Skip to main content
Log in

ISS-based robustness to various neglected damping mechanisms for the 1-D wave PDE

  • Original Article
  • Published:
Mathematics of Control, Signals, and Systems Aims and scope Submit manuscript

Abstract

This paper is devoted to the study of the robustness properties of the 1-D wave equation for an elastic vibrating string under four different damping mechanisms that are usually neglected in the study of the wave equation: (i) friction with the surrounding medium of the string (or viscous damping), (ii) thermoelastic phenomena (or thermal damping), (iii) internal friction of the string (or Kelvin-Voigt damping), and (iv) friction at the free end of the string (the so-called passive damper). The passive damper is also the simplest boundary feedback law that guarantees exponential stability for the string. We study robustness with respect to distributed inputs and boundary disturbances in the context of Input-to-State Stability (ISS). By constructing appropriate ISS Lyapunov functionals, we prove the ISS property expressed in various spatial norms.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Auriol J, Aarsnes UJF, Martin P, Di Meglio F (2018) Delay-robust control design for two heterodirectional linear coupled hyperbolic PDEs. IEEE Trans Autom Control 63:3551–3557

    Article  MathSciNet  MATH  Google Scholar 

  2. Balakrishnan AV (1999) On superstability of semigroups. In: systems modelling and optimization. Chapman & Hall/CRC Res, Detroit

  3. Bekiaris-Liberis N, Krstic M (2014) Compensation of wave actuator dynamics for nonlinear systems. IEEE Trans Autom Control 59:1555–1570

    Article  MathSciNet  MATH  Google Scholar 

  4. Bresch-Pietri D, Krstic M (2014) Output-feedback adaptive control of a wave pde with boundary anti-damping. Automatica 50:1407–1415

    Article  MathSciNet  MATH  Google Scholar 

  5. Brezis H (2011) Functional analysis, Sobolev spaces and partial differential equations. Springer, Berlin

    Book  MATH  Google Scholar 

  6. Chen S, Liu K, Liu Z (1998) Spectrum and stability for elastic systems with global or local kelvin-voigt damping. SIAM J Appl Math 59:651–668

    Article  MathSciNet  MATH  Google Scholar 

  7. Chitour Y, Marx S, Mazanti G (2021) One-dimensional wave equation with set-valued boundary damping: well-posedness, asymptotic stability, and decay rates. ESAIM: Control, Optim Calc Variat 27:84

    MathSciNet  MATH  Google Scholar 

  8. Coron J-M (2007) Control and nonlinearity, mathematical surveys and monographs, vol 136, American Mathematical Society

  9. Datko R, Lagnese J, Polis MP (1986) An example on the effect of time delays in boundary feedback stabilization of wave equations. SIAM J Control Optim 24:152–156

    Article  MathSciNet  MATH  Google Scholar 

  10. Datko R (1988) Not all feedback stabilized hyperbolic systems are robust with respect to small time delays in their feedbacks. SIAM J Control Optim 26:697–713

    Article  MathSciNet  MATH  Google Scholar 

  11. Day WA (1982) Extension of a property of the heat equation to linear thermoelasticity and other theories. Q Appl Math 40:319–330

    Article  MathSciNet  MATH  Google Scholar 

  12. Day WA (1983) A decreasing property of solutions of parabolic equations with applications to thermoelasticity. Q Appl Math 40:468–475

    Article  MathSciNet  MATH  Google Scholar 

  13. Feng H, Li S (2013) The stability for a one-dimensional wave equation with nonlinear uncertainty on the boundary. Nonlinear Anal Theory Methods Appl 89:202–207

    Article  MathSciNet  MATH  Google Scholar 

  14. Freitas P, Zuazua E (1996) Stability results for the wave equation with indefinite damping. J Differ Equ 132:338–353

    Article  MathSciNet  MATH  Google Scholar 

  15. Guenther RB, Lee JW (1996) Partial differential equations of mathematical physics and integral equations. Dover

  16. Guo BZ, Luo YH (2002) Controllability and stability of a second order hyperbolic system with collocated sensor/actuator. Syst Control Lett 46:45–65

    Article  MathSciNet  MATH  Google Scholar 

  17. Guo BZ, Xu CZ (2007) The stabilization of a one-dimensional wave equation by boundary feedback with non-collocated observation. IEEE Trans Autom Control 52:371–377

    Article  MATH  Google Scholar 

  18. Hayat A (2021) Global exponential stability and input-to-state stability of semilinear hyperbolic systems for the \(L^{2} \)-Norm. Syst Control Lett 148:104848

    Article  MATH  Google Scholar 

  19. Kafnemer M, Mebkhout B, Chitour Y (2021) Weak input to state estimates for 2D damped wave equations with localized and nonlinear damping. SIAM J Control Optim 59:1604–1627

    Article  MathSciNet  MATH  Google Scholar 

  20. Karafyllis I, Krstic M (2019) Input-to-state stability for PDEs. Springer-Verlag, London

    Book  MATH  Google Scholar 

  21. Karafyllis I, Kontorinaki M, Krstic M (2021) Boundary-to-displacement asymptotic gains for wave systems with kelvin-Voigt damping. Int J Control 94:2822–2833

    Article  MathSciNet  MATH  Google Scholar 

  22. Karlsen JT, Bruus H (2015) Forces acting on a small particle in an acoustical field in a thermoviscous fluid. Phys Rev E 92:043010

    Article  Google Scholar 

  23. Komornik V, Zuazua E (1990) A direct method for the boundary stabilization of the wave equation. J de Math Pures et Appliquées 69:33–54

    MathSciNet  MATH  Google Scholar 

  24. Krstic M (2010) Adaptive control of an anti-stable wave PDE. Dynamics of continuous, discrete, and impulsive systems, invited paper in the special issue in the honor of Professor Hassan Khalil. vol 17, pp 853–882

  25. Krstic M, Smyshlyaev A (2008) Boundary control of PDEs: A course on backstepping designs. SIAM

  26. Lhachemi H, Saussie D, Zhu G, Shorten R (2020) Input-to-state stability of a clamped-free damped string in the presence of distributed and boundary disturbances. IEEE Trans Autom Control 65:1248–1255

    Article  MathSciNet  MATH  Google Scholar 

  27. Mironchenko A, Prieur C (2020) Input-to-state stability of infinite-dimensional systems: recent results and open questions. SIAM Rev 62:529–614

    Article  MathSciNet  MATH  Google Scholar 

  28. Morgul O (1995) On the stabilization and stability robustness against small delays of some damped wave equations. IEEE Trans Autom Control 40:1626–1630

    Article  MathSciNet  MATH  Google Scholar 

  29. Natarajan V (2021) Compensating PDE actuator and sensor dynamics using Sylvester equation. Automatica 123:109362

    Article  MathSciNet  MATH  Google Scholar 

  30. Nicaise S, Pignotti C (2006) Stability and instability results of the wave equation with a delay term in the boundary or internal feedbacks. SIAM J Control Optim 45:1561–1585

    Article  MathSciNet  MATH  Google Scholar 

  31. Pazy A (1983) Semigroups of linear operators and applications to partial differential equations. Springer, New York

    Book  MATH  Google Scholar 

  32. Smyshlyaev A, Krstic M (2009) Boundary control of an anti-stable wave equation with anti-damping on the uncontrolled boundary. Syst Control Lett 58:617–623

    Article  MathSciNet  MATH  Google Scholar 

  33. Smyshlyaev A, Cerpa E, Krstic M (2010) Boundary stabilization of a 1-D wave equation with in-domain anti-damping. SIAM J Control Optim 48:4014–4031

    Article  MathSciNet  MATH  Google Scholar 

  34. Sontag ED (1989) Smooth stabilization implies coprime factorization. IEEE Trans Autom Control 34:435–443

    Article  MathSciNet  MATH  Google Scholar 

  35. Temam RM, Miranville AM (2005) Mathematical modeling in continuum mechanics, 2nd edn. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  36. Vazquez R, Krstic M (2016) Bilateral boundary control of one-dimensional first- and second-order PDEs using infinite-dimensional backstepping. In: Proceedings of the 55th IEEE conference on decision and control (CDC)

  37. Zhang Q (2010) Exponential stability of an elastic string with local kelvin-Voigt damping. Z Angew Math Phys 61:1009–1015

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Iasson Karafyllis.

Ethics declarations

Conflict of interest

The authors certify that they have no affiliations with or involvement in any organization or entity with any financial interest (such as honoraria; educational grants; participation in speakers’ bureaus; membership, employment, consultancies, stock ownership, or other equity interest; and expert testimony or patent-licensing arrangements), or non-financial interest (such as personal or professional relationships, affiliations, knowledge or beliefs) in the subject matter or materials discussed in this manuscript.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

To Eduardo, with gratitude for his deep, elegant ISS and CLF results, in which he had encoded a general vision that enabled their extensions to stochastic, delay, and PDE systems.

Appendix A Well-posedness of 1-D wave models

Appendix A Well-posedness of 1-D wave models

We next present the well-posedness analysis for all studied models. However, it should be noted that such an analysis provides sufficient and not necessary conditions for the existence/uniqueness of a solution with specific regularity properties. On the other hand, Theorem 1, Theorem 2 and Theorem 3 assume less regularity properties for the external inputs. For example, the well-posedness analysis given below for Model (D) requires that \(f\in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \) is a sufficient condition for the existence/uniqueness of a solution for Model (D), while Theorem 3 simply requires that \(f\in C^{0} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \). This is important because one cannot exclude the possibility that a solution for Model (D) with the regularity properties that are specified in Theorem 3 exists for less regular inputs (see the discussion on pages 105-110 in [31]).

The results that follow use a similar notation (for example, the state is U, the state space is X, the linear unbounded operator is A, etc.). However, the reader should not be tempted, by an overlapping notation for different quantities, to compare the different quantities in the analysis of each one of the models.

(1) Well-posedness analysis of Model (B).

Model (B) consists of equations (2), (5), (6). Applying the transformation

$$\begin{aligned} w(t,x)=u_{t} (t,x)\quad v(t,x)=u(t,x)-d(t)x, \ for \ t\ge 0,x\in [0,1] \end{aligned}$$
(A1)

and assuming that \(d\in C^{1} \left( {\mathbb {R}} _{+} \right) \), we obtain the following first-order in time model

$$\begin{aligned} U_{t}+AU=F \end{aligned}$$
(A2)

where

$$\begin{aligned}{} & {} U=(v,w) \end{aligned}$$
(A3)
$$\begin{aligned}{} & {} F(t,x)=\left( -{\dot{d}}(t)x,f(t,x)\right) , \ for \ t\ge 0,x\in [0,1] \end{aligned}$$
(A4)

and \(A:D(A)\rightarrow X\) is the linear unbounded operator defined by

$$\begin{aligned} AU=\left( -w,-c^{2} v''+\mu w\right) , for \ U=(v,w)\in D(A) \end{aligned}$$
(A5)

with X being the Hilbert space

$$\begin{aligned} X=\left\{ \, v\in H^{1} (0,1)\,:\, v(0)=0\, \right\} \times L^{2} (0,1) \end{aligned}$$
(A6)

with scalar product

$$\begin{aligned} \left( U,{\bar{U}}\right)= & {} c^{2} \int _{0}^{1}v'(x){\bar{v}}'(x)dx +\int _{0}^{1}v(x){\bar{v}}(x)dx +\int _{0}^{1}w(x){\bar{w}}(x)dx, \nonumber \\{} & {} \qquad for \ U=(v,w)\in X, \ {\bar{U}}=({\bar{v}},{\bar{w}})\in X \end{aligned}$$
(A7)

and \(D(A)\subset X\) being the following linear subspace of X

$$\begin{aligned} D(A)=\left\{ \, (v,w)\in H^{2} (0,1)\times H^{1} (0,1)\,:\, v(0)=w(0)=v'(1)+aw(1)=0\, \right\} \nonumber \\ \end{aligned}$$
(A8)

Let \(I:X\rightarrow X\) be the identity operator. We next show that \(A+I\) is a maximal monotone operator (see [5]). Indeed, we have for all \(U=(v,w)\in D(A)\)

$$\begin{aligned} \left( (A+I)U,U\right)= & {} -c^{2} \int _{0}^{1}w'(x)v'(x)dx -\int _{0}^{1}w(x)v(x)dx -c^{2} \int _{0}^{1}v''(x)w(x)dx \\{} & {} {+(\mu +1)\left\| w\right\| _{2}^{2} +c^{2} \left\| v'\right\| _{2}^{2} +\left\| v\right\| _{2}^{2} } \\= & {} -c^{2} w(1)v'(1)-\int _{0}^{1}w(x)v(x)dx +(\mu +1)\left\| w\right\| _{2}^{2} +c^{2} \left\| v'\right\| _{2}^{2} +\left\| v\right\| _{2}^{2} \\\ge & {} ac^{2} w^{2} (1)-\left\| w\right\| _{2} \left\| v\right\| _{2} +(\mu +1)\left\| w\right\| _{2}^{2} +c^{2} \left\| v'\right\| _{2}^{2} +\left\| v\right\| _{2}^{2} \\\ge & {} ac^{2} w^{2} (1)+\left( \mu +\frac{1}{2} \right) \left\| w\right\| _{2}^{2} +c^{2} \left\| v'\right\| _{2}^{2} +\frac{1}{2} \left\| v\right\| _{2}^{2} \ge 0 \end{aligned}$$

In the above we have used (A5), (A7), (A8), integration by parts, the Cauchy-Schwarz inequality and the fact that \(\left\| w\right\| _{2} \left\| v\right\| _{2} \le \frac{1}{2} \left\| w\right\| _{2}^{2} +\frac{1}{2} \left\| v\right\| _{2}^{2} \). Moreover, for every \({\bar{F}}=\left( {\bar{F}}_{1},{\bar{F}}_{2} \right) \in X\) the equation \((A+2I)U={\bar{F}}\in X\) has a unique solution \(U=(v,w)\in D(A)\). Indeed, the equation \((A+2I)U={\bar{F}}\in X\) gives \(w=2v-{\bar{F}}_{1} \) and \(v(x)={\tilde{v}}(x)+\frac{a{\bar{F}}_{1} (1)}{2a+1} x\) for \(x\in [0,1]\), where \({\tilde{v}}\in H^{2} (0,1)\) is the unique solution of the boundary value problem

$$\begin{aligned}{} & {} {-c^{2} {\tilde{v}}''(x)+2(\mu +2){\tilde{v}}(x)=(\mu +2){\bar{F}}_{1} (x)+{\bar{F}}_{2} (x)-2(\mu +2)\frac{a{\bar{F}}_{1} (1)}{2a+1} x, }\nonumber \\{} & {} \quad {\ for \ x\in (0,1) \ a.e., \ with \ {\tilde{v}}(0)=0 \ and \ {\tilde{v}}'(1)=-2a{\tilde{v}}(1) } \end{aligned}$$
(A9)

The fact that the boundary-value problem (A9) has a unique solution follows from the methodology described on pages 221-229 in [5] and the Lax-Milgram Theorem (Corollary 5.8 on page 140 in [5]).

Thus \(A+I\) is a maximal monotone operator and consequently (using the Hille-Yosida Theorem) A is the generator of a continuous semigroup of contractions \(S(t):X\rightarrow X\). It follows from Theorem 7.10 on page 198 in [5] that for every \(F\in C^{1} \left( {\mathbb {R}} _{+};X\right) \) and for every \(U_{0} \in D(A)\) there exists a unique solution \(U\in C^{1} \left( {\mathbb {R}} _{+};X\right) \cap C^{0} \left( {\mathbb {R}} _{+};D(A)\right) \) of the initial-value problem (A.2) with initial condition \(U(0)=U_{0} \).

Going back to the original state variables (and using (A1)), we conclude that for every \(d\in C^{2} \left( {\mathbb {R}} _{+} \right) \), \(f\in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \) and for every \(u_{0} \in H^{2} (0,1)\), \(w_{0} \in H^{1} (0,1)\) with \(u_{0} (0)=w_{0} (0)=0\), \(u'_{0} (1)=-aw_{0} (1)+d(0)\) there exists a unique solution \(u\in C^{1} \left( {\mathbb {R}} _{+};H^{1} (0,1)\right) \cap C^{0} \left( {\mathbb {R}} _{+};H^{2} (0,1)\right) \) with \(u_{t} \in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \cap C^{0} \left( {\mathbb {R}} _{+};H^{1} (0,1)\right) \) of the initial-boundary value problem (2), (5), (6) with \(u[0]=u_{0} \), \(u_{t} [0]=w_{0} \) that additionally satisfies \(u_{t} (t,0)=0\) for all \(t\ge 0\).

(2) Well-posedness analysis of Model (C).

Model (C) consists of equations (2), (6), (7), (8), (9). Applying the transformation (A1) and assuming that \(d\in C^{1} \left( {\mathbb {R}} _{+} \right) \), we obtain the first-order in time model (A2), where

$$\begin{aligned}{} & {} U=(v,w,\theta ) \end{aligned}$$
(A10)
$$\begin{aligned}{} & {} F(t,x)=\left( -{\dot{d}}(t)x,f(t,x),0\right) , \ for \ t\ge 0,x\in [0,1] \end{aligned}$$
(A11)

and \(A:D(A)\rightarrow X\) is the linear unbounded operator defined by

$$\begin{aligned} AU=\left( -w,-c^{2} v''+\mu w+b\theta ',-k\theta ''+\lambda w'\right) , \ for \ U=(v,w,\theta )\in D(A)\nonumber \\ \end{aligned}$$
(A12)

with X being the Hilbert space

$$\begin{aligned} X=\left\{ \, v\in H^{1} (0,1)\,:\, v(0)=0\, \right\} \times L^{2} (0,1)\times L^{2} (0,1) \end{aligned}$$
(A13)

with scalar product

$$\begin{aligned} \left( U,{\bar{U}}\right)= & {} c^{2} \int _{0}^{1}v'(x){\bar{v}}'(x)dx +\int _{0}^{1}v(x){\bar{v}}(x)dx \nonumber \\{} & {} {+\int _{0}^{1}w(x){\bar{w}}(x)dx +\frac{b}{\lambda } \int _{0}^{1}\theta (x){\bar{\theta }}(x)dx,} \nonumber \\{} & {} {\qquad \ for \ U=(v,w,\theta )\in X, \ {\bar{U}}=({\bar{v}},{\bar{w}},{\bar{\theta }})\in X } \end{aligned}$$
(A14)

and \(D(A)\subset X\) being the following linear subspace of X

$$\begin{aligned} D(A)= {\left\{ \, (v,w,\theta )\in H^{2} (0,1)\times H^{1} (0,1)\times H^{2} (0,1)\,:\, \begin{array}{c} {v(0)=w(0)=\theta (0)=0} \\ {v'(1)+aw(1)=\theta (1)=0} \end{array}\, \right\} }\nonumber \\ \end{aligned}$$
(A15)

Let \(I:X\rightarrow X\) be the identity operator. We next show that \(A+I\) is a maximal monotone operator (see [5]). Indeed, we have for all \(U=(v,w,\theta )\in D(A)\)

$$\begin{aligned} \left( (A+I)U,U\right)= & {} -c^{2} \int _{0}^{1}w'(x)v'(x)dx -\int _{0}^{1}w(x)v(x)dx -c^{2} \int _{0}^{1}v''(x)w(x)dx \\{} & {} {+(\mu +1)\left\| w\right\| _{2}^{2} +b\int _{0}^{1}\theta '(x)w(x)dx -\frac{bk}{\lambda } \int _{0}^{1}\theta (x)\theta ''(x)dx } \\{} & {} {+b\int _{0}^{1}\theta (x)w'(x)dx +c^{2} \left\| v'\right\| _{2}^{2} +\left\| v\right\| _{2}^{2} +\frac{b}{\lambda } \left\| \theta \right\| _{2}^{2} } \\= & {} -c^{2} w(1)v'(1)-\int _{0}^{1}w(x)v(x)dx +(\mu +1)\left\| w\right\| _{2}^{2} \\{} & {} {+\frac{bk}{\lambda } \left\| \theta '\right\| _{2}^{2} +c^{2} \left\| v'\right\| _{2}^{2} +\left\| v\right\| _{2}^{2} +\frac{b}{\lambda } \left\| \theta \right\| _{2}^{2} } \\\ge & {} ac^{2} w^{2} (1)-\left\| w\right\| _{2} \left\| v\right\| _{2} +(\mu +1)\left\| w\right\| _{2}^{2} +c^{2} \left\| v'\right\| _{2}^{2} +\left\| v\right\| _{2}^{2} +\frac{bk}{\lambda } \left\| \theta '\right\| _{2}^{2} +\frac{b}{\lambda } \left\| \theta \right\| _{2}^{2} \\\ge & {} ac^{2} w^{2} (1)+\left( \mu +\frac{1}{2} \right) \left\| w\right\| _{2}^{2} +c^{2} \left\| v'\right\| _{2}^{2} +\frac{1}{2} \left\| v\right\| _{2}^{2} +\frac{bk}{\lambda } \left\| \theta '\right\| _{2}^{2} +\frac{b}{\lambda } \left\| \theta \right\| _{2}^{2} \ge 0 \end{aligned}$$

In the above we have used (A12), (A14), (A15), three integrations by parts, the Cauchy-Schwarz inequality and the fact that \(\left\| w\right\| _{2} \left\| v\right\| _{2} \le \frac{1}{2} \left\| w\right\| _{2}^{2} +\frac{1}{2} \left\| v\right\| _{2}^{2} \). Moreover, for every \({\bar{F}}=\left( {\bar{F}}_{1},{\bar{F}}_{2},{\bar{F}}_{3} \right) \in X\) the equation \((A+2I)U={\bar{F}}\in X\) has a unique solution \(U=(v,w,\theta )\in D(A)\). Indeed, the equation \((A+2I)U={\bar{F}}\in X\) gives \(w=2v-{\bar{F}}_{1} \) and \(v(x)={\tilde{v}}(x)+\frac{a{\bar{F}}_{1} (1)}{2a+1} x\) for \(x\in [0,1]\), where \(({\tilde{v}},\theta )\in H^{2} (0,1)\times H^{2} (0,1)\) is the unique solution of the boundary-value problem

$$\begin{aligned}{} & {} {-c^{2} {\tilde{v}}''(x)+2(\mu +2){\tilde{v}}(x)+b\theta '(x)=\varphi _{1} (x)} \nonumber \\{} & {} \quad {-k\theta ''(x)+2\lambda {\tilde{v}}'(x)+2\theta (x)=\varphi _{2} (x)} \nonumber \\{} & {} \quad { \ for \ x\in (0,1) \ a.e. \ with ~{\tilde{v}}(0)=\theta (0)=0 } \nonumber \\{} & {} \quad {\ and \ {\tilde{v}}'(1)+2a{\tilde{v}}(1)=\theta (1)=0 } \end{aligned}$$
(A16)

with \(\varphi _{1} (x)={\bar{F}}_{2} (x)+(\mu +2){\bar{F}}_{1} (x)-2(\mu +2)\frac{a{\bar{F}}_{1} (1)}{2a+1} x\) and \(\varphi _{2} (x)={\bar{F}}_{3} (x)+\lambda {\bar{F}}'_{1} (x)-\frac{2\lambda a{\bar{F}}_{1} (1)}{2a+1} \). Let Y be the Hilbert space

$$\begin{aligned} Y=\left\{ \, ({\tilde{v}},\theta )\in H^{1} (0,1)\times H^{1} (0,1)\,:\, {\tilde{v}}(0)=\theta (0)=\theta (1)=0\, \right\} \end{aligned}$$
(A17)

with the usual scalar product

$$\begin{aligned} \left( ({\tilde{v}},\theta ),(r,p)\right)= & {} \int _{0}^{1}{\tilde{v}}'(x)r'(x)dx +\int _{0}^{1}{\tilde{v}}(x)r(x)dx \nonumber \\{} & {} {+\int _{0}^{1}\theta '(x)p'(x)dx +\int _{0}^{1}\theta (x)p(x)dx, } \nonumber \\{} & {} \qquad \ for \ ({\tilde{v}},\theta )\in Y, \ (r,p)\in Y \end{aligned}$$
(A18)

The fact that the boundary-value problem (A16) has a unique solution \(({\tilde{v}},\theta )\in H^{2} (0,1)\times H^{2} (0,1)\) for every \((\varphi _{1},\varphi _{2} )\in L^{2} (0,1)\times L^{2} (0,1)\) is a direct consequence of the methodology described on pages 221-229 in [5] and the Lax-Milgram Theorem (Corollary 5.8 on page 140 in [5]) applied to the continuous, coercive, bilinear form \(\alpha \) on Y defined by the formula

$$\begin{aligned} \alpha \left( ({\tilde{v}},\theta ),(r,p)\right)= & {} 2ac^{2} {\tilde{v}}(1)r(1)+c^{2} \int _{0}^{1}{\tilde{v}}'(x)r'(x)dx \\{} & {} {+2(\mu +2)\int _{0}^{1}{\tilde{v}}(x)r(x)dx +b\int _{0}^{1}\theta '(x)r(x)dx } \\{} & {} {+\frac{bk}{2\lambda } \int _{0}^{1}\theta '(x)p'(x)dx +b\int _{0}^{1}{\tilde{v}}'(x)p(x)dx +\frac{b}{\lambda } \int _{0}^{1}\theta (x)p(x)dx } \\{} & {} \qquad { \ for \ all \ ({\tilde{v}},\theta )\in Y, \ (r,p)\in Y } \end{aligned}$$

Thus \(A+I\) is a maximal monotone operator and consequently (using the Hille-Yosida Theorem) A is the generator of a continuous semigroup of contractions \(S(t):X\rightarrow X\). It follows from Theorem 7.10 on page 198 in [5] that for every \(F\in C^{1} \left( {\mathbb {R}} _{+};X\right) \) and for every \(U_{0} \in D(A)\) there exists a unique solution \(U\in C^{1} \left( {\mathbb {R}} _{+};X\right) \cap C^{0} \left( {\mathbb {R}} _{+};D(A)\right) \) of the initial-value problem (A2) with initial condition \(U(0)=U_{0} \).

Going back to the original state variables (and using (A1)), we conclude that for every \(d\in C^{2} \left( {\mathbb {R}} _{+} \right) \), \(f\in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \) and for every \(u_{0} \in H^{2} (0,1)\), \(w_{0} \in H^{1} (0,1)\), \(\theta _{0} \in H^{2} (0,1)\) with \(u_{0} (0)=w_{0} (0)=\theta (0)=\theta (1)=0\), \(u'_{0} (1)=-aw_{0} (1) +d(0)\) there exists a unique solution \(u\in C^{1} \left( {\mathbb {R}} _{+};H^{1} (0,1)\right) \cap C^{0} \left( {\mathbb {R}} _{+};H^{2} (0,1)\right) \), \(\theta \in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \cap C^{0} \left( {\mathbb {R}} _{+};H^{2} (0,1)\right) \) with \(u_{t} \in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \cap C^{0} \left( {\mathbb {R}} _{+};H^{1} (0,1)\right) \) of the initial-boundary value problem (2), (6), (7), (8), (9) with \(u[0]=u_{0} \), \(u_{t} [0]=w_{0} \), \(\theta [0]=\theta _{0} \) that additionally satisfies \(u_{t} (t,0)=0\) for all \(t\ge 0\).

(3) Well-posedness analysis of Model (D).

Model (C) consists of equations (2), (8), (9), (10), (11). Applying the transformation

$$\begin{aligned}{} & {} {w(t,x)=u_{t} (t,x)-\sigma u_{xx} (t,x)+gu(t,x)}\nonumber \\{} & {} {v(t,x)=u_{t} (t,x)} \nonumber \\{} & {} {y(t)=u_{t} (t,1)} \nonumber \\{} & {} \quad { for \ t\ge 0, \ x\in [0,1] } \end{aligned}$$
(A19)

where

$$\begin{aligned} g:=\mu -\sigma ^{-1} c^{2}, \ \kappa :=\sigma ^{-1} c^{2} \end{aligned}$$
(A20)

we obtain the first-order in time model (A2), where

$$\begin{aligned}{} & {} U=(u,v,\theta ,w,y) \end{aligned}$$
(A21)
$$\begin{aligned}{} & {} F(t,x)=\left( 0,f(t,x),0,f(t,x),0\right) , \ for \ t\ge 0, \ x\in [0,1] \end{aligned}$$
(A22)

and \(A:D(A)\rightarrow X\) is the linear unbounded operator defined by

$$\begin{aligned}{} & {} AU={\left( -\sigma u''-h,-\sigma v''+gv+\kappa h+b\theta ',-k\theta ''+\lambda v',\kappa h+b\theta ',a^{-1} v'(1)\right) } \nonumber \\{} & {} \quad { \ for \ U=(u,v,\theta ,w,y)\in D(A), \ where \ h=w-gu } \end{aligned}$$
(A23)

with X being the Hilbert space

$$\begin{aligned} X=\left( L^{2} (0,1)\right) ^{4} \times {{\mathbb {R}}} \end{aligned}$$
(A24)

with scalar product

$$\begin{aligned}{} & {} {\left( U,{\bar{U}}\right) =\int _{0}^{1}u(x){\bar{u}}(x)dx +\int _{0}^{1}v(x){\bar{v}}(x)dx } \nonumber \\{} & {} {\quad +\frac{b}{\lambda } \int _{0}^{1}\theta (x){\bar{\theta }}(x)dx +\int _{0}^{1}w(x){\bar{w}}(x)dx +a\sigma y{\bar{y}}, } \nonumber \\{} & {} {\qquad for \ U=(u,v,\theta ,w,y)\in X, \ {\bar{U}}=({\bar{u}},{\bar{v}},{\bar{\theta }},{\bar{w}},{\bar{y}})\in X } \end{aligned}$$
(A25)

and \(D(A)\subset X\) being the following linear subspace of X

$$\begin{aligned}{} & {} {D(A)= \left\{ \, (u,v,\theta ,w,y)\in Z\,:\, \begin{array}{c} {u(0)=\theta (0)=v(0)=0} \\ {v(1)-y=u'(1)+ay=\theta (1)=0} \\ { where \ Z=\left( H^{2} (0,1)\right) ^{3} \times L^{2} (0,1)\times {{\mathbb {R}}} } \end{array}\; \right\} }\nonumber \\ \end{aligned}$$
(A26)

Let \(I:X\rightarrow X\) be the identity operator. We next show that \(A+\beta I\) is a maximal monotone operator (see [5]) for a sufficiently large constant \(\beta >0\). Indeed, we have for all \(U=(u,v,\theta ,w,y)\in D(A)\):

$$\begin{aligned}{} & {} {\left( (A+\beta I)U,U\right) =\int _{0}^{1}u(x)\left( -\sigma u''(x)+gu(x)-w(x)\right) dx } \\{} & {} \qquad {+\int _{0}^{1}v(x)\left( -\sigma v''(x)+gv(x)+\kappa (w(x)-gu(x))+b\theta '(x)\right) dx } \\{} & {} \qquad {+\frac{b}{\lambda } \int _{0}^{1}\theta (x)\left( -k\theta ''(x)+\lambda v'(x)\right) dx +\beta \left\| u\right\| _{2}^{2} +\beta \left\| v\right\| _{2}^{2} } \\{} & {} \qquad { +\beta \frac{b}{\lambda } \left\| \theta \right\| _{2}^{2} +\beta \left\| w\right\| _{2}^{2} +a\sigma \beta y^{2} } \\{} & {} \qquad {+\int _{0}^{1}w(x)\left( \kappa (w(x)-gu(x))+b\theta '(x)\right) dx +\sigma yv'(1)} \\{} & {} \quad {=a\sigma yu(1)+\sigma \left\| u'\right\| _{2}^{2} -\left( 1+\kappa g\right) \int _{0}^{1}u(x)w(x)dx +b\int _{0}^{1}w(x)\theta '(x)dx } \\{} & {} \qquad {+\sigma \left\| v'\right\| _{2}^{2} +\sigma ^{-1} c^{2} \int _{0}^{1}v(x)w(x)dx -\kappa g\int _{0}^{1}v(x)u(x)dx +\frac{bk}{\lambda } \left\| \theta '\right\| _{2}^{2} } \\{} & {} \qquad {+\left( \beta +g\right) \left\| u\right\| _{2}^{2} +\left( \beta +g\right) \left\| v\right\| _{2}^{2} +\beta \frac{b}{\lambda } \left\| \theta \right\| _{2}^{2} +\left( \beta +\sigma ^{-1} c^{2} \right) \left\| w\right\| _{2}^{2} +a\sigma \beta y^{2} } \end{aligned}$$

In the above we have used (A23), (A25), (A26) and integration by parts. Using the fact that \(yu(1) \ge -\beta y^{2} -\frac{u^{2} (1)}{4\beta } \) and the Cauchy-Schwarz inequality we obtain:

$$\begin{aligned} \left( (A+\beta I)U,U\right)\ge & {} -\frac{a\sigma }{4\beta } u^{2} (1)+\sigma \left\| u'\right\| _{2}^{2} -\left| 1+\kappa g\right| \left\| u\right\| _{2} \left\| w\right\| _{2} -b\left\| w\right\| _{2} \left\| \theta '\right\| _{2}\\{} & {} {+\sigma \left\| v'\right\| _{2}^{2} -\kappa \left\| v\right\| _{2} \left\| w\right\| _{2} -\kappa \left| g\right| \left\| u\right\| _{2} \left\| v\right\| _{2} +\frac{bk}{\lambda } \left\| \theta '\right\| _{2}^{2} } \\{} & {} {+\left( \beta +g\right) \left\| u\right\| _{2}^{2} +\left( \beta +g\right) \left\| v\right\| _{2}^{2} +\frac{\beta b}{\lambda } \left\| \theta \right\| _{2}^{2} +\left( \beta +\kappa \right) \left\| w\right\| _{2}^{2} } \end{aligned}$$

Since \(u\in H^{2} (0,1)\) with \(u(0)=0\) we get \(u^{2} (1) =2\int _{0}^{1}u(x)u'(x)dx \le 2\left\| u\right\| _{2} \left\| u'\right\| _{2} \). Using this inequality and the inequalities \(\left\| u\right\| _{2} \left\| w\right\| _{2} \le \left\| u\right\| _{2}^{2} +\left\| w\right\| _{2}^{2} \), \(\left\| v\right\| _{2} \left\| w\right\| _{2} \le \left\| v\right\| _{2}^{2} +\left\| w\right\| _{2}^{2} \), \(\left\| u\right\| _{2} \left\| v\right\| _{2} \le \left\| u\right\| _{2}^{2} +\left\| v\right\| _{2}^{2} \), \(\left\| \theta '\right\| _{2} \left\| w\right\| _{2} \le \frac{k}{\lambda } \left\| \theta '\right\| _{2}^{2} +\frac{\lambda }{4k} \left\| w\right\| _{2}^{2} \), we get:

$$\begin{aligned} \left( (A+\beta I)U,U\right)\ge & {} -\frac{a\sigma }{2\beta } \left\| u\right\| _{2} \left\| u'\right\| _{2} +\sigma \left\| u'\right\| _{2}^{2} +\sigma \left\| v'\right\| _{2}^{2} \\{} & {} {+\left( \beta -\left| g\right| -1-2\kappa \left| g\right| \right) \left\| u\right\| _{2}^{2} +\left( \beta -\left| g\right| -\kappa -\kappa \left| g\right| \right) \left\| v\right\| _{2}^{2} } \\{} & {} {+\left( \beta -\frac{b\lambda }{4k} -1-\kappa \left| g\right| \right) \left\| w\right\| _{2}^{2} } \end{aligned}$$

Finally, using the inequality \(\frac{a\sigma }{2\beta } \left\| u\right\| _{2} \left\| u'\right\| _{2} \le \sigma \left\| u'\right\| _{2}^{2} +\frac{a^{2} \sigma }{16\beta ^{2} } \left\| u\right\| _{2}^{2} \) we get:

$$\begin{aligned}{} & {} {\left( (A+\beta I)U,U\right) \ge \left( \beta -\left| g\right| -1-2\kappa \left| g\right| -\frac{a^{2} \sigma }{16\beta ^{2} } \right) \left\| u\right\| _{2}^{2} } \\{} & {} \quad {+\left( \beta -\left| g\right| -\kappa -\kappa \left| g\right| \right) \left\| v\right\| _{2}^{2} +\left( \beta -\frac{b\lambda }{4k} -1-\kappa \left| g\right| \right) \left\| w\right\| _{2}^{2} } \end{aligned}$$

The above inequality shows that for \(\beta \ge 1+\kappa +\left( 2\kappa +1\right) \left| g\right| +\frac{a^{2} \sigma }{16\beta ^{2} } +\frac{b\lambda }{4k} \) we have \(\left( (A+\beta I)U,U\right) \ge 0\) for all \(U=(u,v,\theta ,w,y)\in D(A)\).

We next show that for sufficiently large \(\beta >0\), the range of the linear operator \((A+(\beta +1)I)\) is X. We show that for every \({\bar{F}}=\left( {\bar{F}}_{1},{\bar{F}}_{2},{\bar{F}}_{3},{\bar{F}}_{4},{\bar{F}}_{5} \right) \in X\) the equation \((A+(\beta +1)I)U={\bar{F}}\in X\) has a unique solution \(U=(u,v,\theta ,w,y)\in D(A)\) provided that \(\beta >0\) is sufficiently large. The equation \((A+(\beta +1)I)U={\bar{F}}\in X\) gives \(y=\frac{a{\bar{F}}_{5} }{1+a(\beta +1)} -\frac{{\tilde{v}}'(1)}{a(\beta +1)} \), \(u(x)={\tilde{u}}(x)-a\left( {\tilde{v}}(1)+\frac{a{\bar{F}}_{5} }{1+a(\beta +1)} \right) x\), \(w(x)=\frac{{\bar{F}}_{4} (x)-b\theta '(x)}{\beta +1+\kappa } +\frac{\kappa g}{\beta +1+\kappa } \left( {\tilde{u}}(x)-a\left( {\tilde{v}}(1)+\frac{a{\bar{F}}_{5} }{1+a(\beta +1)} \right) x\right) \) and \(v(x)={\tilde{v}}(x)+\frac{a{\bar{F}}_{5} x}{1+a(\beta +1)} \) for \(x\in [0,1]\), where \(({\tilde{u}},{\tilde{v}},\theta )\in \left( H^{2} (0,1)\right) ^{3} \) is a solution of the boundary-value problem

$$\begin{aligned}{} & {} {-\sigma {\tilde{u}}''(x)+(\beta +1)\left( 1+\frac{g}{\beta +1+\kappa } \right) {\tilde{u}}(x) } \nonumber \\{} & {} \quad {-a(\beta +1)\left( 1+\frac{g}{\beta +1+\kappa } \right) {\tilde{v}}(1)x+\frac{b}{\beta +1+\kappa } \theta '(x)=\varphi _{1} (x)} \nonumber \\{} & {} \quad {-\sigma {\tilde{v}}''(x)+(\beta +1+g){\tilde{v}}(x)+\frac{b\left( \beta +\kappa \right) }{\beta +1+\kappa } \theta '(x) } \nonumber \\{} & {} \quad {-\frac{\kappa g\left( \beta +\kappa \right) }{\beta +1+\kappa } {\tilde{u}}(x)+\frac{a\kappa g\left( \beta +\kappa \right) }{\beta +1+\kappa } {\tilde{v}}(1)x=\varphi _{2} (x)} \nonumber \\{} & {} \quad {-k\theta ''(x)+\lambda {\tilde{v}}'(x)+(\beta +1)\theta (x)=\varphi _{3} (x)} \nonumber \\{} & {} \qquad {\ for \ x\in (0,1) \ a.e. \ with \ {\tilde{u}}(0)={\tilde{v}}(0)=\theta (0)=0} \nonumber \\{} & {} \qquad { and \ {\tilde{u}}'(1)={\tilde{v}}'(1)+a(\beta +1){\tilde{v}}(1)=\theta (1)=0 } \end{aligned}$$
(A27)

with \(\varphi _{3} (x)={\bar{F}}_{3} (x)-\frac{a\lambda {\bar{F}}_{5} }{1+a(\beta +1)} \), \(\varphi _{1} (x)={\bar{F}}_{1} (x)+\frac{1}{\beta +1+\kappa } {\bar{F}}_{4} (x)+\frac{a^{2} (\beta +1){\bar{F}}_{5} }{1+a(\beta +1)} \left( 1+\frac{g}{\beta +1+\kappa } \right) x\) and \(\varphi _{2} (x)={\bar{F}}_{2} (x)-\frac{1}{\beta +1+\kappa } {\bar{F}}_{4} (x)-\left( \beta +1+g+\frac{a\kappa g\left( \beta +\kappa \right) }{\beta +1+\kappa } \right) \frac{a{\bar{F}}_{5} x}{1+a(\beta +1)} \). Let Y be the Hilbert space

$$\begin{aligned} Y=\left\{ \, ({\tilde{u}},{\tilde{v}},\theta )\in \left( H^{1} (0,1)\right) ^{3} \,:\, {\tilde{u}}(0)={\tilde{v}}(0)=\theta (0)=\theta (1)=0\, \right\} \end{aligned}$$
(A28)

with the usual scalar product

$$\begin{aligned}{} & {} \left( ({\tilde{u}},{\tilde{v}},\theta ),(q,r,p)\right) =\int _{0}^{1}{\tilde{u}}'(x)q'(x)dx +\int _{0}^{1}{\tilde{u}}(x)q(x)dx \nonumber \\{} & {} \quad {+\int _{0}^{1}{\tilde{v}}'(x)r'(x)dx +\int _{0}^{1}{\tilde{v}}(x)r(x)dx +\int _{0}^{1}\theta '(x)p'(x)dx +\int _{0}^{1}\theta (x)p(x)dx } \nonumber \\{} & {} \qquad {for \ ({\tilde{u}},{\tilde{v}},\theta )\in Y, \ (q,r,p)\in Y } \end{aligned}$$
(A29)

We show that the boundary-value problem (A27) has a unique solution \(({\tilde{u}},{\tilde{v}},\theta )\in \left( H^{2} (0,1)\right) ^{3} \) for every \((\varphi _{1},\varphi _{2},\varphi _{3} )\in \left( L^{2} (0,1)\right) ^{3} \) provided that \(\beta >0\) is sufficiently large. Let \(R>0\) be an arbitrary constant and consider the continuous, bilinear form \(\alpha \) on Y defined by the formula

$$\begin{aligned} \alpha \left( ({\tilde{u}},{\tilde{v}},\theta ),(q,r,p)\right)= & {} \frac{\sigma \sqrt{15\sigma } }{a^{2} \sqrt{\beta +1} } \int _{0}^{1}{\tilde{u}}'(x)q'(x)dx \nonumber \\{} & {} {+\frac{\sqrt{15\sigma \left( \beta +1\right) } }{a^{2} } \left( 1+\frac{g}{\beta +1+\kappa } \right) \int _{0}^{1}{\tilde{u}}(x)q(x)dx } \nonumber \\{} & {} {-\frac{\sqrt{15\sigma \left( \beta +1\right) } }{a} \left( 1+\frac{g}{\beta +1+\kappa } \right) {\tilde{v}}(1)\int _{0}^{1}xq(x)dx } \nonumber \\{} & {} { +\frac{b\sqrt{15\sigma } }{a^{2} \left( \beta +1+\kappa \right) \sqrt{\beta +1} } \int _{0}^{1}\theta '(x)q(x)dx }\nonumber \\{} & {} {+\sigma a(\beta +1){\tilde{v}}(1)r(1)+\sigma \int _{0}^{1}{\tilde{v}}'(x)r'(x)dx }\nonumber \\{} & {} {+(\beta +1+g)\int _{0}^{1}{\tilde{v}}(x)r(x)dx +\frac{b\left( \beta +\kappa \right) }{\beta +1+\kappa } \int _{0}^{1}\theta '(x)r(x)dx }\nonumber \\{} & {} {-\frac{\kappa g\left( \beta +\kappa \right) }{\beta +1+\kappa } \int _{0}^{1}{\tilde{u}}(x)r(x)dx +\frac{a\kappa g\left( \beta +\kappa \right) }{\beta +1+\kappa } {\tilde{v}}(1)\int _{0}^{1}xr(x)dx } \nonumber \\{} & {} {+\frac{bk\left( \beta +\kappa \right) }{\lambda (\beta +1+\kappa )} \int _{0}^{1}\theta '(x)p'(x)dx +\frac{b\left( \beta +\kappa \right) }{(\beta +1+\kappa )} \int _{0}^{1}{\tilde{v}}'(x)p(x)dx } \nonumber \\{} & {} { +\frac{b(\beta +1)\left( \beta +\kappa \right) }{\lambda (\beta +1+\kappa )} \int _{0}^{1}\theta (x)p(x)dx }\nonumber \\{} & {} {for \ all \ ({\tilde{u}},{\tilde{v}},\theta )\in Y, \ (q,r,p)\in Y } \end{aligned}$$
(A30)

We notice that continuity of the bilinear form \(\alpha \) on Y defined by (A30) is established by means of the inequality \(\left| {\tilde{v}}(1)\right| \le \sqrt{2\left\| {\tilde{v}}\right\| _{2} \left\| {\tilde{v}}'\right\| _{2} } \le \left\| {\tilde{v}}\right\| _{2} +\left\| {\tilde{v}}'\right\| _{2} \), which holds since \({\tilde{v}}\in H^{1} (0,1)\) with \({\tilde{v}}(0)=0\) (recall (A28)). Completing the squares and using the Cauchy-Schwarz inequality and the fact that \(\left| {\tilde{v}}(1)\right| \le \sqrt{2\left\| {\tilde{v}}\right\| _{2} \left\| {\tilde{v}}'\right\| _{2} } \), we are in a position to establish the following inequality for all \(({\tilde{u}},{\tilde{v}},\theta )\in Y\):

$$\begin{aligned}{} & {} {\alpha \left( ({\tilde{u}},{\tilde{v}},\theta ),({\tilde{u}},{\tilde{v}},\theta )\right) \ge \frac{\sigma \sqrt{15\sigma } }{a^{2} \sqrt{\beta +1} } \left\| {\tilde{u}}'\right\| _{2}^{2} +\frac{\sigma a(\beta +1)}{4} {\tilde{v}}^{2} (1)+\frac{\sigma }{2} \left\| {\tilde{v}}'\right\| _{2}^{2} } \nonumber \\{} & {} \quad {+\frac{\left\| {\tilde{u}}\right\| _{2}^{2} }{2a^{4} k\sqrt{\beta +1} } (\beta +1)a^{2} k\sqrt{15\sigma } } \nonumber \\{} & {} \quad {-\frac{\left\| {\tilde{u}}\right\| _{2}^{2} }{2a^{4} k\sqrt{\beta +1} } \left( \left( 2\sqrt{15\sigma } +a^{2} \kappa ^{2} \right) a^{2} k\left| g\right| +k\kappa a^{4} \left| g\right| \sqrt{\beta +1} +15\sigma b\lambda +10ak\left| g\right| ^{2} \right) } \nonumber \\{} & {} \quad {+\frac{1}{6\sigma } \left( \sigma \left( \beta +1\right) -\left( 6\sigma +3\sigma \kappa +a\kappa ^{2} \left| g\right| \right) \left| g\right| \right) \left\| {\tilde{v}}\right\| _{2}^{2} } \nonumber \\{} & {} \quad {+\frac{bk\left( \beta +\kappa \right) }{2\lambda (\beta +1+\kappa )} \left\| \theta '\right\| _{2}^{2} +\frac{b(\beta +1)\left( \beta +\kappa \right) }{\lambda (\beta +1+\kappa )} \left\| \theta \right\| _{2}^{2} } \end{aligned}$$
(A31)

It follows from (A31) that the bilinear form \(\alpha \) on Y defined by (A30) is coercive when

$$\begin{aligned}{} & {} {\beta +1-\frac{\kappa a^{2} \left| g\right| }{\sqrt{15\sigma } } \sqrt{\beta +1}>\frac{\left( 2\sqrt{15\sigma } +a^{2} \kappa ^{2} \right) a^{2} k\left| g\right| +15\sigma b\lambda +10ak\left| g\right| ^{2} }{a^{2} k\sqrt{15\sigma } } } \\{} & {} \quad {\beta +1>\left( 6+3\kappa +\sigma ^{-1} a\kappa ^{2} \left| g\right| \right) \left| g\right| } \end{aligned}$$

The above inequalities are satisfied when \(\beta >0\) is sufficiently large (notice that left hand sides of the above inequalities tend to \(+\infty \) as \(\beta \rightarrow +\infty \)). Following the methodology described on pages 221-229 in [5] and using the Lax-Milgram Theorem (Corollary 5.8 on page 140 in [5]) applied to the continuous, coercive, bilinear form \(\alpha \) on Y defined by (A30) we conclude that the boundary-value problem (A27) has a unique solution \(({\tilde{u}},{\tilde{v}},\theta )\in \left( H^{2} (0,1)\right) ^{3} \) for every \((\varphi _{1},\varphi _{2},\varphi _{3} )\in \left( L^{2} (0,1)\right) ^{3} \) provided that \(\beta >0\) is sufficiently large.

Thus \(A+\beta I\) is a maximal monotone operator when \(\beta >0\) is sufficiently large and consequently (using the Hille-Yosida Theorem) A is the generator of a continuous semigroup of contractions \(S(t):X\rightarrow X\). It follows from Theorem 7.10 on page 198 in [5] that for every \(F\in C^{1} \left( {\mathbb {R}} _{+};X\right) \) and for every \(U_{0} \in D(A)\) there exists a unique solution \(U\in C^{1} \left( {\mathbb {R}} _{+};X\right) \cap C^{0} \left( {\mathbb {R}} _{+};D(A)\right) \) of the initial-value problem (A2) with initial condition \(U(0)=U_{0} \).

Going back to the original state variables (and using (A19)), we conclude that for every \(f\in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \) and for every \(u_{0} \in H^{2} (0,1)\), \(v_{0} \in H^{2} (0,1)\), \(\theta _{0} \in H^{2} (0,1)\) with \(u_{0} (0)=v_{0} (0)=\theta (0)=\theta (1)=0\), \(u'_{0} (1)=-av_{0} (1)\) there exists a unique solution \(u\in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \cap C^{0} \left( {\mathbb {R}} _{+};H^{2} (0,1)\right) \), \(\theta \in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \cap C^{0} \left( {\mathbb {R}} _{+};H^{2} (0,1)\right) \) with \(u_{t} \in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \cap C^{0} \left( {\mathbb {R}} _{+};H^{2} (0,1)\right) \), \(u_{xx} \in C^{1} \left( {\mathbb {R}} _{+};L^{2} (0,1)\right) \), \(u_{t} (\, \cdot \,,1)\in C^{1} \left( {\mathbb {R}} _{+} \right) \) of the initial-boundary value problem (2), (8), (9), (10), (11) with \(u[0]=u_{0} \), \(u_{t} [0]=v_{0} \), \(\theta [0]=\theta _{0} \) that additionally satisfies \(u_{t} (t,0)=0\), \(u_{xt} (t,1)=-a\frac{d}{d\, t} \left( u_{t} (t,1)\right) \) for all \(t\ge 0\).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Karafyllis, I., Krstic, M. ISS-based robustness to various neglected damping mechanisms for the 1-D wave PDE. Math. Control Signals Syst. 35, 741–779 (2023). https://doi.org/10.1007/s00498-023-00353-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00498-023-00353-6

Keywords

Navigation