
Gene regulatory networks (GRNs) play a crucial role in biological processes, with their dynamic behaviors heavily influenced by the spatial organization of genes. In particular, reaction-diffusion mechanisms govern the coupling between adjacent spatial locations in continuous time GRNs. However, traditional models often ignore the spatial coupling and reaction-diffusion properties of these networks, especially in discrete-time settings. In order to solve this problem, a new discrete-time gene regulatory network model is proposed in this paper, which explicitly considers the mutual coupling between adjacent spatial positions. To ensure the passivity of the proposed model, delay-dependent stability criteria are established by constructing appropriate Lyapunov-Krasovskii functions formulated in terms of linear matrix inequalities. To showcase the effectiveness and validity of this approach, a numerical example is presented in this paper. The results reveal that the model accurately captures the spatial coupling and reaction-diffusion nature of gene regulatory networks in discrete time settings.
Citation: Yongwei Yang, Yang Yu, Chunyun Xu, Chengye Zou. Passivity analysis of discrete-time genetic regulatory networks with reaction-diffusion coupling and delay-dependent stability criteria[J]. Electronic Research Archive, 2025, 33(5): 3111-3134. doi: 10.3934/era.2025136
[1] | Bin Wang . Random periodic sequence of globally mean-square exponentially stable discrete-time stochastic genetic regulatory networks with discrete spatial diffusions. Electronic Research Archive, 2023, 31(6): 3097-3122. doi: 10.3934/era.2023157 |
[2] | Minglei Fang, Jinzhi Liu, Wei Wang . Finite-/fixed-time synchronization of leakage and discrete delayed Hopfield neural networks with diffusion effects. Electronic Research Archive, 2023, 31(7): 4088-4101. doi: 10.3934/era.2023208 |
[3] | Yijun Chen, Yaning Xie . A kernel-free boundary integral method for reaction-diffusion equations. Electronic Research Archive, 2025, 33(2): 556-581. doi: 10.3934/era.2025026 |
[4] | Huan Luo . Heterogeneous anti-synchronization of stochastic complex dynamical networks involving uncertain dynamics: an approach of the space-time discretizations. Electronic Research Archive, 2025, 33(2): 613-641. doi: 10.3934/era.2025029 |
[5] | Xiangwen Yin . A review of dynamics analysis of neural networks and applications in creation psychology. Electronic Research Archive, 2023, 31(5): 2595-2625. doi: 10.3934/era.2023132 |
[6] | Meng Zhao . The longtime behavior of the model with nonlocal diffusion and free boundaries in online social networks. Electronic Research Archive, 2020, 28(3): 1143-1160. doi: 10.3934/era.2020063 |
[7] | Peng Gao, Pengyu Chen . Blowup and MLUH stability of time-space fractional reaction-diffusion equations. Electronic Research Archive, 2022, 30(9): 3351-3361. doi: 10.3934/era.2022170 |
[8] | Jianmin Hou, Quansheng Liu, Hongwei Yang, Lixin Wang, Yuanhong Bi . Stability and bifurcation analyses of p53 gene regulatory network with time delay. Electronic Research Archive, 2022, 30(3): 850-873. doi: 10.3934/era.2022045 |
[9] | Yu-Jing Shi, Yan Ma . Finite/fixed-time synchronization for complex networks via quantized adaptive control. Electronic Research Archive, 2021, 29(2): 2047-2061. doi: 10.3934/era.2020104 |
[10] | Xin Du, Quansheng Liu, Yuanhong Bi . Bifurcation analysis of a two–dimensional p53 gene regulatory network without and with time delay. Electronic Research Archive, 2024, 32(1): 293-316. doi: 10.3934/era.2024014 |
Gene regulatory networks (GRNs) play a crucial role in biological processes, with their dynamic behaviors heavily influenced by the spatial organization of genes. In particular, reaction-diffusion mechanisms govern the coupling between adjacent spatial locations in continuous time GRNs. However, traditional models often ignore the spatial coupling and reaction-diffusion properties of these networks, especially in discrete-time settings. In order to solve this problem, a new discrete-time gene regulatory network model is proposed in this paper, which explicitly considers the mutual coupling between adjacent spatial positions. To ensure the passivity of the proposed model, delay-dependent stability criteria are established by constructing appropriate Lyapunov-Krasovskii functions formulated in terms of linear matrix inequalities. To showcase the effectiveness and validity of this approach, a numerical example is presented in this paper. The results reveal that the model accurately captures the spatial coupling and reaction-diffusion nature of gene regulatory networks in discrete time settings.
Genetic regulatory networks, often abbreviated as GRNs, are complex systems that include several components that work together to regulate gene expression. These networks encompass a wide range of interactions, including those between DeoxyriboNucleic Acid (DNA) and proteins (such as transcription factors), RiboNucleic Acid (RNA) and proteins (like ribosomes and RNA-binding proteins), and even interactions among different types of RNA. GRNs play a pivotal role in understanding various biological processes and have garnered significant attention from researchers [1]. Over the last couple of years, advances in molecular-level biology research have highlighted the significance of genetic networks, which are inherently complex and need simplification. GRNs have successfully described biochemical reactions such as gene transcription, translation, and protein diffusion, similar to the achievements in artificial neural network modeling, which have resulted in a plethora of models. Various models have emerged, including the Bayesian model [2], the Boolean model [3,4], and the differential equation model [5,6]. In particular, the differential equation model is good at describing the continuous evolution of a system state over time, and usually can reflect the real dynamics of the system more accurately. However, continuous systems are not suitable for computer processing, so many studies have converted continuous systems into discrete systems [7,8]. In this context, incorporating delayed discrete-time GRNs becomes crucial for simulation and implementation [9]. Numerous excellent results in [10,11] have analyzed exponential stability, made synchronization control, designed guaranteed cost control, and updated estimation methods for discrete-time GRNs. Recently, some notable contributions have further enriched this field. By resorting to the sliding-mode control, Li et al. [12] effectually extended and modified the existing works on the stabilization problem for delayed uncertain semi-Markovian switching complex-valued networks. They proposed sufficient criteria for stochastic stability based on the generalized Dynkin's formula and Lyapunov stability theory. In addition to these works, Narayanan et al. [13] addressed the problem of Mittag-Leffler synchronization of T-S fuzzy fractional-order discrete-time complex-valued molecular models of mRNA and protein in regulatory mechanisms with two kinds of regulation functions. Wei et al. [14] addressed the dissipation synchronization issue of semi-Markovian jumping delayed neural networks under random deception attacks and explained the specific influence of deception attacks on event trigger time for the first time. Pandiselvi et al. [15] studied the approximation concern for the discrete-time stochastic GRNs with the leakage delays, distributed delays, and probabilistic measurement delays into the problem and modeled the robust H∞ state estimator for a class of discrete-time stochastic GRNs. Zhao and Wu [16] contributed to the field by establishing new fixed/prescribed-time stability criteria for stochastic systems with time delay and designed delay-independent control mechanisms that are applicable to systems with unknown delays. These studies highlight the importance of addressing key issues in GRN control and synchronization. The study of discrete-time GRN with spatial coupling has gained significant attention due to its importance in understanding the spatial and temporal dynamics of gene expression. In recent years, discrete-time GRNs with spatial coupling have emerged as a crucial area of research in systems biology as they integrate temporal and spatial dynamics to provide a more comprehensive understanding of gene regulation processes. Unlike existing literature that limits its focus to discrete-time networks, Zhang and Li [17] explored the double effects of both discrete time and discrete spatial diffusions in switching complex dynamical networks. Their work is ground breaking in its consideration of discrete spatial diffusions, offering a robust theoretical and practical foundation for future research in this area. Liang and Wang [18] studied the passivity for coupled reaction-diffusion neural networks with multiple state couplings or spatial diffusion couplings by employing the proportional-derivative control method. This research provides valuable insights into the control and synchronization of networks with reaction-diffusion terms. Discrete-time GRNs with spatial coupling have also been applied to study gene regulatory networks in disease and aging. Avila et al. [19] delineated different types of GRNs and provided their biological interpretation, highlighting the potential of discrete-time GRNs with spatial coupling to uncover the underlying mechanisms of complex diseases and aging processes.
Passivity, a mathematical principle describing a system or function's ability to absorb or retain energy under specific conditions, is often linked to energy retention and dissipation in control theory. At its core, passivity delineates a system's capacity to absorb, dissipate, or retain energy without amplifying it under specified conditions, thereby ensuring stability and boundedness in its state space. A dynamic system is deemed passive if its energy does not increase within its state space [20]. Recently, the concept of dissipation has found widespread application in various fields, including chemistry, biology, complex networks, signal processing, chaos control, and social sciences. Consequently, exploring the passivity of gene regulatory networks is crucial due to its relevance in numerous research and analytical spheres.
In genetic regulation, biochemical reactions involve the diffusion of regulatory proteins and metabolites between the cytosol and nucleus. It is more accurate to consider reaction-diffusion models for genetic regulatory networks than to assume spatial homogeneity. In fact, it has been suggested that reaction-diffusion systems play an important role in the dynamics of GRNs [21,22]. Mathematical models without considering reaction-diffusion effects may lead to incorrect predictions of protein and mRNA concentrations. Therefore, it is essential to consider reaction-diffusion systems in the models of GRNs. However, there are few relevant articles, especially in the study of discrete-time GRNs with reaction-diffusion in plants. In discrete-time gene regulatory networks, the nature of reaction-diffusion reflects the coupling between adjacent spatial locations. Specifically, it represents the interaction between neighboring spatial locations due to their close proximity, which leads to mutual influence on each other's function. Based on this observation, the present study establishes a gene regulatory network model that incorporates the mutual coupling between spatial locations. The GRNs constructed through differential equations in [5,6] are continuous models that solely reflect the temporal changes in protein and mRNA concentrations. In contrast to the continuous models proposed in [5,6], the proposed discretized GRN model offers additional insights. It can not only capture temporal variations in protein and mRNA concentrations but also reveal their spatial variations through the inclusion of reaction-diffusion terms. Computer simulations demonstrate the effectiveness of the theoretical results and the importance of incorporating reaction-diffusion terms in GRN models.
The main difficulties and challenges of this paper stem from the fact that the diffusion terms in GRNs first appeared in models described by differential equations, with limited research on discrete-time GRNs. For electronic computers, continuous models are not directly describable. Therefore, the modeling of continuous GRNs becomes highly necessary. However, since spatial diffusion is characterized by the spatial divergence of mRNA and protein concentrations, defining the spatial diffusion terms of mRNA and proteins in discrete-time GRN models and proving that such models exhibit passivity is quite challenging. In this paper, we address these challenges by discretizing spatial divergence and leveraging Lyapunov functions and linear matrix inequalities to provide solutions.
The main contributions of this article are as follows:
1) The introduction of the concept of spatial diffusion into a discrete GRN, constructing a discretely coupled gene regulatory network with reaction diffusion.
2) By considering Brownian motion and time delay, the article provides a theorem for determining the passivity in a discretely GRN with Dirichlet boundary conditions and reaction diffusion.
For readability, the meanings of commonly used mathematical symbols involved in this paper are as follows:
RN means the N dimensional real vector, and RN×N stands for the sets of N×N real matrices. ∂f(x)∂x denotes the partial derivative of f(x) with respect to x. N∑i=1γi=γ1+γ2+⋯+γN, N⋃i=1αi={α1,α2,αN}. The superscript T stands for matrix transposition.
A typical nonlinear delayed genetic regulatory network based on a differential equation model with reaction-diffusion terms is described as [21]:
{∂m′(t,x)∂t=l∑k=1∂∂xk(Mk∂m′(t,x)∂xk)−Am′(t,x)+Bf(p′(t−t(k),x))+u(t,x)p′(t,x)=l∑k=1∂∂xk(M∗k∂m′(t,x)∂xk)−Cp′(t,x)+Dm′(t−δ(t),x)+v(t,x) | (2.1) |
where
A=diag{a1,a2,⋯,aN}, |
C=diag{c1,c2,⋯,cN}, |
D=diag{d1,d2,⋯,dN}, |
m′(t,x)=[m′1(t,x),m′2(t,x),⋯,m′N(t,x)]T, |
p′(t,x)=[p′1(t,x),p′2(t,x),⋯,p′N(t,x)]T, |
u(t,x)=[u1(t,x),u2(t,x),⋯,uN(t,x)]T, |
v(t,x)=[v1(t,x),v2(t,x),⋯,vN(t,x)]T, |
f(p′(t−τ(t),x))=[f1(p′1(t−τ(t),x)),⋯,fN(p′N(t−τ(t),x))]T, |
and B=(bnn′)∈RN×N, bnn′ represents the interaction between transcription factors and genes. Specifically, bnn′ can take one of three values: αnn′, 0, or −αnn′. Here, αnn′ represents the magnitude of the interaction occurring between transcription factor n′ and gene n. In detail, a positive αnn′ signifies that there is an activating influence, a zero value indicates the absence of any direct interaction, and a negative αnn′ signifies that there is a repressing influence.
In Eq (2.1), the matrix B represents the interactions between genes in the GRN. The notations m′n(t,x) and p′n(t,x) (n=1,2,⋯,N) represent the concentrations of mRNA and protein at time t for the nth node, respectively. The vector x=(x1,x2,⋯,xl)T belongs to the set Q. Specifically, Q is defined as {x||xα|≤Lα,α=1,2,⋯,l} and represents a compact set in the real vector space Rl with a smooth boundary ∂Q. The terms u(t,x) and v(t,x) stand for the disturbance input signals. The parameters A and C are the degradation rates of the mRNA and protein, respectively, while D signifies the translation rate. The matrices Mα>0 and M∗α>0 demonstrate the diffusion coefficient matrices. The regulatory function fn(x) is in the Hill form and captures the feedback regulation of protein on transcription; its mathematical form is described as
fn(x)=(xvn)Hn1+(xvn)Hn, | (2.2) |
where Hn is the Hill coefficient and vn is a positive constant.
Based on Eq (2.1), we construct a corresponding discrete gene regulatory network model with reaction diffusion terms, as shown in Figure 1.
{m(k+1,i)=A′m(k,i)+Bf(p(k−τ(k),i))+Δ1+u(k,i)p(k+1,i)=C′p(k,i)+Δ2+Dm(k−δ(k),i)+v(k,i) | (2.3) |
where
m(k,i)=[m1(k,i),m2(k,i),⋯,mN(k,i)]T,p(k,i)=[p1(k,i),p2(k,i),⋯,pN(k,i)]T,u(k,i)=[u1(k,i),u2(k,i),⋯,uN(k,i)]T,v(t,x)=[v1(k,i),v2(k,i),⋯,vN(k,i)]T,f(p(k−τ(k),i))=[f1(p1(k−τ(k),i)),⋯,fN(pN(k−τ(k),i))]T, |
mn(k,i) and pn(k,i) represent the concentration of mRNA and protein of the nth node at the time k∈N and space i respectively, where i=1,2,⋯,L. Δ1 and Δ2 represent the reaction diffusion terms of concentrations of mRNA and protein, respectively. The discretization of the spatial diffusion term is obtained by discretizing the spatial diffusion term of the continuous gene regulatory network as shown in Eq (2.3), with the specific correspondence shown in the following equations:
l∑k=1∂∂xk(Mk∂m′(t,x)∂xk)=l∑k=1Mk∂2m′(t,x)∂x2k. | (2.4) |
In the modeling of gene regulatory networks, it is very important to choose a suitable discretization scheme. Different discretization schemes may have advantages in different application scenarios. For example, the forward difference and backward difference methods may be easier to implement or more efficient in some cases, although their accuracy is low. In our model, the discretization formula for the second-order partial derivative ∂2m′(t,x)∂x2k typically employs the central difference method for approximation because this method estimates the derivative of a point by calculating the values on both sides of the functio. This can provide a more accurate derivative approximation and is very important for understanding and predicting the behavior of gene regulatory networks. In addition, the central difference method can better maintain the physical meaning and mathematical properties of the system when dealing with the spatial diffusion term.
For the function m′(t,x), the discretization formula for its second-order partial derivative is as follows:
∂2m′(t,x)∂x2k→m′(t,x+hk)−m′(t,x)−(m′(t,x)−m′(t,x−hk))h2k, | (2.5) |
where hk is the step size in the direction of xk.
After the model is discretized, the spatial diffusion term also needs to be discretized. The term m′(t,x) is transformed into m(t,i), where t corresponds to time, and i corresponds to the spatial position x. The discretization process is carried out to facilitate computer simulation and computation. Clearly, in discrete systems, the step size is typically taken as 1, that is, hk=1, and Eq (2.5) can be simplified as follows:
m′(t,x+hk)−m′(t,x)−(m′(t,x)−m′(t,x−hk)). | (2.6) |
If Mk is set to θ, then the reaction-diffusion term of mRNA concentration
Δ1=θ[m(k,i+1)−m(k,i)−(m(k,i)−m(k,i−1))] | (2.7) |
in the discrete form can be obtained. Similarly, the reaction-diffusion term Δ2 for protein concentration can be obtained in the discrete form, and thus we arrive at Eq (2.8).
{Δ1=θ[m(k,i+1)−m(k,i)−(m(k,i)−m(k,i−1))]Δ2=μ[p(k,i+1)−p(k,i)−(p(k,i)−p(k,i−1))] | (2.8) |
where θ and μ are positive coupling coefficients.
Once Eq (2.8) has been substituted into Eq (2.3), and the subsequent simplification process has been carried out, the final expression is obtained:
{m(k+1,i)=Am(k,i)+Bf(p(k−τ(k),i))+θ(m(k,i+1)+m(k,i−1))+u(k,i)p(k+1,i)=Cp(k,i)+μ(p(k,i+1)+p(k,i−1))+Dm(k−δ(k),i)+v(k,i) | (2.9) |
where A=A′−2θ and C=C′−2μ.
Remark 2.1. To provide a concrete application example, we refer to the research by Zhang et al. [23], which effectively demonstrates the importance of combining regulatory networks with reaction-diffusion terms. Zhang et al. studied the oscillatory expression of Escherichia coli mediated by microRNA, incorporating both time delays and reaction-diffusion components. They demonstrated that the oscillatory expression of Escherichia coli is not only critically dependent on transcriptional and translational delays but also significantly influenced by the diffusion coefficients. This conclusion has been practically verified by numerous biological experiments and observations. They also found that if the diffusion coefficients of microRNA, mRNA, and protein are sufficiently small, non-uniform periodic oscillations can be predicted to occur, unless only spatially uniform periodic oscillations are exhibited. Thus, the application example provided by Zhang et al. [23] serves as a practical illustration of how the combination of regulatory networks and reaction-diffusion terms, as considered in our paper, can be applied to model and understand complex biological phenomena such as the oscillatory expression of Escherichia coli.
The current research focuses on the Dirichlet boundary condition, which is characterized by the following mathematical expressions:
{mn(k,i)=0,i∈∂Ωpn(k,i)=0,i∈∂Ω | (2.10) |
Here, the symbol ∂Ω denotes the boundary of the domain under consideration.
Dirichlet boundary conditions provide a clear and physically meaningful way to define the behavior of the system at the boundaries, which is very important for drawing meaningful conclusions about the passivity and stability of the network. The use of Dirichlet boundary conditions in our model enables us to focus on the internal dynamics of the system instead of being dispersed by the complexities of boundary behavior. This simplification enables us to deduce the analysis results, which provide a deep understanding of the behavior of gene regulatory networks with spatial diffusion. The unique contribution of our research lies in applying these boundary conditions to solve the specific problem of analyzing the passivity of discrete gene regulatory networks with spatial diffusion.
Considering that the concentrations of mRNA and proteins are influenced by molecular Brownian motion, Eq (2.9) is rewritten as
{m(k+1,i)=Am(k,i)+Bf(p(k−τ(k),i))+θ(m(k,i+1)+m(k,i−1))+σ(p(k,i),p(k−τ(k),i))ω(k,i)+u(k,i)p(k+1,i)=Cp(k,i)+μ(p(k,i)+p(k,i)+Dm(k−δ(k),i))+v(k,i) | (2.11) |
where ω(k,i) represents a vector-form scalar Brownian motion with the following properties:
ε{ω(k,i)}=0,ε{ω(k,i)Tω(k,i)}=1,ε{ω(k,i)Tω(k′,i)}=0(k′≠k), |
and σ(p(k,i),p(k−τ(k),i)) is the noise intensity matrix satisfying
σT(pn(k,i),pn(k−τ(k),i))σ(pn(k,i),pn(k−τ(k),i))≤pTn(k,i)H1pn(k,i)+pTn(k−τ(k),i)H2pn(k−τ(k),i) | (2.12) |
where H1 and H2 are known constant matrices of appropriate dimensions.
τ(m) and δ(m) are time-varying delays satisfying
{0<τmin≤τ(m)≤τmax0<δmin≤δ(m)≤δmaxϕ′1≤˙τ≤ϕ1ϕ′2≤˙δ≤ϕ2 | (2.13) |
The nonlinear function fi(⋅) satisfies Inequality (2.14) due to its monotonic increasing nature with saturation
0≤fi(xi)xi≤δi,xi≠0,i=1,2,⋯,n. | (2.14) |
Furthermore, this inequality can be reformulated in matrix form as follows:
fT(x)(f(x)−Kx)≤0, | (2.15) |
where K=diag(k1,k2,⋯kn),∀x∈Rn.
Lemma 2.2 ([24]). For any constant matrix W=WT>0, scalar r>1, there is
rk−1∑l=k−rxT(l)Wx(l)≥k−1∑l=k−rxT(l)Wk−1∑l=k−rx(l). | (2.16) |
Lemma 2.3 ([24]). For every pair of vectors X,Y∈Rn, where H>0 denotes a positive definite matrix, the subsequent inequality is guaranteed to exist
2XTHY≤XTHX+YTH−1Y. | (2.17) |
The general form of a linear matrix inequality (LMI) is as follows:
LMI(x)=A0+x1A1+x2A2+⋯+xnAn≥0, | (3.1) |
where x=(x1,x2,⋯,xn) is a vector of real decision variables, and A0,A1,⋯,An are given symmetric matrices (typically real-valued). Note that the symbol ≥0 indicates that the matrix is positive semidefinite, which means all its eigenvalues are non-negative. The solution set of LMI is a convex set, meaning that its solution space has favorable mathematical properties. However, this also limits the range of feasible solutions. This convexity makes LMI very useful in optimization problems, but it increases its constraint. LMI is often used in system stability analysis, such as for verifying the stability of the system through Lyapunov functions. The application environment requires LMI to satisfy strict mathematical conditions to ensure the stability of the system.
In the subsequent discussions, an exploration will be conducted into a stability criterion pertaining to Eq (2.12) when subjected to Dirichlet boundary conditions.
Theorem 3.1 ([24]). For given scalars δ and τ satisfying Inequality (2.13), and under the condition ω(k,i)=0, the trivial solution of Eq (2.12) without molecular Brownian motion under Dirichlet boundary conditions is passive if there exist scalars γ>0, matrices ΛTh=Λh>0(h=1,2,3), PTh′=Ph′>0, RTh′=Rh′>0, QTh′=Qh′>0(h′=1,2,⋯,4), TT1=T1>0, ST1=S1>0 and ST2=S2>0, such that the following linear matrix inequalities hold:
Ξ1=[Σ1,10Σ1,30Σ1,5Σ1,6Σ1,7∗Σ2,2000∗∗Σ3,3Σ3,4000∗∗∗Σ4,4000∗∗∗∗Σ5,5Σ5,6Σ5,7∗∗∗∗∗Σ6,6Σ6,7∗∗∗∗∗∗Σ7,7]<0, | (3.2) |
Ξ2=[∏1,10∏1,30∏1,50∏1,7∏1,8∏1,9∗Π2,2000∏2,6000∗∗∏3,3∏3,400000∗∗∗∏4,400000∗∗∗∗∏5,50000∗∗∗∗∗∏6,6000∗∗∗∗∗∗∏7,7∏7,8∏7,9∗∗∗∗∗∗∗∏8,8∏8,9∗∗∗∗∗∗∗∗∏9,9]<0, | (3.3) |
where
Σ1,1=2ATP1A−P1+P3+Q1+2δmin(A−I)TR1(A−I)−R1δmin+2(δmax−δmin)(A−I)TR2(A−I)+2λ1S1,
Σ1,3=R1δmin,
Σ1,5=ATP1+δmin(A−I)TR1+(δmax−δmin)(A−I)TR2−Λ3,
Σ1,6=ATP1θ+δmin(A−I)TR1θ+(δmax−δmin)(A−I)TR2θ,
Σ1,7=ATP1θ+δmin(A−I)TR1θ+(δmax−δmin)(A−I)TR2θ,
Σ2,2=5DTP2D−P3+5τminDTR3D+5(τmax−τmin)DTR4D,
Σ3,3=Q2−Q1−R1δmin−R2δmax−δmin,
Σ3,4=R2δmax−δmin,
Σ4,4=−Q2−R2δmax−δmin,
Σ5,5=2P1+2δminR1+2(δmax−δmin)R2−γΛ3,
Σ5,6=P1θ+δminR1θ+(δmax−δmin)R2θ,
Σ5,7=P1θ+δminR1θ+(δmax−δmin)R2θ,
Σ6,6=2θTP1θ+2δminθTR1θ+2(δmax−δmin)R2θ−S1,
Σ6,7=θTP1θ+δminθTR1θ+(δmax−δmin)R2θ,
Σ7,7=2θTP1θ+2δminθTR1θ+2(δmax−δmin)R2θ−S1,
∏1,1=2CTP2C−P2+P4+Q3+2τmin(C−I)TR3(C−I)−R3τmin+2(τmax−τmin)(C−I)TR4(C−I)+2λ2S2,
∏1,3=R3δmin,
∏1,5=KΛ1,
∏1,7=CTP2+τmin(C−I)TR3+(τmax−τmin)(C−I)TR4−Λ3,
∏1,8=CTP2μ+τmin(C−I)TR3μ+(τmax−τmin)(C−I)TR4μ,
∏1,9=CTP2μ+τmin(C−I)TR3μ+(τmax−τmin)(C−I)TR4μ,
∏2,2=−P4,
∏2,6=KΛ2 ∏3,3=Q3−Q4−R3τmin−R4τmax−τmin,
∏3,4=R4δmax−δmin,
∏4,4=−Q4−R4τmax−τmin,
∏5,5=T1−2Λ1,
∏6,6=5BTP1B+5δminBTR1B+5(δmax−δmin)BTR2B−T1−2Λ2,
∏7,7=2P2+2τminR3+2(τmax−τmin)R4−γΛ3,
∏7,8=P2μ+τminR3μ+(τmax−τmin)R4μ,
∏7,9=P2μ+τminR3μ+(τmax−τmin)R4μ,
∏8,8=2μTP2μ+2τminμTR3μ+2(δmax−δmin)μTR4μ−S2,
∏8,9=μTP2μ+τminμTR3μ+(δmax−δmin)μTR4μ,
∏9,9=2μTP2μ+2τminμTR3μ+2(δmax−δmin)μTR4μ−S2.
Proof. Set
η1(k,i)=m(k+1,i)−m(k,i),η2(k,i)=p(k+1,i)−p(k,i). | (3.4) |
Consider a Lyapunov-Krasovskii functional candidate defined as
V(k,i)=5∑g=1Vg(k,i), | (3.5) |
where
V1(k,i)=mT(k,i)P1m(k,i)+pT(k,i)P2p(k,i), | (3.6) |
V2(k,i)=k−1∑l=k−δ(k)mT(l,i)P3m(l,i)+k−1∑l=k−τ(k)pT(l,i)P4p(l,i), | (3.7) |
V3(k,i)=k−1∑l=k−δminmT(l,i)Q1m(l,i)+k−δmin−1∑l=k−δmaxmT(l,i)Q2m(l,i)+k−1∑l=k−τminpT(l,i)Q3p(l,i)+k−τmin−1∑l=k−τmaxpT(l,i)Q4p(l,i), | (3.8) |
V4(k,i)=−1∑ϑ=−δmink−1∑l=k+ϑηT1(l,i)R1η1(l,i)+−1∑ϑ=−δmax+δmink−1∑l=k−δmin+ϑηT1(l,i)R2η1(l,i)+−1∑ϑ=−τmink−1∑l=k+ϑηT2(l,i)R3η2(l,i)+−1∑ϑ=−τmax+τmink−1∑l=k−τmin+ϑηT2(l,i)R4η2(l,i), | (3.9) |
V5(k,i)=k−1∑l=k−τ(k)fT(p(l,i))T1f(p(l,i)). | (3.10) |
Define ΔV(k,i)=V(k+1,i)−V(k,i), then
ε{ΔV(k,i)}=ε{5∑i=1ΔVi(k,i)}, | (3.11) |
where
ε{ΔV1(k,i)}=ε{mT(k+1,i)P1m(k+1,i)−mT(k,i)P1m(k,i)+pT(k+1,i)P2p(k+1,i)−pT(k,i)P2p(k,i)}, | (3.12) |
ε{ΔV2(k,i)}=ε{mT(k,i)P3m(k,i)−mT(k−δ(k),i)P3m(k−δ(k),i)+pT(k,i)P4p(k,i)−pT(k−τ(k),i)P4p(k−τ(k),i)}, | (3.13) |
ε{ΔV3(k,i)}=ε{mT(k,i)Q1m(k,i)+mT(k−δmin,i)(Q2−Q1)m(k−δmin,i)−mT(k−δmax,i)Q2m(k−δmax,i)+pT(k,i)Q3p(k,i)+pT(k−τmin,i)(Q4−Q3)p(k−τmin,i)−pT(k−τmax,i)Q4p(k−τmax,i)}, | (3.14) |
ε{ΔV4(k,i)}=ε{δminηT1(k,i)R1η1(k,i)−k−1∑l=k−δminηT1(l,i)R1η1(l,i)+(δmax−δmin)ηT1(k,i)R2η1(k,i)−k−δmin−1∑l=k−δmaxηT1(l,i)R2η1(l,i)+τminηT2(k,i)R3η2(k,i)−k−1∑l=k−τminηT2(l,i)R3η2(l,i)+(τmax−τmin)ηT2(k,i)R4η2(k,i)−k−δmin−1∑l=k−τmaxηT2(l,i)R4η2(l,i)}, | (3.15) |
ε{ΔV5(k,i)}=ε{fT(p(k,i))T1f(p(k,i))−fT(p(k−τ(k),i))T1f(p(k−τ(k),i))}. | (3.16) |
Taking Eq (3.4) into consideration, for diagonal matrices Λ1>0, Λ2>0, there is
2pT(k,i)KΛ1f(p(k,i))−2fT(p(k,i))Λ1f(p(k,i))≥0, | (3.17) |
2pT(k−τ(k),i)KΛ2f(p(k−τ(k),i))−2fT(p(k−τ(k),i))Λ2f(p(k−τ(k),i))≥0. | (3.18) |
Using Lemma 2.2, we have
−k−1∑l=k−δminηT1(l,i)R1η1(l,i)≤−1δmin[m(k,i)−m(k−δmin,i)]TR1[m(k,i)−m(k−δmin,i)], | (3.19) |
−k−δmin−1∑l=k−δmaxηT1(l,i)R2η1(l,i)≤−1δmax−δmin[m(k−δmin,i)−m(k−δmax,i)]TR2[m(k−δmin,i)−m(k−δmax,i)], | (3.20) |
−k−1∑l=k−τminηT2(l,i)R3η2(l,i)≤−1τmin[p(k,i)−m(k−τmin,i)]TR3[p(k,i)−p(k−τmin,i)], | (3.21) |
−k−τmin−1∑l=k−τmaxηT2(l,i)R4η2(l,i)≤−1τmax−τmin[p(k−τmin,i)−m(k−τmax,i)]TR4[p(k−τmin,i)−p(k−τmax,i)]. | (3.22) |
The following expression is given to show the passivity property
J(kp,i)=E{kp∑k=1[−2(mT(k,i)Λ3u(k,i)+pT(k,i)Λ3v(k,i))−γ(uT(k,i)Λ3u(k,i)+vT(k,i)Λ3v(k,i))]}, | (3.23) |
and it is obvious that
J(kp,i)=E{kp∑k=1[ΔV(k,i)−2(mT(k,i)Λ3u(k,i)+pT(k,i)Λ3v(k,i))−γ(uT(k,i)Λ3u(k,i)+vT(k,i)Λ3v(k,i))]−V(kp+1,i)+V(1,i)}≤E{kp∑k=1[ΔV(k,i)−2(mT(k,i)Λ3u(k,i)+pT(k,i)Λ3v(k,i))−γ(uT(k,i)Λ3u(k,i)+vT(k,i)Λ3v(k,i))]+V(1,i)}, | (3.24) |
where k∈N, such that for the original time k0=1, set the initial states of m and p as m(1,i)=p(1,i)=0, so that taking (3.6)–(3.10) into account, we can obtain
V1(1,i)=V2(1,i)=V3(1,i)=V4(1,i)=V5(1,i)=0. | (3.25) |
Considering the spatial continuity and limitations of GRNs, we can obtain
{mT(k,i+1)S1m(k,i+1)≤λ1mT(k,i)S1m(k,i)mT(k,i−1)S1m(k,i−1)≤λ1mT(k,i)S1m(k,i) | (3.26) |
{pT(k,i+1)S2p(k,i+1)≤λ2pT(k,i)S2p(k,i)pT(k,i−1)S2p(k,i−1)≤λ2pT(k,i)S2p(k,i) | (3.27) |
where
λ1=max(max(⋃k∈Nℏk1),max(⋃k∈Nℓk1)),λ2=max(max(⋃k∈Nℏk2),max(⋃k∈Nℓk2)), |
ℏk1=max(L⋃i=1mT(k,i+1)m(k,i+1)mT(k,i)m(k,i)),ℓk1=max(L⋃i=1mT(k,i−1)m(k,i−1)mT(k,i)m(k,i)), |
ℏk2=max(L⋃i=1pT(k,i+1)p(k,i+1)pT(k,i)p(k,i)),ℓk2=max(L⋃i=1pT(k,i−1)p(k,i−1)pT(k,i)p(k,i)). |
Then, by combining Eq (3.11)–(3.27), we can obtain
E{kp∑k=1[2(mT(k,i)u(k,i)+pT(k,i)v(k,i))−γ(uT(k,i)u(k,i)+vT(k,i)v(k,i))]}≤E{ξTΞξ}<0, | (3.28) |
where ξ=[mT(k,i),mT(k−δ(k),i),mT(k−δmin,i),mT(k−δmax,i),pT(k,i),pT(k−τ(k),i),pT(k−τmin,i),pT(k−τmax,i),fT(p(k,i)),fT(p(k−τ(k),i)),uT(k,i),vT(k,i),mT(k,i+1),mT(k,i−1),pT(k,i+1),pT(k,i−1)].
This completes the proof.
Theorem 3.2. For given scalars δ and τ satisfying inequality (2.13), the trivial solution of Eq (2.12) with molecular Brownian motion under Dirichlet boundary conditions is passive if there exist scalars γ>0 and ρ>0, matrices ΛTh=Λh>0(h=1,2), PTh′=Ph′>0, RTh′=Rh′>0 and QTh′=Qh′>0(h′=1,⋯,4), TT1=T1>0, ST1=S1>0 and ST2=S2>0, such that the following LMIs hold:
P1+δminR1+(δmax−δmin)R2≤ρI, | (3.29) |
Θ1=[∑110∑130∑15∑16∑17∑18∗∑22000000∗∗∑33∑340000∗∗∗∑440000∗∗∗∗∑55∑56∑57∑58∗∗∗∗∗∑66∑67∑68∗∗∗∗∗∗∑77∑78∗∗∗∗∗∗∗∑88]<0, | (3.30) |
Θ2=[∏110∏130∏15∏16∏17∏18∗∏22000000∗∗∏33∏340000∗∗∗∏440000∗∗∗∗∏55000∗∗∗∗∗∏66∏67∏68∗∗∗∗∗∗∏77∏78∗∗∗∗∗∗∗∏88]<0, | (3.31) |
where
∑11=ATP1A+δmin(A−I)TR1(A−I)+(δmax−δmin)(A−I)TR2(A−I)−P1+P3+Q1−R1/δmin+2λ1S1,
∑13=R1δmin,
∑15=ATP1B+δmin(A−I)TR1B+(δmax−δmin)(A−I)TR2B,
∑16=ATP1+δmin(A−I)TR1+(δmax−δmin)(A−I)TR2+Λ3,
∑17=ATP1θ+δmin(A−I)TR1θ+(δmax−δmin)(A−I)TR2θ,
∑18=ATP1θ+δmin(A−I)TR1θ+(δmax−δmin)(A−I)TR2θ,
∑22=3DTP2D−P3+3τminDTR3D+3(τmax−τmin)DTR4D,
∑33=Q2−Q1−R1δmin−R2τmax−τmin,
∑34=R2δmax−δmin,
∑44=−Q2−R2δmax−δmin,
∑55=BTP1B+δminBTR1B+(δmax−δmin)BTR2B−T1+Λ2,
∑56=BTP1+δminBTR1+(δmax−δmin)BTR2,
∑57=BTPθ+δminBTR1θ+(δmax−δmin)BTR2θ,
∑58=BTPθ+δminBTR1θ+(δmax−δmin)BTR2θ,
∑66=P1+δminR1+(δmax−δmin)R2−γΛ3,
∑67=P1θ+δminR1θ+(δmax−δmin)R2θ,
∑68=P1θ+δminR1θ+(δmax−δmin)R2θ,
∑77=θTP1θ+δminθTR1θ+(δmax−δmin)θTR2θ−S1,
∑78=θTP1θ+δminθTR1θ+(δmax−δmin)θTR2θ,
∑88=θTP1θ+δminθTR1θ+(δmax−δmin)θTR2θ−S1,
Π11=2CTP2C−P2+P4+Q3+2τmin(C−I)TR3(C−I)−R3/τmin+2(τmax−τmin)(C−I)TR4(C−I)+ρH1+2λ2S2,
∏13=R3/τmin,
∏15=KΛ1,
∏16=CTP2+τmin(C−I)TR3+(τmax−τmin)(C−I)TR4+Λ3,
∏17=CTP2μ1+τmin(C−I)TR3μ1+(τmax−τmin)(C−I)TR4μ1,
∏18=CTP2μ2+τmin(C−I)TR3μ2+(τmax−τmin)(C−I)TR4μ2,
∏22=−P4+KTΛ2K+ρH2,
∏33=Q3−Q4−R3/τmin−R4/(τmax−τmin),
∏34=R4/(τmax−τmax),
∏44=−Q4−R4/(τmax−τmin),
∏55=−2Λ1+T1,
∏66=P2+τmaxR3+(τmax−τmin)R4−γΛ3,
∏67=P2μ+τminR3μ+(τmax−τmin)R4μ,
∏68=P2μ+τminR3μ+(τmax−τmin)R4μ,
∏77=μTP2μ+τminμTR3μ+(τmax−τmin)μTR4μ−S2,
∏78=μTP2μ+τminμTR3μ+(τmax−τmin)μTR4μ,
∏88=μTP2μ+τminμTR3μ+(τmax−τmin)μTR4μ−S2.
Proof. Using Lemma 2.3, we have
{2pT(k,i)CTP2Dm(k−δ(k),i)≤pT(k,i)CTP2Cp(k,i)+mT(k−δ(k),i)DTP2Dm(k−δ(k),i)2mT(k−δ(k),i)DTP2D2v(k,i)≤mT(k−δ(k),i)DTP2Dm(k−δ(k),i)+vT(k,i)DT2P2D2v(k,i)2τminpT(k,i)(C−I)TR3Dm(k−δ(k),i)≤τminpT(k,i)(C−I)TR3(C−I)p(k,i)+τminmT(k−δ(k),i)DTR3Dm(k−δ(k),i)2τminmT(k−δ(k),n)DTR3v(k,n)≤τminmT(k−δ(k),i)DTR3Dm(k−δ(k),i)+τminvT(k,i)R3v(k,i)2(τmax−τmin)pT(k,i)(C−I)TR4Dm(k−δ(k),i)≤(τmax−τmin)pT(k,i)(C−I)TR4(C−I)p(k,i)+(τmax−τmin)mT(k−δ(k),i)DTR4Dm(k−δ(k),i)2(τmax−τmin)mT(k−δ(k),i)DTR4v(k,i)≤(τmax−τmin)mT(k−δ(k),i)DTR4Dm(k−δ(k),i)+(τmax−τmin)vT(k,n)R4v(k,i)2pT(k−τ(k),i)KΛ2f((pi(k−τ(k),i))≤pT(k−τ(k),i)KTΛ2Kp(k−τ(k),i)+f((p(k−τ(k),i))Λ2f((p(k−τ(k),i)). | (3.32) |
Noting Inequality (2.13), it can be seen that
σT(p(k,i),p(k−τ(k),i))[P1+δminR1+(δmax−δmin)R2]σ(p(k,i),p(k−τ(k),i))≤ρpT(k,i)H1p(k,i)+ρpT(k−τ(k),i)H2p(k−τ(k),i). | (3.33) |
Then combining Eqs (3.11) to (3.25) with Eqs (3.32) and (3.33), we get
E{kp∑k=1[2(mT(k,i)u(k,i)+pT(k,i)v(k,i))−γ(uT(k,i)u(k,i)+vT(k,i)v(k,i))]}≤E{ξT1Θ1ξ1+ξT2Θ2ξ2}<0, | (3.34) |
where
ξ1=[mT(k,i),mT(k−δ(k),i),mT(k−δmin,i),mT(k−δmax,i),fT(p(k−τ(k),i)),uT(k,i),mT(k,i+1),mT(k,i−1)],
ξ2=[pT(k,i),pT(k−τ(k),i),pT(k−τmin,i),pT(k−τmax,i),fT(p(k,i)),vT(k,i),pT(k,i+1),pT(k,i−1)].
The proof of the theorem is now complete.
Remark 3.3. In real genetic regulatory networks, the functions and operations of mRNA and proteins are often affected by noise, which includes physical noise and biological noise. Among these, Brownian motion represents a form of physical noise that arises from the random movements of large molecules such as mRNA and proteins within the cell. This randomness contributes to fluctuations in molecular concentrations and interactions, thereby affecting the overall functionality and stability of the network. Theorem 3.1 presents a foundational understanding of these networks without accounting for certain types of noise. In contrast, Theorem 3.2 takes into account the impact of physical noise generated by Brownian motion on the passivity of the model to a greater extent. This not only enhances the model's predictive power but also provides deeper insights into the role of physical noise in shaping network behavior.
In this section, we demonstrate the validity of Theorems 3.1 and 3.2 through two examples. Example 4.1 showcases the passivity of the proposed GRNs model without Brownian motion, while Example 4.2 illustrates the passivity of the proposed GRNs model with Brownian motion included.
Example 4.1. In this section, we consider a GRN (2.12) with 5 nodes, L=100. The parameters are assumed to be
A=C=diag(0.1,0.1,0.1,0.1,0.1),D=(0.8,0.8,0.8,0.8,0.8),
B=[0−0.50.500−0.5000.50.500.50000.5−0.50000000.50],K=[0.65000000.65000000.65000000.65000000.65].
u1(k,i)=0.5(1−i−L2L2)2sin(2k)×m1(k,i), u2(k,i)=0.5(1.2−i−L2L2)2cos(2k)×m2(k,i),
u3(k,i)=0.5(0.5−i−L2L2)2sin(k)×m3(k,i), u4(k,i)=0.5(0.2−i−L2L2)2cos(0.2k)×m4(k,i),
u5(k,i)=0.5(0.5−i−L2L2)2sin(0.3k)×m5(k,i), v1(k,i)=0.5(1.2−i−L2L2)2sin(3k)×p1(k,i),
v2(k,i)=0.5(1−i−L2L2)2cos(k)×p2(k,i), v3(k,i)=0.5(0.7−i−L2L2)2cos(2k)×p3(k,i),
v4(k,i)=0.5(1−i−L2L2)2cos(k)×p4(k,i), v5(k,i)=0.5(0.7−i−L2L2)2cos(2k)×p5(k,i),
where f(x)=x2/(1+x2), the time delays δ(k)=4+2sin(kπ/2) and τ(k)=4+sin(kπ/2), so that δmin=2, δmax=6, τmin=3, τmax=5, coupling coefficient θ=μ=0.2.
The simulation result of trajectories of mRNA and protein concentrations for Example 4.1 are shown in Figures 2–6.
If ϕ′1=ϕ′2=ϕ′∈{0.1,0.3,0.5,0.7,0.9,1.0}, ϕ1=ϕ2=ϕ>1. The maximum delay τmax=δmax is obtained based on ([21], Theorem 1), ([25], Remark 2), ([22], Theorem 1), and Theorem 3.1; the following illustrations are made for Table 1:
● If μ≤0.7, the LMI conditions outlined in ([21], Theorem 1), ([25], Remark 2), ([22], Theorem 1), and Theorem 3.1 are considered to be achievable.
● When μ=0.9, the LMI conditions given in ([22], Theorem 1) are feasible, but the LMI conditions presented in ([21], Theorem 1), ([25], Remark 2), and Theorem 3.1 are ∞.
● When μ=1, the LMI conditions discussed in ([21], Theorem 1) and ([25], Remark 2) are feasible, the LMI conditions discussed in ([22], Theorem 1) are unfeasible, and only the LMI conditions discussed in Theorem 3.1 are ∞.
Case | 0.1 | 0.3 | 0.5 | 0.7 | 0.9 | 1.0 |
([21], Theorem 1) | ∞ | ∞ | ∞ | ∞ | ∞ | 3.9616 |
([25], Remark 2) | ∞ | ∞ | ∞ | ∞ | ∞ | 5.4571 |
([22], Theorem 1) | ∞ | ∞ | ∞ | ∞ | 2.8994 | \ |
Theorem 3.1 | ∞ | ∞ | ∞ | ∞ | ∞ | ∞ |
Therefore, within the range of μ≤1, Theorem 3.1 of this paper exhibits a reduced level of conservativeness compared to ([21], Theorem 1), ([25], Remark 2), and ([22], Theorem 1).
Example 4.2. Considering the Brownian motion for GRNs (2.12), H1=H2=1, λ1=λ2=2, by using the Toolbox YALMIP in MATLAB to solve the LMIs (3.2) and (3.3), we can obtain the following feasible solution:
P1=[1.3576−0.07300.4901−0.29580.0314−0.07301.03780.14990.4016−0.45980.49010.14992.20790.5861−0.0700−0.29580.40160.58611.5541−0.17950.0314−0.4598−0.0700−0.17951.8931], |
P2=[2.96980.00090.0251−0.0237−0.00060.00092.8923−0.00160.0261−0.02560.0251−0.00163.09360.0260−0.0007−0.02370.02610.02603.04360.0023−0.0006−0.0256−0.00070.00232.9869], |
P3=[0.1386−0.00580.0370−0.0200.0028−0.00580.11320.01240.0308−0.03560.03700.01240.20660.0448−0.0059−0.02070.03080.04480.1554−0.01480.0028−0.0356−0.0059−0.01480.1789], |
P4=[0.57730.00290.00120.00290.00500.00290.60930.00360.00030.00120.00120.00360.55150.00010.00040.00290.00030.00010.6105−0.00450.00500.00120.0004−0.00450.5039], |
R1=[0.0141−0.00090.0059−0.00320.0004−0.00090.01050.00210.0049−0.00560.00590.00210.02500.0074−0.0011−0.00320.00490.00740.0167−0.00250.0004−0.0056−0.0011−0.00250.0204], |
R2=[0.0107−0.00060.0044−0.00240.0003−0.00060.00800.00160.0037−0.00420.00440.00160.01890.0055−0.0009−0.00240.00370.00550.0127−0.00190.0003−0.0042−0.0009−0.00190.0155], |
R3=[0.06080.00060.00070.00030.00120.00060.05710.00080.0005−0.00010.00070.00080.06660.00040.00010.00030.00050.00040.0617−0.00100.0012−0.00010.0001−0.00100.0654], |
R4=[0.02620.00040.00030.00030.00070.00040.02410.00050.00020.00000.00030.00050.02950.00020.00010.00030.00020.00020.0266−0.00070.00070.00000.0001−0.00070.0290], |
Q1=[0.1789−0.01060.0680−0.04110.0044−0.01060.13030.02060.0575−0.06590.06800.02060.29580.0803−0.0092−0.04110.05750.08030.2055−0.02440.0044−0.0659−0.0092−0.02440.2528], |
Q2=[0.1206−0.00690.0444−0.02740.0028−0.00690.08820.01300.0376−0.04300.04440.01300.19570.0519−0.0056−0.02740.03760.05190.1375−0.01540.0028−0.0430−0.0056−0.01540.1688], |
Q3=[0.09670.00270.00110.00270.00450.00270.08540.00330.00030.00110.00110.00330.11280.00010.00040.00270.00030.00010.0967−0.00410.00450.00110.0004−0.00410.1130], |
Q4=[0.75970.00050.00020.00060.00090.00050.75750.00070.00000.00030.00020.00070.7628−0.00000.00010.00060.0000−0.00000.7596−0.00080.00090.00030.0001−0.00080.7630], |
Λ1=diag(0.5739,0.5596,0.5650,0.5694,0.5711), |
Λ2=diag(0.7148,0.7849,0.6454,0.7673,0.5683), |
Λ3=diag(0.4336,0.4259,0.4491,0.4452,0.4326), |
T1=[0.4761−0.0509−0.0297−0.0360−0.0821−0.05090.4777−0.0613−0.0094−0.0221−0.0297−0.06130.4698−0.0061−0.0122−0.0360−0.0094−0.00610.48080.0705−0.0821−0.0221−0.01220.07050.4545], |
S1=[0.2240−0.01230.0836−0.05180.0051−0.01230.17040.02470.0678−0.07750.08360.02470.36660.0995−0.0113−0.05180.06780.09950.2564−0.02970.0051−0.0775−0.0113−0.02970.3157], |
S2=[0.47360.00160.0071−0.00460.00250.00160.45360.00150.0068−0.00590.00710.00150.50250.00670.0001−0.00460.00680.00670.4835−0.00180.0025−0.00590.0001−0.00180.4883], |
γ=0.7430,ρ=0.0119. |
Therefore, it can be concluded by Examples 4.1 and 4.2 that Eq (2.8) is stochastically passive.
Figure 7 presents the boundary between the passive and non-passive behavior regions of the proposed GRN model at μ=0.55 and θ=0.60. It can be clearly observed that when the values of the coupling coefficients θ and μ fall within the parameter range defined by Region Ⅰ, specifically 0.05≤μ≤0.55 and 0.05≤θ≤0.60, the proposed discrete-time gene regulatory network model exhibits passive behavior. Conversely, when the values of θ and μ fall within the range defined by Region Ⅱ, the proposed GRN model does not exhibit passive behavior. The dashed line in Figure 7 precisely delineates the boundary between the regions of passivity and non-passivity, distinguishing Region Ⅰ from Region Ⅱ. This boundary was determined through extensive computer software simulations, identifying the critical values of θ and μ at which the system transitions from passive to non-passive behavior. The shapes of the regional boundaries reflect the complex interplay between the model parameters and their impact on the system's passivity properties.
Remark 4.3. In real biological environments, the concentrations of mRNA and proteins are indeed influenced by a wide array of factors, including gene mutations, interactions between cells, biological enzymes, hormone levels, and environmental conditions. These factors contribute to the complex dynamics and heterogeneity observed in biological systems. However, our proposed model focuses on a more simplified mathematical representation of these processes, specifically considering the interactions between mRNA and proteins, the effects of spatial diffusion terms, and the impact of time delays on their concentrations. While this mathematical simplification allows us to derive analytical results and gain insights into the passivity and stability of the system, it also introduces limitations in the model's ability to fully capture the complexity of real biological environments. For instance, our model does not account for the potential effects of gene mutations or the intricate interactions between cells in a tissue or organ. Additionally, we do not consider the role of biological enzymes or hormone levels, which can significantly impact gene expression and protein synthesis.
It is important to recognize these limitations and to understand that our model represents a first step toward understanding the dynamics of gene regulatory networks in a mathematical framework. Future work could build upon our findings by incorporating additional factors that contribute to the complexity of real biological systems. Such extensions could provide even deeper insights into the mechanisms underlying gene expression and protein synthesis, as well as their implications for health and disease.
In this groundbreaking study, we propose a pioneering approach to reaction-diffusion mechanisms into discrete-time GRNs. Our primary focus is to explore the robust passivity of these networks under the influence of time-varying delay and Dirichlet boundary conditions using advanced Lyapunov-Krasovskii functions. Moreover, we delve into the investigation of asymptotic passivity for GRNs that incorporate reaction-diffusion terms along with Brownian motion. This unique exploration is unprecedented and marks the first-ever attempt to introduce reaction-diffusion into discrete-time GRNs. To demonstrate the efficacy and validity of our novel method, we present comprehensive numerical examples and simulation results. These findings showcase the correctness and effectiveness of our approach, solidifying its potential impact on the field of gene regulatory network modeling and analysis. In the future, the asynchronous behavior of discrete-time systems with reaction-diffusion coupling could be studied, in order to understand how different update schedules affect the passivity and stability of the network. In addition, we can also try to extend the proposed model to more complex spatial domains, such as those with irregular geometries or anisotropic diffusion, which will provide further insights into the spatial organization of gene regulatory networks.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This research was funded by Post-doctoral Natural Science Foundation of China under grant number 2022M720650.
The authors declare there is no conflicts of interest.
[1] |
J. Hasty, D. McMillen, F. Isaacs, J. J. Collins, Computational studies of gene regulatory networks: in numero molecular biology, Nat. Rev. Genet., 2 (2001), 268–279. https://doi.org/10.1038/35066056 doi: 10.1038/35066056
![]() |
[2] |
N. Bernaola, M. Michiels, P. Larrañaga, C. Bielza, Learning massive interpretable gene regulatory networks of the human brain by merging Bayesian networks, PloS Comput. Biol., 19 (2023), 1011443. https://doi.org/10.1371/journal.pcbi.1011443 doi: 10.1371/journal.pcbi.1011443
![]() |
[3] |
J. Musilova, Z. Vafek, B. L. Puniya, R. Zimmer, T. Helikar, K. Sedlar, Augusta: From RNA-Seq to gene regulatory networks and Boolean models, Comput. Struct. Biotechnol. J., 23 (2024), 783–790. https://doi.org/10.1016/j.csbj.2024.01.013 doi: 10.1016/j.csbj.2024.01.013
![]() |
[4] |
G. Karlebach, P. N. Robinson, Computing minimal Boolean models of gene regulatory networks, J. Comput. Biol., 31 (2024), 117–127. https://doi.org/10.1089/cmb.2023.0122 doi: 10.1089/cmb.2023.0122
![]() |
[5] |
I. Hossai, V. Fanfani, J. Fischer, J. Quackenbush, R. Burkholz, Biologically informed NeuralODEs for genome-wide regulatory dynamics, Genome Biol., 25 (2024), 127. https://doi.org/10.1186/s13059-024-03264-0 doi: 10.1186/s13059-024-03264-0
![]() |
[6] |
X. She, L. Wang, Y. Zhang, Finite-time stability of genetic regulatory networks with nondifferential delays, IEEE Trans. Circuits Syst. II: Express Briefs, 70 (2023), 2107–2111. https://doi.org/10.1109/TCSII.2022.3233797 doi: 10.1109/TCSII.2022.3233797
![]() |
[7] |
T. Zhang, Y. Li, Exponential Euler scheme of multi-delay Caputo-Fabrizio fractional-order differential equations, Appl. Math. Lett., 124 (2022), 107709. https://doi.org/10.1016/j.aml.2021.107709 doi: 10.1016/j.aml.2021.107709
![]() |
[8] |
A. Polynikis, S. J. Hogan, M. D. Bernardo, Comparing different ODE modelling approaches for gene regulatory networks, J. Theor. Biol., 261 (2009), 511–530. https://doi.org/10.1016/j.jtbi.2009.07.040 doi: 10.1016/j.jtbi.2009.07.040
![]() |
[9] |
M. Kchaou, G. Narayanan, M. S. Ali, S. Sanober, G. Rajchakit, B. Priya, Finite-time Mittag-Leffler synchronization of delayed fractional-order discrete-time complex-valued genetic regulatory networks: Decomposition and direct approaches, Inf. Sci., 664 (2024), 120337. https://doi.org/10.1016/j.ins.2024.120337 doi: 10.1016/j.ins.2024.120337
![]() |
[10] |
K. Xie, C. Zhang, S. Lee, Y. Liu, C. Zhai, Sector bound-dependent matrix-separation-based inequality and its application to stability analysis of discrete-time delayed genetic regulatory networks, IEEE Trans. Circuits Syst. II: Express Briefs, 71 (2024), 3785–3789. https://doi.org/10.1109/TCSII.2024.3367777 doi: 10.1109/TCSII.2024.3367777
![]() |
[11] |
K. Mathiyalagan, R. Sakthivel, Robust stabilization and H∞ control for discrete-time stochastic genetic regulatory networks with time delays, Can. J. Phys., 90 (2012), 939–953. https://doi.org/10.1139/p2012-088 doi: 10.1139/p2012-088
![]() |
[12] |
Q. Li, H. Q. Wei, D. Hua, J. Wang, J. Yang, Stabilization of semi-Markovian jumping uncertain complex-valued networks with time-varying delay: A sliding-mode control approach, Neural Process Lett., 56 (2024), 111. https://doi.org/10.1007/s11063-024-11585-1 doi: 10.1007/s11063-024-11585-1
![]() |
[13] |
G. Narayanan, M. S. Ali, H. Alsulami, T. Saeed, B. Ahmad Synchronization of T–S fuzzy fractional-order discrete-time complex-valued molecular models of mRNA and protein in regulatory mechanisms with leakage effects, Neural Process Lett., 55 (2023), 3305–3331. https://doi.org/10.1007/s11063-022-11010-5 doi: 10.1007/s11063-022-11010-5
![]() |
[14] |
H. Wei, K. Zhang, M. Zhang, Q. Li, J. Wang, Dissipative synchronization of Semi-Markovian jumping delayed neural networks under random deception attacks: An event-triggered impulsive control strategy, J. Franklin Inst., 361 (2024), 106835. https://doi.org/10.1016/j.jfranklin.2024.106835 doi: 10.1016/j.jfranklin.2024.106835
![]() |
[15] |
S. Pandiselvi, R. Raja, J. Cao, G. Rajchakit, B. Ahmad, Approximation of state variables for discrete-time stochastic genetic regulatory networks with leakage, distributed, and probabilistic measurement delays: A robust stability problem, Adv. Differ. Equations, 2018 (2018), 123. https://doi.org/10.1186/s13662-018-1569-z doi: 10.1186/s13662-018-1569-z
![]() |
[16] |
Y. Zhao, H. Wu, Fixed/Prescribed stability criterions of stochastic system with time-delay, AIMS Math., 9 (2024), 14425–14453. https://doi.org/10.3934/math.2024701 doi: 10.3934/math.2024701
![]() |
[17] |
T. Zhang, Z. Li, Switching clusters' synchronization for discrete space-time complex dynamical networks via boundary feedback controls, Pattern Recognit., 143 (2023), 109763. https://doi.org/10.1016/j.patcog.2023.109763 doi: 10.1016/j.patcog.2023.109763
![]() |
[18] |
R. Liang, J. Wang, PD control for passivity of coupled reaction-diffusion neural networks with multiple state couplings or spatial diffusion couplings, Neurocomputing, 489 (2022), 558–569. https://doi.org/10.1016/j.neucom.2021.12.070 doi: 10.1016/j.neucom.2021.12.070
![]() |
[19] |
P. U. Avila, T. Padvitski, A. C. Leote, H. Chen, J. Saez-Rodriguez, M. Kann, et al., Gene regulatory networks in disease and ageing, Nat. Rev. Nephrol., 20 (2024), 616–633. https://doi.org/10.1038/s41581-024-00849-7 doi: 10.1038/s41581-024-00849-7
![]() |
[20] |
P. Jothiappan, M. Kalidass, Robust passivity analysis of stochastic genetic regulatory networks with Levy noise, Int. J. Control Autom. Syst., 20 (2022), 3241–3251. https://doi.org/10.1007/s12555-021-0552-8 doi: 10.1007/s12555-021-0552-8
![]() |
[21] |
Y. Qin, F. Li, J. Wang, H. Shen, Extended dissipative synchronization of reaction-diffusion genetic regulatory networks based on sampled-data control, Neural Process Lett., 55 (2023), 3169–3183. https://doi.org/10.1007/s11063-022-11003-4 doi: 10.1007/s11063-022-11003-4
![]() |
[22] |
X. Song, X. Li, S. Song, C. K. Ahn, State observer design of coupled genetic regulatory networks with Reaction-diffusion terms via time-space sampled-data communications, IEEE/ACM Trans. Comput. Biol. Bioinf., 19 (2022), 3704–3714. https://doi.org/10.1109/TCBB.2021.3114405 doi: 10.1109/TCBB.2021.3114405
![]() |
[23] |
Y. Zhang, H. Liu, H. Yan, J. Zhou, Oscillatory behaviors in genetic regulatory networks mediated by microRNA with time delays and reaction-diffusion terms, IEEE Trans. Nanobiosci., 16 (2017), 166–176. https://doi.org/10.1109/TNB.2017.2675446 doi: 10.1109/TNB.2017.2675446
![]() |
[24] |
C. Ma, Q. Zeng, L. Zhang, Y. Zhu, Passivity and passification for Markov jump genetic regulatory networks with time-varying delays, Neurocomputing, 136 (2014), 321–326. https://doi.org/10.1016/j.neucom.2013.12.028 doi: 10.1016/j.neucom.2013.12.028
![]() |
[25] |
X. Zhang, Y. Han, L. Wu, Y. Wang, State estimation for delayed genetic regulatory networks with reaction-diffusion terms, IEEE Trans. Neural Networks Learn. Syst., 29 (2018), 299–309. https://doi.org/10.1109/TNNLS.2016.2618899 doi: 10.1109/TNNLS.2016.2618899
![]() |