Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article Special Issues

Impact of online public opinion regarding the Japanese nuclear wastewater incident on stock market based on the SOR model


  • Received: 11 January 2023 Revised: 02 March 2023 Accepted: 07 March 2023 Published: 16 March 2023
  • The exposure of the Japanese nuclear wastewater incident has shaped online public opinion and has also caused a certain impact on stocks in aquaculture and feed industries. In order to explore the impact of network public opinion caused by public emergencies on relevant stocks, this paper uses the stimulus organism response(SOR) model to construct a framework model of the impact path of network public opinion on the financial stock market, and it uses emotional analysis, LDA and grounded theory methods to conduct empirical analysis. The study draws a new conclusion about the impact of online public opinion on the performance of relevant stocks in the context of the nuclear waste water incident in Japan. The positive change of media sentiment will lead to the decline of stock returns and the increase of volatility. The positive change of public sentiment will lead to the decline of stock returns in the current period and the increase of stock returns in the lag period. At the same time, we have proved that media attention, public opinion theme and prospect theory value have certain influences on stock performance in the context of the Japanese nuclear wastewater incident. The conclusion shows that after the public emergency, the government and investors need to pay attention to the changes of network public opinion caused by the event, so as to avoid the possible stock market risks.

    Citation: Wei Hong, Yiting Gu, Linhai Wu, Xujin Pu. Impact of online public opinion regarding the Japanese nuclear wastewater incident on stock market based on the SOR model[J]. Mathematical Biosciences and Engineering, 2023, 20(5): 9305-9326. doi: 10.3934/mbe.2023408

    Related Papers:

    [1] Huawei Huang, Xin Jiang, Changwen Peng, Geyang Pan . A new semiring and its cryptographic applications. AIMS Mathematics, 2024, 9(8): 20677-20691. doi: 10.3934/math.20241005
    [2] Hailun Wang, Fei Wu, Dongge Lei . A novel numerical approach for solving fractional order differential equations using hybrid functions. AIMS Mathematics, 2021, 6(6): 5596-5611. doi: 10.3934/math.2021331
    [3] Shousheng Zhu . Double iterative algorithm for solving different constrained solutions of multivariate quadratic matrix equations. AIMS Mathematics, 2022, 7(2): 1845-1855. doi: 10.3934/math.2022106
    [4] Miloud Sadkane, Roger Sidje . Computing a canonical form of a matrix pencil. AIMS Mathematics, 2024, 9(5): 10882-10892. doi: 10.3934/math.2024531
    [5] Zahra Pirouzeh, Mohammad Hadi Noori Skandari, Kamele Nassiri Pirbazari, Stanford Shateyi . A pseudo-spectral approach for optimal control problems of variable-order fractional integro-differential equations. AIMS Mathematics, 2024, 9(9): 23692-23710. doi: 10.3934/math.20241151
    [6] Yinlan Chen, Min Zeng, Ranran Fan, Yongxin Yuan . The solutions of two classes of dual matrix equations. AIMS Mathematics, 2023, 8(10): 23016-23031. doi: 10.3934/math.20231171
    [7] Waleed Mohamed Abd-Elhameed, Abdullah F. Abu Sunayh, Mohammed H. Alharbi, Ahmed Gamal Atta . Spectral tau technique via Lucas polynomials for the time-fractional diffusion equation. AIMS Mathematics, 2024, 9(12): 34567-34587. doi: 10.3934/math.20241646
    [8] Emrah Polatlı . On some properties of a generalized min matrix. AIMS Mathematics, 2023, 8(11): 26199-26212. doi: 10.3934/math.20231336
    [9] Jin Li . Barycentric rational collocation method for semi-infinite domain problems. AIMS Mathematics, 2023, 8(4): 8756-8771. doi: 10.3934/math.2023439
    [10] Samia Bushnaq, Kamal Shah, Sana Tahir, Khursheed J. Ansari, Muhammad Sarwar, Thabet Abdeljawad . Computation of numerical solutions to variable order fractional differential equations by using non-orthogonal basis. AIMS Mathematics, 2022, 7(6): 10917-10938. doi: 10.3934/math.2022610
  • The exposure of the Japanese nuclear wastewater incident has shaped online public opinion and has also caused a certain impact on stocks in aquaculture and feed industries. In order to explore the impact of network public opinion caused by public emergencies on relevant stocks, this paper uses the stimulus organism response(SOR) model to construct a framework model of the impact path of network public opinion on the financial stock market, and it uses emotional analysis, LDA and grounded theory methods to conduct empirical analysis. The study draws a new conclusion about the impact of online public opinion on the performance of relevant stocks in the context of the nuclear waste water incident in Japan. The positive change of media sentiment will lead to the decline of stock returns and the increase of volatility. The positive change of public sentiment will lead to the decline of stock returns in the current period and the increase of stock returns in the lag period. At the same time, we have proved that media attention, public opinion theme and prospect theory value have certain influences on stock performance in the context of the Japanese nuclear wastewater incident. The conclusion shows that after the public emergency, the government and investors need to pay attention to the changes of network public opinion caused by the event, so as to avoid the possible stock market risks.



    The importance of existence and uniqueness theorems for initial value and boundary value problems (IVP and BVP) involving the classical derivative operator is indisputable because, without them, one cannot understand modeled systems correctly and make predictions how they behave. Recently, with the popularity of fractional derivative operators such as Riemann-Liouville (R-L), Caputo (C), etc., the equations involving these operators have begun to be studied in detail (See, [1,2,3,4,5,6]). However, such a generalization leads to some difficulties and differences especially in R-L case. For instance, unlike the initial value problems involving the classical derivative, the existence of continuous solution to some IVPs in the sense of R-L derivative strictly depends on the initial values and smoothness conditions on nonlinear functions in right-hand side of equations in IVPs. For example, as shown in [7], the following initial value problem for R-L derivative of order σ(0,1)

    {Dσω(x)=f(x,ω(x)),x>0ω(0)=b0 (1.1)

    has no continuous solution when f is continuous on [0,T]×R. Therefore, this equation was investigated when subjected to different initial values (See, Theorem 3.3 in [1]). By considering the aforementioned remark, we investigate the existence and uniqueness of solutions to the following problem

    {Dσω(x)=f(x,ω(x),Dσ1ω(x)),x>0ω(0)=0,Dσ1ω(x)|x=0=b, (1.2)

    where σ(1,2), bR, Dσ represents the Riemann-Liouville fractional derivative of order σ, which is given by

    Dσω(x)=1Γ(2σ)d2dx2x0ω(t)(xt)σ1dt,

    and f fulfills the following condition:

    (C1) Let f(x,t1,t2)C((0,T]×R×R) and xσ1f(x,t1,t2)C([0,T]×R×R),

    where C(X) represents the class of continuous functions defined on the subspace X of R3.

    The equation in (1.2) was first considered by Yoruk et. al. [8], when the second initial value is also homogeneous (b=0) and the right-hand side function is continuous on [0,T]×R×R. They gave Krasnoselskii-Krein, Roger and Kooi-type uniqueness results. As seen from condition (C1), the equation in (1.2) has a singularity at x=0. Such singular equations involving R-L and Caputo derivatives were recently investigated in [7,9,10,11,12] and they proved local or global existence theorem for initial value problems involving singular equations. In the investigation of existence of solutions to these problems, converting the problem into the Volterra-type integral equation is one of the referenced tools. These integral equations have weakly and double singularities since the corresponding fractional differential equations are singular. For this reason, some new techniques or lemmas had to be developed to reveal the existence and uniqueness of solutions to integral equations (See for example [10,11]). In this paper, we also encounter a Volterra-type integral equations having a single singularity when we transform the problem (1.2) into a integral equation. For the uniqueness of solutions to the integral equation we need to generalize the definitions and methods previously given in [7,8,13] and use the tools of Lebesgue spaces such as Hölder inequality. These uniqueness theorems are the type of Nagumo-type, Krasnoselskii-Krein and Osgood which are well-known in the literature (See, [13,14,15,16]) Moreover, we give a Peano-type existence theorem for this problem as well.

    We first give definition of the space of functions where we investigate the existence of solutions to the problem (1.2). The space is defined below as in [17] :

    Theorem 2.1. The space of continuous functions defined on [0,T], whose R-L fractional derivative of order σ1, 1<σ<2 are continuous on [0,T] is a Banach space when endowed with the following norm:

    ||ω||σ1=||ω||+||Dσ1ω||,

    where ||.|| is the supremum norm defined on the class of continuous functions. This space will be denoted by Cσ1([0,T]).

    According to this space, we define R-L integral and R-L derivative of higher order [11]. The lower terminal points of integrals in their formulas will be taken as zero.

    Definition 2.1. The R-L integral of order σ>0 of a function ω(x)C0[0,T] is defined for all x[0,T] by

    Iσω(x):=1Γ(σ) x0ω(t)(xt)1σdt. (2.1)

    Definition 2.2. For 1<σ<2 and ω(x)Cσ1[0,T] with Dσ1ω(x)C1(0,T]L1[0,T] the R-L fractional derivative Dβω is defined for all x(0,T] by

    Dσω(x)=1Γ(2σ)d2dx2x0ω(t)(xt)σ1dt. (2.2)

    The local existence of solutions to problem (1.2) will be proved with the aid of Schauder fixed point theorem [18]:

    Theorem 2.2. Let C be a closed, bounded, convex subset of a Banach space X:={u:IR continuous:IR closed  and  bounded  interval}. If operator S:CC is continuous and, if S(C) is an equicontinuous set on I, then S has at least one fixed point in C.

    The one of mathematical tools used for showing the existence and uniqueness of the desired type of solution to a given initial or boundary value problem is first to convert them into an integral equation. One investigates the existence and uniqueness of the solution to the integral equation instead of the associated problem. Here, we follow this way by taking the aid of the lemma given below:

    Lemma 3.1. Under condition (C1), if ωCσ1[0,T] is a solution of problem (1.2), then ωCσ1[0,T] is a solution of the following integral equation

    ω(x)=bΓ(σ)xσ1+1Γ(σ)x0f(t,ω(t),Dσ1ω(t))(xt)1σdt (3.1)

    and, vice versa.

    Proof. We assume that ωCσ1[0,T] is a solution of problem (1.2). By condition (C1), we have f(x,ω(x),Dσ1ω(x)) is continuous on (0,T] and xσ1f(x,ω(x),Dσ1ω(x)) is continuous on [0,T]. It means that f(x,ω(x),Dσ1ω(x)) is integrable, i.e f(x,ω(x),Dσ1ω(x))C(0,T]L1[0,T]. Then, by integrating the both sides of the equation in (1.2) and using the relation IDσ=IDDσ1

    Dσ1ω(x)=Dσ1ω(0)+If(x,ω(x),Dσ1ω(x))C[0,T]

    is obtained. From here, by integration of the both sides of last equation and by use of IDσ1=IDI2σ we have

    I2σω(x)=bx+I2f(x,ω(x),Dσ1ω(x))C1[0,T] (3.2)

    where we used Dσ1ω(0)=b and I2σω(0)=0 since ω(0)=0.

    If the operator Iσ1 is applied to the both sides of (3.2), then by semigroup and commutative property of R-L derivative we get

    Iω(x)=bxσΓ(1+σ)+I[Iσf(x,ω(x),Dσ1ω(x))] (3.3)

    for all x[0,T]. Differentiating the both sides of (3.3), we have

    ω(x)=bxσ1Γ(σ)+Iσf(x,ω(x),Dσ1ω(x))

    which is the equivalent to the integral Eq (3.1).

    Now we suppose that ωCσ1[0,T] is a solution of integral Eq (3.1), and let us show that ω is a solution of the problem (1.2). If Dσ is applied to the both sides of (3.1), and then, if

    DσIσω(x)=ω(x)  for all  ωCσ1[0,T]

    is used, then one can observe that ωCσ1[0,T] satisfies the equation in (1.2). Moreover, let us prove that ωCσ1[0,T] also fulfills initial value conditions. By change of variables and condition (C1) we have

    ω(0)=limx0+ω(x)=bΓ(σ)xσ1+1Γ(σ)limx0+x0f(t,ω(t),Dσ1ω(t))(xt)1σdt=1Γ(σ)limx0+x0tσ1f(t,ω(t),Dσ1ω(t))tσ1(xt)1σdt=1Γ(σ)limx0+x10(xτ)σ1f(xτ,ω(xτ),Dσ1ω(xτ))τσ1(1τ)1σdτ=0, (3.4)

    since the integral is finite. Thus ω satisfies the first initial condition in (1.2).

    Now let us show that ω provides the second initial condition in (1.2). If Dσ1 is applied to both sides of (3.1), and if the relation Dσ1Iσh(x)=Ih(x) is used, then we can first get

    Dσ1ω(x)=b+x0f(t,ω(t),Dσ1ω(t))dt. (3.5)

    From here, by passing the limit as x0+, we then obtain

    Dσ1ω(0)=b+limx0+x0f(t,ω(t),Dσ1ω(t))dt=b+limx0+x01tσ1tσ1f(t,ω(t),Dσ1ω(t))dt=b+limx0+x2σ101τσ1(xτ)σ1f(xτ,ω(xτ),Dσ1ω(xτ))dτ=b, (3.6)

    since 2σ>0 and the integral is finite due to the continuity of tσ1f(t,ω(t),Dσ1ω(t)) on [0,T]. Consequently, it has been shown that any solution of (3.1) provides the problem (1.2) if condition (C1) is assumed to be satisfied.

    Theorem 3.1 (Existence). Let condition (C1) be satisfied, and assume that there exist positive real numbers r1, r2 and M such that |xσ1f(x,ω,v)|Mfor  all(x,ω,v)I=[0,T]×[r1,r1]×[br2,b+r2]. Then problem (1.2) admits at least one solution in Cσ1[0,T0 ], where

    T0 ={TifT<rC(b,σ,M)rC(b,σ,M)ifTrC(b,σ,M)1[rC(b,σ,M)]σ1,ifTrC(b,σ,M),1rC(b,σ,M)and1<σ1.5[rC(b,σ,M)]2σ,ifTrC(b,σ,M),1rC(b,σ,M)and1.5<σ<2, (3.7)

    and

    r=r1+r2andC(b,σ,M)=[|b|Γ(σ)+M(1+Γ(3σ)2σ)]. (3.8)

    Proof. As it is known from Lemma 3.1, solutions of problem (1.2) are solutions of integral equation (3.1) as well. Moreover, the fixed points of the opeator S:Cσ1[0,T0 ]Cσ1[0,T0 ] defined by

    Sω(x)=bΓ(σ)xσ1+x0f(t,ω(t),Dσ1ω(t))(xt)1σdt (3.9)

    interfere with solutions of the integral equation. For this reason, it is sufficient to prove that operator S admits at least one fixed point. For this, it will be verified that operator S satisfies the hypotheses of Schauder fixed-point theorem. Let us start with showing the following inclusion to be valid:

    S(Br)Br

    where

    Br={ωCσ1[0,T0 ]:||ω||+||Dσ1ωb||r}

    is a closed compact subset of Cσ1[0,T0 ]. Accordingly to the norm on Cσ1[0,T0 ], upper bounds of and \left\| D^{\sigma-1}\mathcal{S}\omega(x)-b\right\|_{\infty} can be determined as follows:

    \begin{align} \left|\mathcal{S}\omega(x)\right| &\leq \frac{\left|b\right| }{\Gamma(\sigma)}x^{\sigma-1}+\frac{1}{\Gamma(\sigma)}\int_{0}^{x}\frac{\left|t^{\sigma-1}f(t, \omega(t), D^{\sigma-1}\omega(t)\right|}{t^{\sigma-1}(x-t)^{1-\sigma }}dt \\ &\leq \frac{\left|b\right| }{\Gamma(\sigma)}x^{\sigma-1}+\frac{\mathcal{M}}{\Gamma(\sigma)}\int_{0}^{1}\frac{x}{\tau^{\sigma-1}(1-\tau)^{1-\sigma }}d\tau \leq \frac{\left|b\right| }{\Gamma(\sigma)}x^{\sigma-1}+\Gamma(2-\sigma)\mathcal{M}x \end{align} (3.10)

    and

    \begin{align} \left|D^{\sigma-1}\mathcal{S}\omega(x)-b\right| &\leq \int_{0}^{x}\frac{\left|t^{\sigma-1}f(t, \omega(t), D^{\sigma-1}\omega(t)\right|}{t^{\sigma-1}}dt\leq\mathcal{M}x^{2-\sigma} \int_{0}^{1}\tau^{1-\sigma}d\tau = \frac{\mathcal{M}x^{2-\sigma}}{2-\sigma}. \end{align} (3.11)

    From (3.10)) and (3.11),

    \begin{align} \left|\mathcal{S}\omega(x)\right|+\left|D^{\sigma-1}\mathcal{S}\omega(x)-b\right| \leq \frac{\left|b\right| }{\Gamma(\sigma)}x^{\sigma-1}+\Gamma(2-\sigma)\mathcal{M}x+ \frac{\mathcal{M}x^{2-\sigma}}{2-\sigma} \end{align} (3.12)

    is obtained. Taking supremum over [0, T_0] for a T_0 > 0 for the right hand-side of the above equation,

    \begin{align} \left|\mathcal{S}\omega(x)\right|+\left|D^{\sigma-1}\mathcal{S}\omega(x)-b\right| \leq C(b, \sigma, \mathcal{M}) T_0 ^{\alpha} \end{align} (3.13)

    can be written, where \alpha\in\Omega = \left\lbrace \sigma-1, 1, 2-\sigma\right\rbrace. \alpha depends on values of b, \mathcal{M}, \sigma, r. To determine T_0 and \alpha , let

    C(b, \sigma, \mathcal{M}) T_0 ^{\alpha} = r.

    If T_0 ^{\alpha} = \frac{r}{C(b, \sigma, \mathcal{M})} < 1, then it is observed that T_0 < 1 for any \alpha\in \Omega. If T_0 ^{\alpha} = \frac{r}{C(b, \sigma, \mathcal{M})}\geq1, it must be T_0 \geq 1 for any \alpha\in \Omega. Thus,

    \begin{align} \sup\limits_{x\in [0, T_0]}\left[ \left|\mathcal{S}\omega(x)\right|+\left|D^{\sigma-1}\mathcal{S}\omega(x)-b\right|\right] \leq C(b, \sigma, \mathcal{M}) T_0 ^{\alpha} = r, \end{align} (3.14)

    where

    T_0: = \left[\frac{r}{C(b, \sigma, \mathcal{M})}\right]^{1/\alpha}

    and

    \begin{equation} \alpha = \begin{cases} \quad 1 \quad &\text{if} \quad \frac{r}{C(b, \sigma, \mathcal{M})}\geq 1 \\ \quad \sigma-1\quad &\text{if} \quad \frac{r}{C(b, \sigma, \mathcal{M})} < 1 \quad \text{and} \quad 1 < \sigma\leq1.5\\ \quad 2-\sigma \quad &\text{if} \quad \frac{r}{C(b, \sigma, \mathcal{M})} < 1 \quad \text{and} \quad 1.5\leq\sigma < 2.\\ \end{cases} \end{equation} (3.15)

    As a result, for all cases we obtain

    ||\mathcal{S}\omega||_{\infty}+||D^{\sigma-1}\mathcal{S}\omega-b||_{\infty}\leq r,

    which is the desired.

    Now, let us prove the equicontinuity of \mathcal{S}(B_{r})\subset C^{\sigma-1} [0, T_{0}~]. Since the composition of uniformly continuous functions is so as well, the function x^{\sigma-1}f(x, \omega(x), D^{\sigma-1}\omega(x)) is uniformly continuous on [0, T_{0}~]. Because for any \omega\in B_{r}, both \omega(x) and D^{\sigma-1}\omega(x) and x^{\sigma-1}f(x, \omega, v) are uniformly continuous on I, respectively. Therefore, for given any \epsilon > 0, one can find a \delta = \delta(\epsilon) > 0 so that for all x_{1}, x_{2}\in [0, T_{0}~] with \left|x_{1}-x_{2}\right| < \delta it is

    \left \vert x_{1}^{\sigma-1}f(x_{1}, \omega(x_{1}), D^{\sigma-1}\omega(x ))-x_{2}^{\sigma-1}f(x_{2}, \omega(x_{2}), D^{\sigma-1}\omega(x_{2} ))\right \vert < K\epsilon ,

    where K = \max\left(\frac{1}{T_{0}~\Gamma(2-\sigma)}, \frac{2-\sigma}{T_{0}~^{2-\sigma}}\right). It follows that

    \begin{align*} \big|\mathcal{S}&\omega\left(x_{1}\right)-\mathcal{S}\omega\left(x_{2}\right) \big|+\big|D^{\sigma-1}\mathcal{S}\omega\left(x_{1}\right)-D^{\sigma-1}\mathcal{S}\omega\left(x_{2}\right) \big| \\ &\leq \int_{0}^{1} \frac{\left \vert h\left( \eta x_{1}\right) -h\left( \eta x_{2}\right) \right \vert}{\Gamma \left(\sigma\right)\eta^{1-\sigma} \left( 1-\eta \right) ^{\sigma-1}}x d\eta+\int_{0}^{1} \frac{\left \vert h\left( \eta x_{1}\right) -h\left( \eta x_{2}\right) \right \vert}{\eta^{1-\sigma} }x^{2-\sigma}d\eta\\ & < T_{0}~\Gamma(2-\sigma)K\epsilon+\frac{T_{0}~^{2-\sigma}}{2-\sigma}K\epsilon = \epsilon, \end{align*}

    where h(x) = x^{\sigma-1}f\left(x, \omega\left(x\right), D^{\sigma-1}\omega(x) \right). This implies that \mathcal{S}(B_{r}) is an equicontinuous set of C^{\sigma-1} [0, T_{0}~].

    Finally, the continuity of \mathcal{S} on B_{r} will be proven. Assume that \left \{\omega_{k}\right \}_{k = 1}^{\infty}\subset B_{r} is a sequence with \omega_{k}\stackrel{C^\sigma [0, T_{0}~]}{\rightarrow } \omega as k\rightarrow \infty. Then, one can easily conclude that \omega_{k} and D^{\sigma-1}\omega_{k}(t) converges uniformly to \omega and D^{\sigma-1}\omega(t), respectively. With these and the uniform continuity of x^{\sigma-1}f(x, \omega, v) on I = \left[0, T\right]\times\left[-r_{1}, r_{1}\right] \times \left[b-r_{2}, b+r_{2}\right], it leads to

    \begin{align*} &\left \|\mathcal{S}\omega_{k}-\mathcal{S}\omega\right \|_{\sigma-1} = \sup\limits_{x\in [0, T_{0}~]} \left \vert \frac{1}{\Gamma \left( \sigma\right)}\int_{0}^{x}\frac{ \left[f\left(t, \omega_{k}(t ), D^{\sigma-1}\omega_{k}(t )\right) -f\left(t , \omega(t), D^{\sigma-1}\omega(t )\right)\right] }{\left(x-t\right) ^{1-\sigma}}dt \right \vert\\ &+\sup\limits_{x\in [0, T_{0}~]} \left \vert \int_{0}^{x} \left[f\left(t, \omega_{k}(t ), D^{\sigma-1}\omega_{k}(t )\right) -f\left(t , \omega(t), D^{\sigma-1}\omega(t )\right)\right] dt \right \vert \\ &\leq \sup\limits_{\eta x\in [0, T_{0}~]} \int_{0}^{1}\frac{(\eta x)^{\sigma-1} \left|f\left(\eta x, \omega_{k}(\eta x), D^{\sigma-1}\omega_{k}(\eta x )\right)-f\left( \eta x , \omega(\eta x), D^{\sigma-1}\omega(\eta x )\right)\right|}{\Gamma \left(\sigma\right)\eta^{\sigma-1}\left( 1-\eta \right) ^{1-\sigma}} xd\eta \\ &+\sup\limits_{\eta x\in [0, T_{0}~]} \int_{0}^{1}\frac{(\eta x)^{\sigma-1} \left|f\left(\eta x, \omega_{k}(\eta x), D^{\sigma-1}\omega_{k}(\eta x )\right)-f\left( \eta x , \omega(\eta x), D^{\sigma-1}\omega(\eta x )\right)\right|}{\eta^{\sigma-1}} x^{2-\sigma}d\eta \\ &\rightarrow 0 \quad \text{as} \quad k\rightarrow \infty. \end{align*}

    In conclusion, since hypotheses of Theorem 2.2 are fulfilled, it implies that operator \mathcal{S} admits at least one fixed point in C^{\sigma-1} [0, T_{0}~], which is a solution of problem (1.2) as well.

    The mean value theorem for R-L derivative of order \sigma\in (0, 1) was correctly given by [7]. Now, its counterparts for order \sigma\in (1, 2) is given as follows:

    Lemma 3.2. Let \sigma \in (1, 2) and \omega\in C^{\sigma-1}\left(\left[0, T\right]\right). Then, there is a function \mu:[0, T]\to [0, T] with 0 < \mu(x) < x so that

    \omega(x) = D^{\sigma-1}\omega(0)\frac{x^{\sigma-1}}{\Gamma(\sigma)}+\Gamma(2-\sigma)x(\mu(x))^{\sigma-1}D^{\sigma-1}\omega(\mu(x)),

    is satisfied.

    The lemma can be proved by following the way used in [7] and so we omit it here. With the aid of this lemma we can obtain the Nagumo-type uniqueness:

    Theorem 3.2. (Nagumo type uniqueness) Let 1 < \sigma < 2, 0 < T < \infty and let condition (C1) be satisfied. Moreover, assume that there exists a positive real number L\leq \frac{2-\sigma}{\max\left( T, T^{2-\sigma}\right) (1+\Gamma(3-\sigma))} such that the inequality

    \begin{align} x^{\sigma-1}\left|f(x, \omega_{1}, v_{1})-f(x, \omega_{2}, v_{2})\right|\leq L\left( \left|\omega_{1}-\omega_{2}\right|+\left|v_{1}-v_{2}\right|\right) \end{align} (3.16)

    is fulfilled for all x\in[0, T] and for all \omega{i}, v{i}\in\mathbb{R} with i = 1, 2. Then, (1.2) has at most one solution in the space of C^{\sigma-1}(\left[0, T_0\right]).

    Proof. We have just showed the existence of the solution to problem (1.2) in the previous theorem. For the uniqueness, we first assume that (1.2) admits two different solutions such as \omega_{1} and \omega_{2} in the space of C^{\sigma-1}(\left[0, T_0\right]). Let us define a function \Phi(x) to be in the form

    \Phi(x): = \begin{cases} \left|\omega_{1}(x)-\omega_{2}(x)\right|+\left|D^{\sigma-1}\omega_{1}(x)-D^{\sigma-1}\omega_{2}(x)\right|, & x > 0 \\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ 0 \ \ \ \ \ \ \ \ \ \ \ , & x = 0. \end{cases}

    Since \omega_{1}, \omega_{2}\in C^{\sigma-1}(\left[0, T\right]), the continuity of \Phi(x) on x\in (0, T_0] can obviously be seen. For its continuity at x = 0,

    \begin{align*} 0\leq\lim\limits_{x\to 0^{+}}\Phi(x)& = \lim\limits_{x\to 0^{+}} \frac{1}{\Gamma(\sigma)}\left|\int_{0}^{x}\frac{f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)}{\left(x-t\right) ^{1-\sigma}} dt\right| \\ &+\lim\limits_{x\to 0^{+}}\left|\int_{0}^{x} f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right) dt\right| \\ &\leq\int_{0}^{1}\frac{\lim\limits_{x\to 0^{+}}x\left| H\left(x\eta, \omega_{1}\left( x\eta\right)\right)-H\left(x\eta, \omega_{2}\left( x\eta\right)\right) \right| }{\eta^{\sigma-1}\left(1-\eta\right)^{1-\sigma}}d\eta \\ &+\int_{0}^{1} \frac{\lim\limits_{x\to 0^{+}}x^{2-\sigma} \left| H\left(x\eta, \omega_{1}\left( x\eta\right)\right)-H\left(x\eta, \omega_{2}\left( x\eta\right)\right) \right|}{\eta^{\sigma-1}}d\eta = 0, \end{align*}

    where H(x, \omega(x)) = x^{\sigma-1}f\left(x, \omega\left(x\right), D^{\sigma-1}\omega(x ) \right) and we made the change of variable t = x\eta and used condition (C1), respectively. Consequently, \lim_{x\to 0^{+}}\Phi(x) = 0 = \Phi(0).

    The fact that \Phi(x)\geq 0 on [0, T] allows us to choose a point x_{0}\in (0, T] so that

    \begin{align*} 0 < \Phi(x_{0})& = \left|\omega_{1}(x_{0})-\omega_{2}(x_{0})\right| +\left|D^{\sigma-1}\omega_{1}(x_{0})-D^{\sigma-1}\omega_{2}(x_{0})\right|. \end{align*}

    By using the mean value theorem given by Lemma 3.2

    \begin{align} &\left|\omega_{1}(x_{0})-\omega_{2}(x_{0})\right| = \Gamma(2-\sigma)x_{0}\left|x_{1}^{\sigma-1}D^{\sigma-1}(\omega_{1}-\omega_{2})(x_{1})\right| \\ & = \Gamma(2-\sigma)x_{0}x_{1}^{\sigma-1}\left|f\left(x_{1}, \omega_{1}\left(x_{1}\right), D^{\sigma-1}\omega_{1}(x_{1})\right)-f\left(x_{1}, \omega_{2}\left(x_{1}\right), D^{\sigma-1}\omega_{2}(x_{1})\right)\right| \end{align} (3.17)

    is obtained for x_{1}\in (0, x_{0}).

    Secondly, for the estimation of \left|D^{\sigma-1}\omega_{1}(x_{0})-D^{\sigma-1}\omega_{2}(x_{0})\right|, we have from the well-known integral mean theorem for the classical calculus

    \begin{align} &\left|D^{\sigma-1}\omega_{1}(x_{0})-D^{\sigma-1}\omega_{2}(x_{0})\right| = \int_{0}^{x_0}\frac{t^{\sigma-1}\left|f(t, \omega_{1}(t), D^{\sigma-1}\omega_{1}(t)-f(t, \omega_{2}(t), D^{\sigma-1}\omega_{2}(t)\right|}{t^{\sigma-1}}dt \\ & = \frac{x^{2-\sigma}_{0}}{2-\sigma}x_{2}^{\sigma-1}\left|f\left(x_{2}, \omega_{1}\left(x_{2}\right), D^{\sigma-1}\omega_{1}(x_{2}) \right)-f\left(x_{2}, \omega_{2}\left(x_{2}\right), D^{\sigma-1}\omega_{2}(x_{2})\right)\right|, \end{align} (3.18)

    where x_{2}\in (0, x_{0}).

    We assign x_{3} as one of the points x_{1} and x_{2} so that \left| H(x_{3}, \omega_{1}(x_{3}))-H(x_{3}, \omega_{2}(x_{3}))\right| : = \max\left(\left| H(x_{1}, \omega_{1}(x_{1}))-H(x_{1}, \omega_{2}(x_{1}))\right|, \left| H(x_{2}, \omega_{1}(x_{2}))-H(x_{2}, \omega_{2}(x_{2}))\right| \right).

    Thus, from (3.17) and (3.18), we have

    \begin{align} \label{unique7} 0& < \Phi(x_{0})\leq\left(\Gamma(2-\sigma)x_{0}+\frac{x^{2-\sigma}_{0}}{2-\sigma}\right) \left| H(x_{3}, \omega_{1}(x_{3}))-H(x_{3}, \omega_{2}(x_{3}))\right|\\ &\leq \max\left( T, T^{2-\sigma}\right) \left( \frac{1+\Gamma(3-\sigma)}{2-\sigma}\right)x_{3}^{\sigma-1} \left|f\left(x_{3}, \omega_{1}\left(x_{3}\right)\right)-f\left(x_{3}, \omega_{2}\left(x_{3}\right)\right)\right|\\ &\leq x_{3}^{\sigma-1} \left( \left|\omega_{1}(x_{3})-\omega_{2}(x_{3})\right| +\left|D^{\sigma-1}\omega_{1}(x_{3})-D^{\sigma-1}\omega_{2}(x_{3})\right|\right) = \Phi(x_{3}) \end{align}

    since L\leq \frac{2-\sigma}{\max\left(T, T^{2-\sigma}\right)~~ (1+\Gamma(3-\sigma))}. Repeating the same procedure for the point x_{3}, it enables us to find some points x_{6}\in(0, x_{3}) so that 0 < \Phi(x_{0})\leq\Phi(x_{3})\leq \Phi(x_{6}). Continuing in the same way, the sequence \left\{x_{3n}\right\}_{n = 1}^{\infty}\subset [0, x_{0}) can be constructed so that x_{3n}\to 0 and

    \begin{equation} 0 < \Phi(x_{0})\leq\Phi(x_{3})\leq\Phi(x_{6})\leq...\leq\Phi(x_{3n})\leq... \end{equation} (3.19)

    However, the fact that \Phi(x) is continuous at x = 0 and x_{3n}\to 0 leads to \Phi(x_{3n})\to \Phi(0) = 0, and this contradicts with (3.19). Consequently, IVP (1.2) possesses a unique solution.

    Theorem 3.3. (Krasnoselskii-Krein type uniqueness) Let 1 < \sigma < 2 and T^{*}_{0} = \min\left\lbrace T_{0}~, 1 \right\rbrace, where T_0 is defined by (3.7). Let condition (C1) be fulfilled. Furthermore, suppose that there exist a L > 0 and an \alpha \in (0, 1) so that

    \begin{align} x^{1-\alpha(\sigma-1)}\left|f(x, \omega_{1}, v_{1})-f(x, \omega_{2}, v_{2})\right|\leq \Gamma(\sigma)\frac{\alpha(\sigma-1)+L}{2}\left( \left|\omega_{1}-\omega_{2}\right|+\left|v_{1}-v_{2}\right|\right) \end{align} (3.20)

    holds for all x\in[0, T] and for all \omega_{i}, v_{i}\in\mathbb{R} with i = 1, 2, and that there exist C > 0 satisfying (1-\sigma)(1-\alpha)-L(1-\alpha)+1 > 0 such that

    \begin{align} x^{\sigma-1}\left|f(x, \omega_{1}, v_{1})-f(x, \omega_{2}, v_{2})\right|\leq C\left( \left|\omega_{1}-\omega_{2}\right|^{\alpha}+x^{\alpha(\sigma-1)}\left|v_{1}-v_{2}\right|^{\alpha}\right) \end{align} (3.21)

    holds for all x\in[0, T] and for all \omega_{i}, v_{i}\in\mathbb{R} with i = 1, 2. Then, problem (1.2) has a unique solution in the space of C^{\sigma-1}(\left[0, T^{*}_{0}\right]).

    Proof. As claimed in the previous theorem, we first assume that problem (1.2) has two different solutions such as \omega_{1}(x) and \omega_{2}(x) in C^{\sigma-1}(\left[0, T^{*}_{0}\right]). However, by contradiction, we will show that they are indeed equal. For this, let us first define \Phi_{1}(x) = \left|\omega_{1}(x)-\omega_{2}(x)\right| and \Phi_{2}(x) = \left|D^{\sigma-1}\omega_{1}(x)-D^{\sigma-1}\omega_{2}(x)\right| and try to find estimates for each functions by using condition (C1) and inequality (3.21). Hence, we first have

    \begin{align*} \Phi_{1}(x)&\leq \frac{1}{\Gamma(\sigma)}\int_{0}^{x}\frac{\left|f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)\right|}{\left(x-t\right) ^{1-\sigma}} dt \\ &\leq \frac{1}{\Gamma(\sigma)}\int_{0}^{x}\frac{t^{\sigma-1}\left|f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)\right|}{t^{\sigma-1}\left(x-t\right) ^{1-\sigma}} dt \\ &\leq \frac{C}{\Gamma(\sigma)} \int_{0}^{x}\frac{\left[\Phi^{\alpha}_{1}(x)+t^{\alpha(\sigma-1)}\Phi^{\alpha}_{2}(x) \right] }{t^{\sigma-1}\left(x-t\right) ^{1-\sigma}} dt \\ &\leq \frac{C}{\Gamma(\sigma)}\left( \int_{0}^{x}\left(\frac{1}{t^{\sigma-1}\left(x-t\right) ^{1-\sigma}}\right) ^{q} dt\right)^{1/q}\left( \int_{0}^{x}\left[\Phi^{\alpha}_{1}(t)+t^{\alpha(\sigma-1)}\Phi^{\alpha}_{2}(t) \right]^{p} dt\right)^{1/p} \\ &\leq \frac{C\Gamma(1+(1-\sigma)q)\Gamma(1+(\sigma-1)q))}{\Gamma(\sigma)}x^{1/q}\Omega^{1/p}(x) \end{align*}

    where we used Hölder inequality with q > 1 satisfying (1-\sigma)q+1 > 0 and p = q/(q-1), and \Omega(x) is defined by

    \Omega(x) = \int_{0}^{x}\left[\Phi^{\alpha}_{1}(t)+t^{\alpha(\sigma-1)}\Phi^{\alpha}_{2}(t) \right]^{p}dt.

    From here, we have the following estimation

    \begin{align} \Phi^{p}_{1}(x)\leq Cx^{p/q}\Omega(x), \end{align} (3.22)

    where C is not specified here and throughout the proof. In addition to this, the upper bound for \Phi_{2}(x) can be found as follows:

    \begin{align*} \Phi_{2}(x)&\leq \int_{0}^{x} \left|f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)\right| dt \\ &\leq \int_{0}^{x}\frac{t^{\sigma-1}\left|f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)\right|}{t^{\sigma-1} } dt \\ &\leq \frac{C}{\Gamma(\sigma)} \int_{0}^{x}\frac{\left[\Phi^{\alpha}_{1}(x)+t^{\alpha(\sigma-1)}\Phi^{\alpha}_{2}(x) \right] }{t^{\sigma-1}\left(x-t\right) ^{1-\sigma}} dt \\ &\leq C \left( \int_{0}^{x}\left(\frac{1}{t^{\sigma-1}}\right)^{q} dt\right)^{1/q}\left( \int_{0}^{x}\left[\Phi^{\alpha}_{1}(t)+t^{\alpha(\sigma-1)}\Phi^{\alpha}_{2}(t) \right]^{p} dt\right)^{1/p} \\ &\leq \frac{C\Gamma(1+(1-\sigma)q)}{\Gamma(2+(\sigma-1)q))}x^{(1+q(1-\sigma))/q}\Omega^{1/p}(x). \end{align*}

    From here,

    \begin{align} \Phi^{p}_{2}(x)\leq Cx^{(1+q(1-\sigma))p/q}\Omega(x) \end{align} (3.23)

    is then obtained. By using estimations in (3.22) and (3.23) in the derivative of \Omega(x) we have

    \begin{align} \Omega'(x)& = \left[\Phi^{\alpha}_{1}(x)+t^{\alpha(\sigma-1)}\Phi^{\alpha}_{2}(x) \right]^{p}\leq 2^{p-1}\left[\left( \Phi^{\alpha}_{1}(x)\right) ^{p}+t^{p\alpha(\sigma-1)}\left( \Phi^{\alpha}_{2}(x)\right) ^{p}\right] \\ & = 2^{p-1}\left[\left( \Phi^{p}_{1}(x)\right)^{\alpha} +x^{p\alpha(\sigma-1)}\left( \Phi^{p}_{2}(x)\right)^{\alpha} \right]\\ &\leq 2^{p-1}\left[C^{\alpha}x^{\alpha p/q}\Omega^{\alpha}(x)+ C^{\alpha}x^{p\alpha(\sigma-1)}x^{(1+q(1-\sigma))\alpha p/q}\Omega^{\alpha}(x)\right]\leq Cx^{\alpha p/q}\Omega^{\alpha}(x). \end{align} (3.24)

    If we multiply the both sides of the above inequality with (1-\alpha)\Omega^{-\alpha}(x),

    \begin{align*} (1-\alpha)\Omega^{-\alpha}(x)\Omega'(x) = \frac{d}{dx}\left[\Omega^{1-\alpha}(x)\right] \leq Cx^{\alpha p/q}. \end{align*}

    is then obtained. Integrating the both sides of the inequality over [0, x], we get

    \begin{align*} \Omega^{1-\alpha}(x)\leq Cx^{(\alpha p+q)/q}, \end{align*}

    since \Omega(0) = 0. Consequently, this leads to the following estimation on \Omega(x)

    \begin{align} \Omega(x)\leq Cx^{(\alpha p+q)/(1-\alpha)q}. \end{align} (3.25)

    By considering (3.22) and (3.23) together with (3.25), one can conclude that

    \begin{align*} \Phi^{p}_{1}(x)\leq Cx^{p/q}\Omega(x)\leq Cx^{p/q}x^{(\alpha p+q)/(1-\alpha)pq} = Cx^{p+q/(1-\alpha)q}. \end{align*}

    or

    \begin{align} \Phi_{1}(x)\leq Cx^{(p+q)/(1-\alpha)pq} = x^{1/(1-\alpha)}, \end{align} (3.26)

    and

    \begin{align*} \Phi^{p}_{2}(x)\leq Cx^{p(1+q(1-\sigma))/q}\Omega(x)\leq Cx^{p(1+q(1-\sigma))/q}x^{(\alpha p+q)/(1-\alpha)pq} = Cx^{\frac{(1-\alpha)(1-\sigma)pq+p+q}{(1-\alpha)q}}. \end{align*}

    or

    \begin{align} \Phi_{2}(x)\leq Cx^{(1-\sigma)+\frac{p+q}{(1-\alpha)pq}} = x^{(1-\sigma)+\frac{1}{(1-\alpha)}}, \end{align} (3.27)

    since \frac{p+q}{pq} = 1.

    Let us now define

    \Psi(x) = x^{-L}\max\left\lbrace \Phi_{1}(x), \Phi_{2}(x)\right\rbrace,

    where L(1-\alpha) < 1+(1-\sigma)(1-\alpha). If \Phi_{1}(x) = \max\left\lbrace \Phi_{1}(x), \Phi_{2}(x)\right\rbrace, then from (3.26) we get

    0\leq \Psi(x)\leq x^{\frac{1}{1-\alpha}-L},

    or in the case of \Phi_{2}(x) = \max\left\lbrace \Phi_{1}(x), \Phi_{2}(x)\right\rbrace, by the inequality (3.27) we have the following

    0\leq \Psi(x)\leq x^{(1-\sigma)+\frac{1}{1-\alpha}-L} = x^{\frac{(1-\sigma)(1-\alpha)-L(1-\alpha)+1}{1-\alpha}}.

    In both cases, \Psi(x) is continuous on [0, T^{*}_{0}] and \Psi(0) = 0. Let us now show that \Psi(x)\equiv 0 on [0, T^{*}_{0}]. For this, suppose otherwise and let \Psi(x)\not\equiv 0. This means \Psi(x) > 0, and from its continuity one can say that there exists a point x_{1}\in [0, T^{*}_{0}] so that \Psi(x) takes its maximum value at that point. Thus, let

    M = \Psi(x_{1}) = \max\limits_{x\in [0, T^{*}_{0}]}\Psi(x).

    By assuming \Psi(x) = x^{-L}\Phi_{1}(x),

    \begin{align} M& = \Psi(x_{1}) = x_{1}^{-L}\Phi_{1}(x_{1}) \\ &\leq\frac{x_{1}^{-L}}{\Gamma(\sigma)}\int_{0}^{x_{1}}\frac{t^{1-\alpha(\sigma-1)}\left|f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)\right|}{t^{1-\alpha(\sigma-1)}\left(x-t\right) ^{1-\sigma}}dt \\ &\leq\frac{(\alpha(\sigma-1)+L)x_{1}^{-L}}{2}\int_{0}^{x_{1}}\left(x_{1}-t\right) ^{\sigma-1}t^{-1+\alpha(\sigma-1)}\left[\Phi_{1}(t)+\Phi_{2}(t) \right] dt \\ &\leq M(\alpha(\sigma-1)+L)x_{1}^{-L+\sigma-1}\int_{0}^{x_{1}}t^{-1+\alpha(\sigma-1)+L} dt \\ & < Mx^{(\alpha+1)(\sigma-1)}_{1}\leq M \end{align} (3.28)

    is obtained for t\in [0, x_{1}). However, we get to a contradiction.

    On the other hand, when \Psi(x) = x^{-L}\Phi_{2}(x), we get

    \begin{align} M& = \Psi(x_{1}) = x_{1}^{-L}\Phi_{2}(x_{1})\\ &\leq x_{1}^{-L} \int_{0}^{x_{1}}\frac{t^{1-\alpha(\sigma-1)}\left|f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)\right|}{t^{1-\alpha(\sigma-1)}}dt \\ &\leq\Gamma(\sigma)\frac{(\alpha(\sigma-1)+L)x_{1}^{-L}}{2}\int_{0}^{x_{1}}t^{-1+\alpha\left( \sigma-1\right)+L}\left[\Phi_{1}(t)+\Phi_{2}(t) \right] dt \\ & < Mx^{\alpha(\sigma-1)}_{1}\leq M, \end{align} (3.29)

    which is contraction as well.

    Consequently, \Psi(x) vanishes identically on [0, T^{*}_{0}]. This gives us the uniqueness of solutions to the considered problem.

    Theorem 3.4 (Osgood-type uniqueness). Let 1 < \sigma < 2, and T_0 be defined by (3.7), and let condition (C1) be satisfied. Furthermore, suppose that the equality

    \begin{align} x^{\sigma-1}\left|f(x, \omega_{1}, v_{1})-f(x, \omega_{2}, v_{2})\right|\leq C\left( g\left(\left|\omega_{1}-\omega_{2}\right|^{p}+\left|v_{1}-v_{2}\right|^{p}\right)\right)^{1/p} \end{align} (3.30)

    is fulfilled for all x\in[0, T] and for all \omega_{i}, v_{i}\in\mathbb{R} with i = 1, 2, where p > 1 is conjugate of q > 1 satisfying 1+(1-\sigma)q > 0 and

    \begin{align*} C^{q}\geq 2\max\left( b\Gamma^{q}(\sigma)[T_0\Gamma(1+(1-\sigma)q)\Gamma(1+(\sigma-1)q)]^{-1} , (1+(1-\sigma)q) T_{0}~^{-1-q(1-\sigma)} \right). \end{align*}

    Moreover, assume that g is a continuous, non-negative and non-decreasing function in \left[0, \infty\right) so that g(0) = 0 and it satisfies

    \begin{align} \lim\limits_{\epsilon\to 0^{+}}\int_{\epsilon}^{\gamma}\frac{du}{g(u)} = \infty \end{align} (3.31)

    for any \gamma\in\mathbb{R}. Then, (1.2) has a unique solution in the space C^{\sigma-1}(\left[0, T_{0}~\right]).

    Proof. As made in previously given uniqueness theorems, we assume that there exist two different solutions such as \omega_{1}(x) and \omega_{2}(x) to problem (1.2) in C^{\sigma-1}(\left[0, T^{*}_{0}\right]). Moreover, let \Phi_{1}(x) = \left| \omega_{1}(x)-\omega_{2}(x)\right| and \Phi_{2}(x) = \left| D^{\sigma-1}\omega_{1}(x)-D^{\sigma-1}\omega_{2}(x)\right|. At first, we get the estimation on \Phi_{1}(x) as follows:

    \begin{align} \Phi_{1}(x)&\leq \frac{1}{\Gamma(\sigma)}\int_{0}^{x}\frac{t^{\sigma-1}\left|f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)\right|}{t^{\sigma-1}\left(x-t\right) ^{1-\sigma}} dt \\ &\leq \frac{C}{\Gamma(\sigma)} \int_{0}^{x}\frac{\left[ g\left(\left|\omega_{1}\left(t\right)-\omega_{2}\left(t\right)\right|^{p}+\left|D^{\sigma-1}\omega_{1}\left(t\right)-D^{\sigma-1}\omega_{2}\left(t\right)\right|^{p}\right)\right] ^{1/p}}{t^{\sigma-1}\left(x-t\right) ^{1-\sigma}} dt \\ &\leq \frac{C}{\Gamma(\sigma)}\left( \int_{0}^{x}\left(\frac{1}{t^{\sigma-1}\left(x-t\right) ^{1-\sigma}}\right) ^{q} dt\right)^{1/q}\left( \int_{0}^{x} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right) dt\right)^{1/p} \\ &\leq C\left[ \frac{\Gamma(1+(1-\sigma)q)\Gamma(1+(\sigma-1)q)}{\Gamma^{q}(\sigma)}\right]^{1/q}x^{1/q}\left( \int_{0}^{x} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right) dt\right)^{1/p} \\ &\leq \frac{1}{2^{1/p}}\left( \int_{0}^{x} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right) dt\right)^{1/p} \end{align} (3.32)

    where we used the inequality (3.30), Hölder inequality and the assumption on C, respectively. From here, it follows that

    \begin{align} \Phi^{p}_{1}(x)\leq \frac{1}{2}\int_{0}^{x} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right) dt . \end{align} (3.33)

    Similarly to above, we have

    \begin{align} \Phi_{2}(x)&\leq \int_{0}^{x}\frac{t^{\sigma-1}\left|f\left(t , \omega_{1}\left(t\right), D^{\sigma-1}\omega_{1}\left(t\right)\right)-f\left(t , \omega_{2}\left(t\right), D^{\sigma-1}\omega_{2}\left(t\right)\right)\right|}{t^{\sigma-1}} dt \\ &\leq C \int_{0}^{x}\frac{\left[ g\left(\left|\omega_{1}\left(t\right)-\omega_{2}\left(t\right)\right|^{p}+\left|D^{\sigma-1}\omega_{1}\left(t\right)-D^{\sigma-1}\omega_{2}\left(t\right)\right|^{p}\right)\right] ^{1/p}}{t^{\sigma-1}} dt \\ &\leq C\left( \int_{0}^{x}\left(\frac{1}{t^{\sigma-1}}\right) ^{q} dt\right)^{1/q}\left( \int_{0}^{x} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right) dt\right)^{1/p} \\ &\leq C\left[ \frac{1}{(1+(1-\sigma)q)}\right]^{1/q} x^{(1+q(1-\sigma))/q}\left( \int_{0}^{x} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right) dt\right)^{1/p} \\ &\leq \frac{1}{2^{1/p}}\left( \int_{0}^{x} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right) dt\right)^{1/p}. \end{align} (3.34)

    This leads to

    \begin{align} \Phi^{p}_{2}(x)\leq \frac{1}{2}\int_{0}^{x} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right) dt . \end{align} (3.35)

    Now, set

    \Psi(x): = \max\limits_{0\leq t\leq x} \left[ \Phi^{p}_{1}(x)+\Phi^{p}_{2}(x)\right],

    and assume that \Psi(x) > 0 for x\in (0, T_{0}~]. We will show that it can not be possible under assumptions.

    From the definition of \Psi(x) one easily conclude that for each x\in [0, T_0], it is \Phi^{p}_{1}(x)+\Phi^{p}_{2}(x)\leq \Psi(x) and there exists a x_{1}\leq x so that \Psi(x) = \Phi^{p}_{1}(x_{1})+\Phi^{p}_{2}(x_{1}). Then, from estimations (3.33)–(3.35) and from the fact that g is non-decreasing function

    \begin{align} \Psi(x) = \Phi^{p}_{1}(x_{1})+\Phi^{p}_{2}(x_{1})\leq \int_{0}^{x_{1}} g\left(\Phi^{p}_{1}(t)+\Phi^{p}_{2}(t)\right)\leq \int_{0}^{x} g\left(\Psi(t)\right)dt: = \Psi_{*}(x) \end{align} (3.36)

    is then obtained. It can be seen that \Psi(x)\leq \Psi_{*}(x). Moreover, we have

    \frac{d}{dx}\Psi_{*}(x) = g\left(\Psi(x)\right)\leq g\left(\Psi_{*}(x)\right)

    for all x\in [0, T_0]. From this fact, for sufficiently small \delta > 0, we have

    \int_{\delta}^{T_{0}~}\frac{\Psi^{'}_{*}(x)}{g\left(\Psi_{*}(x)\right)}dx\leq T_{0}~-\delta.

    Furthermore, by changing variables u = \Psi_{*}(x) in the above integral and by using the continuity of \Psi_{*}(x) and \Psi_{*}(0) = 0 , we have

    \int_{\epsilon}^{\gamma}\frac{du}{g\left(u\right)} \leq T_{0}~-\delta.

    for sufficiently small \epsilon > 0 with \epsilon = \Psi_{*}(\delta) and for \gamma = \Psi_{*}(T_{0}~). However, this contradicts with the assumption on g given in (3.31). Consequently, \Psi(x) = 0 for x\in [0, T_{0}~], i.e. \omega_{1} = \omega_{2}.

    Remark 3.1. It must be pointed out that, as noted in Theorem 1.4.3 in [13], the condition that function g(u) is non-decreasing can be dropped.

    Example 3.1. Let us consider the following problem

    \begin{equation} D^{\frac{4}{3}}\omega(x) = \begin{cases} x^{-\frac{1}{3}}\left[-\omega \ln\omega-(D^{\frac{1}{3}}\omega)\ln(D^{\frac{1}{3}}\omega) \right], &if \quad 0 < \omega, D^{\frac{1}{3}}\omega\leq 1/e \\ 0, &if\quad \omega = 0 \quad or \quad D^{\frac{1}{3}}\omega = 0, \end{cases} \end{equation} (3.37)

    with initial conditions \omega(0) = 0 and D^{\sigma-1}\omega\left(x\right)|_{x = 0} = 1. Let T = 1 in Theorem 3.1. Then, \mathcal{M} = \max_ {(x, u, v)\in [0, 1]\times [0, 1/e]\times [0, 1/e]} \left|x^{\frac{1}{3}}f(x, \omega, v) \right|\cong \frac{2}{e}, C(b, \sigma, \mathcal{M}) = C(1, \frac{4}{3}, \frac{2}{e})\cong 3.2 and T_{0}~ = \frac{r}{C(b, \sigma, \mathcal{M})}\cong 0.22. Hence, problem (3.37) has a solution in C^{\frac{1}{3}}(\left[0, 0.22 \right]). Now, we investigate the uniqueness of the solution to the problem in C^{\frac{1}{3}}(\left[0, 0.22 \right]). Let the function g in the previous theorem defined by

    \begin{equation} g(u) = \begin{cases} 0 &if \quad u = 0 \\ -\sqrt{u}\ln\sqrt{u} &if\quad 0 < u\leq 1/e^{2} , \\ 1/e &if\quad u\geq 1/e^{2} , \\ \end{cases} \end{equation} (3.38)

    It is obvious that g is positive for u > 0 . Since g'(u) = -\frac{1}{2\sqrt{u}}\left(1+\ln(\sqrt{u})\right) > 0 for 0 < u\leq 1/e^{2}, it is non-decreasing. Also,

    \begin{align} \lim\limits_{\epsilon\to 0^{+}}\int_{\epsilon}^{\gamma}\frac{du}{g(u)} = \infty \end{align} (3.39)

    for any \gamma\in\mathbb{R}, which can be seen by considering the inequality \frac{1}{g(u)} = -\frac{1}{\sqrt{u}\ln\sqrt{u}}\leq \frac{1}{u} in the neighborhood of u = 0 and the divergence of the integral \lim_{\epsilon\to 0^{+}}\int_{\epsilon}^{\gamma}\frac{du}{u} and by applying comparison test. Moreover, by using the concavity of the nonlinear function f with respect to second and third variables, we have

    \begin{align} x^{1/3}\left|f(x, \omega_{1}, v_{1})-f(x, \omega_{2}, v_{2})\right|&\leq \left|-\omega_{1}\ln \omega_{1}-v_{1}\ln v_{1}-\left( -\omega_{2}\ln \omega_{2}-v_{2}\ln v_{2}\right) \right| \\ &\leq -\left|\omega_{1}\ln \omega_{1}-\omega_{2}\ln \omega_{2} \right|-\left|v_{1}\ln v_{1}-v_{2}\ln v_{2} \right| \\ &\leq -\left|\omega_{1}-\omega_{2}\right| \left| \ln \left( \omega_{1}- \omega_{2}\right) \right|-\left|v_{1}-v_{2}\right| \left| \ln \left( v_{1}-v_{2}\right) \right| \\ &\leq -2\sqrt{ \left|\omega_{1}-\omega_{2}\right|^{2}+\left|v_{1}-v_{2}\right|^{2}} \ln \sqrt{ \left|\omega_{1}-\omega_{2}\right|^{2}+\left|v_{1}-v_{2}\right|^{2}} \\ &\leq C \left( g\left(\left|\omega_{1}-\omega_{2}\right|^{2}+\left|v_{1}-v_{2}\right|^{2}\right)\right) ^{1/2} \end{align} (3.40)

    where v_{i} = D^{\frac{1}{3}}\omega_{i} for i = 1, 2, p = q = 2 and

    C\geq 1.725\cong \sqrt{2}\max\left( \frac{\Gamma(4/3)}{0.22\times\Gamma(1/3)\Gamma(5/3)}, \frac{1} {3\times 0.22^{1/3}} \right).

    Hence, assumptions of Theorem 3.4 are satisfied. So, the problem has a unique solution in C^{\frac{1}{3}}(\left[0, 0.22 \right]).

    In this research, we gave some sufficient conditions for the existence and uniqueness of a problem involving a nonlinear differential equations in the sense of R-L derivative when the right-hand side function has a discontinuity at zero. We presented an example associated with two theorems. Considering the literature, these results can be generalized and improved. Besides, one can obtain another uniqueness results for this problem as well.

    There is no conflict of interest.



    [1] B. Wang, S. Zhang, J. Dong, Y. Li, Y. Jin, H. Xiao, et al., Ambient temperature structures the gut microbiota of zebrafish to impact the response to radioactive pollution, Environ. Pollut., 293 (2022), 118539. https://doi.org/10.1016/j.envpol.2021.118539 doi: 10.1016/j.envpol.2021.118539
    [2] K. Liu, J. Zhou, D. Dong, Improving stock price prediction using the long short-term memory model combined with online social networks, J. Behav. Exp. Finance, 30 (2021). https://doi.org/10.1016/j.jbef.2021.100507 doi: 10.1016/j.jbef.2021.100507
    [3] Y. Lv, J. Piao, B. Li, M. Yang, Does online investor sentiment impact stock returns? Evidence from the Chinese stock market, Appl. Econ. Lett., 29 (2022), 1434–1438. https://doi.org/10.1080/13504851.2021.1937490 doi: 10.1080/13504851.2021.1937490
    [4] G. Huberman, T. Regev, Contagious speculation and a cure for cancer: A nonevent that made stock prices soar, J. Finance, 56 (2001), 387–396. https://doi.org/10.1111/0022-1082.00330 doi: 10.1111/0022-1082.00330
    [5] U. Bhattacharya, N. Galpin, R. Ray, X. Yu, The role of the media in the internet IPO bubble, J. Finance Quant. Anal., 44 (2009), 657–682. https://doi.org/10.1017/S0022109009990056 doi: 10.1017/S0022109009990056
    [6] H. J. V. Heerde, E. Gijsbrechts, K. Pauwels, Fanning the flames? how media coverage of a price war affects retailers, consumers, and investors, J. Mark. Res., 52 (2015), 674–693. https://doi.org/10.1509/jmr.13.0260 doi: 10.1509/jmr.13.0260
    [7] L. Fang, J. Peress, Media coverage and the cross-section of stock returns, J. Finance, 64 (2009), 2023–2052. https://doi.org/10.1111/j.1540-6261.2009.01493.x doi: 10.1111/j.1540-6261.2009.01493.x
    [8] P. C. Tetlock, Giving content to investor sentiment: The role of media in the stock market, J. Finance, 62 (2007), 1139–1168. https://doi.org/10.1111/j.1540-6261.2007.01232.x doi: 10.1111/j.1540-6261.2007.01232.x
    [9] J. Engelberg, Costly information processing: evidence from earnings announcements, AFA 2009 San Francisco Meetings Paper, 2008. http://dx.doi.org/10.2139/ssrn.1107998.
    [10] H. Du, J. Hao, F. He, W. Xi, Media sentiment and cross-sectional stock returns in the Chinese stock market, Res. Int. Bus. Finance, 60 (2022), 101590, https://doi.org/10.1016/j.ribaf.2021.101590 doi: 10.1016/j.ribaf.2021.101590
    [11] M. W. Uhl, The Long-run impact of sentiment on stock returns, Working Paper, 2011.
    [12] M. T. Suleman, Stock market reaction to good and bad political news, Asian J. Finance Account., 4 (2012), 299–312. https://doi.org/10.5296/ajfa.v4i1.1705 doi: 10.5296/ajfa.v4i1.1705
    [13] G. W. Brown, M. T. Cliff., Investor sentiment and the near-term stock market, J. Empir. Finance, 11 (2004), 1–27. https://doi.org/10.1016/j.jempfin.2002.12.001 doi: 10.1016/j.jempfin.2002.12.001
    [14] J. Wurgler, M. Baker, Investor sentiment and the cross-section of stock returns, J. Finance, 61 (2006), 1645–1680. https://doi.org/10.1111/j.1540-6261.2006.00885.x doi: 10.1111/j.1540-6261.2006.00885.x
    [15] F. M. Statman, Investor sentiment and stock returns, Finance Anal. J., 56 (2000), 16–23. https://doi.org/10.2469/faj.v56.n2.2340 doi: 10.2469/faj.v56.n2.2340
    [16] Z. Li, S. Wang, M. Hu, International investor sentiment and stock returns: Evidence from China, Invest. Anal. J., 50 (2021), 60–76. https://doi.org/10.1080/10293523.2021.1876968 doi: 10.1080/10293523.2021.1876968
    [17] Y. Kim, K. Y. Lee, Impact of investor sentiment on stock returns, Asia-Pac. J. Finance Stud., 51 (2022), 132–162. https://doi.org/10.1111/ajfs.12362 doi: 10.1111/ajfs.12362
    [18] T. Renault, Intraday online investor sentiment and return patterns in the US stock market, J. Bank Finance, 84 (2017), 25–40. https://doi.org/10.1016/j.jbankfin.2017.07.002 doi: 10.1016/j.jbankfin.2017.07.002
    [19] E. Bartov, L. Faurel, P. S. Mohanram, Can twitter help predict firm-level earnings and stock returns?, Account. Rev., 93 (2018), 25–57. https://doi.org/10.2308/accr-51865 doi: 10.2308/accr-51865
    [20] Y. Shynkevich, T. M. Mcginnity, S. Coleman, A. Belatreche, Stock price prediction based on stock-specific and sub-industry-specific news articles, in 2015 International Joint Conference on Neural Networks (IJCNN), (2015), 1–8. https://doi.org/10.1109/IJCNN.2015.7280517
    [21] H. Yun, G. Sim, J. Seok, Stock prices prediction using the title of newspaper articles with Korean natural language processing, in 2019 International Conference on Artificial Intelligence in Information and Communication (ICAⅡC), (2019). https://doi.org/10.1109/ICAⅡC.2019.8668996
    [22] M. Zhang, J. Yang, M. Wan, X. Zhang, J. Zhou., Predicting long-term stock movements with fused textual features of Chinese research reports, Expert Syst. Appl., 210 (2022), 118312. https://doi.org/10.1016/j.eswa.2022.118312 doi: 10.1016/j.eswa.2022.118312
    [23] N. Barberis, M. Huang, Stock as lotteries: The implications of probability weighting for security prices, Am. Econ. Rev., 98 (2008), 2066–2100. https://doi.org/10.1257/aer.98.5.2066 doi: 10.1257/aer.98.5.2066
    [24] B. Boyer, T. Mitton, K. Vorkink, Expected idiosyncratic skewness, Rev. Finance Stud., 23 (2010), 169–202. https://doi.org/10.1093/rfs/hhp041 doi: 10.1093/rfs/hhp041
    [25] T. G. Bali, N. Cakici, R. F. Whitelaw, Maxing out: Stock as lotteries and the cross-section of expected returns, J. Finance Econ., 99 (2011), 427–446. https://doi.org/10.1016/j.jfineco.2010.08.014 doi: 10.1016/j.jfineco.2010.08.014
    [26] J. Conrad, R. F. Dittmar, E. Ghysels, Ex ante skewness and expected stock returns, J. Finance, 68 (2013), 85–124. https://doi.org/10.1111/j.1540-6261.2012.01795.x doi: 10.1111/j.1540-6261.2012.01795.x
    [27] N. Barberis, A. Mukherjee, B. Wang, Prospect theory and stock returns: An empirical test, Rev. Finance Stud., 29 (2016), 3068–3107. https://doi.org/10.1093/rfs/hhw049 doi: 10.1093/rfs/hhw049
    [28] J. Wang, C. Wu, X. Zhong, Prospect theory and stock returns: Evidence from foreign share markets, Pac.-Basin Finance J., 69 (2021), 101644. https://doi.org/10.1016/j.pacfin.2021.101644 doi: 10.1016/j.pacfin.2021.101644
    [29] A. J. N. Junior, M. C. Klotzle, L. E. T. Brandão, A. C. F. Pinto, Prospect theory and narrow framing bias: Evidence from emerging markets, Q. Rev. Econ. Finance, 80 (2021), 90–101. https://doi.org/10.1016/j.qref.2021.01.016 doi: 10.1016/j.qref.2021.01.016
    [30] X. Yang, D. Gu, J. Wu, C. Liang, Y. Ma, J. Li, Factors influencing health anxiety: The stimulus–organism-response model perspective, Internet Res., 31 (2021), 2033–2054. https://doi.org/10.1108/INTR-10-2020-0604 doi: 10.1108/INTR-10-2020-0604
    [31] Z. Tang, M. Warkentin, L. Wu, Understanding employees' energy saving behavior from the perspective of stimulus-organism-responses, Resour. Conserv. Recycl., 140 (2019), 216–223. https://doi.org/10.1016/j.resconrec.2018.09.030 doi: 10.1016/j.resconrec.2018.09.030
    [32] B. J. Bushee, J. E. Core, W. Guay, S. Hamm, The role of the business press as an information intermediary, J. Account. Res., 48 (2010), 1–19. https://doi.org/10.1111/j.1475-679X.2009.00357.x doi: 10.1111/j.1475-679X.2009.00357.x
    [33] Z. Da, J. Engelberg, P. J. Gao, In search of attention, J. Finance, 66 (2011), 1461–1499. https://doi.org/10.1111/j.1540-6261.2011.01679.x doi: 10.1111/j.1540-6261.2011.01679.x
    [34] F. Comiran, T. Fedyk, J. Ha, Accounting quality and media attention around seasoned equity offerings, Int. J. Account. Inf. Manage., 26 (2017), 443–462. https://doi.org/10.1108/IJAIM-02-2017-0029 doi: 10.1108/IJAIM-02-2017-0029
    [35] W. S. Chan., Stock price reaction to news and no-news: drift and reversal after headlines, J. Finance Econ., 70 (2003), 223–260. https://doi.org/10.1016/S0304-405X(03)00146-6 doi: 10.1016/S0304-405X(03)00146-6
    [36] P. C. Tetlock, M. Saar-Tsechansky, S. Macskassy, More than words: Quantifying language to measure firms' fundamentals, J. Finance, 63 (2008), 1437–1467. https://doi.org/10.1111/j.1540-6261.2008.01362.x doi: 10.1111/j.1540-6261.2008.01362.x
    [37] P. Jiao, A. Veiga, A. Walther, Social media, news media and the stock market, J. Econ. Behav. Organ., 176 (2020), 63–90. https://doi.org/10.1016/j.jebo.2020.03.002 doi: 10.1016/j.jebo.2020.03.002
    [38] T. Huang, X. Zhang, Industry-level media tone and the cross-section of stock returns, Int. Rev. Econ. Finance, 77 (2021), 59–77. https://doi.org/10.1016/j.iref.2021.09.002 doi: 10.1016/j.iref.2021.09.002
    [39] Y. He, L. Qu, R. Wei, X. Zhao, Media-based investor sentiment and stock returns: A textual analysis based on newspapers, Appl. Econ., 54 (2022), 774–792. https://doi.org/10.1080/00036846.2021.1966369 doi: 10.1080/00036846.2021.1966369
    [40] W. Wang, C. Su, D. Duxbury, The conditional impact of investor sentiment in global stock markets: A two-channel examination, J. Bank Finance, 138 (2022), 106458. https://doi.org/10.1016/j.jbankfin.2022.106458 doi: 10.1016/j.jbankfin.2022.106458
    [41] R. B. Cohen, C. Polk, T. Vuolteenaho, The price is (almost) right, J. Finance, 64 (2009), 2739–2782. https://doi.org/10.1111/j.1540-6261.2009.01516.x doi: 10.1111/j.1540-6261.2009.01516.x
    [42] Z. Da, J. Engelberg, P. Gao, The sum of all FEARS investor sentiment and asset prices, Rev. Finance Stud., 28 (2015), 1–32. https://doi.org/10.1093/rfs/hhu072 doi: 10.1093/rfs/hhu072
    [43] H. Yang, D. Ryu, D. Ryu, Investor sentiment, asset returns and firm characteristics: Evidence from the Korean stock market, Invest. Anal. J., 46 (2017), 1–16. https://doi.org/10.1080/10293523.2016.1277850 doi: 10.1080/10293523.2016.1277850
    [44] J. Li, Y. Zhang, L. Wang, Information transmission between large shareholders and stock volatility, N. Am. Econ. Finance, 58 (2021), 101551. https://doi.org/10.1016/j.najef.2021.101551 doi: 10.1016/j.najef.2021.101551
    [45] M. Ammann, R. Frey, M. Verhofen, Do newspaper articles predict aggregate stock returns?, J. Behav. Finance, 15 (2014), 195–213. https://doi.org/10.1080/15427560.2014.941061 doi: 10.1080/15427560.2014.941061
    [46] F. Wong, Z. Liu, M. Chiang, Stock market prediction from WSJ: text mining via sparse matrix factorization, in Proceedings of the 2014 IEEE International Conference on Data, (2014), 430–439. https://doi.org/10.48550/arXiv.1406.7330
    [47] N. C. Barberis, Thirty years of prospect theory in economics: A review and assessment, J. Econ. Perspect., 27 (2013), 173–195. https://doi.org/10.1257/jep.27.1.173 doi: 10.1257/jep.27.1.173
    [48] D. Kahneman, A. Tversky, Prospect theory: An analysis of decision under risk, Econometrica, 47 (1979), 263–291. http://www.jstor.org/stable/1914185
    [49] H. Chen, P. De, Y. Hu, B. H. Hwang, Wisdom of crowds: The value of stock opinions transmitted through social media, Rev. Finance Stud., 27 (2013), 1367–1403. https://doi.org/10.1093/rfs/hhu001 doi: 10.1093/rfs/hhu001
  • This article has been cited by:

    1. Peng E, Tingting Xu, Linhua Deng, Yulin Shan, Miao Wan, Weihong Zhou, Solutions of a class of higher order variable coefficient homogeneous differential equations, 2025, 20, 1556-1801, 213, 10.3934/nhm.2025011
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2400) PDF downloads(253) Cited by(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog