9
Transfer Matrix Method for Controlled Multibody Systems

9.1 Introduction

Research on the dynamics of multibody systems began more than 40 years ago and international commercial software based on a great many theories and methods for multibody systems has been developed. The transfer matrix method for multibody systems (MSTMM) does not need global dynamic equations of the system and has a lot of advantages, such as easy modeling, high programming stylization and high computational efficiency. In principle, the MSTMM can be combined with any other mechanics method, including all kinds of multibody system dynamics methods, the finite element method (FEM), the classic mechanics method and the analytical mechanics method, so that the advantages of different methods can be exploited. The MSTMM can be combined with all kinds of software to improve its computational speed. Precise performance prediction and effective controls of the multibody system are very important in control engineering problems, especially in fields such as the firing control of the launch system of high‐performance weapons, spatial structure control and robots. Because the problems are complex and the computation speed for the multi‐rigid‐flexible‐body system (MRFS) is too slow to satisfy the required response time for quick control, the dynamic models of multibody systems have to be simplified when dealing with those problems, and the model error needs to be compensated for through complex control design. The excessive simplification of the dynamic model usually results in difficulty when predicting the dynamic law of the system and the redundant high‐frequency behavior of the system, and therefore the control precision of the system is deeply affected. For example, when the firing control system of the launch system of a modern weapon is designed, the launch system is modeled as a simple multi‐rigid‐body system, where the deformation vibration of the system is not considered, resulting in poor control precision and low sensitivity. The best way to solve this problem is to introduce the dynamic method of a controlled multibody system which describes the dynamic of an MRFS perfectly and also satisfies the demand for rapid computation speed for real‐time control. The modeling of the multibody system and the controller are established synchronously and completely.

In this chapter, the following methods developed by the authors will be introduced: the mixed transfer matrix method for multibody systems and other multibody system dynamic methods [71, 72], the mixed MSTMM and the FEM [76], the finite segment discrete time MSTMM [81], the linear transfer matrix method (TMM) for controlled multibody systems [77], the discrete time TMM for controlled multibody systems [78] and the Riccati discrete time MSTMM [78, 80]. These methods make it easier to study the dynamics of multi‐dimensional systems and complex structures by using the MSTMM.

9.2 Mixed Transfer Matrix Method for Multibody Systems

In this section, the mixed TMM for multibody systems and other multibody system dynamic methods [71, 72] developed by the authors is introduced. In this method, any multibody system can be decomposed into subsystems, and the connection relation among the subsystems can be regarded as the boundary of each subsystem, as if the other subsystems do not exist. We usually establish the “global” dynamic equations using the ordinary method of multibody system dynamics for some subsystems and the “overall” transfer equation using the MSTMM for other subsystems. The unknown state variables on these “boundary subsystems” are considered to be external forces in the subsystem handled by the ordinary dynamic method and the internal forces in the subsystem handled by the MSTMM. Combining the global dynamic equations (usually hybrid differential‐algebra equations) with the overall transfer equations (algebra equations), the corresponding equations of the overall system can be assembled. Once all the global dynamic equations and the overall transfer equations of an MRFS are solved, the time history of the system dynamics can be obtained. It is an advantage of the method that each subsystem can adopt an optimum mathematic model and software.

The main steps for a numerical algorithm that can be used to solve MRFSs using the mixed MSTMM and other multibody system dynamics methods can be summarized as follows:

  1. Give the system parameters and the initial state parameters etc.
  2. Set the simulation step sequence number c9-i0001 (discrete time step).
  3. According to the angular coordinates of a rigid body and their first‐ and second‐order derivatives with respect to time at time c9-i0002, calculate the coefficients for the step‐by‐step time integration method at time ti and estimate the angular coordinates of a rigid body at time ti. Take this as the initial value of the next iterative procedure.
  4. Set the iterative variable c9-i0003.
  5. Regard the motion state of a rigid body as the boundary condition of a flexible body and compute the dynamic response of a flexible body using the multibody system dynamic method.
  6. Set the iterative variable for the angular coordinates of the rigid body c9-i0004.
  7. Compute the transfer matrices of each body element and hinge element.
  8. Assemble the overall transfer equations according to the structure of the system.
  9. Regard the boundary conditions of the system and the motion state and force state of the subsystem boundary as the boundary conditions of the overall transfer equations. Solve the overall transfer equations to obtain the state variable of each element at time ti.
  10. Judge the convergence criterion. If the square sum of the errors of all angular coordinates of rigid bodies from the kth to the (c9-i0005)th iterations is bigger than the required supposed precision value or the iterative loop does not reach the given iteration number, set c9-i0006 and return to (7).
  11. Compute the end‐point velocity, the acceleration of each element and the first‐ and second‐order derivatives with respect to time of the rigid‐body angular coordinates using the step‐by‐step time integration method.
  12. Let c9-i0007. If j does not reach the required times, return to (5).
  13. If the required simulation time has been reached, exit the program; if not, let c9-i0008 and return to (3).

The following two examples are given to validate the proposed method.

Image described by caption and surrounding text.

Figure 9.1 MRFS model with a reticulated particle–spring system.

Reticulated lumped mass–spring system with five rows and columns of circles labeled m1,1; m2,1; m3,1; m4,1; m5,1; m1,2; m2,2; m3,2; m4,2; m5,2; m5,3; m4,3; m3,3; etc. interconnected by springs (zigzag lines).

Figure 9.2 Reticulated lumped mass–spring system.

Force analysis of a lumped mass, represented by a circle labeled ri,j with 4 springs labeled K(i–1,i),j; Ki,(j,j+1); K(i,i+1),j; and Ki,(j–1,j) and an outward arrow labeled mi,j g.

Figure 9.3 Force analysis of a lumped mass.

image
image

Figure 9.4 Dynamic simulation of a multibody system with a reticulated particle–spring system: (a) rotation angle of rigid body 1, (b) rotation angle of rigid body 5, (c) rotation angle of rigid body 3, (d) rotation angle of rigid body 4, (e) coordinates of particle 2, (f) coordinates of particle 17, (g) coordinates of particle 24 and (h) total mechanical energy of the system.

image
image

Figure 9.5 Dynamic simulation of a multibody system with a plate: (a) rotation angle of rigid body 1, (b) rotation angle of rigid body 5, (c) rotation angle of rigid body 3, (d) rotation angle of rigid body 4, (e) coordinates of particle 2, (f) coordinates of particle 17, (g) coordinates of particle 24 and (h) total mechanical energy of the system.

Equivalent flexible plate with five rows and columns of circles interconnected by springs (zigzag lines).

Figure 9.6 Equivalent flexible plate.

Image described by caption and surrounding text.

Figure 9.7 The positions of the system at 0.5 s in Examples 9.1 and 9.2: (a) system in Example 9.1 and (b) system in Example 9.2.

Image described by caption and surrounding text.

Figure 9.8 MRFS model.

Graphs of ϑ2/rad and ϑ6/rad vs. t (top), coordinates of 2 and 3 vs. t (middle), and coordinates of 4 and 5 vs. t (bottom), each with 2 sets of 2 coinciding waves representing Newton–Euler method and the mixed method.

Figure 9.9 Time history of the dynamics of an MRFS: (a) time history of the rotations of rigid body 2 and rigid body 6, (b) coordinates of particle 2 and particle 3, and (c) coordinates of particle 4 and particle 5.

9.3 Finite Element Transfer Matrix Method for Multibody Systems

The dynamics of linear multibody systems are solved by combining the MSTMM and the FEM, and the FEM is “reconstructed” by the MSTMM, or the MSTMM is “reconstructed” by the FEM. The mixed MSTMM and the FEM [76] introduced in this section is different from the “reconstructed method” mentioned above. In the proposed method, the two independent methods are applied in different subsystems of the same multibody system, that is, the multibody system is divided into subsystems and each of these subsystems could be readily modeled by the MSTMM or the FEM. The overall transfer equation is obtained by the MSTMM, and the global dynamic equations of other two‐ or three‐dimensional complex subsystems are obtained by the FEM. The subsequent processing method is the same as in section 9.2.

Image described by caption and surrounding text.

Figure 9.10 Chain MRFS.

A horizontal bar divided into 5 segments with labels 0–5, with boxes labeled 1, 2, 3, and 4 below the second, third, fourth, and fifth bars from the left. Below the boxes is another bar with a 2-headed arrow labeled l.

Figure 9.11 The element partition of beam 1.

Graph of y/m vs. t/s displaying 2 coinciding waves representing the mixed method (2 horizontal lines) and Newton–Euler method (“x” marker), illustrating the transverse displacement of the connected point.

Figure 9.12 The transverse displacement of the connected point.

Graph of y/m vs. t/s displaying 2 coinciding waves representing the mixed method (2 horizontal lines) and Newton–Euler method (“x” marker), illustrating the transverse displacements of node 1 of the beam.

Figure 9.13 The transverse displacements of node 1 of the beam.

Graph of y/m vs. t/s illustrating the node 2 of the beam depicting transverse dashed waves representing the mixed method and "X" marks for Newton–Euler method.

Figure 9.14 The transverse displacements of node 2 of the beam.

Graph of y/m vs. t/s illustrating the node 3 of the beam depicting transverse dashed waves representing the mixed method and "X" marks for Newton–Euler method.

Figure 9.15 The transverse displacements of node 3 of the beam.

Graph of θ3/rad vs. t/s of the azimuth angle of rigid body 3 depicting transverse dashed waves representing the mixed method and "X" marks for Newton–Euler method.

Figure 9.16 Time history of the azimuth angle of rigid body 3.

Graph of θ5/rad vs. t/s of the azimuth angle of rigid body 5 depicting transverse dashed waves representing the mixed method and "X" marks for Newton–Euler method.

Figure 9.17 Time history of the azimuth angle of rigid body 5.

Graph of qy/N vs. t/s of the interaction forces between rigid body 3 and beam 1 depicting transverse dashed waves representing the mixed method and "X" marks for Newton–Euler method.

Figure 9.18 Time history of the interaction forces between rigid body 3 and beam 1.

9.4 Finite Segment Transfer Matrix Method for Multibody Systems

This section introduces the finite segment discrete time transfer matrix method for multibody systems [81], which was developed by the authors. The beam is divided into finite rigid segments by the finite segment method, and the segments are connected by a torsional spring and a linear spring with a parallel damper. The inertia characteristic of the beam is described by a rigid segment, and the elasticity and damp characteristics are described by springs and dampers between the rigid segments, as shown in Figure 9.19. After being discretized by the finite segment method, the beam can be regarded as a chain‐type MRS connected by springs and dampers. The finite segment model is similar to the continuous beam model, with an increase in the number of rigid segments.

Finite segment model with a thick downward arrow labeled Equivalent from a beam to boxes connected by linear spring and damper.

Figure 9.19 The finite segment model of a beam.

The stiffness coefficient of each spring connecting the segments can be determined as follows. The relation between the internal moment of the beam and the rotation angle generated by bending the beam can be obtained according to Euler–Bernoulli beam theory,

(9.1) equation

where M is the internal moment of the beam, EI is the bending stiffness and θ is the rotation angle generated by bending the beam.

Substituting difference for differential, we obtain

(9.2) equation

where the subscript i denotes the ith finite segment, Δ denotes difference, c9-i0055 is the torsional stiffness coefficient of a torsional spring with respect to the corresponding finite segment and li is the length of the finite segment.

The elasticity of the finite segment is regarded as two series torsional springs. For a homogenous beam with uniform cross‐section, it becomes

(9.3) equation

where c9-i0056 and c9-i0057 are the stiffness coefficients of two series torsional springs corresponding to c9-i0058.

According to the finite segment model, the torsional springs c9-i0059 and c9-i0060 are introduced for two rigid segments of the series. The torsional stiffness coefficient between two rigid segments is

(9.4) equation

The moment and rotation angle between the two rigid segments can be obtained from

(9.5) equation

where c9-i0061 is the elastic moment between two rigid segments, and c9-i0062 and θi are the orientation angles of the (c9-i0063)th rigid segment and the ith rigid segment, respectively.

Similarly, considering the torsion and compression deformation of the beam, we obtain

(9.6) equation

where c9-i0064 is the elastic force between rigid segments, namely the internal force between rigid segments, and c9-i0065 and xi are the position coordinates of the hinge point of the (c9-i0066)th rigid segment and the ith rigid segment, respectively.

Similar to the above derivation, the elastic coefficient of the linear spring between two rigid segments is

(9.7) equation

where EA is the extensional rigidity coefficient.

Image described by caption and surrounding text.

Figure 9.20 Plane‐motion cantilever mounted on a rotary base.

Graph of drive moment/N.m vs. t/s of the driving moment, displaying a bell-shaped curve representing Kane method with diamond markers for finite segment MSDTTMM.

Figure 9.21 Time history of the driving moment.

Graph of deformation/(m) vs. t/s of a tip end, displaying an inverted bell-shaped curve representing the Kane method with diamond markers for finite segment MSDTTMM.

Figure 9.22 Time history of the transverse deformation of a tip end.

Graph of the longitudinal deformation of a tip end displaying an inverted bell-shaped curve from the origin representing the Kane method with diamond markers for finite segment MSDTTMM.

Figure 9.23 Time history of the longitudinal deformation of a tip end.

Graph of time history of rotation angular velocity displaying an ascending curve representing Kane method with diamond markers for finite segment MSDTTMM.

Figure 9.24 Time history of rotation angular velocity.

Graph of the relation between computational time and number of segments with an ascending line from 0.16 of the origin with square markers.

Figure 9.25 Relation between computational time and number of segments.

9.5 Transfer Matrix Method for Controlled Multibody Systems I

In reference [251], the control signal is regarded as the state variable of a state vector. The TMM is extended to control of a chain multibody system for modal analysis of the controlled multi‐link mechanical arms. The computational method for the dynamics of a real‐time controlled system, whose control force is related to the current system state, is introduced in this section. By taking the control characteristic parameters of the system as special mechanical parameters and re‐deriving the transfer matrices of controlled elements, there is no need to add the state variables corresponding to the control signal. The dynamics of controlled multibody systems can be studied by the MSTMM [77]. For a time‐delay controlled multibody system, the control force can be denoted by the previous system state and regarded as an external force, and only the control forces in the terms of the external force column matrix of the MSTMM need to be considered. The dynamics of the controlled system can be solved using the same methods of the MSDTTMM.

9.5.1 Transfer Matrix Method for Linear Real‐time Controlled Multibody Systems

9.5.1.1 Dynamic Model of Linear Controlled Multibody Systems

For the linear controlled multibody system shown in Figure 9.26, Kj and c9-i0075 denote the elastic coefficient of the spring and viscous damping coefficient of element j, respectively. The simple harmonic external force with frequency Ω acting on lumped mass mjc9-i0076 is

(9.12) equation
Dynamic model of a linear controlled multibody system with series of boxes labeled m2, m4, mp, mk, and m2n connected by spring and damper, with rightward arrows at the bottom for x2,1; x4,3; xp,p−1; xk,k−1; etc.

Figure 9.26 Dynamic model of a linear controlled multibody system.

The sensor of the controlled system is fixed on lumped mass mk and the real‐time control force acting on the lumped mass mp produced by the controller is

9.5.1.2 Transfer Equation and Extended Transfer Matrix of Elements

According to the method given in Section 2.7, when the system is in steady‐state motion, the motion of every lumped mass may be indicated by

where c9-i0077 is the complex amplitude of the steady‐state forced vibration.

The complex amplitude of the simple‐harmonic external force can be written as

(9.15) equation

Substituting Equation (9.14) into Equation (9.13), the control force of the controlled element is

(9.16) equation

The complex amplitude of the control force is

According to the geometrical relationship and the force balance condition of the two ends of the controlled lumped mass p, we can obtain

Substituting Equation (9.17) into Equation (9.18), these equations can be written in the following matrix form

Thus, the transfer equation of the real‐time controlled element p is

The extended transfer equation of other elements without control is

where

(9.22) equation

9.5.1.3 Overall Transfer Equation and Overall Transfer Matrix

From Equations (9.20) and (9.21), we obtain

The second formula of Equation (9.23) yields

Substituting Equation (9.24) into the first formula of Equation (9.23), the overall transfer equation is obtained

where the overall transfer matrix is

(9.26) equation

9.5.1.4 Solving the Steady‐state Motion of the System

The boundary conditions of the system are

Substituting Equation (9.27) into Equation (9.25) yields

Solving Equation (9.28), the unknown state variables in the state vector of the system boundary can be obtained

By using Equation (9.21) in sequence and Equation (9.20), the state vectors at any point of the system can be determined. Therefore, the complex amplitude c9-i0078 of each lumped mass can be obtained. From Equation (9.14), the steady‐state motion can be obtained

(9.30) equation
Linear multibody vibration system with 3 boxes labeled m2, m4, and m6 connected with springs and dampers, with a controlled loop depicted by a line from box m6 to box m2.

Figure 9.27 A linear multibody vibration system with a controlled loop.

Graphs of X21/m (top), X43/m (middle), and X65/m (bottom) vs. t/s, each has curves for multibody system dynamic method with diamond markers for transfer matrix method and “X” marks for non-controlled system.

Figure 9.28 The computational results of system dynamics. (a) Time history of the location coordinate of lumped mass 2. (b) Time history of the location coordinate of lumped mass 4. (c) Time history of the location coordinate of lumped mass 6.

2 Graphs illustrating the frequency response for Ka = 0 (top) and Ka = 5.0 (bottom), each with 3 intersecting curves representing X2,1 (solid); X4,3 (dashed); and X6,5 (dash-dotted).

Figure 9.29 Frequency response for a linear controlled multibody system with a controlled loop. (a) Frequency response for Ka = 0. (b) Frequency response for Ka = 5.0.

Linear controlled multibody system with 3 boxes for m2, m4, and m6 connected by strings and dampers labeled K1,C1; K3,C3; and K5,C5, with a controlled loop depicted by a line from m6 to K3,C3.

Figure 9.30 Linear controlled multibody system with a controlled loop.

Graphs of X21/m (top), X43/m (middle), and X65/m (bottom) vs. t/s, each has dashed curves representing TMM for linear controlled multibody system and scattered “X” marks for multibody system dynamics method.

Figure 9.31 Time history of the dynamics of a system: (a) time history of the displacement of lumped mass 2, (b) time history of the displacement of lumped mass 4 and (c) time history of the displacement of lumped mass 6.

Image described by caption and surrounding text.

Figure 9.32 Model of a branch multibody system.

Graph of time history of angular velocity of element 5 around x axis displaying a transverse wave and intersecting curve from the origin for Newton–Euler method, with diamond markers for TMM and “X” marks for MSTMM.

Figure 9.33 Time history of the angular velocity of element 5 around the x axis.

Graph of time history of angular velocity of element 5 around y axis displaying a transverse wave and intersecting curve from the origin for Newton–Euler method, with diamond markers for TMM and “X” marks for MSTMM.

Figure 9.34 Time history of the angular velocity of element 5 around the y axis.

Graph of time history of angular velocity of element 5 around z axis displaying a transverse wave and intersecting curve from the origin for Newton–Euler method, with diamond markers for TMM and “X” marks for MSTMM.

Figure 9.35 Time history of the angular velocity of element 5 around the z axis.

9.5.2 Transfer Matrix Method for Linear Delay Controlled Multibody Systems

A linear controlled multibody system is shown in Figure 9.26. If the system has a time delay τ and steady‐state motion, the motion of a lumped mass can be denoted by

where c9-i0124 is the complex amplitude of the steady‐state forced vibration.

The control force is

Equation (9.41) can be written as

(9.42) equation

The complex amplitude of the control force is

(9.43) equation

Taking the known control force Fp,c of the current time into the transfer matrix of the controlled element, the transfer equation of the controlled element is

(9.44) equation

namely

(9.45) equation

The overall transfer matrix is

(9.46) equation

According to the boundary conditions of the system and the transfer relation, the complex amplitude of each lumped mass c9-i0125 can be obtained. The steady‐state motion can be obtained from Equation (9.40):

(9.47) equation
3 Graphs illustrating X21/m (top), X43/m(middle), and X65/m (bottom) vs. t/s, each with transverse dashed curves representing TMM for linear multibody system and scattered “X” marks for Newton–Euler method.

Figure 9.36 The time history of the motion of a controlled system with time delay: (a) time history of the displacement of lumped mass 2, (b) time history of the displacement of lumped mass 4 and (c) time history of the displacement of lumped mass 6.

9.6 Transfer Matrix Method for Controlled Multibody Systems II

In this section, using the system in Figure 9.37 as an example, the discrete time transfer matrix method (DTTMM) for controlled multibody systems [78], developed by the authors, is introduced.

Controlled multibody system with 5 boxes labeled m2, m4, mp, mk, and m2n connected by springs and dampers with line from mk to mp and 5 rightward arrows for x2,1; x4,3; xp,p–1; xk,k–1; etc. at the bottom.

Figure 9.37 Controlled multibody system.

9.6.1 Discrete Time Transfer Matrix Method of Real‐time Controlled Multibody Systems

The DTTMM for multi‐rigid‐body systems was introduced in Chapter 7. The transfer equations of a lumped mass and a spring with longitudinal motion (see Figure 9.38) are obtained as follows

Image described by caption.

Figure 9.38 The subsystem composed of a spring and a lumped mass.

The state vector is

(9.50) equation

The transfer matrix of the spring is

(9.51) equation

The transfer matrix of the lumped mass is

(9.52) equation

Linearizing the real‐time control force as shown in Equation (9.13), this can be written as

From the geometry relation and the force equilibrium condition of element p, we have

Substituting Equation (9.53) into Equation (9.54), the transfer equation of the controlled element p is

where

(9.56) equation

From the system structure we can obtain

From the second formula of Equation (9.57), we have

Substituting Equation (9.58) into Equation (9.57), the overall transfer equation is

(9.59) equation

The overall transfer matrix is

Substituting the boundary conditions c9-i0126 and c9-i0127 into Equation (9.53) yields

(9.61) equation

Hence

(9.62) equation

The state vector of each element in the system at any time can be solved using Equations (9.48) and (9.49) repeatedly, and Equation (9.55).

3 Graphs illustrating X21/m (top), X43/m (middle), and X65/m (bottom) vs. t/sof the system motion with or without control, each with transverse solid curves representing ka = 5.0 and scattered “X” marks for ka = 0.0.

Figure 9.39 Time history of system motion with or without control: (a) time history of the displacement of lumped mass 2, (b) time history of the displacement of lumped mass 4 and (c) time history of the displacement of lumped mass 6.

Graphs of x21/m, x43/m, and x65/m vs. t/s of the system motion, each with transverse curves for Newton–Euler method, with diamond markers for TMM linear controlled and “X” marks for DTTMM controlled.

Figure 9.40 Time history of system motion: (a) time history of the displacement of lumped mass 2, (b) time history of the displacement of lumped mass 4 and (c) time history of the displacement of lumped mass 6.

Graphs of x21/m, x43/m, and x65/m vs. t/s of the system motion, each with transverse curves for Newton–Euler method with diamond markers for TMM linear controlled and “X” marks for DTTMM controlled.

Figure 9.41 Time history of system motion: (a) time history of the displacement of lumped mass 2, (b) time history of the displacement of lumped mass 4 and (c) time history of the displacement of lumped mass 6.

Image described by caption and surrounding text.

Figure 9.42 Model of a controlled planar flexible manipulator.

The following assumptions are made for system modeling:

  1. The thickness of the bond for each PZT actuator and the flexible arm is too thin to have a deep effect on dynamics of the system, thus is neglected.
  2. The PZT actuators are rigidly bonded to the flexible arm.
  3. The thickness of each PZT actuator is transparent, and their effects on the mass distribution and stiffness distribution of system are neglected. The flexible arm is considered to be a uniform cross‐section Euler–Bernoulli beam.
  4. The voltage of each PZT actuator is uniformly distributed. The polarization direction of each PZT is the same as the transverse vibration direction of the corresponding flexible arm. The trajectory tracking and active vibration control of the flexible arm are studied by combining the proportional‐differential (PD) feedback control and the modal velocity feedback control.

The control moments acting on the hub rigid body and the driving voltage applied to the ith segmented PZT are, respectively

(9.64) equation
(9.65) equation

where c9-i0138 and c9-i0139 are the desired angle and angular velocity of hub rigid body 1, respectively, θ1 and c9-i0140 are the actual angle and angular velocity, respectively, Kp and Kv are the proportion gain and velocity gain, respectively, τ0(t) is the control moment for drive motor, Kai is the gain coefficient of the ith segmented PZT actuator, Vi is the driving voltage applied to the ith segmented PZT actuator, u is the transverse deformation of the flexible arm, Yk(x2) is the kth c9-i0141 model function of the flexible arm, and c9-i0142. c9-i0143 is the position coordinate of each PZT on the corresponding flexible arm in the body‐fixed reference coordinate system.

Considering only the control moment on the rigid body related to its feedback state, the transfer matrix of the controlled rigid body vibrating in a plane is

where

equation
equation

u41, u42, u45, u46, u57 and u67 are the same as in Equation (7.101).

Using Equations (9.66) to (9.68), and considering the control moment of the PZT actuators, the transfer matrix of the controlled piezoelectric beam vibrating in a plane is

where

equation
equation

equation
equation

b and tb are the width and thickness of the flexible arm, respectively. ta is the thickness of the segmented PZT actuator.

For the fixed hinge whose inboard body is a rigid body and whose outboard body is a piezoelectric beam with large planar motion the position coordinates, orientation angles, internal moments and internal forces of the rigid body’s input and output are equal. The transfer relation between the generalized coordinates describing the deformation of the outboard flexible body and the input end of the fixed hinge only needs to be determined. The dynamics equation of the transverse vibration of the beam is given by Equation (8.91) and the highest order of the mode shapes is chosen to be 3. The distributed moments of the PZT actuators and the control equation of the PZT actuators should be taken into consideration when we deduce the transfer equation of the beam. The state vectors of the fixed hinge whose input end is a rigid body and output end is a piezoelectric beam with large planar motion are defined as

equation

The transfer equation is

(9.69) equation

The transfer matrix is

where

equation

For compound control by the PD controller and modal velocity feedback control based on PZT actuators, the transfer matrix c9-i0144 of the control element is

where c9-i0145 is the feedback parameter matrix related to the feedback parameter from feedback beam 3 to hub 1.

For the modal velocity feedback control, we have

(9.72) equation

According to the DTTMM for controlled multibody systems, the state vectors of the system are

(9.73) equation

The boundary conditions are

(9.74) equation

The overall transfer equation is

(9.75) equation

where c9-i0146, c9-i0147, c9-i0148 and c9-i0149 are determined by Equations (9.67), (9.70), (9.66) and (9.71), respectively.

The dynamics of the controlled flexible manipulator system calculated by the DTTMM for controlled multibody systems are shown in Figure 9.43. The time history of the expected orientation angle of the body‐fixed reference system of flexible arm 3 is shown in Figure 9.43a. The time history of the transverse deformation at the tip of arm 3 with and without PZT active control is shown in Figure 9.43b. The driving voltage applied to segmented PZT 1 is shown in Figure 9.43c. The simulation results show that the orientation angle of the flexible manipulator and expected track have good agreement. The transverse deformation at the tip of the arm decreases quickly. This validates the DTTMM for controlled multibody systems for the solution of the trajectory tracking of the flexible manipulator and active control of vibration.

image
image
image

Figure 9.43 Dynamics of a controlled flexible manipulator system: (a) time history of the orientation angle of a flexible arm, (b) time history of the transverse deformation at the tip of the arm and (c) time history of the driven voltage applied to segmented PZT 1.

Multibody system with a cuboid labeled 2 with an arrow to a box for K, leading to a circle with 2 ellipses intersecting inside labeled 1, linked to the ground symbol for 0.

Figure 9.44 Multibody system composing of a rigid body and an elastic hinge moving in space.

Graph θz/rad vs. t/s of rigid body around z axis displaying a transverse wave for Newton–Euler method with intersecting curve from the origin, with diamond markers for TMM controlled and “X” marks for MSTMM.

Figure 9.45 Time history of the angle of a rigid body around the z axis.

9.6.2 Discrete Time Transfer Matrix Method for Controlled Multibody Systems with Time Delay

For a controlled system with time delay, the control force in Equation (9.13) with time delay can be described as the function of the former motion parameters c9-i0167, c9-i0168 and c9-i0169 as follows

where τ is the time delay.

For the controlled system with time delay, substituting Equation (9.81) into Equation (9.54) and replacing u23 of the control transfer matrix by the known control force Fp,c in current time yields

equation

For the controlled system with time delay, the current control force is the function of the former force and can be regarded as an external force. The elements of c9-i0170 are all known functions of the previous time step, and Equation (9.60) can be written as

(9.83) equation

Applying the boundary conditions of the system and using Equations (9.48) and (9.49) repeatedly, the state vectors of each element at any time can be obtained.

Graphs illustrating x21/m, x43/m, and x65/m vs. t/s of the system motion, each with transverse curves for Newton–Euler method with diamond markers for TMM linear controlled and “X” marks for DTTMM controlled.

Figure 9.46 Time history of system motion: (a) time history of the displacement of lumped mass 2, (b) time history of the displacement of lumped mass 4 and (c) time history of the displacement of lumped mass 6.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.21.100.34