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

Exponential Stability and Design of Sensor Feedback Amplifiers for Fast Stabilization of Magnetizable Piezoelectric Beam Equations

Ahmet Özkan Özer1    \IEEEmembershipMember, IEEE    Ahmet Kaan Aydın2 \IEEEmembershipMember, IEEE    and Rafi Emran3 1Department of Mathematics, Western Kentucky University, Bowling Green, KY 42101, USA. [email protected] 2Department of Mathematics, University of Maryland, Baltimore County, Baltimore, MD 21250, USA. [email protected]3Department of Mathematical Sciences, New Jersey Institute of Technology, Newark, NJ 07102, USA. [email protected]
Abstract

The dynamic partial differential equation (PDE) model governing longitudinal oscillations in magnetizable piezoelectric beams exhibits exponentially stable solutions when subjected to two boundary state feedback controllers. An analytically established exponential decay rate by the Lyapunov approach ensures stabilization of the system to equilibrium, though the actual decay rate could potentially be improved. The decay rate of the closed-loop system is highly sensitive to the choice of material parameters and the design of the state feedback amplifiers. This paper focuses on investigating the design of state feedback amplifiers to achieve a maximal exponential decay rate, which is essential for effectively suppressing oscillations in these beams. Through this design process, we explicitly determine the safe intervals of feedback amplifiers that ensure the theoretically found maximal decay rate, with the potential for even better rates. Our numerical results reaffirm the robustness of the decay rate within the chosen range of feedback amplifiers, while deviations from this range significantly impact the decay rate. To underscore the validity of our results, we present various numerical experiments.

{IEEEkeywords}

Intelligent Materials, Distributed Parameter Systems, Feedback Stabilization, Maximal Decay Rate, Lyapunov Approach

1 Introduction

\IEEEPARstart

Piezoelectric ceramics, including lead-free materials like barium titanate, sodium potassium niobate, and sodium bismuth titanate, are renowned multifunctional smart materials that generate electric displacement in response to mechanical stress [26]. Their small size and high power density make them ideal for various industrial applications, including implantable biomedical devices [5, 25], wearable interfaces with PVDF sensors [7], biocompatible sensors [32], and ultrasound imagers and cleaners [24]. Their fast response, large mechanical force, and fine resolution contribute to their effectiveness [10].

Consider a piezoelectric beam, clamped on one side and free on the other, with sensors for tip velocity and current. he beam is of length LL and thickness hh. Assume that the transverse oscillations of the beam are negligible, so the longitudinal vibrations, in the form of expansion and compression of the center line of the beam, are the only oscillations of note together with the electromagnetic effects. Existing mathematical models often oversimplify intrinsic mechanical or electromagnetic interactions, impacting boundary feedback stabilizability [1, 17, 20, 6, 27, 29].

While electrostatic and quasi-static approaches based on Maxwell’s equations are typically sufficient for non-magnetizable piezoelectric beams, assessing radiated electromagnetic power requires consideration of electromagnetic waves generated by mechanical fields [28]. Therefore, fully dynamic models of piezoelectric beams are essential. Denoting v(x,t)v(x,t) as the longitudinal oscillations and p(x,t)p(x,t) as the total charge accumulated at the electrodes, the equations of motion form following system of partial differential equations [17]

[ρ00μ][vttptt][αγβγββ][vxxpxx]=[00],(v,p)(0,t)=0,[αγβγββ][vxpx](L,t)=[u1(t)u2(t)],t+[v,p,vt,pt](x,0)=[v0,p0,v1,p1](x),x[0,L]\displaystyle\begin{array}[]{ll}\begin{array}[]{ll}\begin{bmatrix}\rho&0\\ 0&\mu\end{bmatrix}\begin{bmatrix}v_{tt}\\ p_{tt}\end{bmatrix}-\begin{bmatrix}\alpha&-\gamma\beta\\ -\gamma\beta&\beta\end{bmatrix}\begin{bmatrix}v_{xx}\\ p_{xx}\end{bmatrix}=\begin{bmatrix}0\\ 0\end{bmatrix},&\\ \end{array}\\ \begin{array}[]{ll}(v,p)(0,t)=0,&\\ \begin{bmatrix}\alpha&-\gamma\beta\\ -\gamma\beta&\beta\end{bmatrix}\begin{bmatrix}v_{x}\\ p_{x}\end{bmatrix}(L,t)=\begin{bmatrix}u_{1}(t)\\ u_{2}(t)\end{bmatrix},\quad t\in\mathbb{R}^{+}\end{array}\\ \left[v,p,v_{t},p_{t}\right]\left(x,0\right)=\left[v_{0},p_{0},v_{1},p_{1}\right]\left(x\right),~{}x\in\left[0,L\right]\end{array} (7)

where ρ\rho, α\alpha, β\beta, γ\gamma, and μ\mu are the mass density per unit volume, the elastic stiffness, the impermeability, the piezoelectric constant, and the magnetic permeability, respectively, and u1(t)u_{1}(t) and u2(t)u_{2}(t) are strain and voltage actuators, respectively.

Define α:=α1+γ2β>0\alpha:=\alpha_{1}+\gamma^{2}\beta>0 with α1>0\alpha_{1}>0, and

ζ=12αμα1β+ρα1(αμα1β+ρα1)24ρμβα1.\displaystyle\begin{array}[]{l}\zeta^{\mp}=\frac{1}{\sqrt{2}}\sqrt{\frac{\alpha\mu}{\alpha_{1}\beta}+\frac{\rho}{\alpha_{1}}\mp\sqrt{\left(\frac{\alpha\mu}{\alpha_{1}\beta}+\frac{\rho}{\alpha_{1}}\right)^{2}-\frac{4\rho\mu}{\beta\alpha_{1}}}}.\end{array} (9)

Note that ρα\sqrt{\frac{\rho}{\alpha}} and βμ\sqrt{\frac{\beta}{\mu}} are non-identical and represent significantly different wave propagation speeds in (7). The natural energy of the solutions is defined as

E(t)=120L[ρ|vt|2+μ|pt|2+α1|vx|2+β|γvxpx|2]dx.\displaystyle\begin{array}[]{l}E(t)=\frac{1}{2}\int^{L}_{0}\left[\rho\left|v_{t}\right|^{2}+\mu\left|p_{t}\right|^{2}+\alpha_{1}|v_{x}|^{2}\right.\\ \hskip 128.0374pt\left.+\beta\left|\gamma v_{x}-p_{x}\right|^{2}\right]dx.\end{array} (12)

The exact observability result for the model (7) with u1(t),u2(t)=0u_{1}(t),~{}u_{2}(t)=0 and with only one sensor measurement is not possible [17]. However, by employing two sensor measurements, vt(L,t)v_{t}(L,t) (tip velocity) and pt(L,t)p_{t}(L,t) (total current accumulated at electrodes), a suboptimal observation time is achieved [23]. This result is later refined to include the optimal observation time [21].

Theorem 1

[21, Theorem 2.2] Define the Hilbert spaces (H1(0,L))={zH1(0,L):z(0)=0}(H^{1}_{*}(0,L))=\{z\in H^{1}(0,L):z(0)=0\} and =(H1(0,L))2×(L2(0,L))2.\mathcal{H}=(H^{1}_{*}(0,L))^{2}\times(L^{2}(0,L))^{2}. For all initial data [v0,p0,v1,p1](x)\left[v_{0},p_{0},v_{1},p_{1}\right](x)\in\mathcal{H}, and for any T>2Lmax{ζ,ζ+}T>\frac{2L}{\max\{\zeta^{-},\zeta^{+}\}}, there exists a constant C(T)>0C(T)>0 such that the weak solutions (v,p,vt,pt)(v,p,v_{t},p_{t})\in\mathcal{H} of the control-free system (7), i.e., u1(t),u2(t)=0,u_{1}(t),~{}u_{2}(t)=0, satisfy

0T(ρ|vt(L,t)|2+μ|pt(L,t)|2)𝑑tC(T)E(0).\displaystyle\begin{array}[]{l}\int^{T}_{0}\left(\rho\left|v_{t}(L,t)\right|^{2}+\mu\left|p_{t}(L,t)\right|^{2}\right)dt\geq C(T)E(0).\end{array} (14)

When the observed sensor signals vt(L,t)v_{t}(L,t) and pt(L,t)p_{t}(L,t) are amplified and fed back to (7), a closed-loop system is formed. The amplification range depends on the sensor limits. While using only one sensor feedback can make the system energy dissipative, it is insufficient for exponential stabilization in \mathcal{H}. This is because the closed-loop system with this control design imposes strict conditions on stabilization results [17]. Exponential stability is jeopardized for a large class of material parameters and is only achievable for a small subset [18].

Let ξ1,ξ2>0\xi_{1},\xi_{2}>0 be the feedback amplifiers for the two sensor measurements of the closed-loop system vt(L,t)v_{t}(L,t) and pt(L,t)p_{t}(L,t), respectively. The strain and voltage actuators u1(t)u_{1}(t) and u2(t)u_{2}(t) are chosen to be proportional to the sensor feedback amplifiers,

[u1(t)u2(t)]=[ξ100ξ2][v˙(L,t)p˙(L,t)].\begin{array}[]{ll}\begin{bmatrix}u_{1}(t)\\ u_{2}(t)\end{bmatrix}=-\begin{bmatrix}\xi_{1}&0\\ 0&\xi_{2}\end{bmatrix}\begin{bmatrix}\dot{v}(L,t)\\ \dot{p}(L,t)\end{bmatrix}.\end{array} (15)

The exponential stability of the closed-loop system (7), (15) has been studied in [23], where the proof relies on decomposing the system into conservative and dissipative components. However, this method does not explicitly describe the decay rate or allow optimization of the feedback amplifiers.

The primary objective of this paper is to establish the existence of an exponential decay rate for the system described by (7) and (15) through the meticulous construction of a Lyapunov function. Given the impracticality of traditional spectral analysis due to the strong coupling in the wave system, we adopt a multiplier approach combined with an optimization argument to define a safe range of intervals for each feedback amplifier ξ1\xi_{1} and ξ2\xi_{2}. This ensures that the decay rate provided by the Lyapunov approach is achievable for any type of initial conditions \mathcal{H}.

Our methodology provides a maximal decay rate for the system, serving as an upper bound to the optimal decay rate since the actual decay rate could potentially be even better. Existing literature on optimal actuator designs commonly offers different approaches for infinite-dimensional systems [8, 13, 15, 22], and finite-dimensional systems [4, 9]. Our approach is particularly valuable for model reductions by Finite Differences or Finite Elements [21] for (7). Notably, the Lyapunov-based exponential stability proof [2], uniformly as the discretization parameter h0h\to 0 with the recently proposed order-reduced Finite Differences [12], closely resembles the approach employed here.

2 Exponential Stability Result

For the solutions of the system (7),(15) to stabilize exponentially, the energy must be dissipative. The proof of the following dissipativity theorem is omitted.

Lemma 1

For all t>0,t>0, the energy E(t)E(t) in (12) is dissipative. In other words,

dEdt=ξ1|v˙(L,t)|2ξ2|p˙(L,t)|20.\frac{dE}{dt}=-\xi_{1}\left|\dot{v}(L,t)\right|^{2}-\xi_{2}\left|\dot{p}(L,t)\right|^{2}\leq 0.

Now, let’s define an energy-like functional F(t)F(t) in order to establish a perturbed energy functional Eδ(t)E_{\delta}(t) as follows

F(t):=0L(ρvtxvx+μptxpx)𝑑x,\displaystyle F(t):=\int^{L}_{0}\left(\rho v_{t}xv_{x}+\mu p_{t}xp_{x}\right)dx, (16)
Eδ(t):=E(t)+δF(t).\displaystyle E_{\delta}(t):=E(t)+\delta F(t). (17)

Here, δ>0\delta>0 will be determined as a function of sensor feedback amplifiers ξ1\xi_{1} and ξ2\xi_{2} later.

The following two lemmas for F(t)F(t) and Eδ(t)E_{\delta}(t) are needed to prove our main exponential stability result. Let

η:=max(ρα1+μγ2α1,μβ+μγ2α1),\eta:=\max\left(\sqrt{\frac{\rho}{\alpha_{1}}}+\sqrt{\frac{\mu\gamma^{2}}{\alpha_{1}}},\sqrt{\frac{\mu}{\beta}}+\sqrt{\frac{\mu\gamma^{2}}{\alpha_{1}}}\right), (18)
Lemma 2

Letting 0<δ<1ηL0<\delta<\frac{1}{\eta L}, for all t>0t>0, Eδ(t)E_{\delta}(t) in (17) is equivalent to E(t)E(t) in (12). In other words,

(1δηL)E(t)Eδ(t)(1+δηL)E(t).\left(1-\delta\eta L\right)E(t)\leq E_{\delta}(t)\leq\left(1+\delta\eta L\right)E(t). (19)
Proof 2.1.

By the Hölder’s, Minkowski’s, Triangle inequalities, as well as algebraic manipulations, F(t)F(t) satisfies

|F(t)|L[(0Lρ|vt|2dx)12(ρα10Lα1|vx|2dx)12+(0Lμ|pt|2dx)12{(0Lμ|γvxpx|2dx)12+(μ0Lγ2|vx|2dx)12}]L2[ρα10Lρ|vt|2dx+(ρα1+μγ2α1)0Lα1|vx|2dx+(μβ+μγ2α1)0Lμ|pt|2dx+μβ0Lβ|γvxpx|2dx]LηE(t).\begin{array}[]{ll}&|F(t)|\leq L\left[\left(\int_{0}^{L}\rho|v_{t}|^{2}{dx}\right)^{\frac{1}{2}}\left(\frac{\rho}{\alpha_{1}}\int_{0}^{L}\ \alpha_{1}|v_{x}|^{2}{dx}\right)^{\frac{1}{2}}\right.\\ &\quad\left.+\left(\int_{0}^{L}\mu|p_{t}|^{2}{dx}\right)^{\frac{1}{2}}\left\{\left(\int_{0}^{L}\mu|\gamma v_{x}-p_{x}|^{2}dx\right)^{\frac{1}{2}}\right.\right.\\ &\qquad\left.\left.+\left(\mu\int_{0}^{L}\gamma^{2}|v_{x}|^{2}dx\right)^{\frac{1}{2}}\right\}\right]\\ &\leq\frac{L}{2}\left[\sqrt{\frac{\rho}{\alpha_{1}}}\int_{0}^{L}\rho|v_{t}|^{2}{dx}+\left(\sqrt{\frac{\rho}{\alpha_{1}}}+\sqrt{\frac{\mu\gamma^{2}}{\alpha_{1}}}\right)\int_{0}^{L}\alpha_{1}|v_{x}|^{2}{dx}\right.\\ &+\left(\sqrt{\frac{\mu}{\beta}}+\sqrt{\frac{\mu\gamma^{2}}{\alpha_{1}}}\right)\int_{0}^{L}\mu|p_{t}|^{2}{dx}\left.+\sqrt{\frac{\mu}{\beta}}\int_{0}^{L}\beta|\gamma v_{x}-p_{x}|^{2}{dx}\right]\\ &\leq L\eta E(t).\end{array}

Since |F(t)|LηE(t),|F(t)|\leq L\eta E(t), this leads to |Eδ(t)||E(t)|+δ|F(t)|(1+Lηδ)E(t),\left|E_{\delta}(t)\right|\leq\left|E(t)\right|+\delta\left|F(t)\right|\leq\left(1+L\eta\delta\right)E(t), and analogously, |Eδ(t)|(1Lηδ)E(t),\left|E_{\delta}(t)\right|\geq\left(1-L\eta\delta\right)E(t), and therefore, (19) is immediate

Lemma 3

For any ϵ,t>0\epsilon,t>0, F(t)F(t) in (16) satisfies the following inequalities,

dFdtE(t)+L2[ρ+(1+ϵ)ξ12α1]|vt(L,t)|2+L2[μ+(1+1ϵ)ξ22γ2α1+ξ22β]|pt(L,t)|2.\displaystyle\begin{array}[]{ll}&\frac{dF}{dt}\leq-E(t)+\frac{L}{2}\left[\rho+\frac{(1+\epsilon)\xi_{1}^{2}}{\alpha_{1}}\right]\left|v_{t}(L,t)\right|^{2}\\ &\qquad\quad+\frac{L}{2}\left[\mu+{\left(1+\frac{1}{\epsilon}\right)}\frac{\xi_{2}^{2}\gamma^{2}}{\alpha_{1}}+\frac{\xi_{2}^{2}}{\beta}\right]\left|p_{t}(L,t)\right|^{2}.\end{array} (22)
Proof 2.2.

Recalling α1=αγ2β\alpha_{1}=\alpha-\gamma^{2}\beta, and (7),

dFdt=L2β(γvx(L,t)px(L,t))2+L2α1(vx(L,t))2+L2[ρ|vt(L,t)|2+μ|pt(L,t)|2]E(t).\displaystyle\begin{array}[]{l}\frac{dF}{dt}=\frac{L}{2}\beta\left(\gamma v_{x}(L,t)-p_{x}(L,t)\right)^{2}+\frac{L}{2}\alpha_{1}(v_{x}(L,t))^{2}\\ \qquad+\frac{L}{2}\left[\rho\left|v_{t}(L,t)\right|^{2}+\mu\left|p_{t}(L,t)\right|^{2}\right]-E(t).\end{array}

Next, the boundary conditions are used so that

dFdt=E(t)+L2(μ+ξ22β)|pt(L,t)|2+Lρ2|vt(L,t)|2+L2α1(ξ1|vt(L,t)|+ξ2γ|pt(L,t)|)2.\displaystyle\begin{array}[]{l}\frac{dF}{dt}=-E(t)+\frac{L}{2}\left(\mu+\frac{\xi_{2}^{2}}{\beta}\right)\left|p_{t}(L,t)\right|^{2}\\ \quad+\frac{L\rho}{2}\left|v_{t}(L,t)\right|^{2}+\frac{L}{2\alpha_{1}}(\xi_{1}\left|v_{t}(L,t)\right|+\xi_{2}\gamma{\left|p_{t}(L,t)\right|)^{2}}.\end{array}

Finally, by the generalized Young’s inequality with ϵ\epsilon (or Peter–Paul inequality), (22) is obtained for any ϵ>0\epsilon>0.

Now, the exponential stability result takes the following form

Theorem 2.

The energy E(t)E(t) of solutions decays exponentially, i.e. for any ϵ>0\epsilon>0,

E(t)ME(0)eσt,t>0σ(δ)=δ(1δLη),M(δ)=1+δLη1δLη,\begin{array}[]{ll}&E(t)\leq ME(0)e^{-\sigma t},\qquad\forall t>0\\ &\sigma(\delta)=\delta\left(1-\delta L\eta\right),\qquad M(\delta)=\frac{1+\delta L\eta}{1-\delta L\eta},\end{array} (25)
{δ(ξ1,ξ2,ϵ)<1Lmin(1η,f1(ξ1,ϵ),f2(ξ2,ϵ)),f1(ξ1,ϵ):=2ξ1α1ρα1+(1+ϵ)ξ12,f2(ξ2,ϵ):=2ξ2ϵα1βϵμα1β+(ϵα+γ2β)ξ22.\begin{cases}&{\delta(\xi_{1},\xi_{2},\epsilon)<\frac{1}{L}\min\left(\frac{1}{\eta},f_{1}(\xi_{1},\epsilon),f_{2}(\xi_{2},\epsilon)\right),}\\ &f_{1}(\xi_{1},\epsilon):=\frac{2\xi_{1}\alpha_{1}}{\rho\alpha_{1}+(1+\epsilon)\xi_{1}^{2}},\\ &f_{2}(\xi_{2},\epsilon):=\frac{2\xi_{2}\epsilon\alpha_{1}\beta}{\epsilon\mu\alpha_{1}\beta+(\epsilon\alpha+\gamma^{2}\beta)\xi_{2}^{2}}.\end{cases} (26)
Proof 2.3.

Considering dEδdt=dEdt+δdFdt\frac{dE_{\delta}}{dt}=\frac{dE}{dt}+\delta\frac{dF}{dt} with Lemma 3

dEδdt[δL2[ρ+(1+ϵ)ξ12α1]ξ1]0|vt(L,t)|2\displaystyle\frac{dE_{\delta}}{dt}\leq\underbrace{\left[\frac{\delta L}{2}\left[\rho+\frac{(1+\epsilon)\xi_{1}^{2}}{\alpha_{1}}\right]-\xi_{1}\right]}_{\leq 0}\left|v_{t}(L,t)\right|^{2}
+[δL2[μ+(ϵα+γ2β)ξ22ϵα1β]ξ2]0|pt(L,t)|2δE(t)\displaystyle+\underbrace{\left[\frac{\delta L}{2}\left[\mu+\frac{(\epsilon\alpha+\gamma^{2}\beta)\xi_{2}^{2}}{\epsilon\alpha_{1}\beta}\right]-\xi_{2}\right]}_{\leq 0}\left|p_{t}(L,t)\right|^{2}-\delta E(t)

where δ\delta is chosen to make the coefficients nonpositive, i.e. δ<1Lmin(1η,f1(ξ1,ϵ),f2(ξ2,ϵ)).\delta<\frac{1}{L}\min\left(\frac{1}{\eta},f_{1}(\xi_{1},\epsilon),f_{2}(\xi_{2},\epsilon)\right). Next, by the equivalence of E(t)E(t) and Eδ(t)E_{\delta}(t) from Lemma 2,

dEδdtδ(1δLη)Eδ(t).\displaystyle\frac{dE_{\delta}}{dt}\leq-\delta\left(1-\delta L\eta\right)E_{\delta}(t). (27)

By choosing σ=δ(1δLη)>0,\sigma=\delta\left(1-\delta L\eta\right)>0, together with (27) lead to

Eδ(t)Eδ(0)eσt.E_{\delta}(t)\leq E_{\delta}(0)e^{-\sigma t}. (28)

Hence, (28) with Lemma 2 lead to the desired result.

3 Optimization of Sensor Feedback Amplifiers

The decay rate σ-\sigma in Theorem 2 provides an upper bound for the exponential decay rate of the system. In this section, we provide an optimization process for feedback amplifiers ξ1\xi_{1} and ξ2\xi_{2} that ensure the maximal value of σ\sigma, given by

σmax(δ)=14ηLachievedatδ=12ηL.\sigma_{max}(\delta)=\frac{1}{4\eta L}~{}{\rm achieved~{}at}~{}\delta=\frac{1}{2\eta L}. (29)

Since δ\delta and σ\sigma are functions of ϵ\epsilon and the sensor feedback amplifiers ξ1\xi_{1} and ξ2\xi_{2} in (15), the following results provide intervals for the feedback amplifiers and ϵ\epsilon that ensures the decay rate (29) is attained.

Theorem 3.

Define non-negative constants

c1±:=2α1η±4α12η2(1+ϵ)ρα11+ϵ,c2±:=2ϵα1βη±4(ϵα1βη)2(ϵα+βγ2)ϵμα1βϵα+βγ2,\begin{array}[]{ll}c_{1}^{\pm}:=\frac{{2}\alpha_{1}\eta\pm\sqrt{{4}\alpha_{1}^{2}\eta^{2}-(1+\epsilon)\rho\alpha_{1}}}{1+\epsilon},\\ c_{2}^{\pm}:=\frac{{2}\epsilon\alpha_{1}\beta\eta\pm\sqrt{{4}(\epsilon\alpha_{1}\beta\eta)^{2}-(\epsilon\alpha+\beta\gamma^{2})\epsilon\mu\alpha_{1}\beta}}{\epsilon\alpha+\beta\gamma^{2}},\end{array} (30)

with any ϵ\epsilon such that

βγ2μ4α1βη2αμ<ϵ<4α1η2ρρ.\frac{\beta\gamma^{2}\mu}{{4}\alpha_{1}\beta\eta^{2}-\alpha\mu}{<}\epsilon{<}\frac{{4}\alpha_{1}\eta^{2}-\rho}{\rho}. (31)

The decay rate σmax(δ)-\sigma_{max}(\delta) is achieved for the closed-loop system (7),(15) when the feedback amplifiers are chosen such as ξ1(c1,c1+),ξ2(c2,c2+).\xi_{1}\in{(}c_{1}^{-},c_{1}^{+}{)},\xi_{2}\in{(}c_{2}^{-},c_{2}^{+}{)}.

Proof 3.1.

Observe that to achieve the σmax(δ)\sigma_{max}(\delta) in (29), it is sufficient to have f1(ξ1,ϵ)>12ηf_{1}(\xi_{1},\epsilon){>}\frac{1}{{2}\eta} and f2(ξ2,ϵ)>12ηf_{2}(\xi_{2},\epsilon){>}\frac{1}{{2}\eta}.

Thus, f1(ξ1,ϵ)>12η,f_{1}(\xi_{1},\epsilon){>}\frac{1}{{2}\eta}, i.e. 2ξ1α1ρα1+(1+ϵ)ξ12>12η,\frac{2\xi_{1}\alpha_{1}}{\rho\alpha_{1}+(1+\epsilon)\xi_{1}^{2}}{>}\frac{1}{{2}\eta}, implies that

ϵ<h1(ξ1):=4α1ξ1ηρα1ξ12ξ12.\displaystyle\begin{array}[]{ll}\epsilon{<}h_{1}(\xi_{1}):=\frac{{4}\alpha_{1}\xi_{1}\eta-\rho\alpha_{1}-\xi_{1}^{2}}{\xi_{1}^{2}}.\end{array} (33)

Noting that 4α12η2ρα1>0{4}\alpha_{1}^{2}\eta^{2}-\rho\alpha_{1}>0 by (18), h1(ξ1)h_{1}(\xi_{1}) defines an upper bound for ϵ\epsilon, and observe that ξ1\xi_{1} must be chosen in between the following roots of h1h_{1} to ensure ϵ>0\epsilon>0, see Fig. 1

a1±:=2α1η±4α12η2ρα1,\displaystyle\begin{array}[]{l}a_{1}^{\pm}:={2}\alpha_{1}\eta\pm\sqrt{{4}\alpha_{1}^{2}\eta^{2}-\rho\alpha_{1}},\end{array}

Observe that h1(ξ1)0h_{1}(\xi_{1})\geq 0 if and only if ξ1(a1,a1+)\xi_{1}\in(a_{1}^{-},a_{1}^{+}). Seeking the critical points of h1(ξ1),h_{1}(\xi_{1}), h1ξ1=4ξ12α1η+2ξ1ρα1ξ4=0,\frac{\partial h_{1}}{\partial\xi_{1}}=\frac{-{4}\xi_{1}^{2}\alpha_{1}\eta+2\xi_{1}\rho\alpha_{1}}{\xi^{4}}=0, leads to ξ1=ρ2η,\xi_{1}=\frac{\rho}{{2}\eta}, for which h1h_{1} achieves its maximum value. Substituting h1(ξ1=ρη)h_{1}\left(\xi_{1}=\frac{\rho}{\eta}\right) into (33) yields the following upper bound for ϵ\epsilon

ϵ<4α1η2ρρ.\epsilon{<}\frac{{4}\alpha_{1}\eta^{2}-\rho}{\rho}. (35)

By f2(ξ2,ϵ)>12η,f_{2}(\xi_{2},\epsilon){>}\frac{1}{{2}\eta}, ϵ>h2(ξ2):=βγ2ξ224α1βξ2ημα1βαξ22.\epsilon{>}h_{2}(\xi_{2}):=\frac{\beta\gamma^{2}\xi_{2}^{2}}{{4}\alpha_{1}\beta\xi_{2}\eta-\mu\alpha_{1}\beta-\alpha\xi_{2}^{2}}. Since ϵ>0\epsilon>0, the denominator 4α1βξ2ημα1βαξ22{4}\alpha_{1}\beta\xi_{2}\eta-\mu\alpha_{1}\beta-\alpha\xi_{2}^{2} is chosen to be strictly positive. This leads to ξ2(a2,a2+)\xi_{2}\in(a_{2}^{-},a_{2}^{+}) where

a2±:=2α1βη±4α12β2η2αμα1βα,\displaystyle\begin{array}[]{ll}a_{2}^{\pm}:=&\frac{{2}\alpha_{1}\beta\eta\pm\sqrt{{4}\alpha_{1}^{2}\beta^{2}\eta^{2}-\alpha\mu\alpha_{1}\beta}}{\alpha},\end{array} (37)

and ξ2=a2\xi_{2}=a_{2}^{-} and ξ2=a2+\xi_{2}=a_{2}^{+} are the two vertical asymptotes of h2(ξ2),h_{2}(\xi_{2}), see dashed lines in Fig. 1. Note that this condition ensures h2(ξ2)0h_{2}(\xi_{2})\geq 0. Seeking the critical points of h2(ξ2)h_{2}(\xi_{2}) leads to ξ2=μη,\xi_{2}=\frac{\mu}{\eta}, for which h2(ξ2)h_{2}(\xi_{2}) takes its minimum value. Substituting h2(μ2η)h_{2}\left(\frac{\mu}{{2}\eta}\right) into (33) yields the lower bound for ϵ\epsilon

ϵ>βγ2μ4α1βη2αμ\displaystyle\begin{array}[]{l}\epsilon{>}\frac{\beta\gamma^{2}\mu}{{4}\alpha_{1}\beta\eta^{2}-\alpha\mu}\end{array} (39)

where 4α1βη2αμ>0{4}\alpha_{1}\beta\eta^{2}-\alpha\mu>0 by (18). Restricting h1(ξ1)h_{1}(\xi_{1}) and h2(ξ2)h_{2}(\xi_{2}) in between roots and asymptotes, respectively, ϵ\epsilon values corresponding to the filled regions in Fig. 1 satisfy conditions (35) and (39), respectively.

For any ϵ\epsilon that satisfies (31), f1(ξ1)>12ηf_{1}(\xi_{1}){>}\frac{1}{{2}\eta}, and f2>12ηf_{2}{>}\frac{1}{{2}\eta}, i.e.

(1+ϵ)ξ12+4α1ηξ1ρα10,(ϵα+βγ2)ξ22+4ϵα1βηξ2ϵμα1β0,\begin{array}[]{ll}-(1+\epsilon)\xi_{1}^{2}+{4}\alpha_{1}\eta\xi_{1}-\rho\alpha_{1}\geq 0,\\ -(\epsilon\alpha+\beta\gamma^{2})\xi_{2}^{2}+{4}\epsilon\alpha_{1}\beta\eta\xi_{2}-\epsilon\mu\alpha_{1}\beta\geq 0,\end{array}

yield ξ1(c1,c1+)\xi_{1}\in{(}c_{1}^{-},c_{1}^{+}{)} and ξ2(c2,c2+)\xi_{2}\in{(}c_{2}^{-},c_{2}^{+}{)}, respectively.

Refer to caption
Refer to caption
Figure 1: For the material parameters in Table I, bounds of ϵ\epsilon depend on h1(ξ1)h_{1}(\xi_{1}) and h2(ξ2)h_{2}(\xi_{2}). Note that the h2h_{2}-plot is logarithmic.

We prove that such ξ1(c1,c1+)\xi_{1}\in{(}c_{1}^{-},c_{1}^{+}{)} and ξ2(c2,c2+)\xi_{2}\in{(}c_{2}^{-},c_{2}^{+}{)} exist. Note that the sufficient condition for (c1,c1+),(c2,c2+){(}c_{1}^{-},c_{1}^{+}{)},{(}c_{2}^{-},c_{2}^{+}{)}\neq\emptyset is the existence of ϵ\epsilon satisfying (31) which is ensured by

4α1η2ρρ3α1η2ρ3,βγ2μ4α1βη2αμβγ2μ3αμ13.\displaystyle\begin{array}[]{ll}\frac{4\alpha_{1}\eta^{2}-\rho}{\rho}\geq\frac{3\alpha_{1}\eta^{2}}{\rho}\geq 3,&\frac{\beta\gamma^{2}\mu}{{4}\alpha_{1}\beta\eta^{2}-\alpha\mu}\leq\frac{\beta\gamma^{2}\mu}{3\alpha\mu}\leq\frac{1}{3}.\end{array} (41)
Refer to caption
Refer to caption
Figure 2: For the material parameters in Table I, graphs of c1±,c2±c_{1}^{\pm},c_{2}^{\pm} with respect to ϵ\epsilon satisfy the inequalities in (31).

The amplifier ξ1\xi_{1} maximizing h1h_{1} and the amplifier ξ2\xi_{2} minimizing h2h_{2} are always in the respective intervals, i.e. ξ1=ρ2η(c1,c1+)\xi_{1}=\frac{\rho}{{2}\eta}\in{(}c_{1}^{-},c_{1}^{+}{)} and ξ2=μ2η(c2,c2+)\xi_{2}=\frac{\mu}{{2}\eta}\in{(}c_{2}^{-},c_{2}^{+}{)}. Moreover, as ϵ\epsilon approaches to the upper (lower) bound (c1,c1+){(}c_{1}^{-},c_{1}^{+}{)} gets smaller (larger) and (c2,c2+){(}c_{2}^{-},c_{2}^{+}{)} gets larger (smaller), see Fig. 2. Feedback amplifiers chosen within the intervals ensure that f1(ξ1)12ηf_{1}(\xi_{1})\geq\frac{1}{{2}\eta} and f2(ξ2)12ηf_{2}(\xi_{2})\geq\frac{1}{{2}\eta} are satisfied, see Figs. 3 and 4.

Refer to caption
Figure 3: For ϵ1015,\epsilon\approx 10^{-15}, f1(ξ1)>12ηf_{1}(\xi_{1})>\frac{1}{{2}\eta} when ξ1(c1,c1+).\xi_{1}\in{(}c_{1}^{-},c_{1}^{+}{)}.
Refer to caption
Figure 4: For ϵ1015,\epsilon\approx 10^{-15}, f2(ξ2)>12ηf_{2}(\xi_{2})>\frac{1}{{2}\eta} when ξ2(c2,c2+).\xi_{2}\in{(}c_{2}^{-},c_{2}^{+}{)}.

4 Simulations and Numerical Experiments

In this section, we present numerical simulations primarily aimed at demonstrating the following:

  • The exponential stability of the system (7) with (15) is showcased using feedback amplifiers ξ1(c1,c1+)\xi_{1}\in(c_{1}^{-},c_{1}^{+}) and ξ2(c2,c2+)\xi_{2}\in(c_{2}^{-},c_{2}^{+}), as obtained in Theorem 3.

  • Comparison of the reduced model exponential decay rates for a selection of feedback amplifiers.

Moreover, we address a common challenge encountered in standard model reductions, such as Finite Differences (FD) and Finite Elements (FE), which introduce spurious high-frequency vibrational modes to the system. These spurious modes significantly impact the decay rate of the system. In fact, reduced models lack the uniform observability property as the mesh parameter approaches zero, necessitating filtering techniques such as the direct Fourier filtering [3, 11] or the indirect filtering [30].

To mitigate the adverse effects of approximation methods on the decay rate, we consider a recently introduced technique known as Order-Reduced Finite Differences (ORFD) [12]. The ORFD method has been observed to retain the decay rate of the infinite-dimensional system. For rigorous stability analyses of all three approximations, interested readers can refer to [2].

Consider a given natural number NN\in\mathbb{N}, which denotes the number of nodes in the spatial semi-discretization. Let’s introduce the mesh size as h:=1N+1h:=\frac{1}{N+1}, and discretize the interval [0,L][0,L] as 0=x0<x1<<xj=jh<<xN<xN+1=L.0=x_{0}<x_{1}<\dots<x_{j}=j*h<\ldots<x_{N}<x_{N+1}=L. Let (vj,pj)=(vj,pj)(t)(v,p)(xj,t)(v_{j},p_{j})=(v_{j},p_{j})(t)\approx(v,p)(x_{j},t) be the approximation of the solution (v,p)(x,t)(v,p)(x,t) of (7)-(15) at the point space xj=jhx_{j}=j\cdot h for any j=0,1,,N,N+1j=0,1,...,N,N+1, and let v=[v1,v2,,vN+1]T\vec{v}=[v_{1},v_{2},...,{v_{N+1}}]^{T} and p=[p1,p2,,pN+1]T\vec{p}=[p_{1},p_{2},...,{p_{N+1}}]^{T}.

In the ORFD method, in addition to the standard nodes {xj}j=0N,\left\{x_{j}\right\}_{j=0}^{N}, we also consider the in-between middle nodes of each subinterval, denoted by {xj+12:=xj+1+xj2}j=0N.\left\{x_{j+\frac{1}{2}}:=\frac{x_{j+1}+x_{j}}{2}\right\}_{j=0}^{N}. Define the average, vj+12:=vj+1+uj2v_{j+\frac{1}{2}}:=\frac{v_{j+1}+u_{j}}{2} and difference operators

δxvj+12:=vj+1vjh,δx2vj:=vj+12vj+vj1h2,\displaystyle\begin{array}[]{ll}\delta_{x}v_{j+\frac{1}{2}}:=\frac{v_{j+1}-v_{j}}{h},&\delta_{x}^{2}v_{j}:=\frac{v_{j+1}-2v_{j}+v_{j-1}}{h^{2}},\end{array} (43)

It is worth noting that by considering odd-number derivatives at the in-between nodes within the uniform discretization of [0,L][0,L], higher-order approximations are achieved [12].

The ORFD approximation of equations (7) with (15) is

(𝑪1𝑴)[v¨p¨]+(𝑪2𝑨h)[vp]+(𝑪3)[v˙p˙]=0,\left(\bm{C}_{1}\otimes\bm{M}\right)\begin{bmatrix}{\ddot{\vec{v}}}\\ {\ddot{\vec{p}}}\\ \end{bmatrix}+\left(\bm{C}_{2}\otimes\bm{A}_{h}\right)\begin{bmatrix}{{\vec{v}}}\\ {{\vec{p}}}\\ \end{bmatrix}\\ +\left(\bm{C}_{3}\otimes\mathcal{B}\right)\begin{bmatrix}{\dot{\vec{v}}}\\ {\dot{\vec{p}}}\\ \end{bmatrix}=\vec{0}, (44)

where C1C_{1} and C2C_{2} are matrices for material parameters, and C3C_{3} is the matrix for the state feedback amplifiers, defined as

𝑪1=[ρ00μ],𝑪2=[αγβγββ],𝑪3=[ξ100ξ2],\bm{C}_{1}=\begin{bmatrix}\rho&0\\ 0&\mu\end{bmatrix},\quad\bm{C}_{2}=\begin{bmatrix}\alpha&-\gamma\beta\\ -\gamma\beta&\beta\end{bmatrix},\quad\bm{C}_{3}=\begin{bmatrix}\xi_{1}&0\\ 0&\xi_{2}\end{bmatrix},

the (N+1)×(N+1)(N+1)\times(N+1) mass matrix 𝑴\bm{M} is defined by

𝑴=14[210012100001210011],\bm{M}=\frac{1}{4}\begin{bmatrix}2&1&0&\dots&\dots&\dots&0\\ 1&2&1&0&\dots&\dots&0\\ &\ddots&\ddots&\ddots&\ddots&\ddots&\\ 0&\dots&\dots&0&1&2&1\\ 0&\dots&\dots&\dots&0&1&1\\ \end{bmatrix},

the (N+1)×(N+1)(N+1)\times(N+1) central difference matrix 𝑨h\bm{A}_{h} is

𝑨h=1h2[210012100001210011],\bm{A}_{h}=\frac{1}{h^{2}}\begin{bmatrix}2&-1&0&\dots&\dots&\dots&0\\ -1&2&-1&0&\dots&\dots&0\\ &\ddots&\ddots&\ddots&\ddots&\ddots&\\ 0&\dots&\dots&0&-1&2&-1\\ 0&\dots&\dots&\dots&0&-1&1\\ \end{bmatrix},

and the (N+1)×(N+1)(N+1)\times(N+1) boundary node matrix \mathcal{B} is a zero matrix except for the (N+1)×(N+1)(N+1)\times(N+1)-th entry, which is 1h,\frac{1}{h}, due to the boundary damping being injected at the last node. Here, \otimes denotes the matrix Kronecker product. It’s worth noting that equation (44) can be rewritten in the first-order form as

ddt[vpv˙p˙]T=𝓐[vpv˙p˙]T,\frac{d}{dt}\begin{bmatrix}{{\vec{v}}}&{{\vec{p}}}&{\dot{\vec{v}}}&{\dot{\vec{p}}}\end{bmatrix}^{T}=\bm{\mathcal{A}}\begin{bmatrix}{{\vec{v}}}&{{\vec{p}}}&{\dot{\vec{v}}}&{\dot{\vec{p}}}\end{bmatrix}^{T}, (45)
𝓐=[𝟎2N+2I2N+2(𝑪11𝑪2)(𝑴1𝑨h)(𝑪11𝑪3)(𝑴1)]\bm{\mathcal{A}}=\begin{bmatrix}\bm{0}_{2N+2}&I_{2N+2}\\ -\left(\bm{C}_{1}^{-1}\bm{C}_{2}\right)\otimes\left(\bm{M}^{-1}\bm{A}_{h}\right)&-\left(\bm{C}_{1}^{-1}\bm{C}_{3}\right)\otimes\left(\bm{M}^{-1}\mathcal{B}\right)\end{bmatrix} (46)

The discretized energy corresponding to (45) is

Eh(t):=h2(𝑪1𝑴)[v˙p˙],[v˙p˙]+h2(𝑪1𝑨h)[vp],[vp].\begin{split}E_{h}(t):=&\frac{h}{2}\left\langle(\bm{C}_{1}\otimes\bm{M})\begin{bmatrix}{\dot{\vec{v}}}\\ {\dot{\vec{p}}}\end{bmatrix},\begin{bmatrix}{\dot{\vec{v}}}\\ {\dot{\vec{p}}}\end{bmatrix}\right\rangle\\ &\qquad\qquad+\frac{h}{2}\left\langle(\bm{C}_{1}\otimes\bm{A}_{h})\begin{bmatrix}{{\vec{v}}}\\ {{\vec{p}}}\end{bmatrix},\begin{bmatrix}{{\vec{v}}}\\ {{\vec{p}}}\end{bmatrix}\right\rangle.\end{split} (47)

To show the importance of our analysis for the choices of sensor feedback amplifiers via the optimization process outlined in the previous section, sample numerical simulations are presented for the material constants in Table 1.

Table 1: Realistic Piezoelectric Material Parameters
Parameter Symbol Value Unit
Length of the beam LL 11 m
Mass density ρ\rho 60006000 kg/m3{\rm m}^{3}
Magnetic permeability μ\mu 10610^{-6} H/m
Elastic stiffness α\alpha 10910^{9} N/m2{\rm m}^{2}
Piezoelectric constant γ\gamma 10310^{-3} C/m3{\rm m}^{3}
Impermittivity β\beta 101210^{12} m/F

Note that σmax(δ)\sigma_{max}(\delta) in (29) depends solely on the material parameters. In particular, σmax=14ηL102.04\sigma_{\rm max}=\frac{1}{4\eta L}\approx 102.04 for the material parameters in Table 1. Letting ϵ=1\epsilon=1, which is a valid choice for any system according to equations (41), the intervals for optimal sensor feedback amplifiers in (30), corresponding to σmax,\sigma_{\rm max}, are as follows:

(c1,c1+)(7.17×105,4.18×106),(c2,c2+)(1.02×104,9.78×109).\begin{split}(c_{1}^{-},c_{1}^{+})&\approx(7.17\times 10^{5},~{}4.18\times 10^{6}),\\ (c_{2}^{-},c_{2}^{+})&\approx(1.02\times 10^{-4},~{}9.78\times 10^{9}).\end{split} (48)

Here note that both the maximal decay rate σmax-\sigma_{\rm max} and the corresponding intervals for feedback amplifiers in (48) are independent of the choice of initial conditions.

For simulations, we take N=80N=80 nodes, and we choose triangular (hat-type) initial conditions that are composed of high and low-frequency vibrational nodes. The simulations are performed for T=0.1T=0.1 sec to show the rapid exponential decay of solutions. Indeed, the solutions of the system (44) with the choices of optimal feedback amplifiers ξ1=106(c1,c1+)\xi_{1}=10^{6}\in(c_{1}^{-},c_{1}^{+}) and ξ2=109(c2,c2+)\xi_{2}=10^{9}\in(c_{2}^{-},c_{2}^{+}), are observed drive all states to zero state rapidly, e.g. see Fig. 5, and the total energy follows the exponentially decay in a higher rate than claimed σmax\sigma_{\rm max}, Fig. 7.

Our analysis indicates that even a relatively small perturbation in the values of the feedback amplifiers ξ1\xi_{1} can lead to suboptimal performance. Indeed, when ξ1=104\xi_{1}=10^{4} and does not fall within the range (c1,c1+)(c_{1}^{-},c_{1}^{+}), the system still continues to exhibit low and high-frequency longitudinal oscillations for T=0.1T=0.1 seconds, as demonstrated in Figure 6. Furthermore, the energy of the system decays in an unstable manner, with a decay rate significantly larger than σmax-\sigma_{\rm max}, as depicted in Figure 7.

Refer to caption
Figure 5: System behavior for the optimal choice of feedback amplifiers: ξ1=106\xi_{1}=10^{6} and ξ2=109\xi_{2}=10^{9}. Both solutions v(x,t)v(x,t) and p(x,t)p(x,t) of Equation (44) decay rapidly to the zero state in much less than 0.10.1 sec.
Refer to caption
Figure 6: For a non-optimal choice of ξ1=104\xi_{1}=10^{4}, which control longitudinal strains, and an optimal choice of ξ2=109\xi_{2}=10^{9}, controlling the voltage, the longitudinal oscillations v(x,t)v(x,t) for eq. (44) exhibit a significantly slower decay rate.
ξ1\xi_{1} ξ2\xi_{2} 10510^{5} 105.510^{5.5} 10610^{6} 106.510^{6.5} 10710^{7} 107.510^{7.5}
10710^{-7} 0.1-0.1 0.1-0.1 0.1-0.1 0.1-0.1 0.1-0.1 0.1-0.1
10510^{-5} 10-10 10-10 10-10 10-10 10-10 10-10
10310^{-3} 17-17 53-53 177-177 421-421 101.9-101.9 32-32
10610^{6} 17-17 53-53 177-177 421-421 101.9-101.9 31-31
10910^{9} 17-17 53-53 177-177 421-421 101-101 30-30
101010^{10} 17-17 52-52 100-100 100-100 97-97 23-23
101110^{11} 10-10 10-10 10-10 10-10 10-10 9-9
Table 2: The max{(μk)}\max\{\Re(\mu_{k})\} values for different choices of feedback amplifiers ξ1\xi_{1} and ξ2\xi_{2}. Red boxes indicate the intervals (c1,c1+)(c_{1}^{-},c_{1}^{+}) and (c2,c2+)(c_{2}^{-},c_{2}^{+}), where the maximal decay rate is achieved.

To demonstrate the robustness of our theoretical findings, we examine the spectrum of the numerical approximation. Let μk\mu_{k} denote the eigenvalues of 𝓐\bm{\mathcal{A}} as defined in Equation (46). The decay rate of the system (7) with (15), and the theoretically derived decay rate σ-\sigma in Theorem (2), can be approximated as the maximal real part of the eigenvalues of 𝓐\bm{\mathcal{A}}, specifically σmax{(μk)}-\sigma\approx\max\{\Re(\mu_{k})\}. A contour plot of max{(μk)}\max\{\Re(\mu_{k})\} in terms of the feedback amplifiers ξ1\xi_{1} and ξ2\xi_{2} is presented in Figure 8. It is notable that as the feedback amplifiers ξ1\xi_{1} and ξ2\xi_{2} fall within the intervals (c1,c1+)(c_{1}^{-},c_{1}^{+}) and (c2,c2+)(c_{2}^{-},c_{2}^{+}), respectively shown as red lines, the maximal decay rate is achieved. Conversely, the decay rate significantly diminishes outside of these intervals. Exact values of max{(μk)}\max\{\Re(\mu_{k})\} for selected feedback amplifiers are also provided in Table 2. As a final note, the decay rates are much smaller than σmax-\sigma_{\rm max} when both feedback amplifiers are chosen from the intervals (c1,c1+)(c_{1}^{-},c_{1}^{+}) and (c2,c2+)(c_{2}^{-},c_{2}^{+}).

5 Conclusions & Future Work

In summary, using the Lyapunov approach, we derived the exponential decay rate ``σ"``-\sigma" for system (7), with the design of state feedback amplifiers ξ1\xi_{1} and ξ2\xi_{2} as described in (15). This decay rate ensures fast stabilization, though improvements are possible. The rate depends on material parameters and feedback amplifiers. Our numerical results confirm the robustness of this decay rate, with deviations affecting it from the theoretical value. Interactive simulations are available through the Wolfram Demonstrations Project [31].

A full numerical analysis and proof of exponential stability for the semi-discretized model (44), as h0h\to 0, and determining optimal designs of feedback amplifiers remain beyond this paper’s scope and are ongoing research. However, the Lyapunov approach here serves as the basis for results in [2], where the discretized model achieves the same decay rate as the PDE model.

Future work can extend these concepts to multi-layer magnetizable piezoelectric beams, which have numerous practical applications [19]. The design of feedback amplifiers may become more delicate when there are more than two state feedback amplifiers, as highlighted in [19]. There is also potential to explore general hyperbolic PDE systems with multiple boundary dampers. Relevant studies address systems with random inputs [14] and the optimal design and placement of actuators in higher-dimensional systems [16, 15], both of which are valuable for future research.

Refer to caption
Figure 7: Normalized total energies for (44) with optimal and non-optimal feedback amplifiers, along with the theoretically derived energy decay as stated in Theorem (3).
Refer to caption
Figure 8: Decay rates for the ORFD approximation of the system (7) with (15), showing dependence on feedback amplifiers ξ1\xi_{1} and ξ2\xi_{2}. Red lines mark the intervals (c1,c1+)(c_{1}^{-},c_{1}^{+}) and (c2,c2+)(c_{2}^{-},c_{2}^{+}), where the maximal decay rate is achieved. Lighter blue and brown areas indicate better decay rates, while darker blue regions show slower rates outside these intervals.

References

  • [1] A. K. Aydın, A.Ö Özer, J. Walterman, A Novel Finite Difference-Based Model Reduction and a Sensor Design for a Multilayer Smart Beam With Arbitrary Number of Layers, IEEE Control Syst. Lett., 7 (2023), 1548-553.
  • [2] A.Ö. Özer, A.K. Aydin, J. Walterman, A Robust Finite-Difference Model Reduction for the Boundary Feedback Stabilization of Fully-dynamic Piezoelectric Beams, arXiv:2309.07492, under revision, 2023.
  • [3] H.E. Boujaoui, H. Bouslous, L. Maniar, Boundary Stabilization for 1-d Semi-Discrete Wave Equation by Filtering Technique Bull. TICMI, 17-1 (2013), 1–18.
  • [4] N. Cîndea, A. Münch, A mixed formulation for the direct approximation of the control of minimal L2L_{2}-norm for linear type wave equations, Calcolo, 52-3 (2015), 245-288.
  • [5] C. Dagdeviren, et al., Recent Progress in Flexible and Stretchable Piezoelectric Devices for Mechanical Energy Harvesting, Sensing and Actuation, Extreme Mech. Lett., 9(1) (2016), 269-281.
  • [6] M.C. de Jong, et al., On control of voltage-actuated piezoelectric beam: A Krasovskii passivity-based approach, Eur. J. Control, 69 (2023), 100724.
  • [7] W. Dong, et al., Wearable human–machine interface based on PVDF piezoelectric sensor, Trans. Inst. Meas. Control, 39-4 (2017), 398–403.
  • [8] I. Ftouhi, E. Zuazua, Optimal design of sensors via geometric criteria, J. Geom. Anal., 33-8 (2023), 253.
  • [9] B. Geshkovski, E. Zuazua, Optimal actuator design via Brunovsky’s normal form, IEEE Trans. Autom. Control, (2022) pp. 6641-6650.
  • [10] G.Y. Gu, et al., Modeling and Control of Piezo-Actuated Nanopositioning Stages: A Survey, IEEE Trans. Autom. Sci. Eng., 13-1 (2016), 313–332.
  • [11] J.A Infante and E. Zuazua, Boundary observability for the space semi-discretizations of the 1-D wave equation, Math. Model. Num. Anal., 33 (1999), 407-438.
  • [12] J. Liu, B.Z. Guo, A new semi-discretized order reduction finite difference scheme for uniform approximation of 1-D wave equation, SIAM J. Control Optim., 58 (2020), 2256-2287.
  • [13] K. Kunisch, P. Trautmann, B. Vexler, Optimal control of the undamped linear wave equation with measure valued controls, SIAM Journal on Control and Optimization, 2016, 1212-1244.
  • [14] F.J. Marín, J. Martínez-Frutos, F. Periago Robust averaged control of vibrations for the Bernoulli-Euler beam equation. Journal of Optimization Theory and Applications. 2017, 174-428.
  • [15] K. Morris, M.A. Demetriou, S.D. Yang, Using H2-control performance metrics for the optimal actuator location of distributed parameter systems, IEEE Trans. Autom. Control, 60-2 (2015), 450-462.
  • [16] A. Münch, P. Pedregal, F. Periago, Optimal Internal Stabilization of the Linear System of Elasticity. Arch Rational Mech Anal 193 (2009), 171–193.
  • [17] K. Morris, A.Ö. Özer, Modeling and stability of voltage-actuated piezoelectric beams with electric effect, SIAM J. Control Optim., 52-4 (2014), 2371-2398.
  • [18] A.Ö. Özer, Further stabilization and exact observability results for voltage-actuated piezoelectric beams with magnetic effects, Math. Control Signals Syst., 27 (2015), 219-244.
  • [19] A.Ö. Özer, Modeling and controlling an active constrained layered (ACL) beam actuated by two voltage sources with/without magnetic effects, IEEE Trans. Autom. Control, 62-12 (2017), 6445–6450.
  • [20] A.Ö. Özer, K.A. Morris, Modeling and related results for current-actuated piezoelectric beams by including magnetic effects, ESAIM: COCV, 26-8 (2019), 1-24.
  • [21] A.Ö. Özer, W. Horner, Uniform boundary observability of Finite Difference approximations of non-compactly-coupled piezoelectric beam equations, Appl. Anal., 101-5 (2022), 1571-1592.
  • [22] Y. Privat, E. Trélat, E. Zuazua, Actuator design for parabolic distributed parameter systems with the moment method, SIAM J. Control Optim., 55-2 (2017), 1128-1152.
  • [23] A.J.A. Ramos, M.M. Freitas, D.S. Almeida Jr., S.S. Jesus, T.R.S. Moura, Equivalence between exponential stabilization and boundary observability for piezoelectric beams with magnetic effect, Z. Angew. Math. Phys., 70-60 (2019).
  • [24] Ultrasound imaging of the brain and liver, Science Daily, Acoust. Soc. Am., 26 June 2017.
  • [25] Q. Shi, T. Wang, C. Lee, MEMS Based Broadband Piezoelectric Ultrasonic Energy Harvester (PUEH) for Enabling Self-Powered Implantable Biomedical Devices, Sci. Rep., 6 (2016), 24946.
  • [26] R.C. Smith, Smart Material Systems, (SIAM, 2005).
  • [27] T. Voss, J.M.A. Scherpen, Stabilization and shape control of a 1D piezoelectric Timoshenko beam, Automatica, 47-12 (2011), 2780-85.
  • [28] J. Yang, An Introduction to the Theory of Piezoelectricity, Springer, New York, 2005.
  • [29] B. Yubo, C. Prieur, Z. Wang, Stabilization of a Rayleigh beam with collocated piezoelectric sensor/actuator. Evol. Equ. Control Theory, (2023).
  • [30] L.T. Tebou, E. Zuazua, Uniform boundary stabilization of the finite difference space discretization of the 1-D wave equation. Adv. Comput. Math., 26 (2007), 337-365.
  • [31] J. Walterman, A.K. Aydin, A.Ö. Özer, Piezoelectric beam dynamics, Wolfram Demonstrations Project, Published: June 13, 2023.
  • [32] H.S. Wang, et al., Biomimetic and flexible piezoelectric mobile acoustic sensors with multiresonant ultrathin structures for machine learning biometrics, Sci. Adv., 7-7 (2021).