Over the past few decades, the fractional calculus made great progress, and it was widely used in various fields of science and engineering. There were numbers applications in electromagnetics, control theory, viscoelasticity and so on. There was a high-speed development in fractional differential equations in recent years, and we referred the reader to the monographs Podlubny [1], Kilbas et al. [2] and Zhou [3]. In the current theory of fractional differential equations, much of the work is based on Riemann-Liouville and Caputo fractional derivatives, but the research of Caputo-Hadamard fractional derivatives of differential equations is very few, which includes logarithmic function and arbitrary exponents. Motivated by this fact, we consider a class of Caputo-Hadamard fractional differential equations with boundary value problems (BVPs).
Nowadays, some authors studied the existence and uniqueness of solutions for nonlinear fractional differential equation with boundary value problems. For the recent development of the topic, we referred the reader to a series papers by Ahmad et al. [4-6], Mahmudov et al. [7] and the references therein. Details and properties of the Hadamard fractional derivative and integral can be found in [8-12].
Wafa Shammakh [13] studied the existence and uniqueness results for the following three-point BVPs
where $ {}_H^CD^{\alpha} $ is the Caputo-Hadamard fractional derivative of order $ 1\leq \alpha\leq 2 $, $ 0\leq\gamma<1, $ $ \xi\in(1,e), $ $ {}_H^CD = t\frac{d}{dt} $, and $ f:[1,e]\rightarrow[0,\infty). $
Yacine Arioua and Nouredine Benhamidouche [14] studied the existence of solutions for the following BVPs of nonlinear fractional differential equations
where $ {}_H^CD_{1+}^{\alpha} $ is the Caputo-Hadamard fractional derivative of order $ \alpha $, and $ f:[1,e]\times{\mathbb{R}}\rightarrow{\mathbb{R}} $ is a given continuous function.
Yunru Bai and Hua Kong [15] used the method of upper and lower solutions, proved the existence of solutions to nonlinear Caputo-Hadamard fractional differential equations
where $ {}_H^CD_{a+}^\alpha $ and $ {}^HI_{a+}^\alpha $ stand for the Caputo-Hadamard fractional derivative and Hadamard integral operators, $ f:[a,b]\times{\mathbb{R}}\times{\mathbb{R}}\rightarrow{\mathbb{R}} $ and $ 1<a<b<\infty $.
The purpose of this paper is to discuss the existence and uniqueness of solutions for nonlinear Caputo-Hadamard fractional differential equations
where $ {}_H^CD^{\alpha}_{1+} $ is the Caputo-Hadamard fractional derivative of order $ 2<\alpha\leq3 $, and $ f $ is a continuous function.
In this section, we introduce some necessary definitions, lemmas and notations that will be used later.
Definition 2.1 [2] The Hadamard fractional integral of order $ \alpha\in\mathbb{R}_+ $ for a continuous function $ g:[1,\infty)\rightarrow{\mathbb{R}} $ is given by
where $ \Gamma(\cdot) $ stands for the Gamma function.
Definition 2.2 [2] The Hadamard fractional derivative of order $ \alpha\in\mathbb{R}_+ $ for a continuous function $ g:[1,\infty)\rightarrow{\mathbb{R}} $ is given by
where $ n-1<\alpha<n, n = [\alpha]+1, \delta = t\frac{d}{dt} $, and $ [\alpha] $ denotes the integer part of the real number $ \alpha $.
Definition 2.3 [16, 17] The Caputo-Hadamard fractional derivative of order $ \alpha\in\mathbb{R}_+ $ for at least $ n $-times differentiable function $ g:[1,\infty)\rightarrow{\mathbb{R}} $ is defined as
Lemma 2.4 [16, 17] Let $ u\in C_{\delta}^n([1,e],\mathbb{R}) $, then
here $ C_{\delta}^n([1,e],\mathbb{R}) = \{u:[1,e]\rightarrow{\mathbb{R}}:\delta^{n-1}u\in C([1,e],\mathbb{R})\} $.
Lemma 2.5 Let $ h\in C([1,e],\mathbb{R}),u\in C_\delta^3([1,e],\mathbb{R}) $. Then the unique solution of the linear Caputo-Hadamard fractional differential equation
is equivalent to the following integral equation
where $ A = \int_1^e(\ln t)^2dt = (e-2) $.
Proof In view of Lemma 2.4, applying $ {}^HI^{\alpha}_{1+} $ to both sides of (2.2),
where $ c_0, c_1, c_2\in\mathbb{R} $.
The boundary condition $ u(1) = u'(1) = 0 $ implies that $ c_0 = c_1 = 0 $. Thus
In view of the boundary condition $ u(e) = \lambda \int_1^eu(s)ds $, we conclude that
Substituting (2.5) in (2.4), we obtain (2.3). This completes the proof.
Based on Lemma 2.5, the solution of problems (1.1)–(1.2) can be expressed as
Let $ E: = C([1,e],\mathbb{R}) $ be the Banach space of all continuous functions from $ [1,e] $ to $ \mathbb{R} $ with the norm $ \|u\| = \max\limits_{t\in[1,e]}|u(t)| $. Due to Lemma 2.5, we define an operator $ \mathbb{A}:E\rightarrow E $ as
It should be noticed that BVPs (1.1) has solutions if and only if the operator $ \mathbb{A} $ has fixed points.
First, we obtain the existence and uniqueness results via Banach fixed point theorem.
Theorem 3.1 Assume that $ f:[1,e]\times\mathbb{R}\rightarrow\mathbb{R} $ is a continuous function, and there exists a constant $ L>0 $ such that
(H1) $ |f(t,u)-f(t,v)|\leq L|u-v|, \forall\; t\in [1,e], u, v\in\mathbb{R} $. If
then problem (1.1) has a unique solution on $ [1,e] $.
Proof Denote $ Q = \frac{1}{\Gamma(\alpha+1)}+\frac{1+\lambda(e-1)}{\Gamma(\alpha+1)|1-\lambda A|} $, we set $ B_r: = \{u\in C([1,e],\mathbb{R}):\|u\|\leq r\} $ and choose $ r\geq\frac{MQ}{1-LQ} $, where $ M = \max\limits_{t\in[1,e]}|f(t,0)|<\infty $.
Obviously it is concluded that
Now, we show that $ \mathbb{A}B_r\subseteq B_r $. For any $ u\in B_r,t\in [1,e] $, we have
which implies that $ \mathbb{A}B_r\subseteq B_r $. Let $ u,v\in B_r $, and for each $ t\in[1,e] $, we have
Therefore,
From assumption (3.2), it follows that $ \mathbb{A} $ is a contraction mapping. Hence problem (1.1) has a unique solution by using Banach fixed point theorem. This completes the proof.
Next, we will use the method of upper and lower solutions to obtain the existence result of BVPs (1.1).
Definition 3.2 Functions $ \overline{u},\underline{u}\in C([1,e],\mathbb R) $ are called upper and lower solutions of fractional integral equation (2.6), respactively, if it satisfies for any $ \; t\in[1,e]\; $,
Define
Theorem 3.3 Let $ f\in C([1,e]\times\mathbb{R},\mathbb{R}) $. Assume that $ \overline{u},\underline{u}\in C([1,e],\mathbb R) $ are upper and lower solutions of fractional integral equation (2.6) with $ \underline{u}(t)\leq\overline{u}(t) $ for $ t\in[1,e] $. If $ f $ is nondecreasing with respect to $ u $ that is $ \begin{aligned} f(t,u_1)\leq f(t,u_2),\ u_1\leq u_2, \end{aligned} $ then there exist maximal and minimal solutions $ u_M,u_L \in X_{(\underline{u},\overline{u})} $ in $ X_{(\underline{u},\overline{u})} $, moreover, for each $ u\in X_{(\underline{u},\overline{u})} $, one has
Proof Constructing two sequences $ \{p_n\},\{q_n\} $ as follows
This proof divides into three steps.
Step 1 Finding the monotonicity of the two sequences, that is, the sequences $ \{p_n\},\{q_n\} $ satisfy the following relation
for $ t\in[1,e] $.
First, we verify that the sequence $ \{p_n\} $ is nondecreasing and satisfies
Since $ \overline{u},\underline{u} $ are upper and lower solutions respectively, we know that $ \underline{u}(t) = p_0(t)\leq\overline{u}(t) = q_0(t) $ for $ t\in[1,e] $,
Since $ f $ is nondecreasing respect to the second variable, this implies that
This deduces
Therefore, we assume inductively
In view of definition of $ \{p_n\},\{q_n\} $, we have
By means of the monotonicity of $ f $, it is obvious that
We show that
For $ n = 0 $, it is obvious that $ \underline{u}(t) = p_0(t)\leq\overline{u}(t) = q_0(t) $ for all $ t\in[1,e] $. Now, we also suppose inductively
Analogously, we easily conclude from the monotonicity of $ f $ with respect to the second variables that
In a similar way, we know that the sequence $ \{q_n\} $ is nonincreasing.
Step 2 The sequences constructed by (3.3), (3.4) are both relatively compact in $ C([1,e],\mathbb R) $.
According to that $ f $ is continuous and $ \overline{u},\underline{u}\in C([1,e],\mathbb R) $, from Step 1, we have $ \{p_n\} $ and $ \{q_n\} $ also belong to $ C([1,e],\mathbb R) $. Moreover, it follows from (3.5) that $ \{p_n\} $ and $ \{q_n\} $ are uniformly bounded. For any $ t_1,t_2\in[1,e] $, without loss of generality, let $ t_1\leq t_2 $, we know that
approaches zero as $ t_2-t_1\rightarrow 0 $, where $ W>0 $ is a constant independent of $ n $, $ t_1 $ and $ t_2 $, $ |f(t,p_n(t))|\leq W $. It implies that $ \{p_n\} $ is equicontinuous in $ C([1,e],\mathbb R) $. By Arzelà-Ascoli theorem, we imply that $ \{p_n\} $ is relatively compact in $ C([1,e],\mathbb R) $. In the same way, we conclude that $ \{q_n\} $ is also relatively compact in $ C([1,e],\mathbb R) $.
Step 3 There exist maximal and minimal solutions in $ X_{(\underline{u},\overline{u})} $.
The sequences $ \{p_n\} $ and $ \{q_n\} $ are both monotone and relatively compact in $ C([1,e],\mathbb R) $ by Step 1 and Step 2. There exist continuous functions $ p $ and $ q $ such that $ p_n(t)\leq p(t ) \leq q(t)\leq q_n(t) $ for all $ t\in[1,e] $ and $ n\in\mathbb N $. $ \{p_n\} $ and $ \{q_n\} $ converge uniformly to $ p $ and $ q $ in $ C([1,e],\mathbb R) $, severally. Therefore, $ p $ and $ q $ are two solutions of (2.6), i.e.,
for $ t\in[1,e] $. However, fact (3.5) determines that
Finally, we shall show that $ p $ and $ q $ are the minimal and maximal solutions in $ X_{(\underline{u},\overline{u})} $, respectively. For any $ u\in X_{(\underline{u},\overline{u})} $, then we have
Because $ f $ is nondecreasing with respect to the second parameter, we conclude
Taking limits as $ n\rightarrow\infty $ into the above inequality, we have
which means that $ u_L = p $ and $ u_M = q $ are the minimal and maximal solutions in $ X_{(\underline{u},\overline{u})} $. This completes the proof.
Theorem 3.4 Assume that assumptions of Theorem 3.3 are satisfied. Then fractional nonlinear differential equation (1.1) has at least one solution in $ C([1,e],\mathbb R) $.
Proof By the hypotheses and Theorem 3.3, we induct $ X_{(\underline{u},\overline{u})}\neq\emptyset $, then the solution set of fractional integral equation (2.6) is nonempty in $ C([1,e],\mathbb R) $. It follows from the solution set of (2.6) together with Lemma 2.5 that problem (1.1) has at least one solution in $ C([1,e],\mathbb R) $. This completes the proof.
In this section, we present two examples to explain our main results.
Example 1 Consider the following nonlinear Caputo-Hadamard fractional differential equation
here $ \alpha = \frac{5}{2}, $ $ \lambda = 1, $ $ f(t,u) = \frac{(\sqrt{t}+\ln t)}{(t+3)^3}\frac{u^2}{|u|+1}, $ $ 1\leq t\leq e. $ One can easily calculate $ Q = \frac{24}{15\sqrt{\pi}(3-e)}\approx3.2 $. Clearly $ f $ is a continuous function and we have
Therefore $ \begin{aligned} LQ<1. \end{aligned} $ Thus all conditions of Theorem 3.1 satisfy which implies the existence of uniqueness solution of the the boundary value problem (4.1).
Example 2 Consider the problem
Proof Where $ \alpha = \frac{5}{2}, $ $ \lambda = \frac{1}{(6-2e)}, $ $ f(t,u) = \frac{t^4}{16\sqrt{\pi}}(|u|+1), $ $ t\in[1,e] $, $ f $ is continuous and nondecreasing with respect to $ u $. Thus
It is easy to check that $ (\underline{u}(t),\overline{u}(t)) = (0,(\ln t)^3) $ is a pair of upper and lower solutions of (4.3) and that all assumptions of Theorem 3.2 are satisfied. So $ u_L = p $ and $ u_M = q $ are the minimal and maximal solutions of the boundary problem (4.3), and the iteration sequences is as follows
$ \lim\limits_{n\rightarrow\infty}p_n = p $ and $ \lim\limits_{n\rightarrow\infty}q_n = q $. Applying Theorem 3.4, this boundary value problem (4.2) has at least one solution.