This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: Corresponding author: Huan Yu.

, , ,

Neural Operators for Adaptive Control of Freeway Traffic

Kaijing Lv [email protected]    Junmin Wang [email protected]    Yihuai Zhang [email protected]    Huan Yu [email protected] School of Mathematics and Statistics, Beijing Institute of Technology, Beijing,China, 100081 Hong Kong University of Science and Technology (Guangzhou), Nansha, Guangzhou, Guangdong, China, 511400
Abstract

Uncertainty and delayed reactions in human driving behavior lead to stop-and-go traffic congestion on freeways. The freeway traffic dynamics are governed by the Aw-Rascle-Zhang (ARZ) traffic Partial Differential Equation (PDE) models with unknown relaxation time. Motivated by the adaptive traffic control problem, this paper presents a neural operator (NO) based adaptive boundary control design for the coupled 2×\times2 hyperbolic systems with uncertain spatially varying in-domain coefficients and boundary parameter. In traditional adaptive control for PDEs, solving backstepping kernel online is computationally intensive, as it requires significant resources at each time step to update the estimation of coefficients. To address this challenge, we use operator learning, i.e. DeepONet, to learn the mapping from system parameters to the kernels functions. DeepONet, a class of deep neural networks designed for approximating operators, has shown strong potential for approximating PDE backstepping designs in recent studies. Unlike previous works that focus on approximating single kernel equation associated with the scalar PDE system, we extend this framework to approximate PDE kernels for a class of the first-order coupled 2×\times2 hyperbolic kernel equations. Our approach demonstrates that DeepONet is nearly two orders of magnitude faster than traditional PDE solvers for generating kernel functions, while maintaining a loss on the order of 10310^{-3}. In addition, we mathematically validate the stability of the system through Lyapunov analysis when using DeepONet-approximated kernels in the adaptive controller. The proposed adaptive control is compared with reinforcement learning (RL) methods. Our approach guarantees stability and does not rely on initial values, which is essential for rapidly changing traffic scenarios. This is the first time this operator learning framework has been applied to the adaptive control of the ARZ traffic model, significantly enhancing the real-time applicability of this design framework for mitigating traffic congestion.

keywords:
Traffic Flow model; 2×\times2 Hyperbolic system; PDE backstepping; Neural operators; Adaptive control

1 Introduction

Stop-and-go traffic congestion is a very common phenomenon in major cities around the world. The traffic congestion on highways leads to many unsafe driving behaviors, as well as increased fuel emissions, environmental pollution, and increased commuting time [1][2]. The traffic congestion is characterized by the propagation of shock waves on road, caused by delayed driver response. There have been many studies on traffic stabilization using PDE models, such as the first-order hyperbolic PDE model proposed by Ligthill and Whitham and Richards (LWR) [3][4] to describe traffic density waves on highways. Then Aw and Rascle [5] and Zhang [6] proposed the second-order nonlinear hyperbolic PDE model to describe the evolution of velocity and density states in traffic flow. The ARZ model is a 2×\times2 hyperbolic PDE system and widely used for describing dynamics of the stop-and-go traffic oscillations. In this paper, we adopt the ARZ model and develop adaptive boundary control designs for traffic stabilization.

1.1 PDE backstepping for traffic control

The control strategy for freeway traffic congestion is usually based on static road infrastructure to regulate traffic flow, such as ramp metering and varying speed limits. Various traffic boundary control designs have been proposed to smooth traffic in the works of Bekiaris-Liberis and Delis [7], Zhang [8] as well as Karafyllis, Bekiaris-Liberis, and Papageorgiou [9]. While Bekiaris-Liberis and Delis utilize Adaptive Cruise Control vehicles for in-domain actuation as control inputs [7], Karafyllis et al. design a boundary feedback law to manage inlet demand [9]. The boundary control strategy using PDE backstepping is first proposed in [10] to stabilize the linearized ARZ system, including full state feedback and output feedback. Recent efforts [11, 12, 13, 14, 15, 16] have further developed backstepping controllers for various traffic scenarios including multi-lane, multi-class and mixed-autonomy traffic. This paper primarily focuses on adaptive control of traffic PDE systems with uncertain parameters.

In traffic flow modeling, relaxation time is a critical parameter representing drivers’ reaction delays to evolving traffic conditions. However, heterogeneity and unpredictability of individual driver behavior makes it impossible to obtain the relaxation time in practice. This uncertainty in relaxation time can significantly impact the stability and performance of traffic systems. Traditional control methods struggle to handle such uncertainties, making it difficult to ensure system stability and optimal performance under varying traffic conditions. To address these challenges, we adopt adaptive control strategies that allow for real-time adjustment of the controller gains to accommodate unknown or time-varying system characteristics, ensuring the desired system performance.

Adaptive control for PDEs first appears in the late 2000s, primarily focused on parabolic PDEs [17, 18]. Adaptive control methods [19, 20, 21] can be categorized into Lyapunov-based design, identifier-based design and swapping-based design. After a decade of research, advancements in adaptive control have begun to be applied to coupled hyperbolic PDEs [22].

Although adaptive control for PDE systems with unknown parameters has been extensively studied [23, 24, 25, 26, 27, 28] and was first applied for the ARZ PDE model in [12]. The practical implementation of the adaptive controller for the traffic systems still faces challenge. This is because the adaptive control process simultaneously requires the estimation of unknown system parameters and PDE states. After each time step, it is necessary to recalculate the solution to the PDE corresponding to the gain kernel function in order to update the estimated system parameter functions. This places extremely high demands on real-time computation. The computational resources required for calculation of the gain function increase significantly with spatial sampling precision when applying traditional finite difference and finite element methods. In this paper, we adopt neural operators to accelerate computation of adaptive PDE backstepping controllers.

1.2 Advances in machine learning for PDE traffic control

With rapid advances in machine learning, data-driven methods for solving, modeling and control of PDEs have received widespread attention including physics-informed learning, reinforcement learning and operator learning. Physics-Informed Neural Networks (PINNs) directly incorporates physical constraints into neural networks training by embedding the physical laws of PDEs into the loss function. This enables PINNs to solve PDEs without large amounts of training data. Mowlavi and Nabi extend PINNs method to PDE optimal control problems in [29]. Zhao proposed a novel hybrid Traffic state estimation (TSE) approach called Observer-Informed Deep Learning (OIDL), which integrates a PDE observer and deep learning paradigm to estimate spatial-temporal traffic states from boundary sensing data in [30]. However, PINNs need to be retrained for each new set of boundary and initial conditions, which poses limitations in adaptive control applications.

Reinforcement learning (RL) has also been increasingly applied for PDE control problems, particularly in boundary and feedback control. RL continuously optimizes strategies to achieve real-time control of complex PDE systems. In the field of traffic management, researchers have been applying RL to various traffic issues. Wu et al. used the city mobility traffic micro-simulator SUMO to design a deep RL framework for hybrid autonomous traffic in various experimental scenarios [31]. Under the same framework, [32] proposed a reinforcement learning-based car-following model for electric, connected, and automated vehicles to reduce traffic oscillations and improve energy efficiency. [33] presented the exploration using RL for traffic PDE boundary control. However, RL has limited generalization ability in practical applications. RL may perform well under the specific initial conditions. However, for initial conditions outside the training range, there may be performance degradation or even failure. RL may be sensitive to hyperparameters and exhibit unpredictable behavior, making it difficult to ensure consistent and stable performance in different scenarios.

Traditional neural networks typically learn mappings between finite dimensional Euclidean spaces, but with the advancement of research, this method has been extended to the field of NO [34]. NO-based learning methods focus on mapping between function spaces and are specifically designed for solving PDEs and dynamical systems. Compared with traditional machine learning methods, NO have two unique advantages. Firstly, theoretically speaking, NO can learn the mapping of the entire system parameter set, rather than being limited to a single system parameter like standard neural networks. Secondly, from an empirical perspective, research work [34][35] has shown that NO have significantly better accuracy than traditional deep learning methods when simulating complex functions. Therefore, NO not only solves individual equation instances, but can also handle the problems of the entire PDE family.

Recent research has effectively utilized DeepONet for one-dimensional transport PDEs[36] and reaction-diffusion equations and observer designs[37]. These studies establish the stability of PDEs under approximate kernels by employing the general operator approximation theorem. Subsequent developments focus on hyperbolic PDEs[38], parabolic PDEs with delays[39], and the ARZ PDE system for traffic flow control[40]. NO are applied for gain scheduling to enable real-time control of nonlinear PDEs[41]. This framework integrates PDE backstepping and DeepONet methods for offline learning, ensuring closed-loop stability and enhancing the accuracy in managing more complex PDEs. The application of NO-approximated gain kernels becomes even more valuable for adaptive control, where the kernel must be recomputed online at each time step to accommodate updated estimates of the plant parameters. This was first explored for first-order hyperbolic PDE in [42] and extended to the reaction-diffusion equation in [43]. Different from [42][43], where the kernel equation involves a single kernel, in this work, we extend the results of [42] to the ARZ traffic models which involved the coupled heterogeneous hyperbolic PDEs. The technical challenges arise from both the more complex kernel computations and the proof analysis of the higher-order PDE systems with the approximated controllers.

Contributions: The main contributions are summarized as follows:

  • \bullet

    We present an NO-based adaptive control method to stabilize the ARZ traffic PDE model with unknown relaxation time. Additionally, we extend stability schemes for more general 2×\times2 hyperbolic systems with uncertain spatially varying in-domain coefficients and boundary parameter. Compared to the relevant works [42][43], which approximate single kernel, a key technical challenge is dealing with the approximation of coupled 2×\times2 Goursat-form PDE kernels in the stabilization of coupled 2×\times2 hyperbolic PDEs.

  • \bullet

    To address the computational challenges associated with solving gain kernel equations, we integrate DeepONet into the adaptive control framework. It is shown that the NO is almost two orders of magnitude faster than the PDE solver in solving kernel functions, and the loss remains on the order of 10310^{-3}. To the best of our knowledge, this is the first study to integrate DeepONet with adaptive control in traffic flow systems, demonstrating its potential to improve the computational efficiency of control schemes in congested traffic scenarios.

  • \bullet

    Through comparative experiments with RL, it has been proven that our method does not rely on initial values compared to RL and provides a model-based solution with guaranteed stability. In addition, we theoretically prove the system’s stability through Lyapunov analysis when replacing with the DeepONet approximation kernels in the adaptive controller.

Organization of paper:The paper is organized as follows. Section 2 introduces ARZ traffic PDE model and a nominal adaptive backstepping control scheme designed for 2×\times2 hyperbolic PDEs. Section 3 gives a series of properties for the gain kernel and its time derivative and introduces the approximation of feedback kernel operators. Section 4 presents the stabilization achieved through the application of approximate controller gain functions via DeepONet. Numerical simulations are presented in Section 5. Section 6 presents the conclusion.

Notation.

exact operator 𝒦\mathcal{K}
neural operator 𝒦^\hat{\mathcal{K}}
unknown model parameters (c1,c2,c3,c4,r)({c}_{1},{c}_{2},{c}_{3},{c}_{4},{r})
estimated model parameters (c^1,c^2,c^3,c^4,r^)(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})
exact kernel (Ku,Kv)=𝒦(c1,c2,c3,c4,r)(K^{u},K^{v})=\mathcal{K}({c}_{1},{c}_{2},{c}_{3},{c}_{4},{r})
exact estimated kernel (K˘u,K˘v)=𝒦(c^1,c^2,c^3,c^4,r^)(\breve{K}^{u},\breve{K}^{v})=\mathcal{K}(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})
approximate estimated kernel (K^u,K^v)=𝒦^(c^1,c^2,c^3,c^4,r^)(\hat{K}^{u},\hat{K}^{v})=\hat{\mathcal{K}}(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})
Table 1: Nomenclature for kernel learning with exact and approximate operators

We present the nomenclature for kernel learning with exact and approximate operators in Table 1. We define the L2L^{2}-norm for χ(x)L2[0,1]\chi(x)\in L^{2}[0,1] as χL22=01|χ(x)|2dx\rVert\chi\rVert_{L^{2}}^{2}=\int_{0}^{1}|\chi(x)|^{2}dx. For the convenience, we set χ2=χL22\rVert\chi\rVert^{2}=\rVert\chi\rVert_{L^{2}}^{2}. The supremum norm is denoted \rVert\cdot\rVert_{\infty}. Proj\operatorname{Proj} denotes the projection operator

Projω¯{τ,ω^}={0ω^aandτ0,0ω^bandτ0,τotherwise.\displaystyle\operatorname{Proj}_{\bar{\omega}}\{\tau,\hat{\omega}\}=\left\{\begin{array}[]{rcl}0{}{}{}{}&&{\hat{\omega}\leq a~{}\mbox{and}~{}\tau\leq 0},\\ 0{}{}{}{}&&{\hat{\omega}\geq b~{}\mbox{and}~{}\tau\geq 0},\\ \tau{}{}{}{}&&\mbox{otherwise}.\end{array}\right. (4)

2 Nominal Adaptive Control Design

2.1 ARZ PDE Traffic Model

The ARZ PDE model is used to describe the formation and dynamics of the traffic oscillations which refer to variations of traffic density and speed around equlibrium values. It consists of a set of 2×\times2 hyperbolic PDEs for traffic density and velocity. The ARZ model of (ρ1(x,t),v1(x,t))(\rho_{1}(x,t),v_{1}(x,t))-system is given by

tρ1+x(ρ1v1)\displaystyle\partial_{t}\rho_{1}+\partial_{x}(\rho_{1}v_{1}) =0,\displaystyle=0, (5)
t(v1V(ρ1))+v1x(v1V(ρ1))\displaystyle\partial_{t}(v_{1}-V(\rho_{1}))+v_{1}\partial_{x}\left(v_{1}-V(\rho_{1})\right) =V(ρ1)v1τ,\displaystyle=\frac{V(\rho_{1})-v_{1}}{\tau}, (6)
ρ1(0,t)\displaystyle\rho_{1}(0,t) =q1v1(0,t),\displaystyle=\frac{q_{1}^{\star}}{v_{1}(0,t)}, (7)
v1(L,t)\displaystyle v_{1}(L,t) =U(t)+v1,\displaystyle=U(t)+v_{1}^{\star}, (8)

where (x,t)[0,L]×+(x,t)\in[0,L]\times\mathbb{R}_{+}, ρ1(x,t)\rho_{1}(x,t) represents the traffic density, v1(x,t)v_{1}(x,t) represents the traffic speed, and τ\tau denotes the relaxation time, which refers to the time required for driver behavior to adapt to equilibrium. This parameter is used to describe the process by which vehicle speed adjusts to match the traffic density. The variable p1(ρ1)p_{1}(\rho_{1}), defined as the traffic system pressure, is related to the density by the equation

p1(ρ1)=c0(ρ1)γ,p_{1}(\rho_{1})=c_{0}(\rho_{1})^{\gamma}, (9)

and c0,γ+c_{0},\gamma\in\mathbb{R}_{+}. The equilibrium velocity-density relationship V(ρ1)V(\rho_{1}) is given in Greenshield model:

V(ρ1)=vf(1(ρ1ρm)0γ),V(\rho_{1})=v_{f}\left(1-\left(\frac{\rho_{1}}{\rho_{m}}\right)^{\gamma}_{0}\right), (10)

where vfv_{f} is the free flow velocity, ρm\rho_{m} is the maximum density, and (ρ1,v1)(\rho_{1}^{\star},v_{1}^{\star}) are the equilibrium points of the system with v1=V(ρ1)v_{1}^{\star}=V(\rho_{1}^{\star}). We consider a constant traffic flux q1=ρ1V(ρ1)q_{1}^{\star}=\rho_{1}^{\star}V(\rho_{1}^{\star}) entering the domain from x=0x=0. We can apply the change of coordinates introduced in [14] to rewrite it in the Riemann coordinates and then map it to a decoupled first-order 2×\times2 hyperbolic system.

tu~1+v1xu~1\displaystyle\partial_{t}\tilde{u}_{1}+v_{1}^{\star}\partial_{x}\tilde{u}_{1} =0,\displaystyle=0, (11)
tv~1(γ0p1v1)xv~1\displaystyle\partial_{t}\tilde{v}_{1}-(\gamma_{0}p_{1}^{\star}-v_{1}^{\star})\partial_{x}\tilde{v}_{1} =c(x)u~1,\displaystyle={c}(x)\tilde{u}_{1}, (12)
u~1(0,t)\displaystyle\tilde{u}_{1}(0,t) =r0v~1(0,t),\displaystyle=r_{0}\tilde{v}_{1}(0,t), (13)
v~1(L,t)\displaystyle\tilde{v}_{1}(L,t) =U(t),\displaystyle=U(t), (14)

where

c(x)\displaystyle{c}(x) =1τexp(xτv1),r0=ρ1V(ρ1)+v1v1.\displaystyle=-\dfrac{1}{\tau}\exp(-\dfrac{x}{\tau v_{1}^{\star}}),r_{0}=\frac{\rho_{1}^{\star}V(\rho_{1}^{\star})+v_{1}^{\star}}{v_{1}^{\star}}. (15)

The relaxation time τ\tau describes how fast drivers adapt their speed to equilibrium speed-density relations. Its value is usually difficult to measure in practice and is easily affected by various external factors. Therefore, we propose adaptive control law. Motivated by the second-order ARZ model, we first propose NO-based adaptive design for a more general framework of 2×\times2 hyperbolic PDEs with spatially varying coefficients, as the linearized ARZ model is a special case of such systems.

2.2 Adaptive Control for Coupled 2×\times2 Hyperbolic PDEs

We consider the first-order coupled 2×\times2 hyperbolic PDE system with four spatially variable coefficients,

tu(x,t)+λxu(x,t)\displaystyle\partial_{t}u(x,t)+\lambda\partial_{x}u(x,t) =c1(x)u(x,t)+c2(x)v(x,t),\displaystyle=c_{1}(x)u(x,t)+c_{2}(x)v(x,t), (16)
tv(x,t)μxv(x,t)\displaystyle\partial_{t}v(x,t)-\mu\partial_{x}v(x,t) =c3(x)u(x,t)+c4(x)v(x,t),\displaystyle=c_{3}(x)u(x,t)+c_{4}(x)v(x,t), (17)
u(0,t)\displaystyle u(0,t) =rv(0,t),\displaystyle=rv(0,t), (18)
v(1,t)\displaystyle v(1,t) =U(t),\displaystyle=U(t), (19)

where t+t\in\mathbb{R}_{+} is the time, x[0,1]x\in[0,1] is the space, the states are given by u,vu,v and the initial conditions are u(x,0)=u0(x)u(x,0)=u_{0}(x), v(x,0)=v0(x)v(x,0)=v_{0}(x) where

u0,v0L2([0,1]).u_{0},v_{0}\in L^{2}([0,1]). (20)

The transport speeds

λ,μ,λ,μ>0,\displaystyle\lambda,\mu\in\mathbb{R},\quad\lambda,\mu>0, (21)

are known, while the spatially variable coefficients and boundary coefficient

c1(x),c2(x),c3(x),c4(x)\displaystyle c_{1}(x),c_{2}(x),c_{3}(x),c_{4}(x) C0([0,1]),\displaystyle\in C^{0}([0,1]), (22)
r\displaystyle r ,\displaystyle\in\mathbb{R}, (23)

are unknown.

Note that system (16)-(19) is 2×\times2 hyperbolic system with spatially variable coefficients in domain, which is different from the system in [25] with the constant coefficients. System (16)-(19) is a direct extension of the system in [25], where the difference lies in the designed adaptive update law.

To ensure the well-posedness of the kernel PDEs, the adaptive control estimation requires bounded assumptions. Our basic assumption is as follows.

Assumption 1.

Bounds are known on all uncertain parameters, that is, there exists some constants c¯i\bar{c}_{i}, i=14i=1\cdots 4, and r¯\bar{r} so that

cic¯i,i=14,|r|r¯.\|c_{i}\|_{\infty}\leq\bar{c}_{i},i=1\cdots 4,\quad|r|\leq\bar{r}. (24)

We first propose an adaptive control design using passive identifier design method, which includes the exact estimated backstepping kernels K˘u,K˘v\breve{K}^{u},\breve{K}^{v}.

We consider the identifier

tu^(x,t)=\displaystyle\partial_{t}\hat{u}(x,t)= λxu^(x,t)+c^1(x,t)u(x,t)\displaystyle-\lambda\partial_{x}\hat{u}(x,t)+\hat{c}_{1}(x,t)u(x,t)
+c^2(x,t)v(x,t)+ρe(x,t)ϖ(t)2,\displaystyle+\hat{c}_{2}(x,t)v(x,t)+\rho e(x,t)\|\varpi(t)\|^{2}, (25)
tv^(x,t)=\displaystyle\partial_{t}\hat{v}(x,t)= μxv^(x,t)+c^3(x,t)u(x,t)\displaystyle\mu\partial_{x}\hat{v}(x,t)+\hat{c}_{3}(x,t)u(x,t)
+c^4(x,t)v(x,t)+ρϵ(x,t)ϖ(t)2,\displaystyle+\hat{c}_{4}(x,t)v(x,t)+\rho\epsilon(x,t)\|\varpi(t)\|^{2}, (26)
u^(0,t)=\displaystyle\hat{u}(0,t)= r^u(0,t)+u(0,t)v2(0,t)1+v2(0,t),\displaystyle\frac{\hat{r}u(0,t)+u(0,t)v^{2}(0,t)}{1+v^{2}(0,t)}, (27)
v^(1,t)=\displaystyle\hat{v}(1,t)= U(t),\displaystyle U(t), (28)

where

e(x,t)=u(x,t)u^(x,t),ϵ(x,t)=v(x,t)v^(x,t),e(x,t)=u(x,t)-\hat{u}(x,t),\epsilon(x,t)=v(x,t)-\hat{v}(x,t), (29)

are errors between uu and vv and their estimates u^\hat{u} and v^\hat{v} , c^i\hat{c}_{i} and r^\hat{r} are estimates cic_{i} and rr. We define

ϖ(x,t)=[u(x,t),v(x,t)]T,\varpi(x,t)=[u(x,t),v(x,t)]^{T}, (30)

for some initial conditions

u^0,v^0L2([0,1]).\hat{u}_{0},\hat{v}_{0}\in L^{2}([0,1]).

The error signals (29) can straightforwardly be shown to have dynamics

te(x,t)=\displaystyle\partial_{t}{e}(x,t)= λxe(x,t)+c~1(x,t)u(x,t)\displaystyle-\lambda\partial_{x}{e}(x,t)+\tilde{c}_{1}(x,t)u(x,t)
+c~2(x,t)v(x,t)ρe(x,t)ϖ(t)2,\displaystyle+\tilde{c}_{2}(x,t)v(x,t)-\rho e(x,t)\|\varpi(t)\|^{2}, (31)
tϵ(x,t)=\displaystyle\partial_{t}{\epsilon}(x,t)= μxϵ(x,t)+c~3(x,t)u(x,t)\displaystyle\mu\partial_{x}{\epsilon}(x,t)+\tilde{c}_{3}(x,t)u(x,t)
+c~4(x,t)v(x,t)ρϵ(x,t)ϖ(t)2,\displaystyle+\tilde{c}_{4}(x,t)v(x,t)-\rho\epsilon(x,t)\|\varpi(t)\|^{2}, (32)
e(0,t)=\displaystyle e(0,t)= r~(t)v(0,t)1+v2(0,t),\displaystyle\frac{\tilde{r}(t)v(0,t)}{1+v^{2}(0,t)}, (33)
ϵ(1,t)=\displaystyle\epsilon(1,t)= 0.\displaystyle 0. (34)

where c~i=cic^i,i=1,,4\tilde{c}_{i}=c_{i}-\hat{c}_{i},i=1,\cdots,4, r~=rr^\tilde{r}=r-\hat{r}. We choose the following update laws

c^1t(x,t)\displaystyle\hat{c}_{1t}(x,t) =Projc¯1{γ1eγxe(x,t)u(x,t),c^1(x,t)},\displaystyle=\operatorname{Proj}_{\bar{c}_{1}}\left\{\gamma_{1}e^{-\gamma x}e(x,t)u(x,t),\hat{c}_{1}(x,t)\right\}, (35)
c^2t(x,t)\displaystyle\hat{c}_{2t}(x,t) =Projc¯2{γ2eγxe(x,t)v(x,t),c^2(x,t)},\displaystyle=\operatorname{Proj}_{\bar{c}_{2}}\left\{\gamma_{2}e^{-\gamma x}e(x,t)v(x,t),\hat{c}_{2}(x,t)\right\}, (36)
c^3t(x,t)\displaystyle\hat{c}_{3t}(x,t) =Projc¯3{γ3eγxε(x,t)u(x,t),c^3(x,t)},\displaystyle=\operatorname{Proj}_{\bar{c}_{3}}\left\{\gamma_{3}e^{\gamma x}\varepsilon(x,t)u(x,t),\hat{c}_{3}(x,t)\right\}, (37)
c^4t(x,t)\displaystyle\hat{c}_{4t}(x,t) =Projc¯4{γ4eγxε(x,t)v(x,t),c^4(x,t)},\displaystyle=\operatorname{Proj}_{\bar{c}_{4}}\left\{\gamma_{4}e^{\gamma x}\varepsilon(x,t)v(x,t),\hat{c}_{4}(x,t)\right\}, (38)
r^˙(t)\displaystyle\dot{\hat{r}}(t) =Projr¯{γ5e(0,t)v(0,t),r^(t)},\displaystyle=\operatorname{Proj}_{\bar{r}}\left\{\gamma_{5}e(0,t)v(0,t),\hat{r}(t)\right\}, (39)

where ρ,γ,γ1,γ2,γ3,γ4,γ5>0\rho,\gamma,\gamma_{1},\gamma_{2},\gamma_{3},\gamma_{4},\gamma_{5}>0 are scalar design gains. The adaptive law (35)-(39) have the following properties for all t>0t>0 [44]

ω~TProjω¯{τ,ω^}ω~Tτ,\displaystyle-\tilde{\omega}^{T}\operatorname{Proj}_{\bar{\omega}}\{\tau,\hat{\omega}\}\leq-\tilde{\omega}^{T}\tau, (40)

where

ω~=ωω^.\displaystyle\tilde{\omega}=\omega-\hat{\omega}. (41)
Lemma 1.

[Properties of passive identifier ] Consider the system (16)-(19) ,the identifier (2.2)-(28), with an arbitrary initial condition u^0=u^(,0),v^0=v^(,0)\hat{u}_{0}=\hat{u}(\cdot,0),\hat{v}_{0}=\hat{v}(\cdot,0) such that u^0<,v^0<\|\hat{u}_{0}\|<\infty,\|\hat{v}_{0}\|<\infty, along with the update law (35)-(39) with an arbitrary Lipschitz initial conditions satisfying the bounds (24), guarantees the following properties

c^i(,t)c¯i,|r^|\displaystyle\|\hat{c}_{i}(\cdot,t)\|_{\infty}\leq\bar{c}_{i},|\hat{r}| \displaystyle\leq r¯,t0,i=1,,4,\displaystyle\bar{r},\quad\forall t\leq 0,i=1,\cdots,4, (42)
e,ε\displaystyle\|e\|,\|\varepsilon\| \displaystyle\in 2,\displaystyle\mathcal{L}_{\infty}\cap\mathcal{L}_{2}\,, (43)
eϖ,εϖ\displaystyle\|e\|\|\varpi\|,\|\varepsilon\|\|\varpi\| \displaystyle\in 2,\displaystyle\mathcal{L}_{2}, (44)
|e(0,)|,|ε(0,)|,e|u(0,)|\displaystyle|e(0,\cdot)|,|\varepsilon(0,\cdot)|,\|e\||u(0,\cdot)| \displaystyle\in 2,\displaystyle\mathcal{L}_{2}\,, (45)
c^i,|r^˙|\displaystyle\|\hat{c}_{i}\|,|\dot{\hat{r}}| \displaystyle\in 2,\displaystyle\mathcal{L}_{2}, (46)
r~v(0,)1+v2(0,)\displaystyle\dfrac{\tilde{r}v(0,\cdot)}{\sqrt{1+v^{2}(0,\cdot)}} \displaystyle\in 2.\displaystyle\mathcal{L}_{2}. (47)
Proof.

The proof of this lemma follows a similar approach to the proof of Theorem 9.2 in [22]. The details can be found in the Appendix. ∎

Considering the plant (16)-(19) with unknown parameters ci,i=1,,4c_{i},i=1,\cdots,4 and rr, we will design a nominal adaptive control law to achieve global stability.

We consider the following adaptive backstepping transformation

w(x,t)=\displaystyle w(x,t)= u^(x,t),\displaystyle\hat{u}(x,t), (48)
z(x,t)=\displaystyle z(x,t)= v^(x,t)0xK˘u(x,ξ,t)u^(ξ,t)𝑑ξ\displaystyle\hat{v}(x,t)-\int_{0}^{x}\breve{K}^{u}(x,\xi,t)\hat{u}(\xi,t)d\xi
0xK˘v(x,ξ,t)v^(ξ,t)𝑑ξ=T[u^,v^](x,t),\displaystyle-\int_{0}^{x}\breve{K}^{v}(x,\xi,t)\hat{v}(\xi,t)d\xi=T[\hat{u},\hat{v}](x,t), (49)

where the kernels K˘u\breve{K}^{u} and K˘v\breve{K}^{v} satisfy the following kernel functions

μK˘xu(x,ξ,t)=\displaystyle\mu\breve{K}_{x}^{u}(x,\xi,t)= λK˘ξu(x,ξ,t)+c^3(ξ,t)K˘v(x,ξ,t)\displaystyle\lambda\breve{K}_{\xi}^{u}(x,\xi,t)+\hat{c}_{3}(\xi,t)\breve{K}^{v}(x,\xi,t)
+(c^1(ξ,t)c^4(ξ,t))K˘u(x,ξ,t),\displaystyle+(\hat{c}_{1}(\xi,t)-\hat{c}_{4}(\xi,t))\breve{K}^{u}(x,\xi,t), (50)
μK˘xv(x,ξ,t)=\displaystyle\mu\breve{K}_{x}^{v}(x,\xi,t)= μK˘ξv(x,ξ,t)+c^2(ξ,t)K˘u(x,ξ,t),\displaystyle-\mu\breve{K}_{\xi}^{v}(x,\xi,t)+\hat{c}_{2}(\xi,t)\breve{K}^{u}(x,\xi,t), (51)
K˘u(x,x,t)=\displaystyle\breve{K}^{u}(x,x,t)= c^3(x,t)λ+μ,\displaystyle-\frac{\hat{c}_{3}(x,t)}{\lambda+\mu}, (52)
K˘v(x,0,t)=\displaystyle\breve{K}^{v}(x,0,t)= λr^(t)μK˘u(x,0,t),\displaystyle\dfrac{\lambda\hat{r}(t)}{\mu}\breve{K}^{u}(x,0,t), (53)

The coupled 2×22\times 2 Goursat-form PDEs, governed by two gain kernels, are defined over the triangular domain 𝒯1\mathcal{T}_{1}, given by:

𝒯1={(x,ξ)| 0ξx1}.\displaystyle\mathcal{T}_{1}=\{(x,\xi)\;|\;0\leq\xi\leq x\leq 1\}. (54)

Using the transformation (48) and (49), we get the following target system

wt(x,t)=\displaystyle w_{t}(x,t)= λwx(x,t)+c^1w(x,t)+c^1e(x,t)+c^2z(x,t)\displaystyle-\lambda w_{x}(x,t)+\hat{c}_{1}w(x,t)+\hat{c}_{1}e(x,t)+\hat{c}_{2}z(x,t)
+0xω(x,ξ,t)w(ξ,t)𝑑ξ\displaystyle+\int_{0}^{x}\omega(x,\xi,t)w(\xi,t)d\xi
+0xκ(x,ξ,t)z(ξ,t)𝑑ξ\displaystyle+\int_{0}^{x}\kappa(x,\xi,t)z(\xi,t)d\xi
+c^2ε(x,t)+ρe(x,t)ϖ(t)2,\displaystyle+\hat{c}_{2}\varepsilon(x,t)+\rho e(x,t)\|\varpi(t)\|^{2}, (55)
zt(x,t)=\displaystyle z_{t}(x,t)= μzx(x,t)+c^4z(x,t)λK˘u(x,0,t)r(t)ϵ(0,t)\displaystyle\mu z_{x}(x,t)+\hat{c}_{4}z(x,t)-\lambda\breve{K}^{u}(x,0,t)r(t)\epsilon(0,t)
λK˘u(x,0,t)r~(t)z(0,t)+K˘u(x,0,t)e(0,t)\displaystyle-\lambda\breve{K}^{u}(x,0,t)\tilde{r}(t)z(0,t)+\breve{K}^{u}(x,0,t)e(0,t)
0xK˘tu(x,ξ,t)w(ξ,t)𝑑ξ\displaystyle-\int_{0}^{x}\breve{K}_{t}^{u}(x,\xi,t)w(\xi,t)d\xi
0xK˘tv(x,ξ,t)T1[w,z](ξ,t)𝑑ξ\displaystyle-\int_{0}^{x}\breve{K}_{t}^{v}(x,\xi,t)T^{-1}[w,z](\xi,t)d\xi
+T[c^1e+c^2ϵ,c^3e+c^4ε](x,t)\displaystyle+T\left[\hat{c}_{1}e+\hat{c}_{2}\epsilon,\hat{c}_{3}e+\hat{c}_{4}\varepsilon\right](x,t)
+ρT[e,ϵ](x,t)ϖ(t)2,\displaystyle+\rho T[e,\epsilon](x,t)\|\varpi(t)\|^{2}, (56)
w(0,t)=\displaystyle w(0,t)= r(t)z(0,t)+r(t)ϵ(0,t)e(0,t),\displaystyle r(t)z(0,t)+r(t)\epsilon(0,t)-e(0,t), (57)
z(1,t)=\displaystyle z(1,t)= 0,\displaystyle 0, (58)

where the coefficient ω\omega and κ\kappa are chosen to satisfy

ω(x,ξ,t)\displaystyle\omega(x,\xi,t) =c^1(x,t)K˘u(x,ξ,t)+ξxκ(x,s,t)K˘u(s,ξ,t)𝑑s,\displaystyle=\hat{c}_{1}(x,t)\breve{K}^{u}(x,\xi,t)+\int_{\xi}^{x}\kappa(x,s,t)\breve{K}^{u}(s,\xi,t)ds, (59)
κ(x,ξ,t)\displaystyle\kappa(x,\xi,t) =c^1(x,t)K˘v(x,ξ,t)+ξxκ(x,s,t)K˘v(s,ξ,t)𝑑s.\displaystyle=\hat{c}_{1}(x,t)\breve{K}^{v}(x,\xi,t)+\int_{\xi}^{x}\kappa(x,s,t)\breve{K}^{v}(s,\xi,t)ds. (60)

From the boundary conditions (19), (49) and (58), the nominal stabilizing controller is straightforwardly derived as follows [25]

U(t)=01K˘u(1,ξ,t)u^(ξ,t)𝑑ξ+01K˘v(1,ξ,t)v^(ξ,t)𝑑ξ.U(t)=\int_{0}^{1}\breve{K}^{u}(1,\xi,t)\hat{u}(\xi,t)d\xi+\int_{0}^{1}\breve{K}^{v}(1,\xi,t)\hat{v}(\xi,t)d\xi. (61)

Next, we present the stability of exact adaptive backstepping control, which serves as a guide for what we aim to achieve under the NO-based approximate adaptive backstepping design.

Theorem 1.

[Stability of exact adaptive backstepping control] Consider the plant (16)-(19) in feedback with the adaptive control law

U(t)=01K˘u(1,ξ,t)u^(ξ,t)𝑑ξ+01K˘v(1,ξ,t)v^(ξ,t)𝑑ξ,U(t)=\int_{0}^{1}\breve{K}^{u}(1,\xi,t)\hat{u}(\xi,t)d\xi+\int_{0}^{1}\breve{K}^{v}(1,\xi,t)\hat{v}(\xi,t)d\xi, (62)

along with the update law for c^1,c^2,c^3,c^4,r^\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r} given by (35)-(39) and the passive identifier u^,v^\hat{u},\hat{v} given by (2.2)-(28) satisfies the following properties for all solutions for all time:

u,v,u^,v^,u^,v^L2L,\displaystyle\|u\|,\|v\|,\|\hat{u}\|,\|\hat{v}\|,\|\hat{u}\|_{\infty},\|\hat{v}\|_{\infty}\in L_{2}\cap L_{\infty}, (63)
u^,v^,u^,v^0.\displaystyle\|\hat{u}\|_{\infty},\|\hat{v}\|_{\infty},\|\hat{u}\|,\|\hat{v}\|\mapsto 0. (64)
Proof.

The proof of this theorem follows a similar approach to the proof of Theorem 9.2 in [22]. ∎

In summary, the exact adaptive backstepping feedback law (61) can achieve global stability at the equilibrium point, with the system states u(x,t),v(x,t)u(x,t),v(x,t) converging pointwise to zero. However, this method is computationally intensive because it requires solving Volterra equations (2.2)-(53) at each time step t. To simplify the computation, we propose using the NO 𝒦^:(c^1,c^2,c^3,c^4,r^)(K^u,K^v)\hat{\mathcal{K}}:(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})\mapsto(\hat{K}^{u},\hat{K}^{v}) to approximate the exact adaptive backstepping gain operator 𝒦:(c^1,c^2,c^3,c^4,r^)(K˘u,K˘v)\mathcal{K}:(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})\mapsto(\breve{K}^{u},\breve{K}^{v}). This approach allows for neural network evaluation at each time step instead of solving the complex equations. In the following section, we will introduce how to approximate the operator 𝒦\mathcal{K} using DeepONet and use the resulting approximate control gain functions for boundary stabilization of plant (16)-(19). And we will use the universal approximation theorem of DeepONet [45] to derive the stability theorem from control gain kernel approximations.

3 Neural Operator for Approximating Gain Kernels

To ensure the effectiveness of the control strategy, the time and spatial derivatives of the gain kernel PDEs must be well-defined, continuous, and bounded. By proving the continuity and boundedness mentioned above, it can be further deduced that for a set of continuously differentiable coefficients within a certain supremely bounded norm, there exists a NO with arbitrary accuracy. This means that the NO can be used to accurately approximate these reaction coefficients.

3.1 Properties of the gain kernel functions

Lemma 2.

(Existence and bound for K˘u(x,y,t)\breve{K}^{u}(x,y,t), K˘v(x,y,t)\breve{K}^{v}(x,y,t), K˘tw(x,y,t)\breve{K}^{w}_{t}(x,y,t), K˘tz(x,y,t)\breve{K}^{z}_{t}(x,y,t)[25]) Let c^i(,t)c¯i,|r^|r¯\|\hat{c}_{i}(\cdot,t)\|_{\infty}\leq\bar{c}_{i},|\hat{r}|\leq\bar{r}, (x,t)[0,1]×+\forall(x,t)\in[0,1]\times\mathbb{R}_{+}. Then, for any fixed t+t\in\mathbb{R}_{+} and for any λ,μ+\lambda,\mu\in\mathbb{R}_{+}, c^iC1([0,1])\hat{c}_{i}\in C^{1}([0,1]) and r+r\in\mathbb{R}_{+}, the gain kernels Ku,KvK^{u},K^{v} satisfying the PDE systems (2.2)–(53), have unique C1(𝒯)C^{1}(\mathcal{T}) solution with the following property

|K˘u(x,ξ,t)|\displaystyle|\breve{K}^{u}(x,\xi,t)|\leq K¯,t0,\displaystyle\bar{K},\quad t\geq 0, (65)
|K˘v(x,ξ,t)|\displaystyle|\breve{K}^{v}(x,\xi,t)|\leq K¯,t0,\displaystyle\bar{K},\quad t\geq 0, (66)
K˘tu(x,ξ,t)\displaystyle\|\breve{K}^{u}_{t}(x,\xi,t)\|\leq M1c^1t+M2c^2t+M3c^3t\displaystyle M_{1}\|\hat{c}_{1t}\|+M_{2}\|\hat{c}_{2t}\|+M_{3}\|\hat{c}_{3t}\|
+M4c^4t+M5|r^˙|,t0,\displaystyle+M_{4}\|\hat{c}_{4t}\|+M_{5}|\dot{\hat{r}}|,\quad\forall t\geq 0, (67)
K˘tv(x,ξ,t)\displaystyle\|\breve{K}^{v}_{t}(x,\xi,t)\|\leq M6c^1t+M7c^2t+M8c^3t0,\displaystyle M_{6}\|\hat{c}_{1t}\|+M_{7}\|\hat{c}_{2t}\|+M_{8}\|\hat{c}_{3t}\|0,
+M9c^4t+M10|r^˙|,t0,\displaystyle+M_{9}\|\hat{c}_{4t}\|+M_{10}|\dot{\hat{r}}|,\quad\forall t\geq 0, (68)

where K¯>0,Mj>0,j=110\bar{K}>0,M_{j}>0,j=1\cdots 10 are constants depending on the parameter bounds (24).

Proof.

The proof of Lemma 2 can be found in [46]. ∎

3.2 Approximation of the NO

In the following discussion, we will first introduce the universal approximation theorem of DeepONet. This theorem demonstrates DeepONet’s ability to approximate operators, enabling us to use it to learn the backstepping gain kernel mapping of PDEs and theoretically guarantee the stability of the adaptive control system.

Theorem 2.

[Universal approximation theorem of DeepONet[45]].For sets XdxX\subset\mathbb{R}^{d_{x}} and YdyY\subset\mathbb{R}^{d_{y}}, they are compact sets of vectors xXx\in X and yYy\in Y, respectively. Define 𝒰:XUdu\mathcal{U}:X\mapsto U\subset\mathbb{R}^{d_{u}} and 𝒱:YVdv\mathcal{V}:Y\mapsto V\subset\mathbb{R}^{d_{v}} as sets of continuous functions u(x)u(x) and v(y)v(y), respectively, and assume that 𝒰\mathcal{U} is also a compact set. If the operator 𝒢:𝒰𝒱\mathcal{G}:\mathcal{U}\mapsto\mathcal{V} is continuous, then for any >0\mathcal{E}>0, there exist aa^{\star} and bb^{\star}\in\mathbb{N} such that when aa,bba\geq a^{\star},b\geq b^{\star}, there exist neural networks f𝒩(;θ(i)),g𝒩(;θ(i)),i=1,,bf^{\mathcal{N}}(\cdot;\theta^{(i)}),g^{\mathcal{N}}(\cdot;\theta^{(i)}),i=1,\dots,b with parameters θ(i),ϑ(i)\theta^{(i)},\vartheta^{(i)}, and their corresponding xjX,j=1,,ax_{j}\in X,j=1,\dots,a, such that 𝐮a=(u(x1),u(x2),,u(xa))T\mathbf{u}_{a}=(u(x_{1}),u(x_{2}),\dots,u(x_{a}))^{T} satisfies the following conditions.

|𝒢(u)(y)𝒢(𝐮a)(y)|,|\mathcal{G}(u)(y)-\mathcal{G}_{\mathbb{N}}(\mathbf{u}_{a})(y)|\leq\mathcal{E}, (69)

where

𝒢(𝐮a)(y)=i=1bg𝒩(𝐮;ϑ(i))f𝒩(y;θ(i)),\mathcal{G}_{\mathbb{N}}(\mathbf{u}_{a})(y)=\sum\limits_{i=1}^{b}g^{\mathcal{N}}(\mathbf{u};\vartheta^{(i)})f^{\mathcal{N}}(y;\theta^{(i)}), (70)

for all u𝒰u\in\mathcal{U} and yYy\in Y of 𝒢(u)𝒱\mathcal{G}(u)\in\mathcal{V}.

Figure 1 presents a schematic diagram of the control circuit, illustrating the utilization of neural operators to accelerate the generation process of gain kernel functions in PDE adaptive control.

We use the deepxde package to train DeepONet. As shown in Figure 2, we constructed a branch network consisting of two layers of convolutional neural networks and two layers of fully connected networks, and a trunk network consisting of two layers of fully connected networks. This framework contains a total of 32,094,465 parameters. nxn_{x} represents the number of discrete lattices in the xx direction. This framework consists of a branch network and a trunk network, encoding the input function space and output function domain respectively. Firstly, based on the kernel functions (2.2)-(53), multiple input-output data pairs are calculated offline using a traditional PDE numerical solver. Subsequently, the computed kernel functions are used as training data for training the neural network. Through the combination of multi-layer neural networks, DeepONet learns the hidden nonlinear relationships in the system, namely the mapping relationship between system parameters and kernel functions. During the training process, DeepONet continuously adjusts the network parameters through the backpropagation algorithm to minimize the error between the predicted output and the true kernels.

Refer to caption
Figure 1: The NO learning framework for adaptive control.

Next, we define the operator 𝒦:(C0([0,1]))4×(C1(𝒯))2\mathcal{K}:(C^{0}([0,1]))^{4}\times\mathbb{R}\mapsto(C^{1}(\mathcal{T}))^{2}

𝒦(c^1,c^2,c^3,c^4,r^)(,t):=(K˘u(x,ξ),K˘v(x,ξ)).\displaystyle\mathcal{K}(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})(\cdot,t):=(\breve{K}^{u}(x,\xi),\breve{K}^{v}(x,\xi)). (71)
Theorem 3.

[Existence of a NO to approximating the kernel] Consider the NO defined in (71), along with (2.2)-(53). Fixed t>0t>0, Let (c^1,c^2,c^3,c^4,r^)×C0([0,1])(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})\in\mathbb{R}\times C^{0}([0,1]). Then for all >0\mathcal{E}>0 there exists a NO 𝒦^:3×(C0([0,1]))4(C1(𝒯))2\mathcal{\hat{K}}:\mathbb{R}^{3}\times(C^{0}([0,1]))^{4}\mapsto(C^{1}(\mathcal{T}))^{2} such that for all (x,ξ)𝒯\forall(x,\xi)\in\mathcal{T},

|𝒦(c^1,c^2,c^3,c^4,r^)(,t)𝒦^(c^1,c^2,c^3,c^4,r^)(,t)|,\displaystyle|\mathcal{K}(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})(\cdot,t)-\hat{\mathcal{K}}(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})(\cdot,t)|\leq\mathcal{E},

namely,

|K~u(,t)|+|K~v(,t)|.|\tilde{K}^{u}(\cdot,t)|+|\tilde{K}^{v}(\cdot,t)|\leq\mathcal{E}. (72)
Proof.

The continuity of the operator 𝒦\mathcal{K} is derived from Lemma 2. And this result is based on Theorem 2.1 proposed by B. Deng et al. in their study [45]. ∎

Refer to caption

Figure 2: The DeepONet structure for operator 𝒦^\hat{\mathcal{K}}

4 Stabilization under DeepONet-Approximated Gain Feedback

We will demonstrate that although the adaptive controller uses approximate estimated kernel functions, the stability of the system is still guaranteed. Based on Theorem 1, we present the system stability proof for the adaptive backstepping controller using approximate estimation kernels K^u,K^v\hat{K}^{u},\hat{K}^{v} in the following theorem.

Theorem 4.

[Stabilization under approximate adaptive backstepping control] For all c¯1,c¯2,c¯3,c¯4,r¯,>0\bar{c}_{1},\bar{c}_{2},\bar{c}_{3},\bar{c}_{4},\bar{r},\mathcal{E}>0 there exists a 0:=d2μekΓ¯\mathcal{E}_{0}:=\frac{\sqrt{d}}{\sqrt{2\mu e^{k}}\bar{\Gamma}} such that for all NO approximations K^u,K^v\hat{K}^{u},\hat{K}^{v} of accuracy (0,0)\mathcal{E}\in(0,\mathcal{E}_{0}) provided by Theorem 3, the plant (16)-(19) in feedback with the adaptive control law

U(t)=01K^u(1,ξ,t)u^(ξ,t)𝑑ξ+01K^v(1,ξ,t)v^(ξ,t)𝑑ξ,U(t)=\int_{0}^{1}\hat{K}^{u}(1,\xi,t)\hat{u}(\xi,t)d\xi+\int_{0}^{1}\hat{K}^{v}(1,\xi,t)\hat{v}(\xi,t)d\xi, (73)

along with the update law for c^1,c^2,c^3,c^4\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4} and r^\hat{r} given by (35)-(39) with any Lipschitz initial condition c^10=c^1(,0),c^20=c^2(,0),c^30=c^3(,0),c^40=c^4(,0),r^0=r^(0)\hat{c}_{10}=\hat{c}_{1}(\cdot,0),\hat{c}_{20}=\hat{c}_{2}(\cdot,0),\hat{c}_{30}=\hat{c}_{3}(\cdot,0),\hat{c}_{40}=\hat{c}_{4}(\cdot,0),\hat{r}_{0}=\hat{r}(0) such that c^10c¯1,c^20c¯2,c^30c¯3,c^40c¯4,r^10r¯\hat{c}_{10}\leq\bar{c}_{1},\hat{c}_{20}\leq\bar{c}_{2},\hat{c}_{30}\leq\bar{c}_{3},\hat{c}_{40}\leq\bar{c}_{4},\hat{r}_{10}\leq\bar{r} and the passive identifier u^\hat{u}, v^\hat{v} given by (2.2)-(28) with any initial condition u^0=u^(,0),v^0=v^(,0)\hat{u}_{0}=\hat{u}(\cdot,0),\hat{v}_{0}=\hat{v}(\cdot,0) such that u^0<,v^0<\|\hat{u}_{0}\|<\infty,\|\hat{v}_{0}\|<\infty, the following properties hold:

u,v,u^,v^,u,v,u^,v^,L2L,\displaystyle\|u\|,\|v\|,\|\hat{u}\|,\|\hat{v}\|,\|{u}\|_{\infty},\|{v}\|_{\infty},\|\hat{u}\|_{\infty},\|\hat{v}\|_{\infty},\in L_{2}\cap L_{\infty}, (74)
u,v,u^,v^0.\displaystyle\|{u}\|_{\infty},\|{v}\|_{\infty},\|\hat{u}\|_{\infty},\|\hat{v}\|_{\infty}\mapsto 0. (75)

Moreover, for the equilibrium (u,v,u^,v^,c^1,c^2,c^3,c^4,r^)=(0,0,0,0,c1,c2,c3,c4,r)(u,v,\hat{u},\hat{v},\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})=(0,0,0,0,{c}_{1},{c}_{2},{c}_{3},{c}_{4},r) the following global stability estimate holds

S(t)2k2k1θ2S(0)eθ1k2S(0),t>0,S(t)\leq 2\dfrac{k_{2}}{k_{1}}\theta_{2}S(0)e^{\theta_{1}k_{2}S(0)},t>0, (76)

where

S(t):=\displaystyle S(t):= u2+v2+u^2+v^2+c~12+c~22\displaystyle\|u\|^{2}+\|v\|^{2}+\|\hat{u}\|^{2}+\|\hat{v}\|^{2}+\|\tilde{c}_{1}\|^{2}+\|\tilde{c}_{2}\|^{2}
+c~32+c~42+r~2,\displaystyle+\|\tilde{c}_{3}\|^{2}+\|\tilde{c}_{4}\|^{2}+\tilde{r}^{2}, (77)

and k1k_{1}, k2k_{2}, θ1\theta_{1} and θ2\theta_{2} are strictly positive constants.

Proof.

This proof mainly refers to [22, Chperter 9], and makes necessary supplements to the gain approximation error while reducing repetition.

Part A: DeepONet-perturbed target system

We consider the following adaptive backstepping transformation (48) and (49)

w(x,t)=\displaystyle w(x,t)= u^(x,t),\displaystyle\hat{u}(x,t), (78)
z(x,t)=\displaystyle z(x,t)= v^(x,t)0xK˘u(x,ξ,t)u^(ξ,t)𝑑ξ\displaystyle\hat{v}(x,t)-\int_{0}^{x}\breve{K}^{u}(x,\xi,t)\hat{u}(\xi,t)d\xi
0xK˘v(x,ξ,t)v^(ξ,t)𝑑ξ=T[u^,v^](x,t),\displaystyle-\int_{0}^{x}\breve{K}^{v}(x,\xi,t)\hat{v}(\xi,t)d\xi=T[\hat{u},\hat{v}](x,t), (79)

where K˘u\breve{K}^{u} and K˘v\breve{K}^{v} are exact solutions to kernel function (2.2)-(53). The transformation is an invertible backstepping transformation, with inverse in the same form

u^(x,t)=\displaystyle\hat{u}(x,t)= w(x,t),\displaystyle w(x,t), (80)
v^(x,t)=\displaystyle\hat{v}(x,t)= z(x,t)+0xL˘uw(ξ,t)𝑑ξ\displaystyle z(x,t)+\int_{0}^{x}\breve{L}^{u}w(\xi,t)d\xi
+0xL˘vz(ξ,t)𝑑ξ=T1[w,z](x,t)\displaystyle+\int_{0}^{x}\breve{L}^{v}z(\xi,t)d\xi=T^{-1}[w,z](x,t) (81)

where T1T^{-1} is an operator similar to TT. According to [47], from Lemma 2, Ku,KvK^{u},K^{v} are continuous, there exists a unique continuous inverse kernels Lu,LvL^{u},L^{v} defined on 𝒯1\mathcal{T}_{1} and there exits a constant L¯\bar{L} so that L˘uL¯\|\breve{L}^{u}\|_{\infty}\leq\bar{L}, L˘vL¯\|\breve{L}^{v}\|_{\infty}\leq\bar{L}. We will derive the DeepONet-perturbed target system with exact estimated kernels. Because the controller we have chosen is (73), where the kernels K^u\hat{K}^{u} and K^v\hat{K}^{v} are approximated by NO. This transformation lead to the following target system

wt(x,t)=\displaystyle w_{t}(x,t)= λwx(x,t)+c^1w(x,t)+c^1e(x,t)+c^2z(x,t)\displaystyle-\lambda w_{x}(x,t)+\hat{c}_{1}w(x,t)+\hat{c}_{1}e(x,t)+\hat{c}_{2}z(x,t)
+0xω(x,ξ,t)w(ξ,t)𝑑ξ\displaystyle+\int_{0}^{x}\omega(x,\xi,t)w(\xi,t)d\xi
+0xκ(x,ξ,t)z(ξ,t)𝑑ξ\displaystyle+\int_{0}^{x}\kappa(x,\xi,t)z(\xi,t)d\xi
+c^2ε(x,t)+ρe(x,t)ϖ(t)2,\displaystyle+\hat{c}_{2}\varepsilon(x,t)+\rho e(x,t)\|\varpi(t)\|^{2}, (82)
zt(x,t)=\displaystyle z_{t}(x,t)= μzx(x,t)+c^4z(x,t)λK˘u(x,0,t)r(t)ϵ(0,t)\displaystyle\mu z_{x}(x,t)+\hat{c}_{4}z(x,t)-\lambda\breve{K}^{u}(x,0,t)r(t)\epsilon(0,t)
λK˘u(x,0,t)r~(t)z(0,t)+K˘u(x,0,t)e(0,t)\displaystyle-\lambda\breve{K}^{u}(x,0,t)\tilde{r}(t)z(0,t)+\breve{K}^{u}(x,0,t)e(0,t)
0xK˘tu(x,ξ,t)w(ξ,t)𝑑ξ\displaystyle-\int_{0}^{x}\breve{K}_{t}^{u}(x,\xi,t)w(\xi,t)d\xi
0xK˘tv(x,ξ,t)T1[w,z](ξ,t)𝑑ξ\displaystyle-\int_{0}^{x}\breve{K}_{t}^{v}(x,\xi,t)T^{-1}[w,z](\xi,t)d\xi
+T[c^1e+c^2ϵ,c^3e+c^4ε](x,t)\displaystyle+T\left[\hat{c}_{1}e+\hat{c}_{2}\epsilon,\hat{c}_{3}e+\hat{c}_{4}\varepsilon\right](x,t)
+ρT[e,ϵ](x,t)ϖ(t)2,\displaystyle+\rho T[e,\epsilon](x,t)\|\varpi(t)\|^{2}, (83)
w(0,t)=\displaystyle w(0,t)= r(t)z(0,t)+r(t)ϵ(0,t)e(0,t),\displaystyle r(t)z(0,t)+r(t)\epsilon(0,t)-e(0,t), (84)
z(1,t)=\displaystyle z(1,t)= 01K~u(1,ξ,t)w(ξ,t)𝑑ξ\displaystyle-\int_{0}^{1}\tilde{K}^{u}(1,\xi,t)w(\xi,t)d\xi
01K~v(1,ξ,t)T1[w,z](ξ,t)𝑑ξ:=Γ(t),\displaystyle-\int_{0}^{1}\tilde{K}^{v}(1,\xi,t)T^{-1}[w,z](\xi,t)d\xi:=\Gamma(t), (85)

where

ω(x,ξ,t)\displaystyle\omega(x,\xi,t) =c^2(x,t)K˘u(x,ξ,t)+ξxκ(x,s,t)K˘u(s,ξ,t)𝑑s,\displaystyle=\hat{c}_{2}(x,t)\breve{K}^{u}(x,\xi,t)+\int_{\xi}^{x}\kappa(x,s,t)\breve{K}^{u}(s,\xi,t)ds, (86)
κ(x,ξ,t)\displaystyle\kappa(x,\xi,t) =c^2(x,t)K˘v(x,ξ,t)+ξxκ(x,s,t)K˘v(s,ξ,t)𝑑s.\displaystyle=\hat{c}_{2}(x,t)\breve{K}^{v}(x,\xi,t)+\int_{\xi}^{x}\kappa(x,s,t)\breve{K}^{v}(s,\xi,t)ds. (87)

The main difference between the current system (82)-(85) and the system described in (55)-(58) is the perturbation Γ(t)\Gamma(t) in the boundary conditions (85). This difference is due to the controller (73) using an approximated estimated kernels K^u\hat{K}^{u} and K^v\hat{K}^{v} instead of the exact estimated kernels K˘u\breve{K}^{u} and K˘v\breve{K}^{v}. The specific derivation process of (85) is as follows

z(1,t)=\displaystyle z(1,t)= v^(1,t)01K˘u(1,ξ,t)u^(ξ,t)𝑑ξ\displaystyle\hat{v}(1,t)-\int_{0}^{1}\breve{K}^{u}(1,\xi,t)\hat{u}(\xi,t)d\xi
01K˘v(1,ξ,t)v^(ξ,t)𝑑ξ\displaystyle-\int_{0}^{1}\breve{K}^{v}(1,\xi,t)\hat{v}(\xi,t)d\xi
=\displaystyle= U(t)01K˘u(1,ξ,t)u^(ξ,t)𝑑ξ\displaystyle U(t)-\int_{0}^{1}\breve{K}^{u}(1,\xi,t)\hat{u}(\xi,t)d\xi
01K˘v(1,ξ,t)v^(ξ,t)𝑑ξ\displaystyle-\int_{0}^{1}\breve{K}^{v}(1,\xi,t)\hat{v}(\xi,t)d\xi
=\displaystyle= 01K^u(1,ξ,t)u^(ξ,t)𝑑ξ+01K^v(1,ξ,t)v^(ξ,t)𝑑ξ\displaystyle\int_{0}^{1}\hat{K}^{u}(1,\xi,t)\hat{u}(\xi,t)d\xi+\int_{0}^{1}\hat{K}^{v}(1,\xi,t)\hat{v}(\xi,t)d\xi
01K˘u(1,ξ,t)u^(ξ,t)𝑑ξ01K˘v(1,ξ,t)v^(ξ,t)𝑑ξ\displaystyle-\int_{0}^{1}\breve{K}^{u}(1,\xi,t)\hat{u}(\xi,t)d\xi-\int_{0}^{1}\breve{K}^{v}(1,\xi,t)\hat{v}(\xi,t)d\xi
=\displaystyle= 01K~u(1,ξ,t)u^(ξ,t)𝑑ξ01K~v(1,ξ,t)v^(ξ,t)𝑑ξ.\displaystyle-\int_{0}^{1}\tilde{K}^{u}(1,\xi,t)\hat{u}(\xi,t)d\xi-\int_{0}^{1}\tilde{K}^{v}(1,\xi,t)\hat{v}(\xi,t)d\xi.

By using (81), we obtained (85).

Spatial 2\mathcal{L}^{2} boundedness and regulation of plant and
observer states

We use the following Lyapunov function candidate

V(t):=V4(t)+aV5(t),\displaystyle V(t):=V_{4}(t)+aV_{5}(t), (88)

where a>0a>0 and

V4(t)\displaystyle V_{4}(t) :=w(t)δ2=01eδxw2(x,t)𝑑x,\displaystyle:=\|w(t)\|_{-\delta}^{2}=\int_{0}^{1}e^{-\delta x}w^{2}(x,t)dx, (89)
V5(t)\displaystyle V_{5}(t) :=z(t)k2=01ekxz2(x,t)𝑑x.\displaystyle:=\|z(t)\|_{k}^{2}=\int_{0}^{1}e^{kx}z^{2}(x,t)dx. (90)

Before we start the formal calculations of the Lyapunov function, we will present the inequalities derived from Lemma 1.

K˘uK¯,K˘vK¯,\displaystyle\|\breve{K}^{u}\|_{\infty}\leq\bar{K},\|\breve{K}^{v}\|_{\infty}\leq\bar{K}, (91)
K~u,K~v,\displaystyle\|\tilde{K}^{u}\|_{\infty}\leq\mathcal{E},\|\tilde{K}^{v}\|_{\infty}\leq\mathcal{E}, (92)
w(t)=u^(t),\displaystyle\|w(t)\|=\|\hat{u}(t)\|, (93)
z(t)(1+K¯v)v^(t)+K¯uu^(t),\displaystyle\|z(t)\|\leq(1+\bar{K}^{v})\|\hat{v}(t)\|+\bar{K}^{u}\|\hat{u}(t)\|, (94)
v^(t)(1+L¯)z(t)+L¯w(t),\displaystyle\|\hat{v}(t)\|\leq(1+\bar{L})\|z(t)\|+\bar{L}\|w(t)\|, (95)
Γ(t)Γ¯(w+z),\displaystyle\Gamma(t)\leq\mathcal{E}\bar{\Gamma}(\|w\|+\|z\|), (96)

where

Γ¯=1+L¯.\displaystyle\bar{\Gamma}=1+\bar{L}.

The current work is based on the previous work [22, Chapter 9], with a key difference being that z2(1,t)=Γ(t)20z^{2}(1,t)=\Gamma(t)^{2}\neq 0. This leads to the terms 2μek2Γ¯2V(t)2\mu e^{k}\mathcal{E}^{2}\bar{\Gamma}^{2}V(t) in (98) as follows. Let δ1\delta\geq 1. There exist positive constants h1,h2,,h6h_{1},h_{2},\cdots,h_{6} and nonnegative, integrable function l1,l2,,l5l_{1},l_{2},\cdots,l_{5} such that

V˙4(t)\displaystyle\dot{V}_{4}(t)\leq h1w2(0,t)[λδh2]V4(t)+h3V5(t)\displaystyle h_{1}w^{2}(0,t)-\left[\lambda\delta-h_{2}\right]V_{4}(t)+h_{3}V_{5}(t)
+l1(t)V4(t)+l2(t),\displaystyle+l_{1}(t)V_{4}(t)+l_{2}(t), (97)
V˙5(t)\displaystyle\dot{V}_{5}(t)\leq [μekh4q~2(t)]w2(0,t)+h5V4(t)\displaystyle-\left[\mu-e^{k}h_{4}\tilde{q}^{2}(t)\right]w^{2}(0,t)+h_{5}V_{4}(t)
[kμh6]V5(t)+l3(t)V4(t)+l4(t)V5(t)\displaystyle-\left[k\mu-h_{6}\right]V_{5}(t)+l_{3}(t)V_{4}(t)+l_{4}(t)V_{5}(t)
+l5(t)+2μek2Γ¯2V(t).\displaystyle+l_{5}(t)+2\mu e^{k}\mathcal{E}^{2}\bar{\Gamma}^{2}V(t). (98)

Thus, we obtain the following upper bound calculation

V˙(t)[d2μek2Γ¯2]V(t)+l6(t)V(t)+l10(t),\dot{V}(t)\leq-\left[d-2\mu e^{k}\mathcal{E}^{2}\bar{\Gamma}^{2}\right]V(t)+l_{6}(t)V(t)+l_{10}(t), (99)

for positive constant dd and the nonnegative, integrable functions l10(t)l_{10}(t) and l6(t)l_{6}(t)

l10(t)=\displaystyle l_{10}(t)= 2r~2z(0,t)21+v(0,t)2+8br~2ϵ(0,t)2+l2(t)+al5(t)\displaystyle\dfrac{2\tilde{r}^{2}z(0,t)^{2}}{1+v(0,t)^{2}}+8b\tilde{r}^{2}\epsilon(0,t)^{2}+l_{2}(t)+al_{5}(t)
+br~2z(0,t)21+v(0,t)2z(0,t)2,\displaystyle+b\dfrac{\tilde{r}^{2}z(0,t)^{2}}{1+v(0,t)^{2}}z(0,t)^{2}, (100)
l2(t)=\displaystyle l_{2}(t)= (c¯12+4eδ)e(t)2+(c¯22+4eδ)ϵ(t)2\displaystyle\left(\bar{c}_{1}^{2}+4e^{-\delta}\right)\|e(t)\|^{2}+\left(\bar{c}_{2}^{2}+4e^{-\delta}\right)\|\epsilon(t)\|^{2}
+3λe2(0,t)+3λr¯2ϵ2(0,t),\displaystyle+3\lambda e^{2}(0,t)+3\lambda\bar{r}^{2}\epsilon^{2}(0,t), (101)
l5(t)=\displaystyle l_{5}(t)= λ2K¯2r¯2ekϵ2(0,t)+λ2K¯2eke2(0,t)\displaystyle\lambda^{2}\bar{K}^{2}\bar{r}^{2}e^{k}\epsilon^{2}(0,t)+\lambda^{2}\bar{K}^{2}e^{k}e^{2}(0,t)
+2ekA12c¯12e(t)2+2ekA12c¯22ϵ(t)2\displaystyle+2e^{k}A_{1}^{2}\bar{c}_{1}^{2}\|e(t)\|^{2}+2e^{k}A_{1}^{2}\bar{c}_{2}^{2}\|\epsilon(t)\|^{2}
+2ekA22c¯32e(t)2+2ekA22c¯42ϵ(t)2\displaystyle+2e^{k}A_{2}^{2}\bar{c}_{3}^{2}\|e(t)\|^{2}+2e^{k}A_{2}^{2}\bar{c}_{4}^{2}\|\epsilon(t)\|^{2}
+4eδe(t)2+4eδϵ(t)2,\displaystyle+4e^{-\delta}\|e(t)\|^{2}+4e^{-\delta}\|\epsilon(t)\|^{2}, (102)
l6(t)=\displaystyle l_{6}(t)= max{l1(t)+al3(t),l4(t)},\displaystyle\mathop{\max}\left\{l_{1}(t)+al_{3}(t),l_{4}(t)\right\}, (103)
l1(t)=\displaystyle l_{1}(t)= e2δρ2e(t)2ϖ(t)2,\displaystyle e^{2\delta}\rho^{2}\|e(t)\|^{2}\|\varpi(t)\|^{2}, (104)
l3(t)=\displaystyle l_{3}(t)= eδ+kK^tu(t)2+2eδ+kK^tv(t)2A32,\displaystyle e^{\delta+k}\left\|\hat{K}_{t}^{u}(t)\right\|^{2}+2e^{\delta+k}\left\|\hat{K}_{t}^{v}(t)\right\|^{2}A_{3}^{2}, (105)
l4(t)=\displaystyle l_{4}(t)= 2ρ2eδ+2k(A12e(t)2+A22ϵ(t)2)ϖ(t)2\displaystyle 2\rho^{2}e^{\delta+2k}\left(A_{1}^{2}\|e(t)\|^{2}+A_{2}^{2}\|\epsilon(t)\|^{2}\right)\|\varpi(t)\|^{2}
+2ekK^tv(t)2A42,\displaystyle+2e^{k}\left\|\hat{K}_{t}^{v}(t)\right\|^{2}A_{4}^{2}, (106)

where a,b,k,Aia,b,k,A_{i}, i=14i=1\cdots 4 are positive constants. We introduce

0:=d2μekΓ¯.\mathcal{E}_{0}:=\dfrac{\sqrt{d}}{\sqrt{2\mu e^{k}}\bar{\Gamma}}. (107)

Thus, if we choose (0,0)\mathcal{E}\in(0,\mathcal{E}_{0}) we have d2μek2Γ¯2>12>0d-2\mu e^{k}\mathcal{E}^{2}\bar{\Gamma}^{2}>\frac{1}{2}>0. It then follows from Lemma B.6 in [48] that

V1,V\in\mathcal{L}_{1}\cap\mathcal{L}_{\infty},

and hence

u,v2.\|u\|,\|v\|\in\mathcal{L}_{2}\cap\mathcal{L}_{\infty}.

Due to the invertibility of the backstepping transformation

u^,v^2.\|\hat{u}\|,\|\hat{v}\|\in\mathcal{L}_{2}\cap\mathcal{L}_{\infty}.

From Lemma 1 it follows that

u,v2.\|u\|,\|v\|\in\mathcal{L}_{2}\cap\mathcal{L}_{\infty}.

Part B: Pointwise-in-space boundedness and regulation

The paper [49] proved that the system (16)-(19) is equivalent to the following system through an invertible backstepping transformation.

αt(x,t)\displaystyle\alpha_{t}(x,t) =λαx(x,t)+g(x)β(0,t),\displaystyle=-\lambda\alpha_{x}(x,t)+g(x)\beta(0,t), (108)
βt(x,t)\displaystyle\beta_{t}(x,t) =μβx(x,t),\displaystyle=\mu\beta_{x}(x,t), (109)
α(0,t)\displaystyle\alpha(0,t) =rβ(0,t),\displaystyle=r\beta(0,t), (110)
β(1,t)\displaystyle\beta(1,t) =U(t)01G1(ξ)u(ξ)𝑑ξ01G2(ξ)v(ξ)𝑑ξ,\displaystyle=U(t)-\int_{0}^{1}G_{1}(\xi)u(\xi)d\xi-\int_{0}^{1}G_{2}(\xi)v(\xi)d\xi, (111)

for some bounded functions g,G1,G2𝒞g,G_{1},G_{2}\in\mathcal{C} fo the unknown parameters. Equation (108)-(111) can be explicitly be solved for t>λ1+μ1t>\lambda^{-1}+\mu^{-1} to yield

α(x,t)=\displaystyle\alpha(x,t)= rβ(1,tμ1λ1x)\displaystyle r\beta\left(1,t-\mu^{-1}-\lambda^{-1}x\right)
+λ10xh(τ)β(1,tμ1λ1(xτ))𝑑τ,\displaystyle+\lambda^{-1}\int_{0}^{x}h(\tau)\beta\left(1,t-\mu^{-1}-\lambda^{-1}(x-\tau)\right)d\tau, (112)
β(x,t)=\displaystyle\beta(x,t)= β(1,tμ1(1x)).\displaystyle\beta\left(1,t-\mu^{-1}(1-x)\right). (113)

From (111), the control law U(t)U(t) and u,v,u^,v^\|u\|,\|v\|,\|\hat{u}\|,\|\hat{v}\|\in 2\mathcal{L}_{2}\cap\mathcal{L}_{\infty}, it follows that β(1,)\beta(1,\cdot)\in 2\mathcal{L}_{2}\cap\mathcal{L}_{\infty}. Since β\beta and α\alpha are simple, cascaded transport equations, this implies

α,β2,α,β0.\displaystyle\|\alpha\|_{\infty},\|\beta\|_{\infty}\in\mathcal{L}_{2}\cap\mathcal{L}_{\infty},\quad\|\alpha\|_{\infty},\|\beta\|_{\infty}\rightarrow 0. (114)

With the invertibility of the transformation, then yields

u,v2,u,v0.\|u\|_{\infty},\|v\|_{\infty}\in\mathcal{L}_{2}\cap\mathcal{L}_{\infty},\quad\|u\|_{\infty},\|v\|_{\infty}\rightarrow 0. (115)

From the structure of the identifier (16), we will also have u^(x,),v^(x,)2\hat{u}(x,\cdot),\hat{v}(x,\cdot)\in\mathcal{L}_{\infty}\cap\mathcal{L}_{2} , and hence

u^,v^2,u^,v^0.\displaystyle\|\hat{u}\|_{\infty},\|\hat{v}\|_{\infty}\in\mathcal{L}_{\infty}\cap\mathcal{L}_{2},\quad\|\hat{u}\|_{\infty},\|\hat{v}\|_{\infty}\mapsto 0. (116)

Part C: Global stability

Here, we will prove the global stability of the system, specifically by proving (76), and thus we introduce the following function

S(t):=\displaystyle S(t):= u2+v2+u^2+v^2+c~12+c~22\displaystyle\|u\|^{2}+\|v\|^{2}+\|\hat{u}\|^{2}+\|\hat{v}\|^{2}+\|\tilde{c}_{1}\|^{2}+\|\tilde{c}_{2}\|^{2} (117)
+c~32+c~42+r~2.\displaystyle+\|\tilde{c}_{3}\|^{2}+\|\tilde{c}_{4}\|^{2}+\tilde{r}^{2}. (118)

The goal of the proof is to demonstrate the existence of a function θ𝒦\theta\in\mathcal{K}_{\infty} such that the following inequality holds.

S(t)θ(S(0)),t0.S(t)\leq\theta(S(0)),t\geq 0. (119)

We will reuse the Lyapunov function from Appendix to show that the system’s state remains stable over time.

V1(t)=\displaystyle V_{1}(t)= V2(t)+γ11c~12+γ21c~22+γ31c~32\displaystyle V_{2}(t)+\gamma_{1}^{-1}\|\tilde{c}_{1}\|^{2}+\gamma_{2}^{-1}\|\tilde{c}_{2}\|^{2}+\gamma_{3}^{-1}\|\tilde{c}_{3}\|^{2}
+γ41c~42+12γ5r~2(t),\displaystyle+\gamma_{4}^{-1}\|\tilde{c}_{4}\|^{2}+\dfrac{1}{2\gamma_{5}}\tilde{r}^{2}(t), (120)

where

V2(t)=01eγxe2(x,t)𝑑x+01eγxε2(x,t)𝑑x,V_{2}(t)=\int_{0}^{1}e^{-\gamma x}e^{2}(x,t)dx+\int_{0}^{1}e^{\gamma x}\varepsilon^{2}(x,t)dx, (121)

leads to the following upper bound:

V˙1(t)\displaystyle\dot{V}_{1}(t)\leq λeγe2(1,t)λe2(0,t)v2(0,t)λγeγe(t)2\displaystyle-\lambda e^{-\gamma}e^{2}(1,t)-\lambda e^{2}(0,t)v^{2}(0,t)-\lambda\gamma e^{-\gamma}\|e(t)\|^{2}
2ρeγe(t)2ϖ(t)2μϵ2(0,t)\displaystyle-2\rho e^{-\gamma}\|e(t)\|^{2}\|\varpi(t)\|^{2}-\mu\epsilon^{2}(0,t)
μγϵ(t)22ρeγϵ(t)2ϖ(t)2,\displaystyle-\mu\gamma\|\epsilon(t)\|^{2}-2\rho e^{\gamma}\|\epsilon(t)\|^{2}\|\varpi(t)\|^{2}, (122)

which shows that V1(t){V}_{1}(t) is non-increasing and hence bounded. Thus implies that the V1(t)<V1(0){V}_{1}(t)<V_{1}(0) and limit limtV1(t)=V1,\lim_{t\to\infty}V_{1}(t)=V_{1,\infty} exists. By integrating (122) from zero to infinity, we obtain the following upper bound:

λeγ0e2(1,τ)𝑑τ+λ0e2(0,τ)v2(0,τ)𝑑τ\displaystyle\lambda e^{-\gamma}\int_{0}^{\infty}e^{2}(1,\tau)d\tau+\lambda\int_{0}^{\infty}e^{2}(0,\tau)v^{2}(0,\tau)d\tau
+λγeγ0e(τ)2𝑑τ\displaystyle+\lambda\gamma e^{-\gamma}\int_{0}^{\infty}\|e(\tau)\|^{2}d\tau
+2ρeγ0e(τ)2ϖ(τ)2𝑑τ+μ0ϵ2(0,τ)𝑑τ\displaystyle+2\rho e^{-\gamma}\int_{0}^{\infty}\|e(\tau)\|^{2}\|\varpi(\tau)\|^{2}d\tau+\mu\int_{0}^{\infty}\epsilon^{2}(0,\tau)d\tau
+μγ0ϵ(τ)2𝑑τ+2ρeγ0ϵ(τ)2ϖ(τ)2𝑑τ\displaystyle+\mu\gamma\int_{0}^{\infty}\|\epsilon(\tau)\|^{2}d\tau+2\rho e^{\gamma}\int_{0}^{\infty}\|\epsilon(\tau)\|^{2}\|\varpi(\tau)\|^{2}d\tau
V1(0).\displaystyle\leq V_{1}(0). (123)

From (100) and (103), it can be concluded that there are constants θ1>0\theta_{1}>0 and θ2>1\theta_{2}>1 such that

l61\displaystyle\|l_{6}\|_{1}\leq θ1V1(0),\displaystyle\theta_{1}V_{1}(0), (124)
l101\displaystyle\|l_{10}\|_{1}\leq θ2V1(0).\displaystyle\theta_{2}V_{1}(0). (125)

Recalling (99), we have that

V˙(t)12V(t)+l6(t)V(t)+l10(t).\displaystyle\dot{V}(t)\leq-\frac{1}{2}V(t)+l_{6}(t)V(t)+l_{10}(t). (126)

We also have from Lemma B.6 in [48] that

V(t)(e12tV(0)+l101)el61.V(t)\leq(e^{-\frac{1}{2}t}V(0)+\|l_{10}\|_{1})e^{\|l_{6}\|_{1}}. (127)

We then introduce the function

V3(t):=V1(t)+V(t).V_{3}(t):=V_{1}(t)+V(t). (128)

Noticing that

V1(t)V1(0)θ2V1(0)eθ1V1(0),V_{1}(t)\leq V_{1}(0)\leq\theta_{2}V_{1}(0)e^{\theta_{1}V_{1}(0)}, (129)

we achieve from (127), (129), (124) and (125) the following

V3(t)=\displaystyle V_{3}(t)= V(t)+V1(t)\displaystyle V(t)+V_{1}(t)
\displaystyle\leq (e12tV(0)+l101)el61+θ2V1(0)eθ1V1(0)\displaystyle(e^{-\frac{1}{2}t}V(0)+\|l_{10}\|_{1})e^{\|l_{6}\|_{1}}+\theta_{2}V_{1}(0)e^{\theta_{1}V_{1}(0)}
\displaystyle\leq (θ2V(0)+θ2V1(0))el61+θ2V1(0)eθ1V1(0)\displaystyle(\theta_{2}V(0)+\theta_{2}V_{1}(0))e^{\|l_{6}\|_{1}}+\theta_{2}V_{1}(0)e^{\theta_{1}V_{1}(0)}
\displaystyle\leq 2θ2V3(0)eθ1V3(0).\displaystyle 2\theta_{2}V_{3}(0)e^{\theta_{1}V_{3}(0)}. (130)

This Lyapunov functional can be represented by an equivalent norm, and the bounds of this equivalent norm are determined by two positive constants k1>0k_{1}>0 and k2>0k_{2}>0.

k1S(t)V3(t)k2S(t).k_{1}S(t)\leq V_{3}(t)\leq k_{2}S(t). (131)

So we have

S(t)2k2k1θ2S(0)eθ1k2S(0).S(t)\leq 2\dfrac{k_{2}}{k_{1}}\theta_{2}S(0)e^{\theta_{1}k_{2}S(0)}. (132)

5 Simulations

This section will present and analyze the performance of the proposed NO-based adaptive controllers for two PDE models: (i) a general 2×\times2 hyperbolic system (16)-(19) (ii) the ARZ PDE system. Through these examples, we will demonstrate the effectiveness of the NO-based adaptive control design.

5.1 Simulation of the Coupled 2×\times2 Hyperbolic System

A. Simulation configuration

The coefficients of system (16)-(19) are defined as c1(x)=cos(σ1cos1(x))c_{1}(x)=\cos(\sigma_{1}\cos^{-1}(x)), c2(x)=cos(σ2cos1(x))c_{2}(x)=\cos(\sigma_{2}\cos^{-1}(x)), c3(x)=sin(1σ3x)+1c_{3}(x)=\sin(1-\sigma_{3}x)+1 and c4(x)=cos(σ4x)c_{4}(x)=\cos(\sigma_{4}x) with the shape parameters σ1\sigma_{1}, σ2\sigma_{2}, σ3\sigma_{3} and σ4\sigma_{4}. We choose the spatially varying coefficient from three classes of functions—namely the Chebyshev polynomial forms, sine function and cosine function. Because these coefficients are both bounded and equicontinuous, satisfying the compactness requirement of Theorem 3. And We can generate a rich and diverse set of kernel functions for training data by varying a single parameter σ1\sigma_{1}, σ2\sigma_{2}, σ3\sigma_{3} and σ4\sigma_{4}. Although this paper uses specific Chebyshev polynomial forms, sine functions, and cosine functions, our framework is applicable to any compact set of continuous functions. We use the finite difference scheme to solve PDEs, where the time step dt=0.005sdt=0.005s, the spatial step dx=0.05mdx=0.05m, the total time T=10sT=10s, and the length L=1mL=1m. The initial conditions are u0=sin(2πx)u_{0}=\sin(2\pi x), v0=xv_{0}=x. We demonstrate in the Figure 3 that the system is open-loop unstable.

B. Dataset generation and NO training

In order to establish a mapping (c^1,c^2,c^3,c^4,r^)(K˘u,K˘v)(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r})\mapsto(\breve{K}^{u},\breve{K}^{v}), one must build a sufficiently large dataset containing the (c^1(x),c^2(x),c^3(x),c^4(x),r^)(\hat{c}_{1}(x),\hat{c}_{2}(x),\hat{c}_{3}(x),\hat{c}_{4}(x),\hat{r}) values that may be encountered. In this work, we choose 10 sets of (c1,c2,c3,c4,r)(c_{1},c_{2},c_{3},c_{4},r) randomly sampled with σ1U(3.5,4.5)\sigma_{1}\sim U(3.5,4.5), σ2U(0.8,1)\sigma_{2}\sim U(0.8,1), σ3U(20,21)\sigma_{3}\sim U(20,21), σ4U(10,11)\sigma_{4}\sim U(10,11) and rU(2,5)r\sim U(2,5), where U(a,b)U(a,b) denotes the uniform distribution over the interval [a,b][a,b]. We simulate trajectories using adaptive control methods and calculate the corresponding kernel functions using numerical solvers. Each trajectory was sampled at 10001000 time points, resulting in a dataset of 1000010000 sets of (c^1,c^2,c^3,c^4,r^,K˘u,K˘v)(\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4},\hat{r},\breve{K}^{u},\breve{K}^{v}) for training. We trained the model on an Nvidia RTX 4060 Ti GPU. After 600 epochs of training, the L2L_{2} error of kernel K˘u\breve{K}^{u} reached 1.2×1031.2\times 10^{-3}, and the test error was 1.1×1031.1\times 10^{-3}. The L2L_{2} error of kernel K˘v\breve{K}^{v} reached 1.3×1031.3\times 10^{-3}, and the testing error was 1.2×1031.2\times 10^{-3}, as shown in Figure 4.

C. Computation time comparison

We begin to discuss the computational acceleration performance of NO. Table 2 provides a comparison of the computation time of solving kernels at each time step using the numerical solver and the trained DeepONet model. The last column of Table 2 represents the multiple of acceleration calculation. As the sampling accuracy increases (i.e.,the discrete space step size dxdx decreases), the calculation time of the finite-difference method will significantly increase. In contrast, the calculation time of NO only slightly increases with the decrease of spatial step size. We can see that as the sampling accuracy improves, the acceleration obtained by the NO becomes substantial. Especially when the spatial step size dx=0.0005, the kernel computation time is shortened from 3.191s to 4.631 ×103\times 10^{-3}s. We computed the average absolute error ξ10x(|K˘uK^u|+|K˘vK^v|)𝑑ξ𝑑x\int_{\xi}^{1}\int_{0}^{x}(|\breve{K}^{u}-\hat{K}^{u}|+|\breve{K}^{v}-\hat{K}^{v}|)d\xi dx between numerical solutions and NO solutions with different step sizes. Although the error slightly increases with the decrease of step size, they are quite small at all step sizes. Because adaptive control requires calculating control gain at every step of updating parameter estimation, quickly solving the kernel function can help improve the performance of adaptive control.

D. Simulation results

We test the performance of the closed-loop system stability with test values (σ1=4,σ2=0.9,σ3=20.1,σ4=10.1,r=4)(\sigma_{1}=4,\sigma_{2}=0.9,\sigma_{3}=20.1,\sigma_{4}=10.1,r=4) unseen during training. Figure 5 shows the kernels K˘u,K˘v\breve{K}^{u},\breve{K}^{v} calculated by the numerical solver, the kernels K^u,K^v\hat{K}^{u},\hat{K}^{v} learned by DeepONet, and the error between them. In Figure 6, we demonstrate closed-loop stability with the NO approximated kernel function for the control feedback law. The NO approximation of the kernels is illustrated in Figure 5, where most of the absolute pointwise errors for the learned kernel do not exceed 0.03. Figure 5 and Figure 6 confirm that the kernels K^u,K^v\hat{K}^{u},\hat{K}^{v} approximated by NO can effectively simulate the backstepping kernels K˘u,K˘v\breve{K}^{u},\breve{K}^{v} while maintaining the stability of the system. All estimated parameters c^i\hat{c}_{i} and r^\hat{r} are shown in Figure LABEL:coefficient_ci. We emphasize that although in adaptive control the system parameters c^1,c^2,c^3,c^4\hat{c}_{1},\hat{c}_{2},\hat{c}_{3},\hat{c}_{4} and r^\hat{r} may not precisely converge to their true values, this does not affect the control performance. This phenomenon is not a problem but rather a characteristic of adaptive control. The goal of adaptive control is not perfect system identification, but rather the estimation of parameters that ensure system stability.

E. Comparative experiment with RL

We will evaluate the performance of NO-based (DeepONet) adaptive control method and RL method for stabilization results under different initial conditions. Specifically, we assume that the initial condition of state uu is a sine function, and the specific form is

u0=sin(ω0πx),u_{0}=\sin(\omega_{0}\pi x), (133)

where ω0\omega_{0} is the frequency of a sine wave. To evaluate the performance of these two methods, we train DeepONet and RL at the same frequency ω0=2\omega_{0}=2, ensuring all other parameters remained consistent with those in Figure 5. In the testing phase, we will use sine initial conditions of different frequencies ω0=2,10\omega_{0}=2,10 to verify the model stability of NO-based adaptive control and RL. Figure LABEL:Comparison_with_RL shows the stabilization results of the RL and NO control under different initial conditions. The comparative experiments highlight a significant advantage of the NO-based adaptive control method, which consistently demonstrates robustness across different initial conditions. Specifically, the NO-based adaptive control method maintains system stability without requiring retraining even when the initial conditions are changed. This characteristic underscores its adaptability in dynamic environments. In contrast, the RL method shows a significant dependency on initial conditions. Although it performs well under specific conditions encountered during training, it is unstable when faced with unforeseen initial conditions(ω0=10\omega_{0}=10).In real-world scenarios where initial conditions are often variable and unpredictable, DeepONet ensures stability and adaptability without the need for retraining. In summary, this demonstrates DeepONet’s potential for more reliable applications in adaptive control systems, where maintaining performance across diverse conditions is crucial.

Refer to caption
Figure 3: Open-loop u(x,t)u(x,t) and v(x,t)v(x,t). We note that the plant is open-loop unstable.
Spatial Step Size
Numerical solver (s)
NO(s)
Speedup Error
dx=0.010.01 8.221×1038.221\times 10^{-3} 4.38×1034.38\times 10^{-3} 2×2\times 0.0240.024
dx=0.050.05 2.432×1022.432\times 10^{-2} 4.42×1034.42\times 10^{-3} 58×58\times 0.0310.031
dx=0.0010.001 8.701×1018.701\times 10^{-1} 4.513×1034.513\times 10^{-3} 192×192\times 0.0370.037
dx=0.00050.0005 3.1913.191 4.631×1034.631\times 10^{-3} 689×689\times 0.0450.045
Table 2: Comparison of computation time of kernels K˘u\breve{K}^{u} and K˘v\breve{K}^{v}.
Refer to caption
Figure 4: The train and test loss for K˘u\breve{K}^{u} and K˘v\breve{K}^{v}
Refer to caption
Figure 5: The first row shows the exact estimated kernels K˘u\breve{K}^{u} and K˘v\breve{K}^{v}. The second row shows the NO estimated kernels K^u\hat{K}^{u} and K^v\hat{K}^{v}. The last row shows the kernel errors K˘uK^u\breve{K}^{u}-\hat{K}^{u} and K˘vK^v\breve{K}^{v}-\hat{K}^{v}. All kernels are plotted at the final time TT.
Refer to caption
Figure 6: Simulation of the close-loop system with the update law (35)-(39), feedback controller (62) and (73). The left columns of the first and second rows show close-loop system states u(x,t)u(x,t), v(x,t)v(x,t) with the analyzed kernels K˘u\breve{K}^{u} and K˘v\breve{K}^{v}. The middle columns of the first and second rows show close-loop system states uNO(x,t)u_{NO}(x,t), vNO(x,t)v_{NO}(x,t) with the approximated kernels K^u\hat{K}^{u} and K^v\hat{K}^{v}. The right columns of the first and second rows show the errors between u(x,t)u(x,t) and uNO(x,t)u_{NO}(x,t), and between v(x,t)v(x,t) and vNO(x,t)v_{NO}(x,t), respectively.

5.2 Application Simulation of the ARZ Traffic System

A. NO-based adaptive controller

Following the steps in the section 2, we can obtain the adaptive controller for ARZ traffic system (5)-(8) as follows

U(t)=01K˘u1(1,ξ,t)u~1(ξ,t)𝑑ξ+01K˘v1(1,ξ,t)v~1(ξ,t)𝑑ξ,U(t)=\int_{0}^{1}\breve{K}^{u_{1}}(1,\xi,t)\tilde{u}_{1}(\xi,t)d\xi+\int_{0}^{1}\breve{K}^{v_{1}}(1,\xi,t)\tilde{v}_{1}(\xi,t)d\xi, (134)

with the parameter update law

c^t(x,t)=Projc¯{γ3eγxε1(x,t)u~1(x,t),c^(x,t)},\hat{c}_{t}(x,t)=\operatorname{Proj}_{\bar{c}}\left\{\gamma_{3}e^{\gamma x}\varepsilon_{1}(x,t){\tilde{u}_{1}}(x,t),\hat{c}(x,t)\right\}, (135)

where

ϵ1(x,t)=v1(x,t)v^1(x,t),\epsilon_{1}(x,t)=v_{1}(x,t)-\hat{v}_{1}(x,t), (136)

and the kernels satisfy the following kernel functions

(γpv1)K˘xu1(x,ξ,t)=\displaystyle(\gamma p^{\star}-{v_{1}}^{\star})\breve{K}_{x}^{u_{1}}(x,\xi,t)= v1K˘ξu1(x,ξ,t)\displaystyle{v_{1}}^{\star}\breve{K}_{\xi}^{u_{1}}(x,\xi,t)
+c^(x,t)K˘v1(x,ξ,t),\displaystyle+\hat{c}(x,t)\breve{K}^{v_{1}}(x,\xi,t), (137)
(γpv1)K˘xv1(x,ξ,t)=\displaystyle(\gamma p^{\star}-{v_{1}}^{\star})\breve{K}_{x}^{v_{1}}(x,\xi,t)= (γpv1)K˘ξv1(x,ξ,t),\displaystyle-(\gamma p^{\star}-{v_{1}}^{\star})\breve{K}_{\xi}^{v_{1}}(x,\xi,t), (138)
K˘u1(x,x,t)=\displaystyle\breve{K}^{u_{1}}(x,x,t)= c^(x,t)γp,\displaystyle-\frac{\hat{c}(x,t)}{\gamma p^{\star}}, (139)
K˘v1(x,0,t)=\displaystyle\breve{K}^{v_{1}}(x,0,t)= v1γpv1rK˘u1(x,0,t).\displaystyle\dfrac{{v_{1}}^{\star}}{\gamma p^{\star}-{v_{1}}^{\star}}r\breve{K}^{u_{1}}(x,0,t). (140)

According to the approximation of NO in Theorem 3, we get the NO-based adaptive controller

U(t)=01K^u1(1,ξ,t)u~1(ξ,t)𝑑ξ+01K^v1(1,ξ,t)v~1(ξ,t)𝑑ξ.U(t)=\int_{0}^{1}\hat{K}^{u_{1}}(1,\xi,t)\tilde{u}_{1}(\xi,t)d\xi+\int_{0}^{1}\hat{K}^{v_{1}}(1,\xi,t)\tilde{v}_{1}(\xi,t)d\xi. (141)
Method
Average
Computation Time (s)
Error
Nominal Adaptive Controller 1.921.92 0
NO-based adaptive Controller 0.0360.036 0.0320.032
Table 3: The computation time and error of nominal adaptive controller and NO-based adaptive controller.

B. Simulation results

Then, we analyze the performance of the proposed NO-based adaptive control law for the ARZ traffic PDE system through simulations on a L=600m road over T=300 seconds. The parameters are set as follows: free-flow velocity vm=40v_{m}=40 m/s, maximum density ρm=160\rho_{m}=160 veh/km, equilibrium density ρ1=120\rho_{1}^{\star}=120 veh/km, driver reaction time τ=60\tau=60 s. Let γ0=1\gamma_{0}=1. Initial conditions are sinusoidal inputs ρ1(x,0)=ρ1+0.1sin(3πxL)ρ1\rho_{1}(x,0)=\rho_{1}^{\star}+0.1\sin(\frac{3\pi x}{L})\rho_{1}^{\star} and v1(x,0)=v10.01sin(3πxL)vv_{1}(x,0)=v_{1}^{\star}-0.01\sin(\frac{3\pi x}{L})v^{\star} to mimic stop-and-go traffic. To generate a sufficient dataset for training, we use 10 different c(x)c(x) functions with τU[50,70]\tau\in U[50,70] and simulate the resulting PDEs under the adaptive controller for T=300T=300 seconds. We sub-sample each (c,K^u1,K^v1)(c,\hat{K}^{u_{1}},\hat{K}^{v_{1}}) pair every 0.1 seconds, resulting in a total of 30,000 distinct (c,K^u1,K^v1)(c,\hat{K}^{u_{1}},\hat{K}^{v_{1}}) pairs for training the NO. Using the trained NO, we simulate with the same parameters. Figure 11 shows the ARZ system is open-loop unstable. Figures LABEL:rho_v show the density and velocity of ARZ traffic system with the adaptive controller (134) with the update law (135) and NO-based adaptive controller (141). The blue line indicates the initial condition, whereas the red line represents the boundary condition of the system. The results indicate that both the NO-based adaptive method and the adaptive backstepping control method effectively stabilize the transportation system. The traffic density and velocity converge to the equilibrium values of ρ1=120\rho_{1}^{\star}=120 veh/km and v1=36v_{1}^{\star}=36 m/s, respectively. The relative L2L_{2} error between the closed-loop states of the nominal adaptive controller and the NO-based adaptive controller is shown in Figures LABEL:rho_v. The maximum error does not exceed 10%\%. The estimated parameter c^\hat{c} is shown in Figure 17.

Table 3 presents the computation times for both the nominal adaptive controller and the NO-based adaptive controller. As the baseline result, the nominal adaptive control method is compared with the NO-based adaptive control method. Notably, the NO-based adaptive control method not only achieves significantly faster average computation times but also maintains superior accuracy with a lower relative L2L_{2} error. These advantages of the NO-based adaptive control method not only enhance computational efficiency but also make it highly suitable for real-time traffic system applications. The NO method’s efficiency and accuracy represent a substantial advancement, promising more effective and scalable traffic control strategies in practical scenarios.

Refer to caption
Figure 11: Density and velocity evolution of open-loop ARZ traffic system
Refer to caption
Figure 17: Parameter estimation of c^\hat{c} in ARZ traffic system and comparision of true cc and c^\hat{c}

6 Conclusion

This paper builds upon prior studies such as [42] and [43], which primarily focused on approximating a single kernel PDE. In contrast, our work accelerates the computation of 2×\times2 coupled Goursat-form PDEs that arise from the application of adaptive control designs to a 2×\times2 linear first-order hyperbolic system. In this paper, the DeepONet is used to learn the adaptive control gains for stabilizing the traffic PDE system, and it is shown that under the DeepONet-approximated kernels the stabilization of 2×\times2 hyperbolic PDEs can still be achieved with significant improvement for computational speeds. Experimental results show that compared to traditional numerical solvers, our method improves computational efficiency by two orders of magnitude. Additionally, compared with RL, the NO-based adaptive control strategy is independent of the system’s initial conditions, making it more robust for rapidly changing traffic scenarios. Our method significantly accelerates the process of obtaining adaptive controllers in PDE systems, greatly improving the real-time applicability of adaptive control strategies for mitigating traffic congestion. In the future, we will incorporate real traffic data into the training of the neural operator.

Appendix

The proof of Lemma 1.

Property (42) follows trivially from projection in (35)-(38) and Lemma A.1 in [22]. The result can be easily obtained using the following Lyapunov function candidate:

V1(t)=\displaystyle V_{1}(t)= V2(t)+γ11c~12+γ21c~22+γ31c~32\displaystyle V_{2}(t)+\gamma_{1}^{-1}\|\tilde{c}_{1}\|^{2}+\gamma_{2}^{-1}\|\tilde{c}_{2}\|^{2}+\gamma_{3}^{-1}\|\tilde{c}_{3}\|^{2}
+γ41c~42+12γ51r~2(t),\displaystyle+\gamma_{4}^{-1}\|\tilde{c}_{4}\|^{2}+\dfrac{1}{2}\gamma_{5}^{-1}\tilde{r}^{2}(t), (142)

where

V2(t)=01eγxe2(x,t)𝑑x+01εγxe2(x,t)𝑑x.\displaystyle V_{2}(t)=\int_{0}^{1}e^{-\gamma x}e^{2}(x,t)dx+\int_{0}^{1}\varepsilon^{\gamma x}e^{2}(x,t)dx. (143)

Computing the time derivative of (142) along (31)-(34) as

V˙1(t)=\displaystyle\dot{V}_{1}(t)= 201eγxe(x,t)et(x,t)+eγxε(x,t)εt(x,t)dx\displaystyle 2\int_{0}^{1}e^{-\gamma x}e(x,t)e_{t}(x,t)+e^{\gamma x}\varepsilon(x,t)\varepsilon_{t}(x,t)dx
+2γ1101c~1(x,t)c~1t(x,t)𝑑x\displaystyle+2\gamma_{1}^{-1}\int_{0}^{1}\tilde{c}_{1}(x,t)\tilde{c}_{1t}(x,t)dx
+2γ2101c~2(x,t)c~2t(x,t)𝑑x\displaystyle+2\gamma_{2}^{-1}\int_{0}^{1}\tilde{c}_{2}(x,t)\tilde{c}_{2t}(x,t)dx
+2γ3101c~3(x,t)c~3t(x,t)𝑑x\displaystyle+2\gamma_{3}^{-1}\int_{0}^{1}\tilde{c}_{3}(x,t)\tilde{c}_{3t}(x,t)dx
+2γ4101c~4(x,t)c~4t(x,t)𝑑x+λγ51r~(t)r~˙(t).\displaystyle+2\gamma_{4}^{-1}\int_{0}^{1}\tilde{c}_{4}(x,t)\tilde{c}_{4t}(x,t)dx+\lambda\gamma_{5}^{-1}\tilde{r}(t)\dot{\tilde{r}}(t). (144)

We put into the dynamics (31)-(34), integrating by parts, thus we have

V˙1(t)=\displaystyle\dot{V}_{1}(t)= λeγe2(1,t)+λe2(0,t)λγ01eγxe2(x,t)𝑑x\displaystyle-\lambda e^{-\gamma}e^{2}(1,t)+\lambda e^{2}(0,t)-\lambda\gamma\int_{0}^{1}e^{-\gamma x}e^{2}(x,t)dx
+201eγxe(x,t)ϖT(x,t)c~1(x,t)𝑑x\displaystyle+2\int_{0}^{1}e^{-\gamma x}e(x,t)\varpi^{T}(x,t)\tilde{c}_{1}(x,t)dx
2ρ01eγxe2(x,t)ϖ(t)2𝑑x\displaystyle-2\rho\int_{0}^{1}e^{-\gamma x}e^{2}(x,t)\|\varpi(t)\|^{2}dx
μϵ2(0,t)μγ01eγxϵ2(x,t)𝑑x\displaystyle-\mu\epsilon^{2}(0,t)-\mu\gamma\int_{0}^{1}e^{\gamma x}\epsilon^{2}(x,t)dx
+201eγxϵ(x,t)ϖT(x,t)c~2(x,t)𝑑x\displaystyle+2\int_{0}^{1}e^{\gamma x}\epsilon(x,t)\varpi^{T}(x,t)\tilde{c}_{2}(x,t)dx
2ρ01eγxϵ2(x,t)ϖ(t)2𝑑x\displaystyle-2\rho\int_{0}^{1}e^{\gamma x}\epsilon^{2}(x,t)\|\varpi(t)\|^{2}dx
+2γ1101c~1(x,t)c~1t(x,t)𝑑x\displaystyle+2\gamma_{1}^{-1}\int_{0}^{1}\tilde{c}_{1}(x,t)\tilde{c}_{1t}(x,t)dx
+2γ2101c~2(x,t)c~2t(x,t)𝑑x\displaystyle+2\gamma_{2}^{-1}\int_{0}^{1}\tilde{c}_{2}(x,t)\tilde{c}_{2t}(x,t)dx
+2γ3101c~3(x,t)c~3t(x,t)𝑑x\displaystyle+2\gamma_{3}^{-1}\int_{0}^{1}\tilde{c}_{3}(x,t)\tilde{c}_{3t}(x,t)dx
+2γ4101c~4(x,t)c~4t(x,t)𝑑x+λγ51r~(t)r~˙(t).\displaystyle+2\gamma_{4}^{-1}\int_{0}^{1}\tilde{c}_{4}(x,t)\tilde{c}_{4t}(x,t)dx+\lambda\gamma_{5}^{-1}\tilde{r}(t)\dot{\tilde{r}}(t). (145)

Inserting the adaptive laws (35)-(39), and using the property (40), give

V˙1(t)λeγe2(1,t)+λe2(0,t)λγ01eγxe2(x,t)𝑑x2ρ01eγxe2(x,t)ϖ(t)2𝑑xμϵ2(0,t)μγ01eγxϵ2(x,t)𝑑x2ρ01eγxϵ2(x,t)ϖ(t)2𝑑xλr~(t)e(0,t)v(0,t).\begin{split}\dot{V}_{1}(t)\leq&-\lambda e^{-\gamma}e^{2}(1,t)+\lambda e^{2}(0,t)-\lambda\gamma\int_{0}^{1}e^{-\gamma x}e^{2}(x,t)dx\\ &-2\rho\int_{0}^{1}e^{-\gamma x}e^{2}(x,t)\|\varpi(t)\|^{2}dx\\ &-\mu\epsilon^{2}(0,t)-\mu\gamma\int_{0}^{1}e^{\gamma x}\epsilon^{2}(x,t)dx\\ &-2\rho\int_{0}^{1}e^{\gamma x}\epsilon^{2}(x,t)\|\varpi(t)\|^{2}dx-\lambda\tilde{r}(t)e(0,t)v(0,t).\end{split} (146)

From the boundary condition (33), we have

e(0,t)=r~(t)v(0,t)e(0,t)v2(0,t),e(0,t)=\tilde{r}(t)v(0,t)-e(0,t)v^{2}(0,t), (147)

By substituting this, we obtain

V˙1(t)\displaystyle\dot{V}_{1}(t)\leq λeγe2(1,t)λe2(0,t)v2(0,t)λγeγe(t)2\displaystyle-\lambda e^{-\gamma}e^{2}(1,t)-\lambda e^{2}(0,t)v^{2}(0,t)-\lambda\gamma e^{-\gamma}\|e(t)\|^{2}
2ρeγe(t)2ϖ(t)2μϵ2(0,t)\displaystyle-2\rho e^{-\gamma}\|e(t)\|^{2}\|\varpi(t)\|^{2}-\mu\epsilon^{2}(0,t)
μγϵ(t)22ρeγϵ(t)2ϖ(t)2,\displaystyle-\mu\gamma\|\epsilon(t)\|^{2}-2\rho e^{\gamma}\|\epsilon(t)\|^{2}\|\varpi(t)\|^{2}, (148)

This establishes that V1V_{1} is bounded. By the definitions of V1V_{1} and V2V_{2}, it follows that e,ε\|e\|,\|\varepsilon\|\in\mathcal{L}_{\infty}. When (Appendix) is integrated over time from zero to infinity, we conclude that e,ε2\|e\|,\|\varepsilon\|\in\mathcal{L}_{2}. Additionally, from the properties (44), |e(0,)v(0,)|2|e(0,\cdot)v(0,\cdot)|\in\mathcal{L}_{2} and the adaptive laws (35)-(38), we derive that (45). we choose the following Lyapunov function candidate

V2(t)=12γ51r~2(t),V_{2}(t)=\frac{1}{2}\gamma_{5}^{-1}\tilde{r}^{2}(t), (149)

and use the property (40), we find

V˙2(t)r~(t)e(0,t)v(0,t)r~2(t)v2(0,t)1+v2(0,t).\dot{V}_{2}(t)\leq-\tilde{r}(t)e(0,t)v(0,t)\leq-\frac{\tilde{r}^{2}(t)v^{2}(0,t)}{1+v^{2}(0,t)}. (150)

This implies that V2V_{2} is upper-bounded, and hence we have V3V_{3}\in\mathcal{L}_{\infty}. By integrating (150) from zero to infinity, we obtain (47). Using (147) and (33), we derive that

e2(0,t)=e(0,t)(r~(t)v(0,t)e(0,t)v2(0,t))=r~2(t)v2(0,t)1+v2(0,t)e2(0,t)v2(0,t),\begin{split}e^{2}(0,t)&=e(0,t)\left(\tilde{r}(t)v(0,t)-e(0,t)v^{2}(0,t)\right)\\ &=\frac{\tilde{r}^{2}(t)v^{2}(0,t)}{1+v^{2}(0,t)}-e^{2}(0,t)v^{2}(0,t),\end{split} (151)

and from |e(0,)v(0,)|2|e(0,\cdot)v(0,\cdot)|\in\mathcal{L}_{2} and (47), (147) and (33), we have e(0,)2e(0,\cdot)\in\mathcal{L}_{2}.

References

  • [1] F. Belletti, M. Huo, X. Litrico, A. M. Bayen, Prediction of traffic convective instability with spectral analysis of the aw–rascle–zhang model, Physics Letters A 379 (38) (2015) 2319–2330.
  • [2] A. De Palma, R. Lindsey, Traffic congestion pricing methodologies and technologies, Transportation Research Part C: Emerging Technologies 19 (6) (2011) 1377–1399.
  • [3] M. J. Lighthill, G. B. Whitham, On kinematic waves ii. a theory of traffic flow on long crowded roads, Proceedings of the royal society of london. series a. mathematical and physical sciences 229 (1178) (1955) 317–345.
  • [4] P. I. Richards, Shock waves on the highway, Operations research 4 (1) (1956) 42–51.
  • [5] A. Aw, M. Rascle, Resurrection of” second order” models of traffic flow, SIAM journal on applied mathematics 60 (3) (2000) 916–938.
  • [6] H. M. Zhang, A non-equilibrium traffic model devoid of gas-like behavior, Transportation Research Part B: Methodological 36 (3) (2002) 275–290.
  • [7] N. Bekiaris-Liberis, A. Delis, Feedback control of freeway traffic flow via time-gap manipulation of acc-equipped vehicles: A pde-based approach, IFAC-PapersOnLine 52 (6) (2019) 1–6.
  • [8] L. Zhang, C. Prieur, J. Qiao, Pi boundary control of linear hyperbolic balance laws with stabilization of arz traffic flow models, Systems & Control Letters 123 (2019) 85–91.
  • [9] I. Karafyllis, N. Bekiaris-Liberis, M. Papageorgiou, Feedback control of nonlinear hyperbolic pde systems inspired by traffic flow models, IEEE Transactions on Automatic Control 64 (9) (2018) 3647–3662.
  • [10] H. Yu, M. Krstic, Traffic congestion control for aw–rascle–zhang model, Automatica 100 (2019) 38–51.
  • [11] M. Burkhardt, H. Yu, M. Krstic, Stop-and-go suppression in two-class congested traffic, Automatica 125 (2021) 109381.
  • [12] H. Yu, M. Krstic, Adaptive output feedback for aw-rascle-zhang traffic model in congested regime, in: 2018 Annual American Control Conference (ACC), IEEE, 2018, pp. 3281–3286.
  • [13] H. Yu, M. Krstic, Traffic congestion control by PDE backstepping, Springer, 2022.
  • [14] H. Yu, M. Krstic, Varying speed limit control of aw-rascle-zhang traffic model, in: 2018 21st international conference on intelligent transportation systems (ITSC), IEEE, 2018, pp. 1846–1851.
  • [15] H. Yu, J. Auriol, M. Krstic, Simultaneous downstream and upstream output-feedback stabilization of cascaded freeway traffic, Automatica 136 (2022) 110044.
  • [16] Y. Zhang, H. Yu, J. Auriol, M. Pereira, Mean-square exponential stabilization of mixed-autonomy traffic pde system, arXiv preprint arXiv:2310.15547 (2023).
  • [17] M. Krstic, A. Smyshlyaev, Adaptive boundary control for unstable parabolic pdes—part i: Lyapunov design, IEEE Transactions on Automatic Control 53 (7) (2008) 1575–1591.
  • [18] A. Smyshlyaev, M. Krstic, Adaptive boundary control for unstable parabolic pdes—part ii: Estimation-based designs, Automatica 43 (9) (2007) 1543–1556.
  • [19] F. Di Meglio, D. Bresch-Pietri, U. J. F. Aarsnes, An adaptive observer for hyperbolic systems with application to underbalanced drilling, IFAC Proceedings Volumes 47 (3) (2014) 11391–11397.
  • [20] M. C. Belhadjoudja, M. Maghenem, E. Witrant, C. Prieur, Adaptive stabilization of the kuramoto-sivashinsky equation subject to intermittent sensing, in: 2023 American Control Conference (ACC), IEEE, 2023, pp. 1608–1613.
  • [21] C. Kawan, A. Mironchenko, M. Zamani, A lyapunov-based iss small-gain theorem for infinite networks of nonlinear systems, IEEE Transactions on Automatic Control 68 (3) (2022) 1447–1462.
  • [22] H. Anfinsen, O. M. Aamo, Adaptive control of hyperbolic PDEs, Springer, 2019.
  • [23] M. Krstic, A. Smyshlyaev, Backstepping boundary control for first-order hyperbolic pdes and application to systems with actuator and sensor delays, Systems & Control Letters 57 (9) (2008) 750–758.
  • [24] A. Smyshlyaev, E. Cerpa, M. Krstic, Boundary stabilization of a 1-d wave equation with in-domain antidamping, SIAM journal on control and optimization 48 (6) (2010) 4014–4031.
  • [25] H. Anfinsen, O. M. Aamo, Adaptive control of linear2×\times 2 hyperbolic systems, Automatica 87 (2018) 69–82.
  • [26] L. Hu, F. Di Meglio, R. Vazquez, M. Krstic, Control of homodirectional and general heterodirectional linear coupled hyperbolic pdes, IEEE Transactions on Automatic Control 61 (11) (2015) 3301–3314.
  • [27] J. Wang, M. Krstic, Event-triggered backstepping control of 2×\times 2 hyperbolic pde-ode systems, IFAC-PapersOnLine 53 (2) (2020) 7551–7556.
  • [28] J. Auriol, Output feedback stabilization of an underactuated cascade network of interconnected linear pde systems using a backstepping approach, Automatica 117 (2020) 108964.
  • [29] S. Mowlavi, S. Nabi, Optimal control of pdes using physics-informed neural networks, Journal of Computational Physics 473 (2023) 111731.
  • [30] C. Zhao, H. Yu, Observer-informed deep learning for traffic state estimation with boundary sensing, IEEE Transactions on Intelligent Transportation Systems (2023).
  • [31] C. Wu, A. Kreidieh, E. Vinitsky, A. M. Bayen, Emergent behaviors in mixed-autonomy traffic, in: Conference on Robot Learning, PMLR, 2017, pp. 398–407.
  • [32] X. Qu, Y. Yu, M. Zhou, C.-T. Lin, X. Wang, Jointly dampening traffic oscillations and improving energy consumption with electric, connected and automated vehicles: A reinforcement learning based approach, Applied Energy 257 (2020) 114030.
  • [33] H. Yu, S. Park, A. Bayen, S. Moura, M. Krstic, Reinforcement learning versus pde backstepping and pi control for congested freeway traffic, IEEE Transactions on Control Systems Technology 30 (4) (2021) 1595–1611.
  • [34] L. Lu, P. Jin, G. Pang, Z. Zhang, G. E. Karniadakis, Learning nonlinear operators via deeponet based on the universal approximation theorem of operators, Nature machine intelligence 3 (3) (2021) 218–229.
  • [35] Y. Shi, Z. Li, H. Yu, D. Steeves, A. Anandkumar, M. Krstic, Machine learning accelerated pde backstepping observers, in: 2022 IEEE 61st Conference on Decision and Control (CDC), IEEE, 2022, pp. 5423–5428.
  • [36] L. Bhan, Y. Shi, M. Krstic, Neural operators for bypassing gain and control computations in pde backstepping, arXiv preprint arXiv:2302.14265 (2023).
  • [37] M. Krstic, L. Bhan, Y. Shi, Neural operators of backstepping controller and observer gain functions for reaction–diffusion pdes, Automatica 164 (2024) 111649.
  • [38] J. Qi, J. Zhang, M. Krstic, Neural operators for pde backstepping control of first-order hyperbolic pide with recycle and delay, Systems & Control Letters 185 (2024) 105714.
  • [39] S. Wang, M. Diagne, M. Krstić, Deep learning of delay-compensated backstepping for reaction-diffusion pdes, arXiv preprint arXiv:2308.10501 (2023).
  • [40] Y. Zhang, R. Zhong, H. Yu, Neural operators for boundary stabilization of stop-and-go traffic, in: 6th Annual Learning for Dynamics & Control Conference, PMLR, 2024, pp. 554–565.
  • [41] M. Lamarque, L. Bhan, R. Vazquez, M. Krstic, Gain scheduling with a neural operator for a transport pde with nonlinear recirculation, arXiv preprint arXiv:2401.02511 (2024).
  • [42] M. Lamarque, L. Bhan, Y. Shi, M. Krstic, Adaptive neural-operator backstepping control of a benchmark hyperbolic pde, arXiv preprint arXiv:2401.07862 (2024).
  • [43] L. Bhan, Y. Shi, M. Krstic, Adaptive control of reaction-diffusion pdes via neural operator-approximated gain kernels, arXiv preprint arXiv:2407.01745 (2024).
  • [44] M. Krstic, Delay compensation for nonlinear, adaptive, and pde systems (2009).
  • [45] B. Deng, Y. Shin, L. Lu, Z. Zhang, G. E. Karniadakis, Approximation rates of deeponets for learning operators arising from advection–diffusion equations, Neural Networks 153 (2022) 411–426.
  • [46] F. Di Meglio, R. Vazquez, M. Krstic, Stabilization of a system of n+1n+1 coupled first-order hyperbolic linear pdes with a single boundary input, IEEE Transactions on Automatic Control 58 (12) (2013) 3097–3111.
  • [47] R. V. Valenzuela, Boundary control laws and observer design for convective, turbulent and magnetohydrodynamic flows, University of California, San Diego, 2006.
  • [48] M. Krstic, P. V. Kokotovic, I. Kanellakopoulos, Nonlinear and adaptive control design, John Wiley & Sons, Inc., 1995.
  • [49] R. Vazquez, M. Krstic, J.-M. Coron, Backstepping boundary stabilization and state estimation of a 2×\times 2 linear hyperbolic system, in: 2011 50th IEEE conference on decision and control and european control conference, IEEE, 2011, pp. 4937–4942.