Flatness-based motion planning for a non-uniform moving cantilever Euler-Bernoulli beam with a tip-mass
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:
(1) | |||
(2) | |||
(3) | |||
(4) |
Here is the length of the beam, is the displacement of the beam at the location and time , and are the mass per unit length and flexural rigidity, respectively, of the beam at , and and are the mass and moment of inertia, respectively, of the tip-mass located at the end of the beam. We suppose that , and they are strictly positive, i.e and . The displacement of the cantilever joint at the end of the beam is determined by the scalar control input . 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.
![]() |
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 for the non-uniform moving cantilever Euler-Bernoulli beam model (1)-(4) in which it has a unique state trajectory for every initial state and each smooth control input which is compatible with , see Section II. In this paper, we study the motion planning problem of computing a control input which transfers (1)-(4) from an initial state to a final state over a prescribed time-interval . We solve this problem for a set of initial and final states in , 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 denote the usual Sobolev space of order on the interval . A natural choice of state space for (1)-(4) is the Hilbert space
For and in , the inner product . We now define the notion of classical solutions for (1)-(4) in .
Definition II.1
Consider the following dense subspace of :
(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 and the input 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
Proof:
Choose such that and . Replacing with formally in (1)-(4) we get that for and ,
(7) | |||
(8) | |||
(9) | |||
(10) |
The subspace is a closed subset of and is itself a Hilbert space. The coupled PDE-ODE system (7)-(10) can be written as an abstract evolution equation on as follows:
(11) |
The operators and are defined as follows: The domain of is and for and is a bounded linear operator with . The operator generates a -semigroup on , see [15, Section 5].
Recall and and their properties from the statement of the proposition. Define . Then . From [12, Chapter 4, Section 2] it follows that there exists a unique function , given by the expression
(12) |
such that and satisfies (11) for each . Equivalently, there exists a unique function with such that satisfies and and satisfy (7)-(10) for each . Defining , it follows that there exists a unique function with such that satisfies and and satisfy (1)-(4) for each . This completes the proof of the proposition. ∎
It is easy to see using (12) that the unique classical solution in the above proof is given by the following formula: The expression on the right side of this formula makes sense even for inputs and initial states satisfying . 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 , the Gevrey class is the space of all the functions which satisfy the estimate for some and all integers . Here denotes the -derivative of .
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 of (1)-(4) on the interval can be expressed as
(13) |
for all and . Here and belonging to are the generating functions (see Proposition III.1) and with are the flat outputs. We remark that while the solution of 1D parabolic PDEs can be expressed using a single flat output , see [8], we need two flat outputs to express the solution of the beam model (1)-(4). These flat outputs cannot be chosen independently and must satisfy (35).
The generating functions are obtained by solving a sequence of fourth-order linear ODEs recursively, on the interval , as follows:
(14) |
is obtained by solving the ODE
(15) | |||
(16) | |||
(17) |
and for is obtained by solving the ODE
(18) | |||
(19) | |||
(20) |
The generating functions are also obtained by solving another sequence of fourth-order linear ODEs recursively, on the interval , as follows:
(21) |
is obtained by solving the ODE
(22) | |||
(23) | |||
(24) |
and for is obtained by solving the ODE
(25) | |||
(26) | |||
(27) |
In the following proposition we show that the generating functions and belong to and derive some estimates for them. Using these estimates, in Proposition III.2 we show that if and satisfy (35), then the function determined by the function in (13) via the expression (5) is the classical solution of (1)-(4) for the initial state and input .
Proposition III.1
For each the generating functions and belong to and there exist positive constants and independent of such that the following estimates hold for :
(28) | |||
(29) |
Proof:
Recall from (14). Solving (15)-(17) we get
(30) |
Since and are strictly positive functions in it follows from the above equation that and the estimates in (28) hold for with
(31) |
Solving (18)-(20) for we get that for ,
(32) |
Let be as defined in (31). Suppose that and the estimates in (28) hold for some with . We can then conclude that the same is true for using (32) with , the estimates in (28) with and the fact that are strictly positive. So from the principle of mathematical induction it follows that and the estimates in (28) hold for all .
Next we will prove the estimates in (29). Recall from (21) that . Solving (22)-(24) we get
(33) |
Since and are strictly positive functions in it follows from the above equation that and the estimates in (29) hold for with
(34) |
Solving (25)-(27) for we get that for ,
The claim that and the estimates in (29) hold for all with given in (34) can be established by mimicking the induction argument given below (32). ∎
Proposition III.2
Proof:
Differentiating the expression for in (32) as required and then using the estimate for from (28) it follows that the functions , , , and are uniformly bounded on by . Here is independent of . Similar estimates can be obtained for , , , and . Using these estimates and the fact that are in for some (which implies that and are uniformly bounded on by for some and all ), it follows via the Weierstrass M-test and the ratio test that the series for and its derivatives with respect to and (obtained by termwise differentiation of the series in (13)) converge uniformly on . This implies that the derivatives of with respect to (up to four times) and (any number of times) are nothing but the series obtained by termwise differentiation of the series in (13) and 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 and are the same and hence satisfies the PDE (1). Taking in the series for , , and and using the initial conditions for in (16), (17), (19), (20) and the initial conditions for by in (23), (24), (26), (27) it follows that satisfies the boundary conditions (2)-(3). Finally taking and using (35) (which means ) we get that satisfies the boundary condition (4). In summary, satisfies (1)-(4) and has the desired regularity so that given by (5) is a classical solution of (1)-(4) for the initial state and input . Recall the set from (6). Differentiating (35) with respect to we get . Using this, (35) and the regularity of we can conclude that . Also is compatible with by definition. The uniqueness of the classical solution 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 and that satisfy (35). Below we will need the estimate
(36) |
This estimate follows from the fact that the -term in the binomial expansion of is less than .
Proposition IV.1
Fix . Let the operators and be defined as follows: For ,
Then and belong to for each and satisfy (35).
Proof:
Let . Then
(37) |
and some constant . Using this estimate and the estimate for in (29) we can conclude by applying the Weierstrass M-test and the ratio test that for each the series for , obtained by differentiating the series for termwise -times with respect to , converges uniformly on . Hence and using (37) and (29) we get
Using (36) with and to bound in the above inequality we get for some . Since , applying the ratio test it follows that the series converges and therefore for a and all , i.e. . We can similarly show that .
Note that and for each . So and are both double sums which only differ in the order of summation. We claim that the double sum is absolutely convergent for each . Indeed, using (37), (28) and (29) we get
Here is some constant, the second inequality is derived by using the estimate in (36) with and and the last inequality is obtained by applying the ratio test. Therefore from Fubini’s theorem it follows that the double sum is independent of the order of summation, i.e. . So and satisfy or equivalently (35). This completes the proof of the proposition. ∎
Remark IV.2
We can rewrite (13) concisely as
(38) |
for and . Here the operators and are defined as follows: and for with . For any with , it follows from Proposition IV.1 that and are in and satisfy (35). Letting and in (38) and appealing to Proposition III.2 we can conclude that is in . Moreover, the function determined by via the expression for all is the unique classical solution of (1)-(4) for the initial state and input .
Recall the operators , , , 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 which transfers the beam model (1)-(4) between any two states belonging to a certain subspace of over a prescribed time-interval.
Theorem IV.3
Proof:
From Remark IV.2 it is evident that is a well-defined and non-empty set. Since and for , it follows from the definition of that there exist with such that , , and .
For let
(39) |
where for and . Then and
(40) |
for all , see [11]. For all define
(41) |
Since and is closed under addition and multiplication of functions [14, Proposition 1.4.5] we get that . Let . Then is the unique classical solution of (1)-(4) for the initial state and input , see Remark IV.2. We will now complete the proof of this theorem by establishing that and .
The expression is the same as (13) with and . The series for , , and can be obtained by termwise differentiation of the series in (13) (see the proof of Proposition III.2) so that can be written as for . Here for . Since and (see the proof of Proposition IV.1), using the definition of the operators and we get for ,
(42) |
Applying the above argument to and (instead of ) we get that , and . Differentiating (41) -times and then using (40) we get and for all . It now follows from the expressions for and from (42) and the expressions for and given above that and . This completes the proof.∎
Remark IV.4
Let . For the initial state with for all and the constant input with for all , note that the constant function for is the classical solution of (1)-(4), see Definition II.1. We call the steady-state of (1)-(4) corresponding to the constant input . Each such steady-state is a rest configuration of the beam corresponding to some fixed position of the cantilever joint.
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 , , and and for .
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 and the final state to be . In Problem 1 we take the initial state to be , where with for . Note that 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 , which is obtained from the above expressions for and by taking for all , see Remark IV.4. We solve both the problems using the procedure described in the proof of Theorem IV.3. Accordingly we choose to be the function in (39) with and and define via (41) by taking (since ). The required control input is , where . Using the expressions for , , , and changing the order of the double summations we have . (Note that and are computed using the expressions in the proof of Proposition III.1.) This series converges rapidly and by truncating it with we compute a very good approximation for the inputs which solve the two problems being considered, see Figure 2.
![]() |
Figure 2. Plot of the inputs and which solve Problem 1 and Problem 2, respectively.
![]() |
Figure 3. Displacement profile 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 ) 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 and the control inputs shown in Figure 2. Figure 3 shows the beam displacement profile obtained from our simulation for Problem 1. As expected (recall ), the displacement profile settles down to the zero function within 3 seconds. Figure 4 shows the displacement trajectory of the tip-mass obtained from our simulation for Problem 2. As expected the tip-mass is initially at rest with and again at rest finally with . 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 0.4 m) within 3 seconds. We recorded the experiment using a camera at 240 fps 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.
![]() |
Figure 4. Tip-mass position 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. 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,” 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.