Loading [MathJax]/jax/output/SVG/jax.js
Research article Topical Sections

Steady-state bifurcation and regularity of nonlinear Burgers equation with mean value constraint

  • In this paper, we focus on the steady-state bifurcation problem of the nonlinear Burgers equation within a bounded domain, considering both homogeneous Dirichlet boundary conditions and homogeneous Neumann boundary conditions with a mean value constraint. Unlike previous studies, we develop an enhanced turbulence model by incorporating nonlinear higher-order terms (such as u2 and u3) and linear source terms λu into the one-dimensional Burgers equation. Our steady-state bifurcation analysis establishes for the first time how the coupled forward energy cascade and inverse energy transfer mechanisms collectively govern the dynamics of initial flow instability. By combining the spectral theorem for a linear compact operator with the normalized Lyapunov–Schmidt reduction method and the implicit function theorem, we derive the complete criterion for the critical bifurcation condition, the explicit form of the bifurcation solution, and its regularity.

    Citation: Qingming Hao, Wei Chen, Zhigang Pan, Chao Zhu, Yanhua Wang. Steady-state bifurcation and regularity of nonlinear Burgers equation with mean value constraint[J]. Electronic Research Archive, 2025, 33(5): 2972-2988. doi: 10.3934/era.2025130

    Related Papers:

    [1] Qing Li, Junfeng He . Pattern formation in a ratio-dependent predator-prey model with cross diffusion. Electronic Research Archive, 2023, 31(2): 1106-1118. doi: 10.3934/era.2023055
    [2] Xianjun Wang, Huaguang Gu, Bo Lu . Big homoclinic orbit bifurcation underlying post-inhibitory rebound spike and a novel threshold curve of a neuron. Electronic Research Archive, 2021, 29(5): 2987-3015. doi: 10.3934/era.2021023
    [3] Xiaojun Huang, Zigen Song, Jian Xu . Amplitude death, oscillation death, and stable coexistence in a pair of VDP oscillators with direct–indirect coupling. Electronic Research Archive, 2023, 31(11): 6964-6981. doi: 10.3934/era.2023353
    [4] Junhai Ma, Hui Jiang . Dynamics of a nonlinear differential advertising model with single parameter sales promotion strategy. Electronic Research Archive, 2022, 30(4): 1142-1157. doi: 10.3934/era.2022061
    [5] 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
    [6] Cunbin An . Global bifurcation and structure of stationary patterns of a diffusive system of plant-herbivore interactions with toxin-determined functional responses. Electronic Research Archive, 2023, 31(4): 2095-2107. doi: 10.3934/era.2023107
    [7] Zhili Zhang, Aying Wan, Hongyan Lin . Spatiotemporal patterns and multiple bifurcations of a reaction- diffusion model for hair follicle spacing. Electronic Research Archive, 2023, 31(4): 1922-1947. doi: 10.3934/era.2023099
    [8] Bin Long, Shanshan Xu . Persistence of the heteroclinic loop under periodic perturbation. Electronic Research Archive, 2023, 31(2): 1089-1105. doi: 10.3934/era.2023054
    [9] Mufit San, Seyma Ramazan . A study for a higher order Riemann-Liouville fractional differential equation with weakly singularity. Electronic Research Archive, 2024, 32(5): 3092-3112. doi: 10.3934/era.2024141
    [10] Christos Sourdis . A Liouville theorem for ancient solutions to a semilinear heat equation and its elliptic counterpart. Electronic Research Archive, 2021, 29(5): 2829-2839. doi: 10.3934/era.2021016
  • In this paper, we focus on the steady-state bifurcation problem of the nonlinear Burgers equation within a bounded domain, considering both homogeneous Dirichlet boundary conditions and homogeneous Neumann boundary conditions with a mean value constraint. Unlike previous studies, we develop an enhanced turbulence model by incorporating nonlinear higher-order terms (such as u2 and u3) and linear source terms λu into the one-dimensional Burgers equation. Our steady-state bifurcation analysis establishes for the first time how the coupled forward energy cascade and inverse energy transfer mechanisms collectively govern the dynamics of initial flow instability. By combining the spectral theorem for a linear compact operator with the normalized Lyapunov–Schmidt reduction method and the implicit function theorem, we derive the complete criterion for the critical bifurcation condition, the explicit form of the bifurcation solution, and its regularity.



    Consider the nonlinear Burgers equation:

    {ut=γ2ux2+λuauux+bu2+cu3,(x,t)(0,π)×(0,),u(x,0)=ψ(x),x(0,π). (1.1)

    Here, (0,π) is a continuous bounded domain in R, and u(x,t) represents the fluid velocity at position x and time t. The term λu denotes the linear source term, where λ>0 is a system parameter. The term γ2ux2 represents the momentum diffusion effect caused by fluid viscosity, with γ>0 being the viscosity coefficient satisfying 0<γ<1. Furthermore, bu2 and cu3 describe the nonlinear effects that dominate the fluid dynamics, where b>0 and c>0 indicate that the nonlinear force increases with the flow velocity u. The nonlinear convective term auux characterizes the self-convection effect of the fluid velocity field, with a>0 representing the mechanism of energy transfer from large-scale vortices to small-scale vortices. Notably, when a=1 and λ=b=c=0, the equation reduces to the form introduced initially by Bateman in his paper [1], as shown below.

    {ut=γ2ux2uux,(x,t)(0,π)×(0,),u(x,0)=ψ(x),x(0,π). (1.2)

    Burgers [2] expanded on the mathematical modeling of turbulence using Eq (1.2), making a significant contribution to fluid mechanics. This equation, subsequently named the Burgers equation, is widely recognized as one of the most prominent models incorporating nonlinear propagation and diffusion effects when γ0, Eq (1.2) simplifies to the inviscid Burgers equation. Conversely, when u0, the viscous Burgers equation (1.2) can be converted to the linear heat equation.

    The Burgers equation is characterized by its nonlinear convection and diffusion terms. As a one-dimensional simplified model of the Navier–Stokes equations, it provides an essential theoretical tool for studying complex phenomena such as turbulence and shock waves. Widely applied in fluid mechanics [3], nonlinear acoustics [4] and gas dynamics [5]. Furthermore, its exact solutions and simple form make it a crucial benchmark for developing and validating numerical methods, particularly in turbulence and shock wave simulations under high Reynolds number conditions.

    Given that the exact solution of the Burgers equation is known, it can serve as a benchmark solution in numerical simulations of fluid dynamics, enabling the evaluation and comparison of numerical methods. Numerous studies have employed the Burgers equation for comparative analysis and improvement of numerical methods. These include classical approaches such as the Galerkin finite element method [6], cubic Hermite finite element method [7], standard finite element method [8], finite difference method [9], and finite volume method [10], along with more recent advancements. For instance, Kaur et al. [11] developed a compact finite difference scheme for the 1D nonlinear Burgers equation, achieving first-order temporal and fourth-order spatial accuracy, validated theoretically and numerically. Further advancing the field of numerical solutions, Zhang and Yu [12] developed a novel MQ quasi-interpolation meshless method, demonstrating superior accuracy to finite difference methods in solving Burgers' equation shock problems at high Reynolds numbers through theoretical convergence proofs and numerical validation. Similarly, Shi and Yang [13] first propose a temporal two-grid difference method for the nonlinear viscous Burgers equation. The Crank–Nicolson-based scheme proves more efficient than standard finite difference methods while maintaining rigorous L2 and L stability.

    Beyond these advancements in numerical methodologies, fundamental theoretical understanding of the Burgers equation's dynamical behavior has also been significantly advanced. For example, Ortiz et al. [14] integrated boundary layer theory with PINNS for viscous Burgers solutions, proving their capability to resolve both shock formation and viscous effects. Mouktonglang et al. [15] proved periodic solution uniqueness for the Rosenau–RLW–Burgers equation, showing viscous effects control convergence and oscillation dynamics. Li et al. [16] investigated the dynamic transition behavior of the generalized Burgers equation under one-dimensional periodic boundary conditions, demonstrating that the bifurcation type is determined by parameter b while elucidating the influence of length scale l, dispersion parameter δ, and viscosity coefficient υ on transition characteristics. These findings also offer insights into turbulence control mechanisms.

    Nevertheless, several critical questions remain unresolved. To our knowledge, the steady-state bifurcation behavior of nonlinear Burgers equations has not been thoroughly investigated—a research gap with significant implications for elucidating the laminar-to-turbulent transition. To better capture such flow characteristics, and inspired by Li's work [16], we introduce both nonlinear higher-order terms and linear source terms λu into the primitive Burgers equation.

    In mathematical physics, bifurcation refers to the fascinating phenomenon of an abrupt transition in the steady state of a dynamic system, triggered by changes in the system's parameters. Specifically, steady-state bifurcation arises from the study of the stability of nonlinear evolution equations. When the system parameters λ reach critical values, the system loses its original stability, leading to the emergence of new fixed points, limit cycles, and other dynamic behaviors. This theory is widely applied in fields such as chemistry, biology, ecology, and engineering, serving as an essential tool for analyzing changes in system stability.

    Recent advances in bifurcation research of nonlinear evolution equations demonstrate the universality of bifurcation theory. For instance, Zhang et al. [17] illustrated the existence of a double eigenvalue bifurcation for nonlinear equations with singularities that are fully degenerate of the second order and nondegenerate of the third order, employing the normalized Lyapunov–Schmidt reduction method. Similarly, Wei [18] employed the same methods to investigate a nonlinear parabolic system under nonlinear boundary conditions exhibiting steady-state bifurcation behavior. Guo [19] conducted a study on the steady-state bifurcation of Langford's PDE system emanating from both simple and double eigenvalues. By integrating central manifold theory, Guo further analyzed the direction of the Hopf bifurcation within the PDE system. Pan [20] applied dynamic transition theory to the problem of convection in couple-stress fluid-saturated porous media. The study yielded significant results, including the approximate bifurcation solution, the number of global attractors, and the finding that these attractors encompass four steady-state convection solutions. Wang [21] investigated the chemotaxis-fluid coupled model, while Chen et al. [22,23] further analyzed the existence and regularity of weak and classical solutions.

    Previous studies have focused on the steady-state bifurcation behavior in two-dimensional dynamical systems, while recent work has extended these investigations to one-dimensional settings. Ma et al. [24] employed topological degree theory and bifurcation theory to analyze the Robin problem for the mean curvature equation in a one-dimensional Minkowski space. They provided a calculation of the bifurcation curvature, which not only aids in understanding the existence of solutions for specific types of Minkowski-curvature equations in higher dimensions but also elucidates the properties of these solutions. Extending these analytical approaches to 1D reaction-diffusion systems. Taylan et al. [25] investigated one-dimensional non-self-adjoint reaction-diffusion equations, revealing a unique Hopf bifurcation under zero-mean boundary conditions, and further characterized the primary transition behavior in the Burgers equation.

    In this paper, we investigate the bifurcation problem of one-dimensional homogeneous boundary value problems related to the Burgers equation. We develop a simplified model to simulate turbulent behavior, which also provides a foundation for understanding high-dimensional turbulence properties. Fan and Feng [26,27] integrate linear stability analysis with nonlinear dynamical systems approaches, employing eigenvalue analysis and reduction techniques to elucidate parameter threshold-induced dynamic transition mechanisms and external field coupling effects. Building upon these investigations of dynamic transitions, parallel theoretical advances have been made in steady-state bifurcation research. The seminal works by Li et al. [28,29,30] established a bifurcation-theoretic framework employing Lyapunov–Schmidt reduction and eigenvalue analysis to characterize steady-state bifurcation phenomena in diffusion systems. Inspired by these findings, we examine the influence of nonlinear, higher-order viscous effects in turbulence on the emergence of steady-state bifurcation phenomena. Using the bifurcation and transition theory developed by Ma and Wang [31,32], we analyze the steady-state bifurcation of the nonlinear Burgers equation under various homogeneous boundary conditions.

    Considering the System (1.1) with Dirichlet boundary conditions as follows:

    {ut=γ2ux2+λuauux+bu2+cu3,(x,t)(0,π)×(0,),u(0)=0,u(π)=0,u(x,0)=ψ(x),x(0,π). (1.3)

    The Dirichlet boundary condition enforces zero fluid velocity at the boundary, corresponding physically to the no-slip condition between fluid and solid walls where complete momentum exchange occurs. Significantly, the steady-state bifurcation in System (1.3) captures the essential dynamics of the initial instability stage in wall-bounded turbulence [33], providing critical insights into the formation mechanisms of turbulent boundary layers.

    Researching the System (1.1) with mean value constraint Neumman boundary conditions as follows

    {ut=γ2ux2+λuauux+bu2+cu3,(x,t)(0,π)×(0,),uxx=0=0,uxx=π=0,π0udx=0,t(0,),u(x,0)=ψ(x),x(0,π), (1.4)

    The Neumann boundary condition with mean constraint specifies a zero gradient for the velocity field u at the boundary, while the mean constraint enforces a zero spatial average of u. In turbulent flow simulations, these mean-constrained Neumann conditions rigorously satisfy the mass conservation law for shear flows. By constraining the total system momentum, they effectively suppress numerical oscillations at high Reynolds numbers and significantly enhance solution stability. Importantly, the steady-state bifurcation in System (1.4) directly captures the shear flow instability [34], with this simplified model providing crucial theoretical insight into the onset of flow instabilities.

    The rest of the paper is organized as follows: Section 2 formulates the abstract operator equation for Eq (2.1) using bifurcation theory. Section 3 rigorously derives the explicit bifurcation solutions for the Burgers equation under two types of homogeneous boundary conditions and establishes their regularity. Building on these results, Section 4 reveals the universality of critical parameters, the stabilizing effect of dissipation terms, and the influence of higher-order nonlinear terms on laminar-turbulent transition under different boundary conditions, with numerical simulations demonstrating both subcritical and supercritical bifurcation scenarios.

    In this paper, we focus on the existence of a bifurcation solution of (1.1), which satisfies the following equation

    γd2udx2+λuaududx+bu2+cu3=0,x(0,π). (2.1)

    (2.1) with Dirichlet boundary and Neumann boundary conditions with mean value constraint, respectively, are as follows:

    {γd2udx2+λuaududx+bu2+cu3=0,x(0,π),u(0)=u(π)=0, (2.2)
    {γd2udx2+λuaududx+bu2+cu3=0,x(0,π),uxx=0=uxx=π=0,π0udx=0,int(0,). (2.3)

    First, we denote by L2(0,π) the Lebesgue space of square productible functions defined in (0,π); let H be the Hilbert space H=L2(0,π). We define H1 under the Dirichlet boundary and the Neumann boundary conditions with mean value constraint, respectively, as follows:

    H1={uH2[0,π]u(0)=u(π)=0},
    H1={uH2[0,π]π0u(x)dx=0,dudxx=0=dudxx=π=0}.

    Then, we define the linear operators Lλ=A+Bλ and the nonlinear operator G:H1H by

    Au=γd2udx2,Bλu=λu,G(u)=aududx+bu2+cu3.

    It is easy to see that Lλ is a completely continuous field.

    We thus obtain the equivalent operator equation of the Burgers equation (2.1) as follows:

    Lλu+G(u)=0. (2.4)

    Definition 2.1. [31] Suppose (0,λ),λR1 is a trivial solution of Eq (2.4). If there exists λ0R1 such that when λ<λ0 or λ>λ0, Eq (2.4) has a nontrivial solution (uλ,λ)(0,λ) and limλλ0(uλ,λ)=(0,λ0), limλλ0uλH1=0, then it is said that the Eq (2.4) undergoes a bifurcation solution at (0,λ0).

    Definition 2.2. [31] Let H and H1 be Hilbert spaces, and H1H being dense and compactly embedded. A linear operator Lλ:H1H is called a completely continuous field if

    {Lλ=A+Bλ:H1H,A:H1Hisalinearisomorphismwitheigenvalueshavingpositiverealparts,Bλ:H1H is a linear compact operator.

    Definition 2.3. [31] Let uλH1 be a bifurcation solution of Eq (2.4) at λ=λ0. The bifurcation solution is called regular, or nondegenerate, if the differential operator of Lλ+G(,λ) at uλ

    Lλ+DuG(uλ,λ):H1H,

    is a linear isomorphism for all 0<|λλ0|<ε, where ε is sufficiently small.

    Lemma 2.1. [31] Let L:H1H be a linear completely continuous field. Then the following conclusions hold:

    (i) if {λkk1}C are the eigenvalues (counting multiplicity) of L, then we can choose eigenvectors {ek}H1 of L and eigenvectors {ek}H1 of L such that

    ei,ejH{=0ij,0i=j;

    (ii) if ρ=λk==λnk(n1) is an eigenvalue of L with algebraic multiplicity m=n+1 and

    geometric multiplicity r=1, then for any nonzero constant σ0, we can choose eigenvectors {ek,,ek+n} of L and eigenvectors {ek,,ek+n} of L such that we have

    {Lek=ρek,Lek+1=ρek+1+σek,Lek+n=ρek+n+σek+n1,
    {Lek+n=ρek+n,Lek+n1=ρek+n1+σek+n,Lek=ρek+σek+1;

    (iii) H can be decomposed into the following direct sum of spaces

    H=¯E1¯E2,E1=span{ekk1},E2={υH1υ,ekH=0,k1},

    when ¯E1 and ¯E2 are the closures of E1 and E2 in H, respectively.

    (iv) E1 and E2 are invariant subspaces of L

    L:Ei¯Ei,i=1,2,

    Moreover, =LE2 has an inverse 1=¯E2E2¯E2, such that

    limnn1nH=0,u¯E2.

    (v)For any uH, there exists a generalized Fourier expansion as follows

    u=kxkek+υ,υ¯E2,xk=u,ek,H.

    In particular, if L:H1H has a complete spectrum, there exists a complete Fourier expansion as follows:

    u=kxkek,xk=u,ekH.

    Lemma 2.2. [31] Let xλ be the bifurcation equation of Eq (2.4), where

    L1λx+P1G(x+ϕ(x,λ),λ)=0, (2.5)

    has a bifurcation solution at λ=λ0. The bifurcation solution uλ=x+ϕ(x,λ) of Eq (2.3) is regular if and only if xλ is regular with respect to Eq (2.5).

    We are ready to state the main result and the proof process in this section. For the System (2.2), we have the following bifurcation theorem.

    Theorem 3.1. System (2.2) bifurcates a bifurcation solution from (u,λ)=(0,γ) under the Dirichlet boundary conditions, and the expression of the bifurcation solution is given by

    ˉu=3π8b(λγ)sinx+o(|λγ|2).

    Proof. We compute the eigenvalues and eigenvectors of Lλ.

    Let ρk(k=1,2) and ek(k=1,2) be the eigenvalues and eigenvectors of the following eigenvalue problem

    {d2ekdx2=ρkek,x(0,π),ek(0)=ek(π)=0,π0e2kdx=1. (3.1)

    We obtain the specific form of the eigenvalues ρk(k=1,2) and the corresponding eigenvectors ek(k=1,2) of Eq (3.1)

    ρk=k2,ek=2πsinkx.

    Hence, the eigenvalues and the corresponding eigenvectors of the operator Lλ in Eq (2.4) are

    {βk=λγρkk=1,2},{ek=2πsinkxk=1,2}.

    According to Lemma (2.1), it can be obtained that the eigenvectors {ekk=1,2} of Lλ form an orthogonal basis for H1. Therefore, it is easy to obtain that the first eigenvalue of Lλ and the corresponding eigenvector are

    β1(λ)=λγ,βj(γ)0,j2,e1=2πsinx.

    By the spectral theorem for the linear completely continuous fields (Lemma 2.1), we can decompose the space H1 and H in a neighborhood of λ=γ as follows:

    H1=E1E2,H=E1¯E2,

    where

    E1=span{e1},E2=span{e2,e3,},

    Then, the linear operator Lλ can be decomposed in a neighborhood of λ=γ as

    Lλ=L1λL2λ,L1λ:E1E1,L2λ:E2¯E2.

    Here, E1denotes the finite-dimensional kernel space of operatorLλ, whileE2represents its corresponding infinite-dimensional complementary subspace (with its closure denoted by ¯E2). Under this decomposition, the restricted operatorL1λacting on E1 constitutes a finite-dimensional linear operator that reduces to the zero operator at the critical parameter λ0, whereas the restricted operator L2λ acting on E2 forms a bounded linear operator that preserves its invertibility at λ0.

    We have u=u1+u2 with uH1, where u1E1 and u2E2. Now assume

    u1=x1e1,u2=j=2yjej,x1,yjR.

    When the linear operator Lλ possesses a nontrivial kernel E1 at the critical parameter λ0, the direct solution becomes problematic due to the operator's non-invertibility. To address this, we employ the Lyapunov–Schmidt reduction method, decomposing Eq (2.2) into two subproblems on the kernel space E1 and its complementary subspace E2. The equation on E2 is first solved using the implicit function theorem, thereby reducing the original problem to a finite-dimensional equation on E1. This method essentially transforms an infinite-dimensional problem into a finite-dimensional one through dimensional reduction.

    By applying the normalized Lyapunov–Schmidt reduction method, we obtain the bifurcation solution for Eq (2.2). Substituting u1 and u2 into Eq (2.1), we have

    β1(λ)x1a(x1e1+j=2yjej)(x1de1dx+j=2yjdejdx),e1+b(x1e1+j=2yjej)2,e1+c(x1e1+j=2yjej)3,e1=0, (3.2)

    and

    βj(λ)yja(x1e1+j=2yjej)(x1de1dx+j=2yjdejdx),ej+b(x1e1+j=2yjej)2,ej+c(x1e1+j=2yjej)3,ej=0,j2. (3.3)

    Let us get the approximate expression for the reduction Eq (3.3) as

    βj(λ)yjax21e1de1dx,ej+bx21e21,ej+(x21)=0,j2, (3.4)

    Note that

    e1de1dx,e2=12π,e1de1dx,ej=0,j3,e21,e1=823ππ,e21,ej=0,j2.

    Now, with Eq (3.3), we can calculate

    y2=ax21β1(λ)2π,yj=(x21),j3,

    substituting yj(j=2,3,) into Eq (3.2) yields

    β1(λ)x1+bx21(e21,e1+(x21)=0,

    namely,

    β1(λ)x1+82b3ππx21+(x21)=0, (3.5)

    the approximate equation corresponding to Eq (3.5) is

    β1(λ)x1+82b3ππx21=0, (3.6)

    therefore, Eq (3.6) has a bifurcation solution in a neighborhood of (x,λ) = (0,γ), which indicates that Eq (3.6) undergoes a bifurcation at (x,λ) = (0,γ), and the expression for the bifurcation solution branch is as follows:

    x1=3ππβ1(λ)82b,

    now, we give the expression for the bifurcation solution of Eq (2.2)

    ˉu=3π8b(λγ)sinx+(|λγ|2).

    Remark 3.1. Theorem 3.1 establishes that under Dirichlet boundary conditions, the Burgers equation exhibits a steady-state bifurcation at the critical parameter λ0=γ, yielding a bifurcated solution that provides a one-dimensional simplified representation of the steady streak velocity profile in wall-bounded turbulence. For λ<γ, the negative real part of the eigenvalue guarantees system stability, whereas when λ>γ, the laminar base state becomes unstable and transitions to a new steady flow configuration characterized by the sinx modal structure.

    Building upon the existence and analytical expressions of bifurcation solutions for System (2.2) established in Theorem 3.1, we further investigate their regularity properties. The regularity analysis of the bifurcation solution ˉu is of fundamental importance as it directly governs the spatiotemporal evolutionary dynamics of system solutions and numerical computation accuracy. We further study the regularity of the bifurcation solution ˉu of Eq (2.2), which gives the following conclusion.

    Theorem 3.2. Let x1 be a bifurcation solution of Eq (3.6) at λ=γ. So the bifurcation solution ˉu of Eq (3.3) is regular if and only if x1 is regular with respect to Eq (3.5).

    Proof. First, considering the regularity of the bifurcation solution of Eq (3.6), we consider the derivative of Eq (3.6) with respect to x1, as follows:

    β1(λ)+162b3ππx1.

    This shows that the bifurcation solution of Eq (3.6) is regular. Lemma 2.2 further establishes that the bifurcation solution ˉu of Eq (2.4) is also regular. The bifurcation solution ˉu of Eq (2.2) is regular.

    Then, we consider the Neumann boundary condition with mean value constraints. For system (2.3) we have the following bifurcation theorem.

    Theorem 3.3. System (2.3) bifurcates a bifurcation solution from (u,λ)=(0,γ) under the Neumann boundary condition with the mean value constraint, and the expression of the bifurcation solution is as follows:

    ˉu=3π4a(λγ)cosx+o(|λγ|2).

    Proof. By solving Lλ=A+Bλ for all eigenvalues and eigenvectors, following the proof of Theorem 3.1, we can obtain the eigenvalues and eigenvectors of Eq (2.4) as follows:

    {βk(λ)=λγk2k=1,2},{ek=2πcoskxk=1,2}.

    Through Lemma 2.1, we obtain that the eigenvectors {ekk=1,2} of Lλ form an orthogonal basis for H1. Hence, it can be easily obtained that

    β1(λ)=λγ,βj(γ)0,j2,

    the corresponding eigenvectors are

    e1=2πcosx.

    Using the spectral theorem (Lemma 2.1), we can decompose the space H1 and the operator Lλ in a neighborhood of λ=γ as follows:

    H1=E1E2,H=E1¯E2,

    where

    E1=span{e1},E2=span{e2,e3,},

    the linear operator Lλ can be decomposed near λ=γ as

    Lλ=L1λL2λ,L1λ:E1E1,L2λ:E2¯E2,

    Now, we let uH1, and u=u1+u2, and assume that

    u1=x1e1,u2=j=2yjej,xj,yjR,

    where u1E1 and u2E2.

    We use the Lyapunov-Schmidt reduction method to obtain a bifurcated solution to Eq (2.4). First we substitute u1 and u2 into Eq (1.4) to obtain

    β1(λ)x1a(x1e1+j=2yjej)(x1de1dx+j=2yjdejdx),e1+b(x1e1+j=2yjej)2,e1+c(x1e1+j=2yjej)3,e1=0, (3.7)

    and

    βj(λ)yja(x1e1+j=2yjej)(x1de1dx+j=2yjdejdx),ej+b(x1e1+j=2yjej)2,ej+c(x1e1+j=2yjej)3,ej=0,j2 (3.8)

    the approximate equation for Eq (3.8) is

    βj(λ)yj+a22ππx1yj+bx21e21,ej+o(x21)=0,j2,

    Note that

    e21,e2=12π,e21,ej=0,j3,

    hence, we solve for

    y2=bx212π[2πβ2λ+a(2π)32x1],yj=o(x21),j3,

    substitute yj(j=2,3,) into Eq (3.7), noting that

    e1de1dx,e1=423ππ,e21,e1=0,

    we have

    β1(λ)x1ax21e1de1dx,e1+o(x21)=0,

    that is

    β1(λ)x1+42a3ππx21+o(x21)=0, (3.9)

    the approximate equation corresponding to Eq (3.9) is

    β1(λ)x1+42a3ππx21=0. (3.10)

    Then, Eq (3.9) has a bifurcation solution in the neighborhood of (x,λ) = (0,γ). This indicates that Eq (3.9) undergoes a bifurcation at (x,λ) = (0,γ), and the expression for the bifurcation solution branch is as follows

    x1=3ππβ1(λ)42a,

    thus, we obtain the expression for the bifurcation solution of Eq (2.3)

    ˉu=3π4a(λγ)cosx+o(|λγ|2).

    Based on the above analysis, we have successfully proven Theorem 3.3.

    Remark 3.2. Theorem 3.3 establishes that the system undergoes bifurcation at the first eigenvalue, which consequently determines the stability threshold. The mean-constrained boundary condition plays a pivotal role in the Lyapunov–Schmidt reduction, enabling the exact determination of the bifurcation solution's analytical form. This solution captures the dynamical transition of confined shear flow from steady-state destabilization to finite-amplitude coherent mode formation.

    Regarding the regularity of the bifurcation solution ˉu of Eq (2.3), we have the following conclusions.

    Theorem 3.4. Let x1 be a bifurcation solution of Eq (3.9) at λ=γ. The bifurcation solution ˉu of Eq (2.4) is regular if and only if x1 is regular with respect to Eq (3.10).

    Proof. First, let us consider the regularity of the bifurcation solution of Eq (3.6). Then, we consider the derivative of Eq (3.6) with respect to x1, which is as follows:

    β1(λ)+82a3ππx1.

    Substituting the bifurcation solution of Eq (3.6) into the above expression yields (λγ). If a sufficiently small decentered neighborhood (λγ)0 of λ=γ holds, this shows that the bifurcation solution of Eq (3.6) is regular. According to Lemma 2.2, we prove that the bifurcation solution ˉu of Eq (2.4) is regular. Hence, the bifurcation solution ˉu of Eq (2.3) is regular.

    Based on the spectral theorem analysis of the linear fully continuous field, this paper uses the normalized Lyapunov–Schmidt reduction method to rigorously derive the expressions of the bifurcation solutions of the one-dimensional nonlinear Burgers equation under Dirichlet boundary conditions and Neumann boundary conditions with mean-value constraints and demonstrates the regularity of the bifurcation solutions. We find that the bifurcation points of Systems (2.2) and (2.3) are (0,γ), and the bifurcation solutions both maintain regularity. This also shows that when the system undergoes bifurcation, the critical parameter λ0 has universality independent of boundary conditions, and the dissipation term plays a stabilizing role under both boundary conditions.

    However, the expressions of the bifurcation solutions corresponding to the two types of boundaries are entirely different. From the structure of the bifurcation solutions, the sinx modal solution generated by the Dirichlet boundary reflects the characteristics of wall-constrained flow, with its amplitude showing an inverse proportionality to the coefficient b of the nonlinear square term clear manifestation of the energy dissipation role played by the u2 term. The cosx modal solution generated by the Neumann boundary with mean-value constraint describes the characteristics of confined shear flow, where the amplitude is regulated by the coefficient a of the convection term, revealing the energy transport process dominated by non-convection terms. In turbulent motion, the spatial modal differences of the bifurcation solutions indicate different transition paths: the Dirichlet boundary case corresponds to the formation of wall-turbulence streak structures [35], whereas the Neumann case corresponds to the evolution process of coherent structures in confined shear flow [36].

    To better visualize the dependence of the bifurcation solution ˉu on the system parameter λ, we set the viscous coefficient γ=0.5, nonlinear convection coefficient a=1, and higher-order nonlinear coefficients b=c=1. Based on the conclusions of Theorems 3.1 and 3.2, we employ MATLAB to generate three-dimensional plots illustrating the bifurcation behavior of solutions ˉu for Systems (2.2) and (2.3) at the critical parameter λ0, demonstrating both subcritical and supercritical bifurcation scenarios. Figure 1 presents a three-dimensional visualization of the System (2.2) dynamical transition: When λ<γ, viscous dissipation dominates, resulting in a smooth parabolic velocity profile characteristic of stable laminar flow; when λ>γ, the system develops sinx modulated streak structures with alternating high- and low-speed bands between near-wall regions and channel center, where nonlinear effects induce velocity profile distortion, accurately reproducing wall-turbulence features.

    Figure 1.  Dirichlet boundary condition: (a) subcritical λ<γ; (b) supercritical λ>γ.

    Figure 2 presents the three-dimensional visualization of the dynamical transition in System (2.3): when λ<γ, the velocity field maintains a uniform distribution; when λ>γ, the system develops an antisymmetric structure dominated by the cosx mode, generating streamwise vortex pairs and coherent shear-layer structures.

    Figure 2.  Neumann boundary condition: (a) subcritical λ<γ; (b) supercritical λ>γ.

    Under both boundary conditions, the solutions develop pronounced multi-scale features, which are characteristic hallmarks of turbulence onset. This process shares identical physical mechanisms with the transition phenomenon observed in classical fluid mechanics beyond critical Reynolds numbers. Notably, at γ=0.5, the corresponding critical equivalent Reynolds number Re=2 corresponds to the transition threshold range established in one-dimensional dynamical systems [37].

    It should be noted that the one-dimensional Burgers equation is inherently limited to modeling fundamental characteristics of one-dimensional turbulent bifurcation behavior. Nevertheless, these findings establish a theoretical foundation for investigating bifurcation mechanisms in high-dimensional turbulent systems.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    We are grateful to the editors and referees for their helpful comments and suggestions. This work is supported by the National Nature Science Foundation of China (NSFC) (Grant No. 11901408), the Sichuan Natural Science Youth Foundation (NSFYSE) (Grant No. 22NSFSC16338) and the Central University for Excellence in Research of Science and Technology (CERST) (Grant No. 2682022ZTPY063).

    The authors declare there are no conflicts of interest.



    [1] H. Bateman, Some recent researches on the motion of fluids, Mon. Weather Rev., 43 (1915), 163–170. https://doi.org/10.1175/1520-0493(1915)43 doi: 10.1175/1520-0493(1915)43
    [2] J. Burgers, A mathematical model illustrating the theory of turbulence, Adv. Appl. Mech., 1 (1948), 171–199. https://doi.org/10.1016/S0065-2156(08)70100-5 doi: 10.1016/S0065-2156(08)70100-5
    [3] J. D. Murray, On Burgers' model equations for turbulence, J. Fluid Mech., 59 (1973), 263–279. https://doi.org/10.1017/S0022112073001564 doi: 10.1017/S0022112073001564
    [4] N. T. Eldabe, E. M. Elghazy, A. Ebaid, Closed form solution to a second order boundary value problem and its application in fluid mechanics, Phys. Lett. A, 363 (2007), 257–259. https://doi.org/10.1016/j.physleta.2006.11.010 doi: 10.1016/j.physleta.2006.11.010
    [5] K. Fujisawa, A. Asada, Nonlinear parametric sound enhancement through different fluid layer and its application to noninvasive measurement, Measurement, 94 (2016), 726–733. https://doi.org/10.1016/j.measurement.2016.09.004 doi: 10.1016/j.measurement.2016.09.004
    [6] Y. L. Chen, T. Zhang, A weak Galerkin finite element method for Burgers' equation, Comput. Appl. Math., 348 (2019), 103–119. https://doi.org/10.1016/j.cam.2018.08.044 doi: 10.1016/j.cam.2018.08.044
    [7] M. A. Shallal, A. H. Taqi, B. F. Jumaa, H. Rezazadeh, M. Inc, Numerical solutions to the 1-D Burgers' equation by a cubic Hermite finite element method, Indian J. Phys., 96 (2022), 3831–3836. https://doi.org/10.1007/s12648-022-02304-4 doi: 10.1007/s12648-022-02304-4
    [8] Y. S. Uçar, N. Yağmurlu, İ. Çelikkaya, Operator splitting for numerical solution of the modified Burgers' equation using finite element method, Numer. Meth. Partial Differ. Equations, 35 (2019), 478–492. https://doi.org/10.1002/num.22309 doi: 10.1002/num.22309
    [9] L. Wang, H. Li, Y. Meng, Numerical solution of coupled Burgers' equation using finite difference and sinc collocation method, Eng. Lett., 29 (2021), 668–674.
    [10] J. Zhang, Q. Yang, The finite volume element method for time fractional generalized Burgers' equation, Fractal Fract., 8 (2024), 1–17. https://doi.org/10.3390/fractalfract8010053 doi: 10.3390/fractalfract8010053
    [11] K. Kaur, G. Singh, An efficient non-standard numerical scheme Coupled with a compact finite difference method to solve the one-dimensional Burgers' equation, Axioms, 12 (2023), 593–609. https://doi.org/10.3390/axioms12060593 doi: 10.3390/axioms12060593
    [12] J. Zhang, J. Yu, A Multi-Quadrics quasi-interpolation scheme for numerical solution of Burgers' equation, Appl. Numer. Math., 208 (2025), 38–44. https://doi.org/10.1016/j.apnum.2024.09.025 doi: 10.1016/j.apnum.2024.09.025
    [13] Y. Shi, X. Yang, A time two-grid difference method for nonlinear generalized viscous Burgers' equation, J. Math. Chem., 62 (2024), 1323–1356.
    [14] R. D. Ortiz, O. M. Nunez, A. M. M. Ramirez, Solving viscous Burgers' equation: Hybrid approach combining Boundary Layer theory and physics-informed neural networks, Mathematics, 12 (2024), 3430. https://doi.org/10.3390/math12213430 doi: 10.3390/math12213430
    [15] T. Mouktonglang, S. Yimnet, N. Sukantamala, B. Wongsaijai, Dynamical behaviors of the solution to a periodic initial-boundary value problem of the generalized Rosenau-RLW-Burgers equation, Math. Comput. Simul., 196 (2022), 114–136. https://doi.org/10.1016/j.matcom.2022.01.004 doi: 10.1016/j.matcom.2022.01.004
    [16] L. Li, K. W. Ong, Dynamic transitions of generalized Burgers equation, J. Math. Fluid Mech., 18 (2016), 89–102. https://doi.org/10.1007/s00021-015-0240-7 doi: 10.1007/s00021-015-0240-7
    [17] Q. Zhang, D. Yan, Z. Pan, Bifurcation from double eigenvalue for nonlinear equation with third-order nondegenerate singularity, Appl. Math. Comput., 220 (2013), 549–559. https://doi.org/10.1016/j.amc.2013.06.024 doi: 10.1016/j.amc.2013.06.024
    [18] D. Wei, S. Guo, Steady-state bifurcation of a nonlinear boundary problem, Appl. Math. Lett., 128 (2022), 107902. https://doi.org/10.1016/j.aml.2021.107902 doi: 10.1016/j.aml.2021.107902
    [19] G. Guo, X. Wang, X. Lin, M. Wei, Steady-state and Hopf bifurcations in the Langford ODE and PDE systems, Nonlinear Anal. Real World Appl., 34 (2017), 343–362. https://doi.org/10.1016/j.nonrwa.2016.09.008 doi: 10.1016/j.nonrwa.2016.09.008
    [20] Z. Pan, L. Jia, Y. Mao, Q. Wang, Transitions and bifurcations in couple stress fluid saturated porous media using a thermal non-equilibrium model, Appl. Math. Comput., 415 (2022), 126727. https://doi.org/10.1016/j.amc.2021.126727 doi: 10.1016/j.amc.2021.126727
    [21] Y. Wang, The Progress of Chemotaxis-fluid Models, Journal of Xihua University (Natural Science Edition), 35 (2016), 30–34, 38. https://doi.org/10.3969/j.issn.1673-159X.2016.04.006
    [22] J. Chen, Y. Han, Global Weak Solvability for a Chemotaxis-Fluid Model with Low Regular Initial Data, Journal of Xihua University (Natural Science Edition), 43 (2024), 97–105. https://doi.org/10.12198/j.issn.1673-159X.5354
    [23] W. Kuang, Y. Dong, Global Well-posedness of 2D Chemotaxis-fluid System, Journal of Xihua University (Natural Science Edition), 43 (2024), 103–108. https://doi.org/10.12198/j.issn.1673-159X.5233
    [24] R. Ma, L. Wei, Z. Chen, Evolution of bifurcation curves for one-dimensional Minkowski-curvature problem, Appl. Math. Lett., 103 (2020), 106176. https://doi.org/10.1016/j.aml.2019.106176 doi: 10.1016/j.aml.2019.106176
    [25] T. Sengul, B. Tiryakioglu, Dynamic transitions and bifurcations of 1D reaction-diffusion equations: The non-self-adjoint case, Math. Anal. Appl., 523 (2023), 127114. https://doi.org/10.1016/j.jmaa.2023.127114 doi: 10.1016/j.jmaa.2023.127114
    [26] Z. Feng, Y. Su, Traveling wave phenomena of inhomogeneous half-wave equation, J. Differ. Equations, 400 (2024), 248–277. https://doi.org/10.1016/j.jde.2024.04.029 doi: 10.1016/j.jde.2024.04.029
    [27] Y. Fan, L. Li, Z. Pan, Q. Wang, On dynamics of double-diffusive magneto-convection in a non-Newtonian fluid layer, Math. Methods Appl. Sci., 46 (2023), 14596–14621. https://doi.org/10.1002/mma.9337 doi: 10.1002/mma.9337
    [28] S. Li, J. Wu, H. Nie, Steady-state bifurcation and Hopf bifurcation for a diffusive Leslie–Gower predator–prey model, Comput. Math. Appl., 70 (2015), 3043–3056. https://doi.org/10.1016/j.camwa.2015.10.017 doi: 10.1016/j.camwa.2015.10.017
    [29] M. Wei, J. Wu, G. Guo, Steady state bifurcations for a glycolysis model in biochemical reaction, Nonlinear Anal. Real World Appl., 22 (2015), 155–175. https://doi.org/10.1016/j.nonrwa.2014.08.003 doi: 10.1016/j.nonrwa.2014.08.003
    [30] L. Ma, Z. Feng, Stability and bifurcation in a two-species reaction–diffusion–advection competition model with time delay, Nonlinear Anal. Real World Appl., 61 (2021), 103327. https://doi.org/10.1016/j.nonrwa.2021.103327 doi: 10.1016/j.nonrwa.2021.103327
    [31] T. Ma, S. Wang, Stability and Bifurcation of Nonlinear Evolution Equations, 1nd edition, Science Press, China, 2007.
    [32] T. Ma, S. Wang, Bifurcation theory and applications, World Sci, Singapore, 2005.
    [33] W. Cao, X. Shan, S. Tang, W. Ouyang, W. Zhang, Solving parametric high-Reynolds-number wall-bounded turbulence around airfoils governed by Reynolds-averaged Navier-Stokes equations using time-stepping-oriented neural network, Phys. Fluids, 37 (2025), 015151. https://doi.org/10.1063/5.0245918 doi: 10.1063/5.0245918
    [34] B. Tripathi, P. W. Terry, A. E. Fraser, E. G. Zweibel, M. J. Pueschel, Three-dimensional shear-flow instability saturation via stable modes, Phys. Fluids, 35 (2023), 105151. https://doi.org/10.1063/5.0167092 doi: 10.1063/5.0167092
    [35] N. Ciola, P. D. Palma, J. C. Robinet, S. Cherubini, Large-scale coherent structures in turbulent channel flow: a detuned instability of wall streaks, Fluid Mech., 997 (2024), A18. https://doi.org/10.1017/jfm.2024.726 doi: 10.1017/jfm.2024.726
    [36] J. Zhu, Y. Peng, W. Zhang, S. Niu, Y. Huang, Y. Li, et al. Research on the high-integration vortex-shape flow channel of a hydraulic servo valve based on additive manufacturing, in CSAA/IET International Conference on Aircraft Utility Systems (AUS 2024), 2024 (2025), 1700–1705. https://doi.org/10.1049/icp.2024.3137
    [37] B. Eckhardt, H. Faisst, A. Schmiegel, T. M. Schneider, Dynamical systems and the transition to turbulence in linearly stable shear flows, Phil. Trans. R. Soc. A, 366 (2008), 1297–1315. https://doi.org/10.1098/rsta.2007.2132 doi: 10.1098/rsta.2007.2132
  • Reader Comments
  • © 2025 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(162) PDF downloads(31) Cited by(0)

Figures and Tables

Figures(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog