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

Performance of protein-ligand docking with CDK4/6 inhibitors: a case study

  • It is widely believed that tertiary protein-ligand interactions are essential in determining protein function. Currently, the structure sampling and scoring function in traditional docking methods still have limitations. Therefore, new methods for protein-ligand docking are desirable. The accurate docking can speed up the early-stage development of new drugs. Here we present a multi-source information-based protein-ligand docking approach (pmDock). In the CDK4/6 inhibitor case study, pmDock produces a substantial accuracy increases between the predicted geometry centers of ligands and experiments compared to AutoDock and SwissDock alone. Also, pmDock improves predictions for critical binding sites and captures more tertiary binding interactions. Our results demonstrate that pmDock is a reliable docking method for accurate protein-ligand prediction.

    Citation: Linlu Song, Shangbo Ning, Jinxuan Hou, Yunjie Zhao. Performance of protein-ligand docking with CDK4/6 inhibitors: a case study[J]. Mathematical Biosciences and Engineering, 2021, 18(1): 456-470. doi: 10.3934/mbe.2021025

    Related Papers:

    [1] Tianyuan Xu, Shanming Ji, Chunhua Jin, Ming Mei, Jingxue Yin . EARLY AND LATE STAGE PROFILES FOR A CHEMOTAXIS MODEL WITH DENSITY-DEPENDENT JUMP PROBABILITY. Mathematical Biosciences and Engineering, 2018, 15(6): 1345-1385. doi: 10.3934/mbe.2018062
    [2] Wenjie Zhang, Lu Xu, Qiao Xin . Global boundedness of a higher-dimensional chemotaxis system on alopecia areata. Mathematical Biosciences and Engineering, 2023, 20(5): 7922-7942. doi: 10.3934/mbe.2023343
    [3] Sunwoo Hwang, Seongwon Lee, Hyung Ju Hwang . Neural network approach to data-driven estimation of chemotactic sensitivity in the Keller-Segel model. Mathematical Biosciences and Engineering, 2021, 18(6): 8524-8534. doi: 10.3934/mbe.2021421
    [4] Qianhong Zhang, Fubiao Lin, Xiaoying Zhong . On discrete time Beverton-Holt population model with fuzzy environment. Mathematical Biosciences and Engineering, 2019, 16(3): 1471-1488. doi: 10.3934/mbe.2019071
    [5] Chichia Chiu, Jui-Ling Yu . An optimal adaptive time-stepping scheme for solving reaction-diffusion-chemotaxis systems. Mathematical Biosciences and Engineering, 2007, 4(2): 187-203. doi: 10.3934/mbe.2007.4.187
    [6] Xu Song, Jingyu Li . Asymptotic stability of spiky steady states for a singular chemotaxis model with signal-suppressed motility. Mathematical Biosciences and Engineering, 2022, 19(12): 13988-14028. doi: 10.3934/mbe.2022652
    [7] Tingting Yu, Sanling Yuan . Dynamics of a stochastic turbidostat model with sampled and delayed measurements. Mathematical Biosciences and Engineering, 2023, 20(4): 6215-6236. doi: 10.3934/mbe.2023268
    [8] Lin Zhang, Yongbin Ge, Zhi Wang . Positivity-preserving high-order compact difference method for the Keller-Segel chemotaxis model. Mathematical Biosciences and Engineering, 2022, 19(7): 6764-6794. doi: 10.3934/mbe.2022319
    [9] Changwook Yoon, Sewoong Kim, Hyung Ju Hwang . Global well-posedness and pattern formations of the immune system induced by chemotaxis. Mathematical Biosciences and Engineering, 2020, 17(4): 3426-3449. doi: 10.3934/mbe.2020194
    [10] Marcin Choiński, Mariusz Bodzioch, Urszula Foryś . A non-standard discretized SIS model of epidemics. Mathematical Biosciences and Engineering, 2022, 19(1): 115-133. doi: 10.3934/mbe.2022006
  • It is widely believed that tertiary protein-ligand interactions are essential in determining protein function. Currently, the structure sampling and scoring function in traditional docking methods still have limitations. Therefore, new methods for protein-ligand docking are desirable. The accurate docking can speed up the early-stage development of new drugs. Here we present a multi-source information-based protein-ligand docking approach (pmDock). In the CDK4/6 inhibitor case study, pmDock produces a substantial accuracy increases between the predicted geometry centers of ligands and experiments compared to AutoDock and SwissDock alone. Also, pmDock improves predictions for critical binding sites and captures more tertiary binding interactions. Our results demonstrate that pmDock is a reliable docking method for accurate protein-ligand prediction.


    Baló's concentric sclerosis (BCS) was first described by Marburg [1] in 1906, but became more widely known until 1928 when the Hungarian neuropathologist Josef Baló published a report of a 23-year-old student with right hemiparesis, aphasia, and papilledema, who at autopsy had several lesions of the cerebral white matter, with an unusual concentric pattern of demyelination [2]. Traditionally, BCS is often regarded as a rare variant of multiple sclerosis (MS). Clinically, BCS is most often characterized by an acute onset with steady progression to major disability and death with months, thus resembling Marburg's acute MS [3,4]. Its pathological hallmarks are oligodendrocyte loss and large demyelinated lesions characterized by the annual ring-like alternating pattern of demyelinating and myelin-preserved regions. In [5], the authors found that tissue preconditioning might explain why Baló lesions develop a concentric pattern. According to the tissue preconditioning theory and the analogies between Baló's sclerosis and the Liesegang periodic precipitation phenomenon, Khonsari and Calvez [6] established the following chemotaxis model

    ˜uτ=DΔX˜udiffusion ofactivated macrophagesX(˜χ˜u(ˉu˜u)˜v)chemoattractant attractssurrounding activated macrophages+μ˜u(ˉu˜u)production of activated macrophages,˜ϵΔX˜vdiffusion of chemoattractant=˜α˜v+˜β˜wdegradationproduction of chemoattractant,˜wτ=κ˜uˉu+˜u˜u(ˉw˜w)destruction of oligodendrocytes, (1.1)

    where ˜u, ˜v and ˜w are, respectively, the density of activated macrophages, the concentration of chemoattractants and density of destroyed oligodendrocytes. ˉu and ˉw represent the characteristic densities of macrophages and oligodendrocytes respectively.

    By numerical simulation, the authors in [6,7] indicated that model (1.1) only produces heterogeneous concentric demyelination and homogeneous demyelinated plaques as χ value gradually increases. In addition to the chemoattractant produced by destroyed oligodendrocytes, "classically activated'' M1 microglia also can release cytotoxicity [8]. Therefore we introduce a linear production term into the second equation of model (1.1), and establish the following BCS chemotaxis model with linear production term

    {˜uτ=DΔX˜uX(˜χ˜u(ˉu˜u)˜v)+μ˜u(ˉu˜u),˜ϵΔX˜v+˜α˜v=˜β˜w+˜γ˜u,˜wτ=κ˜uˉu+˜u˜u(ˉw˜w). (1.2)

    Before going to details, let us simplify model (1.2) with the following scaling

    u=˜uˉu,v=μˉu˜ϵD˜v,w=˜wˉw,t=μˉuτ,x=μˉuDX,χ=˜χ˜ϵμ,α=D˜α˜ϵμˉu,β=˜βˉw,γ=˜γˉu,δ=κμ,

    then model (1.2) takes the form

    {ut=Δu(χu(1u)v)+u(1u),xΩ,t>0,Δv+αv=βw+γu,xΩ,t>0,wt=δu1+uu(1w),xΩ,t>0,ηu=ηv=0,xΩ,t>0,u(x,0)=u0(x),w(x,0)=w0(x),xΩ, (1.3)

    where ΩRn(n1) is a smooth bounded domain, η is the outward normal vector to Ω, η=/η, δ balances the speed of the front and the intensity of the macrophages in damaging the myelin. The parameters χ,α and δ are positive constants as well as β,γ are nonnegative constants.

    If δ=0, then model (1.3) is a parabolic-elliptic chemotaxis system with volume-filling effect and logistic source. In order to be more line with biologically realistic mechanisms, Hillen and Painter [9,10] considered the finite size of individual cells-"volume-filling'' and derived volume-filling models

    {ut=(Du(q(u)q(u)u)uq(u)uχ(v)v)+f(u,v),vt=DvΔv+g(u,v). (1.4)

    q(u) is the probability of the cell finding space at its neighbouring location. It is also called the squeezing probability, which reflects the elastic properties of cells. For the linear choice of q(u)=1u, global existence of solutions to model (1.4) in any space dimension are investigated in [9]. Wang and Thomas [11] established the global existence of classical solutions and given necessary and sufficient conditions for spatial pattern formation to a generalized volume-filling chemotaxis model. For a chemotaxis system with generalized volume-filling effect and logistic source, the global boundedness and finite time blow-up of solutions are obtained in [12]. Furthermore, the pattern formation of the volume-filling chemotaxis systems with logistic source and both linear diffusion and nonlinear diffusion are shown in [13,14,15] by the weakly nonlinear analysis. For parabolic-elliptic Keller-Segel volume-filling chemotaxis model with linear squeezing probability, asymptotic behavior of solutions is studied both in the whole space Rn [16] and on bounded domains [17]. Moreover, the boundedness and singularity formation in parabolic-elliptic Keller-Segel volume-filling chemotaxis model with nonlinear squeezing probability are discussed in [18,19].

    Very recently, we [20] investigated the uniform boundedness and global asymptotic stability for the following chemotaxis model of multiple sclerosis

    {ut=Δu(χ(u)v)+u(1u),χ(u)=χu1+u,xΩ,t>0,τvt=Δvβv+αw+γu,xΩ,t>0,wt=δu1+uu(1w),xΩ,t>0,

    subject to the homogeneous Neumann boundary conditions.

    In this paper, we are first devoted to studying the local existence and uniform boundedness of the unique classical solution to system (1.3) by using Neumann heat semigroup arguments, Banach fixed point theorem, parabolic Schauder estimate and elliptic regularity theory. Then we discuss that exponential asymptotic stability of the positive equilibrium point to system (1.3) by constructing Lyapunov function.

    Although, in the pathological mechanism of BCS, the initial data in model (1.3) satisfy 0<u0(x)1,w0(x)=0, we mathematically assume that

    {u0(x)C0(ˉΩ)with0,u0(x)1inΩ,w0(x)C2+ν(ˉΩ)with0<ν<1and0w0(x)1inΩ. (1.5)

    It is because the condition (1.5) implies u(x,t0)>0 for any t0>0 by the strong maximum principle.

    The following theorems give the main results of this paper.

    Theorem 1.1. Assume that the initial data (u0(x),w0(x)) satisfy the condition (1.5). Then model (1.3) possesses a unique global solution (u(x,t),v(x,t),w(x,t)) satisfying

    u(x,t)C0(ˉΩ×[0,))C2,1(ˉΩ×(0,)),v(x,t)C0((0,),C2(ˉΩ)),w(x,t)C2,1(ˉΩ×[0,)), (1.6)

    and

    0<u(x,t)1,0v(x,t)β+γα,w0(x)w(x,t)1,inˉΩ×(0,).

    Moreover, there exist a ν(0,1) and M>0 such that

    uC2+ν,1+ν/2(ˉΩ×[1,))+vC0([1,),C2+ν(ˉΩ))+wCν,1+ν/2(ˉΩ×[1,))M. (1.7)

    Theorem 1.2. Assume that β0,γ0,β+γ>0 and

    χ<{min (1.8)

    Let (u, v, w) be a positive classical solution of the problem (1.3), (1.5). Then

    \begin{equation} \|u(\cdot, t)-u^{\ast}\|_{L^{\infty}(\Omega)}+\|v(\cdot, t)-v^{\ast}\|_{L^{\infty}(\Omega)} +\|w(\cdot, t)-w^{\ast}\|_{L^{\infty}(\Omega)}\rightarrow 0, \; \; \mathit{\text{as}}\, t\rightarrow \infty. \end{equation} (1.9)

    Furthermore, there exist positive constants \lambda = \lambda(\chi, \alpha, \gamma, \delta, n) and C = C(|\Omega|, \chi, \alpha, \beta, \gamma, \delta) such that

    \begin{equation} \|u-u^{\ast}\|_{L^{\infty}(\Omega)}\leq C e^{-\lambda t}, \, \|v-v^{\ast}\|_{L^{\infty}(\Omega)}\leq C e^{-\lambda t}, \, \|w-w^{\ast}\|_{L^{\infty}(\Omega)} \leq C e^{-\lambda t}, \; \; t \gt 0, \end{equation} (1.10)

    where (u^{\ast}, v^{\ast}, w^{\ast}) = (1, \frac{\beta+\gamma}{\alpha}, 1) is the unique positive equilibrium point of the model (1.3).

    The paper is organized as follows. In section 2, we prove the local existence, the boundedness and global existence of a unique classical solution. In section 3, we firstly establish the uniform convergence of the positive global classical solution, then discuss the exponential asymptotic stability of positive equilibrium point in the case of weak chemotactic sensitivity. The paper ends with a brief concluding remarks.

    The aim of this section is to develop the existence and boundedness of a global classical solution by employing Neumann heat semigroup arguments, Banach fixed point theorem, parabolic Schauder estimate and elliptic regularity theory.

    Proof of Theorem 1.1 (ⅰ) Existence. For p\in (1, \infty) , let A denote the sectorial operator defined by

    Au: = -\Delta u \; \mathrm{for}\; u\in D(A): = \Big\{\varphi\in W^{2, p}(\Omega)\Big|\frac{\partial}{\partial \eta}\varphi\Big|_{\partial\Omega} = 0\Big\}.

    \lambda_{1} > 0 denote the first nonzero eigenvalue of -\Delta in \Omega with zero-flux boundary condition. Let A_{1} = -\Delta+\alpha and X^{l} be the domains of fractional powers operator A^{l}, \; l\geq 0 . From the Theorem 1.6.1 in [21], we know that for any p > n and l\in(\frac{n}{2p}, \frac{1}{2}) ,

    \begin{equation} \|z\|_{L^{\infty}(\Omega)}\leq C\|A_{1}^{l}z\|_{ L^{p}(\Omega)}\, \, \mathrm{for\; all}\, \, z\in X^{l}. \end{equation} (2.1)

    We introduce the closed subset

    S: = \left\{u\in X\big| \|u\|_{L^{\infty}((0, T);L^{\infty}(\Omega))}\leq R+1\right\}

    in the space X: = C^{0}([0, T];C^{0}(\bar{\Omega})) , where R is a any positive number satisfying

    \|u_{0}(x)\|_{L^{\infty}(\Omega)}\leq R

    and T > 0 will be specified later. Note F(u) = \frac{u}{1+u} , we consider an auxiliary problem with F(u) replaced by its extension \tilde{F}(u) defined by

    \tilde{F}(u) = \begin{cases} F(u)u\; \; &\text{if}\; \; u\geq 0, \\ -F(-u)(-u)\; \; &\text{if}\; \; u \lt 0. \end{cases}

    Notice that \tilde{F}(u) is a smooth globally Lipshitz function. Given \hat{u}\in S , we define \Psi\hat{u} = u by first writing

    \begin{equation} w(x, t) = (w_{0}(x)-1)e^{-\delta\int_{0}^{t}\tilde{F} (\hat{u})\hat{u}ds}+1, \; \; x\in\Omega, \; t \gt 0, \end{equation} (2.2)

    and

    w_{0}\leq w(x, t)\leq 1, \; \; x\in\Omega, \; t \gt 0,

    then letting v solve

    \begin{equation} \left\{\begin{array}{ll} -\Delta v+\alpha v = \beta w+\gamma \hat{u}, &x\in\Omega, \; t\in(0, T), \\ \partial_{\eta}v = 0, &x\in\partial\Omega, \; t\in(0, T), \\ \end{array} \right. \end{equation} (2.3)

    and finally taking u to be the solution of the linear parabolic problem

    \left\{\begin{array}{ll} u_{t} = \Delta u-\chi \nabla\cdot(\hat{u}(1-\hat{u})\nabla v)+\hat{u}(1-\hat{u}), &x\in\Omega, \; t\in(0, T), \\ \partial_{\eta}u = 0, &x\in\partial\Omega, \; t\in(0, T), \\ u(x, 0) = u_{0}(x), &x\in\Omega.\\ \end{array} \right.

    Applying Agmon-Douglas-Nirenberg Theorem [22,23] for the problem (2.3), there exists a constant C such that

    \begin{equation} \begin{aligned} \|v\|_{W_{p}^{2}(\Omega)}&\leq C(\beta\|w\|_{L^{p}(\Omega)}+\gamma\|\hat{u}\|_{L^{p}(\Omega)})\\ &\leq C(\beta|\Omega|^{\frac{1}{p}}+\gamma (R+1)) \end{aligned} \end{equation} (2.4)

    for all t\in(0, T) . From a variation-of-constants formula, we define

    \Psi(\hat{u}) = e^{t\Delta}u_{0}-\chi\int^{t}_{0}e^{(t-s)\Delta}\nabla\cdot\left(\hat{u}(1-\hat{u})\nabla v(s)\right)ds+\int^{t}_{0}e^{(t-s)\Delta}\hat{u}(s)(1-\hat{u}(s))ds.

    First we shall show that for T small enough

    \|\Psi(\hat{u})\|_{L^{\infty}((0, T);L^{\infty}(\Omega))}\leq R+1

    for any \hat{u}\in S . From the maximum principle, we can give

    \begin{equation} \|e^{t\Delta}u_{0}\|_{L^{\infty}(\Omega)}\leq \|u_{0}\|_{L^{\infty}(\Omega)}, \end{equation} (2.5)

    and

    \begin{equation} \begin{aligned} \int^{t}_{0}\|e^{t\Delta}\hat{u}(s)(1-\hat{u}(s))\|_{L^{\infty} (\Omega)}ds \leq& \int^{t}_{0}\|\hat{u}(s)(1-\hat{u}(s))\|_{L^{\infty} (\Omega)}ds\\ \leq&(R+1)(R+2)T \end{aligned} \end{equation} (2.6)

    for all t\in(0, T) . We use inequalities (2.1) and (2.4) to estimate

    \begin{equation} \begin{aligned} &\chi\int^{t}_{0}\|e^{(t-s)\Delta}\nabla\cdot(\hat{u}(1-\hat{u})\nabla v(s))\|_{L^{\infty}(\Omega)}ds\\ \leq& C\int^{t}_{0}(t-s)^{-l} \|e^{\frac{t-s}{2}\Delta}\nabla\cdot(\hat{u}(1-\hat{u})\nabla v(s))\|_{L^{p}(\Omega)}ds \\ \leq& C\int^{t}_{0}(t-s)^{-l-\frac{1}{2}} \|(\hat{u}(1-\hat{u})\nabla v(s)\|_{L^{p}(\Omega)}ds \\ \leq& C T^{\frac{1}{2}-l}(R+1)(R+2)(\beta|\Omega|^{\frac{1}{p}}+\gamma (R+1)) \end{aligned} \end{equation} (2.7)

    for all t\in(0, T) . This estimate is attributed to T < 1 and the inequality in [24], Lemma 1.3 iv]

    \| e^{t\Delta}\nabla z\|_{L^{p}(\Omega)}\leq C_{1}(1+t^{-\frac{1}{2}})e^{-\lambda_{1}t}\| z\|_{L^{p}(\Omega)}\; \mathrm{for\; all}\; \; z\in C^{\infty}_{c}(\Omega).

    From inequalities (2.5), (2.6) and (2.7) we can deduce that \Psi maps S into itself for T small enough.

    Next we prove that the map \Psi is a contractive on S . For \hat{u}_{1}, \hat{u}_{2}\in S , we estimate

    \begin{align*} \; \; \; \; \; \; \; \; &\|\Psi(\hat{u}_{1})-\Psi(\hat{u}_{2})\|_{L^{\infty}(\Omega)} \\ \leq & \chi \int^{t}_{0}(t-s)^{-l-\frac{1}{2}}\|\left[\hat{u}_{2}(s)(1-\hat{u}_{2}(s))-\hat{u}_{1}(s)(1-\hat{u}_{1}(s))\right] \nabla v_{2}(s)\|_{L^{p}(\Omega)}ds\\ &+\chi \int^{t}_{0}\|\hat{u}_{1}(s)(1-\hat{u}_{1}(s))(\nabla v_{1}(s)- \nabla v_{2}(s))\|_{L^{p}(\Omega)}ds \\ &+\int^{t}_{0}\|e^{(t-s)\Delta}[\hat{u}_{1}(s)(1-\hat{u}_{1}(s))-\hat{u}_{2}(s)(1-\hat{u}_{2}(s))]\|_{L^{\infty}(\Omega)}ds \\ \leq & \chi \int^{t}_{0}(t-s)^{-l-\frac{1}{2}}(2R+1)\|\hat{u}_{1}(s)-\hat{u}_{2}(s)\|_{X}\|\nabla v_{2}(s)\|_{L^{p}(\Omega)}ds\\ &+\chi \int^{t}_{0}(R+1)(R+2)\left(\beta \|w_{1}(s)-w_{2}(s)\|_{L^{p}(\Omega)}+\gamma \|\hat{u}_{1}(s)-\hat{u}_{2}(s)\|_{L^{p}(\Omega)}\right)ds \\ & +\int^{t}_{0}(2R+1)\|\hat{u}_{1}(s)-\hat{u}_{2}(s)\|_{X}ds \\ \leq & \chi \int^{t}_{0}(t-s)^{-l-\frac{1}{2}}(2R+1)\|\hat{u}_{1}(s)-\hat{u}_{2}(s)\|_{X}\|\nabla v_{2}(s)\|_{L^{p}(\Omega)}ds\\ &+2\beta\delta \chi \int^{t}_{0}(R+1)(R+2)t\|\hat{u}_{1}(s)-\hat{u}_{2}(s)\|_{L^{p}(\Omega)}+\gamma \|\hat{u}_{1}(s)-\hat{u}_{2}(s)\|_{L^{p}(\Omega)}ds\\ & +\int^{t}_{0}(2R+1)\|\hat{u}_{1}(s)-\hat{u}_{2}(s)\|_{X}ds \\ \leq & \left(C\chi T^{\frac{1}{2}-l}(2R+1)(\beta|\Omega|^{\frac{1}{p}}+\gamma (R+1))+2\beta\delta \chi T(R^{2}+3R+\gamma+2)+T(2R+1)\right)\|\hat{u}_{1}(s)-\hat{u}_{2}(s)\|_{X}. \end{align*}

    Fixing T\in(0, 1) small enough such that

    \left(C\chi T^{\frac{1}{2}-l}(2R+1)(\beta|\Omega|^{\frac{1}{p}}+\gamma (R+1))+2\beta\delta \chi T(R^{2}+3R+\gamma+2)+T(2R+1)\right)\leq \frac{1}{2}.

    It follows from the Banach fixed point theorem that there exists a unique fixed point of \Psi .

    (ⅱ) Regularity. Since the above of T depends on \|u_{0}\|_{L^{\infty}(\Omega)} and \|w_{0}\|_{L^{\infty}(\Omega)} only, it is clear that (u, v, w) can be extended up to some maximal T_{\max}\in(0, \infty] . Let Q_{T} = \Omega \times (0, T] for all T\in (0, T_{\max}) . From u\in C^{0}(\bar{Q}_{T}) , we know that w\in C^{0, 1}(\bar{Q}_{T}) by the expression (2.2) and v\in C^{0}([0, T], W_{p}^{2}(\Omega)) by Agmon-Douglas-Nirenberg Theorem [22,23]. From parabolic L^{p} -estimate and the embedding relation W_{p}^{1}(\Omega)\hookrightarrow C^{\nu}(\bar{\Omega}), \; p > n , we can get u\in W^{2, 1}_{p}(Q_{T}) . By applying the following embedding relation

    \begin{equation} W^{2, 1}_{p}(Q_{T})\hookrightarrow C^{\nu, \nu/2}(\bar{Q}_{T}), \; p \gt \frac{n+2}{2}, \end{equation} (2.8)

    we can derive u(x, t)\in C^{\nu, \nu/2}(\bar{Q}_{T}) with 0 < \nu\leq 2-\frac{n+2}{p} . The conclusion w\in C^{\nu, 1+\nu/2}(\bar{Q}_{T}) can be obtained by substituting u\in C^{\nu, \nu/2}(\bar{Q}_{T}) into the formulation (2.2). The regularity u\in C^{2+\nu, 1+\nu/2}(\bar{Q}_{T}) can be deduced by using further bootstrap argument and the parabolic Schauder estimate. Similarly, we can get v\in C^{0}((0, T), C^{2+\nu}(\bar{\Omega})) by using Agmon-Douglas-Nirenberg Theorem [22,23]. From the regularity of u we have w\in C^{2+\nu, 1+\nu/2}(\bar{Q}_{T}) .

    Moreover, the maximal principle entails that 0 < u(x, t)\leq 1 , 0\leq v(x, t)\leq\frac{\beta+\gamma}{\alpha} . It follows from the positivity of u that \tilde{F}(u) = F(u) and because of the uniqueness of solution we infer the existence of the solution to the original problem.

    (ⅲ) Uniqueness. Suppose (u_{1}, v_{1}, w_{1}) and (u_{2}, v_{2}, w_{2}) are two deferent solutions of model (1.3) in \Omega\times [0, T] . Let U = u_{1}-u_{2} , V = v_{1}-v_{2} , W = w_{1}-w_{2} for t\in (0, T) . Then

    \begin{equation} \begin{aligned} &\frac{1}{2}\frac{d}{dt}\int_{\Omega}U^{2}dx+\int_{\Omega}|\nabla U|^{2}dx\\ \leq& \chi\int_{\Omega}|u_{1}(1-u_{1})-u_{2}(1-u_{2})|\nabla v_{1}||\nabla U|+u_{2}(1-u_{2})|\nabla V||\nabla U| dx\\ &+\int_{\Omega}|u_{1}(1-u_{1})-u_{2}(1-u_{2})||U|dx \\ \leq & \chi\int_{\Omega}|U||\nabla v_{1}||\nabla U|+\frac{1}{4}|\nabla V||\nabla U| dx +\int_{\Omega}|U|^{2}dx \\ \leq &\int_{\Omega}|\nabla U|^{2}dx+\frac{\chi^{2}}{32}\int_{\Omega}|\nabla V|^{2}dx+ \frac{\chi^{2} K^{2}+2}{2}\int_{\Omega}|U|^{2}dx, \end{aligned} \end{equation} (2.9)

    where we have used that |\nabla v_{1}|\leq K results from \nabla v_{1}\in C^{0}([0, T], C^{0}(\bar{\Omega})).

    Similarly, by Young inequality and w_{0}\leq w_{1}\leq 1 , we can estimate

    \begin{equation} \int_{\Omega}|\nabla V|^{2}dx+\frac{\alpha}{2}\int_{\Omega}| V|^{2}dx\leq\frac{\beta^{2}}{\alpha} \int_{\Omega}|W|^{2}dx+\frac{\gamma^{2}}{\alpha} \int_{\Omega}|U|^{2}dx, \end{equation} (2.10)

    and

    \begin{equation} \frac{d}{dt}\int_{\Omega}W^{2}dx\leq \delta\int_{\Omega}|U|^{2}+|W|^{2}dx. \end{equation} (2.11)

    Finally, adding to the inequalities (2.9)–(2.11) yields

    \frac{d}{dt}\left(\int_{\Omega}U^{2}dx+\int_{\Omega}W^{2}dx\right)\leq C\left(\int_{\Omega}U^{2}dx+\int_{\Omega}W^{2}dx\right)\; \mathrm{for\; all}\; t \in (0, T).

    The results U\equiv 0 , W\equiv0 in \Omega\times(0, T) are obtained by Gronwall's lemma. From the inequality (2.10), we have V\equiv 0 . Hence (u_{1}, v_{1}, w_{1}) = (u_{2}, v_{2}, w_{2}) in \Omega\times(0, T) .

    (ⅳ) Uniform estimates. We use the Agmon-Douglas-Nirenberg Theorem [22,23] for the second equation of the model (1.3) to get

    \begin{equation} \|v\|_{C^{0}([t, t+1], W_{p}^{2}(\Omega))}\leq C\left(\|u\|_{L^{p}(\Omega \times [t, t+1])}+\|w\|_{L^{p}(\Omega \times [t, t+1])}\right) \leq C_{2} \end{equation} (2.12)

    for all t\geq 1 and C_{2} is independent of t . From the embedded relationship W_{p}^{1}(\Omega)\hookrightarrow C^{0}({\bar{\Omega}}), \; p > n , the parabolic L^{p} -estimate and the estimation (2.12), we have

    \|u\|_{W_{p}^{2, 1}(\Omega\times[t, t+1])}\leq C_{3}

    for all t\geq 1 . The estimate \|u\|_{C^{\nu, \frac{\nu}{2}}(\bar{\Omega}\times [t, t+1])}\leq C_{4} for all t\geq 1 obtained by the embedded relationship (2.8). We can immediately compute \|w\|_{C^{\nu, 1+\frac{\nu}{2}}(\bar{\Omega}\times [t, t+1])}\leq C_{5} for all t\geq 1 according to the regularity of u and the specific expression of w . Further, bootstrapping argument leads to \|v\|_{C^{0}([t, t+1], C^{2+\nu}(\bar{\Omega}))}\leq C_{6} and \|u\|_{C^{2+\nu, 1+\frac{\nu}{2}}(\bar{\Omega}\times [t, t+1])}\leq C_{7} for all t\geq 1 . Thus the uniform estimation (1.7) is proved.

    Remark 2.1. Assume the initial data 0 < u_{0}(x)\leq 1 and w_{0}(x) = 0 . Then the BCS model (1.3) has a unique classical solution.

    In this section we investigate the global asymptotic stability of the unique positive equilibrium point (1, \frac{\beta+\gamma}{\alpha}, 1) to model (1.3). To this end, we first introduce following auxiliary problem

    \begin{equation} \left\{\begin{array}{ll} u_{\epsilon t} = \Delta u_{\epsilon}-\nabla\cdot(u_{\epsilon}(1-u_{\epsilon})\nabla v_{\epsilon})+u_{\epsilon}(1-u_{\epsilon}), & x\in\Omega, \; t \gt 0, \\ -\Delta v_{\epsilon}+\alpha v_{\epsilon} = \beta w_{\epsilon}+\gamma u_{\epsilon}, &x\in\Omega, \; t \gt 0, \\ w_{\epsilon t} = \delta\frac{u_{\epsilon}^{2}+\epsilon}{1+u_{\epsilon}} (1-w_{\epsilon}), &x\in\Omega, \; t \gt 0, \\ \partial_{\eta}u_{\epsilon} = \partial_{\eta}v_{\epsilon} = 0, &x\in\partial\Omega, \; t \gt 0, \\ u_{\epsilon}(x, 0) = u_{0}(x), \; w_{\epsilon}(x, 0) = w_{0}(x), &x\in\Omega. \end{array} \right. \end{equation} (3.1)

    By a similar proof of Theorem 1.1, we get that the problem (3.1) has a unique global classical solution (u_{\epsilon}, v_{\epsilon}, w_{\epsilon}) , and there exist a \nu\in(0, 1) and M_{1} > 0 which is independent of \epsilon such that

    \begin{equation} \|u_{\epsilon}\|_{C^{2+\nu, 1+\nu/2}(\bar{\Omega}\times[1, \infty))}+\|v_{\epsilon}\|_{C^{2+\nu, 1+\nu/2}(\bar{\Omega}\times[1, \infty))} +\|w_{\epsilon}\|_{C^{\nu, 1+\nu/2}(\bar{\Omega}\times[1, \infty))}\leq M_{1}. \end{equation} (3.2)

    Then, motivated by some ideas from [25,26], we construct a Lyapunov function to study the uniform convergence of homogeneous steady state for the problem (3.1).

    Let us give following lemma which is used in the proof of Lemma 3.2.

    Lemma 3.1. Suppose that a nonnegative function f on (1, \infty) is uniformly continuous and \int_{1}^{\infty}f(t)dt < \; \infty . Then f(t)\rightarrow 0 as t\rightarrow \infty.

    Lemma 3.2. Assume that the condition (1.8) is satisfied. Then

    \begin{equation} \|u_{\epsilon}(\cdot, t)-1\|_{L^{2}(\Omega)}+ \|v_{\epsilon}(\cdot, t)-v^{\ast}\|_{L^{2}(\Omega)} +\|w_{\epsilon}(\cdot, t)-1\|_{L^{2}(\Omega)}\rightarrow 0, \; \; t\rightarrow \infty, \end{equation} (3.3)

    where v^{\ast} = \frac{\beta+\gamma}{\alpha} .

    Proof We construct a positive function

    E(t): = \int_{\Omega}(u_{\varepsilon}-1-\ln u_{\epsilon}) +\frac{1}{2\delta\epsilon}\int_{\Omega}(w_{\epsilon}-1)^{2}, \; \; t \gt 0.

    From the problem (3.1) and Young's inequality, we can compute

    \begin{equation} \frac{d}{dt}E(t)\leq {\frac{\chi^{2}}{4}}\int_{\Omega}|\nabla v_{\epsilon}|^{2}dx-\int_{\Omega}(u_{\epsilon}-1)^{2}dx-\int_{\Omega}(w_{\epsilon}-1)^{2}dx, \; \; t \gt 0. \end{equation} (3.4)

    We multiply the second equations in system (3.1) by v_{\epsilon}-v^{\ast} , integrate by parts over \Omega and use Young's inequality to obtain

    \begin{equation} \int_{\Omega}|\nabla v_{\epsilon}|^{2}dx\leq\frac{\gamma^{2}}{2\alpha}\int_{\Omega}(u_{\epsilon}-1)^{2}dx +\frac{\beta^{2}}{2\alpha}\int_{\Omega}(w_{\epsilon}-1)^{2}dx, \; \; t \gt 0, \end{equation} (3.5)

    and

    \begin{equation} \int_{\Omega}(v_{\epsilon}-v^{\ast})^{2}dx\leq\frac{2\gamma^{2}}{\alpha^{2}}\int_{\Omega}(u_{\epsilon}-1)^{2}dx+\frac{2 \beta^{2}}{\alpha^{2}}\int_{\Omega}(w_{\epsilon}-1)^{2}dx, \; \; t \gt 0. \end{equation} (3.6)

    Substituting inequality (3.5) into inequality (3.4) to get

    \begin{equation} \nonumber \frac{d}{dt}E(t)\leq -C_{8}\left(\int_{\Omega}(u_{\epsilon}-1)^{2}dx+\int_{\Omega}(w_{\epsilon}-1)^{2}dx\right), \; \; t \gt 0, \end{equation}

    where C_{8} = \min\left\{1-\frac{\chi^{2}\beta^{2}}{8\alpha}, 1-\frac{\chi^{2}\gamma^{2}}{8\alpha}\right\} > 0.

    Let f(t): = \int_{\Omega}(u_{\epsilon}-1)^{2}+(w_{\epsilon}-1)^{2}dx . Then

    \int_{1}^{\infty}f(t)dt\leq \frac{E(1)}{C_{8}} \lt \infty, \; \; t \gt 1.

    It follows from the uniform estimation (3.2) and the Arzela-Ascoli theorem that f(t) is uniformly continuous in (1, \infty) . Applying Lemma 3.1, we have

    \begin{equation} \int_{\Omega}(u_{\epsilon}(\cdot, t)-1)^{2}+ (w_{\epsilon}(\cdot, t)-1)^{2}dx\rightarrow 0, \; \; t\rightarrow \infty. \end{equation} (3.7)

    Combining inequality (3.6) and the limit (3.7) to obtain

    \int_{\Omega}(v_{\epsilon}(\cdot, t)-v^{\ast})^{2}dx \rightarrow 0, \; \; t\rightarrow \infty.

    Proof of Theorem 1.2 As we all known, each bounded sequence in C^{2+\nu, 1+\frac{\nu}{2}}(\bar{\Omega}\times[1, \infty)) is precompact in C^{2, 1}(\bar{\Omega}\times[1, \infty)) . Hence there exists some subsequence \{u_{\epsilon_{n}}\}_{n = 1}^{\infty} satisfying \epsilon_{n}\rightarrow0 as n\rightarrow \infty such that

    \lim\limits_{n\rightarrow \infty}\|u_{\epsilon_{n}}-u_{\ast}\|_{C^{2, 1}(\bar{\Omega}\times[1, \infty))} = 0.

    Similarly, we can get

    \lim\limits_{n\rightarrow \infty}\|v_{\epsilon_{n}}-v_{\ast}\|_{C^{2}(\bar{\Omega})} = 0,

    and

    \lim\limits_{n\rightarrow \infty}\|w_{\epsilon_{n}}-w_{\ast}\|_{C^{0, 1}(\bar{\Omega}\times[1, \infty))} = 0.

    Combining above limiting relations yields that (u_{\ast}, v_{\ast}, w_{\ast}) satisfies model (1.3). The conclusion (u_{\ast}, v_{\ast}, w_{\ast}) = (u, v, w) is directly attributed to the uniqueness of the classical solution of the model (1.3). Furthermore, according to the conclusion, the strong convergence (3.3) and Diagonal line method, we can deduce

    \begin{equation} \|u(\cdot, t)-1\|_{L^{2}(\Omega)}+ \|v(\cdot, t)-v^{\ast}\|_{L^{2}(\Omega)} +\|w(\cdot, t)-1\|_{L^{2}(\Omega)}\rightarrow 0, \; \; t\rightarrow \infty. \end{equation} (3.8)

    By applying Gagliardo-Nirenberg inequality

    \begin{equation} \|z\|_{L^{\infty}}\leq C\|z\|_{L^{2}(\Omega)}^{2/(n+2)}\|z\|_{W^{1, \infty}(\Omega)}^{n/(n+2)}, \; \; z\in W^{1, \infty}(\Omega), \end{equation} (3.9)

    comparison principle of ODE and the convergence (3.8), the uniform convergence (1.9) is obtained immediately.

    Since \lim_{t\rightarrow \infty}\|u(\cdot, t)-1\|_{L^{\infty}(\Omega)} = 0 , so there exists a t_{1} > 0 such that

    \begin{equation} u(x, t)\geq \frac{1}{2}\; \; \mathrm{for\; all}\; \; x\in \Omega, \; \; t \gt t_{1}. \end{equation} (3.10)

    Using the explicit representation formula of w

    w(x, t) = (w_{0}(x)-1)e^{-\delta\int_{0}^{t}F(u)uds}+1, \; \; x\in\Omega, \; t \gt 0

    and the inequality (3.10), we have

    \begin{equation} \|w(\cdot, t)-1\|_{L^{\infty}(\Omega)}\leq e^{-\frac{\delta}{6}(t-t_{1})}, \; \; t \gt t_{1}. \end{equation} (3.11)

    Multiply the first two equations in model (1.3) by u-1 and v-v^{\ast} , respectively, integrate over \Omega and apply Cauchy's inequality, Young's inequality and the inequality (3.10), to find

    \begin{equation} \frac{d}{dt}\int_{\Omega}(u-1)^{2}dx\leq \frac{\chi^{2}}{32}\int_{\Omega}|\nabla v|^{2}dx-\int_{\Omega}(u-1)^{2}dx, \; \; t \gt t_{1}. \end{equation} (3.12)
    \begin{equation} \int_{\Omega}|\nabla v|^{2}dx+\frac{\alpha}{2}\int_{\Omega}(v-v^{\ast})^{2}dx\leq \frac{\beta^{2}}{\alpha}\int_{\Omega}(w-1)^{2}dx +\frac{\gamma^{2}}{\alpha}\int_{\Omega}(u-1)^{2}dx, \; \; t \gt 0. \end{equation} (3.13)

    Combining the estimations (3.11)–(3.13) leads us to the estimate

    \begin{equation} \nonumber \frac{d}{dt}\int_{\Omega}(u-1)^{2}dx\leq \left(\frac{\chi^{2}\gamma^{2}}{32\alpha}-1\right)\int_{\Omega}(u-1)^{2}dx +\frac{\chi^{2}\beta^{2}}{32\alpha}e^{-{\frac{\delta}{3}(t-t_{1})}}, \; \; t \gt t_{1}. \end{equation}

    Let y(t) = \int_{\Omega}(u-1)^{2}dx . Then

    y'(t)\leq \left(\frac{\chi^{2}\gamma^{2}}{32\alpha}-1\right)y(t) +\frac{\chi^{2}\beta^{2}}{32\alpha}e^{-{\frac{\delta}{3}(t-t_{1})}}, \; \; t \gt t_{1}.

    From comparison principle of ODE, we get

    y(t)\leq \left(y(t_{1})-\frac{3\chi^{2}\beta^{2}}{32\alpha(3-\delta)-\chi^{2}\gamma^{2}}\right) e^{-\left(1-\frac{\chi^{2}\gamma^{2}}{32\alpha}\right)(t-t_{1})} +\frac{3\chi^{2}\beta^{2}}{32\alpha(3-\delta)-\chi^{2}\gamma^{2}}e^{-\frac{\delta}{3}(t-t_{1})}, \; \; t \gt t_{1}.

    This yields

    \begin{equation} \int_{\Omega}(u-1)^{2}dx\leq C_{9} e^{-\lambda_{2} (t-t_{1})}, \; \; t \gt t_{1}, \end{equation} (3.14)

    where \lambda_{2} = \min\{1-\frac{\chi^{2}\gamma^{2}}{32\alpha}, \frac{\delta}{3}\} and C_{9} = \max\left\{|\Omega|-\frac{3\chi^{2}\beta^{2}}{32\alpha(3-\delta)-\chi^{2}\gamma^{2}}, \frac{3\chi^{2}\beta^{2}}{32\alpha(3-\delta)-\chi^{2}\gamma^{2}}\right\} .

    From the inequalities (3.11), (3.13) and (3.14), we derive

    \begin{equation} \int_{\Omega}\left(v-\frac{\beta+\gamma}{\alpha}\right)^{2}dx\leq C_{10}e^{-\lambda_{2}(t-t_{1})}, \; \; t \gt t_{1}, \end{equation} (3.15)

    where C_{10} = \max\left\{\frac{2\gamma^{2}}{\alpha^{2}}C_{9}, \frac{2\beta^{2}}{\alpha^{2}}\right\} . By employing the uniform estimation (1.7), the inequalities (3.9), (3.14) and (3.15), the exponential decay estimation (1.10) can be obtained.

    The proof is complete.

    In this paper, we mainly study the uniform boundedness of classical solutions and exponential asymptotic stability of the unique positive equilibrium point to the chemotactic cellular model (1.3) for Baló's concentric sclerosis (BCS). For model (1.1), by numerical simulation, Calveza and Khonsarib in [7] shown that demyelination patterns of concentric rings will occur with increasing of chemotactic sensitivity. By the Theorem 1.1 we know that systems (1.1) and (1.2) are {uniformly} bounded and dissipative. By the Theorem 1.2 we also find that the constant equilibrium point of model (1.1) is exponentially asymptotically stable if

    \tilde{\chi} \lt \frac{2}{\bar{w}\tilde{\beta}} \sqrt{\frac{2D\mu\tilde{\alpha}\tilde{\epsilon}}{\bar{u}}},

    and the constant equilibrium point of the model (1.2) is exponentially asymptotically stable if

    \tilde{\chi} \lt 2\sqrt{\frac{2D\mu\tilde{\alpha}\tilde{\epsilon}}{\bar{u}}}\min \left\{\frac{1}{\bar{w}\tilde{\beta}}, \frac{1}{\bar{u}\tilde{\gamma}}\right\}.

    According to a pathological viewpoint of BCS, the above stability results mean that if chemoattractive effect is weak, then the destroyed oligodendrocytes form a homogeneous plaque.

    The authors would like to thank the editors and the anonymous referees for their constructive comments. This research was supported by the National Natural Science Foundation of China (Nos. 11761063, 11661051).

    We have no conflict of interest in this paper.



    [1] S. Pernas, S. M. Tolaney, E. P. Winer, S. Goel, CDK4/6 inhibition in breast cancer: current practice and future directions, Ther. Adv. Med. Oncol., 10 (2018), 1758835918786451.
    [2] H. Xu, S. Yu, Q. Liu, X. Yuan, S. Mani, R. G. Pestell, et al., Recent advances of highly selective CDK4/6 inhibitors in breast cancer, J. Hematol. Oncol., 10 (2017), 97. doi: 10.1186/s13045-017-0467-2
    [3] S. F. Dowdy, M. Kaulich, Abstract 1304: Cyclin D:Cdk4/6 activates RB by mono-phosphorylation during early G1 phase, Cancer Res., 74 (2014), 1304–1304.
    [4] A. N. Omstead, D. Matsui, J. E. Kosovec, S. A. Martin, B. A. Jobe, Antitumor efficacy of CDK 4/6 dual inhibitor, abemaciclib, in an esophageal adenocarcinoma model, J. Clin. Oncol., 35 (2017), e15598–e15598.
    [5] M. Tutone, A. M. Almerico, Recent advances on CDK inhibitors: An insight by means of in silico methods, Eur. J. Med. Chem., 142 (2017), 300–315. doi: 10.1016/j.ejmech.2017.07.067
    [6] S. Müller, A. Chaikuad, N. S. Gray, S. Knapp, The ins and outs of selective kinase inhibitor development, Nat. Chem. Biol., 11 (2015), 818–821. doi: 10.1038/nchembio.1938
    [7] P. Ayaz, D. Andres, D. A. Kwiatkowski, C. C. Kolbe, P. Lienau, G. Siemeister, et al., Conformational adaption may explain the slow dissociation kinetics of roniciclib (BAY 1000394), a yype I CDK inhibitor with kinetic selectivity for CDK2 and CDK9, ACS Chem. Biol., (2016), acschembio.6b00074.
    [8] T. Dale, P. A. Clarke, C. Esdar, D. Waalboer, O. Adeniji-Popoola, M. J. Ortiz-Ruiz, et al., A selective chemical probe for exploring the role of CDK8 and CDK19 in human disease, Nat. Chem. Biol., 2015.
    [9] M. Schreuer, V. Kruse, Y. Jansen, B. Neyns, COMBI-rechallenge: a phase II clinical trial on dabrafenib plus trametinib in BRAFV600-mutant melanoma patients who previously experienced progression on BRAF(+MEK)-inhibition, Ann. Oncolo., 27 (2016).
    [10] R. B. Corcoran, G. S. Falchook, J. R. Infante, O. Hamid, W. A. Messersmith, E. L. Kwak, et al., BRAF V600 mutant colorectal cancer (CRC) expansion cohort from the phase I/II clinical trial of BRAF inhibitor dabrafenib (GSK2118436) plus MEK inhibitor trametinib (GSK1120212), J. Clin. Oncol., 2012.
    [11] T. Wang, Z. Yang, Y. Zhang, W. Yan, F. Wang, L. He, et al., Discovery of novel CDK8 inhibitors using multiple crystal structures in docking-based virtual screening, Eur. J. Med. Chem., 129 (2017), 275–286. doi: 10.1016/j.ejmech.2017.02.020
    [12] S. E. Dixon-Clarke, S. N. Shehata, T. Krojer, T. D. Sharpe, F. Von Delft, K. Sakamoto, et al., Structure and inhibitor specificity of the PCTAIRE-family kinase CDK16, Biochem. J., 474 (2017), 699–713. doi: 10.1042/BCJ20160941
    [13] N. Canela, M. Orzaez, R. Fucho, F. Mateo, R. Gutierrez, A. Pineda-Lucena, et al., Identification of an hexapeptide that binds to a surface pocket in cyclin A and inhibits the catalytic activity of the complex cyclin-dependent kinase 2-cyclin A, J. Biol. Chem., 281 (2006), 35942–35953. doi: 10.1074/jbc.M603511200
    [14] Orzáez, Guevara, Sancho, Pérez-Payá, Intrinsic caspase-8 activation mediates sensitization of erlotinib-resistant tumor cells to erlotinib/cell-cycle inhibitors combination treatment, Cell Death Dis., 2012.
    [15] R. S. Finn, A. Aleshin, D. J. Slamon, Targeting the cyclin-dependent kinases (CDK) 4/6 in estrogen receptor-positive breast cancers, Breast Cancer Res.: BCR, 18 (2016), 17. doi: 10.1186/s13058-015-0661-5
    [16] T. Otto, P. Sicinski, Cell cycle proteins as promising targets in cancer therapy, Nat. Rev. Cancer, 17 (2017), 93–115. doi: 10.1038/nrc.2016.138
    [17] L. Spring, A. Bardia, S. Modi, Targeting the cyclin D-cyclin-dependent kinase (CDK) 4/6-retinoblastoma pathway with selective CDK 4/6 inhibitors in hormone receptor-positive breast cancer: rationale, current status, and future directions, Discovery Med., 21 (2016), 65.
    [18] M. W. Landis, B. S. Pawlyk, T. Li, P. Sicinski, P. W. Hinds, Cyclin D1-dependent kinase activity in murine development and mammary tumorigenesis, Cancer Cell, 9 (2006), 13–22. doi: 10.1016/j.ccr.2005.12.019
    [19] B. Laderian, T. Fojo, CDK4/6 inhibition as a therapeutic strategy in breast cancer: palbociclib, ribociclib, and abemaciclib, Semin. Oncol., (2018), S0093775418300812.
    [20] S. Parylo, A. Vennepureddy, V. Dhar, P. Patibandla, A. Sokoloff, Role of cyclin-dependent kinase 4/6 inhibitors in the current and future eras of cancer treatment, J. Oncol. Pharm. Pract., (2018), 107815521877090.
    [21] A. Patnaik, L. S. Rosen, S. M. Tolaney, A. W. Tolcher, J. W. Goldman, L. Gandhi, et al., Efficacy and safety of abemaciclib, an inhibitor of CDK4 and CDK6, for patients with breast cancer, non–small cell lung cancer, and other solid tumors, Cancer Discovery, (2016), 740–753.
    [22] H. Wang, K. Wang, Z. Guan, Y. Jian, Y. Jia, F. Kashanchi, et al., Computational study of non-catalytic T-loop pocket on CDK proteins for drug development, Chin. Phys. B, 2017.
    [23] H. W. Wang, Z. Y. Guan, J. D. Qiu, Y. Jia, C. Zeng, Y. J. Zhao, Novel method to identify group-specific non-catalytic pockets of human kinome for drug design, RSC Adv., 4 (2020).
    [24] Y. Zhao, H. Chen, C. Du, Y. Jian, H. Li, Y. Xiao, et al., Design of tat-activated CDK9 inhibitor, Int. J. Peptide Res. Therapeutics, 25 (2018), 807–817.
    [25] A. M. Almerico, M. Tutone, A. Lauria, 3D-QSAR pharmacophore modeling and in silico screening of new Bcl-xl inhibitors, Eur. J. Med. Chem., 45 (2010), 4774–4782. doi: 10.1016/j.ejmech.2010.07.042
    [26] A. M. Almerico, M. Tutone, A. Lauria, Receptor-guided 3D-QSAR approach for the discovery of c-kit tyrosine kinase inhibitors, J. Mol. Model., 18 (2012), 2885–2895. doi: 10.1007/s00894-011-1304-0
    [27] Z. Shentu, M. A. Hasan, C. Bystroff, M. J. Zaki, Context shapes: Efficient complementary shape matching for protein-protein docking, Proteins-Struct. Funct. Bioinformatics, 70 (2010), 1056–1073.
    [28] D. W. Ritchie, Evaluation of protein docking predictions using Hex 3.1 in CAPRI rounds 1 and 2, Proteins: Struct., Funct., Bioinformatics, 2003.
    [29] K. Wiehe, B. Pierce, J. Mintseris, W. W. Tong, R. Anderson, R. Chen, et al., ZDOCK and RDOCK performance in CAPRI rounds 3, 4, and 5, Proteins-Struct. Funct. Bioinformatics, 60 (2005), 207–213. doi: 10.1002/prot.20559
    [30] A. Caflisch, P. Niederer, M. Anliker, Monte Carlo docking of oligopeptides to proteins, Proteins-Struct. Funct. Bioinformatics, 13 (2010), 223–230.
    [31] T. N. Hart, R. J. Read, A multiple-start Monte Carlo docking method, J. Mol. Graphics, 13 (2010), 206–222.
    [32] P. Reigan, W. Guo, D. Siegel, D. Ross, Molecular docking studies investigating the interaction of a series of benzoquinone ansamycin Hsp90 inhibitors with NAD(P)H: quinone oxidoreductase 1 (NQO1), Cancer Res., 66 (2006), 457–457.
    [33] C. M. Venkatachalam, X. Jiang, T. Oldfield, M. Waldman, LigandFit: a novel method for the shape-directed rapid docking of ligands to protein active sites, J. Mol. Graphics Model., 21 (2003), 289–307. doi: 10.1016/S1093-3263(02)00164-X
    [34] L. Kang, H. L. Li, H. L. Jiang, X. C. Wang, An improved adaptive genetic algorithm for protein–ligand docking, J. Comput. Aided Mol. Des., 23 (2009), 1–12.
    [35] F. Sterberg, G. M. Morris, M. F. Sanner, A. J. Olson, D. S. Goodsell, Automated docking to multiple target structures: Incorporation of protein mobility and structural water heterogeneity in AutoDock, Protns Struct. Funct. Bioinformatics, 46 (2002), 34–40. doi: 10.1002/prot.10028
    [36] G. Jones, P. Willett, R. C. Glen, A. R. Leach, R. Taylor, Development and validation of a genetic algorithm for flexible docking, J. Mol. Biol., 267 (1997), 727–748. doi: 10.1006/jmbi.1996.0897
    [37] H. Jing, X. Zhou, X. Dong, J. Cao, H. Zhu, J. Lou, et al., Abrogation of Akt signaling by Isobavachalcone contributes to its anti-proliferative effects towards human cancer cells, Cancer Lett., 294 (2010), 167–177. doi: 10.1016/j.canlet.2010.01.035
    [38] H. Li, C. Li, C. Gui, X. Luo, K. Chen, J. Shen, et al., GAsDock: a new approach for rapid flexible docking based on an improved multi-population genetic algorithm, Bioorg. Med. Chem. Lett., 14 (2004), 4671–4676. doi: 10.1016/j.bmcl.2004.06.091
    [39] G. Culletta, A. M. Almerico, M. Tutone, Comparing molecular dynamics-derived pharmacophore models with docking: a study on CDK-2 inhibitors, Chem. Data Collect., 2020.
    [40] P. N. Sekhar, Software for molecular docking: a review, Biophys. Rev., 9 (2016), 91–102.
    [41] Z. Bikadi, E. Hazai, Application of the PM6 semi-empirical method to modeling proteins enhances docking accuracy of AutoDock, J. Cheminformatics, 1 (2009), 1–16. doi: 10.1186/1758-2946-1-1
    [42] A. Grosdidier, V. Zoete, O. Michielin, SwissDock, a protein-small molecule docking web service based on EADock DSS, Nucleic Acids Res., 39 (2011), W270–W277.
    [43] G. M. Morris, R. Huey, W. Lindstrom, M. F. Sanner, A. J. Olson, AutoDock4 and AutoDockTools4: Automated docking with selective receptor flexibility, J. Comput. Chem., 30 (2010), 2785–2791.
    [44] A. Grosdidier, V. Zoete, O. Michielin, Fast docking using the CHARMM force field with EADock DSS, J. Comput. Chem., 32 (2011), 2149–2159. doi: 10.1002/jcc.21797
    [45] R. Huey, G. M. Morris, A. J. Olson, D. S. Goodsell, A semi-empirical free energy force field with charge-based desolvation, J. Comput. Chem., 28 (2010), 1145–1152.
    [46] S. J. Weiner, P. A. Kollman, D. A. Case, U. C. Singh, C. Ghio, G. Alagona, et al., A new force field for molecular mechanical simulation of nucleic acids and proteins, J. Am. Chem. Soc., 106 (1984), 765–784. doi: 10.1021/ja00315a051
    [47] P. J. Goodford, A computational procedure for determining energetically favorable binding sites on biologically important macromolecules, J. Med. Chem., 28 (1985), 849–857. doi: 10.1021/jm00145a002
    [48] E. L. Mehler, T. Solmajer, Electrostatic effects in proteins: comparison of dielectric and charge models, Protn. Eng., (1991), 903–910.
    [49] G. M. Verkhivker, D. Bouzida, D. K. Gehlhaar, P. A. Rejto, S. Arthurs, A. B. Colson, et al., Deciphering common failures in molecular docking of ligand-protein complexes, J. Comput.-Aided Mol. Des., 14 (2000), 731–751. doi: 10.1023/A:1008158231558
    [50] B. R. Brooks, R. E. Bruccoleri, B. D. Olafson, D. J. States, M. Karplus, CHARMM: A program for macromolecular energy, minimization, and dynamics calculations, J. Comput. Chem., 4 (2010), 187–217.
    [51] A. Grosdidier, V. Zoete, O. Michielin, Fast docking using the CHARMM force field with EADock DSS, J. Comput. Chem., 32 (2011), 2149–2159. doi: 10.1002/jcc.21797
    [52] H. J. C. Berendsen, J. R. Grigera, T. P. Straatsma, The missing term in effective pair potentials, J. Phys. Chem., 91 (1987), 6269–6271. doi: 10.1021/j100308a038
    [53] B. R. R. Brooks, C. L. B. Brooks, A. D. Mackerell, L. Nilsson, M. J. Karplus, CHARMM: the biomolecular simulation program, J. Comput. Chem., 30 (2009), 1545. doi: 10.1002/jcc.21287
    [54] J. A. Hartigan, M. A. Wong, A K-Means clustering algorithm, Appl. Stats., 28 (1979).
    [55] A. K. Jain, Data clustering: 50 years beyond K-means, Pattern Recognit. Lett., 31 (2010), 651–666. doi: 10.1016/j.patrec.2009.09.011
    [56] A. B. Chorin, G. Masrati, A. Kessel, A. Narunsky, ConSurf‐DB: An accessible repository for the evolutionary conservation patterns of the majority of PDB proteins, Protein Sci., 29 (2020).
    [57] O. Goldenberg, E. Erez, G. Nimrod, N. Ben-Tal, The ConSurf-DB: pre-calculated evolutionary conservation profiles of protein structures, Nucleic Acids Res., 37 (2009), D323–D327. doi: 10.1093/nar/gkn822
    [58] M. Jaina, R. D. Finn, S. R. Eddy, B. Alex, P. Marco, Challenges in homology search: HMMER3 and convergent evolution of coiled-coil regions, Nucleic Acids Res., 41 (2013), e121–e121. doi: 10.1093/nar/gkt263
    [59] K. Kazutaka, D. M. Standley, MAFFT multiple sequence alignment software version 7: improvements in performance and usability, Mol. Biol. Evol., 30 (2013), 772–780. doi: 10.1093/molbev/mst010
    [60] P. Tal, R. E. Bell, M. Itay, G. Fabian, B. T. Nir, Rate4Site: an algorithmic tool for the identification of functional regions in proteins by surface mapping of evolutionary determinants within their homologues, Bioinformatics, (2002), S71.
    [61] P. Chen, N. V. Lee, W. Hu, M. Xu, B. W. Murray, Spectrum and degree of CDK drug interactions predicts clinical performance, Mol. Cancer Therapeutics, 15 (2016), 2273. doi: 10.1158/1535-7163.MCT-16-0300
    [62] N. M. O'Boyle, M. Banck, C. A. James, C. Morley, G. R. Hutchison, Open babel: an open chemical toolbox, J. Cheminformatics, 3 (2011), 33. doi: 10.1186/1758-2946-3-33
    [63] H. Wang, J. Qiu, H. Liu, Y. Xu, Y. Jia, Y. Zhao, HKPocket: human kinase pocket database for drug design, BMC Bioinformatics, 20 (2019), 617. doi: 10.1186/s12859-019-3254-y
    [64] K. Wang, Y. Jian, H. Wang, C. Zeng, Y. Zhao, RBind: computational network method to predict RNA binding sites, Bioinformatics, 34 (2018).
    [65] Y. Jian, X. Wang, J. Qiu, H. Wang, Z. Liu, Y. Zhao, C. Zeng, DIRECT: RNA contact predictions by integrating structural patterns, BMC Bioinformatics, 20 (2019), 497. doi: 10.1186/s12859-019-3099-4
    [66] H. Wang, Y. Zhao, RBinds: a user-friendly server for RNA binding site prediction, Comput. Struct. Biotechnol. J., 18 (2020), 3762–3765. doi: 10.1016/j.csbj.2020.10.043
    [67] H. Wang, Y. Zhao, Methods and applications of RNA contact prediction, Chin. Phys. B, 29 (2020), 108708. doi: 10.1088/1674-1056/abb7f3
    [68] K. Rascon, G. Flajc, C. De Angelis, X. Liu, M. V. Trivedi, E. Ekinci, Ribociclib in HR+/HER2- advanced or metastatic breast cancer patients, Ann. Pharmacotherapy, 2019.
    [69] R. J. Cersosimo, Cyclin-dependent kinase 4/6 inhibitors for the management of advanced or metastatic breast cancer in women, (2019), 1183–1202.
    [70] A. F. D. Groot, C. J. Kuijpers, J. R. Kroep, CDK4/6 inhibition in early and metastatic breast cancer: A review, Cancer Treatment Rev., 60 (2017), 130–138. doi: 10.1016/j.ctrv.2017.09.003
    [71] M. Poratti, G. Marzaro, Third-generation CDK inhibitors: A review on the synthesis and binding modes of Palbociclib, Ribociclib and Abemaciclib, Eur. J. Med. Chem., 172 (2019), 143–153. doi: 10.1016/j.ejmech.2019.03.064
  • mbe-18-01-025-supplementary.pdf
  • This article has been cited by:

    1. Lu Xu, Chunlai Mu, Qiao Xin, Global boundedness and asymptotic behavior of solutions for a quasilinear chemotaxis model of multiple sclerosis with nonlinear signal secretion, 2023, 28, 1531-3492, 1215, 10.3934/dcdsb.2022118
  • Reader Comments
  • © 2021 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(5198) PDF downloads(419) Cited by(2)

Figures and Tables

Figures(7)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog