7
Discrete Time Transfer Matrix Method for Multibody Systems

7.1 Introduction

7.1.1 Characteristics of Ordinary Multibody System Dynamics Methods

With the development of modern industrial technology, many complex mechanical systems have appeared, such as weaponry, aeronautics, astronautics, vehicles, robots and precision machines, which are composed of many components with large relative motion. Under the condition of strong engineering demands, a new branch, namely multibody system dynamics, appeared. The numerical simulation of complex mechanical systems became possible with the appearance and development of the computer. Multibody system dynamics is an engineering application essential to the study of mechanical system dynamics, and systems consisting of many bodies undergoing large relative motion are its main study object. In general, the components of multibody systems include bodies and joints. Body components include rigid bodies, flexible bodies and lumped masses. A system is a multi‐rigid‐body system if the bodies involved are all rigid bodies, it is a multi‐flexible‐body system if the bodies involved are all flexible bodies and a multi‐rigid‐flexible‐body system if both rigid bodies and flexible bodies are involved. The joints are the components connecting the bodies.

The deduction of the dynamic equations of multibody systems and their numerical solutions are two main research areas of multibody system dynamics. Generally, there are two major methods for deriving dynamic equations, and their basic forms are different. The first method uses Lagrangian equations of the first kind, in which Lagrange multipliers are used to define unknown generalized constraint forces and obtain an augmented formulation. This approach leads to a mixed set of nonlinear differential and algebraic equations that have to be solved simultaneously, and it is the theoretical foundation of some commercial analysis software, such as ADAMS and DADS. It must solve a set of differential‐algebraic equations. The second method describes the motion using a set of first‐order differential equations (kinematics and dynamics) with a minimal set of unknown variables without constraint forces. From a point of view of computational efficiency, its numerical solution is more sophisticated. Differential‐algebraic equations can be transformed into a new form with a minimal number of coordinates, and a typical code is the MBOSS [246] developed by Arizona University.

With developments in engineering technology, researchers are still proposing and improving various multibody system dynamics methods, which are promoting the development of modern engineering technology. These methods have different forms, but in general they all have the two same characteristics:

  1. It is necessary to build the global dynamic equations of the system, which is the most interesting and brilliant part of various methods, but is also the difficult part for engineers.
  2. The orders of matrices involved in the global dynamic equations of the system are rather high (generally speaking not less than the degrees of freedom (DOFs) of the system). It has become one of the main research directions in multibody system dynamics to search for a new method to reduce the orders of relevant matrices and to avoid computational difficulty caused by high matrix order.

7.1.2 Features of the Discrete Time Transfer Matrix Method for Multibody Systems

To improve the computational efficiency of multibody system dynamics, a new method, the discrete time transfer matrix method for multibody systems (MSDTTMM) [3, 64–68], has been developed, which is different from other ordinary methods of multibody system dynamics. This method combines the advantages of the transfer matrix method and numerical integration. The dynamics equations of typical elements are derived first, followed by their discretization and linearization in the time domain based on the basic thought of numerical integration. Then the transfer matrices of these elements are obtained. The overall transfer equation and overall transfer matrix of the multibody system can be assembled. The system motion can be obtained using step‐by‐step time integration. The method is suitable for linear time‐variable, nonlinear, large motion and general multibody systems. Compared with ordinary multibody system dynamics methods, the MSDTTMM has the following features:

  1. The global dynamic equations of a system are not needed, hence its application and computation are very convenient.
  2. The order of involved matrices is much lower than for ordinary multibody system dynamics methods. For example, the order of relevant matrices for chain multibody systems is determined only by the highest order of differential equations of its subsystems. The computational cost is small, resulting in high computational efficiency.
  3. It provides a flexible modeling instrument and the modeling procedure is highly programmable. The modeling of a multibody system looks like building blocks, the overall transfer matrix of the system can be assembled by the transfer matrices of each element directly, and its solution procedure is very simple and straightforward.
  4. Complex multibody systems, such as branch, closed‐loop or network systems, can be conveniently dealt with by this method in a similar way to the transfer matrix method for linear multibody systems.
  5. The study method for multi‐flexible‐body systems and the MRFS is the same as that for multi‐rigid‐body systems.
  6. The transfer matrix is always a real number matrix, even when considering dampers. This simplifies the numerical algorithm.
  7. The natural frequency of linear multibody systems can be obtained easily by using the MSDTTMM, and the transfer matrix method for linear multibody systems can be regarded as a special case.
  8. An arbitrary appropriate numerical integration method can be introduced into the MSDTTMM, which gives the method some flexibility.

The basic principles, steps and algorithm of the MSDTTMM developed by the authors are introduced in this chapter, and the transfer matrices of typical elements are deduced.

7.2 State Vector, Transfer Equation and Transfer Matrix

Enlightened by the idea of matrix “transfer” in the transfer matrix method for linear time‐invariant systems, the concepts of state vector, transfer equation and transfer matrix for nonlinear time‐variant systems are introduced in this section. These new concepts are not the same as those in the transfer matrix method for linear time‐invariant systems. We start to address these concepts with a simple example.

Figure 7.1 shows a chain multi‐DOF system composed of springs and lumped masses moving along the ox axis. As shown in Figure 7.2, each spring and the lumped mass on its right can be regarded as a subsystem. The position coordinate of the left side of spring j is denoted by c7-i0001, the corresponding force is denoted by c7-i0002. Their positive direction is the same as the positive direction of the ox axis, and the position coordinate of the right side of spring j is denoted by c7-i0003, the force is denoted by c7-i0004 and their positive direction is the same as the negative direction of the ox axis. For a massless spring, we have

and for a linear spring, according to Hooke’s law, we obtain

where lj is the initial length of the spring and Kj is the stiffness coefficient.

A chain with multi-DOFs composed of 4 springs connected to lumped masses along the ox axis, with ellipsis between the lumped mass labeled 6 and a spring labeled n–1.

Figure 7.1 A chain system with multiple DOFs.

Subsystem model composed of a spring labeled Kj connected to a lumped mass labeled mj+1 with a rightward arrow. Vertical lines over the ends of the spring and over the arrow are labeled xj–1, j; xj+1, j; and xj+1, j+2.

Figure 7.2 Subsystem model.

For lumped mass c7-i0005, the position coordinate of its left side is denoted by c7-i0006 and the corresponding force denoted by c7-i0007. Their positive direction is the same as the positive direction of the ox axis, and the position coordinate of its right side is denoted by c7-i0008. The force is denoted by c7-i0009 and the positive direction is the same as the negative direction of the ox axis. According to Newton’s second law of motion, we obtain

where c7-i0010 is the mass of body c7-i0011 and c7-i0012 is the external force acting on the body.

There are many numerical methods which describe c7-i0013 and c7-i0014 as linear functions of x. For example, by using the Taylor series, we find

equation

c7-i0015 and c7-i0016 can be expressed as linear functions of c7-i0017:

(7.5)equation

where c7-i0018 is a time step.

Substituting Equation (7.4) into Equation (7.3) yields

According to the continuous condition of displacement of lumped mass, this becomes

Combining Equations (7.1) and (7.2) yields

The state vector at the connection point is defined as

(7.9)equation

Equation (7.8) can be written as

where

Equation (7.8) or Equation (7.10) is the transfer equation of a spring, and Equation (7.11) is the transfer matrix of a spring. Combining Equations (7.6) and (7.7) yields

namely

where

Equation (7.12) or Equation (7.13) is the transfer equation of a lumped mass, and Equation (7.14) is the transfer matrix of a lumped mass.

According to Equations (7.10) and (7.13), the transfer equation of the subsystem including the spring and lumped mass is

where c7-i0019 is the transfer matrix of the subsystem:

(7.16)equation

that is

(7.17)equation

By simply assembling the transfer matrices of the subsystems, the overall transfer equation of the chain multibody system shown in Figure 7.1 is

The overall transfer matrix of the system is

(7.19)equation

Substituting the boundary conditions c7-i0020, c7-i0021 and the initial condition at time t0 into Equation (7.18), the unknown quantities of state vectors of the system boundary at time t1 can be obtained. For the chain multibody system shown in Figure 7.1, c7-i0022 and q0,1 at time t1 can be determined. The dynamics of all elements can be computed by using Equation (7.15), which can be considered as the initial condition for computing the dynamic response at time t2. The system dynamic response at an arbitrary time point is determined via the iterative method.

For the model shown in Figure 7.1, we choose c7-i0023, with structure parameters and initial conditions according to Table 7.1. The computational results of motion by the transfer matrix method for linear multibody systems, the MSDTTMM and the Newton mechanics method, respectively, are shown in Figure 7.3. The solid line denotes the results computed by the Newton mechanics method, the symbol + denotes the results computed by the MSDTTMM and the symbol * denotes the results computed by the transfer matrix method for linear multibody systems. The results obtained by these three methods have good agreement and validate the proposed method.

Table 7.1 Structure parameters and initial conditions

Spring number 1 3 5 7 9
Stiffness K (N/m) 1000 1000 1000 1000 1000
Initial length l (m) 0.1 0.1 0.1 0.1 0.1
Lumped mass number 2 4 6 8 10
Mass m (kg) 0.1 0.1 0.1 0.1 0.1
External force f (N) 100 0 0 0 0
Initial displacement x0 (m) 0.1 0.2 0.3 0.4 0.5
Initial velocity c7-i0024 0 0 0 0 0
Graph of t (s) vs. x (m) displaying 3 discrete waves representing Newton–Euler method (solid line), MS-Dt-TMM (plus sign), and linear MS-TMM (asterisk).

Figure 7.3 Computational results by three methods.

The state vector of a rigid body moving in space is defined as

where x, y, z, θx, θy and θz are the position coordinates of the connection point with respect to the inertial coordinate system and the corresponding rotation angles about the x, y and z axes. mx, my, c7-i0025, qx, qy and qz are the corresponding internal moments and forces in the same coordinate system.

The state vector of a rigid body moving in a plane is defined as

For the body element j with one input end and one output end, the transfer equation which relates the state vectors at its two points is

The transfer equation which relates the state vectors at the two points of a joint j is

where c7-i0026 is the transfer matrix of element j.

For a chain system composed of n elements, the overall transfer equation which relates the state vectors at the two points can be obtained by using Equations (7.22) and (7.23) repeatedly,

(7.24)equation

where

(7.25)equation

c7-i0027 is the overall transfer matrix of the system.

Once the overall transfer matrix of the system is obtained, the boundary conditions of the system can be applied and the unknown quantities in the boundary state vectors can be computed. The state vectors of each element at time ti can be obtained using Equations (7.22) and (7.23) repeatedly and the system motion at time ti can be obtained. The entire procedure can be repeated for time c7-i0028 and so on.

7.3 Step‐by‐step Time Integration Method and Linearization

The linear MSTMM is established based on the fact that the internal forces at the same point are equal in magnitude, opposite in direction and act on different objects, resulting in a linear relationship between different points’ state vectors. For a nonlinear system, this linear relationship does not exist. Even for a linear time‐variant system, this linear relationship also does not exist. If we still want to use the transfer matrix method to study the dynamics of such a system, the linear relationship of state vectors between different points must be developed. Therefore, two questions must be answered: the first is whether physically nonlinear relations of state vectors involving transcendent functions or other complex relations can be expressed as linear forms; the second is how these complex relations can be linearly expressed to make the computation convergent and stable. The answers which come from modern science are affirmative. In fact, the increment transfer matrix method of a nonlinear time‐variant system was presented in Chapter 4 to compute the steady‐state response. For a time‐variant system, during the physical process of any computational time step size, if the time step size is small enough the relationship among many physical quantities can be approximately expressed by linear formulas. Modern numerical integral methods [222, 242, 247, 248] provide many linear forms to ensure computational convergence and stability. This section introduces these linearization methods of motion parameters and special functions that are based on step‐by‐step time integration methods.

7.3.1 Linearization of Velocity (Angular Velocity) and Acceleration (Angular Acceleration)

In dynamic equations of general elements, the unknown quantities include velocities (angular volocities), accelerations (angular accelerations), trigonometric functions and internal forces at connection points, as well as cross terms, high‐order terms and so on. For the free vibration of linear time‐invariant multibody systems, the state vectors satisfy the following relation:

equation

where z is the physical coordinate and Z is the corresponding modal coordinate.

We can obtain

If the motion of a multibody system is aperiodic, Equation (7.26) does not exist. In order to develop the transfer matrix of an element, its dynamic equation must be linearized. Using numerical integral methods, velocity c7-i0029 (angular velocity) and acceleration c7-i0030 (angular acceleration) can be expressed as linear functions of the position coordinate (angular coordinate), that is

where c7-i0031, c7-i0032, c7-i0033 and c7-i0034 at time ti are known functions with respect to time c7-i0035, and can be summarized as A, Bz, C and Dz.

Based on the linearization of velocities, accelerations and other variables in a small time segment during time interval c7-i0036, the nonlinear dynamic equation at time ti can be expressed by linear formulas of the motion variables at time ti. The quantities above can be determined using the following methods.

7.3.2 Step‐by‐step Time Integration Method

7.3.2.1 Forward–Euler Method

According to the Taylor expansion theorem, x(ti) can be approximately expressed as

(7.29)equation

From Equation (7.30), it follows that

Thus, the acceleration can be expressed approximately by the first‐order difference as c7-i0037 from Equation (7.30), that is

By comparing Equations (7.30) and (7.31) with Equations (7.27) and (7.28), we obtain

(7.32)equation

7.3.2.2 Newmark‐β Method

Expanding the variable x(ti) at time c7-i0038 by Taylor expansion theorem yields

(7.33)equation

Let

(7.34)equation

then

The velocity can be expressed by using Taylor expansion theorem as

Let

Substituting Equations (7.35) and (7.37) into Equation (7.36) yields

From Equations (7.35) and (7.38) we obtain

(7.39)equation

The Newmark‐β method originates from the trapezoid method, where only the first and second‐order derivatives of the Taylor series are adopted. The two parameters γ and β are used to compensate the contributions of the abandoned higher‐order quantities. If γ and β have different values, many formulas can be obtained. When c7-i0039 and c7-i0040, this is the average acceleration method. c7-i0041 leads to the constant acceleration method (center difference method), and c7-i0042 yields the linear acceleration method. It can be proved that if γ ≥ 1/2, βγ/2, and the formulas are unconditionally stable. If the value of β is increased, the computational precision will reduced, and if c7-i0043, the computational precision is the highest, but the formula is conditionally stable. Newmark has pointed out that c7-i0044 induces negative damping to a linear system, i.e. the vibration amplitude will be amplified in the integral computation. When c7-i0045, artificial damp will be introduced and finally induce reduction in the vibration amplitude. Generally, let γ ≥ 1/2. The most common choice is to let c7-i0046 and then change β accordingly, so the method is called the Newmark‐β method. If c7-i0047 and c7-i0048, this is the Forward–Euler method.

7.3.2.3 Wilson‐θ Method

The Wilson‐θ method was developed on the basis of the linear acceleration assumption. The parameter θ ≥ 1 can be introduced in the time interval c7-i0049, assuming acceleration is linear with respect to time, as shown in Figure 7.4. A set of equations at time c7-i0050, called predicted equations, is deduced first. Then the correction equations of displacement, velocity and acceleration at time c7-i0051 are deduced. When c7-i0052, this is the linear acceleration method. If τ is the time variable limited by 0 ≤ τθΔT, the acceleration during this time interval can be written as c7-i0053, according to the assumption of linear acceleration. Later the acceleration can be integrated in [0, τ].

Linear acceleration, depicted by a trapezoid divided into 2 parts, with vertices labeled ẍn, ẍn+1, ẍn+ϑ, t, t +Δt, and t +ϑΔt.

Figure 7.4 Linear acceleration.

If c7-i0054, according to Equations (7.40) and (7.41), the velocity and displacement at time c7-i0055 can be computed by the following two formulas

Combining Equations (7.42) and (7.43) yields

According to Equations (7.44) and (7.45), we obtain

equation
(7.46)equation

7.3.2.4 Houbolt Method

The basic idea of the Houbolt method is that the velocity and acceleration at c7-i0056 can be expressed approximately by the position coordinates of four points at previous times c7-i0057, as shown in Figure 7.5. The relative coordinate c7-i0058 is introduced, and c7-i0059, c7-i0060, xn and c7-i0061 are the position coordinates corresponding to c7-i0062, 1, 2 and 3, respectively. When 0 ≤ ζ ≤ 3, the position coordinate x(ζ) can be expressed approximately by the Lagrange interpolation polynomial:

(7.47)equation
Houbolt method, illustrated by a curve with divided by 4 vertical lines labeled xn−2, xn−1, xn, and xn+1 with their corresponding points 0, 1, 2, and 3, respectively, along the relative coordinate ζ = t/ΔT.

Figure 7.5 Houbolt method.

It is not difficult to figure out the rules of this interpolation polynomial: any individual term of this polynomial is the product of a position coordinate function and three terms of the four basic functions ζ, c7-i0063, c7-i0064 and c7-i0065. Its coefficients are obtained by the value of the function x(ζ) corresponding to each point. The first‐ and second‐order derivatives of x(ζ) are

(7.48)equation
(7.49)equation

If c7-i0066, then

Three previous time points are used in this linearization method, hence the first and second steps cannot be implemented in these formulations. The method can only be used if the information for three time points is known. This is the disadvantage of the multistep method. The two‐step and single‐step Houbolt methods are described here. Equation (7.50) yields

as

Adding Equations (7.53) and (7.54), then

Substituting Equation (7.55) into Equations (7.51) and (7.52), the linearization coefficients of the two‐step Houbolt method can be obtained:

(7.56)equation

As

(7.57)equation

then

(7.58)equation
(7.59)equation

Therefore, the linearization coefficients of the one‐step Houbolt method become

(7.60)equation

In summary, the higher‐order Houbolt method can be constructed by a higher‐order Lagrange interpolation polynomial.

For different step‐by‐step time integration methods, different formulations of A, Bz, C and Dz can be obtained. Linearization coefficient formulations for different step‐by‐step time integration methods are shown in Table 7.2.

Table 7.2 Linearization coefficient for different step‐by‐step time integration methods

Method A B z
TMM c7-i0067 0
Forward–Euler c7-i0068 c7-i0069
Newmark‐β c7-i0070 c7-i0071
Wilson‐θ c7-i0072 c7-i0073
Houbolt for i ≥ 3 c7-i0074 c7-i0075
Houbolt for i = 2 c7-i0076 c7-i0077
Houbolt for i = 1 c7-i0078 c7-i0079
Method C Dz
TMM 0 0
Forward–Euler c7-i0080 c7-i0081
Newmark‐β c7-i0082 c7-i0083
Wilson‐θ c7-i0084 c7-i0085
Houbolt for i ≥ 3 c7-i0086 c7-i0087
Houbolt for i = 2 c7-i0088 c7-i0089
Houbolt for i = 1 c7-i0090 c7-i0091

Let the variable z represent a column matrix, such as the position coordinates c7-i0092 or orientation angles c7-i0093. At present, of the four linearization coefficients, A and C are irrespective of the state vector coefficient z, hence their expressions are invariant. Bz and Dz, however, should be rewritten as column matrices c7-i0094 and c7-i0095 or c7-i0096 and c7-i0097, respectively. For example, in the Newmark‐β method

equation

7.3.3 Linearization of Nonlinear Functions

Correct to ΔT2, we expand the trigonometric functions at time ti as an explicit function of previous time c7-i0098 by the Taylor expansion theorem, that is

(7.61)equation

where

(7.62)equation

In the geometric relationship from input end to output end, which includes trigonometric functions, the trigonometric functions can be linearized by the second‐order Taylor expansion, that is

For higher‐order terms in body dynamic equations, the second‐order terms can be expressed approximately as linear functions, that is

(7.65)equation

The linear expression of a third‐order function is

The above expansion formulations neglect the effect of terms that are higher than ΔT3 order. For higher‐order terms, their linearization expression can be obtained by using the same method.

The linearization expression of a general polynomial can be written as

(7.67)equation

where c7-i0099 and c7-i0100 at time ti are constants or functions of variables at previous times.

Substituting differential for finite difference

(7.68)equation

yields

(7.69)equation

Hence

(7.70)equation

Linearization formulas of the trigonometric functions sin θ(ti), cos θ(ti) and higher‐order functions a(ti)b(ti) and a(ti)b(ti)c(ti) are used to derive the transfer matrices of elements. For convenience, all of these formulations are listed in Table 7.3.

Table 7.3 Linearization formulas of trigonometric functions and higher order functions

Function Linearization formulas
sin θ(ti) c7-i0101
cos θ(ti) c7-i0102
sin θ(ti) c7-i0103
cos θ(ti) c7-i0104
a(ti)b(ti) c7-i0105
a(ti)b(ti)c(ti) c7-i0106

7.3.4 Linearization of Coordinate Transformation Matrices

The angular motion of a rigid body with respect to a fixed point can be realized by three rotations along three spatial fixed axes successively. The proof can be found in Appendix A.

A motion of a rigid body B relative to reference frame A is shown in Figure 7.6, where c7-i0107 is any vector fixed in reference system A, and c7-i0108 is a vector fixed on rigid body B. Before B rotates relative to A, c7-i0109. One fixed point’s 3 DOF rotation can be described by three angles which rotate in sequence around three arbitrary non‐coplanar axes. Herein we consider the angles θx, θy, θz that rotate in sequence around three spatial fixed axes in inertial coordinates. Therefore, the transform matrix, from the body‐fixed coordinate system to the inertial coordinate system, can be expressed by trigonometric functions of angles θx, θy, θz as follows:

A rectangle labeled A has a circle inside labeled B plotted on a coordinate plane labeled L, LA, and LB, which has a double-headed arrow labeled ϑ between LA and LB. Other arrows labeled a, b, and λ are indicated.

Figure 7.6 Simple rotation.

Where

(7.72)equation

Using multiple Taylor expansion, for transform matrix c7-i0110 at c7-i0111, keeping second‐order terms, we obtain

According to Equation (7.71)

(7.74)equation
(7.75)equation
(7.76)equation

Let

Then

Similarly, it is easy to prove that

(7.79)equation

Combining Equations (7.71) and (7.78) yields

(7.80)equation
(7.81)equation

Let

equation

Substituting Equations (7.78) to (7.82) into Equation (7.73) yields

where

7.4 Transfer Matrix of a Planar Rigid Body

A multi‐rigid‐body system is a system composed of several rigid bodies connected by all kinds of joints. Based on the MSDTTMM, once the transfer matrices of the elements are obtained, the multibody system dynamics can be computed and studied by simple programming matrix operations. It is important to point out that the transfer matrices of these elements remain the same for any multi‐rigid‐body systems and do not change with the structural variation of the multi‐rigid‐body system. The method provides flexibility in the modeling and computation of the dynamics of multibody systems with varying configuration, and this is one of the characteristics of the proposed method compared with any other method of multibody system dynamics. In this section, the dynamic equations of a rigid body are derived relative to the inertial coordinate system oxyz. The series uniform transfer equations and transfer matrices are deduced according to the input and output situations of all kinds of rigid bodies.

7.4.1 Planar Rigid Body with One Input End and One Output End

For the planar rigid body shown in Figure 7.7 the input end is I, the output end is O and the mass center is C. The inertial coordinate system is oxy and the body‐fixed coordinate system is O2x2y2 with origin I. In the coordinate system O2x2y2, the position coordinates of O are (b1, b2) and for the mass center C they are (cc1, cc2). The body‐fixed coordinate system is obtained by rotating o1x1y1 about the z2 axis with angle θ. Hence,

where

equation
Image described by caption and surrounding text.

Figure 7.7 Rigid body moving in a plane with one input and one output end.

The external forces and moments acting on the rigid body are equivalent to the external forces and moments acting on the mass center of the rigid body. Therefore, we only consider the external forces and moments acting on the mass center of rigid bodies in the following analysis. The motion equation of the mass center is

where m is the mass of the rigid body and (xC, yC) are the position coordinates relative to the inertial coordinate system, while qx,I and qy,I are internal forces acting on the input end of the rigid body, and qx,O and qy,O are internal forces acting on the output end. fx,C and fy,C are external forces acting on the mass center of the rigid body.

According to the theorem of absolute angular momentum with respect to moving point I

The projection of Equation (7.92) on the inertial coordinate system can be written as

where c7-i0112 is the absolute moment of momentum of the rigid body with respect to origin I of the body‐fixed coordinate system, and JI is the rotational inertia relative to point I. c7-i0113 is the absolute angular velocity of the rigid body. rIC is the radius vector of the mass center C of the rigid body with respect to the point I, and aI is the absolute acceleration of point I.

Linearizing Equations (7.88) and (7.89) using Equations (7.63) and (7.64) yields

(7.94)equation

Substituting Equations (7.86) and (7.87) into Equations (7.90) and (7.91), and linearizing using Equations (7.27), (7.63) and (7.64), we obtain

Substituting Equations (7.96) and (7.97) into Equation (7.93), and linearizing using Equation (7.27), we obtain

Combining Equations (7.85) and (7.95) to (7.98), the transfer equation of the rigid body moving in a plane with one input and one output end can be expressed as

(7.99)equation

The state vector is

(7.100)equation

and the transfer matrix is

where

(7.102)equation
(7.103)equation

7.4.2 Planar Rigid Body with Multiple Input and Multiple Output Ends

In this section, we derive the transfer equations and transfer matrices of rigid bodies with multiple input and multiple output ends, rigid bodies with one input end and multiple output ends, and rigid bodies with one output end and multiple input ends. For rigid bodies with multiple input and multiple output ends, we cannot obtain the main transfer equation from one point to another point. Only the matrix relation between input and output ends can be obtained.

Figure 7.8 is the dynamic model of a rigid body moving in a plane. Assuming the input ends are c7-i0114 and the output ends are c7-i0115, the mass center is C. The inertial coordinate system is oxy. The coordinate system o1x1y1 is fixed on the first input end I1, and parallels the inertial coordinate system. The coordinate system O2x2y2 is the body‐fixed coordinate system and its origin is the same as o1x1y1. The position coordinates of the input and output ends in the coordinate system O2x2y2 are (a1,1, a2,1), (a1,2, a2,2), c7-i0116 and (b1,1, b2,1), (b1,2, b2,2) c7-i0117, respectively. The position coordinates of the mass center of the rigid body in the coordinate system O2x2y2 are (cc1, cc2).

Image described by caption and surrounding text.

Figure 7.8 Rigid body with multiple input and multiple output ends moving in a plane.

Suppose we rotate o1x1y1 by angle θ to obtain body‐fixed coordinate O2x2y2. If c7-i0118, it is a one input end and one output end rigid body. When c7-i0119 and c7-i0120, it is a rigid body with one input end and multiple output ends. When c7-i0121 and c7-i0122, it is a rigid body with multiple input ends and one output end. When c7-i0123 and c7-i0124, it is a rigid body with multiple input ends and multiple output ends.

Similarly to Equations (7.85) to (7.89), considering Equations (7.63) and (7.64) we obtain

where

(7.109)equation
(7.110)equation
(7.111)equation
(7.112)equation
(7.113)equation
(7.114)equation

(xC, yC) are the position coordinates of the mass center of a rigid body with respect to the inertial coordinate system. c7-i0125 and c7-i0126 are the position coordinates of the input end Ij and the output end Oj in the inertial coordinate system.

In the following analysis, we only consider the external force and external moment acting on the mass center of the rigid body. According to the dynamic equation of the mass center of a rigid body

where m is the mass of the rigid body, c7-i0127 and c7-i0128 are the internal forces acting on the input end of the rigid body, c7-i0129 and c7-i0130 are the internal forces acting on the output end of the rigid body, and fx,C and fy,C are the external forces acting on the mass center of the rigid body.

Based on the theorem of the absolute angular momentum on the moving momentum center, which is the first input end I1, we obtain

The projection of Equation (7.117) on the inertial coordinate system can be written as

Substituting Equations (7.27), (7.106), (7.63) and (7.64) into Equations (7.115) and (7.116) yields

Substituting Equations (7.27), (7.119) and (7.120) into Equation (7.118) yields

Equations (7.107) and (7.104) can be written in matrix form

where

(7.123)equation
(7.124)equation

Equations (7.105) and (7.108) can be written in matrix form

where

(7.126)equation

Equations (7.119) to (7.121) can be written in matrix form

where

(7.128)equation
(7.129)equation
(7.130)equation

The state vector of the rigid body with multiple input ends and multiple output ends can be defined as

(7.131)equation

According to Equations (7.122), (7.125) and (7.127), the transfer equation of the rigid body with multiple input ends and multiple output ends is

(7.132)equation

The transfer matrix is

Similarly, the transfer equations and transfer matrices of the rigid body with multiple input ends and one output end, and the rigid body with one input end and multiple output ends can be obtained, which are provided in the following sections without the detailed deduction procedures.

7.4.2.1 Transfer Equation and Transfer Matrix of a Rigid Body moving in a Plane with Multiple Input Ends and One Output End

State vector

(7.134)equation

Transfer equation

(7.135)equation

Transfer matrix

(7.136)equation

All submatrices are the same as those in Equation (7.133).

For computation convenience, c7-i0131 can be expressed by c7-i0132 as

(7.137)equation

7.4.2.2 Transfer Equation and Transfer Matrix of a Rigid Body Moving in a Plane with One Input End and Multiple Output Ends

State vector

(7.138)equation

Transfer equation

(7.139)equation

Transfer matrix

(7.140)equation

All submatrices are the same as those in Equation (7.133).

For computation convenience, c7-i0133 can be expressed by c7-i0134 as

(7.141)equation

where

equation

7.5 Transfer Matrices of Spatial Rigid Bodies

In this book, the spatial motion of a multibody system is studied in the inertial coordinate system using three spatial angles about the x–y–z axis defined in reference [9], which is very convenient for the transformation of state vectors.

7.5.1 Spatial Rigid Body with One Input End and One Output End

The column matrix of the angular velocity of a rigid body with respect to the inertial coordinate system in the body‐fixed coordinate system of a rigid body is

(7.142)equation

and its skew symmetric matrix is

(7.143)equation

The first‐order derivative with respect to time of transformation matrix c7-i0135 defined in Equation (7.71) is

equation

The dynamics equation and transfer matrix are deduced using the three spatial axis angles about x–y–z, that is

where

equation

It can be proved that

(7.146)equation

The first‐order derivative with respect to the time of the angular velocity ω of the rigid body in the inertial coordinate system is

(7.147)equation

The time derivative of Equation (7.144) is

(7.148)equation

where

(7.149)equation

For a rigid body moving in space, as shown in Figure 7.9, the column vectors of the rotation angle and positions are defined as c7-i0136 and c7-i0137, respectively. For the input end and output end

where c7-i0138 is the coordinate transformation matrix from O2x2y2z2 to o1x1y1z1.

Graph displaying a cube with 6 intersecting arrows labeled y2, y1, x2, x1, z1, and z2 and circles labeled C and O. The common point of the arrows is labeled I, which has a southwest arrow labeled z.

Figure 7.9 Rigid body moving in space.

If only the external force and moment acting on the mass center of the rigid body are considered, according to the motion equation of the mass center

where m is the mass of the rigid body and c7-i0139 is the radius vector of the mass center of the rigid body with respect to the origin of the inertial coordinate system. c7-i0140 is the internal force acting on the input end of the rigid body, c7-i0141 is the internal force acting on the output end of rigid body and c7-i0142 is the external force acting on the mass center of the rigid body.

The projection of Equation (7.152) on the inertial coordinate system can be written as

where c7-i0143 is the coordinate column matrix of the mass center of the rigid body with respect to the inertial coordinate system:

(7.154)equation

According to the theorem of absolute angular momentum to the moving momentum center, we obtain

The projection of Equation (7.155) on the inertial coordinate system can be written as

(7.157)equation

where c7-i0144, c7-i0145 and c7-i0146 is the absolute moment of momentum of the rigid body with respect to point I. c7-i0147 and c7-i0148 are the internal moments acting on points I and O, respectively, c7-i0149 is the internal forces acting on point O, c7-i0150 is the external moment acting on the mass center of the rigid body, c7-i0151 is the acceleration of the point I and c7-i0152 is the inertia matrix of the rigid body with respect to point I in the body‐fixed coordinate system.

The transfer matrix is now derived. Substituting Equation (7.83) into Equation (7.151), we obtain

where

equation

c7-i0153, c7-i0154 and c7-i0155 can be obtained by Equations (7.84), (7.77) and (7.20), respectively.

Equation (7.150) can be written as

Linearizing the left‐hand terms of Equation (7.153) yields

Replacing the subscript O in Equation (7.158) by C, we obtain

Substituting Equation (7.161) into Equation (7.160) yields

where

equation

The coordinate column matrix of angular acceleration in the body‐fixed coordinate system can be obtained, that is

equation

Let

(7.163)equation

then the coordinate column matrix of angular acceleration in the body‐fixed coordinate system can be rewritten as

Combining Equations (7.144) and (7.77) yields

According to Equations (7.145), (7.164) and (7.165)

(7.166)equation

Let

(7.167)equation
(7.168)equation

then

where

(7.170)equation
(7.171)equation

equation

Substituting Equations (7.27) and (7.169) into Equation (7.156) yields

where

equation

Combining Equations (7.158), (7.159), (7.162) and (7.172), the transfer equation of a rigid body moving in space with one input end and one output end can be obtained:

(7.173)equation

The transfer matrix is

7.5.2 Spatial Rigid Body with Multiple Input and Multiple Output Ends

7.5.2.1 Transfer Equation and Transfer Matrix of a Rigid Body moving in Space with Multiple Input and Multiple Output Ends

The dynamics model of a rigid body moving in space with multiple input and multiple output ends is shown in Figure 7.10. c7-i0156 are input ends, c7-i0157 are output ends and C is the mass center of the rigid body. The inertial coordinate system is oxyz. The origin of the coordinate system o1x1y1z1 is the first input end I1 of the rigid body, which is parallel to the inertial coordinate system. The origin of the body‐fixed coordinate system O2x2y2z2 is the same as the coordinate system o1x1y1z1. The coordinates of the input and output ends in the body‐fixed coordinate system O2x2y2z2 are (a1,j, a2,j, a3,j) and (b1,j, b2,j, b3,j), respectively, and the coordinates of the mass center are (cc1, cc2, cc3). Then;

(7.178)equation

where c7-i0158 and c7-i0159 are the coordinates column matrices of the input and output ends in the inertial coordinate system, rC is the vector of the mass center relative to the origin of the inertial coordinate system, c7-i0160 are the coordinates column matrix of the mass center of the rigid body in the inertial coordinate system and A is the transformation matrix from the coordinate system O2x2y2z2 to o1x1y1z1.

Image described by caption and surrounding text.

Figure 7.10 Rigid body moving in space with multiple input and multiple output ends.

Linearizing Equations (7.175) and (7.176) yields

where

equation

Linearizing Equations (7.175) and (7.177) yields

where

equation

According to the dynamic equation of the mass center of the rigid body, we obtain

Projecting Equation (7.181) into the inertial coordinate system yields

where m is the mass of the rigid body. c7-i0161 are the internal forces at the input ends and c7-i0162 are the internal forces at the output ends. fC is the external force acting on the rigid body mass center.

According to the theorem of absolute angular momentum to the moving momentum center, the dynamics equation of a rigid body with multiple input and multiple output ends can be written as

Projecting Equation (7.183) into the inertial coordinate system yields

equation

where

equation

Linearizing Equations (7.182) and (7.184) yields

where

equation

The state vector of a rigid body with multiple input ends and multiple output ends can be defined as

(7.186)equation

Combining Equations (7.179), (7.180) and (7.185), the transfer equation of a rigid body with multiple input ends and multiple output ends can be obtained

(7.187)equation

The transfer matrix is

In a similar manner, the transfer equations and transfer matrices of a rigid body with multiple input ends and one output end, and a rigid body with one input end and multiple output ends can be found. The results are given in the following sections.

7.5.2.2 Transfer Matrix of a Rigid Body with Multiple Input Ends and One Output End

State vector

(7.189)equation

Transfer equation

(7.190)equation

Transfer matrix

(7.191)equation

All submatrices are the same as those in Equation (7.188).

For ease of calculation, c7-i0163 can be expressed by c7-i0164, hence

(7.192)equation

7.5.2.3 Transfer Matrix of a Rigid Body with One Input End and Multiple Output Ends

State vector

(7.193)equation

Transfer equation

(7.194)equation

Transfer matrix

(7.195)equation

All submatrices are the same as those in Equation (7.188).

For ease of calculation, c7-i0165 can be expressed by c7-i0166, hence

(7.196)equation

where

equation

7.6 Transfer Matrices of Planar Hinges

The interaction (such as damper, spring, driven constraint or motion and contact) between different bodies can be regarded as a force element in the ordinary dynamics methods of multibody systems. However, in MSDTTMM, the interactions between different bodies can be regarded as hinges, and that makes dynamic modeling more convenient and flexible.

7.6.1 Elastic Hinges

Here, the elastic hinge consists of linear (or nonlinear) spring hinges and linear (or nonlinear) rotary spring hinges. Figures 7.11 and 7.12 illustrate the models for linear spring and nonlinear spring forces, respectively.

Graph of q vs. Δ l illustrating the linear spring force model, depicted by an ascending line.

Figure 7.11 Linear spring force model.

Graph of q vs. Δ l illustrating the linear spring force model, depicted by an ascending curve.

Figure 7.12 Nonlinear spring force model.

Ignore the mass and the initial length of the spring and assume

where c7-i0167 is the change in length of a spring in the xy plane. O denotes the output end and I denotes the input end. c7-i0168 is the spring force, and K1 and K2 are the stiffness coefficients of a nonlinear spring.

Projection of Equation (7.197) into the ox and oy directions results in

Substituting Equation (7.66) into Equation (7.198) yields

where

equation

According to Equation (7.199), we obtain

where

equation

The moment of a nonlinear rotary spring is

where c7-i0169 and c7-i0170 are the stiffness coefficients.

Similarly, linearizing Equation (7.202), we obtain

For linear spring and rotary spring hinges whose masses are neglected, we obtain

(7.204)equation
(7.205)equation

Rewriting Equations (7.200), (7.201) and (7.203) to (7.206) in terms of matrices, the transfer equation of a spring moving in a plane is

(7.207)equation

The state vector is

equation

and the transfer matrix is

where

equation

For a linear spring, that is, c7-i0171, the transfer matrix can be derived from Equation (7.208) as

(7.209)equation

7.6.2 Damper Hinges

For a viscous damper, the viscous damped forces and damped moment are

where Cd and c7-i0172 are the damping coefficients of translational dampers and rotary dampers, respectively.

Linearizing Equation (7.210), we obtain

(7.212)equation

From Equations (7.211) to (7.213), the transfer equation of a viscous damper hinge moving in a plane is

(7.214)equation

The state vector is

equation

and the transfer matrix is

(7.215)equation

where

equation

7.6.3 Smooth Hinges

For the smooth hinge shown in Figure 7.13, at the two ends the position coordinates and internal forces are equal and the internal torques are constant zero:

Image described by caption and surrounding text.

Figure 7.13 Model of a smooth hinge.

The numbers of inboard and outboard bodies of the smooth hinge j are c7-i0173 and c7-i0174, respectively. In the following, the rotation angle θO of the outboard body c7-i0175 of a smooth hinge is discussed for two cases.

7.6.3.1 Smooth Hinge whose Outboard is a Rigid Body and with Internal Torques at the Output End of Zero

If another point of the outboard rigid body is also connected to a smooth hinge or a free boundary, then we can obtain the transfer equation of the outboard rigid body of the smooth hinge

where u41, u42, u43, u45, u46 and u47 are elements of the transfer matrix of the outboard rigid body.

Substituting Equation (7.216) into Equation (7.217) yields

Combining Equations (7.216) and (7.218), and referring to the state vectors defined by Equation (7.21), we obtain the transfer equation of the smooth hinge whose outboard is a rigid body and whose internal torques of the output end are zero, namely

equation

The transfer matrix is

In Equation (7.219), the element u4,4 could also be written as “1”. However, if we set c7-i0176, the computational precision will be improved. According to the above derivation, if the internal moments of the outboard body at the two ends of the smooth hinge are equal to zero, the above transfer equation and transfer matrix are valid.

7.6.3.2 Smooth Hinge whose Outboard is a Rigid Body and with Nonzero Internal Torques at the Output End

The output end of the outboard rigid body of a smooth hinge is connected to an elastic hinge or a damper hinge, as shown in Figure 7.14. Among the outboard elements of the smooth hinge, if there is such an element whose internal moment is always zero or with free boundaries, then the transfer equation of this smooth hinge can be deduced in the following way. The transfer equation of the subsystem from the smooth hinge to the hinge whose internal moment is zero can be derived as

equation
Model of a multibody system with 3 sets of 2 concentric circles labeled j+n, j+2, and j, connected to bars labeled j+n–1, j+3, j+1, and j–1, with ellipsis between bars j+3 and j+n–1.

Figure 7.14 Model of a multibody system.

The transfer matrix is

where c7-i0177 are the transfer matrices of the outboard elements of the smooth hinge.

Because the internal moments in the state vectors c7-i0178 of the output end of the smooth hinge and the state vector c7-i0179 of the output end of the subsystem are all equal to zero, we can obtain the transfer equation and transfer matrix of the smooth hinge which are similar to the form of Equation (7.219), where, c7-i0180 are elements of the transfer matrix c7-i0181.

7.7 Transfer Matrices of Spatial Hinges

7.7.1 Elastic and Damper Hinges

The elastic damper hinge is composed of a linear spring hinge with a parallel connected damper, combining a rotary spring with a parallel connected damper. As shown in Figure 7.15, element c7-i0182 is the elastic joint and damper, whose inboard body and outboard body are rigid bodies j and c7-i0183, respectively. According to the force equilibrium at points O and I, we obtain

(7.221)equation

where

equation
Image described by caption and surrounding text.

Figure 7.15 Elastic and damper hinges.

Kx, Ky and Kz are the stiffness coefficients of the spring along the three axes of the inertial coordinate system, respectively, and Cd,x, Cd,y and Cd,z are the damper coefficients of the translation damper in the three axes of the inertial coordinate system, respectively.

Linearizing the velocity item yields

(7.222)equation

where

equation

and c7-i0184 is defined as in Equation (7.20).

For the rotational spring and damper we obtain

where

equation

The definitions of c7-i0185 and c7-i0186 can be seen in Equations (7.71) and (7.144). c7-i0187, c7-i0188 and c7-i0189 are the stiffness coefficients of the rotational spring along the three axes of the inertial system, respectively, and c7-i0190, c7-i0191 and c7-i0192 are the damper coefficients of the rotational damper about the three axes of the inertial system, respectively.

Linearizing Equation (7.223), we obtain

(7.224)equation

where

equation

The transfer equation of the spring joint and damper moving in space can be written as

equation

The transfer matrix is

(7.225)equation

7.7.2 Smooth Ball‐and‐socket Hinge whose Outboard is a Rigid Body and with Internal Torques at the Output End of Zero

For a smooth ball‐and‐socket hinge, using a similar method as for the smooth hinge moving in a plane, its mass and size can be neglected. The position coordinates and internal forces are equal and the internal moments are constant zero at its two ends, that is

The c7-i0193 and c7-i0194 denote the numbers of the inboard body and outboard body of the smooth ball‐and‐socket hinge j. For a smooth ball‐and‐socket hinge, if its outboard is a rigid body and the internal moment of the output end of this rigid body is zero, then by using the transfer equation of the outboard body of the smooth ball‐and‐socket joint, we obtain:

where c7-i0195, c7-i0196, c7-i0197 and c7-i0198 are submatrices of the transfer matrix of the smooth ball‐and‐socket joint’s outboard body. c7-i0199 is defined in Equation (7.20).

Substituting Equation (7.226) into Equation (7.227) yields

Combining Equations (7.226) and (7.228), the state vectors are derived by Equation (7.20), and the transfer equation of the smooth ball‐and‐socket joint can be derived as

(7.229)equation

The transfer matrix is

7.7.3 Smooth Ball‐and‐socket Hinge whose Outboard is a Rigid Body and with Nonzero Internal Torques at the Output End

If the internal moment of the output end of the outboard rigid body of the smooth ball‐and‐socket joint is nonzero, the transfer matrix and transfer equation of smooth ball‐and‐socket joint can be obtained by using the same method as in section 7.6.3. The form is the same as in Equation (7.220).

7.8 Algorithm of the Discrete Time Transfer Matrix Method for Multibody Systems

The algorithm to compute multi‐rigid‐body system dynamics using MSDTTMM is as follows:

  1. Determine the initial conditions and the boundary conditions of the system and set i = 1.
  2. Compute the coefficients A, Bz (or Bz), C, Dz (or Dz), c7-i0200, c7-i0201, c7-i0202, c7-i0203, c7-i0204, c7-i0205, c7-i0206 etc. at time ti for each connecting element by the linearization method, and determine the initial conditions c7-i0207, c7-i0208 and c7-i0209 at time ti as well as the system parameters.
  3. Formulate the transfer matrix for each subsystem and the overall transfer matrix at time ti, respectively.
  4. Apply the boundary conditions and the overall transfer equation to compute the unknown quantities in the boundary state vectors at time ti.
  5. By using the transfer equations of each element, compute the state vectors of each element at time ti.
  6. Use the state vectors at time ti to compute the first‐ and second‐order derivatives of the position coordinates and rotation angles with respect to time.
  7. Let c7-i0210, use the computational result of the last step as the initial conditions, and return to step (2). This procedure is repeated until the final time step is accomplished.

7.9 Numerical Examples of Multibody System Dynamics

7.9.1 Chain Multibody System Dynamics

For the chain multibody system shown in Figure 7.16, the transfer equations of hinge element j and body element c7-i0211 are

(7.231)equation

where c7-i0212 are the transfer matrices of hinge j and body c7-i0213.

Chain multibody system with hatched pattern connected to a series of 5 sets of 2 concentric circles connected to bars, with labels 1,2, 3, 4, j–1, j, j+1, n–1, n, and n+1.

Figure 7.16 Chain multibody system.

The overall transfer equation of the system can be assembled by only using the transfer matrices of these elements, that is

(7.232)equation

where the overall transfer matrix is

For a chain multibody system, the overall transfer matrix of the system can be obtained by successive multiplication of the transfer matrices of the elements. The order of the overall transfer matrix of the system is equal to the highest order of the transfer matrices of the elements, and it does not change as the number of DOFs in the system increases. For a chain multi‐rigid‐body system moving in a plane, the highest order of the transfer matrix is (7 × 7), while for a chain multi‐rigid‐body system moving in space, the highest order is (13 × 13).

Substituting the boundary conditions into the overall transfer equation of the system, the unknown quantities in the state vectors of the system boundaries can be computed. The state vectors and the motion quantities of each element at time ti can be computed by repeatedly using the corresponding transfer equations of elements. The quantities of velocity, angular velocity, acceleration and angular acceleration at time ti can be obtained. Repeating the entire procedure can allow the system motion to be obtained.

7.9.2 Numerical Example of a Chain Multi‐rigid‐body System Moving in a Plane

7.9.3 Numerical Example of a Chain Multi‐rigid‐body System Moving in Space

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

You can't read the all page of ebook, please click here login for view all page.
Reset
18.220.137.164