1. Introduction
Applications of the maximum principle in functional analysis are well known and we refer the interested reader to the authoritative account [
1]. In recent years, the maximum principle has become an important tool in the study of boundary value problems for fractional differential equations. Early applications appear in [
2,
3] where explicit Green’s functions, expressed in terms of power functions, were constructed; sign properties of the Green’s function were analyzed so that fixed point theorems could be applied to give sufficient conditions for the existence of positive solutions. More recently, Green’s functions, expressed in terms of Mittag-Leffler functions, have been constructed so that fixed-point theorems and the maximum principle can be applied. See, for example, Refs. [
4,
5,
6,
7].
Credit for the discovery of an anti-maximum principle is given to Clément and Peletier [
8]. Although primarily interested in partial differential equations, they initially illustrated the anti-maximum principle with the boundary value problem,
,
with
They showed, if
and if
, then the boundary value problem is uniquely solvable and
implies
where
y is the unique solution associated with
f.
At
the boundary value problem,
,
is at resonance, and
is a simple eigenvalue of the homogeneous problem. Moreover, for
then
implies
that is, for
the boundary value problem obeys a maximum principle. Thus, there has been a change in the sign property, maximum principle or anti-maximum principle, through the simple eigenvalue
In more succinct terms, if
, and if
, then the boundary value problem is uniquely solvable and
implies
where
y is the unique solution associated with
f. Since the publication of [
8], the change in behavior from maximum to anti-maximum principles as a function of the parameter has received considerable attention. For partial differential equations, see [
9,
10,
11,
12,
13,
14,
15,
16]. For ordinary differential equations, see [
17,
18,
19,
20,
21]. More recently, this change in behavior from maximum to anti-maximum principles has also been noticed and studied in fractional differential equations. For equations analyzing the fractional
Laplacian, see [
22,
23]; for fractional differential equations of one independent variable, see [
24].
In [
9], the authors studied the nature of the maximum principle for boundary value problems for an abstract differential equation,
defined on
with
under a fundamental assumption that
was a simple eigenvalue for the homogeneous problem. Under mild sufficient conditions, they proved the existence of
and a constant
, independent of
f, such that
where
y is the unique solution of the boundary value problem associated with
and
If (
1) holds and
then
implies
that is, the boundary value problem for (
1) obeys a maximum principle. If (
1) holds and
then
implies
that is, the boundary value problem for (
1) obeys an anti-maximum principle [
8].
The methods of [
9] were recently adapted to apply to a boundary value problem with a parameter for a Riemann–Liouville fractional differential equation [
24]. During the review process for [
24], those authors were asked by one referee if the methods of [
9] could be successfully adapted to apply to analogues of Neumann or periodic boundary value problems for Riemann–Liouville fractional differential equations. In [
24], the eigenspace generated by
is contained in the space of continuous functions on
The corresponding eigenspace for boundary value problems analogous to Neumann or periodic type boundary value problems will contain a singularity. Thus, the question is interesting. The purpose of this study is to address that question with a positive response.
In
Section 2, we shall introduce preliminary notations and concepts from fractional calculus. We shall also introduce four boundary value problems for which the general theorem, stated in
Section 3, applies. In
Section 3, we introduce the notations adapted from [
9] and state and prove the abstract theorem. The proof of the abstract theorem closely models the proofs of analogous theorems in [
9,
24]; with subtle differences in the technical details due to the specific function space, we shall produce a proof here for the self-containment of the manuscript. In
Section 4, we shall apply the abstract theorem to each of the four examples introduced in
Section 2. In
Section 5, to illustrate an application of the abstract theorem, we develop a monotone method motivated by the abstract theorem and apply the monotone method to a nonlinear problem related to one of the examples introduced in
Section 2. The monotone method closely models one that has been developed in [
24] with subtle differences in the convergence argument. In
Section 6, we illustrate the monotone method with a specific example. In this example, a Green’s function is constructed using Mittag-Leffler functions. The purpose of introducing the Green’s function is not to produce an explicit function on which to analyze sign properties, as is the case in say, [
2] or [
3]; the purpose is to obtain a verifiable bound on
so that if
then
implies
2. Preliminaries
In this section, we introduce notations from fractional calculus and state common properties that we shall employ throughout. For authoritative accounts on the development of fractional calculus, we refer to the monographs [
25,
26,
27].
Assume
For
the space of Lebesgue integrable functions, a Riemann–Liouville fractional integral of
y of order
is defined by
where
denotes the special gamma function. For
is defined to be the identity operator. Let
n denote a positive integer and assume
A Riemann–Liouville fractional derivative of
y of order
is defined by
, where
if this expression exists. In the case
is a positive integer, we may write
or
since the Riemann–Liouville derivative or integral agrees with the classical derivative or integral if
is a positive integer.
For the sake of self-containment, we state properties that we shall employ in this study. It is well known that the Riemann–Liouville fractional integrals commute; that is, if
and
then
A power rule is valid for the Riemann–Liouville fractional integral; if
and
then
A power rule is valid for the Riemann–Liouville fractional derivative; if
and
, then
If
and if
exists, then
exists and
Thus, it is clear that for each
exists and
A Green’s function will be constructed in
Section 6. The two-parameter Mittag-Leffler function
will be employed in those calculations. Many properties and identities for the two-parameter Mittag-Leffler are derived in [
26].
In [
24], a boundary value problem,
was studied. This is an example of a boundary value problem at resonance since
the linear span of
, denotes the solution space of the homogeneous problem,
, with the given homogeneous boundary conditions; moreover,
is a simple eigenvalue of the homogeneous problem. There, an abstract theorem was proved that gave the existence of
and a constant
, independent of
f, such that
where
y is the unique solution associated with
Thus,
implies
. It was also proved in [
24] that
implies
Thus, with control of the sign of both
and
y, a monotone method was developed to obtain sufficient conditions for a solution of the nonlinear problem,
Since the purpose of this study is to modify the methods developed in [
9] to apply to Neumann-like or periodic-like boundary conditions, we shall focus on a differential equation,
where
is an integer.
Consider the fractional differential equation To study the Neumann-like boundary conditions, assume
Consider the fractional differential equation
We shall refer to the boundary conditions
as Neumann boundary conditions. The first exhibited boundary value problem is the boundary value problem, (
3), (
4).
To study periodic-like boundary conditions we shall consider a fractional differential equation
or
In the second exhibited example, we study the boundary value problem, (
5), with boundary conditions
in the third exhibited example, we study the boundary value problem, (
6), with the boundary conditions
and in the final exhibited boundary value problem we study the boundary value problem, (
6), with the boundary conditions
3. The Abstract Theorem
Let
denote the Banach space of continuous functions defined on
with the supremum norm,
and let
denote the space of Lebesgue integrable functions with the usual
norm. Let
denote an integer. Assume
Employing notation introduced in [
28], define
It is clear that
if, and only if, there exists
such that
for
Define
and
with norm
is a Banach space.
The following definition is motivated by Definition 1 found in [
9].
Definition 1. Assume is a linear operator with and Im . For the operator , where denotes the identity operator, satisfies a signed maximum principle in if for each , the equationhas unique solution y, and implies The operator satisfies a strong signed maximum principle in if and a.e. implies Remark 1. In [9], the authors employed the phrase, maximum principle. We have taken the liberty to employ the phrase signed maximum principle to distinguish further from classical usage of maximum principle or anti-maximum principle. Remark 2. The phrases “maximum principle” or “anti-maximum principle” are used loosely and we mean the following. Maximum principle means implies This is precisely the case for the classical second order ordinary differential equation with Dirichlet boundary conditions. Anti-maximum principle means implies This is the case observed in [8] for , where the phrase anti-maximum principle was coined. For
(or
), let
and define
Define
Assume
denotes a linear operator satisfying
where
denotes the linear span of
Assume further that for
the problem
is uniquely solvable with solution
Dom
and such that
In particular, define
and then
is invertible. Moreover, if
for
Dom
, assume there exists a constant
depending only on
such that
For
, define
which implies
and for
Dom
define
which implies
Dom
Since Ker
, with the decompositions
and
, it follows that
which decouples as follows:
Denote the inverse of
, if it exists, by
and denote the inverse of
by
. So,
and
Note that (
15) implies that since
Dom
,
Note that (
11) implies that
is continuous, and hence,
is a bounded linear operator with
To note the continuity, if
and
as
then
as
Since
, we can also consider
Equation (
11) also implies that
is continuous and hence, bounded. To see this, assume
as
Then,
uniformly as
For each
and
, eventually; in particular,
as
, which implies
as
Theorem 1. Assume denotes a linear operator satisfying (9). Define by (10) and assumeis invertible. Finally, if for Dom, assume there exists a constant depending only on such that (11) is satisfied. Then there exists such that if then , the inverse of , exists. Moreover, if , if where denotes the inverse of and if thenFurther, there exists such that if then the operator satisfies a strong signed maximum principle in Proof. Employ (
16) and apply
to (
13) to obtain
It has been established that (
11) implies that each of
and
are bounded linear operators. Since
it follows that
is invertible and
Assume
and assume
Then,
exists. Since
and
it follows that
and so the triangle inequality implies
Thus, (
17) is proved since
Now assume
and assume
a.e. Then,
Let
, write
, and consider
Note that
since
Thus,
Continuing to assume that
it now follows from (
17) that
Since
, and
the theorem is proved with
In particular, if
then
□
4. Four Examples
To apply Theorem 1, there are two primary tasks. First, if , we must show there exists a unique solution Dom of satisfying . In the case of ordinary differential equations or partial differential equations, one can often appeal to a Fredholm alternative to complete this task. For the Riemann–Liouville fractional differential equation, we only know to construct explicitly, and show uniqueness to complete this task. Second, we must show the existence of a constant such that This will be a straightforward task since we will have constructed explicitly.
Example 1. Let and consider the linear boundary value problem, with a Riemann–Liouville analogue of Neumann boundary conditions, (3), (4); that is, consider,For the boundary value problem (3), (4), and Ker We show that the operator satisfies the hypotheses of Theorem 1. One can show directly that Im If Im, then there exists a solution y ofwhich impliesand Likewise, if thenis a solution ofand To verify that satisfies these properties, note that any solution of , , has the form, Thus, To see that the boundary conditions are satisfied, writeand note that Thus, and since ; in particular, the boundary conditions are satisfied. To see that , note thatand so, To argue that is uniquely solvable with solution Dom , (18) implies the solvability. For uniqueness, if and are two such solutions, then and implies Finally, (18) implies (11) is satisfied with Theorem 1 applies and there exists such that if then satisfies a signed maximum principle in that is, implies
Example 2. For the second example, let and let Consider the linear boundary value problem, with a Riemann–Liouville analogue of periodic boundary conditions, (5), (7); that is, consider,Now, and Ker We show that the operator satisfies the hypotheses of Theorem 1. We show directly that Im If Im , thenthus, since y satisfies the periodic boundary conditions. In particular, Now assume We first construct a general solution ofSince , apply an integrating factor, andwhich impliesThen, Apply the periodic boundary conditions. Then,and the boundary condition impliesis uniquely determined. Now,Thus, andAt this point in the construction, c is still undetermined andis a general solution of To obtain the parameter c uniquely, Theorem 1 requires that Thus,andis uniquely determined. Note thatThus,implies (11) is satisfied. This concludes the second example.
Before proceeding to the third example, we observe that Theorem 1 does not apparently apply to a Neumann boundary value problem (5), (4) in the case . Assume and begin the construction of a general solution. As before, one obtainsTake for example, Then, Example 3. For the third example, let let and consider the linear boundary value problem, with a Riemann–Liouville analogue of periodic boundary conditions, (6), (7); that is, consider, For the boundary value problem (6), (7), and Ker Again, we show Im First, note that if the boundary value problem (6), (7) is solvable, then the boundary condition implies since Thus, Now assume If Dom , thenWe show the coefficients are uniquely determined. The condition implieswhich implies is undetermined at this point in the construction. Let Then,Apply the boundary conditions in the order At Thus, is uniquely determined. Employ (19) inductively and for Inductively, , have been uniquely determined and so,is uniquely determined. To summarize, the boundary conditions uniquely determine the coefficients, To determine the coefficient, employ the boundary condition Sinceit follows thatis uniquely determined. Finally, the application of Theorem 1 requires that Thus,Hence, is uniquely determined and the proof that implies Dom is uniquely determined is complete. To see that M in (11) can be computed, recall thatand employ (20) and (21). Note that is a multiple of , which implies that is a linear combination of for Thus, M is computable. Thus, Theorem 1 applies and there exists such that if , then satisfies the strong signed maximum principle in Example 4. Theorem 1 can also apply to the boundary value problem with boundary conditions analogous to periodic boundary conditions, (6), (8); that is, consider, The unique determination of proceeds precisely as in Example (3). Apply the boundary condition to to obtainand is uniquely determined. Then, as in Example 3, is uniquely determined by the requirement that Thus, Theorem 1 applies and there exists such that if then satisfies the strong signed maximum principle in
5. A Monotone Method
The application of monotone methods in the presence of a maximum principle or in the presence of an anti-maximum principle to construct approximate solutions of initial value or boundary value type problems enjoys a long history. The purpose of this section is to employ (
1) to quickly recognize the presence of the maximum principle or the anti-maximum principle. There are recent applications of monotone methods to periodic-like boundary value problems for Riemann–Liouville fractional differential equations; see, for example, [
6,
7]. In each of those application,
, and the anti-maximum principle is observed by the explicit construction of a corresponding Green’s function in terms of Mittag-Leffler functions.
Assume
is continuous and consider the boundary value problem
Assume that
and assume further that
f satisfies the following monotonicity property,
Thus,
f is monotone decreasing in the second component.
Apply Theorem 1 and find
such that if
then
satisfies a strong signed maximum principle in
Apply a shift [
29] to (
22) and consider the equivalent boundary value problem,
with boundary conditions (
23) where
and
is shown to exist in Theorem 1. Note that if
and
f satisfies (
24) and (
25), then
g satisfies (
24) and
g satisfies (
25) if
Assume the existence of solutions,
, of the following boundary value problems for differential inequalities
Assume further that
Since
, define a partial order
on
by
Then, the assumption (
27) implies
Define iteratively the sequences
,
, where
and
Inductively, Theorem 1 implies the existence of each
,
since
implies the inverse of
exists, and, for example,
Theorem 2. Assume is continuous, assume that f satisfies (24), and assume f satisfies the monotonicity properties (25). Assume the existence of functions satisfying (26) and (27). Define the sequences of iterates , by (28) and (29), respectively. Then, for each positive integer Moreover, converges in to a solution of the boundary value problem (22), (23) and converges in to a solution of the boundary value problem (22), (23) satisfying Proof. Since
satisfies a differential inequality given in (
27), then for
Set
and
u satisfies a boundary value problem for a differential inequality,
The signed maximum principle applies and
; in particular,
Similarly,
Now set
and
Since
f satisfies (
25) and
then
and again the signed maximum principle applies and
. In particular,
Thus, (
30) is proved for
Before applying a straightforward induction to obtain (
30), we must show
and
for
Since
and
, it follows that
Similarly,
and (
30) is valid.
To obtain the existence of limiting solutions
v and
w satisfying (
31), note that the sequence
is monotone decreasing and bounded below by
. Thus, the sequence
is converging pointwise to some
for each
. Moreover, if
the sequence
is converging pointwise to some
where
converges monotonically to
. At this point in the argument, the convergence is pointwise. Since
if follows that
is converging pointwise to
for each
Since
Thus, by the dominated convergence theorem
in particular,
and
v satisfies the fractional differential equation. To see that
v satisfies the Neumann type boundary conditions, again observe
Since
, it follows that
Again, the dominated convergence theorem implies that
. Thus,
which implies
and
Note that since
on
and
, then
is uniformly continuous on any compact subinterval of
Thus,
implies
and
for each
Moreover, Dini’s theorem now applies and the convergence of
is uniform.
Similar details apply to and the theorem is proved. □
Suppose now
f satisfies the “anti”-inequalities to (
25); that is, suppose
f satisfies
One can appeal to the signed maximum principle, apply a shift to (
22), and consider the equivalent boundary value problem,
where
and
is given by Theorem 1. Note, if
f satisfies (
32) and
then
satisfies (
32).
Now, assume the existence of solutions,
, of the following differential inequalities
Assume further that
Noting that
defines a partial order
on
by
In particular, in (
34), assume
Theorem 3. Assume is continuous, assume that f satisfies (24), and assume f satisfies the monotonicity properties, (32). Assume the existence of satisfying (33) and (34). Define the sequences of iterates , by (28) and (29), respectively. Then, for each positive integer Moreover, converges in to a solution of the boundary value problem (22), (23) and converges in to a solution of the boundary value problem (22), (23) satisfying 6. Example
We close the article with an example in which Theorem 3 applies and in which upper and lower solutions,
and
are explicitly produced. To do so, we construct an explicit Green’s function to obtain an estimate on
and we exhibit verifiable conditions on
f so that (
24) is satisfied.
The two-parameter Mittag-Leffler function
will be employed to construct an appropriate Green’s function.
Assume
, assume
, and consider a Neumann boundary value problem for nonhomogenous linear Equations (
3) and (
4). We restate the boundary value problem for convenience.
Thus,
where
c is still undetermined or
Employ the Neumann series to see that if
, then
Thus,
where
Employ the identity
and note that
To calculate
we have
and
Thus,
Employ the boundary condition
and obtain
if
The solution
y in (
36) satisfies
or
Define
where an identity
has been employed. Then,
where
One can see from this construction that a maximum principle will be valid for
For the anti-maximum principle, it is shown in ([
30], Corollary 3) that
has the smallest in modulus root which is a positive root. From the identity,
and integrating from 0 to 1, it is clear that
has the smallest positive root which is smaller than the smallest root of
Then, the identity
implies that
has the smallest positive root which is smaller than the smallest positive root of
Thus, from the construction, an anti-maximum principle will be valid for
where
is the smallest positive real root of the Mittag-Leffler function,
Now, consider a boundary value problem for nonlinear fractional differential Equations (
22) and (
23). Assume
is continuous, assume
f satisfies the monotonicity property (
25), and assume there exists
such that
and
is bounded and continuous on
Then,
f satisfies (
24).
Corollary 1. Assume Assume is continuous, and assume f satisfies the monotonicity property (25). Assume there exists such that and is bounded and continuous on Then, there exists a solution of the boundary value problem Proof. As noted above, the boundedness condition on
g implies that
f satisfies (
24). Let
denote an upper bound on
Set
and set
Thus,
and
satisfy the boundary conditions (
4). Moreover,
or
Similarly,
and Theorem 2 applies. □
Corollary 2. Assume Assume is continuous and assume f satisfies the monotonicity property (32). Let denote the smallest positive real root of Assume there exists such that and is bounded and continuous on Then, there exists a solution of the boundary value problem Proof. Let
denote an upper bound on
Set
and set
and
satisfy (
33) and Theorem 3 applies. □