
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
[1] | Guohui Zhang, Jinghe Sun, Xing Liu, Guodong Wang, Yangyang Yang . Solving flexible job shop scheduling problems with transportation time based on improved genetic algorithm. Mathematical Biosciences and Engineering, 2019, 16(3): 1334-1347. doi: 10.3934/mbe.2019065 |
[2] | Ruiping Yuan, Jiangtao Dou, Juntao Li, Wei Wang, Yingfan Jiang . Multi-robot task allocation in e-commerce RMFS based on deep reinforcement learning. Mathematical Biosciences and Engineering, 2023, 20(2): 1903-1918. doi: 10.3934/mbe.2023087 |
[3] | Shixuan Yao, Xiaochen Liu, Yinghui Zhang, Ze Cui . An approach to solving optimal control problems of nonlinear systems by introducing detail-reward mechanism in deep reinforcement learning. Mathematical Biosciences and Engineering, 2022, 19(9): 9258-9290. doi: 10.3934/mbe.2022430 |
[4] | Kongfu Hu, Lei Wang, Jingcao Cai, Long Cheng . An improved genetic algorithm with dynamic neighborhood search for job shop scheduling problem. Mathematical Biosciences and Engineering, 2023, 20(9): 17407-17427. doi: 10.3934/mbe.2023774 |
[5] | Jianguo Duan, Mengting Wang, Qinglei Zhang, Jiyun Qin . Distributed shop scheduling: A comprehensive review on classifications, models and algorithms. Mathematical Biosciences and Engineering, 2023, 20(8): 15265-15308. doi: 10.3934/mbe.2023683 |
[6] | Zilong Zhuang, Zhiyao Lu, Zizhao Huang, Chengliang Liu, Wei Qin . A novel complex network based dynamic rule selection approach for open shop scheduling problem with release dates. Mathematical Biosciences and Engineering, 2019, 16(5): 4491-4505. doi: 10.3934/mbe.2019224 |
[7] | Shaofeng Yan, Guohui Zhang, Jinghe Sun, Wenqiang Zhang . An improved ant colony optimization for solving the flexible job shop scheduling problem with multiple time constraints. Mathematical Biosciences and Engineering, 2023, 20(4): 7519-7547. doi: 10.3934/mbe.2023325 |
[8] | Zichen Wang, Xin Wang . Fault-tolerant control for nonlinear systems with a dead zone: Reinforcement learning approach. Mathematical Biosciences and Engineering, 2023, 20(4): 6334-6357. doi: 10.3934/mbe.2023274 |
[9] | Jin Zhang, Nan Ma, Zhixuan Wu, Cheng Wang, Yongqiang Yao . Intelligent control of self-driving vehicles based on adaptive sampling supervised actor-critic and human driving experience. Mathematical Biosciences and Engineering, 2024, 21(5): 6077-6096. doi: 10.3934/mbe.2024267 |
[10] | Lu-Wen Liao . A branch and bound algorithm for optimal television commercial scheduling. Mathematical Biosciences and Engineering, 2022, 19(5): 4933-4945. doi: 10.3934/mbe.2022231 |
In this paper, we consider the following diffusion equation on
$ {−∇⋅(α∇u)=f,inΩ,u=0,on∂Ω. $ | (1) |
To approximate (1), taking advantage of the adaptive mesh refinement (AMR) to save valuable computational resources, the adaptive finite element method on quadtree mesh is among the most popular ones in the engineering and scientific computing community [20]. Compared with simplicial meshes, quadtree meshes provide preferable performance in the aspects of the accuracy and robustness. There are lots of mature software packages (e.g., [1,2]) on quadtree meshes. To guide the AMR, one possible way is through the a posteriori error estimation to construct computable quantities to indicate the location that the mesh needs to be refined/coarsened, thus to balance the spacial distribution of the error which improves the accuracy per computing power. Residual-based and recovery-based error estimators are among the most popular ones used. In terms of accuracy, the recovery-based error estimator shows more appealing attributes [28,3].
More recently, newer developments on flux recovery have been studied by many researchers on constructing a post-processed flux in a structure-preserving approximation space. Using (1) as an example, given that the data
However, these
More recently, a new class of methods called the virtual element methods (VEM) were introduced in [4,8], which can be viewed as a polytopal generalization of the tensorial/simplicial finite element. Since then, lots of applications of VEM have been studied by many researchers. A usual VEM workflow splits the consistency (approximation) and the stability of the method as well as the finite dimensional approximation space into two parts. It allows flexible constructions of spaces to preserve the structure of the continuous problems such as higher order continuities, exact divergence-free spaces, and many others. The VEM functions are represented by merely the degrees of freedom (DoF) functionals, not the pointwise values. In computation, if an optimal order discontinuous approximation can be computed elementwisely, then adding an appropriate parameter-free stabilization suffices to guarantee the convergence under common assumptions on the geometry of the mesh.
The adoption of the polytopal element brings many distinctive advantages, for example, treating rectangular element with hanging nodes as polygons allows a simple construction of
The major ingredient in our study is an
If
$ (α∇uT,∇vT)=(f,vT),∀vT∈Qk(T)∩H10(Ω), $ | (2) |
in which the standard notation is opted.
$ \mathcal{Q}_k(\mathcal{T}) : = \{v\in H^1(\Omega): v|_K\in \mathbb{Q}_k(K), \;\forall K\in \mathcal{T} \}. $ |
and on
$ \mathbb{Q}_k(K) : = \mathbb{P}_{k,k}(K) = \big\{p(x)q(y), \; p\in \mathbb{P}_k([a,b]), q\in \mathbb{P}_k([c,d]) \big\}, $ |
where
On
$ NH:={z∈N:∃K∈T,z∈∂K∖NK} $ | (3) |
Otherwise the node
For each edge
$ \{v\}^{\gamma}_e : = \gamma v^- + (1-\gamma) v^+. $ |
In this subsection, the quadtree mesh
For the embedded element
Subsequently,
On
$ Vk(K):={τ∈H(div;K)∩H(rot;K):∇⋅τ∈Pk−1(K),∇×τ=0,τ⋅ne∈Pk(e),∀e⊂∂K}. $ | (4) |
An
$ Vk:={τ∈H(div):τ|K∈Vk(K),onK∈Tpoly}. $ | (5) |
Next we turn to define the degrees of freedom (DoFs) of this space. To this end, we define the set of scaled monomials
$ Pk(e):=span{1,s−mehe,(s−mehe)2,…,(s−mehe)k}, $ | (6) |
where
$ Pk(K):=span{mα(x):=(x−xKhK)α,|α|≤k}. $ | (7) |
The degrees of freedom (DoFs) are then set as follows for a
$ (e)k≥1∫e(τ⋅ne)mds,∀m∈Pk(e),one⊂Epoly.(i)k≥2∫Kτ⋅∇mdx,∀m∈Pk−1(K)/RonK∈Tpoly. $ | (8) |
Remark 1. We note that in our construction, the degrees of freedom to determine the curl of a VEM function originally in [8] are replaced by a curl-free constraint thanks to the flexibility to virtual element. The reason why we opt for this subspace is that the true flux
As the data
Consider
On each
$ {−α∇uT}γee⋅ne:=(γe(−αK−∇uT|K−)+(1−γe)(−αK+∇uT|K+))⋅ne, $ | (9) |
where
$ γe:=α1/2K+α1/2K++α1/2K−. $ | (10) |
First for both
$ σT⋅ne={−α∇uT}γee⋅ne. $ | (11) |
In the lowest order case
$ |K|∇⋅σT=∫K∇⋅σTdx=∫∂KσT⋅n∂Kds=∑e⊂∂K∫eσT⋅n∂K|eds. $ | (12) |
If
$ ∇⋅σT=Πk−1f+cK. $ | (13) |
The reason to add
$ cK=1|K|(−∫KΠk−1fdx+∑e⊂∂K∫e{−α∇uT}γee⋅n∂K|eds), $ | (14) |
Consequently for
$ (σT,∇q)K=−(Πk−1f+cK,q)K+∑e⊂∂K({−α∇uT}γee⋅n∂K|e,q)e. $ | (15) |
To the end of constructing a computable local error indicator, inspired by the VEM formulation [8], the recovered flux is projected to a space with a much simpler structure. A local oblique projection
$ (Πτ,∇p)K=(τ,∇p)K,∀p∈Pk(K)/R. $ | (16) |
Next we are gonna show that this projection operator can be straightforward computed for vector fields in
When
$ (τ,∇p)K=−(∇⋅τ,p)K+(τ⋅n,p)∂K. $ | (17) |
By definition of the space (4) when
When
$ (τ,∇p)K=−(∇⋅τ,Πk−1p)K+(τ⋅n,p)∂K=(τ,∇Πk−1p)K+(τ⋅n,p−Πk−1p)∂K, $ | (18) |
which can be evaluated using both DoF sets
Given the recovered flux σT in Section 3, the recovery-based local error indicator
$ ηflux,K:=‖α−1/2(σT+α∇uT)‖K,andηres,K:=‖α−1/2(f−∇⋅σT)‖K, $ | (19) |
then
$ ηK={ηflux,Kwhenk=1,(η2flux,K+η2res,K)1/2whenk≥2. $ | (20) |
A computable
$ ˆηflux,K:=‖α−1/2KΠ(σT+αK∇uT)‖K, $ | (21) |
with the oblique projection
$ ˆηstab,K:=|α−1/2K(I−Π)(σT+αK∇uT)|S,K. $ | (22) |
Here
$ SK(v,w):=∑e⊂∂Khe(v⋅ne,w⋅ne)e+∑α∈Λ(v,∇mα)K(w,∇mα)K, $ | (23) |
where
The computable error estimator
$ ˆη2={∑K∈T(ˆη2flux,K+ˆη2stab,K)=:∑K∈Tˆη2Kwhenk=1,∑K∈T(ˆη2flux,K+ˆη2stab,K+η2res,K)=:∑K∈Tˆη2Kwhenk≥2. $ | (24) |
In this section, we shall prove the proposed recovery-based estimator
Theorem 4.1. Let
$ ˆη2flux,K≲ $ | (25) |
where
$ \begin{align*} \mathrm{osc}(f;K) & = \alpha_K^{-1/2}h_K \big\Vert f- \Pi_{k-1} f \big\Vert_K, \\ \eta_{\mathrm{elem},K} &: = \alpha_K^{-1/2}h_K \big\Vert f + \nabla\cdot(\alpha \nabla u_{\mathcal{T}}) \big\Vert_K, \\ \mathit{\;{\rm{and}}\;}\; \eta_{\mathrm{edge},K} &: = \left(\sum\limits_{e\subset \partial K} \frac{h_e}{\alpha_K + \alpha_{K_e}} \big\Vert \lbrack\lbrack {{\alpha \nabla u_{\mathcal{T}}\cdot\mathit{\boldsymbol{n}}_e}} \rbrack\rbrack_{{ {} }} \big\Vert_e^2\right)^{1/2}. \end{align*} $ |
In the edge jump term,
Proof. Let
$ \begin{equation} \begin{aligned} \widehat{\eta}_{\mathrm{flux},K}^2 & = \bigl({\Pi}(\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K \nabla u_{\mathcal{T}}), \nabla p \bigr)_K = \bigl(\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K \nabla u_{\mathcal{T}}, \nabla p \bigr)_K \\ & = -\bigl(\nabla \cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} +\alpha_K \nabla u_{\mathcal{T}}), p \bigr)_K + \sum\limits_{e\subset \partial K}\int_e \big( \mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K \nabla u_{\mathcal{T}}\big)\cdot \mathit{\boldsymbol{n}}_{\partial K}\big|_{e} \, p \, \mathrm{d} s. \end{aligned} \end{equation} $ | (26) |
By (11), without loss of generality we assume
$ \begin{equation} \begin{aligned} \big( \mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K \nabla u_{\mathcal{T}}\big)\cdot \mathit{\boldsymbol{n}}_e & = \Big( (1-\gamma_e) \alpha_{K} \nabla u_{\mathcal{T}}|_K - (1-\gamma_e)\alpha_{K_e} \nabla u_{\mathcal{T}}|_{K_e} \Big)\cdot \mathit{\boldsymbol{n}}_e \\ & = \frac{\alpha_{K}^{1/2}}{\alpha_{K}^{1/2} + \alpha_{K_e}^{1/2}}\lbrack\lbrack {{\alpha \nabla u_{\mathcal{T}}\cdot \mathit{\boldsymbol{n}}_e}} \rbrack\rbrack_{{ {e} }}. \end{aligned} \end{equation} $ | (27) |
The boundary term in (26) can be then rewritten as
$ \begin{equation} \begin{aligned} & \int_e \big( \mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K \nabla u_{\mathcal{T}}\big)\cdot \mathit{\boldsymbol{n}}_e \,p\, \mathrm{d} s \\ = &\;\int_e \frac{1}{\alpha_{K}^{1/2} + \alpha_{K_e}^{1/2}}\lbrack\lbrack {{\alpha \nabla u_{\mathcal{T}}\cdot \mathit{\boldsymbol{n}}_e}} \rbrack\rbrack_{{ {e} }} \,\alpha_{K}^{1/2}p\, \mathrm{d} s \\ \lesssim & \; \frac{1}{(\alpha_{K}+ \alpha_{K_e})^{1/2}} h_e^{1/2} \big\Vert \lbrack\lbrack {{\alpha \nabla u_{\mathcal{T}}\cdot\mathit{\boldsymbol{n}}_e}} \rbrack\rbrack_{{ {} }} \big\Vert_e \alpha_{K}^{1/2} h_e^{-1/2} \left\Vert{p}\right\Vert_e. \end{aligned} \end{equation} $ | (28) |
By a trace inequality on an edge of a polygon (Lemma 7.1), and the Poincaré inequality for
$ h_e^{-1/2}\|p\|_e \lesssim h_K^{-1} \|p\|_K + \|\nabla p\|_K \lesssim \|\nabla p\|_K. $ |
As a result,
$ \sum\limits_{e\subset \partial K}\int_e \big( \mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K \nabla u_{\mathcal{T}}\big)\cdot \mathit{\boldsymbol{n}}_e \,p\, \mathrm{d} s \lesssim \eta_{\mathrm{edge},K}\, \alpha_{K}^{1/2} \left\Vert{\nabla p}\right\Vert_e = \eta_{\mathrm{edge},K} \,\widehat{\eta}_{\mathrm{flux},K}. $ |
For the bulk term on
$ \begin{aligned} &-\bigl(\nabla \cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} +\alpha_K \nabla u_{\mathcal{T}}), p \bigr)_K \leq \left|\nabla \cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} +\alpha_K \nabla u_{\mathcal{T}}) \right| |K|^{1/2} \left\Vert{p}\right\Vert_K \\ \leq & \; \frac{1}{ |K|^{1/2}}\left|\int_K \nabla \cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} +\alpha_K \nabla u_{\mathcal{T}}) \, \mathrm{d} \mathit{\boldsymbol{x}}\right| \left\Vert{p}\right\Vert_K \\ = & \; \frac{1}{ |K|^{1/2}} \left|\sum\limits_{e\subset \partial K} \int_{e} (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} +\alpha_K \nabla u_{\mathcal{T}})\cdot \mathit{\boldsymbol{n}}_e \, \mathrm{d} s\right| \left\Vert{p}\right\Vert_K \\ \leq & \; \left(\sum\limits_{e\subset \partial K} \frac{1}{\alpha_{K}^{1/2} + \alpha_{K_e}^{1/2}} \left\Vert{\lbrack\lbrack {{\alpha \nabla u_{\mathcal{T}}\cdot \mathit{\boldsymbol{n}}_e}} \rbrack\rbrack_{{ {} }}}\right\Vert_e \,\alpha_{K}^{1/2}h_e \right) \left\Vert{\nabla p}\right\Vert \\ \lesssim &\; \eta_{\mathrm{edge},K} \, \widehat{\eta}_{\mathrm{flux},K}. \end{aligned} $ |
When
$ \begin{equation} \begin{aligned} & -\bigl(\nabla \cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} +\alpha_K \nabla u_{\mathcal{T}}), p \bigr)_K = -\bigl(\Pi_{k-1} f + c_K + \nabla\cdot(\alpha_K \nabla u_{\mathcal{T}}), p \bigr)_K \\ \leq & \; \left( \big\Vert f- \Pi_{k-1} f \big\Vert_K + \big\Vert f + \nabla\cdot(\alpha \nabla u_{\mathcal{T}}) \big\Vert_K + |c_K| |K|^{1/2}\right)\left\Vert{p}\right\Vert_K. \end{aligned} \end{equation} $ | (29) |
The first two terms can be handled by combining the weights
$ \begin{equation} \begin{aligned} c_K |K|^{1/2} & = \frac{1}{ |K|^{1/2} }\Big(-\int_K (\Pi_{k-1} f -f) \mathrm{d} \mathit{\boldsymbol{x}} - \int_K \big(f + \nabla\cdot (\alpha \nabla u_{\mathcal{T}})\big) \mathrm{d} \mathit{\boldsymbol{x}} \\ &\quad + \int_K \nabla\cdot (\alpha \nabla u_{\mathcal{T}}) \mathrm{d} \mathit{\boldsymbol{x}} + \sum\limits_{e\subset\partial K} \int_e \left\{-\alpha \nabla u_{\mathcal{T}} \right\}^{\gamma_e}_e \cdot\mathit{\boldsymbol{n}}_e \mathrm{d} s\Big) \\ &\leq \big\Vert f- \Pi_{k-1} f \big\Vert_K + \big\Vert f + \nabla\cdot(\alpha \nabla u_{\mathcal{T}}) \big\Vert_K \\ & \quad + \frac{1}{ |K|^{1/2}}\sum\limits_{e\subset\partial K} \int_e (\alpha_K \nabla u_{\mathcal{T}} - \left\{\alpha \nabla u_{\mathcal{T}} \right\}^{\gamma_e}_e) \cdot\mathit{\boldsymbol{n}}_e \mathrm{d} s \\ & \leq \big\Vert f- \Pi_{k-1} f \big\Vert_K + \big\Vert f + \nabla\cdot(\alpha \nabla u_{\mathcal{T}}) \big\Vert_K \\ & \quad + \sum\limits_{e\subset\partial K} \frac{\alpha_{K}^{1/2}}{\alpha_{K}^{1/2} + \alpha_{K_e}^{1/2}} \left\Vert{\lbrack\lbrack {{\alpha \nabla u_{\mathcal{T}}\cdot \mathit{\boldsymbol{n}}_e}} \rbrack\rbrack_{{ {} }}}\right\Vert_e . \end{aligned} \end{equation} $ | (30) |
The two terms on
$ -\bigl(\nabla \cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} +\alpha_K \nabla u_{\mathcal{T}}), p \bigr)_K \lesssim \Big( \mathrm{osc}(f; K) + \eta_{\mathrm{elem},K} + \eta_{\mathrm{edge},K} \Big) \alpha_K^{1/2}\left\Vert{\nabla p}\right\Vert $ |
and the theorem follows.
Theorem 4.2. Under the same setting with Theorem 4.1, let
$ \begin{equation} \widehat{\eta}_{\mathrm{stab},K}^2 \lesssim \mathrm{osc}(f; K)^2 + \eta_{\mathrm{elem},K}^2 + \eta_{\mathrm{edge},K}^2 , \end{equation} $ | (31) |
The constant depends on
Proof. This theorem follows directly from the norm equivalence Lemma 7.3:
$ \big\vert{\alpha_K^{-1/2}(\operatorname{I}-{\Pi})(\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K \nabla u_{\mathcal{T}})} \big\vert_{S,K} \lesssim \big\vert{\alpha_K^{-1/2}(\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K \nabla u_{\mathcal{T}})} \big\vert_{S,K}, $ |
while evaluating the DoFs
Theorem 4.3. Under the same setting with Theorem 4.1, on any
$ \begin{equation} \widehat{\eta}_{K} \lesssim \mathrm{osc}(f;K) + \big\Vert \alpha^{1/2}\nabla (u-u_{\mathcal{T}})\big\Vert_{\omega_K}, \end{equation} $ | (32) |
with a constant independent of
Proof. This is a direct consequence of Theorem 4.1 and 4.2 and the fact that the residual-based error indicator is efficient by a common bubble function argument.
In this section, we shall prove that the computable error estimator
Assumption 1 (
By Assumption 1, we denote the father
Assumption 2 (Quasi-monotonicity of
Denote
$ \begin{equation} \pi_{z} v = \left\{\begin{array}{ll} \frac{\int_{\omega_{z} \cap \omega_{m(z)}} v \phi_z }{\int_{\omega_{z} \cap \omega_{m(z)}} \phi_z }& {\rm { if }}\; \mathit{\boldsymbol{z}} \in \Omega, \\ 0 & {\rm { if }}\; \mathit{\boldsymbol{z}} \in \partial \Omega. \end{array}\right. \end{equation} $ | (33) |
We note that if
$ \begin{equation} \mathcal{I} v : = \sum\limits_{z\in \mathcal{N}_1} (\pi_z v)\phi_z. \end{equation} $ | (34) |
Lemma 4.4 (Estimates for
$ \begin{equation} \alpha_K^{1/2} h_K^{-1} \left\Vert{v - \mathcal{I}v}\right\Vert_{K} + \alpha_K^{1/2} \left\Vert{\nabla \mathcal{I}v}\right\Vert_{K} \lesssim \big\Vert \alpha^{1/2} \nabla v\big\Vert_{\omega_K}, \end{equation} $ | (35) |
and for
$ \begin{equation} \sum\limits_{K\subset\omega_z} h_{z}^{-2} \|\alpha^{1/2}(v-\pi_{z} v)\phi_z\|_{K}^2 \lesssim \big\Vert \alpha^{1/2} \nabla v\big\Vert_{\omega_z}^2, \end{equation} $ | (36) |
in which
Proof. The estimate for
Denotes the subset of nodes
$ \begin{equation} \begin{aligned} \mathrm{osc}(f;\mathcal{T})^2 : = & \sum\limits_{z \in \mathcal{N}_1\cap( \mathcal{N}_{\partial \Omega} \cup \mathcal{N}_I)} h_{z}^{2} \big\|\alpha^{-1/2} f\big\|_{\omega_{z}}^2 \\ +& \sum\limits_{z \in \mathcal{N}_1 \backslash ( \mathcal{N}_{\partial \Omega} \cup \mathcal{N}_I)} h_{z}^{2} \big\|\alpha^{-1/2} (f - f_z)\big\|_{\omega_{z}}^2, \end{aligned} \end{equation} $ | (37) |
with
Theorem 4.5. Let
$ \begin{equation} \big\Vert{\alpha^{1/2}\nabla (u - u_{\mathcal{T}})}\big\Vert \lesssim \left(\widehat{\eta}^2 +\mathrm{osc}(f;\mathcal{T})^2 \right)^{1/2}. \end{equation} $ | (38) |
For
$ \begin{equation} \big\Vert{\alpha^{1/2}\nabla (u - u_{\mathcal{T}})}\big\Vert \lesssim \widehat{\eta}, \end{equation} $ | (39) |
where the constant depends on
Proof. Let
$ \begin{align*} &\big\Vert{\alpha^{1/2}\nabla \varepsilon}\big\Vert^2 = \big(\alpha\nabla (u - u_{\mathcal{T}}), \nabla(\varepsilon -\mathcal{I}\varepsilon)\big) \\ = & \big(\alpha \nabla u + \mathit{\boldsymbol{\sigma}}_{\mathcal{T}}, \nabla(\varepsilon -\mathcal{I}\varepsilon)\big) - \big(\alpha \nabla u_{\mathcal{T}} + \mathit{\boldsymbol{\sigma}}_{\mathcal{T}}, \nabla(\varepsilon -\mathcal{I}\varepsilon)\big) \\ = & \big(f -\nabla\cdot \mathit{\boldsymbol{\sigma}}_{\mathcal{T}}, \varepsilon -\mathcal{I}\varepsilon\big) - \big(\alpha \nabla u_{\mathcal{T}} + \mathit{\boldsymbol{\sigma}}_{\mathcal{T}}, \nabla(\varepsilon -\mathcal{I}\varepsilon)\big) \\ \leq& \left(\sum\limits_{K\in\mathcal{T}} \alpha_K^{-1}h_K^2\left\Vert{f - \nabla\cdot\mathit{\boldsymbol{\sigma}}_{\mathcal{T}}}\right\Vert_{K}^2 \right)^{1/2} \left(\sum\limits_{K\in\mathcal{T}} \alpha_K h_K^{-2}\left\Vert{\varepsilon - \mathcal{I}\varepsilon}\right\Vert_{K}^2 \right)^{1/2} \\ & \left(\sum\limits_{K\in\mathcal{T}} \alpha_K^{-1}\big\Vert{\alpha \nabla u_{\mathcal{T}} + \mathit{\boldsymbol{\sigma}}_{\mathcal{T}}} \big\Vert_{K}^2 \right)^{1/2} \left(\sum\limits_{K\in\mathcal{T}} \alpha_K \left\Vert{\nabla(\varepsilon - \mathcal{I}\varepsilon)}\right\Vert_{K}^2 \right)^{1/2}. \\ & \lesssim \left(\sum\limits_{K\in\mathcal{T}} (\eta_{\mathrm{res},K}^2+\eta_{\mathrm{flux},K}^2) \right)^{1/2} \left(\sum\limits_{K\in\mathcal{T}} \big\Vert{\alpha^{1/2}\nabla \varepsilon}\big\Vert_{\omega_K} \right)^{1/2}. \end{align*} $ |
Applying the norm equivalence of
When
$ \begin{equation} (f, \varepsilon-\mathcal{I}\varepsilon) = \sum\limits_{z \in \mathcal{N}_1} \sum\limits_{K \subset \omega_{z}} \big(f,\left(\varepsilon-\pi_{z} \varepsilon\right) \phi_{z}\big)_{K}, \end{equation} $ | (40) |
in which a patch-wise constant
$ \begin{align*} & \big(f -\nabla\cdot \mathit{\boldsymbol{\sigma}}_{\mathcal{T}}, \varepsilon - \mathcal{I}\varepsilon\big) = \big(f, \varepsilon - \mathcal{I}\varepsilon\big) - \big(\nabla\cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K\nabla u_{\mathcal{T}}), \varepsilon - \mathcal{I}\varepsilon\big) \\ = & \sum\limits_{z \in \mathcal{N}} \sum\limits_{K \subset \omega_{z}} \big(f,\left(\varepsilon-\pi_{z} \varepsilon\right) \phi_{z}\big)_{K} - \big(\nabla\cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K\nabla u_{\mathcal{T}}), \varepsilon - \mathcal{I}\varepsilon\big) \\ \leq & \left(\mathrm{osc}(f;\mathcal{T})^2 \right)^{1/2} \left(\sum\limits_{z \in \mathcal{N}_1} \sum\limits_{K\subset\omega_z} h_{z}^{-2} \|\alpha^{1/2}(\varepsilon-\pi_{z} \varepsilon )\phi_z\|_{K}^2\right)^{1/2} \\ & \; + \left(\sum\limits_{K\in\mathcal{T}} \alpha_K^{-1} h_K^{2} \big\Vert \nabla\cdot (\mathit{\boldsymbol{\sigma}}_{\mathcal{T}} + \alpha_K\nabla u_{\mathcal{T}}) \big\Vert_{K}^2 \right)^{1/2} \left(\sum\limits_{K\in\mathcal{T}} \alpha_K h_K^{-2}\left\Vert{\varepsilon - \mathcal{I}\varepsilon}\right\Vert_{K}^2 \right)^{1/2}. \end{align*} $ |
Applied an inverse inequality in Lemma 7.2 on
The numerics is prepared using the bilinear element for common AMR benchmark problems. The codes for this paper are publicly available on https://github.com/lyc102/ifem implemented using
The adaptive finite element (AFEM) iterative procedure is following the standard
$ \texttt{SOLVE}\longrightarrow \texttt{ESTIMATE} \longrightarrow \texttt{MARK} \longrightarrow \texttt{REFINE}. $ |
The linear system is solved by MATLAB
$ \sum\limits_{K \in \mathcal{M}} \widehat{\eta}^{2}_{K} \geq \theta \sum\limits_{K \in \mathcal{T}} \widehat{\eta}^{2}_{K}, \quad \rm { for } \theta \in(0,1). $ |
Throughout all examples, we fix
$ \eta_{\;{\rm{Residual}}\;,K}^2 : = \alpha_K^{-1}h_K^2 \big\Vert f + \nabla\cdot(\alpha \nabla u_{\mathcal{T}}) \big\Vert_K^2 + \frac{1}{2}\sum\limits_{e\subset \partial K} \frac{h_e}{\alpha_K + \alpha_{K_e}} \big\Vert \lbrack\lbrack {{\alpha \nabla u_{\mathcal{T}}\cdot\mathit{\boldsymbol{n}}_e}} \rbrack\rbrack_{{ {} }} \big\Vert_e^2, $ |
Let
$ \;{\rm{effectivity index}}\; : = {\eta}/{\big\Vert{\alpha^{1/2}\nabla \varepsilon }\big\Vert}, \quad \;{\rm{ where }}\;\; \varepsilon: = u - u_{\mathcal{T}}, \; \eta = \eta_{\;{\rm{Residual}}\;} \;{\rm{ or }}\; \widehat{\eta}, $ |
i.e., the closer to 1 the effectivity index is, the more accurate this estimator is to measure the error of interest. We use an order
$ \ln \eta_n \sim -r_{\eta} \ln N_n + c_1,\quad\;{\rm{and}}\;\quad \ln \Vert{\alpha^{1/2} \nabla (u - u_{\mathcal{T}}) } \Vert \sim -r_{\;{\rm{err}}\;} \ln N_n + c_2, $ |
where the subscript
In this example, a standard AMR benchmark on the L-shaped domain is tested. The true solution
The solution
This example is a common benchmark test problem introduced in [9], see also [17,12]) for elliptic interface problems. The true solution
$ \mu(\theta) = \left\{\begin{array}{ll} \cos ((\pi / 2-\delta) \gamma) \cdot \cos ((\theta-\pi / 2+\rho) \gamma) & {\rm { if }}\; 0 \leq \theta \leq \pi / 2 \\ \cos (\rho \gamma) \cdot \cos ((\theta-\pi+\delta) \gamma) & {\rm { if }}\; \pi / 2 \leq \theta \leq \pi \\ \cos (\delta \gamma) \cdot \cos ((\theta-\pi-\rho) \gamma) & {\rm { if }}\; \pi \leq \theta < 3 \pi / 2 \\ \cos ((\pi / 2-\rho) \gamma) \cdot \cos ((\theta-3 \pi / 2-\delta) \gamma) & {\rm { if }}\; 3 \pi / 2 \leq \theta \leq 2 \pi \end{array}\right. $ |
While
$ \gamma = 0.1,\;\; R \approx 161.4476387975881, \;\; \rho = \pi / 4,\;\; \delta \approx -14.92256510455152, $ |
By this choice, this function is very singular near the origin as the maximum regularity it has is
The AFEM procedure for this problem stops when the relative error reaches
A postprocessed flux with the minimum
However, we do acknowledge that the technical tool involving interpolation is essentially limited to
The author is grateful for the constructive advice from the anonymous reviewers. This work was supported in part by the National Science Foundation under grants DMS-1913080 and DMS-2136075, and no additional revenues are related to this work.
Unlike the identity matrix stabilization commonly used in most of the VEM literature, for
$ \begin{equation} (\!(\mathit{\boldsymbol{\sigma}}, {\mathit{\boldsymbol{\tau}}})\!)_{K} : = \big({\Pi} \mathit{\boldsymbol{\sigma}}, {\Pi} \mathit{\boldsymbol{\tau}} \big)_K + {S}_K\big(({\rm I}-{\Pi} )\mathit{\boldsymbol{\sigma}}, ({\rm I}-{\Pi} )\mathit{\boldsymbol{\tau}}\big), \end{equation} $ | (41) |
where
To show the inverse inequality and the norm equivalence used in the reliability bound, on each element, we need to introduce some geometric measures. Consider a polygonal element
Proposition 1. Under Assumption 1,
Lemma 7.1 (Trace inequality on small edges [13]). If Proposition 1 holds, for
$ \begin{equation} h_e^{-1/2}\left\Vert{v}\right\Vert_{e} \lesssim h_K^{-1} \left\Vert{v}\right\Vert_{K} + \left\Vert{\nabla v}\right\Vert_{K}, \quad \mathit{on} \;e\subset K. \end{equation} $ | (42) |
Proof. The proof follows essentially equation (3.9) in [13,Lemma 3.3] as a standard scaled trace inequality on
$ h_e^{-1/2}\left\Vert{v}\right\Vert_{e} \lesssim h_e^{-1} \left\Vert{v}\right\Vert_{T_e} + \left\Vert{\nabla v}\right\Vert_{T_e} \lesssim h_K^{-1} \left\Vert{v}\right\Vert_{K} + \left\Vert{\nabla v}\right\Vert_{K}. $ |
Lemma 7.2 (Inverse inequalities). Under Assumption 1, we have the following inverse estimates for
$ \begin{equation} \|\nabla \cdot \mathit{\boldsymbol{\tau}}\|_K \lesssim h_K^{-1} \|\mathit{\boldsymbol{\tau}}\|_K, \quad \mathit{and} \quad \|\nabla \cdot \mathit{\boldsymbol{\tau}}\|_K \lesssim h_K^{-1} S_K\big(\mathit{\boldsymbol{\tau}},\mathit{\boldsymbol{\tau}}\big)^{1/2}. \end{equation} $ | (43) |
Proof. The first inequality in (43) can be shown using a bubble function trick. Choose
$ \|\nabla \cdot \mathit{\boldsymbol{\tau}}\|_K^2 \lesssim (\nabla \cdot \mathit{\boldsymbol{\tau}}, p b_K) = -(\mathit{\boldsymbol{\tau}}, \nabla (p b_K)) \leq \left\Vert{\mathit{\boldsymbol{\tau}}}\right\Vert_K \left\Vert{ \nabla (p b_K)}\right\Vert_K, $ |
and then
$ \left\Vert{ \nabla (p b_K)}\right\Vert \leq \left\Vert{ b_K \nabla p }\right\Vert_K + \left\Vert{p\nabla b_K}\right\Vert_K \leq \left\Vert{ b_K }\right\Vert_{\infty,\Omega} \left\Vert{\nabla p }\right\Vert_K + \left\Vert{p}\right\Vert_K \left\Vert{\nabla b_K}\right\Vert_{\infty,K}. $ |
Consequently, the first inequality in (43) follows above by the standard inverse estimate for polynomials
To prove the second inequality in (43), by integration by parts we have
$ \begin{equation} \left\Vert{\nabla\cdot\mathit{\boldsymbol{\tau}}}\right\Vert^2 = (\nabla\cdot\mathit{\boldsymbol{\tau}}, p) = -(\mathit{\boldsymbol{\tau}},\nabla p) + \sum\limits_{e\subset\partial K} (\mathit{\boldsymbol{\tau}}\cdot \mathit{\boldsymbol{n}}_e, p). \end{equation} $ | (44) |
Expand
$ \begin{equation} \left\Vert{p}\right\Vert_K^2 = \mathbf{p}^{\top} \mathbf{M} \mathbf{p} \geq \mathbf{p}^{\top} \operatorname{diag}(\mathbf{M}) \mathbf{p} \geq \min\limits_j \mathbf{M}_{jj}\left\Vert{\mathbf{p}}\right\Vert_{\ell^2}^2 \simeq h_K^2 \left\Vert{\mathbf{p}}\right\Vert_{\ell^2}^2, \end{equation} $ | (45) |
since
$ \begin{aligned} \left\Vert{\nabla\cdot\mathit{\boldsymbol{\tau}}}\right\Vert^2 & \leq \left(\sum\limits_{\alpha\in \Lambda} (\mathit{\boldsymbol{\tau}}, m_{\alpha})_K^2 \right)^{1/2} \left(\sum\limits_{\alpha\in \Lambda} p_{\alpha}^2 \right)^{1/2} \\ & \quad + \left(\sum\limits_{e\subset \partial K} h_e \left\Vert{\mathit{\boldsymbol{\tau}}\cdot \mathit{\boldsymbol{n}}_e}\right\Vert_e^2 \right)^{1/2} \left(\sum\limits_{e\subset \partial K} h_e^{-1} \left\Vert{p}\right\Vert_e^2 \right)^{1/2} \\ & \lesssim S_K(\mathit{\boldsymbol{\tau}},\mathit{\boldsymbol{\tau}})^{1/2} \left(\left\Vert{\mathbf{p}}\right\Vert_{\ell^2} + h_K^{-1} \left\Vert{p}\right\Vert_K + \left\Vert{\nabla p}\right\Vert_K \right). \end{aligned} $ |
As a result, the second inequality in (43) is proved when apply an inverse inequality for
Remark 2. While the proof in Lemma 7.2 relies on
Lemma 7.3 (Norm equivalence). Under Assumption 1, let
$ \begin{equation} \gamma_* \Vert{\mathit{\boldsymbol{\tau}}}\Vert_K \leq \Vert{ \mathit{\boldsymbol{\tau}}}\Vert_{h,K} \leq \gamma^*\Vert{\mathit{\boldsymbol{\tau}}}\Vert_K, \end{equation} $ | (46) |
where both
Proof. First we consider the lower bound, by triangle inequality,
$ \Vert{\mathit{\boldsymbol{\tau}}}\Vert_{K}\leq \big\Vert{{\Pi}\mathit{\boldsymbol{\tau}}}\big\Vert_{K} + \big\Vert{(\mathit{\boldsymbol{\tau}} - {\Pi}\mathit{\boldsymbol{\tau}}) }\big\Vert_{K}. $ |
Since
$ \begin{equation} \Vert{\mathit{\boldsymbol{\tau}} }\Vert_{K}^2 \leq S_K\big(\mathit{\boldsymbol{\tau}},\mathit{\boldsymbol{\tau}}\big), \quad \;{\rm{ for }}\; \mathit{\boldsymbol{\tau}}\in \mathcal{V}_k(K). \end{equation} $ | (47) |
To this end, we consider the weak solution to the following auxiliary boundary value problem on
$ \begin{equation} \left\{ \begin{aligned} \Delta \psi & = \nabla\cdot \mathit{\boldsymbol{\tau}}&\;{\rm{ in }}\; K, \\ \frac{\partial \psi}{\partial n} & = \mathit{\boldsymbol{\tau}} \cdot\mathit{\boldsymbol{n}}_{\partial K} &\;{\rm{ on }}\;\partial K. \end{aligned} \right. \end{equation} $ | (48) |
By a standard Helmholtz decomposition result (e.g. Proposition 3.1, Chapter 1[23]), we have
$ \left\Vert{\mathit{\boldsymbol{\tau}} - \nabla \psi}\right\Vert_K^2 = (\mathit{\boldsymbol{\tau}} - \nabla \psi, \nabla^{\perp} \phi) = 0. $ |
Consequently, we proved essentially the unisolvency of the modified VEM space (4) and
$ \begin{equation} \begin{aligned} & \big\Vert{\mathit{\boldsymbol{\tau}} }\big\Vert_{K}^2 = (\mathit{\boldsymbol{\tau}}, \nabla \psi)_K = \big(\mathit{\boldsymbol{\tau}}, \nabla \psi \big)_K \\ = & \; -\big(\nabla\cdot\mathit{\boldsymbol{\tau}}, \psi \big)_K+ (\mathit{\boldsymbol{\tau}}\cdot\mathit{\boldsymbol{n}}_{\partial K} ,\psi )_{\partial K} \\ \leq & \;\|\nabla \cdot \mathit{\boldsymbol{\tau}}\|_K \| \psi\|_K + \sum\limits_{e\subset \partial K} \|\mathit{\boldsymbol{\tau}}\cdot\mathit{\boldsymbol{n}}_e\|_e\| \psi \|_e \\ \leq &\; \|\nabla \cdot \mathit{\boldsymbol{\tau}}\|_K \| \psi\|_K + \left(\sum\limits_{e\subset \partial K} h_e\|\mathit{\boldsymbol{\tau}}\cdot\mathit{\boldsymbol{n}}_e\|_e^2\right)^{1/2} \left(\sum\limits_{e\subset \partial K} h_e^{-1}\|\psi\|_e^2\right)^{1/2} \end{aligned} \end{equation} $ | (49) |
Proposition 1 allows us to apply an isotropic trace inequality on an edge of a polygon (Lemma 7.1), combining with the Poincaré inequality for
$ h_e^{-1/2}\|\psi\|_e \lesssim h_K^{-1} \|\psi\|_K + \|\nabla \psi\|_K \lesssim \|\nabla \psi\|_K. $ |
Furthermore applying the inverse estimate in Lemma 7.2 on the bulk term above, we have
$ \big\Vert{\mathit{\boldsymbol{\tau}} }\big\Vert_{K}^2 \lesssim S_K\big(\mathit{\boldsymbol{\tau}},\mathit{\boldsymbol{\tau}}\big)^{1/2} \|\nabla \psi\|_K, $ |
which proves the validity of (47), thus yield the lower bound.
To prove the upper bound, by
$ \begin{equation} h_e\|\mathit{\boldsymbol{\tau}}\cdot\mathit{\boldsymbol{n}}_e\|_e^2 \lesssim \left\Vert{\mathit{\boldsymbol{\tau}}}\right\Vert_K,\quad \;{\rm{ and }}\; \quad |(\mathit{\boldsymbol{\tau}}, \nabla m_{\alpha})_K| \leq \left\Vert{\mathit{\boldsymbol{\tau}}}\right\Vert_K. \end{equation} $ | (50) |
To prove the first inequality, by Proposition 1 again, consider the edge bubble function
$ \begin{equation} \left\Vert{\nabla b_e}\right\Vert_{\infty,K} = O(1/h_e), \;{\rm{ and }}\; \left\Vert{b_e}\right\Vert_{\infty, K} = O(1). \end{equation} $ | (51) |
Denote
$ \begin{aligned} \|\mathit{\boldsymbol{\tau}}\cdot\mathit{\boldsymbol{n}}_e\|_e^2 & \lesssim \big(\mathit{\boldsymbol{\tau}}\cdot\mathit{\boldsymbol{n}}_e, b_e q_e \big)_e = x\big(\mathit{\boldsymbol{\tau}}\cdot\mathit{\boldsymbol{n}}_e, b_e q_e \big)_{\partial K} \\ & = \big(\mathit{\boldsymbol{\tau}}, q_e\nabla b_e \big)_K + \big(\nabla\cdot\mathit{\boldsymbol{\tau}}, b_e q_e\big)_K \\ & \leq \left\Vert{\mathit{\boldsymbol{\tau}}}\right\Vert_K \left\Vert{q_e\nabla b_e}\right\Vert_{T_e} + \left\Vert{\nabla\cdot\mathit{\boldsymbol{\tau}}}\right\Vert_K \left\Vert{q_e b_e}\right\Vert_{T_e}, \\ & \leq \left\Vert{\mathit{\boldsymbol{\tau}}}\right\Vert_K \left\Vert{q_e}\right\Vert_{T_e} \left\Vert{\nabla b_e}\right\Vert_{\infty,K} + \left\Vert{\nabla\cdot\mathit{\boldsymbol{\tau}}}\right\Vert_K \left\Vert{q_e}\right\Vert_{T_e} \left\Vert{b_e}\right\Vert_{\infty,K}. \end{aligned} $ |
Now by the fact that
The second inequality in (50) can be estimated straightforward by the scaling of the monomials (7)
$ \begin{equation} \left|(\mathit{\boldsymbol{\tau}}, \nabla m_{\alpha})_K\right| \leq \left\Vert{\mathit{\boldsymbol{\tau}}}\right\Vert_K \left\Vert{\nabla m_{\alpha}}\right\Vert_K \leq \left\Vert{\mathit{\boldsymbol{\tau}}}\right\Vert_K . \end{equation} $ | (52) |
Hence, (46) is proved.
[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
![]() |
1. | Hengliang Tang, Jinda Dong, Solving Flexible Job-Shop Scheduling Problem with Heterogeneous Graph Neural Network Based on Relation and Deep Reinforcement Learning, 2024, 12, 2075-1702, 584, 10.3390/machines12080584 | |
2. | Chen Han, Xuanyin Wang, TPN:Triple network algorithm for deep reinforcement learning, 2024, 591, 09252312, 127755, 10.1016/j.neucom.2024.127755 | |
3. | Miguel S. E. Martins, João M. C. Sousa, Susana Vieira, A Systematic Review on Reinforcement Learning for Industrial Combinatorial Optimization Problems, 2025, 15, 2076-3417, 1211, 10.3390/app15031211 | |
4. | Tianhua Jiang, Lu Liu, A Bi-Population Competition Adaptive Interior Search Algorithm Based on Reinforcement Learning for Flexible Job Shop Scheduling Problem, 2025, 24, 1469-0268, 10.1142/S1469026824500251 | |
5. | Tianyuan Mao, A Review of Scheduling Methods for Multi-AGV Material Handling Systems in Mixed-Model Assembly Workshops, 2025, 5, 2710-0723, 227, 10.54691/p4x5a536 | |
6. | Peng Zhao, You Zhou, Di Wang, Zhiguang Cao, Yubin Xiao, Xuan Wu, Yuanshu Li, Hongjia Liu, Wei Du, Yuan Jiang, Liupu Wang, 2025, Dual Operation Aggregation Graph Neural Networks for Solving Flexible Job-Shop Scheduling Problem with Reinforcement Learning, 9798400712746, 4089, 10.1145/3696410.3714616 | |
7. | Yuxin Peng, Youlong Lyu, Jie Zhang, Ying Chu, Heterogeneous Graph Neural-Network-Based Scheduling Optimization for Multi-Product and Variable-Batch Production in Flexible Job Shops, 2025, 15, 2076-3417, 5648, 10.3390/app15105648 |