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

Flatness-based motion planning for a non-uniform moving cantilever Euler-Bernoulli beam with a tip-mass

Soham Chatterjee, Aman Batra and Vivek Natarajan A. Batra is supported by the Prime Minister’s Research Fellowship via the grant RSPMRF0262.S. Chatterjee ([email protected]), A. Batra ([email protected]) and V. Natarajan ([email protected]) are with the Centre for Systems and Control Engineering, Indian Institute of Technology Bombay, Mumbai 400076, India.
Abstract

Consider a non-uniform Euler-Bernoulli beam with a tip-mass at one end and a cantilever joint at the other end. The cantilever joint is not fixed and can itself be moved along an axis perpendicular to the beam. The position of the cantilever joint is the control input to the beam. The dynamics of the beam is governed by a coupled PDE-ODE model with boundary input. On a natural state-space, there exists a unique state trajectory for this beam model for every initial state and each smooth control input which is compatible with the initial state. In this paper, we study the motion planning problem of transferring the beam from an initial state to a final state over a prescribed time-interval. We address this problem by extending the generating functions approach to flatness-based control, originally proposed in the literature for motion planning of parabolic PDEs, to the beam model. We prove that such a transfer is possible if the initial and final states belong to a cer-tain set, which also contains steady-states of the beam. We illust-rate our theoretical results using simulations and experiments.

Index Terms:
Coupled PDE-ODE model, flatness, generating functions, flexible structures, spatially-varying coefficients.

I Introduction

Consider the following model of a non-uniform moving cantilever Euler-Bernoulli beam with a tip-mass at one end and a cantilever joint at the other end:

ρ(x)wtt(x,t)+(EIwxx)xx(x,t)=0,\displaystyle\rho(x)w_{tt}(x,t)+(EIw_{xx})_{xx}(x,t)=0, (1)
mwtt(0,t)+(EIwxx)x(0,t)=0,\displaystyle\hskip 5.69054ptmw_{tt}(0,t)+(EIw_{xx})_{x}(0,t)=0, (2)
Jwxtt(0,t)EI(0)wxx(0,t)=0,\displaystyle\hskip 5.69054ptJw_{xtt}(0,t)-EI(0)w_{xx}(0,t)=0, (3)
w(L,t)=f(t),wx(L,t)=0.\displaystyle\hskip 5.69054ptw(L,t)=f(t),\qquad w_{x}(L,t)=0. (4)

Here L>0L>0 is the length of the beam, w(x,t)w(x,t) is the displacement of the beam at the location x[0,L]x\in[0,L] and time t(0,)t\in(0,\infty), ρ(x)\rho(x) and EI(x)EI(x) are the mass per unit length and flexural rigidity, respectively, of the beam at x[0,L]x\in[0,L], and m>0m>0 and J>0J>0 are the mass and moment of inertia, respectively, of the tip-mass located at the x=0x=0 end of the beam. We suppose that ρC4[0,L]\rho\in C^{4}[0,L], EIC4[0,L]EI\in C^{4}[0,L] and they are strictly positive, i.e infx[0,L]ρ(x)>0\inf_{x\in[0,L]}\rho(x)>0 and infx[0,L]EI(x)>0\inf_{x\in[0,L]}EI(x)>0. The displacement of the cantilever joint at the x=Lx=L end of the beam is determined by the scalar control input ff. The coupled PDE-ODE model (1)-(4) governs the dynamics of engineering systems which have a moving cantilever beam such as single-axis flexible cartesian robots. Figure 1 shows an experimental setup consisting of a non-uniform moving cantilever beam with a tip-mass.

[Uncaptioned image]

Figure 1. Experimental setup of a non-uniform moving cantilever beam with tip-mass. The green marker on the tip-mass is used to track the tip-mass position with a high speed camera.

There exists a natural state-space ZZ for the non-uniform moving cantilever Euler-Bernoulli beam model (1)-(4) in which it has a unique state trajectory for every initial state z0z_{0} and each smooth control input ff which is compatible with z0z_{0}, see Section II. In this paper, we study the motion planning problem of computing a control input ff which transfers (1)-(4) from an initial state z0z_{0} to a final state zTz_{T} over a prescribed time-interval [0,T][0,T]. We solve this problem for a set of initial and final states in ZZ, which includes the steady-states of the beam.

Over the past three decades, flatness has emerged as a predominant technique for solving motion planning problems for dynamical systems governed by PDEs, including flexible structures. The main idea of flatness is to express the state and the input of the PDE as an infinite linear combination of a flat output and its time derivatives. Then, based on the desired motion, an appropriate trajectory is selected for the flat output using which the input necessary to execute the motion is computed. Using the transform approach (Laplace transform or Mikusiński’s operational calculus) to flatness, motion planning problems have been solved for Euler-Bernoulli beams with fixed cantilever joints in [10], rotating cantilever joints in [1], [5], [7] and [13] and translating cantilever joints in [2]. More recently, the power series approach to flatness has been used in [3], [4], [6] to solve motion planning problems for Euler-Bernoulli beams with some other boundary conditions. The Riesz spectral approach to flatness, see [9], is in general applicable to PDEs with spatially-varying coefficients. However, addressing motion planning problems for non-uniform Euler-Bernoulli beams using this approach requires spectral assumptions that are hard to verify [11]. Moreover, the admissibility assumption on the control operator in [11] does not hold for our beam model ​​ (1)-(4).

Motion planning of 1D PDEs using the generating functions approach to flatness involves solving a sequence of initial value ODEs recursively to obtain the generating functions and then expressing the input and solution of the PDE in terms of these functions. In principle, this approach is well-suited for PDEs with spatially-varying coefficients and it has been used to address a motion planning problem (null control problem) for 1D parabolic PDEs with highly irregular coefficients in [8]. In the present work, we extend this approach to solve the motion planning problem described above for the non-uniform Euler-Bernoulli beam model (1)-(4). The extension inherently leads to a certain infinite-order differential equation which we solve by establishing that a pair of infinite-order differential operators commute. We show that if the initial and final states belong to a certain subspace of the state-space which contains steady-states of the beam, then our motion planning problem has a solution for any prescribed time-interval. We illustrate our theoretical results using both simulations and experiments.

The rest of the paper is organized as follows: In Section II we establish the well-posedness of the Euler-Bernoulli beam model (1)-(4). We define the generating functions for the beam model in Section III and derive some estimates for them. Section IV contains our solution to the motion planning problem and in Section V we present our numerical and experimental results.

II Well-posedness

In this section we establish the existence and uniqueness of solutions to the coupled PDE-ODE model (1)-(4). Let Hn(0,L)H^{n}(0,L) denote the usual Sobolev space of order nn on the interval (0,L)(0,L). A natural choice of state space for (1)-(4) is the Hilbert space

Z={[uvαβ]H2(0,L)×L2(0,L)××|ux(L)=0}.Z=\{[u\ \ v\ \ \alpha\ \ \beta]\in H^{2}(0,L)\times L^{2}(0,L)\times{\mathbb{R}}\times{\mathbb{R}}\big{|}u_{x}(L)=0\}.

For z1=[u1v1α1β1]z_{1}=[u_{1}\ v_{1}\ \alpha_{1}\ \beta_{1}] and z2=[u2v2α2β2]z_{2}=[u_{2}\ v_{2}\ \alpha_{2}\ \beta_{2}] in ZZ, the inner product z1,z2Z=0LEI(x)u1,xx(x)u2,xx(x)d x+0Lu1(x)u2(x)d x+0Lρ(x)v1(x)v2(x)d x+mα1α2+Jβ1β2\langle z_{1},z_{2}\rangle_{Z}=\int_{0}^{L}EI(x)u_{1,xx}(x)u_{2,xx}(x){\rm d}\hbox{\hskip 0.5pt}x+\int_{0}^{L}u_{1}(x)u_{2}(x){\rm d}\hbox{\hskip 0.5pt}x+\int_{0}^{L}\rho(x)v_{1}(x)v_{2}(x){\rm d}\hbox{\hskip 0.5pt}x+m\alpha_{1}\alpha_{2}+J\beta_{1}\beta_{2}. We now define the notion of classical solutions for (1)-(4) in ZZ.

Definition II.1

​Given T>0T\!>\!0, z0Zz_{0}\!\in\!Z and fC([0,T];)f\!\in\!C([0,T];{\mathbb{R}}), a function zC([0,T];Z)z\in C([0,T];Z) is a classical solution of (1)-(4) on the time interval [0,T][0,T] for the initial state z0z_{0} and input ff if z(0)=z0z(0)=z_{0} and

z(t)=[w(,t)wt(,t)wt(0,t)wxt(0,t)] t[0,T],z(t)=[w(\cdot,t)\ \ w_{t}(\cdot,t)\ \ w_{t}(0,t)\ \ w_{xt}(0,t)]\qquad\forall{\hbox{\hskip 1.0pt}}t\in[0,T], (5)

where wC([0,T];H4(0,L))C1([0,T];H2(0,L))C2([0,T];L2(0,L))w\in C([0,T];H^{4}(0,L))\cap C^{1}([0,T];H^{2}(0,L))\cap C^{2}([0,T];L^{2}(0,L)) with w(0,),wx(0,)C2([0,T];)w(0,\cdot),w_{x}(0,\cdot)\in C^{2}([0,T];{\mathbb{R}}) and ww and ff satisfy (1)-(4) for each t(0,T)t\in(0,T). \square

Consider the following dense subspace of ZZ:

V=\displaystyle V= {[uvαβ]H4(0,L)×H2(0,L)×× |\displaystyle\{[u\ \ v\ \ \alpha\ \ \beta]\in H^{4}(0,L)\times H^{2}(0,L)\times{\mathbb{R}}\times{\mathbb{R}}{\hbox{\hskip 1.0pt}}\big{|}{\hbox{\hskip 1.0pt}}
ux(L)=vx(L)=0,v(0)=α,vx(0)=β}.\displaystyle\quad u_{x}(L)=v_{x}(L)=0,\,v(0)=\alpha,\,v_{x}(0)=\beta\}. (6)

In the next proposition we will establish the existence and uniqueness of a classical solution for (1)-(4) when the initial state belongs to VV and the input ff is smooth and compatible with the initial state. Since the proof of the proposition is based on standard techniques, we omit detailed explanations in the proof.

Proposition II.2

Let T>0T>0, an input fC3([0,T];)f\in C^{3}([0,T];{\mathbb{R}}) and an initial state z0=[u0v0α0β0]Vz_{0}=[u_{0}\ \ v_{0}\ \ \alpha_{0}\ \ \beta_{0}]\in V with f(0)=u0(L)f(0)=u_{0}(L) and f˙(0)=v0(L)\dot{f}(0)=v_{0}(L) be given. There exists a unique classical solution zC([0,T];Z)z\in C([0,T];Z) of (1)-(4) on the time interval [0,T][0,T] for the initial state z0z_{0} and input ff.

Proof:

Choose νC[0,L]\nu\in C^{\infty}[0,L] such that ν(0)=νx(0)=νxx(0)=νxxx(0)=νx(L)=0\nu(0)=\nu_{x}(0)=\nu_{xx}(0)=\nu_{xxx}(0)=\nu_{x}(L)=0 and ν(L)=1\nu(L)=1. Replacing w(x,t)w(x,t) with w~(x,t)+ν(x)f(t)\tilde{w}(x,t)+\nu(x)f(t) formally in (1)-(4) we get that for x(0,L)x\in(0,L) and t>0t>0,

ρ(x)w~tt(x,t)+(EIw~xx)xx(x,t)\displaystyle\rho(x)\tilde{w}_{tt}(x,t)+(EI\tilde{w}_{xx})_{xx}(x,t)
+(EI(νxx)xx(x)f(t)+ρ(x)ν(x)f¨(t)=0,\displaystyle\hskip 5.69054pt+(EI(\nu_{xx})_{xx}(x)f(t)+\rho(x)\nu(x)\ddot{f}(t)=0, (7)
mw~tt(0,t)+(EIw~xx)x(0,t)=0,\displaystyle m\tilde{w}_{tt}(0,t)+(EI\tilde{w}_{xx})_{x}(0,t)=0, (8)
Jw~xtt(0,t)EI(0)w~xx(0,t)=0,\displaystyle J\tilde{w}_{xtt}(0,t)-EI(0)\tilde{w}_{xx}(0,t)=0, (9)
w~(L,t)=0,w~x(L,t)=0.\displaystyle\hskip 14.22636pt\tilde{w}(L,t)=0,\quad\tilde{w}_{x}(L,t)=0. (10)

The subspace Z~={[uvαβ]Z | u(L)=0}\tilde{Z}=\{[u\ v\ \alpha\ \beta]\in Z{\hbox{\hskip 1.0pt}}\big{|}{\hbox{\hskip 1.0pt}}u(L)=0\} is a closed subset of ZZ and is itself a Hilbert space. The coupled PDE-ODE system (7)-(10) can be written as an abstract evolution equation on Z~\tilde{Z} as follows:

z~˙(t)=𝒜z~(t)+[f(t)f¨(t)]t>0.\dot{\tilde{z}}(t)={\mathcal{A}}\tilde{z}(t)+{\mathcal{B}}\bigl{[}f(t)\ \ \ddot{f}(t)\bigr{]}\qquad\forall\,t>0. (11)

The operators 𝒜{\mathcal{A}} and {\mathcal{B}} are defined as follows: The domain of 𝒜{\mathcal{A}} is 𝒟(𝒜)={[uvαβ]V | u(L)=v(L)=0}{\mathcal{D}}({\mathcal{A}})=\{[u\ v\ \alpha\ \beta]\in V{\hbox{\hskip 1.0pt}}\big{|}{\hbox{\hskip 1.0pt}}u(L)=v(L)=0\} and 𝒜[uvαβ]=[v(EIuxx)xxρ(EIuxx)x(0)mEI(0)uxx(0)J]{\mathcal{A}}[u\ v\ \alpha\ \beta]=[v\ \ -\frac{(EIu_{xx})_{xx}}{\rho}\ \ -\frac{(EIu_{xx})_{x}(0)}{m}\ \ \frac{EI(0)u_{xx}(0)}{J}] for [uvαβ]𝒟(𝒜)[u\ v\ \alpha\ \beta]\in{\mathcal{D}}({\mathcal{A}}) and :×Z~{\mathcal{B}}:{\mathbb{R}}\times{\mathbb{R}}\mapsto\tilde{Z} is a bounded linear operator with [ab]=[0a(EIνxx)xxbρνρ 0 0]{\mathcal{B}}[a\ b]=[0\ \frac{-a(EI\nu_{xx})_{xx}-b\rho\nu}{\rho}\ 0\ 0]. The operator 𝒜{\mathcal{A}} generates a C0C_{0}-semigroup 𝕋{\mathbb{T}} on Z~\tilde{Z}, see [15, Section 5].

Recall z0z_{0} and ff and their properties from the statement of the proposition. Define z~0=z0[νf(0)νf˙(0) 0 0]\tilde{z}_{0}=z_{0}-[\nu f(0)\ \nu\dot{f}(0)\ 0\ 0]. Then z~0𝒟(𝒜)\tilde{z}_{0}\in{\mathcal{D}}({\mathcal{A}}). From [12, Chapter 4, Section 2] it follows that there exists a unique function z~C1([0,T];Z~)C([0,T];𝒟(𝒜))\tilde{z}\in C^{1}([0,T];\tilde{Z})\cap C([0,T];{\mathcal{D}}({\mathcal{A}})), given by the expression

z~(t)=𝕋tz~0+0t𝕋tτ[f(τ)f¨(τ)]d τ,\tilde{z}(t)={\mathbb{T}}_{t}\tilde{z}_{0}+\int_{0}^{t}{\mathbb{T}}_{t-\tau}{\mathcal{B}}\bigl{[}f(\tau)\ \ \ddot{f}(\tau)\bigr{]}{\rm d}\hbox{\hskip 0.5pt}\tau, (12)

such that z~(0)=z~0\tilde{z}(0)=\tilde{z}_{0} and z~\tilde{z} satisfies (11) for each t(0,T)t\in(0,T). Equivalently, there exists a unique function w~C([0,T];H4(0,L))C1([0,T];H2(0,L))C2([0,T];L2(0,L))\tilde{w}\in C([0,T];H^{4}(0,L))\cap C^{1}([0,T];H^{2}(0,L))\cap C^{2}([0,T];L^{2}(0,L)) with w~(0,),w~x(0,)C2([0,T];)\tilde{w}(0,\cdot),\tilde{w}_{x}(0,\cdot)\in C^{2}([0,T];{\mathbb{R}}) such that z~(t)=[w~(,t)w~t(,t)w~t(0,t)w~xt(0,t)]\tilde{z}(t)=[\tilde{w}(\cdot,t)\ \ \tilde{w}_{t}(\cdot,t)\ \ \tilde{w}_{t}(0,t)\ \ \tilde{w}_{xt}(0,t)] satisfies z~(0)=z~0\tilde{z}(0)=\tilde{z}_{0} and w~\tilde{w} and ff satisfy (7)-(10) for each t(0,T)t\in(0,T). Defining w=w~+νfw=\tilde{w}+\nu f, it follows that there exists a unique function wC([0,T];H4(0,L))C1([0,T];H2(0,L))C2([0,T];L2(0,L))w\in C([0,T];H^{4}(0,L))\cap C^{1}([0,T];H^{2}(0,L))\cap C^{2}([0,T];L^{2}(0,L)) with w(0,),wx(0,)C2([0,T];)w(0,\cdot),w_{x}(0,\cdot)\in C^{2}([0,T];{\mathbb{R}}) such that z(t)=[w(,t)wt(,t)wt(0,t)wxt(0,t)]z(t)=[w(\cdot,t)\ \ w_{t}(\cdot,t)\ \ w_{t}(0,t)\ \ w_{xt}(0,t)] satisfies z(0)=z0z(0)=z_{0} and ww and ff satisfy (1)-(4) for each t(0,T)t\in(0,T). This completes the proof of the proposition. ∎

It is easy to see using (12) that the unique classical solution zz in the above proof is given by the following formula: z(t)=𝕋tz~0+0t𝕋tτ[f(τ)f¨(τ)]d τ+[νf(t)νf˙(t) 0 0].z(t)={\mathbb{T}}_{t}\tilde{z}_{0}+\int_{0}^{t}{\mathbb{T}}_{t-\tau}{\mathcal{B}}\bigl{[}f(\tau)\ \ \ddot{f}(\tau)\bigr{]}{\rm d}\hbox{\hskip 0.5pt}\tau+[\nu f(t)\ \ \nu\dot{f}(t)\ \ 0\ \ 0]. The expression on the right side of this formula makes sense even for inputs fC2([0,T];)f\in C^{2}([0,T];{\mathbb{R}}) and initial states z0=[u0v0α0β0]Zz_{0}=[u_{0}\ v_{0}\ \alpha_{0}\ \beta_{0}]\in Z satisfying f(0)=u0(L)f(0)=u_{0}(L). Indeed it defines the unique strong solution of (1)-(4) for such inputs and initial states. But since the initial states and inputs of (1)-(4) considered in the rest of the paper satisfy the assumptions in Proposition II.2, we will not discuss strong solutions any further.

III Generating functions

For each s>0s>0, the Gevrey class Gs[0,T]G_{s}[0,T] is the space of all the functions yC[0,T]y\in C^{\infty}[0,T] which satisfy the estimate supt[0,T]|y(m)(t)|Dm+1(m!)s\sup_{t\in[0,T]}|y^{(m)}(t)|\leq D^{m+1}(m!)^{s} for some D>0D>0 and all integers m0m\geq 0. Here y(m)y^{(m)} denotes the mthm^{\rm th}-derivative of yy.

In this paper we solve a motion planning problem for the beam model (1)-(4) by building on the generating functions approach to flatness proposed for 1D parabolic PDEs in [8]. Accordingly we suppose that the solution ww of (1)-(4) on the interval [0,T][0,T] can be expressed as

w(x,t)=k0gk(x)y1(2k)(t)+k0hk(x)y2(2k)(t)w(x,t)=\sum_{k\geq 0}g_{k}(x)y_{1}^{(2k)}(t)+\sum_{k\geq 0}h_{k}(x)y_{2}^{(2k)}(t) (13)

for all x[0,L]x\in[0,L] and t[0,T]t\in[0,T]. Here gkg_{k} and hkh_{k} belonging to H4(0,L)H^{4}(0,L) are the generating functions (see Proposition III.1) and y1,y2Gs[0,T]y_{1},y_{2}\in G_{s}[0,T] with 1<s<21<s<2 are the flat outputs. We remark that while the solution of 1D parabolic PDEs can be expressed using a single flat output yy, see [8], we need two flat outputs y1,y2y_{1},y_{2} to express the solution of the beam model (1)-(4). These flat outputs cannot be chosen independently and must satisfy (35).

The generating functions gkg_{k} are obtained by solving a sequence of fourth-order linear ODEs recursively, on the interval x[0,L]x\in[0,L], as follows:

g0(x)=1,g_{0}(x)=1, (14)

g1g_{1} is obtained by solving the ODE

(EIg1,xx)xx(x)+ρ(x)g0(x)=0,\displaystyle\quad(EIg_{1,xx})_{xx}(x)+\rho(x)g_{0}(x)=0, (15)
g1(0)=0,g1,x(0)=0,\displaystyle\qquad g_{1}(0)=0,\qquad g_{1,x}(0)=0, (16)
g1,xx(0)=0,(EIg1,xx)x(0)=m,\displaystyle g_{1,xx}(0)=0,\qquad(EIg_{1,xx})_{x}(0)=-m, (17)

and gkg_{k} for k2k\geq 2 is obtained by solving the ODE

(EIgk,xx)xx(x)+ρ(x)gk1(x)=0,\displaystyle\quad(EIg_{k,xx})_{xx}(x)+\rho(x)g_{k-1}(x)=0, (18)
gk(0)=0,gk,x(0)=0,\displaystyle\qquad g_{k}(0)=0,\qquad g_{k,x}(0)=0, (19)
gk,xx(0)=0,(EIgk,xx)x(0)=0.\displaystyle\ \ g_{k,xx}(0)=0,\qquad(EIg_{k,xx})_{x}(0)=0. (20)

The generating functions hkh_{k} are also obtained by solving another sequence of fourth-order linear ODEs recursively, on the interval x[0,L]x\in[0,L], as follows:

h0(x)=x,h_{0}(x)=x, (21)

h1h_{1} is obtained by solving the ODE

(EIh1,xx)xx(x)+ρ(x)h0(x)=0,\displaystyle\ \ \quad(EIh_{1,xx})_{xx}(x)+\rho(x)h_{0}(x)=0, (22)
h1(0)=0,h1,x(0)=0,\displaystyle\ \ \qquad h_{1}(0)=0,\qquad h_{1,x}(0)=0, (23)
EI(0)h1,xx(0)=J,(EIh1,xx)x(0)=0,\displaystyle EI(0)h_{1,xx}(0)=J,\qquad(EIh_{1,xx})_{x}(0)=0, (24)

and hkh_{k} for k2k\geq 2 is obtained by solving the ODE

(EIhk,xx)xx(x)+ρ(x)hk1(x)=0,\displaystyle\quad(EIh_{k,xx})_{xx}(x)\!+\rho(x)h_{k-1}(x)=0, (25)
hk(0)=0,hk,x(0)=0,\displaystyle\qquad h_{k}(0)=0,\qquad h_{k,x}(0)=0, (26)
hk,xx(0)=0,(EIhk,xx)x(0)=0.\displaystyle h_{k,xx}(0)=0,\qquad(EIh_{k,xx})_{x}(0)=0. (27)

In the following proposition we show that the generating functions gkg_{k} and hkh_{k} belong to H4(0,L)H^{4}(0,L) and derive some estimates for them. Using these estimates, in Proposition III.2 we show that if y1y_{1} and y2y_{2} satisfy (35), then the function zC([0,T];Z)z\in C([0,T];Z) determined by the function ww in (13) via the expression (5) is the classical solution of (1)-(4) for the initial state z0=z(0)z_{0}=z(0) and input f(t)=w(L,t)f(t)=w(L,t).

Proposition III.1

For each k1k\geq 1 the generating functions gkg_{k} and hkh_{k} belong to H4(0,L)H^{4}(0,L) and there exist positive constants R1R_{1} and R2R_{2} independent of kk such that the following estimates hold for x[0,L]x\in[0,L]:

|gk(x)|R1kx4k1(4k1)!,|gk,x(x)|R1kx4k2(4k2)!,\displaystyle|g_{k}(x)|\leq\frac{R_{1}^{k}\,x^{4k-1}}{(4k-1)!},\qquad|g_{k,x}(x)|\leq\frac{R_{1}^{k}\,x^{4k-2}}{(4k-2)!}, (28)
|hk(x)|R2kx4k2(4k2)!,|hk,x(x)|R2kx4k3(4k3)!.\displaystyle|h_{k}(x)|\leq\frac{R_{2}^{k}\,x^{4k-2}}{(4k-2)!},\qquad|h_{k,x}(x)|\leq\frac{R_{2}^{k}\,x^{4k-3}}{(4k-3)!}. (29)
Proof:

​​Recall g0=1g_{0}\!=\!1 from (14). Solving (15)-(17) we get​​

g1(x)=\displaystyle g_{1}(x)= 0x0s10s20s3ρ(s4)EI(s2)d s4d s3d s2d s1\displaystyle-\int_{0}^{x}\!\int_{0}^{s_{1}}\!\int_{0}^{s_{2}}\!\int_{0}^{s_{3}}\!\frac{\rho(s_{4})}{EI(s_{2})}{\rm d}\hbox{\hskip 0.5pt}s_{4}{\rm d}\hbox{\hskip 0.5pt}s_{3}{\rm d}\hbox{\hskip 0.5pt}s_{2}{\rm d}\hbox{\hskip 0.5pt}s_{1}
m0x0s10s21EI(s2)d s3d s2d s1.\displaystyle\hskip 8.53581pt-m\int_{0}^{x}\int_{0}^{s_{1}}\int_{0}^{s_{2}}\frac{1}{EI(s_{2})}{\rm d}\hbox{\hskip 0.5pt}s_{3}{\rm d}\hbox{\hskip 0.5pt}s_{2}{\rm d}\hbox{\hskip 0.5pt}s_{1}. (30)

Since EIEI and ρ\rho are strictly positive functions in C4[0,L]C^{4}[0,L] it follows from the above equation that g1H4(0,L)g_{1}\in H^{4}(0,L) and the estimates in (28) hold for k=1k=1 with

R1=(maxx[0,L]ρ(x)+mminx[0,L]EI(x))max{1,L}.R_{1}=\Bigg{(}\frac{\max_{x\in[0,L]}\rho(x)+m}{\min_{x\in[0,L]}EI(x)}\Bigg{)}\max\{1,L\}. (31)

Solving (18)-(20) for gkg_{k} we get that for k2k\geq 2,

gk(x)=0x0s10s20s3ρ(s4)gk1(s4)EI(s2)d s4d s3d s2d s1.g_{k}(x)=-\int_{0}^{x}\!\int_{0}^{s_{1}}\!\!\int_{0}^{s_{2}}\!\!\int_{0}^{s_{3}}\!\frac{\rho(s_{4})g_{k-1}(s_{4})}{EI(s_{2})}{\rm d}\hbox{\hskip 0.5pt}s_{4}{\rm d}\hbox{\hskip 0.5pt}s_{3}{\rm d}\hbox{\hskip 0.5pt}s_{2}{\rm d}\hbox{\hskip 0.5pt}s_{1}. (32)

Let R1R_{1} be as defined in (31). Suppose that gkH4(0,L)g_{k}\in H^{4}(0,L) and the estimates in (28) hold for some k=nk=n with n1n\geq 1. We can then conclude that the same is true for k=n+1k=n+1 using (32) with k=n+1k=n+1, the estimates in (28) with k=nk=n and the fact that EI,ρC4[0,L]EI,\rho\in C^{4}[0,L] are strictly positive. So from the principle of mathematical induction it follows that gkH4(0,L)g_{k}\in H^{4}(0,L) and the estimates in (28) hold for all k1k\geq 1.

Next we will prove the estimates in (29). Recall from (21) that h0(x)=xh_{0}(x)=x. Solving (22)-(24) we get

h1(x)=\displaystyle h_{1}(x)= 0x0s10s20s3s4ρ(s4)EI(s2)d s4d s3d s2d s1\displaystyle-\int_{0}^{x}\!\int_{0}^{s_{1}}\!\int_{0}^{s_{2}}\!\int_{0}^{s_{3}}\!\frac{s_{4}\rho(s_{4})}{EI(s_{2})}{\rm d}\hbox{\hskip 0.5pt}s_{4}{\rm d}\hbox{\hskip 0.5pt}s_{3}{\rm d}\hbox{\hskip 0.5pt}s_{2}{\rm d}\hbox{\hskip 0.5pt}s_{1}
+J0x0s11EI(s2)d s2d s1.\displaystyle\hskip 14.22636pt+J\int_{0}^{x}\int_{0}^{s_{1}}\frac{1}{EI(s_{2})}{\rm d}\hbox{\hskip 0.5pt}s_{2}{\rm d}\hbox{\hskip 0.5pt}s_{1}. (33)

Since EIEI and ρ\rho are strictly positive functions in C4[0,L]C^{4}[0,L] it follows from the above equation that h1H4(0,L)h_{1}\in H^{4}(0,L) and the estimates in (29) hold for k=1k=1 with

R2=(maxx[0,L]ρ(x)+Jminx[0,L]EI(x))max{1,L3}.R_{2}=\Bigg{(}\frac{\max_{x\in[0,L]}\rho(x)+J}{\min_{x\in[0,L]}EI(x)}\Bigg{)}\max\{1,L^{3}\}.\vspace{-1mm} (34)

Solving (25)-(27) for hkh_{k} we get that for k2k\geq 2,

hk(x)=0x0s10s20s3ρ(s4)hk1(s4)EI(s2)d s4d s3d s2d s1.h_{k}(x)=-\int_{0}^{x}\!\int_{0}^{s_{1}}\!\!\int_{0}^{s_{2}}\!\!\int_{0}^{s_{3}}\!\!\frac{\rho(s_{4})h_{k-1}(s_{4})}{EI(s_{2})}{\rm d}\hbox{\hskip 0.5pt}s_{4}{\rm d}\hbox{\hskip 0.5pt}s_{3}{\rm d}\hbox{\hskip 0.5pt}s_{2}{\rm d}\hbox{\hskip 0.5pt}s_{1}.\vspace{-0.5mm}

The claim that hkH4(0,L)h_{k}\in H^{4}(0,L) and the estimates in (29) hold for all k1k\geq 1 with R2R_{2} given in (34) can be established by mimicking the induction argument given below (32). ∎

Proposition III.2

Fix T>0T>0 and s(1,2)s\in(1,2). Suppose that the functions y1,y2Gs[0,T]y_{1},y_{2}\in G_{s}[0,T] satisfy the following infinite order differential equation: for t[0,T]t\in[0,T],

k0gk,x(L)y1(2k)(t)+k0hk,x(L)y2(2k)(t)=0.\sum_{k\geq 0}g_{k,x}(L)y_{1}^{(2k)}(t)+\sum_{k\geq 0}h_{k,x}(L)y_{2}^{(2k)}(t)=0.\vspace{-1mm} (35)

Then ww given in (13) belongs to C([0,T];C4[0,L])C^{\infty}([0,T];C^{4}[0,L]) so that w(0,)w(0,\cdot), w(L,)w(L,\cdot) and wx(0,)w_{x}(0,\cdot) belong to C([0,T];)C^{\infty}([0,T];{\mathbb{R}}). Furthermore, the function zC([0,T];Z)z\in C([0,T];Z) determined by ww via (5) is the unique classical solution of (1)-(4) for the initial state z0=z(0)z_{0}=z(0) and input f(t)=w(L,t)f(t)=w(L,t).

Proof:

Differentiating the expression for gkg_{k} in (32) as required and then using the estimate for gk1g_{k-1} from (28) it follows that the functions gkg_{k}, gk,xg_{k,x}, gk,xxg_{k,xx}, gk,xxxg_{k,xxx} and gk,xxxxg_{k,xxxx} are uniformly bounded on [0,L][0,L] by Cgk/(4k5)!C_{g}^{k}/(4k-5)!. Here Cg>0C_{g}>0 is independent of kk. Similar estimates can be obtained for hkh_{k}, hk,xh_{k,x}, hk,xxh_{k,xx}, hk,xxxh_{k,xxx} and hk,xxxxh_{k,xxxx}. Using these estimates and the fact that y1,y2y_{1},y_{2} are in Gs[0,T]G_{s}[0,T] for some s(1,2)s\in(1,2) (which implies that y1(k)y_{1}^{(k)} and y2(k)y_{2}^{(k)} are uniformly bounded on [0,T][0,T] by Cyk+1(k!)sC_{y}^{k+1}(k!)^{s} for some Cy>0C_{y}>0 and all k0k\geq 0), it follows via the Weierstrass M-test and the ratio test that the series for ww and its derivatives with respect to xx and tt (obtained by termwise differentiation of the series in (13)) converge uniformly on [0,L]×[0,T][0,L]\times[0,T]. This implies that the derivatives of ww with respect to xx (up to four times) and tt (any number of times) are nothing but the series obtained by termwise differentiation of the series in (13) and ww has the regularity mentioned in the statement of this proposition.

Using (14), (15), (18) and (21), (22), (25) it is easy to check that the series corresponding to wttw_{tt} and (EIwxx)xx-(EIw_{xx})_{xx} are the same and hence ww satisfies the PDE (1). Taking x=0x=0 in the series for wttw_{tt}, (EIwxx)x(EIw_{xx})_{x}, wxttw_{xtt} and EIwxxEIw_{xx} and using the initial conditions for gkg_{k} in (16), (17), (19), (20) and the initial conditions for by hkh_{k} in (23), (24), (26), (27) it follows that ww satisfies the boundary conditions (2)-(3). Finally taking f(t)=w(L,t)f(t)=w(L,t) and using (35) (which means wx(L,t)=0w_{x}(L,t)=0) we get that ww satisfies the boundary condition (4). In summary, ww satisfies (1)-(4) and has the desired regularity so that zz given by (5) is a classical solution of (1)-(4) for the initial state z0=z(0)z_{0}=z(0) and input f(t)f(t). Recall the set VV from (6). Differentiating (35) with respect to tt we get wxt(L,t)=0w_{xt}(L,t)=0. Using this, (35) and the regularity of ww we can conclude that z0Vz_{0}\in V. Also ff is compatible with z0z_{0} by definition. The uniqueness of the classical solution zz now follows from Proposition II.2. ∎

IV Motion planning

We present our main results on the motion planning problem for the beam model (1)-(4) in this section, see Theorem IV.3 and Remark IV.4. In the following proposition, we first describe an approach for constructing functions y1y_{1} and y2y_{2} that satisfy (35). Below we will need the estimate

(a+b)!2a+ba!b!a,b.(a+b)!\leq 2^{a+b}a!b!{\hbox{$\hskip 31.29802pt\forall\;$}}a,b\in{\mathbb{N}}. (36)

This estimate follows from the fact that the (a+1)th(a+1)^{\rm th}-term in the binomial expansion of (1+1)a+b(1+1)^{a+b} is less than (1+1)a+b(1+1)^{a+b}.

Proposition IV.1

Fix s(1,2)s\in(1,2). Let the operators 1{\mathcal{L}}_{1} and 2{\mathcal{L}}_{2} be defined as follows: For pGs[0,T]p\in G_{s}[0,T],

1p=k0gk,x(L)p(2k),2p=k0hk,x(L)p(2k).{\mathcal{L}}_{1}p=\sum_{k\geq 0}g_{k,x}(L)p^{(2k)},\qquad{\mathcal{L}}_{2}p=\sum_{k\geq 0}h_{k,x}(L)p^{(2k)}.

Then y1=2py_{1}={\mathcal{L}}_{2}p and y2=1py_{2}=-{\mathcal{L}}_{1}p belong to Gs[0,T]G_{s}[0,T] for each pGs[0,T]p\in G_{s}[0,T] and satisfy (35).

Proof:

Let pGs[0,T]p\in G_{s}[0,T]. Then

supt[0,T]|p(k)(t)|Dk+1(k !)sk0\sup_{t\in[0,T]}|p^{(k)}(t)|\leq D^{k+1}(k{\hbox{\hskip 1.0pt}}!)^{s}{\hbox{$\hskip 31.29802pt\forall\;$}}k\geq 0 (37)

and some constant D>0D>0. Using this estimate and the estimate for hk,xh_{k,x} in (29) we can conclude by applying the Weierstrass M-test and the ratio test that for each n0n\geq 0 the series for y1(n)y_{1}^{(n)}, obtained by differentiating the series for 2p{\mathcal{L}}_{2}p termwise nn-times with respect to tt, converges uniformly on [0,T][0,T]. Hence y1(n)=2p(n)y_{1}^{(n)}={\mathcal{L}}_{2}p^{(n)} and using (37) and (29) we get

supt[0,T]|y1(n)(t)|k0R2kL4k3D2k+n+1((2k+n)!)s(4k3)!.\sup_{t\in[0,T]}|y_{1}^{(n)}(t)|\leq\sum_{k\geq 0}R_{2}^{k}L^{4k-3}D^{2k+n+1}\frac{((2k+n)!)^{s}}{(4k-3)!}.

Using (36) with a=2ka=2k and b=nb=n to bound (2k+n)!(2k+n)! in the above inequality we get supt[0,T]|y1(n)(t)|2nsDn(n!)sk0C0k+1((2k)!)s/(4k3)!\sup_{t\in[0,T]}|y_{1}^{(n)}(t)|\leq 2^{ns}D^{n}(n!)^{s}\sum_{k\geq 0}C_{0}^{k+1}((2k)!)^{s}\big{/}(4k-3)! for some C0>0C_{0}>0. Since s<2s<2, applying the ratio test it follows that the series k0C0k+1((2k)!)s/(4k3)!\sum_{k\geq 0}C_{0}^{k+1}((2k)!)^{s}\big{/}(4k-3)! converges and therefore supt[0,T]|y1(n)(t)|Cn+1(n!)s\sup_{t\in[0,T]}|y_{1}^{(n)}(t)|\leq C^{n+1}(n!)^{s} for a C>0C>0 and all n0n\geq 0, i.e. y1Gs[0,T]y_{1}\in G_{s}[0,T]. We can similarly show that y2Gs[0,T]y_{2}\in G_{s}[0,T].

Note that 12p=k0j0gk,x(L)hj,x(L)p(2k+2j){\mathcal{L}}_{1}{\mathcal{L}}_{2}p=\sum_{k\geq 0}\sum_{j\geq 0}g_{k,x}(L)h_{j,x}(L)p^{(2k+2j)} and 21p=j0k0gk,x(L)hj,x(L)p(2k+2j){\mathcal{L}}_{2}{\mathcal{L}}_{1}p=\sum_{j\geq 0}\sum_{k\geq 0}g_{k,x}(L)h_{j,x}(L)p^{(2k+2j)} for each pGs[0,T]p\in G_{s}[0,T]. So 12p{\mathcal{L}}_{1}{\mathcal{L}}_{2}p and 21p{\mathcal{L}}_{2}{\mathcal{L}}_{1}p are both double sums which only differ in the order of summation. We claim that the double sum l0j+k=lgj,x(L)hk,x(L)p(2l)(t)\sum_{l\geq 0}\sum_{j+k=l}g_{j,x}(L)h_{k,x}(L)p^{(2l)}(t) is absolutely convergent for each t[0,T]t\in[0,T]. Indeed, using (37), (28) and (29) we get

l0j+k=l|gj,x(L)hk,x(L)p(2l)(t)|\displaystyle\sum_{l\geq 0}\sum_{j+k=l}|g_{j,x}(L)h_{k,x}(L)p^{(2l)}(t)|
\displaystyle\leq l0j+k=lR1jR2kL4j2L4k3D2l+1((2l)!)s(4j2)!(4k3)!\displaystyle\sum_{l\geq 0}\sum_{j+k=l}\frac{R_{1}^{j}R_{2}^{k}L^{4j-2}L^{4k-3}D^{2l+1}((2l)!)^{s}}{(4j-2)!(4k-3)!}
\displaystyle\leq l0j+k=lRl+1((2l)!)s(4l5)!=l0Rl+1((2l)!)s(l+1)(4l5)!<.\displaystyle\sum_{l\geq 0}\sum_{j+k=l}\frac{R^{l+1}((2l)!)^{s}}{(4l-5)!}=\sum_{l\geq 0}\frac{R^{l+1}((2l)!)^{s}(l+1)}{(4l-5)!}<\infty.

Here R>0R>0 is some constant, the second inequality is derived by using the estimate in (36) with a=4j2a=4j-2 and b=4k3b=4k-3 and the last inequality is obtained by applying the ratio test. Therefore from Fubini’s theorem it follows that the double sum j,k0gj,x(L)hk,x(L)p(2l)\sum_{j,k\geq 0}g_{j,x}(L)h_{k,x}(L)p^{(2l)} is independent of the order of summation, i.e. 12p=21p{\mathcal{L}}_{1}{\mathcal{L}}_{2}p={\mathcal{L}}_{2}{\mathcal{L}}_{1}p. So y1=2py_{1}={\mathcal{L}}_{2}p and y2=1py_{2}=-{\mathcal{L}}_{1}p satisfy 1y12y2=0{\mathcal{L}}_{1}y_{1}-{\mathcal{L}}_{2}y_{2}=0 or equivalently (35). This completes the proof of the proposition. ∎

Remark IV.2

We can rewrite (13) concisely as

w(x,t)=[𝒲1y1](x,t)+[𝒲2y2](x,t)w(x,t)=[{\mathcal{W}}_{1}y_{1}](x,t)+[{\mathcal{W}}_{2}y_{2}](x,t) (38)

for x[0,L]x\in[0,L] and t[0,T]t\in[0,T]. Here the operators 𝒲1{\mathcal{W}}_{1} and 𝒲2{\mathcal{W}}_{2} are defined as follows: [𝒲1y](x,t)=k0gk(x)y(2k)(t)[{\mathcal{W}}_{1}y](x,t)=\sum_{k\geq 0}g_{k}(x)y^{(2k)}(t) and [𝒲2y](x,t)=k0hk(x)y(2k)(t)[{\mathcal{W}}_{2}y](x,t)=\sum_{k\geq 0}h_{k}(x)y^{(2k)}(t) for yGs[0,T]y\in G_{s}[0,T] with s(1,2)s\in(1,2). For any pGs[0,T]p\in G_{s}[0,T] with s(1,2)s\in(1,2), it follows from Proposition IV.1 that y1=2py_{1}={\mathcal{L}}_{2}p and y2=1py_{2}=-{\mathcal{L}}_{1}p are in Gs[0,T]G_{s}[0,T] and satisfy (35). Letting y1=2py_{1}={\mathcal{L}}_{2}p and y2=1py_{2}=-{\mathcal{L}}_{1}p in (38) and appealing to Proposition III.2 we can conclude that w=𝒲12p𝒲21pw={\mathcal{W}}_{1}{\mathcal{L}}_{2}p-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p is in C([0,T];C4[0,L])C^{\infty}([0,T];C^{4}[0,L]). Moreover, the function zC([0,T];Z)z\in C([0,T];Z) determined by ww via the expression z(t)=[w(,t)wt(,t)wt(0,t)wxt(0,t)]z(t)=[w(\cdot,t)\ w_{t}(\cdot,t)\ w_{t}(0,t)\ w_{xt}(0,t)] for all t[0,T]t\in[0,T] is the unique classical solution of (1)-(4) for the initial state z0=z(0)z_{0}=z(0) and input f(t)=w(L,t)f(t)=w(L,t). \square

Recall the operators 1{\mathcal{L}}_{1}, 2{\mathcal{L}}_{2}, 𝒲1{\mathcal{W}}_{1}, 𝒲2{\mathcal{W}}_{2} from Proposition IV.1 and Remark IV.2. In the next theorem, building on the results in Propositions III.1, III.2 and IV.1, we prove by construction the existence of a control input ff which transfers the beam model (1)-(4) between any two states belonging to a certain subspace of ZZ over a prescribed time-interval.

Theorem IV.3

Fix a time T>0T>0. Consider the set

M={[v(,0)vt(,0)vt(0,0)vxt(0,0)]Z |\displaystyle\quad M=\Big{\{}[v(\cdot,0)\ v_{t}(\cdot,0)\ v_{t}(0,0)\ v_{xt}(0,0)]\in Z{\hbox{\hskip 1.0pt}}\Big{|}{\hbox{\hskip 1.0pt}}
v=𝒲12p𝒲21pfor somepGs[0,T],s(1,2)}.\displaystyle v={\mathcal{W}}_{1}{\mathcal{L}}_{2}p-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p\ \textrm{for some}\ p\in G_{s}[0,T],\ s\in(1,2)\Big{\}}.

Let z0Mz_{0}\in M and zTMz_{T}\in M be given. Then there exists an fC([0,T];)f\in C([0,T];{\mathbb{R}}) and a unique classical solution zC([0,T];Z)z\in C([0,T];Z) of (1)-(4) on the time interval [0,T][0,T] for the initial state z0z_{0} and input ff such that z(T)=zTz(T)=z_{T}.

Proof:

From Remark IV.2 it is evident that MM is a well-defined and non-empty set. Since z0,zTMz_{0},z_{T}\in M and Gs1[0,T]Gs2[0,T]G_{s_{1}}[0,T]\subset G_{s_{2}}[0,T] for s1s2s_{1}\leq s_{2}, it follows from the definition of MM that there exist p0,pTGs[0,T]p_{0},p_{T}\in G_{s}[0,T] with s(1,2)s\in(1,2) such that v0=𝒲12p0𝒲21p0v_{0}={\mathcal{W}}_{1}{\mathcal{L}}_{2}p_{0}-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p_{0}, vT=𝒲12pT𝒲21pTv_{T}={\mathcal{W}}_{1}{\mathcal{L}}_{2}p_{T}-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p_{T}, z0=[v0(,0)v0,t(,0)v0,t(0,0)v0,tx(0,0)]z_{0}=[v_{0}(\cdot,0)\ v_{0,t}(\cdot,0)\ v_{0,t}(0,0)\ v_{0,tx}(0,0)] and zT=[vT(,0)vT,t(,0)vT,t(0,0)vT,tx(0,0)]z_{T}=[v_{T}(\cdot,0)\ v_{T,t}(\cdot,0)\ v_{T,t}(0,0)\ v_{T,tx}(0,0)].

For t[0,T]t\in[0,T] let

ψ(t)=1(0tψ0(τ)d τ/0Tψ0(τ)d τ),\psi(t)=1-\Big{(}\int_{0}^{t}\psi_{0}(\tau){\rm d}\hbox{\hskip 0.5pt}\tau\bigg{/}\int_{0}^{T}\psi_{0}(\tau){\rm d}\hbox{\hskip 0.5pt}\tau\Big{)}, (39)

where ψ0(t)=exp([(1tT)tT]1s1)\psi_{0}(t)=\exp\left(-\left[\left(1-\frac{t}{T}\right)\frac{t}{T}\right]^{-\frac{1}{s-1}}\right) for t(0,T)t\in(0,T) and ψ0(0)=ψ0(T)=0\psi_{0}(0)=\psi_{0}(T)=0. Then ψGs[0,T]\psi\in G_{s}[0,T] and

ψ(0)=1,ψ(T)=0,ψ(k)(0)=ψ(k)(T)=0\psi(0)=1,\quad\psi(T)=0,\quad\psi^{(k)}(0)=\psi^{(k)}(T)=0 (40)

for all k1k\geq 1, see [11]. For all t[0,T]t\in[0,T] define

p(t)=p0(t)ψ(t)+pT(Tt)ψ(Tt).p(t)=p_{0}(t)\psi(t)+p_{T}(T-t)\psi(T-t). (41)

Since p0,pT,ψGs[0,T]p_{0},p_{T},\psi\in G_{s}[0,T] and Gs[0,T]G_{s}[0,T] is closed under addition and multiplication of functions [14, Proposition 1.4.5] we get that pGs[0,T]p\in G_{s}[0,T]. Let w=𝒲12p𝒲21pw={\mathcal{W}}_{1}{\mathcal{L}}_{2}p-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p. Then z(t)=[w(,t)wt(,t)wxt(,t)wxt(,t)]C([0,T];Z)z(t)=[w(\cdot,t)\ w_{t}(\cdot,t)\ w_{xt}(\cdot,t)\ w_{xt}(\cdot,t)]\in C([0,T];Z) is the unique classical solution of (1)-(4) for the initial state z(0)z(0) and input f(t)=w(L,t)f(t)=w(L,t), see Remark IV.2. We will now complete the proof of this theorem by establishing that z(0)=z0z(0)=z_{0} and z(T)=zTz(T)=z_{T}.

The expression w=𝒲12p𝒲21pw={\mathcal{W}}_{1}{\mathcal{L}}_{2}p-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p is the same as (13) with y1=2py_{1}={\mathcal{L}}_{2}p and y2=1py_{2}=-{\mathcal{L}}_{1}p. The series for w(,t)w(\cdot,t), wt(,t)w_{t}(\cdot,t), wt(0,t)w_{t}(0,t) and wxt(0,t)w_{xt}(0,t) can be obtained by termwise differentiation of the series in (13) (see the proof of Proposition III.2) so that z(t)=[w(,t)wt(,t)wt(0,t)wxt(0,t)]z(t)=[w(\cdot,t)\ w_{t}(\cdot,t)\ w_{t}(0,t)\ w_{xt}(0,t)] can be written as k0Aky1(k)(t)+k0Bky2(k)(t)\sum_{k\geq 0}A_{k}y_{1}^{(k)}(t)+\sum_{k\geq 0}B_{k}y_{2}^{(k)}(t) for t[0,T]t\in[0,T]. Here Ak,BkH4(0,L)×H4(0,L)××A_{k},B_{k}\in H^{4}(0,L)\times H^{4}(0,L)\times{\mathbb{R}}\times{\mathbb{R}} for k0k\geq 0. Since y1(k)(t)=[2p(k)](t)y_{1}^{(k)}(t)=[{\mathcal{L}}_{2}p^{(k)}](t) and y2(k)(t)=[1p(k)](t)y_{2}^{(k)}(t)=-[{\mathcal{L}}_{1}p^{(k)}](t) (see the proof of Proposition IV.1), using the definition of the operators 1{\mathcal{L}}_{1} and 2{\mathcal{L}}_{2} we get for t[0,T]t\in[0,T],

z(t)=k0n0(Akhn,x(L)Bkgn,x(L))p(n+k)(t).z(t)=\sum_{k\geq 0}\sum_{n\geq 0}\bigl{(}A_{k}h_{n,x}(L)-B_{k}g_{n,x}(L)\bigr{)}p^{(n+k)}(t). (42)

Applying the above argument to v0=𝒲12p0𝒲21p0v_{0}={\mathcal{W}}_{1}{\mathcal{L}}_{2}p_{0}-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p_{0} and vT=𝒲12pT𝒲21pTv_{T}={\mathcal{W}}_{1}{\mathcal{L}}_{2}p_{T}-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p_{T} (instead of ww) we get that z0=k0n0(Akhn,x(L)Bkgn,x(L))p0(n+k)(0)z_{0}=\sum_{k\geq 0}\sum_{n\geq 0}\bigl{(}A_{k}h_{n,x}(L)-B_{k}g_{n,x}(L)\bigr{)}p_{0}^{(n+k)}(0), and zT=k0n0(Akhn,x(L)Bkgn,x(L))pT(n+k)(0)z_{T}=\sum_{k\geq 0}\sum_{n\geq 0}\bigl{(}A_{k}h_{n,x}(L)-B_{k}g_{n,x}(L)\bigr{)}p_{T}^{(n+k)}(0). Differentiating (41) nn-times and then using (40) we get p(n)(0)=p0(n)(0)p^{(n)}(0)=p_{0}^{(n)}(0) and p(n)(T)=pT(n)(0)p^{(n)}(T)=p_{T}^{(n)}(0) for all n0n\geq 0. It now follows from the expressions for z(0)z(0) and z(T)z(T) from (42) and the expressions for z0z_{0} and zTz_{T} given above that z(0)=z0z(0)=z_{0} and z(T)=zTz(T)=z_{T}. This completes the proof.∎

Remark IV.4

Let cc\in{\mathbb{R}}. For the initial state [uss 0 0 0]Z[u_{ss}\ 0\ 0\ 0]\in Z with uss(x)=cu_{ss}(x)=c for all x[0,L]x\in[0,L] and the constant input fssC([0,T];)f_{ss}\in C([0,T];{\mathbb{R}}) with fss(t)=cf_{ss}(t)=c for all t[0,T]t\in[0,T], note that the constant function zss(t)=[uss 0 0 0]z_{ss}(t)=[u_{ss}\ 0\ 0\ 0] for t[0,T]t\in[0,T] is the classical solution of (1)-(4), see Definition II.1. We call zss(0)z_{ss}(0) the steady-state of (1)-(4) corresponding to the constant input fssf_{ss}. Each such steady-state is a rest configuration of the beam corresponding to some fixed position of the cantilever joint.

Let v=𝒲12p𝒲21pv={\mathcal{W}}_{1}{\mathcal{L}}_{2}p-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p with p(t)=cp(t)=c for all t[0,T]t\in[0,T] and some cc\in{\mathbb{R}}. From the definition of the operators 1{\mathcal{L}}_{1}, 2{\mathcal{L}}_{2}, 𝒲1{\mathcal{W}}_{1}, 𝒲2{\mathcal{W}}_{2} it follows that [v(,0)vt(,0)vt(0,0)vxt(0,0)]=[uss 0 0 0][v(\cdot,0)\ v_{t}(\cdot,0)\ v_{t}(0,0)\ v_{xt}(0,0)]=[u_{ss}\ 0\ 0\ 0], where uss(x)=cu_{ss}(x)=c for all x[0,L]x\in[0,L]. In other words, steady-states of (1)-(4) belong to the set MM. It now follows from Theorem IV.3 that we can find an input ff to transfer (1)-(4) between any two steady-states. \square

V Numerical and experimental results

Our experimental setup consists of a non-uniform moving cantilever beam, with linearly-varying width, made of stainless steel. One end of the beam supports a tip-mass, while the other end is attached via a cantilever joint to a cart mounted on a Hiwin single axis robot, see Figure 1. The robot is driven by a Yasaka AC servo motor which fixes the cart position as per the control input it receives from a Raspberry Pi 4B microprocessor. The dynamics of the beam is described by the model (1)-(4) with L=0.5 mL=0.5{\hbox{\hskip 1.0pt}}\textrm{m}, m=0.402 kgm=0.402{\hbox{\hskip 1.0pt}}\textrm{kg}, J=1.9×104 kg m2J=1.9\times 10^{-4}{\hbox{\hskip 1.0pt}}\textrm{kg{\hbox{\hskip 1.0pt}}m}^{2} and ρ(x)=0.11(1+3x) kg/m\rho(x)=0.11(1+3x){\hbox{\hskip 1.0pt}}\textrm{kg/m} and EI(x)=0.297(1+3x) N/mEI(x)=0.297(1+3x){\hbox{\hskip 1.0pt}}\textrm{N/m} for x[0,L]x\in[0,L].

We consider two problems to illustrate our solution to the motion planning problem presented in Theorem IV.3. In both the problems we take the time of transfer to be T=3 sT=3{\hbox{\hskip 1.0pt}}\textrm{s} and the final state to be zT=0z_{T}=0. In Problem 1 we take the initial state to be z0=[v0(,0)v0,t(,0)v0,t(0,0)v0,xt(0,0)]z_{0}=[v_{0}(\cdot,0)\ v_{0,t}(\cdot,0)\ v_{0,t}(0,0)\ v_{0,xt}(0,0)], where v0=𝒲12p0𝒲21p0v_{0}={\mathcal{W}}_{1}{\mathcal{L}}_{2}p_{0}-{\mathcal{W}}_{2}{\mathcal{L}}_{1}p_{0} with p0(t)=1+10t2e2tp_{0}(t)=1+10t^{2}e^{-2t} for t0t\geq 0. Note that z0Mz_{0}\in M is not a steady-state of (1)-(4). In Problem 2 we take the initial state to be the steady-state of (1)-(4) given by z0=[0.4 0 0 0]z_{0}=[0.4\ 0\ 0\ 0], which is obtained from the above expressions for z0z_{0} and v0v_{0} by taking p0(t)=0.4p_{0}(t)=0.4 for all t0t\geq 0, see Remark IV.4. We solve both the problems using the procedure described in the proof of Theorem IV.3. Accordingly we choose ψ\psi to be the function in (39) with s=1.5s=1.5 and T=3T=3 and define pp via (41) by taking pT=0p_{T}=0 (since zT=0z_{T}=0). The required control input is f(t)=w(L,t)f(t)=w(L,t), where w=[𝒲12𝒲21]pw=[{\mathcal{W}}_{1}{\mathcal{L}}_{2}-{\mathcal{W}}_{2}{\mathcal{L}}_{1}]p. Using the expressions for 1{\mathcal{L}}_{1}, 2{\mathcal{L}}_{2}, 𝒲1{\mathcal{W}}_{1}, 𝒲2{\mathcal{W}}_{2} and changing the order of the double summations we have f(t)=limNl=0Nj+k=l[gk(L)hj,x(L)+hk(L)gj,x(L)]p(2l)(t)f(t)=\lim_{N\to\infty}\sum_{l=0}^{N}\sum_{j+k=l}\big{[}g_{k}(L)h_{j,x}(L)+h_{k}(L)g_{j,x}(L)\big{]}p^{(2l)}(t). (Note that gkg_{k} and hkh_{k} are computed using the expressions in the proof of Proposition III.1.) This series converges rapidly and by truncating it with N=20N=20 we compute a very good approximation for the inputs which solve the two problems being considered, see Figure 2.

[Uncaptioned image]

Figure 2. Plot of the inputs f1f_{1} and f2f_{2} which solve Problem 1 and Problem 2, respectively.

[Uncaptioned image]

Figure 3. Displacement profile w(x,t)w(x,t) of the beam model (1)-(4) for the initial state and the input in Problem 1. The displacement profile starts from a non-constant (in xx) function and settles down to the zero function in 3 seconds as expected.

We discretized the spatial derivatives in the beam model (1)-(4) using the finite-difference method (with step-size 1/300) to obtain a set of ODEs which serve as a numerical model for the beam model. We validated our solution for the motion planning problems, Problems 1 and 2, presented above by simulating the numerical model with the appropriate initial states z0z_{0} and the control inputs shown in Figure 2. Figure 3 shows the beam displacement profile w(x,t)w(x,t) obtained from our simulation for Problem 1. As expected (recall zT=0z_{T}=0), the displacement profile settles down to the zero function within 3 seconds. Figure 4 shows the displacement trajectory w(0,t)w(0,t) of the tip-mass obtained from our simulation for Problem 2. As expected the tip-mass is initially at rest with w(0,0)=0.4w(0,0)=0.4 and again at rest finally with w(0,3)=0w(0,3)=0. We have implemented the control input in Problem 2 on our experimental setup in Figure 1 and observed that the beam is transferred from one steady-state to another (at a distance of 0.4m) within 3 seconds. We recorded the experiment using a camera at 240fps frame rate; The video of the experiment is available here: https://youtu.be/2IvgK5pK7Og. By tracking the position of a green marker placed on the tip-mass using hue-based segregation and contour detection algorithms, we extracted the trajectory of the tip-mass from the video, see Figure 4.

[Uncaptioned image]

Figure 4. Tip-mass position w(0,t)w(0,t) obtained from the simulation and experiment for Problem 2 match closely.

References

  • [1] Y. Aoustin, M. Fliess, H. Mounier, P. Rouchon and J. Rudolph, “Theory and practice in the motion planning and control of a flexible robot arm using Mikusiński operators,” Proc. 5th IFAC Symp. Robot Control, pp. 267-273, Sep. 3-5, 1997, Nantes, France.
  • [2] M. Bachmayer, H. Ulbrich and J. Rudolph, “Flatness-based control of a horizontally moving erected beam with a point mass,” Math. Comput. Model. Dyn. Syst., vol. 17, pp. 49-69, 2011.
  • [3] A. Badkoubeh and G. Zhu, “A Green’s function-based design for deformation control of a microbeam with in-domain actuation,” J. Dyn. Syst. Meas. Control, vol. 136, 2014.
  • [4] A. Badkoubeh, J. Zheng and G. Zhu, “Flatness-based deformation control of an Euler-Bernoulli beam with in-domain actuation,” IET Control Theory Appl., vol. 10, pp. 2110-2118, 2016.
  • [5] M. Barczyk and A. F. Lynch, “Flatness-based estimated state feedback control for a rotating flexible beam: Experimental results,” IET Control Theory Appl., vol. 2, 288-302, 2008.
  • [6] B. Kolar, N. Gehring and M. Schöberl, “On the calculation of differential parameterizations for the feedforward control of an Euler-Bernoulli beam,” In Dynamics and Control of Advanced Structures and Machines: Contributions from the 4th International Workshop, Linz, pp. 123-136, Springer, 2022.
  • [7] A. F. Lynch and D. Wang, “Flatness-based control of a flexible beam in a gravitional field,” Proc. 2004 Amer. Control Conf., pp. 5449-5454, Jun. 30 - Jul. 2, 2004, Boston, USA.
  • [8] P. Martin, L. Rosier and P. Rouchon, “Null controllability of one-dimensional parabolic equations by the flatness approach,” SIAM J. Control Optim., vol. 54, pp. 198-220, 2016.
  • [9] T. Meurer, Control of Higher-Dimensional PDEs, Springer-Verlag, Berlin, 2013.
  • [10] T. Meurer, D. Thull, and A. Kugi, “Flatness based tracking control of a piezoactuated Euler-Bernoulli beam with non-collocated feedback: theory and experiments,” Int. J. Control, vol. 81, pp. 475-493, 2008.
  • [11] T. Meurer, J. Schröck, and A. Kugi, “Motion planning for a damped Euler-Bernoulli Beam,” Proc. 49th49{\rm th} IEEE Conf. Decision & Control, pp. 2566-2571, Dec. 15-17, 2010, Atlanta, USA.
  • [12] A. Pazy, Semigroups of Linear Operators and Applications to Partial Differential Equations, Springer-Verlag, New York, 1983.
  • [13] G. G. Rigatos and S. G. Tzafestas, “Flatness-based and energy-based control for distributed parameter robotic systems,” 13th13^{\rm th} IFAC Symp. Information Control Problems in Manufacturing, pp. 1847-1852, Jun. 3-5, 2009, Moscow, Russia.
  • [14] L. Rodino, Linear partial differential operators in Gevrey spaces, World Scientific Publishing, Singpore, 1993.
  • [15] X. Zhao and G. Weiss, “Controllability and observability of a well-posed system coupled with a finite-dimensional system,” IEEE Trans. Autom. Control, vol. 56, pp. 88-99, 2011.