You are currently viewing a new version of our website. To view the old version click .
Symmetry
  • Article
  • Open Access

10 November 2025

Symmetry, Conservation Law, Uniqueness and Stability of Optimal Control and Inverse Problems for Burgers’ Equation

,
and
1
School of Science, Jiangsu University of Science and Technology, Zhenjiang 212003, China
2
Zhenjiang Jizhi Ship Technology Co., Ltd., Zhenjiang 212003, China
3
Yangzijiang Shipbuilding Group, Taizhou 212299, China
*
Author to whom correspondence should be addressed.
This article belongs to the Section Mathematics

Abstract

This paper tackles the ill-posed inversion of initial conditions and diffusion coefficient for Burgers’ equation with a source term. Using optimal control theory combined with a finite difference discretization scheme and a dual-functional descent method (DFDM), it sets the unknown boundary function g ( τ ) and diffusion coefficient u as control variables to build a multi-objective functional, proving the existence of the optimal solution via the variational method. Symmetry analysis reveals the intrinsic connection between the equation’s Lie group invariances and conservation laws through Noether’s theorem, providing a natural regularization framework for the inverse problem. Uniqueness and stability are demonstrated by the adjoint equation under cost function convexity. An energy-consistent discrete scheme is created to verify the energy conservation law while preserving the underlying symmetry structure. A comprehensive error analysis reveals dual error sources in inverse problems. A multi-scale adaptive inversion algorithm incorporating symmetry considerations achieves high-precision recovery under noise: boundary error < 1 % , energy conservation error 0.13 % . The symmetry-aware approach enhances algorithmic robustness and maintains physical consistency, with the solution showing linear robustness to noise perturbations.

1. Introduction

The study of conservation laws and symmetry properties in partial differential equations has been fundamental to mathematical physics since Emmy Noether’s seminal 1918 theorem. Noether’s theorem establishes a profound connection: every continuous symmetry of a physical system corresponds to a conserved quantity, and conversely, every conserved quantity arises from an underlying continuous symmetry. This principle provides the theoretical foundation for understanding energy conservation in differential equations through time translation symmetry. Inverse problems for partial differential equations are fundamental across numerous scientific domains. In geophysics, parameter identification techniques are essential for subsurface characterization and seismic inversion []. Medical imaging applications rely heavily on inverse problem formulations for tomographic reconstruction and image enhancement []. Aerospace engineering utilizes inverse methods for design optimization and system identification in complex fluid-structure interactions []. In fluid mechanics, inverse problems naturally arise in determining flow parameters and boundary conditions from limited measurements []. In contrast to the forward problem [], where one seeks the state of a system given all governing equations, boundaries, and initial data. The inverse problem [] asks to recover unknown inputs or parameters from partial or indirect observations of the system’s response. According to the mathematician Keller from Stanford University [], a pair of problems is said to be inverse to each other if the formulation (known data) of one problem requires (partial) information from the solution of the other problem. If one is called the direct problem, the other is called the inverse problem. Closely related to the inverse problem is the well-posedness of the problem. The concept of well-posedness, introduced by Hadamard [] in 1923, defines a problem that must possess three properties: a solution must exist, that solution must be unique, and it must depend continuously on the input data.
The inverse problem of Burgers’ equation [] in fluid mechanics is examined in this work, as detailed below:
v x v v τ u 2 v τ 2 = f ( x , τ ) , ( x , τ ) Q = [ 0 , l ] × [ 0 , T ] , v ( 0 , τ ) = g ( τ ) , τ [ 0 , T ] , v ( x , 0 ) = v ( x , T ) = 0 , x [ 0 , l ] ,
where g ( τ ) denotes the time-varying boundary control, and f ( x , τ ) is the distributed source term. Burgers’ equation in Equation (1) follows the standard convection-diffusion form where the convective term v v τ represents nonlinear wave steepening, and the diffusion term u 2 v τ 2 provides viscous smoothing. The parameter u > 0 is the kinematic viscosity coefficient.
Give the additional condition v ( l , τ ) = q ( τ ) , where q ( τ ) is a known function, and the unknown parameters ( g , u ) along with the state variable v in Equation (1) need to be recovered. In the context of fluid mechanics and Burgers’ equation, Boundary control g ( τ ) represents time-varying inflow conditions, injection/suction profiles, or prescribed velocity/temperature at domain boundaries. In practical applications, this corresponds to controllable physical inputs such as inlet velocity profiles in channel flows, heat flux variations in thermal systems, and injection rates in porous media flow. The Diffusion Parameter Control u represents material properties or process parameters that can be spatially or temporally adjusted: viscosity in fluid systems, thermal conductivity in heat transfer, and permeability in subsurface flow.
The simultaneous optimization of both parameters allows for inverse characterization of coupled transport phenomena where boundary dynamics and material properties must be jointly identified from limited measurements. This approach is particularly valuable in applications where direct measurement of these quantities is difficult or impossible. While existing literature on inverse problems for Burgers’-type equations primarily addresses single-parameter identification, the multi-parameter inversion problem under physical constraints, such as energy conservation, remains relatively unexplored. Burgers’ equation possesses inherent Lie group symmetries that are intimately connected to its conservation laws [,] through Noether’s theorem, providing a natural framework for regularization. Our key innovation lies in a novel multi-objective optimization framework that rigorously integrates energy conservation laws and exploits the underlying symmetry structure as physical constraints. This symmetry-aware approach uniquely resolves the coupling between boundary conditions and diffusion parameters while maintaining the equation’s fundamental invariance properties, providing theoretical error bounds under discretization and perturbations. It bridges numerical implementations with physical consistency for nonlinear ill-posed problems. The inverse problem for Burgers’ equation exhibits the characteristic features of ill-posedness in Hadamard’s sense: (i) Non-uniqueness: Multiple parameter pairs ( g , u ) may produce similar boundary observations q ( τ ) due to the nonlinear coupling term v v τ ; (ii) Instability: Small perturbations in the observed data q ( τ ) can lead to large variations in the recovered parameters, particularly for the diffusion coefficient u; (iii) Non-existence: For inconsistent or noisy data, exact solutions may not exist in the classical sense. The nonlinearity v v τ creates a fundamental coupling between the unknown boundary condition g ( τ ) and the interior dynamics, while the diffusion term u 2 v τ 2 introduces additional parameter sensitivity that compounds the ill-conditioning.
Numerous regularization techniques have been developed to cope with ill-posed inverse problems. The seminal Tikhonov regularization [,] adds a stabilizing penalty on the solution norm, guaranteeing well-posedness at the cost of possible over-smoothing. In the context of parabolic and hyperbolic PDEs, theoretical advances in conditional stability have been established through rigorous mathematical analysis []. In contrast, computational stability bounds have been derived for specific PDE classes []. Practical algorithmic developments include iterative regularization methods that balance convergence speed with solution stability []. Carleman-weight approaches provide theoretical frameworks for quantifying solution uniqueness under limited data []. Variational formulations offer robust mathematical foundations for inverse PDE problems []. Recent extensions to fractional partial differential equations have opened new research directions in anomalous diffusion modeling [], memory-dependent processes [], and non-local inverse problems []. However, the specific case of Burgers’ equation poses additional challenges: the nonlinearity v v τ couples the unknown boundary with the interior dynamics, and energy dissipation through the term u v τ τ requires careful handling to preserve physical consistency.
In this work, we cast the inversion of ( g , u ) as an optimal control problem [,,] in a Sobolev space setting. We augment the classical data-misfit functional, measuring the discrepancy between the computed outflow v ( l , τ ; g , u ) and the observations q ( τ ) with two regularization terms: one enforcing H 1 smoothness of the boundary profile g, and another penalizing deviation of u from a prior estimate u 0 . Our contributions are threefold and as follows:
  • Theoretical framework. We prove the existence of a minimizer ( g , u ) using the direct method in calculating variations and establish necessary first-order optimality conditions via an adjoint equation. Under strict convexity of the regularized cost function, we further show the uniqueness of the control pair and continuous dependence on perturbations in the data and source term.
  • Energy-consistent discretization. Motivated by the physical energy conservation law [] of Burgers’ equation, we derive a discrete energy balance that equates the spatial change im the time-integrated squared state to the combined contributions of the source forcing and diffusion dissipation. Our detailed error analysis decomposes the conservation law error into two distinct sources: state variable error propagation ( ε c + ε e ) and energy discretization truncation error, yielding a total bound of ε i ε c + ε e . This explains why observation noise limits conservation accuracy even with extreme grid refinement. Among them, ε c represents all the errors generated during the discretization process, and ε e represents the errors generated during the iterative process. These two types of errors can be studied separately. This makes the sources of errors in our numerical experiments clearer.
  • Efficient two-stage optimization with physical guidance. To accelerate convergence and improve robustness against noise, we introduce a two-stage procedure: an initial coarse grid search guided by physical priors, followed by a nested, gradient-based optimization (interior-point then SQP) that refines boundary shape and diffusion coefficient. Our error analysis demonstrates that this approach effectively minimizes both error components. Numerical experiments on synthetic noisy data demonstrate that the method recovers g, u, the full state v ( x , τ ) , and the effective source term with relative L 2 errors below 1–2%, even under moderate observation noise, validating our theoretical error bounds.
While recent studies have addressed inverse problems for Burgers’-type equations, most focus on single-parameter identification. Our work distinguishes itself through several key innovations. Comparison with Recent Works: Traditional approaches use standard Tikhonov regularization without exploiting the symmetry structure. Machine learning methods lack theoretical guarantees for well-posedness. Energy-based methods consider only a discrete energy balance without symmetry preservation. Our Novel Contributions: Multi-parameter inversion framework. Unlike existing single-parameter methods, we simultaneously recover boundary conditions g ( τ ) and diffusion coefficient u with theoretical guarantees. Symmetry-guided regularization. First application of Lie group symmetries as natural constraints for Burgers’ equation inverse problems. Dual error decomposition. Novel theoretical framework separating discretization and iterative errors with rigorous bounds.
The rest of the paper is organized as follows. Section 2 formulates the discrete optimal control problem, states the main existence and uniqueness theorems, derives the adjoint equations and necessary optimality conditions, and discusses discrete energy conservation and its use for numerical verification. Section 3 presents detailed multi-scale inversion results, error analyses, and convergence studies. Finally, Section 4 offers concluding remarks and perspectives for extending the approach to more general nonlinear PDEs.
Remark 1
(Boundary Condition Justification). The homogeneous terminal conditions v ( x , 0 ) = v ( x , T ) = 0 are physically motivated by the assumption of quiescent initial and final states in controlled flow scenarios. This transforms the problem into an Initial-Boundary Value Problem (IBVP) where the system evolves from rest, reaches a controlled state through g ( τ ) , and returns to rest. This setup is common in flow control applications where transient behavior is studied over finite time horizons.

2. Optimal Control Theory

Since the inverse problem of Burgers’ equation is inherently ill-posed, an optimal control framework must be introduced to stabilize the solution. Burgers’ equation exhibits scaling and translational symmetries that provide natural constraints for the optimization problem. This section first constructs the optimal control problem J ( g , u ) , proves the existence of its optimal solution, then introduces adjoint variables (Lagrangian multipliers) to derive the optimality conditions, incorporating symmetry considerations to enhance well-posedness, and finally proves the uniqueness and stability of the solution, providing a theoretical foundation for subsequent numerical studies.

2.1. Lie Group Symmetries of Burgers’ Equation

Burgers’ equation admits several Lie point symmetries,
1.
Time translation: ( x , τ , v ) ( x , τ + c , v )
2.
Scaling symmetry: ( x , τ , v ) ( λ x , λ 2 τ , v )
3.
Galilean transformation: ( x , τ , v ) ( x + c τ , τ , v + c )
These symmetries form a Lie group that preserves the equation’s structure and provides natural constraints for the inverse problem regularization.

2.2. Optimal Control

The multi-objective functional J ( g , u ) is designed to balance three competing requirements: data fidelity, boundary smoothness, and parameter regularization. There exists ( g , u ) A 1 × A 2 ; we define the objective function
J ( g , u ) = 1 2 0 T | v ( l , τ ; g , u ) q ( τ ) | 2 d τ + α 2 0 T | g ( τ ) | 2 d τ + β 2 | u u 0 | 2 ,
where α , β > 0 are regularization parameters, M > 0 is a bound ensuring physical realizability, and the admissible sets are A 1 = g H 1 ( 0 , T ) : g { H 1 ( 0 , T ) } M and A 2 = u R + : u min u u max . Here, H 1 ( 0 , T ) denotes the Sobolev space of functions with square-integrable first derivatives, equipped with the norm g { H 1 ( 0 , T ) } = 0 T | g ( τ ) | 2 + | g ( τ ) | 2 d τ 1 / 2 . The prior estimate u 0 represents physical knowledge about the expected diffusion coefficient, typically derived from material properties or previous experimental data. The first term 1 2 0 T | v ( l , τ ; g , u ) q ( τ ) | 2 d τ measures the L 2 discrepancy between the computed terminal condition and observed data, ensuring consistency with available measurements. The second term α 2 0 T | g ( τ ) | 2 d τ enforces H 1 smoothness on the boundary profile g ( τ ) , preventing oscillatory artifacts common in inverse problems. The third term β 2 | u u 0 | 2 provides Tikhonov-type regularization for the diffusion coefficient, stabilizing the solution around a physically reasonable prior estimate u 0 . This formulation implicitly preserves the equation’s symmetry structure by maintaining invariance under the natural transformations of Burgers’ equation.
Theorem 1
(Optimal Control). There exists a control pair ( g , u ) A 1 × A 2 minimizing J ( g , u ) ,
J ( g , u ) = min ( g , u ) A 1 × A 2 J ( g , u ) .
Proof of Theorem 1.
The objective functional J ( g , u ) is composed of squared integral terms, each of which is non-negative.
J ( g , u ) 0 , ( g , u ) A 1 × A 2 .
Therefore, ( g , u ) must have a lower bound on the domain A 1 × A 2 . Let ( g n , u n ) A 1 × A 2 be a minimizing sequence of the objective functional J ( g , u ) , satisfying
lim n J ( g n , u n ) = inf ( g , u ) A 1 × A 2 J ( g , u ) .
Since each term of J ( g n , u n ) is non-negative, we can deduce that
0 T | g ( τ ) | 2 d τ 2 α J ( g n , u n ) , | u u 0 | 2 2 β J ( g n , u n ) .
Thus, with bounded J ( g n , u n ) , both g n in H 1 [ 0 , T ] and u n in A 2 are bounded. By the Banach-Alaoglu theorem [], there exists a weakly convergent subsequence
g n g H 1 ( 0 , T ) , u n u R .
By the weak lower semi-contivity of J ( g , u ) ,
inf ( g , u ) A 1 × A 2 J ( g , u ) = lim n J ( g n , u n ) J ( g , u ) .
Therefore, there exists an optimal solution ( g , u ) that satisfies
J ( g , u ) = min ( g , u ) A 1 × A 2 J ( g , u ) .
   □
Theorem 2
(Necessary condition). If ( g , u ) is the solution to the optimal control problem J ( g , u ) , and v satisfies the constraint conditions of Burgers’ equation. There exists an adjoint variable ξ such that ( v , ξ , g , u ) satisfies the following system:
v x v v τ u 2 v τ 2 = f ( x , τ ) , ( x , τ ) Q , v ( 0 , τ ) = g ( τ ) , τ [ 0 , T ] , v ( x , 0 ) = v ( x , T ) = 0 , x [ 0 , l ] , ξ x v ξ τ + u 2 ξ τ 2 = 0 , ( x , τ ) Q , ξ ( l , τ ) = v ( l , τ ) q ( τ ) , τ [ 0 , T ] , ξ ( x , 0 ) = ξ ( x , T ) = 0 , x [ 0 , l ] ,
with the optimality conditions
g ( τ ) = ξ ( 0 , τ ) α , u = u 0 1 β 0 T 0 l 2 v τ 2 ξ ( x , τ ) d x d τ .
Proof of Theorem 2.
Introduce the adjoint variable ξ ( x , τ ) and construct the Lagrangian functional
L = 0 T 0 l ξ v x u 2 v τ 2 v v τ f d x d τ + J ( g , u ) ,
where the objective functional is
J ( g , u ) = 1 2 0 T | v ( l , τ ) q ( τ ) | 2 d τ + α 2 0 T | g ( τ ) | 2 d τ + β 2 | u u 0 | 2 .
At the optimal solution ( g , u , v ) , we compute variations. Variation with respect to v,
δ v L = 0 T 0 l ξ ( δ v ) x u 2 ( δ v ) τ 2 v ( δ v ) τ δ v v τ d x d τ + 0 T [ v ( l , τ ) q ( τ ) ] δ v ( l , τ ) d τ .
Applying integration by parts
0 l ( ξ δ v ) x d x = [ ξ δ v ] 0 l 0 l ξ x δ v d x = ξ ( l , τ ) δ v ( l , τ ) ξ ( 0 , τ ) δ v ( 0 , τ ) 0 l ξ x δ v d x , 0 T ξ 2 ( δ v ) τ 2 d τ = ξ ( δ v ) τ 0 T ξ τ δ v 0 T + 0 T 2 ξ τ 2 δ v d τ .
Using boundary conditions δ v ( x , 0 ) = δ v ( x , T ) = 0 and ξ ( x , 0 ) = ξ ( x , T ) = 0 ,
δ v L = 0 T 0 l ξ x u 2 ξ τ 2 + v ξ τ δ v d x d τ + 0 T ξ ( l , τ ) + v ( l , τ ) q ( τ ) δ v ( l , τ ) d τ 0 T ξ ( 0 , τ ) δ v ( 0 , τ ) d τ .
Setting δ v L = 0 for arbitrary δ v gives
ξ x + u 2 ξ τ 2 v ξ τ = 0 ξ ( l , τ ) = q ( τ ) v ( l , τ ) ξ ( x , 0 ) = ξ ( x , T ) = 0
Using v ( 0 , τ ) = g ( τ ) so δ v ( 0 , τ ) = δ g ( τ ) ,
δ g L = 0 T ξ ( 0 , τ ) δ g ( τ ) d τ + α 0 T g ( τ ) δ g ( τ ) d τ = 0 g ( τ ) = ξ ( 0 , τ ) α .
Variation with respect to u,
δ u L = 0 T 0 l 2 v τ 2 ξ δ u d x d τ + β ( u u 0 ) δ u = 0 u = u 0 + 1 β 0 T 0 l 2 v τ 2 ξ d x d τ .
   □
Condition 1.
The terminal observation term 1 2 0 T | v ( l , τ ; g , u ) q ( τ ) | 2 d τ is convex with respect to ( g , u ) when the forward operator v ( l , τ ; g , u ) satisfies
2 J 1 g 2 λ 1 I , 2 J 1 u 2 λ 2 > 0 .
where λ 1 , λ 2 are positive constants depending on the regularization parameters.
Condition 2
(Regularization Terms). The H 1 regularization α 2 0 T | g ( τ ) | 2 d τ and Tikhonov term β 2 | u u 0 | 2 are inherently strictly convex
2 J 2 g 2 = α I > 0 , 2 J 3 u 2 = β > 0 .
Condition 3
(Combined Convexity). The total Hessian matrix satisfies
H J = 2 J = α g 2 g 0 0 β λ min I ,
where λ min = min ( α , β ) > 0 ensures positive definiteness.
Theorem 3
(Uniqueness and Stability). If the objective functional J ( g , u ) is strictly convex with respect to the control variables ( g , u ) . The control variables uniquely determine the state variable g ( τ ) , and the regularization parameters α > 0 and β > 0 ; then the optimal solution ( g , u ) is unique. In addition, when there are small perturbations in the source term f ( x , τ ) and the terminal value g ( τ ) , we have
g ϵ g H 1 ( 0 , T ) + | u ϵ u | + v ϵ v L 2 ( q ) c 0 ϵ .
Proof of Theorem 3. 
The strict convexity established in Conditions 1–3 above, combined with the compactness of the admissible sets A 1 A 2 , ensures that since J ( g , u ) is the sum of convex functions for each term, it is strictly convex with respect to ( g , u ) . This strict convexity implies that if there exist two distinct solutions ( g 1 , u 1 ) and ( g 2 , u 2 ) , then J ( g 1 , u 1 ) J ( g 2 , u 2 ) , contradicting their optimality. Therefore, ( g , u ) is unique.
Consider the perturbations of the source term f and the target terminal value q ( τ ) ,
f ϵ ( x , τ ) = f ( x , τ ) + ϵ f 1 ( x , τ ) , q ϵ ( τ ) = q ( τ ) + ϵ q 1 ( τ ) ,
where ϵ > 0 represents the magnitude of the perturbation. The corresponding objective functional is
J ( g ϵ , u ϵ ) = 1 2 0 T v ( l , τ ; g , u ) q ϵ ( τ ) 2 d τ + α 2 0 T g 2 ( τ ) d τ + β 2 ( u u 0 ) 2 .
The optimal solution after the perturbation is ( g ϵ , u ϵ , v ϵ ) . In the objective functional J ( g ϵ , u ϵ ) , since q ϵ ( τ ) = q ( τ ) + ϵ q 1 ( τ ) , the first term can be expanded as
1 2 0 T v ( l , τ ; g , u ) q ϵ ( τ ) 2 d τ = 1 2 0 T v ( l , τ ; g , u ) q ( τ ) ϵ q 1 ( τ ) 2 d τ = 1 2 0 T v ( l , τ ; g , u ) q ( τ ) 2 d τ + ϵ 2 2 0 T q 1 2 ( τ ) d τ ϵ 0 T v ( l , τ ; g , u ) q ( τ ) q 1 ( τ ) d τ .
When ϵ 0 , the objective functional J ( g ϵ , u ϵ , v ϵ ) J ( g , u , v ) . Therefore, the objective function is continuous concerning the perturbation.
Since the objective functional J ( g , u , v ) is strictly convex, its minimum point ( g , u , v ) is unique. According to the theory of variational analysis [], if J ( g ϵ , u ϵ , v ϵ ) is continuous for f ϵ and q ϵ , then its minimum point ( g ϵ , u ϵ , v ϵ ) also continuously depends on the input data concerning the perturbation.
Specifically, there exists a constant c 0 > 0 such that
g ϵ g H 1 ( 0 , T ) + | u ϵ u | + v ϵ v L 2 ( q ) c 0 ϵ ,
where c 0 depends on the regularization parameters α , β and the boundary conditions of the problem.    □
Compared to classical results, Tikhonov only guarantees existence for linear problems, and Lions establishes necessary conditions without stability bounds. Our Contribution: simultaneous uniqueness + quantitative stability estimates for Burgers’ nonlinear equation. The multi-parameter framework handles coupled ( g , u ) identification, unlike single-parameter studies. Symmetry integration exploits Lie group structure for enhanced well-posedness. Maintains energy conservation during parameter recovery.
The above description shows that u e converges to the unique solution u . In numerical experiments, we often use the discrete solution u c . We need to introduce some corresponding estimations to study the relationship among these three quantities further.
Lemma 1
(Discretization error []). Let ( g c , u c ) be the solution of the continuous optimal control problem, and ( g ϵ , v ϵ ) be the solution of the discrete optimal control problem with the grid size e i = ( Δ x , Δ τ ) . Then, when the condition u Δ τ ( Δ x ) 2 1 2 is satisfied, there exist constants a i such that
g c g ϵ 2 a 1 ( Δ x + Δ τ ) , | u c u ϵ | a 2 ( Δ x + Δ τ ) , v c v ϵ 2 a 3 ( Δ x + Δ τ ) .
Among them, the constant a i depends on the regularization parameters α, β, and the boundary data of the problem.
Corollary 1
(System error). For Burgers’ equation, if the solution v ( x , τ ) is continuously differentiable and satisfies the boundary conditions v ( x , 0 ) = v ( x , T ) = 0 , then its energy satisfies the conservation law
g c g 2 + | u c u | + v c v 2 ϵ c + ϵ e ,
where ϵ c is the error caused by the discrete scheme, and ϵ c c 1 ( Δ x + Δ τ ) . ϵ e is the error generated by the iterative process, and ϵ e c 0 ε .
Error Components—throughout this paper, we consistently use the following:
  • ε c (Discretization Error): Errors arising from spatial-temporal grid discretization, bounded by ε c c 1 ( Δ x + Δ τ ) .
  • ε e (Iterative Error): Errors from algorithmic convergence and data perturbations, bounded by ε e c 0 ε where ε is the perturbation level.
Total Error Bound: ε i ε c + ε e for all numerical computations.

3. Energy Conservation Law

Energy conservation laws play a fundamental role in the analysis and numerical solution of partial differential equations, providing both theoretical insights and practical computational benefits. For inverse problems, energy conservation serves multiple purposes: (i) it provides natural regularization constraints that improve problem conditioning [], (ii) it offers quantitative measures for solution quality assessment, and (iii) it enables the development of structure-preserving numerical schemes that maintain physical consistency during discretization. In the context of Burgers’ equation, Noether’s theorem establishes a direct connection between the equation’s time-translation symmetry and energy conservation, creating a theoretical framework that we exploit for enhanced stability in parameter identification. This symmetry-conservation relationship is particularly valuable for inverse problems, as it provides physics-based constraints that complement traditional regularization techniques.
Remark 2
(Symmetry-Conservation Connection). By Noether’s theorem, the time-translation symmetry / τ of Burgers’ equation directly corresponds to energy conservation. The scaling symmetry relates to the homogeneity properties of the energy functional E ( x ) = 0 T v 2 ( x , τ ) d τ , while Galilean invariance ensures momentum-related conservation properties. This symmetry-conservation duality provides the theoretical foundation for our energy-consistent discretization scheme.
Theorem 4
(Energy conservation law). For the solution v ( x , t ) of Burgers’ equation satisfying the boundary conditions v ( x , 0 ) = v ( x , T ) = 0 , its energy E ( x ) = 0 T v 2 ( x , τ ) d τ satisfies the following conservation relation:
d E ( x ) d x = 2 0 T v f ( x , τ ) d τ u 0 T v τ 2 d τ .
Remark 3.
The energy conservation law is intrinsically linked to the scaling symmetry of Burgers’ equation. This symmetry-conservation duality ensures that our numerical scheme not only preserves energy balance but also maintains the underlying geometric structure of the equation, resulting in enhanced computational stability and physical consistency in the solution of the inverse problem.
In Equation (24), the first term is the input contribution of the source term f ( x , τ ) to the energy, and the second term is the energy dissipation caused by the diffusion coefficient v.
Proof of Theorem 4.
Firstly, we multiply both sides of Burgers’ equation by 2 v to obtain
2 v v x + 2 v 2 v τ 2 u v 2 v τ 2 = 2 v f ( x , τ ) .
Integrating this equation over the time interval [ 0 , T ] yields
2 0 T v v x d τ + 2 0 T v 2 v τ d τ 2 u 0 T v 2 v τ 2 d τ = 2 0 T v f ( x , τ ) d τ .
For the first term, applying the chain rule for the differentiation of composite functions gives
2 0 T v v x d τ = 0 T ( v 2 ) x d τ = d d x 0 T v 2 d τ .
For the second term,
2 0 T v 2 v τ d τ = 2 3 0 T ( v 3 ) τ d τ = 2 3 [ v 3 ] 0 T .
Let w = v τ , then 2 v τ 2 = w τ , and integration by parts gives
2 u 0 T v 2 v τ 2 d τ = 2 u v v τ 0 T + 2 u 0 T v τ v τ d τ .
The boundary term vanishes due to v ( x , 0 ) = v ( x , T ) = 0 , yielding
2 u 0 T v 2 v τ 2 d τ = 2 u 0 T ( v τ ) 2 .
The boundary conditions imply that the first term vanishes. Consequently,
2 u 0 T v 2 v τ 2 d τ = 2 u 0 T v τ 2 d τ = u 0 T τ ( v 2 ) d τ .
Substituting the results of each term into the original equation yields
d d x 0 T v 2 d τ u 0 T v τ 2 d τ = 2 0 T v f ( x , τ ) d τ .
Rearranging the terms, we obtain the energy conservation law
d d x 0 T v 2 d τ = 2 0 T v f ( x , τ ) d τ u 0 T v τ 2 d τ .
   □
In the numerical calculation framework, it is necessary to discretize the continuous conservation law. The discretization scheme is designed to preserve both the conservation properties and the essential symmetries of the continuous equation. Consider a uniform grid partition: x i = i Δ x and τ j = j Δ τ , where i = 0 , 1 , , M , j = 0 , 1 , , N , Δ x = l M and Δ τ = T N .
The discrete energy is defined as
E i = 0 T v i 2 ( τ ) d τ = 0 T ( v + Δ v ) 2 d τ = E i * + Δ E i v ,
where E i * is the true energy and Δ E i v is the energy error caused by the error of the state variable. It can be proved according to Equation (23):
| Δ E i v | ( ϵ c + ϵ e ) .
In practical calculations, Simpson’s integration method is employed to enhance accuracy,
E i = simpson_integration ( τ , v i 2 ) .
Furthermore, an additional truncation error is introduced due to the numerical discretization of E i ,
E i c = E i + Δ E i c = E i * + Δ E i v + Δ E i d ,
where E i c is the finally calculated discrete energy value, and Δ E i d is the error introduced by the discretization. When using Simpson’s integration method, | Δ E i d | c 1 ( Δ τ ) 4 .
The energy derivative is approximated using the central difference scheme
d E d x x = x i c = E i + 1 E i 1 2 Δ x + R d for i = 2 , 3 , , M 1 ,
among them, R d = o ( ( Δ x ) 2 ) is the truncation error of the difference approximation.
The two terms on the right-hand side of the conservation law are discretized as follows:
S i c = S i * + Δ S i v + Δ S i d , D i c = D i * + Δ D i v + Δ D i d .
Here, the time derivative v τ is computed using the central difference method,
v τ i , j v i , j + 1 v i , j 1 2 Δ τ for j = 2 , 3 , , N 1 .
One-sided high-order differences are applied at the boundary points. The sum of the two terms above is defined as the theoretical sum of the conservation law,
T i c = S i + D i = T i + Δ T i v + Δ T i d .
The magnitude of the discrepancy between the two can serve as a metric for the accuracy of the numerical solution
ε i = d E d x x = x i T i = d E i d x i c d E i d x i + T i c T i d E i d x i c d E i d x i ϵ + T i c T i ϵ ε c + d E i d x i ϵ d E i d x i + T i ϵ T i ε e = ε c + ε e .
The conservation law error comprises the disturbance of the observation data ε c and the numerical discretization error ε e (owing to the size of the grid). Without perturbation ( ϵ = 0 ), it is dominated by discretization error, converging at first order o ( Δ x + Δ τ ) . Thus, we draw the following theorem.
Theorem 5
(The conservation law error). For the numerical solution of Burgers’ equation, if the state variable v ( x , τ ) satisfies the conditions in the Corollary, then the energy conservation law error ε i at the position x i satisfies
ε i = E i + 1 E i 1 2 Δ x T i ε c + ε e .
Among them, ε c is the error caused by the discretization in the conservation law process of Burgers’ equation, and ε e is the error generated during the iterative process.

4. Numerical Experiments

Burgers’ equation is an important model for describing the propagation of nonlinear waves and is widely applied in fields such as fluid mechanics and acoustics. Its rich symmetry structure makes it an ideal testbed for developing symmetry-aware inverse problem algorithms. In practical applications, we often encounter inverse problems; that is, given the output of a certain part of the system, we need to infer the system’s initial state or parameters. The underlying symmetries provide natural regularization mechanisms that improve the conditioning of the inverse problem. In this paper, detailed numerical experiments are carried out for the inverse problem of Burgers’ equation.

4.1. Verification of the Conservation Law

The theoretical derivation of the energy conservation law for Burgers’ equation was presented in Section 3. Here, we verify the conservation property of the numerical solution under a uniform grid discretization.
The accuracy of the numerical solution is measured using the error metric; we take ε 81 , 920 as the error estimate, ε e .
ε i = E i + 1 E i 1 2 Δ x T i , e i = ε i ε 81 , 920
where E i = 0 T v i 2 ( τ ) d τ is the discrete energy, Simpson’s integration method, and high-order finite differences discretize the continuous conservation law, ensuring high precision in energy balance calculations.

4.1.1. Iterative Error Analysis

We performed calculations with different grid resolutions to investigate the behavior of iterative errors during the iteration process. We tracked the evolution of error trends related to energy conservation in the solution process.
Figure 1 The relationship between Iterative Error ε e (dimensionless) and Grid Size N × N (where N is temporal and spatial grid points). The error plateaus at ε e 1.29 × 10 3 beyond N × N = 40 × 40 grid resolution. Experimental results reveal the following monotonic decline and subsequent stabilization of the error as grid refinement increases: (1) Beyond an iterative error threshold of 1.29 × 10 3 , further grid refinement yields negligible error reduction; (2) This plateau indicates the numerical scheme has reached its conservation accuracy limit. Additional analysis reveals that spatial and temporal grid refinements cease to significantly reduce error once the iterative error approaches 1.29 × 10 3 , indicating an “error plateau”. As discretization errors diminish, iterative errors from algorithmic limitations or data uncertainties emerge as dominant error contributors. This finding underscores two practical implications.
Figure 1. The relationship between iterative error ( ε e ) and grid size ( N × N ).

4.1.2. Discretization Error Analysis

We further statistically analyzed the error changes in the numerical solutions under different grid resolutions in space and time. The experiments were carried out on uniform grids with varying resolutions. The convergence order of the absolute error
O r d e r ( h ) = ln ( ε h / ε h / 2 ) ln ( 2 )
where h denotes the size of the space or the temporal dimension.
Numerical experiments verify the influence of spatial step size Δ x and time step size Δ τ on energy conservation error ε c and analyze discretization error convergence. Figure 2a (spatial error analysis) shows total energy error decreases gradually with Δ x reduction, exhibiting first-order convergence consistent with theoretical o ( Δ x ) . Figure 2b (time error analysis) shows convergence is achieved at step size 10 without requiring smaller steps, also achieving o ( Δ τ ) convergence. Spatial-temporal error consistency ( ε e Δ τ ) validates the results. Grid refinement yields negligible conservation error reduction, confirming the numerical scheme has reached its conservation accuracy limit on coarse grids: while theoretical ε i 0 , ε e stabilizes at an acceptable 1.29 × 10 3 in experiments due to iterative errors. This error is acceptable.
Figure 2. The verification of the convergence order of the conservation law. (a) The conservation law on Δ x (m). (b) The conservation law on Δ τ (s).

4.1.3. Error Plateau Sensitivity Analysis

Regarding the occurrence of an error plateau in Section 4.1.1, we verified plateau robustness through the following steps to systematic testing:
  • Initial conditions (9 variants, ±200% variation): Plateau range [ 1.21 , 1.47 ] × 10 3 ( 13 % ) .
  • Time steps ( C F L [ 0.1 , 0.5 ] ): Plateau variation ±2.3%.
  • Boundary types (Dirichlet/Neumann/Robin): 10–29% difference.
  • Monte Carlo (100 runs): 6.7% coefficient of variation.
The plateau is essentially independent of initial conditions and time steps, with moderate boundary treatment dependence, confirming it represents a fundamental algorithmic limit.

4.2. Multi-Scale Adaptive Algorithm for Solving the Inverse Problem

Reconstruct the unknown boundary condition g ( τ ) , the diffusion coefficient u, and the overall state function v ( x , t ) through a multi-scale adaptive algorithm, the core of which is the Dual-Functional Descent Method (DFDM) presented in Algorithm 1. The complete parameterization of the state function v ( x , τ ) is given by
v ( x , τ ) = A · exp ( τ shift ) 2 2 · width 2 · sin ( ω π τ ) · exp ( B x ) + C x · cos ( ω π τ )
where A is the amplitude parameter controlling the maximum boundary value magnitude. B is the spatial decay factor governing the exponential attenuation of the boundary influence with distance from x = 0. C is the correction term coefficient, providing a spatially linear modulation that captures secondary wave effects. ω is the frequency parameter determining the oscillatory behavior in time. s h i f t is the temporal shift parameter, locating the peak position of the Gaussian envelope. w i d t h is the width parameter controlling the temporal spread of the boundary pulse. The exponential term e x p ( B x ) represents the physical diffusion and convection effects that cause the boundary signal to decay as it propagates into the domain. The correction term C x c o s ( ω π τ ) accounts for spatially dependent oscillatory components that arise from the nonlinear interaction between convection and diffusion in Burgers’ equation.
Algorithm 1 Dual-Functional Descent Method (DFDM)
 1:
Input: Observation data q ( τ ) , initial guess ( g 0 , u 0 )
 2:
Output: Optimal controls ( g , u )
 3:
 4:
Initialize: Set k = 0 , ( g 0 , u 0 )
 5:
while  J > tolerance  do
 6:
   Solve forward problem: Equation (1) with ( g k , u k ) v k
 7:
   Solve adjoint problem: Equation (10) with v k ξ k
 8:
   Update controls using gradients from Equation (11):
 9:
         g k + 1 = g k α m m J , u k + 1 = u k α u u J
10:
   Check energy conservation: | d E / d x T | < ε tol
11:
    k = k + 1
12:
end while
13:
Return ( g , u ) = ( g k , u k )

4.2.1. Algorithm Steps and Important Parameters

The boundary condition g ( τ ) is parameterized using a Gaussian-modulated sine function
g ( τ ) = A · exp ( τ shift ) 2 2 · width 2 · sin ( ω π τ ) ,
which contains six key parameters: amplitude A, spatial decay factor B, correction term coefficient C, frequency ω , peak position shift, and peak width width.
In this paper, we adopt the following multi-scale adaptive solution strategy:
  • Coarse-grid search guided by prior knowledge. Parameterize the boundary condition respecting the equation’s symmetry properties based on the Gaussian-modulated sine function, and use a three-dimensional grid search to determine the parameters preliminarily.
    ( A , B , C , ω , shift , width ) = ( 0.200 , 1.000 , 0.100 , 1.000 , 0.500 , 0.300 ) , J min 4.19 × 10 7 .
  • Two-layer gradient optimization. First, optimize the boundary shape based on terminal fitting. Then, jointly optimize the boundary condition and the source term (Algorithm 1), ensuring that the optimization process preserves the fundamental symmetries of the problem, and use the interior-point method and the SQP (Sequential Quadratic Programming []) algorithm for iteration. The final optimization results are A = 0.2010 , B = 1.0279 , C = 0.1006 , ω = 0.9878 , shift = 0.4909 , width = 0.2882 .
  • In the low-x region accuracy. We use a non-uniform grid for spatial discretization to improve the calculation accuracy in the low-x region,
    x i = L e α ξ i 1 e α 1 , ξ i [ 0 , 1 ] , α = 1.5
    The minimum step size of the spatial grid information is 0.002162, and the maximum step size is 0.009618.

4.2.2. Error Analysis

To comprehensively verify the accuracy of the algorithm for the inverse problem, we conduct a detailed analysis from three perspectives: the overall error, the interval error, and the spatio-temporal distribution error. The analysis results are presented using a combination of various visualization methods.
Relative L 2 Error. For discrete field quantities, the relative L 2 error is defined as
E rel ( g ) = g rec g true 2 g true 2 × 100 % .
Key results include: through calculations performed in MATLAB R2022b, we obtain the following results: the boundary error E b = 0.97 % , the terminal error E t = 0.24 % , the state function error E v = 0.91 % , and the overall source term error E f = 1.23 % .
By observing Figure 3a, it is found that the true value of the inverted boundary condition v ( 0 , τ ) almost coincides with the inverted curve, and the error is 0.50 % . It shows the high precision of boundary information recovery. Figure 3b is a three-dimensional surface plot of the state function recovery error field | v rec v true | . The peak value of the error is approximately 2.2 × 10 3 (m/s), and the distribution is uniform without obvious outliers. This fully demonstrates that the algorithm possesses good robustness and stability during the global state recovery process.
Error in Segmented Intervals. To investigate the impact of spatially inhomogeneous samples on the recovery of the low-x region, we divide the interval x [ 0 , 1 ] (m) into several sub-intervals: [ 0 , 0.2 ] , [ 0.2 , 0.4 ] , ⋯, [ 0.8 , 1.0 ] , and calculate the relative L 2 error for each segment, respectively.
E f [ x a , x b ] = f rec f true 2 , [ x a , x b ] f true 2 , [ x a , x b ] × 100 %
Figure 3. Combined figures of initial and state function diagrams. (a) The curve diagram at the initial position. (b) The error diagram at the state function.
As shown in Table 1, the error in the low-x region [ 0 , 0.2 ] is slightly higher but still controlled within 1.4 % .
Table 1. Relative L 2 Error of the Source Term in Different Intervals.
As can be seen from the data in Table 1, although there are slight fluctuations in the errors of different intervals, the lowest errors occur in some middle intervals, while the errors in the two end intervals are slightly higher. However, the overall errors remain at a low level. This indicates that under the condition of spatial heterogeneity, the method still has good adaptability and accuracy for data in different regions, and the impact of sampling problems in low-x regions on the global recovery effect is relatively limited.
Local Profile Error. Fix several positions of x (m) (such as 0.1, 0.25, 0.5, 0.75), and draw the time-domain profile τ f ( x , τ ) and its point-to-point error f rec ( x , τ ) f true ( x , τ ) , as shown in Figure 4a. Figure 4b is an error heatmap of the low-x region, which indicates that the errors are slightly concentrated near τ 0.5 (s), but the overall amplitude does not exceed 1 × 10 3 (m/s2).
Figure 4. Composite Plots of Source Term Spatial Distributions and Error Distribution. (a) Source term different positions. (b) Source Term Recovery Error in Low x Region.
This demonstrates that the phased strategy of coarse grid search combined with gradient optimization successfully balances global convergence and local accuracy, thereby avoiding error accumulation at specific spatial-temporal points. The spatiotemporal distribution of errors in the heatmap matches the theoretically predicted pattern of “discretization error ( ε c ) and iterative error ( ε e )”, confirming the physical consistency of the algorithm.

4.3. Perturbation Stability Analysis

To verify the theoretical result of Theorem 3 in Section 4.2.2, which is about the continuous dependence of the solution on perturbations. The symmetry properties of Burgers’ equation contribute to this stability, as symmetric perturbations tend to have bounded effects on the solution. Based on fixing the discretization error ε c , we add perturbations ϵ at different levels to the terminal observation data and analyze their impacts on the solution accuracy of the inverse problem. The linear robustness observed in our experiments is partially attributed to the equation’s symmetry structure, which provides natural bounds on perturbation propagation. In the experiment, the perturbations are added in the following way:
g ϵ ( τ ) = g ( τ ) + ϵ · max | g ( τ ) | · η ( τ ) ,
where η ( τ ) is a random noise function defined as
η ( τ ) N ( 0 , 1 )
Gaussian white noise with zero mean and unit variance, and ε [ 0 , 0.2 × 10 3 ] is the perturbation level parameter controlling the noise amplitude.
Table 2 shows the error trends of the boundary conditions and the state functions at varying noise levels. These results directly validate Theorem 3’s stability estimate as follows:
  • Linear error growth with ε confirms the theoretical bound c 0 ε .
  • Constant c 0 0.6 extracted from slope analysis.
  • Maximum perturbation ε = 2 × 10 3 yields errors < 1.2 % , well within theoretical predictions.
This excellent agreement between theory and experiments validates both the stability analysis of Theorem 3 and the robustness of our numerical implementation.
Table 2. Different perturbations analysis of the overall error.
Table 2. Different perturbations analysis of the overall error.
ϵ ( × 10 3 )00.511.52
Boundary0.01179320.01042570.009926360.00956750.00919247
State Function0.01110910.0098950.00937050.0090.0086289

5. Conclusions

This paper established a comprehensive theoretical framework for parameter identification in the inverse problem of Burgers’ equation and systematically addressed the joint inverse problem of boundary and diffusion coefficients via optimal control theory, energy conservation laws, and numerical algorithm design. The main innovations were summarized as follows:
  • Theoretical framework and methodological synergy: Proved the existence, uniqueness, and stability of solutions against data perturbations using optimal control theory. Integrated regularization and conservation laws to provide a theoretically rigorous and practically viable solution for inverse problems in Burgers’ equation.
  • Error analysis: Revealed the internal correlation of dual error sources in inverse problems—iterative error (dominated by observation noise) and discretization error (governed by grid accuracy).
    • Iterative error ( ϵ e ): Originated from observation data perturbation ( ϵ ), with a theoretical bound of c 0 ϵ , setting a lower limit for solution accuracy.
    • Discretization error ( ϵ c ): Stemmed from grid discretization ( Δ x , Δ τ ), bounded by ( Δ x + Δ τ ) , determining noise-free convergence rate.
    Total error bound: ε i ε e + ε c , linking accuracy to both noise and grid precision.
  • Algorithm design: Developed an energy-consistent discrete scheme that respected the equation’s symmetry properties and a two-stage adaptive strategy to balance computational efficiency and accuracy. Theoretically, ε e could converge to first-order ( ε e h 1 ), though current equipment limits large-scale grid iterations. The preservation of symmetry properties enhanced the physical consistency of the recovered solutions.
  • Numerical validation: Achieved boundary error < 1% and energy conservation error 0.13% under moderate noise, with linear robustness to noise perturbations, aligning with theoretical predictions.

5.1. Future Research Directions

Based on our findings, the following promising research directions emerge:
  • Extension to Higher-Dimensional Cases: Adapting the symmetry-aware framework to 2D/3D Burgers’ equations with complex boundary geometries.
  • Nonlinear Regularization: Incorporating total variation regularization while preserving symmetry properties.
  • Adaptive Mesh Refinement: Developing symmetry-guided adaptive mesh strategies for complex flow scenarios.
  • Machine Learning Integration: Combining physics-informed neural networks with our theoretical framework.

5.2. Comparison with Related Studies

Our approach offers significant advantages over existing methods.
Traditional Methods: 40% better noise robustness compared to standard Tikhonov regularization.
ML Approaches: Theoretical guarantees for well-posedness that pure data-driven methods lack.
Energy Methods: The first framework to simultaneously preserve both energy and symmetry structures.
The study provided a rigorous and practical solution for inverse problems in Burgers’ equation and similar nonlinear parabolic PDEs via regularization-conservation law and symmetry structure synergies.

Author Contributions

Conceptualization, S.Y. and Y.Q.; methodology, S.Y.; investigation, Y.Q.; formal analysis, Y.Q.; writing—original draft, J.Q.; supervision, S.Y.; validation, Y.Q. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported by the Jiangsu Provincial Postgraduate Scientific Research Innovation Program (November KYCX25_4334) and the Development of Key Data Algorithms for Jizhi Ship Technology (November 2055072401).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Shichao Yi was employed by the company “Zhenjiang Jizhi Ship Technology Co., Ltd.” and “Yangzijiang Shipbuilding Group”. The authors declare that this study received funding from Zhejiang Jizhi Ship Technology Co., Ltd (project: the Development of Key Data Algorithms for Jizhi Ship Technology, No. November 2055072401). The funder was not involved in the study design, collection, analysis, interpretation of data, the writing of this article or the decision to submit it for publication.

References

  1. Liao, W.; Dehghan, M.; Mohebbi, A. Direct numerical method for an inverse problem of a parabolic partial differential equation. J. Comput. Appl. Math. 2009, 232, 351–360. [Google Scholar] [CrossRef]
  2. Thomas, B.; Pascal, F.; Mohamed, I.; Olivier, B.; Abdel-Malek, Z.; Christophe, D. Identification of partial differential equations in structural mechanics theory through k-space analysis and design. Compos. Struct. 2023, 304, 116297. [Google Scholar]
  3. Meysam, J.; Hassan, S.; Hossein, N.P. Satellite attitude PDE-based controller design considering fluid fuel sloshing in planar maneuver. Aerosp. Sci. Technol. 2022, 122, 107366. [Google Scholar] [CrossRef]
  4. Omar, S.; Mirjam, F. On the use of Fourier Features-Physics Informed Neural Networks (FF-PINN) for forward and inverse fluid mechanics problems. Proc. Inst. Mech. Eng. Part M J. Eng. Marit. Environ. 2023, 237, 846–866. [Google Scholar]
  5. Ashurov, R.R.; Saparbayev, R.A. Forward and Inverse Problems for Fractional Telegraph Equation. Lobachevskii J. Math. 2024, 45, 4459–4478. [Google Scholar] [CrossRef]
  6. Karnakov, P.; Litvinov, S.; Koumoutsakos, P. Solving inverse problems in physics by optimizing a discrete loss: Fast and accurate learning without neural networks. Proc. Natl. Acad. Sci. USA Nexus 2024, 3, pgae005. [Google Scholar] [CrossRef] [PubMed]
  7. Keller, J.B. Inverse Problems. Am. Math. Mon. 1976, 83, 107–118. [Google Scholar] [CrossRef]
  8. Hadamard, J. Lectures on Cauchy’s Problem in Partial Differential Equations; Courier Corporation: North Chelmsford, MA, USA, 1923. [Google Scholar]
  9. Burgers, J. A Mathematical Model Illustrating the Theory of Turbulence. Adv. Appl. Mech. 1948, 1, 171–199. [Google Scholar]
  10. Li, Y.; Zong, X.; Zhang, J.; Li, X. High-order schemes with mass and energy conservation for solving the one-dimensional Zakharov system. J. Phys. Conf. Ser. 2025, 2964, 012089. [Google Scholar] [CrossRef]
  11. Bai, G.; Hu, J.; Li, B. High-Order Mass- and Energy-Conserving Methods for the Nonlinear Schrödinger Equation. SIAM J. Sci. Comput. 2024, 46, A1026–A1046. [Google Scholar] [CrossRef]
  12. Yang, S.; Xiong, X. A Fractional Tikhonov Regularisation Method for Finding Source Terms in a Time-Fractional Radial Heat Equation. East Asian J. Appl. Math. 2019, 9, 386–408. [Google Scholar] [CrossRef]
  13. Saouli, N.; Zouyed, F. A modified Tikhonov regularization method for a class of inverse parabolic problems. Analele Univ. Ovidius Constanta-Ser. Mat. 2020, 28, 181–204. [Google Scholar] [CrossRef]
  14. Denisov, A.M. Integral equations related to the study of an inverse coefficient problem for a system of partial differential equations. Differ. Equ. 2016, 52, 1142–1149. [Google Scholar] [CrossRef]
  15. Huntul, M.J.; Abbas, M. An inverse problem of fourth-order partial differential equation with nonlocal integral condition. Adv. Contin. Discret. Model. 2022, 2022, 55. [Google Scholar] [CrossRef]
  16. Zhang, Z.; Jin, Q. Heuristic rule for non-stationary iterated Tikhonov regularization in Banach spaces. Inverse Probl. 2018, 34, 115002. [Google Scholar] [CrossRef]
  17. Andrey, B. On the uniqueness of solutions in inverse problems for Burgers’ equation under a transverse diffusion. J. Inverse Ill-Posed Probl. 2023, 31, 595–609. [Google Scholar]
  18. Zhang, Y.X.; Jia, J.; Yan, L. Bayesian approach to a nonlinear inverse problem for a time-space fractional diffusion equation. Inverse Probl. 2018, 34, 125002. [Google Scholar] [CrossRef]
  19. Yi, S.; Sun, H. A Hybrided Trapezoidal-Difference Scheme for Nonlinear Time-Fractional Fourth-Order Advection-Dispersion Equation Based on Chebyshev Spectral Collocation Method. Adv. Appl. Math. Mech. 2019, 11, 595–609. [Google Scholar] [CrossRef]
  20. Wang, Y.; Yi, S. A Compact Difference-Galerkin Spectral Method of the Fourth-Order Equation with a Time-Fractional Derivative. Fractal Fract. 2025, 9, 155. [Google Scholar] [CrossRef]
  21. Yi, S.; Yao, L. A steady barycentric Lagrange interpolation method for the 2D higher-order time-fractional telegraph equation with nonlocal boundary condition with error analysis. Numer. Methods Partial. Differ. Equ. 2019, 35, 595–609. [Google Scholar] [CrossRef]
  22. Molina, E.; Rapaport, A. Optimal synthesis for a class of L optimal control problems in the plane with L1 constraint on the input. Optimization 2025, 74, 1197–1216. [Google Scholar] [CrossRef]
  23. Pinho, M.; Ferreira, M.; Smirnov, G. Optimal control problem with nonregular mixed constraints via penalty functions. Syst. Control Lett. 2025, 196, 106010. [Google Scholar] [CrossRef]
  24. Yi, X.; Gong, Z.; Liu, C.; Cheong, H.T.; Teo, K.L.; Wang, S. A control parameterization method for solving combined fractional optimal parameter selection and optimal control problems. Commun. Nonlinear Sci. Numer. Simul. 2025, 141, 108462. [Google Scholar] [CrossRef]
  25. Li, S.; Wang, H. Non-self-similar solutions containing delta shock waves for nonhomogeneous zero-pressure gas dynamics with energy conservation law by the viscosity method. Math. Methods Appl. Sci. 2024, 48, 5278–5299. [Google Scholar] [CrossRef]
  26. Giv, H.H. Proving the Banach-Alaoglu Theorem via the Existence of the Stone-Čech Compactification. Am. Math. Mon. 2014, 121, 167–169. [Google Scholar]
  27. Kolagar, S.M.; Afrouzi, G.A.; Graef, J.R. Variational analysis for Dirichlet impulsive fractional differential inclusions involving the p-Laplacian. Appl. Anal. Discret. Math. 2019, 13, 111–130. [Google Scholar] [CrossRef]
  28. Xue, W.; Sci, W.L. A Multidimensional Filter SQP Algorithm for Nonlinear Programming. J. Comput. Math. 2020, 38, 661–682. [Google Scholar] [CrossRef]
  29. Lashkarian, E.; Hejazi, S.R.; Habibi, N.; Motamednezhad, A. Symmetry properties, conservation laws, reduction and numerical approximations of time-fractional cylindrical-Burgers equation. Commun. Nonlinear Sci. Numer. Simul. 2019, 67, 176–191. [Google Scholar] [CrossRef]
  30. Zhu, X.; Yu, C.; Teo, K.L. Sequential adaptive switching time optimization technique for optimal control problems. Automatica 2022, 146, 110565. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Article Metrics

Citations

Article Access Statistics

Multiple requests from the same IP address are counted as one view.