3)–(4.5) is by the Laplace transform (see Sect. 2.7 of Chap. 2),
which relies on formulae of Laplace transforms of time-fractional derivatives. For
the rigorous treatments, we have to specify the class of solutions u admitting such
formulae for the Laplace transforms, but it is not often clarified in view of the
consistency with the expected regularity of solutions to (4.3)–(4.5).
4.2 Some Results from Theorem 4.1 and Proposition 4.2 79
Within our framework, we can apply Theorem 2.7 of Chap. 2 concerning the
Laplace transform of ∂tα in Vα (0, ∞). Here we sketch such treatments. As related
arguments, see Kian and Yamamoto [14].
For arbitrary T > 0, let u ∈ L2 (0, T ; H 2() ∩ H01 ()) satisfy u − a ∈
Hα (0, T ; L2 ()) and (4.3)–(4.5) with a ∈ H01 () and F = 0. In particular,
∂tα (u − a) + Au = 0 in × (0, T ).
Taking the scalar products in L2 () of both sides with arbitrarily fixed ϕ ∈ C0∞ ()
and integrating by parts, we have
(∂tα (u − a)(·, t), ϕ)L2 () + (Au(·, t), ϕ)L2 () = 0, t > 0.
Then
∂tα {((u − a)(·, t), ϕ)L2 () } + (u(·, t), A∗ ϕ)L2 () = 0, t > 0. (4.12)
Here
n
n
∗
(A v)(x) = − ∂i (aij (x)∂j v) + ∂j (bj (x)v(x)) − c(x)v(x), x ∈ .
i,j =1 j =1
By u − a ∈ Hα (0, T ; L2 ()), we see that ((u − a)(·, t), ϕ)L2 () ∈ Hα (0, T ).
Proposition 4.2 yields u(·, t) L2 () ≤ CeCt a H 1 () for all t ≥ 0. Hence
0
|((u − a)(·, t), ϕ)L2 () |, |(u(·, t), A∗ ϕ)L2 () | ≤ CeC1 t , t ≥ 0.
exists for p > C1 . We take the Laplace transforms of both sides of (4.12) and apply
Theorem 2.7, so that
pα (L(u − a)(·, p), ϕ)L2 () + ((Lu)(·, p), A∗ ϕ)L2 () = 0, p > C1 ,
that is,
in L2 (). Thus we can justify the method by the Laplace transform for an initial
boundary value problem (4.3)–(4.5).
The proof is based on what is called the Galerkin approximation (e.g., Evans [7],
Lions and Magenes [19]). The proof is composed of:
(i) Construction of approximating solutions in a family of finitely dimensional
subspaces: For this step, Theorem 3.6 (iii) plays an important role.
(ii) Uniform boundedness of the approximating solutions: The coercivity Theo-
rems 3.3 and 3.4 are essential.
(iii) Then we can prove the existence of solution as a weak convergent limit of a
subsequence of the sequence of the approximating solutions. The uniqueness
of the solution follows from the coercivity.
First we prove the existence of solutions.
Let 0 < λ1 ≤ λ2 ≤ · · · be the eigenvalues of − with the zero Dirichlet
boundary condition, which are numbered according to the multiplicities, that is, λm
appears -times if the multiplicity of λm is . By ρk ∈ H 2 (), k ∈ N, we denote an
eigenfunction of − for λk such that
1, k = ,
(ρk , ρ )L2 () = δk :=
0, k = .
Then
(ρk , ρ )L2 () = δk , H −1 () ρk , ρ H01 () = δk . (4.13)
Moreover,
⎧ ∞ ∞
⎪
⎪ a= k=1 (a, ρk )L2 () ρk in L2 (), a 2 = k=1 |(a, ρk )L2 () |
2,
⎪
⎪ L2 ()
⎪
⎪ there exists a constant C > 0 such that
⎪
⎪
⎪
⎨ a= ∞ −1 () and
k=1 H −1 () a, ρk H01 () ρk in H
⎪
⎪ C −1 ∞ −1
k=1 λk |H −1 () a, ρk H01 () | ≤ a H −1 ()
2 2
⎪
⎪
⎪
⎪
⎪ ≤C ∞ −1
k=1 λk |H −1 () a, ρk H01 () | ,
2 a ∈ H −1 (),
⎪
⎪
⎩ C −1 ∞ k=1 λk |(a, ρk )| ≤ a
2 2
1 ≤C ∞ k=1 λk |(a, ρk )| ,
2 a ∈ H01 ().
H0 ()
(4.14)
4.3 Proof of Theorem 4.1 81
We note that
H −1 () ϕ, ψH01 () = (ϕ, ψ)L2 () for ϕ ∈ L2 () and ψ ∈ H01 ()
N
N
uεN (x, t) = ε
pN,k (t)ρk (x), FNε (x, t) = H −1 () F
ε
(·, t), ρk H 1 () ρk (x)
0
k=1 k=1
(4.16)
satisfying
(∂tα (uεN − aN ), ρ )L2 () + (A(t)uεN , ρ )L2 () = H −1 () FNε , ρ H 1 () ,
0
uεN − aN ∈ Hα (0, T ; L2 ()), 1 ≤ ≤ N, 0 ≤ t ≤ T ,
(4.17)
where we set
N
aN = ck ρk , ck = (a, ρk )L2 () . (4.18)
k=1
We rewrite (4.17) as
ε
∂tα (pN, − c ) = N k=1 k (t)pN,k (t) + f (t),
ε ε 0 < t < T,
(4.20)
ε
pN, − c ∈ Hα (0, T ), 1 ≤ ≤ N,
82 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
where
⎧
⎪
⎪ fε (t) = H −1 () F ε (·, t), ρ H 1 () ,
⎪
⎪ 0
⎪
⎨ k (t) = ni,j =1 (∂i (aij (·, t)∂j ρk ), ρ (·))L2 ()
+ ni=1 (bi (·, t)∂i ρk , ρ )L2 () + (c(·, t)ρk , ρ )L2 ()
⎪
⎪
⎪
⎪ = − ni,j =1 aij (x, t)(∂j ρk )(x)(∂i ρ )(x)dx
⎪
⎩ n
+ i=1 (bi (·, t)∂i ρk , ρ )L2 () + (c(·, t)ρk , ρ )L2 () .
By the regularity of F ε and aij , we know that k ∈ C 2 [0, T ] and fε ∈ W 1,1 (0, T ).
Consequently, we can apply Theorem 3.6 (iii) to see that there exists a unique
ε
solution pN, ∈ W 1,1 (0, T ) to (4.20) and pN,
ε
(0) = c for = 1, . . . , N.
Next we estimate uN Since pN,k ∈ W (0, T ) and pN,k
ε ε 1,1 ε
(0) = ck , we see that
pN,k − ck ∈ Hα (0, T ) ∩ 0 W (0, T ) and so (2.27) yields
ε 1,1
∂tα (pN,k
ε
− ck ) = dtα (pN,k
ε
− ck ) = dtα pN,k
ε
, N ∈ N, 1 ≤ k ≤ N. (4.21)
Therefore,
N
N
∂tα (uεN − aN ) = ∂tα (pN,k
ε
− ck )ρk = (dtα pN,k
ε
)ρk = dtα uεN . (4.22)
k=1 k=1
By (4.19) we have
n
− (t − s)α−1 (bj (·, s)∂j uεN (·, s), uεN (·, s))L2 ()
j =1
= H −1 () FNε (·, s), uεN (·, s)H 1 () (t − s)α−1 . (4.23)
0
By (4.2) we obtain
n
− (t − s)α−1 (∂i (aij (·, s)∂j uεN (·, s)), uεN (·, s))L2 ()
i,j =1
n
= (t − s)α−1 aij (x, s)(∂j uεN (x, s))∂i uεN (x, s))dx
i,j =1
Moreover, fixing small δ > 0, we can choose a constant Cδ > 0 such that we have
n n
− (b (·, s)∂ u ε
(·, s), u ε
(·, s))
2 () ≤ C |∂j uεN (x, s)||uεN (x, s)|dx
j j N N L
j =1 j =1
≤δ |∇uεN (x, s)|2 dx + Cδ |uεN (x, s)|2 dx. (4.25)
Similarly we see
|H −1 () FNε (·, s), uεN (·, s)H 1 () | ≤ δ ∇uεN (·, s) 2
L2 ()
+ Cδ FNε (·, s) 2
H −1 ()
.
0
(4.26)
On the other hand, by Theorem 3.4 (i) and uεN ∈ W 1,1 (0, T ; L2 ()), we obtain
t (α)
(t − s)α−1 (dtα uεN (·, s), uεN (·, s))L2 () ds ≥ ( uεN (·, t) 2
L2 ()
− aN 2
(L2 ())N
).
0 2
Fixing δ > 0 sufficiently small and absorbing the second term on the right-had side
into the left-hand side, we obtain
2
uεN (·, t) L2 ()
t
≤ uεN (·, t) 2
L2 ()
+ (t − s)α−1 ∇uεN (·, s) 2
L2 ()
ds
0
t
≤C aN 2L2 () +C (t − s)α−1 uεN (·, s) 2
L2 ()
ds
0
t
+C (t − s)α−1 FNε (·, s) 2
H −1 ()
ds. (4.27)
0
84 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
Here
t s
(t − s) α−1
(s − ξ ) α−1
FNε (·, ξ ) 2H −1 () dξ ds
0 0
t t
= FNε (·, ξ ) 2
H −1 ()
(t − s)α−1 (s − ξ )α−1 ds dξ
0 ξ
(α)2 t
= (t − ξ )2α−1 FNε (·, ξ ) 2
H −1 ()
dξ.
(2α) 0
Since
we have
t
uεN (·, t) 2L2 () ≤C aN 2L2 () +C (t − s)α−1 FNε (·, s) 2
H −1 ()
ds.
0
T
Taking 0 · · · dt and applying the Young inequality on the convolution (Lemma A.1
in the Appendix), we have
2 2 2
uεN (·, t) L2 (0,T ;L2 ())
≤ C( aN L2 ()
+ FNε L2 (0,T ;H −1 ())
). (4.28)
On the other hand, estimating the second term on the left-hand side and the right-
hand side of (4.19) by (4.29) and (4.25)–(4.26), we have
T T
uεN (·, t) 2L2 () dt + ∇uεN (·, t) 2
L2 ()
dt
0 0
T T
≤ Cδ ∇uεN (·, t) 2
L2 ()
dt + Cδ uεN (·, t) 2
L2 ()
dt
0 0
T
+ Cδ FNε (·, t) 2
H −1 ()
dt + C aN 2
L2 ()
.
0
Therefore, fixing δ > 0 small and absorbing the first term on the right-hand side
into the left-hand side, we reach
∇uεN 2
L2 (0,T ;L2 ())
≤ C( aN 2
L2 ()
+ uεN 2
L2 (0,T ;L2 ())
+ FNε 2
L2 (0,T ;H −1 ())
).
N
∂tα (uεN − aN ) = ∂tα (pN,k
ε
− ck )ρk ,
k=1
we have
N
H −1 () ∂tα (uεN − aN ), ρ H 1 () = ∂tα (pN,k
ε
(t) − ck )H −1 () ρk , ρ H 1 () = 0
0 0
k=1
N
for ≥ N + 1. For any ψ ∈ H01 (), we set ψN = =1 H −1 () ψ, ρ H01 () ρ .
Therefore (4.17) yields
Hence
∂tα (uεN − aN )(·, t ) H −1 () = sup |H −1 () ∂tα (uεN − aN )(·, t ), ψH 1 () |
0
ψ∈H01 (), ψ H01 ()
=1
∂tα (uεN − aN ) 2
L2 (0,T ;H −1 ())
+ uεN 2
L2 (0,T ;H01 ())
≤ C( a 2
L2 ()
+ Fε 2
L2 (0,T ;H −1 ())
). (4.31)
The sequences {uεN − aN }N∈N and {uεN }N∈N are bounded in H α (0, T ; H −1 ())
and in L2 (0, T ; H01()) respectively. Therefore we can extract a subsequence N
of N ∈ N and uε ∈ L2 (0, T ; H01 ()) and v ε ∈ Hα (0, T ; H −1 ()) such that
uεN −→ uε weakly in L2 (0, T ; H01 ()) and uεN − aN −→ v ε weakly in
Hα (0, T ; H −1 ()). Since aN −→ a strongly in L2 (), we have uεN −→ a + v ε
weakly in Hα (0, T ; H −1 ()). Therefore uεN −→ uε and uεN −→ a + v ε in the
sense of distribution, that is, in (C0∞ ( × (0, T ))) . Hence uε = a + v ε , that is,
v ε = uε − a. Therefore by (4.31) we have
uε − a 2
Hα (0,T ;H −1 ())
+ uε 2
L2 (0,T ;H01 ())
(4.32)
2 2
≤ lim
inf uεN − aN H α (0,T ;H −1 ())
+ uεN L2 (0,T ;H01 ())
N →∞
≤ C( F ε 2
L2 (0,T ;H −1 ())
+ a 2
L2 ()
).
4.3 Proof of Theorem 4.1 87
Since ψN ∈ XN and N ∈ N are chosen arbitrarily, it follows that (4.33) holds for
each ψ ∈ H01 (), so that
then u = 0 in × (0, T ).
From (4.34) and integration by parts in view of u(·, t) ∈ H01 (), we have
⎛ ⎞
n
H −1 () ∂t u(·, s), u(·, s)H01 ()
α
+⎝ aij (·, s)∂j u(·, s), ∂i u(·, s)⎠
i,j =1 L2 ()
n
= H −1 () bj (·, s)∂j u(·, s), u(·, s)H 1 () + H −1 () c(·, s)u, uH 1 () , 0 < s < T.
0 0
j =1
88 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
Therefore
t
H −1 () ∂t u(·, s), u(·, s)H01 () (t − s)α−1 ds
α
0
⎛ ⎞
t
n
+ (t − s)α−1 ⎝ aij (x, s)(∂j u(x, s))∂i u(x, s)dx ⎠ ds
0 i,j =1
⎛ ⎞
t
n
= (t − s)α−1 ⎝ bj (x, s)(∂j u(x, s))u(x, s)dx ⎠ ds
0 j =1
t
2
+ (t − s) α−1
c(x, s)u (x, s)dx ds.
0
By u ∈ Hα (0, T ; H −1 ()) ∩ L2 (0, T ; H01 ()), applying Theorem 3.4 (ii) to the
first term on the left-hand side and choosing δ > 0 sufficiently small, we have
t
(α) (α)
u(·, t) 2L2 () ≤ u(·, t) 2L2 () + (1 − Cδ) (t − s)α−1 ∇u(·, s) 2L2 () ds
2 2 0
t
≤ Cδ (t − s)α−1 u(·, s) 2L2 () ds, 0 < t < T .
0
For F ∈ L2 (0, T ; L2 ()) and ε > 0, let F ε ∈ C0∞ (0, T ; L2 ()) satisfy F ε −
F L2 (0,T ;L2 ()) < ε. In terms of Theorem 3.6 (iii) and (v), we can argue similarly
ε
to the proof of Theorem 4.1 to construct an approximating sequence, that is, pN,k ∈
4.4 Proof of Theorem 4.2 89
ε (0) = c := (a, ρ )
(Wα (0, T ))N and pN,k k k L2 () for 1 ≤ k ≤ N. Here we have to
estimate
N
uεN (x, t) := ε
pN,k (t)ρk (x) in L2 (0, T ; H 2())
k=1
and ∂tα (uεN − aN ) in L2 (0, T : L2 ()), where aN = N k=1 ck ρk .
ε
Using (4.21), multiplying the first equation in (4.20) by dtα pN, , summing over
= 1, . . . , N, we obtain
n
(dtα uεN (·, t), dtα uεN (·, t))L2 () − (∂i (aij (·, t)∂j uεN (·, t)), dtα uεN (·, t))L2 ()
i,j =1
n
− (bj (·, t)∂j uεN (·, t), dtα uεN (·, t))L2 () − (c(·, t)uεN (·, t), dtα uεN (·, t))L2 ()
j =1
The main part of the proof is the estimation of the second term on the left-hand
side. By uεN (·, t) ∈ H01 (), integration by parts yields
⎛ ⎞
n
−⎝ ∂i (aij (·, t)∂j uεN (·, t)), dtα uεN (·, t)⎠
i,j =1 L2 ()
⎛ ⎞
n
=⎝ aij (·, t)∂j uεN (·, t), dtα ∂i uεN (·, t)⎠ .
i,j =1 L2 ()
n
Now we consider ε α ε
i,j =1 aij (∂j uN )∂i dt uN (·, t). Since aij = aj i , 1 ≤ i, j ≤ n,
we have
n
aij (∂j uεN )∂i dtα uεN (·, t)
i,j =1
n
= aij ((∂j uεN )∂i dtα uεN + (∂i uεN )∂j dtα uεN ) + aii (∂i uεN )(∂i dtα uεN ) =: J1 + J2 .
i>j i=1
90 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
We set g(t) = 1 −α . Here we write uεN (s) = uεN (x, s), etc. Then
(1−α) t
J1 = aij ((∂j uεN )∂i dtα uεN + (∂i uεN )∂j dtα uεN )(x, t)
i>j
t
= aij (x, t) g(t − s)((∂j uεN )(t)∂i ∂s uεN (x, s) + (∂i uεN )(t)∂j ∂s uεN (x, s))ds
i>j 0
t
=− aij (x, t) g(t − s)((∂i ∂s uεN (s))(∂j uεN (s) − ∂j uεN (t))
i>j 0
=: J11 + J12 .
We have
t
J11 = − aij (x, t) g(t − s){∂s (∂i uεN (s) − ∂i uεN (t))(∂j uεN (s) − ∂j uεN (t))
i>j 0
Here we used
lim g(t − s)(∂i uεN (s) − ∂i uεN (t))(∂j uεN (s) − ∂j uεN (t)) = 0,
s↑t
which can be justified by ∂i uεN (x, ·), ∂j uεN (x, ·) ∈ Wα (0, T ) following from
Theorem 3.6 (v).
4.4 Proof of Theorem 4.2 91
Consequently,
1
J1 = aij g(t)(∂i uεN (0) − ∂i uεN (t))(∂j uεN (0) − ∂j uεN (t))
2
i=j
t
1 dg
+ − (t − s) aij (x, t)(∂i uεN (s) − ∂i uεN (t))(∂j uεN (s) − ∂j uεN (t))ds
2 0 dξ
i=j
1
+ aij dtα (∂i uεN ∂j uεN )(x, t).
2
i=j
t
=− aii g(t − s)∂i ∂s uεN (s)(∂i uεN (s) − ∂i uεN (t))ds
i 0
t
+ aii g(t − s)∂i ∂s uεN (s)∂i uεN (s)ds
i 0
t
1
=− aii g(t − s)∂s ((∂i uεN (s) − ∂i uεN (t))(∂i uεN (s) − ∂i uεN (t)))ds
2 0 i
1 t
+ aii g(t − s)∂s (|∂i uεN (s)|2 )ds
2 0
i
1 1 t dg
= aii g(t)(∂i uεN (t) − ∂i uεN (0))2 + − (t − s) aii (∂i uεN (s) − ∂i uεN (t))2 ds
2 2 0 dξ
i i
1
+ aii dtα (|∂i uεN (t)|2 ).
2
i
92 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
Thus
n
aij ∂j uεN (x, t)∂i dtα uεN (x, t)
i,j =1
1
n
= g(t)aij (x, t)(∂i uεN (0) − ∂i uεN (t))(∂j uεN (0) − ∂j uεN (t))
2
i,j =1
t
n
1 dg
+ − (t − s) aij (x, t)(∂i uεN (s) − ∂i uεN (t))(∂j uεN (s) − ∂j uεN (t))ds
2 0 dξ
i,j =1
1
n
+ aij dtα ((∂i uεN )∂j uεN )(x, t).
2
i,j =1
By g(t) > 0, − dg
dξ (t − s) > 0 for 0 < s < t < T and (4.2), we obtain
n
n
2 aij (x, t)∂j uεN (x, t)∂i dtα uεN (x, t) ≥ aij (x, t)dtα ((∂i uεN )∂j uεN )(x, t)
i,j =1 i,j =1
t
n
= g(t − s) ∂s (aij (x, s)∂i uεN (s)∂j uεN (s))ds
0 i,j =1
t
n
− g(t − s) (∂s aij (x, s))∂i uεN (s)∂j uεN (s)ds
0 i,j =1
t
n
+ g(t − s) (aij (x, t) − aij (x, s))∂s ((∂i uεN (s))∂j uεN (s))ds =: S1 + S2 + S3 .
0 i,j =1
(4.36)
Here we have
t
n
S1 = g(t − s) ∂s (aij (x, s)∂i uεN (s)∂j uεN (s))ds (4.37)
0 i,j =1
⎛ ⎞
t
n
= g(s)∂t ⎝ aij (x, t − s)∂i uεN (t − s)∂j uεN (t − s)⎠ ds
0 i,j =1
t
n
= ∂t g(s) aij (x, t − s)∂i uεN (t − s)∂j uεN (t − s)ds
0 i,j =1
n
− g(t) aij (x, 0)∂i uεN (x, 0)∂j uεN (x, 0).
i,j =1
4.4 Proof of Theorem 4.2 93
Furthermore
t
n
S2 = − g(t − s) (∂s aij (x, s))∂i uεN (s)∂j uεN (s)ds
0 i,j =1
and
t
n
S3 = rij (x, t, s)∂s (∂i uεN (s)∂j uεN (s))ds,
0 i,j =1
where
(t − s)−α
rij (x, t, s) = (aij (x, t) − aij (x, s)), 1 ≤ i, j ≤ n.
(1 − α)
We directly obtain
t
|S2 | ≤ C (t − s)−α |∇uεN (·, s)|2 ds. (4.38)
0
n
2 aij (x, t)∂j uεN (x, t)∂i dtα uεN (x, t)
i,j =1
t
n
≥ ∂t g(s) aij (x, t − s)∂i uεN (x, t − s)∂j uεN (x, t − s)ds
0 i,j =1
n
− g(t) aij (x, 0)∂i uεN (x, 0)∂j uεN (x, 0)
i,j =1
t
− C|∇uεN (x, 0)|2 −C (t − s)−α |∇uεN (x, s)|2 ds, x ∈ , 0 < t < T .
0
(4.39)
T
Taking 0 · · · dxdt and applying the Young inequality (Lemma A.1 in the
Appendix) to the fourth term on the right-hand side of (4.39), we have
T
n
2 (aij (·, t)∂j uεN (·, t), dtα ∂i uεN (·, t))L2 () dt
0 i,j =1
T
n
≥ g(s) (aij (T − s)∂i uεN (T − s), ∂j uεN (T − s))L2 () ds
0 i,j =1
T
−C g(t)dt ∇a 2
L2 ()
− C( ∇a L2 () + ∇uεN 2
L2 (0,T ;L2 ())
).
0
Since
T
n
g(s) (aij (·, T − s)∂i uεN (·, T − s), ∂j uεN (·, T − s))L2 () ds
0 i,j =1
T
n
= g(T − s) (aij (·, s)∂i uεN (·, s), ∂j uεN (·, s))L2 () ds
0 i,j =1
−α
≥ CT ∇uεN 2L2 (0,T ;L2 ())
4.4 Proof of Theorem 4.2 95
by (4.2), we obtain
T
[the second term on the left-hand side of (4.35)]dt (4.40)
0
⎛ ⎞
T
n
= ⎝ aij (·, t)∂j uεN (·, t), dtα ∂i uεN (·, t)⎠ dt
0 i,j =1 L2 ()
≥ − C( ∇uεN 2
L2 (0,T ;L2 ())
+ ∇a 2
L2 ()
).
By the Poincaré inequality and uεN (·, t) ∈ H01 (), for small δ > 0, we can
estimate
T
|[the third and the fourth terms on the left-hand side of (4.35)]dt
0
T
≤ (Cδ ∇uεN (·, t) 2
L2 ()
+ δ dtα uεN (·, t) 2
L2 ()
)dt (4.41)
0
and
T
(FNε (·, t), dtα uεN (·, t))L2 () dt
0
≤ Cδ FNε 2
L2 (0,T ;L2 ())
+ Cδ dtα uεN (·, t) 2
L2 (0,T ;L2 ())
. (4.42)
Fixing δ > 0 sufficiently small and absorbing the third term on the right-hand
side into the left-hand side and applying (4.30), which was verified in the proof
of Theorem 4.1, we obtain
∂tα (uεN − a) 2
L2 (0,T ;L2 ())
= dtα uεN 2
L2 (0,T ;L2 ())
≤ C( ∇aN 2
L2 ()
+ FNε 2
L2 (0,T ;L2 ())
), (4.43)
We note that the uniqueness of solutions is proved for less regular class, and the
uniqueness holds also within the stronger regularity of solutions in Theorem 4.2.
Then we have
H −1 () ∂s (u − a)(·, s), u − aH 1 () + H −1 () A(s)(u − a)(·, s), u − aH 1 ()
α
0 0
+ H −1 () A(s)a, u − aH 1 () = H −1 () F (·, s), (u − a)(·, s)H 1 ()
0 0
for 0 < s < T . Here we note that for any ε > 0 there exists a constant Cε > 0 such
that
⎧
⎪
⎪ |H −1 () A(s)a, u(·, s) − aH 1 () | ≤ C a H 1 () (u − a)(·, s) H 1 ()
⎪
⎪ 0 0 0
⎪
⎪ ≤ ε (u − a)(·, s) 2H 1 () + Cε a 2H 1 () ,
⎪
⎪
⎪
⎨ n
⎪ 0 0
j =1 H −1 () bj (·, s)∂j (u(·, s) − a), u(·, s) − aH01 ()
⎪
⎪ + |H −1 () c(·, s)(u(·, s) − a), u(·, s) − aH 1 () |
⎪
⎪
⎪
⎪
0
⎪ ≤ C u(·, s) − a L2 () u(·, s) − a H01 ()
⎪
⎪
⎪
⎩ ≤ ε u(·, s) − a 2 1 + Cε u(·, s)0 − a 2 2 .
H () 0
L ()
(4.44)
Here we used
bj (·, s)∂j (u(·, s) − a) H −1 () = sup |H −1 () bj (·, s)∂j (u(·, s) − a), ψH 1 () |
0
ψ H 1 ()
=1
0
Multiplying (t −s)
α−1
(α) and integrating in s over (0, t) and applying (4.2) and
Theorem 3.4 (ii), we obtain
t
u(·, t) − a 2
L2 ()
+ (t − s)α−1 ∇(u(·, s) − a) 2
L2 ()
ds
0
t
≤C (t − s)α−1 |H −1 () A(s)a, u(·, s) − aH 1 () |ds
0
0
t
n
+ (t − s)α−1 −1
H () b j (·, s)∂j (u(·, s) − a), u(·, s) − a 1 ds
H0 ()
0 j =1
t
+ H −1 () c(·, s)(u(·, s) − a), u(·, s) − aH 1 () (t − s)α−1 ds
0
0
t
+C (t − s)α−1 F (·, s) H −1 () u(·, s) − a H 1 () ds.
0
0
we obtain
t
u(·, t) − a 2
L2 ()
+ (t − s)α−1 u(·, s) − a 2
H01 ()
ds
0
t t
≤ε (t − s) α−1
u(·, s) − a 2
H01 ()
ds + Cε (t − s)α−1 u(·, s) − a 2
L2 ()
ds
0 0
t t
+ Cε (t − s) α−1
a 2
H01 ()
ds + Cε (t − s)α−1 F (·, s) 2
H −1 ()
ds.
0 0
Choosing ε > 0 sufficiently small, we absorb the first term on the right-hand side
into the left-hand side, so that
u(·, t) − a 2
L2 ()
t
2 2
≤ u(·, t) − a L2 ()
+C (t − s)α−1 u(·, s) − a H01 ()
ds
0
t
2
≤C (t − s)α−1 u(·, s) − a L2 ()
ds
0
t
+ Ct α a 2
H01 ()
+C (t − s)α−1 F (·, s) 2
H −1 ()
ds.
0
98 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
t t q1 t p2
p
(t − s) α−1
F (·, s) 2
H −1 ()
ds ≤ (t − s)qα−q
ds ( F (·, s) 2
H −1 ()
) 2 ds
0 0 0
θ
≤ Ct q F 2
Lp (0,t;H −1 ())
,
Multiplying both sides by (u − a)(x, s)(t − s)α−1 and integrating over × (0, t),
we obtain
t
(∂s (u − a)(x, s))(u − a)(x, s)dx (t − s)α−1 ds
α
0
⎛ ⎞
t
n
− ⎝ ∂i (aij (x, s)∂j u(x, s))u(x, s)dx ⎠ (t − s)α−1 ds
0 i,j =1
⎛ ⎞
t
n
+ ⎝ ∂i (aij (x, s)∂j u(x, s))a(x)dx ⎠ (t − s)α−1 ds
0 i,j =1
4.5 Proofs of Propositions 4.1 and 4.2 99
⎛ ⎞
t
n
− ⎝ bj (x, s)∂j u(x, s)(u(x, s) − a(x))dx ⎠ (t − s)α−1 ds
0 j =1
t
− c(x, s)u(x, s)(u(x, s) − a(x))dx (t − s)α−1 ds = 0.
0
Since (u − a)(x, ·) ∈ Hα (0, T ) for x ∈ and u(·, s) ∈ H01 () for 0 < s < t, we
integrate by parts and use Theorem 3.4 (ii) where we note
H −1 () ∂s (u(·, s)−a), u(·, s)−aH 1 () =
α
(∂sα (u(x, s)−a(x))(u(x, s)−a(x))dx,
0
t
1 α−1 α−1
− C3 ε|∇u(x, s)|(t − s) |a(x)|(t − s) 2 dxds 2
0 ε
t t
− C3 |u(x, s)| (t − s) dxds − C3
2 α−1
|a(x)|2(t − s)α−1 dxds ≤ 0.
0 0
Therefore
t
u(·, t) 2
L2 ()
+ u(·, s) 2
H01 ()
(t − s)α−1 ds
0
t
≤ Ct α a 2
H01 ()
+C u(·, s) 2
L2 ()
(t − s)α−1 ds, t > 0,
0
that is,
t
u(·, t) 2
L2 ()
≤ Ct α a 2
H01 ()
+C u(·, s) 2
L2 ()
(t − s)α−1 ds, t > 0.
0
Consequently, noting that the constant C > 0 can be chosen uniformly in all t > 0
and applying Lemma A.2 in the Appendix, we can take a constant C0 > 0 such that
t
2 2 2
u(·, t) L2 ()
≤ C0 t α
a H01 ()
+ C0 e C0 t
(t − s) α−1 α
s ds a H01 ()
0
(α)(α + 1) 2α
≤ C0 t α + C0 eC0 t t a 2
(2α + 1) H01 ()
In the case where the coefficients of A(t) are independent of t, we can represent the
solution to the initial boundary value problem by the Mittag-Lefller functions (e.g.,
Sakamoto and Yamamoto [26]). In this section, we show that such a represented
solution coincides with the solution established by Theorem 4.1.
Let
n
−Av(x) = ∂i (aij (x)∂j v(x)) + c(x)v(x), x ∈ ,
i,j =1
0 < λ1 ≤ λ2 ≤ · · · −→ ∞.
Henceforth if there is no fear of confusion, then we write the scalar product (f, g)
for f, g ∈ L2 (), in place of (f, g)L2 () .
Let ϕk , k ∈ N be an eigenfunction of A for λn such that
1, k = ,
(ϕk , ϕ ) := ϕk (x)ϕ(x)dx =
0, k = .
We will verify:
Theorem 4.3 Let a ∈ L2 () and F ∈ L2 (0, T ; H −1 ()). Then v given by (4.45)
satisfies v − a ∈ Hα (0, T ; H −1 ()), v ∈ L2 (0, T ; H01 ()) and (4.3)–(4.5).
The theorem means that (4.45) coincides with the solution by Theorem 4.1. We
can similarly prove that v defined by (4.45) is the same as u in Theorem 4.2 if
a ∈ H01 () and F ∈ L2 (0, T ; L2 ()), but we omit the proof.
Proof It is sufficient to prove that v − a ∈ Hα (0, T ; H −1 ()) and v satisfies (4.3).
First Step First we note that D(A− 2 ) = H −1 () = (H01 ()) and there exist
1
∞
Since a = k=1 (a, ϕk )ϕk in L2 (), we have
∞
v1 (x, t) − a(x) = (a, ϕk )(Eα,1 (−λk t α ) − 1)ϕk (x). (4.46)
k=1
for almost all x ∈ . Moreover, since Eα,1 (−λk t α ) − 1 ∈ Hα (0, T ) ∩ 0 W 1,1 (0, T )
by Proposition 2.2, we see that
so that
∞
∂tα (v1 − a)(x, t) = (a, ϕk )∂tα (Eα,1 (−λk t α ) − 1)ϕk (x)
k=1
∞
=− λk (a, ϕk )Eα,1 (−λk t α )ϕk (x). (4.47)
k=1
that is,
2 2
∂tα (v1 − a) L2 (0,T ;H −1 ())
≤C a L2 ()
. (4.48)
d
Eα,1 (−ηα ) ≤ 0, Eα,α (−ηα ) ≥ 0, η ≥ 0, 0<α<1 (4.49)
dη
(e.g., Gorenflo et al. [10], Miller and Samko [22]). Therefore we obtain
η η
|s α−1 Eα,α (−λk s α )|ds = s α−1 Eα,α (−λk s α )ds
0 0
η
1 d 1
=− Eα,1 (−λk s α )ds = (1 − Eα,1 (−λk ηα )). (4.50)
λk 0 ds λk
Here we recall: since Eα,1 (z) is an entire function in z and (αn + 1) = αn(αn),
we have
+∞ , ∞
d d (−λk )n s nα (−λk )n nαs nα−1
Eα,1 (−λk s ) =
α
=
ds ds (αn + 1) (αn + 1)
n=0 n=1
∞
∞
(−λk )n s nα−1 (−λk )m s mα
= = s α−1 (−λk ) = −λk s α−1 Eα,α (−λk s α ).
(αn) (αm + α)
n=1 m=0
104 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
By (4.51), we have
d 1−α
∂tα wk (t) = J wk (t)
dt
t d
= F (·, t), ϕk + (Eα,1 (−λk (t − ξ )α )F (·, ξ ), ϕk dξ
0 dt
t
= F (·, t ), ϕk −λk (t −s)α−1 Eα,α (−λk (t −s)α )F (·, s), ϕk ds = F (·, t ), ϕk −λk wk (t ).
0
Hence
∞
∂tα v2 (x, t) = F (·, t), ϕk ϕk (x)
k=1
- ∞ .
t
+ − λk (t − s)α−1 Eα,α (−λk (t − s)α )F (·, s), ϕk ds ϕk =: S1 + S2 .
k=1 0
(4.52)
Therefore
∞
1
∂tα v2 (·, t) 2
H −1 ()
= |(F (·, t), ϕk )|2 + S2 (·, t) 2
H −1 ()
λk
k=1
≤ F (·, t) 2
H −1 ()
+ S2 (·, t) 2
H −1 ()
.
Next
∞
1
S2 (·, t) 2
H −1 ()
= (S2 (·, t), ϕk )2
λk
k=1
∞
t 2
= λk (t − s)α−1 Eα,α (−λk (t − s)α )F (·, s), ϕk ds
k=1 0
and the Young inequality for the convolution (Lemma A.1) yields
2
S2 L2 (0,T ;H −1 ())
∞
t 2
T
= λk (t − s)α−1 Eα,α (−λk (t − s)α )F (·, s), ϕk ds dt
k=1 0 0
∞
T 2 T
1
≤ λ2k |s α−1
Eα,α (−λk s )|ds α
|F (·, s), ϕk |2 ds.
0 λk 0
k=1
106 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
Consequently,
∞
T 1
S2 2
L2 (0,T ;H −1 ())
≤ |F (·, s), ϕk |2 ds
0 λk
k=1
T
= F (·, s) 2
H −1 ()
ds = F 2
L2 (0,T ;H −1 ())
. (4.54)
0
Hence
we see that
Fourth Step Here we verify that v ∈ L2 (0, T ; H01 ()) and estimate
1
A v L2 (0,T ;L2 ()), which is equivalent to ∇v
2
L2 (0,T ;L2 ()) .
First Lemma 2.5 (i) yields
∞ 2
1 ∞
A 2 (a, ϕk )Eα,1 (−λk t α )ϕk = λk |(a, ϕk )|2 |Eα,1 (−λk t α )|2
k=1 L2 () k=1
⎛ ⎞2
∞
2 ∞
1 α
2 2
1 λ t
≤C |(a, ϕk )|2 λk = Ct −α |(a, ϕk )|2 ⎝ k α ⎠
1 + λk t α 1 + λk t
k=1 k=1
≤ Ct −α a 2
L2 ()
.
4.6 Case Where the Coefficients of A(t) Are Independent of Time 107
≤C F L2 (0,T ;H −1 ()) .
N
N
aN (x) = (a, ϕk )ϕk (x), v1N (x, t) = (a, ϕk )Eα,1 (−λk t α )ϕk (x),
k=1 k=1
N
t
v2N (x, t) = F (·, s), ϕk (t − s)α−1 Eα,α (−λk (t − s)α )ds ϕk (x)
k=1 0
and
N
∂tα (v1N − aN )(x, t) = − λk (a, ϕk )Eα,1 (−λk t α )ϕk (x)
k=1
+N ,
=−A (a, ϕk )Eα,1 (−λk t α )ϕk (x) = −Av1N
k=1
108 4 Initial Boundary Value Problems for Time-Fractional Diffusion Equations
and
N
∂tα v2N (x, t) = F (·, t), ϕk ϕk
k=1
+N ,
t
−A F (·, s), ϕk (t − s)α−1 Eα,α (−λk (t − s)α )dsϕk
k=1 0
N
= − Av2N (x, t) + F (·, t), ϕk ϕk .
k=1
Hence
N
∂tα (v1N − a)(x, t) + Av N = F (·, t), ϕk ϕk
k=1
We consider
⎧ α
⎨ ∂t (u − a) + A(t)u(x, t) = 0 in L2 (), 0 < t < T ,
u(·, t) ∈ H01 (), 0 < t < T , (5.1)
⎩
u − a ∈ Hα (0, T ; L2 ()).
Here
n
− (A(t)v)(x, t) = ∂i (aij (x, t)∂j v(x)) + c(x, t)v, (5.2)
i,j =1
where
In the case of α = 1, we can prove that there exist constants C > 0 and θ > 0 such
that
for each a ∈ L2 () by the classical energy estimate. For 0 < α < 1, we know that
C
u(·, t) L2 () ≤ a L2 () , t >0 (5.3)
tα
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 109
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5_5
110 5 Decay Rate as t → ∞
in the case where the coefficients of A(t) are independent of t (e.g., [26]). On
the other hand, Vergara and Zacher [28] proved (5.3) for t-dependent A(t). For
t-dependent A(t), in a conventional energy estimate, we multiply the equation by
u − a and integrate over and we may obtain only a weak estimate
C
u(·, t) L2 () ≤ α a H01 () , t > 0.
t2
This is weaker than we can expect in terms of (5.3) for t-independent A(t), because
the decay rate is α2 .
In this chapter, we modify the arguments in Vergara and Zacher [28] to prove
(5.3) within our framework of solutions in the case where the coefficients of A(t)
are dependent on time t. More precisely, we prove:
Theorem 5.1 In (5.2), we assume
n
n
aij (x, t)ξi ξj ≥ μ0 ξj2 , x ∈ , t ≥ 0, ξ1 , ..., ξn ∈ R (5.5)
i,j =1 j =1
and
c(x, t) ≤ 0, x ∈ , t ≥ 0. (5.6)
and
Furthermore
u(·, t) L2 () dtα u(·, t) L2 () ≤ u(x, t)dtα u(x, t)dx. (5.10)
as M → 0 by |∂s u(x, s)| ≤ Cs α−1 by (5.8). Hence dtα u(·, t) 2L2 () ≤ 0, and (5.9)
holds for u(·, t) L2 () = 0.
Finally, we have to prove (5.10). We can assume that u(·, t) L2 () = 0,
because (5.10) is trivial if u(·, t) L2 () = 0. By (5.8), we can directly verify
ds u(·, s) L2 () ∈ L (0, T ). First,
d 1
dtα u(·, t) 2
L2 ()
− 2 u(·, t) α
L2 () dt u(·, t) L2 ()
t
2 d
= (t − s)−α ( u(·, s) L2 () − u(·, t) L2 () ) u(·, s) L2 () ds
(1 − α) 0 ds
+ , + ,
2 t u(·, s) d u(·, s) L2 ()
−α L2 ()
= (t − s) u(·, t) 2
−1 ds
(1 − α) 0
L2 () u(·, t) L2 () ds u(·, t) L2 ()
114 5 Decay Rate as t → ∞
- + ,.
2 t d u(·, s)
−α L2 ()
= (t − s) u(·, t) 2
G ds
(1 − α) 0
L2 () ds u(·, t) L2 ()
+ ,
2α t u(·, s)
−α−1 L2 ()
=− (t − s) u(·, t) ds 2
L2 ()
G
(1 − α) 0 u(·, t) L2 ()
+ ,
2 u(·, s) L2 () s=t
+ (t − s)−α u(·, t) 2L2 () G
(1 − α) u(·, t) L2 () s=0
t + ,
−2α −α−1
u(·, s) L2 ()
≥ (t − s) 2
u(·, t) L2 () G ds
(1 − α) 0 u(·, t) L2 ()
+ ,
2t −α u(·, 0) L2 ()
− 2
u(·, t) L2 () G . Here we used also (5.8). Hence
(1 − α) u(·, t) L2 ()
+ ,
2α t u(·, s)
−α−1 L2 ()
dtα u(·, t) 2
+ (t − s) u(·, t) 2
G ds
L2 () (1 − α) 0
L2 () u(·, t) L2 ()
+ ,
2t −α u(·, 0) L2 ()
+ u(·, t) 2
L2 ()
G
(1 − α) u(·, t) L2 ()
≥ 2 u(·, t) α
L2 () dt u(·, t) L2 () .
Combining with (5.9), we reach (5.10). Thus the proof of Lemma 5.1 is complete.
Moreover, we show a generalized extremum principle under weaker assumptions
than Lemma 1.2.
Lemma 5.2 Let f ∈ W 1,1 (0, T ) attain its maximum over the interval [0, T ] at a
point t0 ∈ (0, T ]. If for each κ ∈ (0, T ), there exists β ∈ (0, 1] such that f ∈
1
W 1, 1−β (κ, T ), then (dtα f )(t0 ) ≥ 0 for every α ∈ (0, β).
Proof Firstly, we introduce the function g(t) := f (t0 ) − f (t) for t ∈ [0, T ].
We notice that g(t) ≥ 0 and (dtα g)(t) = −(dtα f )(t) for t ∈ [0, T ]. Then
g ∈ W 1,1 (0, T ) and g(t0 ) = 0. Hence, for κ ∈ (0, t0 ), the Hölder inequality yields
t0 dg
|g(t)| ≤ (s) ds ≤ dg 1 |t − t0 |β for t ∈ [κ, t0 ]. (5.12)
ds dt 1−β
t L (κ,T )
Young inequality, we deduce that there exists sufficiently small κ > 0 such that the
first integral is smaller than ε. As for the second one, integration by parts, g ≥ 0 on
[0, T ] and g(t0 ) = 0 yield
t0 t0 −h
dg dg
(t0 − s)−α (s)ds = lim (s)ds (t0 − s)−α
κ ds
κ ds
h→0+
t0 −h
−α −α
= − (t0 − κ) g(κ) + lim h g(t0 − h) − lim α (t0 − s)−α−1 g(s)ds
h→0+ h→0+ κ
−α
≤ lim h g(t0 − h) = 0.
h→0+
For the last limit, we used (5.13) by noting 0 < α < β. Thus the proof of Lemma 5.2
is completed.
We apply the argument in Sect. 4.3 of Chap. 4 for proving the existence of u to
(5.1), and we use the same notations. Since the non-homogeneous term F (x, t) is
identically zero, we need not introduce the approximation parameter ε > 0 like
Sect. 4.3 of Chap. 4.
We recall that the function
N
uN (x, t) := pN, (t)ρ (x)
=1
satisfies
N
(dtα uN )(x, t)ρ (x)dx + aij (x, t)∂j uN (x, t)∂i ρ (x)dx
i,j =1
= c(x, t)uN (x, t)ρ (x)dx
116 5 Decay Rate as t → ∞
pN, − c ∈ Wα (0, T ).
N
uN (x, t)(dtα uN )(x, t)dx + aij (x, t)(∂j uN )(x, t)(∂i uN )(x, t)dx
i,j =1
= c(x, t)|uN (x, t)|2 dx.
We define
uN,η = uN + ηρN+1 ,
where η > 0 is a constant. Using the fact that dtα (ηρN+1 ) = 0 and the orthogonality
of {ρN }∞
N=1 , we obtain
uN,η (x, t)dtα uN,η (x, t)dx + μ0 ∇uN (t, ·) 2
L2 ()
≤ 0.
By Theorem 3.6 (iv) in Chap. 3, we see that the functions uN,η satisfy the
assumptions of Lemma 5.1, so that (5.10) and the Poincaré inequality yield
2
uN,η (·, t) α
L2 () dt uN,η (·, t) L2 () + c0 uN (·, t) L2 ()
≤ 0. (5.14)
N
(uN , ρN+1 ) = pN, (t)(ρ , ρN+1 ) = 0,
=1
we have
uN,η (·, t) 2
L2 ()
= uN (·, t) 2
L2 ()
+ η2 ,
5.3 Completion of Proof of Theorem 5.1 117
uN,η (·, t) α
L2 () dt uN,η (·, t) L2 () + c0 uN,η (·, t) 2
L2 ()
= uN,η (·, t) α
L2 () dt uN,η (·, t) L2 () + c0 uN (·, t) 2
L2 ()
+ c0 η 2 ≤ c0 η 2
by (5.14). Consequently,
uN,η (·, t) α
L2 () dt uN,η (·, t) L2 () + c0 uN,η (·, t) 2
L2 ()
≤ η 2 c0 .
(dtα uN,η (·, t) L2 () , ξ ) + (c0 uN,η (·, t) L2 () , ξ ) ≤ (ηc0 , ξ ) (5.16)
for all ξ ∈ C0∞ (0, T ) satisfying ξ ≥ 0. In view of (3.12), by Lemma 2.7 and
integration by parts, we know that (2.51) holds for u ∈ W 1,1 (0, T ) and ξ ∈
C0∞ (0, T ):
By the definition of uN,η , we can readily verify that uN,η (·, t) L2 () −→
uN (·, t) L2 () in L1 (0, T ) and uN,η (·, 0) L2 () −→ uN (·, 0) L2 () as η → 0.
Therefore
T
1
= ( uN (·, t) L2 () , (dtα )∗ ξ ) − t −α ξ(t)dt uN (·, 0) L2 () .
(1 − α) 0
Since uN (·, t) L2 () ∈ W 1,1 (0, T ) by Theorem 3.6 (iv), we apply (2.51) to the
right-hand side of (5.17) for uN (·, t) L2 () , and we obtain
Again using the letter θ , we denote the zero extension of θ to R, and θ ∈ L1 (R).
Setting ξ(t) = χε (τ − t) with arbitrarily fixed τ ∈ (0, T ) and ε > 0, we see
that ξ ≥ 0and ξ ∈ C0∞ (0, T ) with sufficiently small ε > 0. By (5.18) we obtain
∞
θε (τ ) := −∞ θ (t)χε (τ − t)dt ≤ 0 for small ε > 0. We know that θε −→ θ
in L1 (R) as ε → 0 (e.g., [2]), so that we can choose a subsequence {εn }n∈N with
limn→∞ εn = 0 such that limn→∞ θεn (τ ) = θ (τ ) for almost all τ ∈ (0, T ). Hence,
letting n → ∞ in (5.18), we reach
dtα uN (·, t) L2 () + c0 uN (·, t) L2 () ≤0 for almost all t ∈ (0, T ). (5.19)
Using the power series of Eα,1 (−c0 t α ), one can directly verify that
satisfies
Needless to say, one purpose of the linear theory developed in this book is the
applications to the nonlinear theory.
3. Here we argue only the homogeneous Dirichlet boundary condition. For exam-
ple, the Neumann boundary condition should be considered and initial nonhomo-
geneous boundary value problems are demanded. The theory for common partial
differential equations such as α = 1, 2 has been established satisfactorily (e.g.,
Lions and Magenes [19]). Such a theory is not only mathematically interesting
but also is a basis for e.g., control problems, and widely applicable. As for
non-homogeneous boundary value problems for fractional partial differential
equations, we refer only to Yamamoto [29] and the references therein.
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 121
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5_6
122 6 Concluding Remarks on Future Works
4. There are various topics on inverse problems where we are required to determine
parameters such as coefficients and orders of the derivatives of fractional
differential equations by data of solutions which are over-determining in view
of the well-posedness of initial boundary value problems. As for surveys, see
the three chapters [16a–c] in the handbook edited by Kochubei, Luchko and
Machado. The current book has established the foundations for future studies of
inverse problems for fractional differential equations, and we can greatly expect
the comprehensive development of studies of inverse problems.
Appendix A
Proofs of Two Inequalities
For convenience, we prove two inequalities which have been used in this book.
We set
t
(ρ ∗ v)(t) = ρ(t − s)v(s)ds, 0 < t < T .
0
we have
t T t
T
ρ ∗v = ρ(t − s)v(s)ds dt ≤ |ρ(t − s)||v(s)|ds dt
L1 (0,T )
0 0 0 0
T T T T −s
= |ρ(t − s)|dt |v(s)|ds = |ρ(η)|dη |v(s)|ds
0 s 0 0
T T
≤ |ρ(η)|dη |v(s)|ds = ρ L1 (0,T ) v L1 (0,T ) .
0 0
For the second equality from the last, we changed the integral variables η = t − s.
Thus the proof in the case of p = 1 is complete.
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 123
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5
124 A Proofs of Two Inequalities
Then
1
1− q1
|ρ(t − s)v(s)| = |ρ(t − s)| q (|ρ(t − s)| |v(s)|),
1
t p1
≤ ρ q
L1 (0,T )
|ρ(t − s)||v(s)| ds p
.
0
Then
t
u(t) ≤ r(t) + C1 e C2 t
(t − s)α−1 r(s)ds, 0 ≤ t ≤ T. (A.2)
0
Here the constants C1 > 0 and C2 > 0 are dependent on α, C0 , but independent of
T > 0. We note that if C0 > 0 is independent of T > 0, then (A.2) holds for t > 0
with C1 , C2 > 0 which are independent of t > 0.
Proof We set
s
(Lr)(s) = C0 (s − ξ )α−1 r(ξ )dξ, 0<s<T
0
u ≤ r + Lu ≤ r + Lr + L2 u in (0, T ).
N
u≤r+ Lk r + LN+1 u in (0, T ) (A.4)
k=1
Here, by r ∈ L1 (0, T ) and the Young inequality, we note that all the functions
including r(ξ ) in the calculations here are in L1 (0, T ). Continuing the calculations,
we can reach
(C0 (α))k t
(Lk r)(t) = (t − s)kα−1 r(s)ds, k ∈ N. (A.5)
(kα) 0
126 A Proofs of Two Inequalities
1
We set θ = (C0 (α)) α . Substitution of (A.5) into (A.4) yields
t +
N
, t
θ αk (t − s)αk−1 θ α(N +1)
u(t ) ≤ r(t ) + r(s)ds + (t − s)(N +1)α−1 u(s)ds
0 (αk) ((N + 1)α) 0
k=1
t +
N
, t
θ αk s αk−1 θ α(N +1)
= r(t ) + r(t − s)ds + (t − s)(N +1)α−1 u(s)ds.
0 (αk) ((N + 1)α) 0
k=1
(A.6)
On the other hand, since Eα,1 (z) is an entire function in z ∈ C, we can verify
+ ∞
, ∞ ∞
d d (θ s)αk θ αk (αk)s αk−1 θ αk s αk−1
Eα,1 ((θ s)α ) = = =
ds ds (αk + 1) (αk + 1) (αk)
k=0 k=1 k=1
Moreover, we have
1−α 1 C
|Eα,α (η)| ≤ C(1 + η) eη +
α
α
1+η
for all η ≥ 0 (e.g., Theorem 1.5 (p. 35) in Podlubny [24]). Therefore
1−α C
|Eα,α ((θ s)α )| ≤ C(1 + θ α s α ) α eθs + ≤ C1 eC2 s , s > 0,
1 + θ αsα
Let us choose N0 ∈ N such that (N0 + 1)α − 1 > 0. Then for each N ≥ N0 , we
have
t t
(t − s)(N+1)α−1 u(s)ds ≤ t (N+1)α−1 |u(s)|ds ≤ T (N+1)α−1 u L1 (0,T ) .
0 0
Consequently,
t
θ N+1 α N
(t − s) (N+1)α−1
u(s)ds ≤ θ T α−1 (θ T ) u
((N + 1)α) ((N + 1)α) L1 (0,T )
0
(θ T α )N
lim = 0.
N→∞ ((N + 1)α)
∇ = (∂1 , . . . , ∂n ) Gradient
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 129
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5
References
1. N.H. Abel, Résolution d’un problème de mécanique. J. Reine Angew. Math. 1, 97–101 (1826)
2. R.A. Adams, Sobolev Spaces (Academic, New York, 1975)
3. A. Alsaedi, B. Ahmad, M. Kirane, A survey of useful inequalities in fractional calculus. Fract.
Calc. Appl. Anal. 20, 574–594 (2017)
4. E.G. Bajlekova, Fractional evolution equations in Banach spaces, Doctoral thesis, Eindhoven
University of Technology, 2001
5. H. Brezis, Functional Analysis, Sobolev Spaces and Partial Differential Equations (Springer,
Berlin, 2011)
6. K. Diethelm, The Analysis of Fractional Differential Equations. Lecture Note in Mathematics,
vol. 2004 (Springer, Berlin, 2010)
7. L.C. Evans, Partial Differential Equations (American Mathematical Society, Providence, 1998)
8. R. Gorenflo, S. Vessella, Abel Integral Equations. Lecture Note in Mathematics, vol. 1461
(Springer, Berlin, 1991)
9. R. Gorenflo, M. Yamamoto, Operator theoretic treatment of linear Abel integral equations of
first kind. Jpn. J. Ind. Appl. Math. 16, 137–161 (1999)
10. R. Gorenflo, A.A. Kilbas, F. Mainardi, S.V. Rogosin, Mittag-Leffler Functions, Related Topics
and Applications (Springer, Berlin, 2014)
11. R. Gorenflo, Y. Luchko, M. Yamamoto, Time-fractional diffusion equation in the fractional
Sobolev spaces. Fract. Calc. Appl. Anal. 18, 799–820 (2015)
12. B. Jin, B. Li, Z. Zhou, Subdiffusion with a time-dependent coefficient: analysis and numerical
solution. Math. Comput. 88, 2157–2186 (2019)
13. T. Kato, Perturbation Theory for Linear Operators (Springer, Berlin, 1980)
14. Y. Kian, M. Yamamoto, On existence and uniqueness of solutions for semilinear fractional
wave equations. Fract. Calc. Appl. Anal. 20, 117–138 (2017)
15. A.A. Kilbas, H.M. Srivastava, J.J. Trujillo, Theory and Applications of Fractional Differential
Equations (Elsevier, Amsterdam, 2006)
16. A.N. Kochubei, Y. Luchko, J.A. Tenreiro Machado (eds.), Handbook of Fractional Calculus
with Applications, vol. 2 (De Gruyter, Berlin, 2019); (a) Y. Liu, Z. Li, M. Yamamoto, Inverse
problems of determining sources of the fractional partial differential equations, pp. 411–429;
(b) Z. Li, Y. Liu, M. Yamamoto, Inverse problems of determining parameters of the fractional
partial differential equations, pp. 431–442; (c) Z. Li, M. Yamamoto, Inverse problems of
determining coefficients of the fractional partial differential equations, pp. 443–464
17. A. Kubica, M. Yamamoto, Initial-boundary value problems for fractional diffusion equations
with time-dependent coefficients. Fract. Calc. Appl. Anal. 21, 276–311 (2018)
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 131
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5
132 References
18. Z. Li, X. Huang, M. Yamamoto, Initial-boundary value problems for multi-term time-fractional
diffusion equations with x-dependent coefficients. Evol. Equ. Control Theory 9, 153–179
(2020)
19. J.L. Lions, E. Magenes, Non-homogeneous Boundary Value Problems and Applications, vols.
I, II (Springer, Berlin, 1972)
20. Y. Luchko, Maximum principle for the generalized time-fractional diffusion equation. J. Math.
Anal. Appl. 351, 218–223 (2009)
21. Y. Luchko, Some uniqueness and existence results for the initial-boundary value problems for
the generalized time-fractional diffusion equation. Comput. Math. Appl. 59, 1766–1772 (2010)
22. K.S. Miller, S.G. Samko, Completely monotonic functions. Integral Transforms Spec. Funct.
12, 389–402 (2001)
23. A. Pazy, Semigroups of Linear Operators and Applications to Partial Differential Equations
(Springer, Berlin, 1983)
24. I. Podlubny, Fractional Differential Equations (Academic, San Diego, 1999)
25. B. Ross, The development of fractional calculus 1695–1900. Hist. Math. 4, 75–89 (1977)
26. K. Sakamoto, M. Yamamoto, Initial value/boundary value problems for fractional diffusion-
wave equations and applications to some inverse problems. J. Math. Anal. Appl. 382, 426–447
(2011)
27. H. Tanabe, Equations of Evolution (Pitman, London, 1979)
28. V. Vergara, R. Zacher, Optimal decay estimates for time-fractional and other nonlocal
subdiffusion equation via energy methods. SIAM J. Math. Anal. 47, 210–239 (2015)
29. M. Yamamoto, Weak solutions to non-homogeneous boundary value problems for time-
fractional diffusion equations. J. Math. Anal. Appl. 460, 365–381 (2018)
30. R. Zacher, Weak solutions of abstract evolutionary integro-differential equations in Hilbert
spaces. Funkcial. Ekvac. 52, 1–18 (2009)
Index
A G
Caputo derivative 2 H
Closable 25
Closure 25
Coercivity 48 Heinz–Kato inequality 20
Complete monotonicity 103 Homogeneous Dirichlet boundary condition
121
D
I
L
Fractional ordinary differential equation 47
Fractional power 19
Fractional Sobolev space 12, 65 Laplace transform 9, 43
Fredholm alternative 57
© The Author(s), under exclusive licence to Springer Nature Singapore Pte Ltd. 2020 133
A. Kubica et al., Time-Fractional Differential Equations, SpringerBriefs
in Mathematics, https://doi.org/10.1007/978-981-15-9066-5
134 Index
M R
N S
P Y