Research article Special Issues

Associations of CYP1A1 gene polymorphisms and risk of breast cancer in Indian women: a meta-analysis

  • Reported associations of CYP1A1 polymorphisms with breast cancer have been inconsistent. In this meta-analysis examining breast cancer associations of three CYP1A1 polymorphisms (M1, M2 and M4) among Indian women may yield information that may be of clinical and epidemiological use for this particular demography. We searched MEDLINE using PubMed and Embase for association studies. From seven published case-control studies, we estimated overall associations and applied subgroup analysis to explore differential effects. All three polymorphisms exhibited overall increased risk, significant in M1 (OR 1.61-1.65, p = 0.04) and M4 (OR 2.02-3.92, p = 0.02-0.04). Differential effects were observed only in the M1 polymorphism where M1 effects were significant in South Indians (OR 2.20-4.34, p < 0.0001) but not the North population, who were at reduced risk (OR 0.64-0.77, p = 0.03-0.55). These populations were not materially different in regard to M2 and M4 as did the women stratified by menopausal status. In this meta-analysis, M1 and M4 effects may render Indian women susceptible, but may be limited by heterogeneity of the studies. Differential effects of the M1 polymorphism in breast cancer render South Indians susceptible compared to those in the North.

    Citation: Noel Pabalan, Neetu Singh, Eloisa Singian, Caio Parente Barbosa, Bianca Bianco, Hamdi Jarjanazi. Associations of CYP1A1 gene polymorphisms and risk of breast cancer in Indian women: a meta-analysis[J]. AIMS Genetics, 2015, 2(4): 250-262. doi: 10.3934/genet.2015.4.250

    Related Papers:

    [1] Boyu Liu, Wenyan Wang . An efficient numerical scheme in reproducing kernel space for space fractional partial differential equations. AIMS Mathematics, 2024, 9(11): 33286-33300. doi: 10.3934/math.20241588
    [2] Shiyv Wang, Xueqin Lv, Songyan He . The reproducing kernel method for nonlinear fourth-order BVPs. AIMS Mathematics, 2023, 8(11): 25371-25381. doi: 10.3934/math.20231294
    [3] Yingchao Zhang, Yingzhen Lin . An $ {\varepsilon} $-approximation solution of time-fractional diffusion equations based on Legendre polynomials. AIMS Mathematics, 2024, 9(6): 16773-16789. doi: 10.3934/math.2024813
    [4] Zhi-Yuan Li, Mei-Chun Wang, Yu-Lan Wang . Solving a class of variable order nonlinear fractional integral differential equations by using reproducing kernel function. AIMS Mathematics, 2022, 7(7): 12935-12951. doi: 10.3934/math.2022716
    [5] Haroon Ahmad, Om Prakash Chauhan, Tania Angelica Lazăr, Vasile Lucian Lazăr . Some convergence results on proximal contractions with application to nonlinear fractional differential equation. AIMS Mathematics, 2025, 10(3): 5353-5372. doi: 10.3934/math.2025247
    [6] Xiangyun Qiu, Xingxing Yue . Solving time fractional partial differential equations with variable coefficients using a spatio-temporal meshless method. AIMS Mathematics, 2024, 9(10): 27150-27166. doi: 10.3934/math.20241320
    [7] Wei Fan, Kangqun Zhang . Local well-posedness results for the nonlinear fractional diffusion equation involving a Erdélyi-Kober operator. AIMS Mathematics, 2024, 9(9): 25494-25512. doi: 10.3934/math.20241245
    [8] Fethi Bouzeffour . Inversion formulas for space-fractional Bessel heat diffusion through Tikhonov regularization. AIMS Mathematics, 2024, 9(8): 20826-20842. doi: 10.3934/math.20241013
    [9] Siyu Tian, Boyu Liu, Wenyan Wang . A novel numerical scheme for reproducing kernel space of 2D fractional diffusion equations. AIMS Mathematics, 2023, 8(12): 29058-29072. doi: 10.3934/math.20231488
    [10] Amnah E. Shammaky, Eslam M. Youssef . Analytical and numerical techniques for solving a fractional integro-differential equation in complex space. AIMS Mathematics, 2024, 9(11): 32138-32156. doi: 10.3934/math.20241543
  • Reported associations of CYP1A1 polymorphisms with breast cancer have been inconsistent. In this meta-analysis examining breast cancer associations of three CYP1A1 polymorphisms (M1, M2 and M4) among Indian women may yield information that may be of clinical and epidemiological use for this particular demography. We searched MEDLINE using PubMed and Embase for association studies. From seven published case-control studies, we estimated overall associations and applied subgroup analysis to explore differential effects. All three polymorphisms exhibited overall increased risk, significant in M1 (OR 1.61-1.65, p = 0.04) and M4 (OR 2.02-3.92, p = 0.02-0.04). Differential effects were observed only in the M1 polymorphism where M1 effects were significant in South Indians (OR 2.20-4.34, p < 0.0001) but not the North population, who were at reduced risk (OR 0.64-0.77, p = 0.03-0.55). These populations were not materially different in regard to M2 and M4 as did the women stratified by menopausal status. In this meta-analysis, M1 and M4 effects may render Indian women susceptible, but may be limited by heterogeneity of the studies. Differential effects of the M1 polymorphism in breast cancer render South Indians susceptible compared to those in the North.


    In recent years, the study of fractional derivatives has been an important topic. It has been used to model many phenomena in numerous fields such as science and engineering. There are many interpretations for fractional derivatives, such as the definition of Caputo [1], the definition of Riemann-Liouville [2], the definition of Grunwald-Letnikov [3], and most recently, Conformable [4], Atangana-Baleanu [5], Wallström [6], Jumarie [7], Klimek [8] and others.

    In practice, where quantitative results are needed for given real-world problems, numerically approximate solutions can often be demonstrably better, more reliable, more detailed, efficient and cost-effective than analytical ones for certain fractional structures. A number of studies [9,10,11,12,13,14] were therefore involved in developing approaches for providing estimated solutions. One of these approaches is the Hilbert space kernel reproduction (RKHS) method used for the first time by S. Zaremba for the harmonic and biharmonic functions at the beginning of the 20th century to find solutions for boundary value problems (BVPs).

    The RKHS precede the Dirac delta function in many ways, among which we mention providing an important structure for random distribution of multi-round data and, providing accurate approximation of multi-dimensional general functions and the possibility to pick any point in the integration interval.

    The RKHS algorithm has been successfully applied to various fields of numerical analysis, computational mathematics, probability and statistics [15,16], biology [17], quantum mechanics and wave mechanics [18]. Therefore wide range of research works have been directed to its applications in various stochastic categories [19], and defined problems involving operator equations [20], partial differential equations [21,22], integrative equations [23,24], and differential integration equations [24,25,26,27,28,29]. In addition, many studies have focused in recent years on the use of the RKHS method as a framework for seeking approximate numerical solutions to different problems [30,31,32,33,34,35,36,37,38,39].

    Moreover, the numerical solution of the different groups of BVP can be found in [40,41,42]. The two-point BVPs has a strong interest in applied mathematics, this kind of problems arise directly from mathematical models or by turning partial differential equations into ordinary differential equations. As this type of problems does not have an exact solution, many special techniques have been used to solve it, including the shooting method [43,44], the collocation method [45,46], the finite difference method [47,48], and the quasilinearization method [49,50]. The continuous genetic algorithm approach was used to solve these schemes in [51,52,53].

    The present paper is structured as follows: in Section 2, we set out some basic concepts and results from fractional calculus theory. In Section 3, the iterative form of the reproducing kernel algorithm is used to build and measure the solution of the fractional differential method with temporal two points. In Section 4 and 5, the convergence and error estimator are discussed to provide a number of numerical results to demonstrate the efficiency and accuracy of the reproducing kernel Hilbert space method. At last in section 6, a conclusion of the results is made.

    In applied mathematics and mathematical analysis, there are several definitions of fractional derivatives, Riemann-Liouville and Caputo are the most popular of all [54]. In this section, we list some of these definitions in addition to reproducing kernel spaces on finite domain $ [t_0, t_f] $.

    Definition 2.1. [55] Let $ n \in \mathbb{R}^+ $. The operator $ \mathscr{J}^{n}_{t_0} $ defined on $ L_1[t_0, t_f] $ by

    $\mathscr{J}^{n}_{t_0}f(x): = \frac{1}{\Gamma(n)}\int_{t_0}^{x}(x-\zeta)^{n-1}f(\zeta)d\zeta$,

    for $ t_0\leq x\leq t_f $, is called the Riemann-Liouville fractional integral operator of order $ n $. For $ n = 0 $, we set $ \mathscr{J}^{0}_{t_0}: = I $, the identity operator.

    Definition 2.2. [55] Let $ n\in \mathbb{R}^+ $ and $ m = \left[n \right] $. The operator $ \mathcal{D}^{n}_{t_0} $ defined by

    $\mathcal{D}^{n}_{t_0}f: = \mathcal{D}^m \mathscr{J}^{m-n}_{t_0}f = \frac{1}{\Gamma(m-n)}(\frac{d}{dx})^m\int^{x}_{t_0}(x-\zeta)^{m-n-1}f(\zeta)d\zeta$,

    is called the Riemann-Liouville fractional differential operator of order $ n $. For $ n = 0 $, we set $ \mathcal{D}^{0}_{t_0}: = I $, the identity operator.

    Definition 2.3. [55] Let $ \alpha \in \mathbb{R}^+ $ and $ n-1 < \alpha < n $. The operator $ \mathcal{D}^{\alpha}_{*t_0} $ defined by

    $ \mathcal{D}^{\alpha}_{*t_0} f(x) = \mathscr{J}^{n-\alpha}_{t_0}\mathcal{D}^n f(x) = \frac{1}{\Gamma(n-\alpha)}\int^x_{t_0} (x-\zeta)^{n-\alpha-1}(\frac{d}{d\zeta})^n f(\zeta)d\zeta$,

    for $ t_0\leq x\leq t_f $, is called the Caputo differential operator of order $ \alpha $.

    Definition 2.4. [35] Let $ \mathcal{M} $ be nonempty set, the function $ \mathcal{K} :\mathcal{M}\times \mathcal{M}\longrightarrow \mathbb{C} $ is a reproducing kernel of the Hilbert space $ \mathcal{H} $ if the following conditions are met:

    (1) $ \mathcal{K}(., t)\in \mathcal{M}, \forall t \in \mathcal{M} $,

    (2) the reproducing property: $ \forall t\in \mathcal{M}, \forall z\in \mathcal{H}: \langle z(.), \mathcal{K}(., t) \rangle = z(t). $

    The second condition means that the value of $ z $ at the point $ t $ is reproduced by the inner product of $ z $ with $ \mathcal{K} $.

    Note: The reproducing kernel is unique, symmetric and positive definite.

    Definition 2.5. $ L^2[t_0, t_f] = \left\{ \vartheta\; |\int_{t_0}^{t_f}\vartheta^2(t)\; dt < \infty \right\} $.

    Definition 2.6. The space $ \mathcal{W}^{1}_{2}[t_0, t_f] $ is defined as:

    $ \mathcal{W}^{1}_{2}[t_0, t_f] = \left\{ \vartheta|\vartheta\;{ is \;absolutely \;continuous\; real \;value \;function, }\;\vartheta'\in L^{2}[t_0, t_f] \right\}$.

    The inner product and its norm are given by:

    $\left\{ ϑ1(t),ϑ2(t)W12=ϑ1(t0)ϑ2(t0)+tft0ϑ1(t)ϑ2(t)dt,ϑW12=ϑ(t),ϑ(t)W12.
    \right.$

    Definition 2.7. The space $ \mathcal{W}^{2}_{2}[t_0, t_f] $ is defined by:

    $ \mathcal{W}^{2}_{2}[t_0, t_f] = \left\{ \vartheta|\vartheta, \vartheta'\;{ are \;absolutely\; continuous \;real \;value \;functions, }\vartheta''\in L^{2}[t_0, t_f], \vartheta(t_0) = 0 \right\} $.

    The inner product and its norm are given by:

    $\left\{ ϑ1(t),ϑ2(t)W22=ϑ1(t0)ϑ2(t0)+ϑ1(t0)ϑ2(t0)+tft0ϑ1(t)ϑ2(t)dt,ϑW22=ϑ(t),ϑ(t)W22.
    \right.$

    Definition 2.8. $ \mathcal{W}^{3}_{2}[t_0, t_f]\! = \!\left\{ \vartheta|\vartheta, \vartheta', \vartheta''\;{are\; absolutely\; continuous, }\vartheta^{(3)}\!\in L^{2}[t_0, t_f], \vartheta(t_0)\! = \!0, \vartheta(t_f)\! = \!0 \right\} $.

    The inner product and its norm in $ \mathcal{W}^{3}_{2}[t_0, t_f] $ are given by:

    $ \left\{ ϑ1(t),ϑ2(t)W32=2i=0ϑ(i)1(t0)ϑ(i)2(t0)+tft0ϑ(3)1(t)ϑ(3)2(t)dt,ϑW32=ϑ(t),ϑ(t)W32,ϑW32.
    \right. $

    Remark 2.1. The Hilbert space $ \mathcal{W}_{2}^{m}[t_0, t_f] $ is called a reproducing kernel if for any fixed $ t \in [t_0, t_f] $, $ \exists \mathcal{K}_{t}(s)\in \mathcal{W}_{2}^{m}[t_0, t_f] $ such that $ \langle \vartheta(s), \mathcal{K}_{t}(s) \rangle_{\mathcal{W}_{2}^{m}} = \vartheta(t) $ for any $ \vartheta(s) \in \mathcal{W}_{2}^{m}[t_0, t_f] $ and $ s \in [t_0, t_f] $.

    Remark 2.2.

    (1) In [56], $ \mathcal{W}^{1}_{2} $ is RKHS and its reproducing kernel is:

    $ \mathcal{K}_{1}(t, s) = \frac{1}{2 sinh 1}[cosh(t+s-1)+cosh|t-s|-1]$.

    (2) In [57], $ \mathcal{W}^{2}_{2} $ is RKHS and its reproducing kernel is:

    $\mathcal{K}_{2}(s, t) = \frac{1}{6} \left\{ t(t2+3s(2+t)) ts,s(s2+3t(2+s)) t>s.
    \right.$

    In this section, we develop an iterative method for constructing and calculating fractional differential equations with a temporal two-point solution. In order to emphasize the idea, we start by considering the general form of the BVP:

    $ {Dαt0X(t)=F(t,X(t),Y(t)),Dαt0Y(t)=G(t,X(t),Y(t)),t0ttf,0α1.
    $
    (3.1)

    Subject to BC's:

    $ X(t0)=δ,Y(tf)=β.
    $
    (3.2)

    where:

    $ \delta $, $ \beta $ $ \in \mathbb{R} $, and $ \mathcal{D}^{\alpha} $ denotes the Caputo fractional derivative of order $ \alpha $ and

    $ \left\{ X(t)=[x1(t),x2(t),...,xm(t)],Y(t)=[y1(t),y2(t),...,yl(t)],
    \right.\text{ and } \left\{ δ=[δ1(t),δ2(t),...,δm(t)],β=[β1(t),β2(t),...,βl(t)],F=[f1(t),f2(t),...,fm(t)],G=[g1(t),g2(t),...,gl(t)].
    \right.$

    We use the RKHS method to obtain a solution of BVPs (3.1) and (3.2) based on the following methodology:

    ● To attain a problem with homogenous BC's, we first assume that: $ Y(t_0) = \gamma $, ($ \gamma $ arbitrary) and

    $ {U(t)=X(t)X(t0),V(t)=Y(t)Y(t0).
    $
    (3.3)

    We get:

    $ {Dαt0U(t)=Dαt0X(t),Dαt0V(t)=Dαt0Y(t).
    $
    (3.4)

    Subject to:

    $ {U(t0)=0,V(t0)=Y(t0)γ=0.
    $
    (3.5)

    ● Then, we construct the reproducing kernel space $ \mathcal{W}_{2}^{2}[t_0, t_f] $ in which each function satisfies the homogeneous boundary conditions of (3.5) using the space $ \mathcal{W}_{2}^{1}[t_0, t_f] $.

    Take $ \mathcal{K}_t(\tau) $ and $ \mathcal{R}_t(\tau) $ to be the reproducing kernel functions of the spaces $ \mathcal{W}_{2}^{2}[t_0, t_f] $ and $ \mathcal{W}_{2}^{1}[t_0, t_f] $ respectively.

    ● Next, we define the invertible bounded linear operator $ L:\mathcal{W}_{2}^{2}[t_0, t_f]\longrightarrow \mathcal{W}_{2}^{1}[t_0, t_f] $ such that:

    $ {LU(t)=Dαt0U(t),LV(t)=Dαt0V(t).
    $
    (3.6)

    The BVPs (3.4), (3.5) can therefore be transformed to the following form:

    $ {LU(t)=F(t,X(t),Y(t)),LV(t)=G(t,X(t),Y(t)),U(t0)=0,V(t0)=0.
    $
    (3.7)

    Where $ \mathcal{U}(t) $ and $ \mathcal{V}(t) $ are in $ \mathcal{W} _{2}^{2}[t_0, t_f] $ and $ \mathcal{F}, \mathcal{G}\in \mathcal{W} _{2}^{1}[t_0, t_f] $.

    Applying Riemann-Liouville fractional integral operator $ \mathscr{J}^{\alpha}_{t_0} $ to both sides using $ \mathcal{U}(t_{0}) = 0 $ and $ \mathcal{V}(t_{0}) = 0 $, we get:

    $U(t)=1Γ(t0)tt0(tτ)α1F(τ,X(τ),Y(τ))dτ=F(t,X(t),Y(t)),V(t)=1Γ(t0)tt0(tτ)α1G(τ,X(τ),Y(τ))dτ=G(t,X(t),Y(t)).
    $

    Thus, we can notice that: $ L \mathcal{U}(t) = \mathcal{U}(t), $ and so the BVPs are transformed to the equivalent form:

    $ {U(t)=F(t,X(t),Y(t)),V(t)=G(t,X(t),Y(t)),U(t0)=0,V(t0)=0.
    $
    (3.8)

    ● When choosing a countable dense set $ \left\{ t _{i} \right\}_{i = 1}^{\infty } $ from $ [t_0, t_f] $ for the reproducing kernel of the space $ \mathcal{W} _2^{2}[t_0, t_f] $, we define a complete system on $ \mathcal{W} _2^{2}[t_0, t_f] $ as: $ \Psi_i(t) = L^*\Phi_i(t) $ where $ \Phi_{i}(t) = \mathcal{R}_{t _{i}}(\tau) $, and $ L^* $ is the adjoint operator of $ L $.

    Lemma 3.1. $ \Psi_{i}(t) $ can be written on the following form:

    $\Psi_i(t) = L_\tau\mathcal{K}_t(\tau)\vert_{ \tau = t _{i}}$.

    Proof. It is clear that:

    $Ψi(t)=LΦi(t)=LΦi(τ),Kt(τ)W22,=Φi(τ),LKt(τ)W12=LτKt|τ=ti.
    $

    ● The orthonormal function system $ \{\overline{\Psi}_i^\eta(t)\}_{i = 1}^\infty $, $ \eta = 1, 2 $ of the space $ \mathcal{W}_2^{2}[t_0, t_f] $ can be derived from Gram-Schmidt orthogonalization process of $ \{\Psi_i^\eta(t)\}_{i = 1}^\infty $ as follows:

    $ \overline{\Psi}_{i}^\eta(t) = \sum_{k = 1}^{i}\mathcal{B}^{\eta}_{ik}\Psi_{k}^\eta(t), \; \; i = 1, 2, ..., \; \; \eta = 1, 2$,

    where $ \mathcal{B}^{\eta}_{ik} $ are positive orthogonalization coefficients such that:

    $ Bη11=1Ψη1,Bηii=1Ψηi2i1k=1(Cηik)2,Bηij=i1k=1CηikBηkjΨηi2i1k=1(Cηik)2,j<i.
    $
    (3.9)

    $ \mathcal{C}^\eta_{ik} $ given by: $ \langle \Psi_{i}^\eta, \Psi_{k}^\eta\rangle_{\mathcal{W}_{2}^{2}} $.

    Theorem 3.1. If the operator $ L $ is invertible i.e: $ L^{-1} $ exist, and if $ \left\{ t _{i} \right\}_{i = 1}^{\infty } $ is dense on $ [t_0, t_f] $, then $ \left\{ \Psi_{i}^\eta \right\}_{i = 1}^{\infty } $, $ \eta = 1, 2 $ is the complete function system of the space $ \mathcal{W}_2^{2}[t_0, t_f] $.

    Proof. For each fixed $ \mathcal{U}(t), \; \mathcal{V}(t) \in \mathcal{W}_2^{2}[t_0, t_f] $, let $ \langle \mathcal{U}(t), \Psi_{i}^1(t) \rangle = 0, \; \text{and }\; \langle \mathcal{V}(t), \Psi_{i}^2(t) \rangle = 0, \; \; i = 1, 2, ... $ that is:

    $U(t),Ψ1i(t)W22=U(t),LΦ1i(τ)W22=LU(t),Φ1i(t)W12=LU(ti)=0,V(t),Ψ2i(t)W22=V(t),LΦ2i(τ)W22=LV(t),Φ2i(t)W12=LV(ti)=0,
    $

    since $ \{t _{i}\}_{i = 1}^\infty $ is dense on $ [t_0, t_f] $ then $ L\mathcal{U}(t) = 0 $, and $ L \mathcal{V}(t) = 0 $ it follows that $ \mathcal{U}(t) = 0, \; \mathcal{V}(t) = 0 $ since $ L^{-1} $ exist and $ \mathcal{U}(t), \; \mathcal{V}(t) $ are continuous.

    Theorem 3.2. For each $ \mathcal{U}(t), \; \mathcal{V}(t)\in \mathcal{W}_2^{2}[t_0, t_f] $ the series

    $\left\{i=0U(t),¯Ψ1i(t)W22¯Ψ1i(t),i=0V(t),¯Ψ2i(t)W22¯Ψ2i(t),
    \right.$

    are convergent in the sense of the norm of $ \mathcal{W}_2^{2}[t_0, t_f] $. In contrast if $ \left\{ t_{i} \right\}_{i = 1}^{\infty } $ is dense subset on $ [t_0, t_f] $ then the solutions of (3.8) given by:

    $ {U(t)=i=1ik=1B1ikF(tk,U(tk),V(tk))¯Ψ1i(t),V(t)=i=1ik=1B2ikG(tk,U(tk),V(tk))¯Ψ2i(t).
    $
    (3.10)

    Proof. Let $ \mathcal{U}(t), \; \mathcal{V}(t)\in \mathcal{W}_2^{2}[t_0, t_f] $ be the solutions of (3.8), since $ \mathcal{U}(t), \; \mathcal{V}(t)\in \mathcal{W}_2^{2}[t_0, t_f] $, and $ \sum_{i = 1}^{\infty }\langle \mathcal{U}(t), \overline{\Psi}^1_{i}(t) \rangle_{\mathcal{W}_2^{2}[t_0, t_f]}\overline{\Psi}^1_{i}(t) $ and $ \sum_{i = 1}^{\infty }\langle \mathcal{V}(t), \overline{\Psi}^2_{i}(t) \rangle_{\mathcal{W}_2^{2}[t_0, t_f]}\overline{\Psi}^2_{i}(t) $ represent the Fourier series expansion about normal orthogonal system $ \{\overline{\Psi}^\eta_{i}(t)\}_{i = 1}^{\infty} $, $ \eta = 1, 2 $, and $ \mathcal{W}_2^{2}[t_0, t_f] $ is Hilbert space, then the series $ \sum_{i = 1}^{\infty }\langle \mathcal{U}(t), \overline{\Psi}^1_{i}(t) \rangle_{\mathcal{W}_2^{2}[t_0, t_f]}\overline{\Psi}^1_{i}(t), \; \sum_{i = 1}^{\infty }\langle \mathcal{V}(t), \overline{\Psi}^2_{i}(t) \rangle_{\mathcal{W}_2^{2}[t_0, t_f]}\overline{\Psi}^2_{i}(t) $ are convergent in the sense of $ \left\|. \right\|_{\mathcal{W}_2^{2}[t_0, t_f]} $. In contrast, according to the orthogonal basis $ \{\overline{\Psi}^\eta_{i}(t)\}_{i = 1}^\infty $, we have:

    $U(t)=i=1U(t),¯Ψ1i(t)W22¯Ψ1i(t),=i=1U(t),ik=1B1ikΨ1k(t)W22¯Ψ1i(t),=i=1ik=1B1ikU(t),Ψ1k(t)W22¯Ψ1i(t),=i=1ik=1B1ikU(t),LΦ1k(t)W22¯Ψ1i(t),=i=1ik=1B1ikLU(t),Φ1k(t)W12¯Ψ1i(t),=i=1ik=1B1ikF(tk,U(t),V(t)),Φ1k(t)W12¯Ψ1i(t),=i=1ik=1B1ikF(tk,U(tk),V(tk)))¯Ψ1i(t).
    $

    The same for finding $ \mathcal{V}(t) $:

    $\mathcal{V}(t) = \sum_{i = 1}^{\infty }\sum_{k = 1}^{i}\mathcal{B}^2_{ik}\mathrm{G}(t_{k}, \mathcal{U}(t_{k}), \mathcal{V}(t_{k}))\overline{\Psi}^2_{i}(t).$

    The theorem is proved.

    Since $ \mathcal{W}_{2}^{2} $ is Hilbert space we get:

    $ \sum_{i = 1}^{\infty }\sum_{k = 1}^{i}\mathcal{B}^1_{ik}\langle L\mathcal{U}(t), \Phi^1_{k}(t) \rangle_{\mathcal{W}_{2}^{1}}\overline{\Psi}^1_{i}(t) < \infty $ and $ \sum_{i = 1}^{\infty }\sum_{k = 1}^{i}\mathcal{B}^2_{ik}\langle L\mathcal{V}(t), \Phi^2_{k}(t) \rangle_{\mathcal{W}_{2}^{1}}\overline{\Psi}^2_{i}(t) < \infty $.

    Hence:

    $ {Un(t)=ni=1ik=1B1ikF(tk,U(tk),V(tk))¯Ψ1i(t),Vn(t)=ni=1ik=1B2ikG(tk,U(tk),V(tk))¯Ψ2i(t),
    $
    (3.11)

    are convergent in the sense of $ \left\|. \right\|_{\mathcal{W}_{2}^{2}} $ and (3.11) represents the numerical solution of (3.8).

    Remark 3.1.

    (1) If the system (3.7) is linear, then the exact solutions can be found directly from (3.10).

    (2) If the system (3.7) is non linear, then the exact and numerical solutions can be obtained by:

    $ {U(t)=i=1A1i¯Ψ1i(t),V(t)=i=1A2i¯Ψ2i(t),
    $
    (3.12)

    where:

    $ {A1i=ik=1B1ikF(tk,Uk1(tk),Vk1(tk)),A2i=ik=1B2ikG(tk,Uk1(tk),Vk1(tk)).
    $
    (3.13)

    We use the known quantities $ \lambda_{i}^{\eta }, \; \eta = 1, 2 $ to approximate the unknowns $ \mathcal{A}_{i}^{\eta}, \; \eta = 1, 2 $ as follows: we put $ t_{1} = t_{0} $ and set $ \mathcal{U}_{0}(t_{1}) = \mathcal{U}(t_{1}), \; \mathcal{V}_{0}(t_{1}) = \mathcal{V}(t_{1}) $ then $ \mathcal{U}_{0}(t_{1}) = \mathcal{V}_{0}(t_{1}) = 0 $ from the conditions of (3.8), and define the n-term approximation to $ \mathcal{U}(t), \; \mathcal{V}(t) $ by:

    $ {Un(t)=ni=1λ1i¯Ψ1i(t),Vn(t)=ni=1λ2i¯Ψ2i(t),
    $
    (3.14)

    where the coefficient $ \lambda_{i}^{\eta }\; \; (\eta = 1, 2, \; \; i = 1, 2, ..., n $), are presented as follows:

    $ {λ1n=nk=1B1ikF(tk,Uk1(tk),Vk1(tk)),λ2n=nk=1B2ikG(tk,Uk1(tk),Vk1(tk)),
    $
    (3.15)

    and so:

    $ {Un(t)=ni=1λ1i¯Ψ1i(t),Vn(t)=ni=1λ2i¯Ψ2i(t).
    $
    (3.16)

    We can guarantee that the approximations $ \mathcal{U}_n(t), \; \mathcal{V}_n(t) $ satisfies the conditions enjoined by (3.7) through the iterative process of (3.16).

    In this section, we present some convergence theories to emphasize that the approximate solution we got is close to the exact solution. Indeed, this finding is very powerful and efficient to RKHS theory and its applications.

    Lemma 4.1. $ \left\| \mathcal{U}_n(t) \right\|_{n = 1}^{\infty } $, and $ \left\| \mathcal{V}_n(t) \right\|_{n = 1}^{\infty } $ are monotone increasing in the sense of the norm of $ \left\|. \right\|^{2}_{{{\mathcal{W} _{2}^{2}}}} $.

    Proof. Since $ \left\| \overline{\Psi}^\eta _{i}(t) \right\|_{i = 1}^{\infty }, \; \eta = 1, 2 $ are the complete orthonormal systems in the space $ {{\mathcal{W} _{2}^{2}}}[t_0, t_f] $ then we have:

    $ \left\{ Un(t)2W22=Un(t),Un(t)W22=ni=1λ1i¯Ψ1i(t),ni=1λ1i¯Ψ1i(t)W22=ni=1(λ1i)2,Vn(t)2W22=Vn(t),Vn(t)W22=ni=1λ2i¯Ψ2i(t),ni=1λ2i¯Ψ2i(t)W22=ni=1(λ2i)2.
    \right.$

    Thus $ \left\| \mathcal{U}_n(t) \right\|_{{{\mathcal{W} _{2}^{2}}}}, \; \left\| \mathcal{V}_n(t) \right\|_{{{\mathcal{W} _{2}^{2}}}} $ are monotone increasing.

    Lemma 4.2. As $ n\to \infty $, the approximate solutions $ \mathcal{U}_{n}(t), \; \mathcal{V}_{n}(t) $ and its derivatives $ \mathcal{U'}_{n}(t), \; \mathcal{V'}_{n}(t) $ are uniformly convergent to the exact solutions $ \mathcal{U}(t), \; \mathcal{V}(t) $ and its derivatives $ \mathcal{U'}(t), \; \mathcal{V'}(t) $ respectively.

    Proof. For any $ t \in [t_0, t_f] $:

    $|Un(t)U(t)|=|Un(t)U(t),Kt(τ)W22|,Kt(τ)W22Un(t)U(t)W22,N1Un(t)U(t)W22,N1R,
    $

    and

    $\left| \mathcal{V'}_{n}(t)-\mathcal{V'}(t) \right| \le \mathscr{N}_{2}\left\| \mathcal{V}_{n}(t)-\mathcal{V}(t) \right\|_{\mathcal{W} _{2}^{2}}, \; \; \mathscr{N}_{2}\in \mathbb{R}, $

    if $ \left\| \mathcal{U}_{n}(t)-\mathcal{U}(t) \right\|_{\mathcal{W} _{2}^{2}}\longrightarrow 0, \; \left\| \mathcal{V}_{n}(t)-\mathcal{V}(t) \right\|_{\mathcal{W} _{2}^{2}}\longrightarrow 0 $ as $ n\to \infty $, then the approximate solutions $ \mathcal{U}^{(i)}_{n}(t), \; \mathcal{V}^{(i)}_{n}(t) $ are uniformly converges to the exact solutions $ \mathcal{U}^{(i)}(t), \; \mathcal{V}^{(i)}(t)\; \; \\i = 1, 2 $ respectively.

    Theorem 4.1. If

    $\left\{ Un(t)U(t),Vn(t)V(t),
    \right.$

    and $ \mathrm{F}(t, \mathcal{U}(t), \mathcal{V}(t)), \; \mathrm{G}(t, \mathcal{U}(t), \mathcal{V}(t)) $ are continuous in $ [t_0, t_f] $, then:

    $ {F(tn,Un1(tn),Vn1(tn))F(t,U(t),V(t))G(tn,Un1(tn),Vn1(tn))G(t,U(t),V(t))asn.
    $
    (4.1)

    Proof. For the first part, we will prove that:

    $\left\{ Un1(tn)U(t),Vn1(tn))V(t),
    \right. $

    it is easy to see that:

    $ \left\{ |Un1(tn)U(t)|=|Un1(tn)Un1(t)+Un1(t)U(t)||Un1(tn)Un1(t)|+|Un1(t)U(t)|,|Vn1(tn)V(t)|=|Vn1(tn)Vn1(t)+Vn1(t)V(t)||Vn1(tn)Vn1(t)|+|Vn1(t)V(t)|,

    \right. $

    by reproducing property of $ \mathcal{K}_{t}(\tau) $ we have:

    $\left\{ Un1(tn)=Un1(τ),Ktn(τ),Vn1(tn)=Vn1(τ),Ktn(τ),
    \right.$

    and

    $\left\{ Un1(t)=Un1(τ),Kt(τ),Vn1(t)=Vn1(τ),Kt(τ),
    \right.$

    thus

    $ \left\{ |Un1(tn)Un1(t)|=|Un1(τ),Ktn(τ)Kt(τ)W22|Un1(τ)W22Ktn(τ)Kt(τ)W22,|Vn1(tn)Vn1(t)|=|Vn1(τ),Ktn(τ)Kt(τ)W22|Vn1(τ)W22Ktn(τ)Kt(τ)W22,

    \right. $

    and from the symmetric property of $ \mathcal{K}_{t}(\tau) $ we get:

    $\left\| \mathcal{K}_{t_{n}}(\tau )-\mathcal{K}_{t}(\tau ) \right\|_{\mathcal{W}_{2}^{2}}\underset{n\to \infty }{\longrightarrow 0}$,

    hence: $ \left| \mathcal{U}_{n-1}(t_{n})-\mathcal{U}_{n-1}(t) \right|\longrightarrow 0 $ as $ t_{n}\to t $.

    By lemma (4.2)

    $\left\{ Un(t)c.uU(t),Vn(t)c.uV(t),
    \right.$

    thus:

    $\left\{ |Un1(t)U(t)|0|Vn1(t)V(t)|0
    \right.\text{ as } n\longrightarrow \infty. $

    Therefore

    $ \left\{ Un1(tn)U(t),Vn1(tn)V(t),
    \right.$

    in the sense of the $ \left\|. \right\|_{\mathcal{W}_{2}^{2}} $ as $ t_n\longrightarrow t $ and $ n \longrightarrow \infty $ for any $ t \in [t_0, t_f] $.

    Moreover, since $ \mathrm{F} $ and $ \mathrm{G} $ are continuous, we obtain:

    $\left\{ F(tn,Un1(tn),Vn1(tn))F(t,U(t),V(t))G(tn,Un1(tn),Vn1(tn))G(t,U(t),V(t))
    \right.\text{ as } n\longrightarrow \infty.$

    Theorem 4.2. Suppose that $ \left\| \mathcal{U}_{n} \right\|_{\mathcal{W}_{2}^{2}} $ and $ \left\| \mathcal{V}_{n} \right\|_{\mathcal{W}_{2}^{2}} $ are bounded in Eq (3.14), if $ \left\{ t_{i} \right\}_{i = 1}^{\infty } $ is dense on $ [t_0, t_f] $, then the approximate solutions $ \mathcal{U}_{n}(t) $, $ \mathcal{V}_{n}(t) $ in Eq (3.14) convergent to the exact solutions $ \mathcal{U}(t), \; \mathcal{V}(t) $ of Eq (3.7) in the space $ {\mathcal{W}_{2}^{2}}[t_0, t_f] $ and $ \mathcal{U}(t), \; \mathcal{V}(t) $ given by (3.12).

    Proof. We first start by proving the convergence of $ \mathcal{U}_{n}(t) $ and $ \mathcal{V}_{n}(t) $ from Eq (3.14) we conclude that:

    $ \left\{ Un+1(t)=Un(t)+λ1n+1¯Ψ1n+1(t),Vn+1(t)=Vn(t)+λ2n+1¯Ψ2n+1(t),
    \right.$

    by orthogonality of $ \left\{ \overline{\Psi}^\eta _{i}(t) \right\}_{i = 1}^{\infty }, \; (\eta) = 1, 2 $ we get:

    $ \left\{ Un+1(t)2W22=Un(t)2W22+(λ1n+1)2==U0(t)2W22+n+1i=1(λ1i)2,Vn+1(t)2W22=Vn(t)2W22+(λ2n+1)2==V0(t)2W22+n+1i=1(λ2i)2,
    \right.$

    $ \left\| \mathcal{U}_{n}(t) \right\|_{\mathcal{W}_{2}^{2}}, \; \left\| \mathcal{V}_{n}(t) \right\|_{\mathcal{W}_{2}^{2}} $ are monotone increasing by Lemma (2). From the assymption that $ \left\| \mathcal{U}_{n}(t) \right\|_{\mathcal{W}_{2}^{2}}, \; \left\| \mathcal{V}_{n}(t) \right\|_{\mathcal{W}_{2}^{2}} $ are bounded, $ \left\| \mathcal{U}_{n}(t) \right\|_{\mathcal{W}_{2}^{2}}, \; \left\| \mathcal{V}_{n}(t) \right\|_{\mathcal{W}_{2}^{2}} $ are convergent as $ n \to \infty $, then $ \exists \; c, \; d $ constants such that

    $\left\{ i=1(λ1i)2=c,i=1(λ2i)2=d,
    \right.$

    if $ m > n $ using

    $ \left\{ (UmUm1)(Um1Um2)(Un+1Un),(VmVm1)(Vm1Vm2)(Vn+1Vn),
    \right.$

    further that

    $\left\{ Um(t)Um1(t)2W22=(λ1m)2,Vm(t)Vm1(t)2W22=(λ2m)2,
    \right.$

    so:

    $\left\{ Um(t)Un(t)2W22=mi=n+1(λ1i)20Vm(t)Vn(t)2W22=mi=n+1(λ2i)20
    \right.\text{ as } n, m\to \infty , $

    since $ \mathcal{W}_{2}^{2}[t_0, t_f] $ is complete, $ \exists \; \mathcal{U}(t), \; \mathcal{V}(t) $ in $ \mathcal{W}_{2}^{2}[t_0, t_f] $ such that

    $\left\{ Un(t)U(t)Vn(t)V(t)
    \right.\text{ as } n\longrightarrow \infty , $

    in the sense of the norm of $ \mathcal{W}_{2}^{2}[t_0, t_f] $.

    Now, we prove that $ \mathcal{U}(t), \; \mathcal{V}(t) $ are solutions of Eq (3.7). Since $ \left\{ t_{i} \right\}_{i = 1}^{\infty } $ is dense on $ [t_0, t_f], \; \forall\; t \in [t_0, t_f], \; \exists $ subsequence $ \left\{ t_{n_j} \right\} $ such that $ t_{n_j}\underset{j\to \infty }{\longrightarrow }t $. From lemma (3) and (4) in [25] we have:

    $\left\{ LU(tnj)=F(tnj,Unj1(tnj),Vnj1(tnj)),LV(tnj)=G(tnj,Unj1(tnj),Vnj1(tnj)),
    \right.$

    let $ j $ goes to $ \infty $, by theorem (4.1) and the continuity of $ \mathrm{F} $ and $ \mathrm{G} $ we have:

    $\left\{ LU(t)=F(t,U(t),V(t)),LV(t)=G(t,U(t),V(t)),
    \right.$

    that is $ \mathcal{U}(t), \; \mathcal{V}(t) $ are solutions of Eq (3.7).

    Theorem 4.3. Let $ \xi_{n} = \left| \mathcal{U}_n(t)-\mathcal{U}(t) \right| $, $ \xi^{'}_{n} = \left| \mathcal{V}_{n}(t)-\mathcal{V}(t) \right| $, where: $ \mathcal{U}_{n}(t), \; \mathcal{V}_{n}(t), \; \mathcal{U}(t), \; \mathcal{V}(t) $ denote the approximate and the exact solutions respectively, then the sequences of numbers $ \left\{ \xi _{n} \right\}, \; \left\{ \xi^{'} _{n} \right\} $ are decreasing in the sense of the norm $ \left\|. \right\|_{{\mathcal{W}_{2}^{2}}} $ and $ \xi _{n}\underset{n\to \infty }{\longrightarrow }0, \; \xi^{'} _{n}\underset{n\to \infty }{\longrightarrow }0 $.

    Proof. From the extension form of $ \mathcal{U}(t), \; \mathcal{V}(t) $ and $ \mathcal{U}_n(t), \; \mathcal{V}_{n}(t) $ in Eqs (3.12), (3.14) and (3.15) we can write:

    $\left\{ ξn2W22=i=n+1λ1i¯Ψ1i(t)2W22=i=n+1(λ1i)2,ξn2W22=i=n+1λ2i¯Ψ2i(t)2W22=i=n+1(λ2i)2,
    \right.$

    and

    $\left\{ ξn12W22=i=nλ1i¯Ψ1i(t)2W22=i=n(λ1i)2,ξn12W22=i=nλ2i¯Ψ2i(t)2W22=i=n(λ2i)2.
    \right.$

    Clearly: $ \left\| \xi _{n} \right\|^{\infty}_{n = 1}, \; \left\| \xi^{'} _{n} \right\|^{\infty}_{n = 1} $ are decreasing in a sense of $ \left\|. \right\|_{{{\mathcal{W}_{2}^{2}}}} $ from theorem (3.2) the series $ \sum_{i = 1}^{\infty }\lambda_{i}^{1}\overline{\Psi}^1_{i}(t), \; \sum_{i = 1}^{\infty }\lambda_{i}^{2}\overline{\Psi}^2_{i}(t) $ are convergent, thus $ \left\| \xi _{n} \right\|_{{{\mathcal{W}_{2}^{2}}}}\longrightarrow 0, \; \left\| \xi^{'} _{n} \right\|_{{{\mathcal{W}_{2}^{2}}}}\longrightarrow 0 $ as $ n\longrightarrow \infty $.

    Theorem 4.4. The approximate solutions $ \mathcal{U}_n(t), \; \mathcal{V}_n(t) $ of (3.7) converge to its exact solutions $ \mathcal{U}(t), \; \mathcal{V}(t) $ with not less than the second order convergence. That is: $ \left| \mathcal{U}_n-\mathcal{U} \right| \le Mk^{2} $ and $ \left| \mathcal{V}_n-\mathcal{V} \right| \le Nk^{2} $, where $ k = \frac{t_f-t_0}{n} $.

    Proof. See [36].

    Numerical examples are conducted in order to verify the accuracy of this method. Computations are performed using Mathematica 11.0.

    Algorithm 1: Use the following stages to approximate the solutions of BVPs (3.4) and (3.5) based on RKHS method.

    Stage A: Fixed $ t \in [t_0, t_f] $ and set $ \tau \in [t_0, t_f] $

    for $ i = 1, ..., n $ do the following stages:

    - stage 1: set $ t_i = t_0+\frac{(t_f-t_0)i}{n} $;

    - stage 2: if $ \tau \leq t $ let

    $\mathcal{K}_{\tau}(t) = \sum_{i = 0}^{3} p_i(t)\tau^i;$

    else let

    $\mathcal{K}_{\tau}(t) = \sum_{i = 0}^{3} q_i(t)\tau^i.$

    - stage 3: For $ \eta = 1, 2 $;

    set

    $\Psi_i^\eta(t) = L_\tau\mathcal{K}_t(\tau)\vert_{ \tau = t _{i}}.$

    Output the orthogonal functions system $ \Psi_i^\eta(t) $.

    Stage B: Obtain the orthogonalization coefficients $ \mathcal{B}^{\eta}_{ij} $ as follows:

    For $ \; \; \eta = 1, 2 $;

    For $ \; \; i = 1, ..., n $;

    For $ \; \; j = 1, ..., i\; \; $ set $ \; \; \mathcal{C}_{ik}^{\eta} = \langle \Psi_{i}^\eta, \Psi_{j}^\eta\rangle_{\mathcal{W}_{2}^{2}} $ and $ \mathcal{B} _{11} = \frac{1}{Sqrt (\mathcal{C}_{11}^{\eta})} $.

    Output $ \mathcal{C}_{ij}^{\eta} $ and $ \mathcal{B} _{11} $.

    Stage C: For $ \; \; \eta = 1, 2 $;

    For $ i = 1, ..., n $, set $ \mathcal{B}^\eta _{ii} = (\left\| \Psi_{i}^\eta \right\|^{2}_{\mathcal{W}^{2}_{2}}-\sum_{k = 1}^{i-1}(\mathcal{C}_{ik}^\eta)^{2})^{\frac{-1}{2}} $;

    else if $ j\neq i\; \; \; \; \; \; $ set $ \mathcal{B}^{\eta}_{ij} = -(\sum_{k = 1}^{i-1}\mathcal{C}^{\eta}_{ik}\mathcal{B}^{\eta}_{kj}).(\left\| \Psi_{i}^{\eta} \right\|^{2}_{\mathcal{W}^{2}_{2}}-\sum_{k = 1}^{i-1}(\mathcal{C}^{\eta}_{ik})^{2})^\frac{-1}{2} $.

    Output the orthogonalization coefficients $ \mathcal{B}^{\eta}_{ij} $.

    Stage D: For $ \; \; \eta = 1, 2 $;

    For $ \; \; i = 1, ..., n\; \; $ set $ \; \; \overline{\Psi}_i^\eta(t) = \sum_{k = 1}^{i}\mathcal{B}^{\eta}_{ik}\Psi_{i}^{\eta}(t) $.

    Output the orthonormal functions system $ \overline{\Psi}_i^\eta(t). $

    Stage E: Set $ \; \; t_1 = 0\; \; $ and choose $ \mathcal{U}_0(t_1) = 0, \; \; \mathcal{V}_0(t_1) = 0 $;

    For $ \; \; \eta = 1, 2 $;

    For $ \; \; i = 1\; \; $ set

    $\left\{ λ11=B111F(t1,U0(t1),V0(t1)),λ21=B211G(t1,U0(t1),V0(t1)),
    \right.\; $ and $\left\{U1(t)=λ11¯Ψ11(t),V1(t)=λ21¯Ψ21(t).
    \right.$

    For $ \; \; i = 2, 3, ..., n\; \; $ set

    $ \left\{ λ1i=ik=1B1nkF(tk,Uk1(tk),Vk1(tk)),λ2i=ik=1B2nkG(tk,Uk1(tk),Vk1(tk)),
    \right.$

    set

    $\left\{ Un(t)=ni=1λ1i¯Ψ1i(t),Vn(t)=ni=1λ2i¯Ψ2i(t).
    \right.$

    Outcome the numerical solutions $ \mathcal{U}_n(t), \; \mathcal{V}_n(t) $.

    Then we implement the above algorithm using numerical simulations. We arrange the resulting data in tables and graphs for examples discussed on $ [t_0, t_f] $ as follows:

    Example 5.1. Consider the following system:

    $\left\{Dαω(t)=4ω+3Θ+6,DαΘ(t)=2.4ω+1.6Θ+3.6,0t0.5,0α1,
    \right.$

    subject to:

    $\left\{ω(0)=0,Θ(0.5)=2.25e1+2.25e0.2,
    \right.$

    with exact solution when $ \alpha = 1 $ is:

    $\left\{ω(t)=3.375e2t+1.875e0.4t+1.5,Θ(t)=2.25e2t+2.25e0.4t.
    \right.$

    After the initial conditions have been homogenised and algorithm 1 used, apply $ t_i = \frac{0.5i}{n} $, $ \overline{i = 1, n}\; $ and $ n = 40 $, the tables 1 and 2 describe the exact solutions of $ \omega(t) $ and $ \Theta(t) $ and approximate solutions for different values of $ \alpha $.

    Table 1.  Numerical results for $ \omega(t) $ of example 5.1.
    $ t $ Exact Sol of $ \omega(t) $ App Sol of $ \omega(t) $ $ \alpha=0.9 $ $ \alpha=0.8 $ $ \alpha=0.7 $ Abs Error Rel Error
    0. 0. 0. 0. 0. 0. 0. Indeterminate
    0.1 0.538264 0.538235 0.672451 0.820896 0.975714 $ 2.8979\times 10^{-5} $ $ 5.3838\times 10^{-5} $
    0.2 0.968513 0.968496 1.10364 1.22992 1.34045 $ 1.6912\times 10^{-5} $ $ 1.7462\times 10^{-5} $
    0.3 1.31074 1.31073 1.41427 1.49734 1.55882 $ 7.6470\times 10^{-6} $ $ 5.8341\times 10^{-6} $
    0.4 1.58128 1.58128 1.64374 1.68334 1.70372 $ 5.4565\times 10^{-7} $ $ 3.4507\times 10^{-7} $
    0.5 1.79353 1.79353 1.81496 1.8167 1.805 $ 4.8467\times 10^{-6} $ $ 2.7023\times 10^{-6} $

     | Show Table
    DownLoad: CSV
    Table 2.  Numerical results for $ \Theta(t) $ of example 5.1.
    $ t $ Exact Sol $ \Theta(t) $ App Sol of $ \Theta(t) $ $ \alpha=0.9 $ $ \alpha=0.8 $ $ \alpha=0.7 $ Abs Error Rel Error
    0. 0. 0. 0. 0. 0. 0. Complex Infinity
    0.1 0.319632 0.31963 0.397424 0.48192 0.56754 $ 1.7129\times 10^{-6} $ $ 5.3589\times 10^{-6} $
    0.2 0.568792 0.568797 0.643125 0.70952 0.76353 $ 5.6398\times 10^{-6} $ $ 9.9154\times 10^{-6} $
    0.3 0.760745 0.760756 0.812467 0.84948 0.87141 $ 1.1152\times 10^{-5} $ $ 1.4659\times 10^{-5} $
    0.4 0.906333 0.906349 0.930619 0.93950 0.93585 $ 1.5248\times 10^{-5} $ $ 1.6823\times 10^{-5} $
    0.5 1.01442 1.01443 1.01229 0.99765 0.97506 $ 1.8229\times 10^{-5} $ $ 1.7970\times 10^{-5} $

     | Show Table
    DownLoad: CSV

    Graphs of the approximate solutions of $ \omega(t) $ are plotted in Figure 1 (a), for different values of $ \alpha $. It is obvious from Figure 1 (a) that the approximate solutions are in reasonable alignment with the exact solution when $ \alpha = 1 $ and the solutions are continuously based on a fractional derivative. The graph in Figure 1 (b) represent the absolute errors of $ \theta(t) $.

    Figure 1.  Solution and graphical curves of Example 5.1.

    Example 5.2. Consider the following system:

    $ \left\{ Dαω=ω24(ω1)cos2(t)sin(t),DαΘ=ωΘ2Θt2cos(t)+2t,
    \right.0 \leqslant t \leqslant 1$,

    with conditions:

    $\left\{ω(0)=3,Θ(1)=1,
    \right.$

    when $ \alpha = 1 $ the exact solution is:

    $ \left\{ ω(t)=cos(t)+2,Θ(t)=t2.
    \right.$

    After homogenizing the initial conditions and using algorithm 1, apply $ t_i = \frac{i}{n} $, $ \overline{i = 1, n}\; $ and $ n = 35 $, the tables 3 and 4 describe the exact solutions of $ \omega(t) $ and $ \Theta(t) $ and approximate solutions for different values of $ \alpha $.

    Table 3.  Numerical results for $ \omega(t) $ of example 5.2.
    $ t $ Exact Sol of $ \omega(t) $ App Sol of $ \omega(t) $ $ \alpha=0.9 $ $ \alpha=0.8 $ $ \alpha=0.7 $ Abs Error
    0. 3. 3. 3. 3. 3. 0.
    0.2 2.98007 2.98008 2.97308 2.96301 2.94782 $ 1.71231717\times 10^{-5} $
    0.4 2.92106 2.92108 2.89745 2.86363 2.81286 $ 2.336775593\times 10^{-5} $
    0.6 2.82534 2.82537 2.77748 2.71004 2.61271 $ 3.05778094\times 10^{-5} $
    0.8 2.69671 2.69675 2.6203 2.51736 2.38161 $ 3.84719926\times 10^{-5} $
    1. 2.5403 2.54035 2.43686 2.3082 2.15981 $ 4.636294967\times 10^{-5} $

     | Show Table
    DownLoad: CSV
    Table 4.  Numerical results for $ \Theta(t) $ of example 5.2.
    $ t $ Exact Sol of $ \Theta(t) $ App Sol of $ \Theta(t) $ $ \alpha=0.9 $ $ \alpha=0.8 $ $ \alpha=0.7 $ Abs Error
    0. 0. 0. 0. 0. 0. 0.
    0.2 0.04 0.0399862 0.0526074 0.0695245 0.0925163 $ 1.379651819\times 10^{-5} $
    0.4 0.16 0.159985 0.198264 0.246476 0.306978 $ 1.465169064\times 10^{-5} $
    0.6 0.36 0.359987 0.428814 0.509138 0.597815 $ 1.348748932\times 10^{-5} $
    0.8 0.64 0.639991 0.735269 0.833398 0.919358 $ 9.089432173\times 10^{-6} $
    1. 1. 1. 1.10687 1.19567 1.24369 0.

     | Show Table
    DownLoad: CSV

    Graphs of the approximate solutions of $ \theta(t) $ are plotted in Figure 2 (b) for different values of $ \alpha $. The graph in Figure 2 (a) represent the absolute errors of $ \omega(t) $.

    Figure 2.  Solution and graphical curves of Example 5.2.

    Example 5.3. Consider the following fractional system:

    $\left\{ Dαω=Θρ+t,DαΘ=3t2,Dαρ=Θ+et,0t1,
    \right.$

    subject to:

    $ \left\{ ω(0)=1,Θ(0)=1,ρ(1)=1.25e1,
    \right.$

    with exact solution:

    $\left\{ ω(t)=0.05t5+0.25t4+t+2et,Θ(t)=t3+1,ρ(t)=0.25t4+tet.
    \right.$

    After the initial conditions have been homogenised and algorithm 1 used, apply $ t_i = \frac{i}{n} $, $ \overline{i = 1, n}\; $ and $ n = 30 $, the tables 5-7 describe the exact solutions of $ \omega(t) $, $ \Theta(t) $ and $ \rho $ and approximate solutions for different values of $ \alpha $.

    Table 5.  Numerical results for $ \omega(t) $ of example 5.3.
    $ t $ Exact Sol of $ \omega(t) $ App Sol of $ \omega(t) $ $ \alpha=0.9 $ $ \alpha=0.8 $ $ \alpha=0.7 $ Absolute Error
    0. 1. 1. 1. 1. 1. 0.
    0.2 1.38165 1.38163 1.45132 1.52412 1.59527 $ 2.778531454\times 10^{-5} $
    0.4 1.73557 1.73552 1.80231 1.85993 1.90359 $ 4.434730105\times 10^{-5} $
    0.6 2.0797 2.07964 2.12935 2.16496 2.18472 $ 5.613341879\times 10^{-5} $
    0.8 2.43669 2.43662 2.46954 2.48922 2.49658 $ 6.305905268\times 10^{-5} $
    1. 2.83212 2.83206 2.8543 2.86685 2.87146 $ 6.508054219\times 10^{-5} $

     | Show Table
    DownLoad: CSV
    Table 6.  Numerical results for $ \Theta(t) $ of example 5.3.
    $ t $ Exact Sol of $ \Theta(t) $ App Sol of $ \Theta(t) $ $ \alpha=0.9 $ $ \alpha=0.8 $ $ \alpha=0.7 $ Absolute Error
    0. 1. 1. 1. 1. 1. 0.
    0.2 1.008 1.38163 1.01064 1.01411 1.01865 $ 7.327471963\times 10^{-15} $
    0.4 1.064 1.73552 1.07942 1.09826 1.1212 $ 1.720845688\times 10^{-13} $
    0.6 1.216 2.07964 1.25738 1.30578 1.36221 $ 8.968381593\times 10^{-13} $
    0.8 1.512 2.43662 1.59278 1.6843 1.78757 $ 1.98951966\times 10^{-12} $
    1. 2. 2.83206 2.13222 2.27818 2.43862 $ 1.869615573\times 10^{-12} $

     | Show Table
    DownLoad: CSV
    Table 7.  Numerical results for $ \rho(t) $ of example 5.3.
    $ t $ Exact Sol of $ \rho(t) $ App Sol of $ \rho(t) $ $ \alpha=0.9 $ $ \alpha=0.8 $ $ \alpha=0.7 $ Absolute Error
    0. -1. -1. -1. -1. -1. 0.
    0.2 -0.618331 -0.61828 -0.534717 -0.436751 -0.323107 $ 5.101539186\times 10^{-5} $
    0.4 -0.26392 -0.263882 -0.161995 -0.0516111 0.0670132 $ 3.825972939\times 10^{-5} $
    0.6 0.0835884 0.0836139 0.191075 0.304226 0.423533 $ 2.55064838\times 10^{-5} $
    0.8 0.453071 0.453084 0.568489 0.69138 0.823913 $ 1.275324046\times 10^{-5} $
    1. 0.882121 0.882121 1.01733 1.16604 1.3321 0.

     | Show Table
    DownLoad: CSV
    Table 8.  Error in $ \omega (t) $ of the first example.
    $ t $ Error in $ \omega (t) $ Error in $ \omega (t) $ Error in $ \omega (t) $ Error in $ \omega (t) $
    by RKHS for n=40 by RKHS for n=100 by Finite difference by Collocation
    0 0. Indeterminate 0. 0.
    0.1 $ 2.89\times 10^{-5} $ $ 3.93\times 10^{-6} $ $ 2.08\times 10^{-2} $ $ 1.62\times 10^{-4} $
    0.2 $ 1.69\times 10^{-5} $ $ 1.65\times 10^{-6} $ $ 3.25\times 10^{-2} $ $ 8.06\times 10^{-4} $
    0.3 $ 7.64\times 10^{-6} $ $ 8.48\times 10^{-7} $ $ 3.79\times 10^{-2} $ $ 4.04\times 10^{-4} $
    0.4 $ 5.45\times 10^{-7} $ $ 4.29\times 10^{-9} $ $ 3.89\times 10^{-2} $ $ 1.61\times 10^{-4} $
    0.5 $ 4.84\times 10^{-6} $ $ 4.19\times 10^{-7} $ $ 3.70\times 10^{-2} $ $ 1.73\times 10^{-4} $

     | Show Table
    DownLoad: CSV
    Table 9.  Error in $ \Theta (t) $ of the first example.
    $ t $ Error in $ \Theta (t) $ Error in $ \Theta (t) $ Error in $ \Theta (t) $ Error in $ \Theta (t) $
    by RKHS for n=40 by RKHS for n=100 by Finite difference by Collocation
    0 0. Indeterminate $ 2.81\times 10^{-2} $ $ 1.41\times 10^{-4} $
    0.1 $ 1.71\times 10^{-6} $ $ 4.40\times 10^{-6} $ $ 1.40\times 10^{-2} $ $ 2.77\times 10^{-5} $
    0.2 $ 5.63\times 10^{-6} $ $ 1.86\times 10^{-6} $ $ 5.01\times 10^{-3} $ $ 7.64\times 10^{-5} $
    0.3 $ 1.11\times 10^{-5} $ $ 8.21\times 10^{-7} $ $ 7.39\times 10^{-4} $ $ 9.81\times 10^{-5} $
    0.4 $ 1.52\times 10^{-5} $ $ 9.84\times 10^{-7} $ $ 6.04\times 10^{-4} $ $ 1.31\times 10^{-4} $
    0.5 $ 1.82\times 10^{-5} $ $ 1.14\times 10^{-6} $ $ 0. $ $ 0. $

     | Show Table
    DownLoad: CSV
    Table 10.  Error in $ \omega (t) $ of the second example.
    $ t $ Error in $ \omega (t) $ Error in $ \omega (t) $ Error in $ \omega (t) $ Error in $ \omega (t) $ Error in $ \omega (t) $
    by RKHS for n=35 by RKHS for n=60 by RKHS for n=100 by Finite difference by Collocation
    0 0. 0. 0. 0. Failed
    0.2 $ 1.71\times 10^{-5} $ $ 3.44\times 10^{-6} $ $ 2.51\times 10^{-7} $ $ 1.09\times 10^{-2} $ Failed
    0.4 $ 2.33\times 10^{-5} $ $ 4.69\times 10^{-6} $ $ 3.49\times 10^{-7} $ $ 2.62\times 10^{-2} $ Failed
    0.6 $ 3.05\times 10^{-5} $ $ 6.14\times 10^{-6} $ $ 4.73\times 10^{-7} $ $ 4.63\times 10^{-2} $ Failed
    0.8 $ 3.84\times 10^{-5} $ $ 7.73\times 10^{-6} $ $ 5.98\times 10^{-7} $ $ 7.11\times 10^{-2} $ Failed
    1 $ 4.63\times 10^{-5} $ $ 9.73\times 10^{-6} $ $ 7.45\times 10^{-7} $ $ 9.91\times 10^{-2} $ Failed

     | Show Table
    DownLoad: CSV
    Table 11.  Error in $ \Theta (t) $ of the second example.
    $ t $ Error in $ \Theta (t) $ Error in $ \Theta (t) $ Error in $ \Theta (t) $ Error in $ \Theta (t) $ Error in $ \Theta (t) $
    by RKHS for n=35 by RKHS for n=60 by RKHS for n=100 by Finite difference by Collocation
    0 0. 0. 0. $ 5.54\times 10^{-2} $ Failed
    0.2 $ 1.37\times 10^{-5} $ $ 2.82\times 10^{-6} $ $ 2.51\times 10^{-7} $ $ 4.49\times 10^{-2} $ Failed
    0.4 $ 1.46\times 10^{-5} $ $ 3.00\times 10^{-6} $ $ 3.49\times 10^{-7} $ $ 3.34\times 10^{-2} $ Failed
    0.6 $ 1.34\times 10^{-5} $ $ 2.79\times 10^{-6} $ $ 4.73\times 10^{-7} $ $ 2.04\times 10^{-2} $ Failed
    0.8 $ 9.08\times 10^{-6} $ $ 2.03\times 10^{-6} $ $ 5.98\times 10^{-7} $ $ 7.77\times 10^{-3} $ Failed
    1 0. 0. 0. 0. Failed

     | Show Table
    DownLoad: CSV
    Table 12.  Error in $ \omega (t) $ of the third example.
    $ t $ Error in $ \omega (t) $ Error in $ \omega (t) $ Error in $ \omega (t) $ Error in $ \omega (t) $
    by RKHS for n=30 by RKHS for n=60 by Finite difference by Collocation
    0 0. 0. 0. 0.
    0.2 $ 2.77\times 10^{-5} $ $ 1.43\times 10^{-6} $ $ 2.83\times 10^{-3} $ $ 5.86\times 10^{-4} $
    0.4 $ 4.43\times 10^{-5} $ $ 1.42\times 10^{-6} $ $ 1.24\times 10^{-2} $ $ 5.29\times 10^{-4} $
    0.6 $ 5.61\times 10^{-5} $ $ 1.12\times 10^{-6} $ $ 3.13\times 10^{-2} $ $ 5.44\times 10^{-4} $
    0.8 $ 6.30\times 10^{-5} $ $ 6.62\times 10^{-6} $ $ 6.06\times 10^{-2} $ $ 4.86\times 10^{-4} $
    1 $ 6.50\times 10^{-5} $ $ 9.69\times 10^{-6} $ $ 1.00\times 10^{-2} $ $ 1.08\times 10^{-3} $

     | Show Table
    DownLoad: CSV
    Table 13.  Error in $ \Theta (t) $ of the third example.
    $ t $ Error in $ \Theta (t) $ Error in $ \Theta (t) $ Error in $ \Theta (t) $ Error in $ \Theta (t) $
    by RKHS for n=30 by RKHS for n=60 by Finite difference by Collocation
    0 0. 0. 0. 0.
    0.2 $ 7.32\times 10^{-15} $ $ 6.28\times 10^{-13} $ $ 5.00\times 10^{-3} $ 0.
    0.4 $ 1.72\times 10^{-13} $ $ 2.80\times 10^{-13} $ $ 2.20\times 10^{-2} $ 0.
    0.6 $ 8.96\times 10^{-13} $ $ 4.10\times 10^{-13} $ $ 5.10\times 10^{-2} $ 0.
    0.8 $ 1.98\times 10^{-12} $ $ 5.94\times 10^{-13} $ $ 9.20\times 10^{-2} $ 0.
    1 $ 1.86\times 10^{-12} $ $ 5.37\times 10^{-14} $ $ 1.45\times 10^{-1} $ 0.

     | Show Table
    DownLoad: CSV
    Table 14.  Error in $ \rho(t) $ of the third example.
    $ t $ Error in $ \rho(t) $ Error in $ \rho(t) $ Error in $ \rho(t) $ Error in $ \rho(t) $
    by RKHS for n=30 by RKHS for n=60 by Finite difference by Collocation
    0 0. 0. $ 5.59\times 10^{-2} $ $ 1.29\times 10^{-4} $
    0.2 $ 5.10\times 10^{-5} $ $ 1.48\times 10^{-6} $ $ 6.47\times 10^{-2} $ $ 5.49\times 10^{-5} $
    0.4 $ 3.82\times 10^{-5} $ $ 3.07\times 10^{-6} $ $ 6.80\times 10^{-2} $ $ 6.03\times 10^{-5} $
    0.6 $ 2.55\times 10^{-5} $ $ 4.66\times 10^{-6} $ $ 6.14\times 10^{-2} $ $ 5.73\times 10^{-5} $
    0.8 $ 1.27\times 10^{-5} $ $ 6.26\times 10^{-6} $ $ 4.03\times 10^{-2} $ $ 6.23\times 10^{-5} $
    1 0. 0. 0. 0.

     | Show Table
    DownLoad: CSV

    Graphs of the approximate solutions of $ \omega(t) $ and $ \theta(t) $ are plotted in Figure 3 (a), Figure 3 (b) for different values of $ \alpha $. The graph in Figure 3 (c) represent the absolute errors of $ \rho(t) $.

    Figure 3.  Solution and graphical curves of Example 5.3.

    Now, we consider the following tables where the RKHS method has been applied in order to give numerical approximations with other values of n, and then compare it with finite difference and collocation methods.

    In this article, we effectively utilize the RKHSM to develop an approximate solution of differential fractional equations with temporal two-point BVP. The results of examples demonstrate reliability and consistency of the method. In the future, we recommend further research on the RKHS method, as solving the temporal two-point boundary value problems with the conformable and the Atangana-Baleanu derivatives. We expect to achieve better results and good approximations for the solutions.

    The authors state that they have no conflict of interest. All authors have worked in an equal sense to find these results.

    [1] Dikshit R, Gupta PC, Ramasundarahettige C, et al. (2012) Cancer mortality in India: a nationally representative survey. Lancet 379: 1807-1816. doi: 10.1016/S0140-6736(12)60358-4
    [2] Vargo-Gogola T, Rosen JM (2007) Modelling breast cancer: one size does not fit all. Nat Rev Cancer 7: 659-672. doi: 10.1038/nrc2193
    [3] Nickels S, Truong T, Hein R, et al. (2013) Evidence of gene-environment interactions between common breast cancer susceptibility loci and established environmental risk factors. PLoS Genet 9: e1003284. doi: 10.1371/journal.pgen.1003284
    [4] Chen C, Huang Y, Li Y, et al. (2007) Cytochrome P450 1A1 (CYP1A1) T3801C and A2455G polymorphisms in breast cancer risk: a meta-analysis. J Hum Genet 52: 423-435. doi: 10.1007/s10038-007-0131-8
    [5] Sergentanis TN, Economopoulos KP (2010) Four polymorphisms in cytochrome P450 1A1 (CYP1A1) gene and breast cancer risk: a meta-analysis. Breast Cancer Res Treat 122: 459-469. doi: 10.1007/s10549-009-0694-5
    [6] Yao L, Yu X, Yu L (2010) Lack of significant association between CYP1A1 T3801C polymorphism and breast cancer risk: a meta-analysis involving 25,087 subjects. Breast Cancer Res Treat 122: 503-507. doi: 10.1007/s10549-009-0717-2
    [7] Syamala VS, Syamala V, Sheeja VR, et al. (2010) Possible risk modification by polymorphisms of estrogen metabolizing genes in familial breast cancer susceptibility in an Indian population. Cancer Invest 28: 304-311. doi: 10.3109/07357900902744494
    [8] Surekha D, Sailaja K, Rao DN, et al. (2009) Association of CYP1A1*2 polymorphisms with breast cancer risk: a case control study. Indian J Med Sci 63: 13-20. doi: 10.4103/0019-5359.49077
    [9] Singh V, Rastogi N, Sinha A, et al. (2007) A study on the association of cytochrome-P450 1A1 polymorphism and breast cancer risk in north Indian women. Breast Cancer Res Treat 101: 73-81. doi: 10.1007/s10549-006-9264-2
    [10] Singh N, Mitra AK, Garg VK, et al. (2007) Association of CYP1A1 polymorphisms with breast cancer in North Indian women. Oncol Res 16: 587-597. doi: 10.3727/000000007783629972
    [11] Naushad SM, Reddy CA, Rupasree Y, et al. (2011) Cross-talk between one-carbon metabolism and xenobiotic metabolism: implications on oxidative DNA damage and susceptibility to breast cancer. Cell Biochem Biophys 61: 715-723. doi: 10.1007/s12013-011-9245-x
    [12] Chacko P, Joseph T, Mathew BS, et al. (2005) Role of xenobiotic metabolizing gene polymorphisms in breast cancer susceptibility and treatment outcome. Mutat Res 581: 153-163. doi: 10.1016/j.mrgentox.2004.11.018
    [13] Kiruthiga PV, Kannan MR, Saraswathi C, et al. (2011) CYP1A1 gene polymorphisms: lack of association with breast cancer susceptibility in the southern region (Madurai) of India. Asian Pac J Cancer Prev 12: 2133-2138.
    [14] Masson LF, Sharp L, Cotton SC, et al. (2005) Cytochrome P-450 1A1 gene polymorphisms and risk of breast cancer: a HuGE review. Am J Epidemiol 161: 901-915. doi: 10.1093/aje/kwi121
    [15] Kawajiri K, Nakachi K, Imai K, et al. (1990) Identification of genetically high risk individuals to lung cancer by DNA polymorphisms of the cytochrome P450IA1 gene. FEBS Lett 263: 131-133. doi: 10.1016/0014-5793(90)80721-T
    [16] Hayashi SI, Watanabe J, Nakachi K, et al. (1991) PCR detection of an A/G polymorphism within exon 7 of the CYP1A1 gene. Nucleic Acids Res 19: 4797.
    [17] Cascorbi I, Brockmoller J, Roots I (1996) A C4887A polymorphism in exon 7 of human CYP1A1: population frequency, mutation linkages, and impact on lung cancer susceptibility. Cancer Res 56: 4965-4969.
    [18] Li Y, Millikan RC, Bell DA, et al. (2005) Polychlorinated biphenyls, cytochrome P450 1A1 (CYP1A1) polymorphisms, and breast cancer risk among African American women and white women in North Carolina: a population-based case-control study. Breast Cancer Res 7: R12-18.
    [19] Crofts F, Taioli E, Trachman J, et al. (1994) Functional significance of different human CYP1A1 genotypes. Carcinogenesis 15: 2961-2963. doi: 10.1093/carcin/15.12.2961
    [20] Kiyohara C, Hirohata T, Inutsuka S (1996) The relationship between aryl hydrocarbon hydroxylase and polymorphisms of the CYP1A1 gene. Jpn J Cancer Res 87: 18-24. doi: 10.1111/j.1349-7006.1996.tb00194.x
    [21] Li Y, Millikan RC, Bell DA, et al. (2004) Cigarette smoking, cytochrome P4501A1 polymorphisms, and breast cancer among African-American and white women. Breast Cancer Res 6: R460-473. doi: 10.1186/bcr814
    [22] Moher D, Liberati A, Tetzlaff J, et al. (2009) Preferred reporting items for systematic reviews and meta-analyses: the PRISMA statement. Ann Intern Med 151: 264-269, W264. doi: 10.7326/0003-4819-151-4-200908180-00135
    [23] Wells S PJ, Welch V. The newcastle-ottawa scale (NOS) for assessing the quality of nonrandomised studies in meta-analyses. Ottawa Health Research Institute, 2011. Available from: www.ohri.ca/programs/clinical_epidemiology/oxford/asp.
    [24] Mantel N, Haenszel W (1959) Statistical aspects of the analysis of data from retrospective studies of disease. J Natl Cancer Inst 22: 719-748.
    [25] DerSimonian R, Laird N (1986) Meta-analysis in clinical trials. Control Clin Trials 7: 177-188. doi: 10.1016/0197-2456(86)90046-2
    [26] Lau J, Ioannidis JP, Schmid CH (1997) Quantitative synthesis in systematic reviews. Ann Intern Med 127: 820-826. doi: 10.7326/0003-4819-127-9-199711010-00008
    [27] Higgins JP, Thompson SG (2002) Quantifying heterogeneity in a meta-analysis. Stat Med 21: 1539-1558. doi: 10.1002/sim.1186
    [28] Higgins JP, Thompson SG, Deeks JJ, et al. (2003) Measuring inconsistency in meta-analyses. Bmj 327: 557-560. doi: 10.1136/bmj.327.7414.557
    [29] Gautham M, Shyamprasad KM, Singh R, et al. (2014) Informal rural healthcare providers in North and South India. Health Policy Plan 29 Suppl 1: i20-29.
    [30] Ravindran RD, Vashist P, Gupta SK, et al. (2011) Prevalence and risk factors for vitamin C deficiency in north and south India: a two centre population based study in people aged 60 years and over. PLoS One 6: e28588. doi: 10.1371/journal.pone.0028588
    [31] Ioannidis JP, Trikalinos TA (2007) The appropriateness of asymmetry tests for publication bias in meta-analyses: a large survey. Cmaj 176: 1091-1096. doi: 10.1503/cmaj.060410
    [32] Gadgil MaG, R. (1992) The fissure land: An ecological history of India; Press OU, editor. New Delhi: Oxford University Press.
    [33] He XF, Wei W, Liu ZZ, et al. (2014) Association between the CYP1A1 T3801C polymorphism and risk of cancer: evidence from 268 case-control studies. Gene 534: 324-344. doi: 10.1016/j.gene.2013.10.025
    [34] Wacholder S, Chanock S, Garcia-Closas M, et al. (2004) Assessing the probability that a positive report is false: an approach for molecular epidemiology studies. J Natl Cancer Inst 96: 434-442. doi: 10.1093/jnci/djh075
    [35] Thakkinstian A, McElduff P, D'Este C, et al. (2005) A method for meta-analysis of molecular association studies. Stat Med 24: 1291-1306. doi: 10.1002/sim.2010
  • This article has been cited by:

    1. Shuoshi Liu, Yu Qian, Dan Li, Jiří Jaromír Klemeš, Siyu Yang, Multi-Scenario scheduling optimisation for a novel Double-Stage ammonia absorption refrigeration system incorporating an organic Rankine cycle, 2022, 270, 01968904, 116170, 10.1016/j.enconman.2022.116170
    2. Ali Akgül, Nourhane Attia, Advancing differential equation solutions: A novel Laplace-reproducing kernel Hilbert space method for nonlinear fractional Riccati equations, 2024, 0129-1831, 10.1142/S0129183124501869
    3. Nourhane Attia, Ali Akgül, 2023, Chapter 5, 978-981-99-5000-3, 117, 10.1007/978-981-99-5001-0_5
    4. Nourhane Attia, Ali Akgül, Abdon Atangana, Yusuf Gurefe, Reproducing Kernel Method with Global Derivative, 2023, 2023, 2314-8888, 1, 10.1155/2023/2726735
  • Reader Comments
  • © 2015 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(5542) PDF downloads(1289) Cited by(2)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog