Abstract

Multibody distributed dynamic systems are seen in many engineering applications. Developed in this investigation is a new analytical method for a class of branched multibody distributed systems, which is called the augmented distributed transfer function (DTFM). This method adopts an augmented state formulation to describe the interactions among multiple distributed and lumped bodies, which resolves the problems with conventional transfer function methods in modeling and analysis of multibody distributed systems. As can be seen, the augmented DTFM, without the need for orthogonal system eigenfunctions, produces exact and closed-form solutions of various dynamic problems, in both frequency and time domains.

Introduction

Multibody distributed dynamic systems are seen in many engineering applications, such as buildings, bridges, railways, automobile suspension systems, light-weight deployable space structures, complex rotor systems, piping systems, and multilayer heat-conduction systems. The dynamics of multibody distributed systems thus have been extensively studied, and are of continued research interest today. For better understanding of the physical behaviors of these systems and for their optimal design and operation, both analytical and numerical tools for modeling and analysis have been in constant development. Among the available methods are the transfer matrix method [1], Lagrange multiplier method [2,3], wave propagation method [4,5], modal analysis [6], the spectral element method [7,8], the Rayleigh–Ritz method [9], the distributed transfer function method [10,11], and of course the finite element method (FEM).

Although analytical and semianalytical methods can produce accurate results and provide deep physical insight into the problem under investigation, most of them are limited to single-body distributed systems with simple configurations. For branched multibody distributed systems with complicate configurations, numerical tools such as the FEM often have to be used. The distributed transfer function method (DTFM), which combines the closed form of distributed transfer functions and the concept of finite element analysis, is capable of modeling certain types of multibody distributed systems [12,13]. This method, however, has been limited to frequency–domain solutions, mainly because of the singularities of the global dynamic equilibrium equation that is assembled from the transfer functions of distributed subsystems. Although the DTFM-based analysis has been recently extended to obtain exact transient solutions for stepped distributed systems [14,15], it is not directly applicable to general branched multibody systems.

The objective of the current investigation is to develop a new analytical method for branched multibody distributed dynamic systems that can systematically deliver closed-form solutions in both frequency and time domains. The branched multibody system in consideration is composed of one-dimensional distributed parameter subsystems, lumped parameter subsystems, and feedback controllers. These subsystems are interconnected at certain points that are called nodes. Each distributed subsystem is bounded by two nodes. Shown in Fig. 1 is an example of multibody systems (a flexible structure in vibration) from Ref. [12], which has 19 nodes, 14 distributed subsystems, two lumped subsystems (one oscillator at node 3, and one rigid-body mounted at nodes 10 and 11), and one feedback controller with a sensor at node 6 and an actuator at node 5. Nodes 1, 13, 14, 15, and 19 form the boundary of the structure. Additionally, a spring and a damper are placed at node 2, a distributed constraint (say elastic foundation) is imposed on the subsystem with nodes 7 and 8, and a viscoelastic bounding is reinforced between the subsystem with nodes 10 and 11 and the subsystem with nodes 16 and 17.

Fig. 1
Schematic of a branched multibody distributed system [12]
Fig. 1
Schematic of a branched multibody distributed system [12]
Close modal

Mathematically, the dynamic responses of branched multibody systems like the one shown in Fig. 1 are governed by a mixture of partial differential equations (for distributed subsystems) and ordinary differential equations (for lumped subsystems), which are coupled through multibody interactions (“force balance” and “displacement compatibility”) at interconnecting nodes. Because of the branched configuration, in which more than two distributed subsystems are interconnected at one node, single-body solutions by conventional analytical methods (such as eigenfunction expansion, the Rayleigh–Ritz method and transfer matrix method) cannot be applied herein. Moreover, a feedback controller with noncollocated sensors and actuators, which are also described by ordinary differential equations, can arbitrarily link several distributed subsystems at multiple nodes, rendering multibody interactions more complicated. In addition, the partial differential equations for distributed subsystems in general are nonself-joint due to damping, gyroscopic and circulatory forces, and other effects, yielding nonorthogonal eigenfunctions. These issues make the determination of analytical solutions for a branched multibody distributed system extremely difficult, and as a result, numerical methods almost always have to be used. In the literature, no analytical method is available for systematic determination of closed-form solutions of branched multibody distributed systems in both frequency and time domains.

The new modeling and analysis method proposed in this work is called the augmented DTFM. The novelty of this method lies in its utility of an augmented state formulation for description of multibody interactions, by which the global distributed transfer functions of the multibody distributed system are obtained. This approach does not need to assemble a global dynamic equilibrium equation from subsystem transfer functions, and as such totally eliminates the singularities with the previous DTFM [12]. Another advantage of the proposed method is that the augmented state formulation is flexible in describing the combination of distributed subsystems, lumped subsystems, and feedback controllers, which, as discussed previously, involves a set of coupled partial and ordinary differential equations. With the augmented DTFM, eigensolutions, frequency response, and transient time response of a wide class of branched multibody distributed systems can be systematically obtained in exact and closed form.

The remainder of the paper is organized as follows. A brief review of the previous DTFM in Sec. 2 reveals two major issues that have limited the utility of the method in dynamic analysis of multibody distributed systems. Presented in Sec. 3 is an augmented s-domain state formulation, which allows systematic description of the interactions of distributed and lumped bodies. Also presented in the section is a comparison study that explains why the issues with the previous DTFM can be resolved by the proposed method. Outlined in Sec. 4 are the solution formulas for dynamic problems of multibody distributed systems. The treatment of certain specific connections of distributed and lumped parameter bodies is given in Sec. 5. The proposed method is illustrated in Sec. 6, where the eigensolutions of two beam structures are computed.

Issues With the Previous DTFM

To understand the significance of this work, two issues with the previous distributed transfer function method (DTFM) are discussed in this section. To this end, some results in Refs. [11] and [12] are briefly reviewed.

Let the dynamic response w(x, t) of a uniformly distributed subsystem be governed by the nth order, linear partial differential equation
k=0n(ak2t2+bkt+ck)kxkw(x,t)=f(x,t),for0<x<1   and   t>0
(1)

where ak, bk, ck are constants representing the physical properties of the subsystem, such as inertia, damping, stiffness, gyroscopic effects, and circulatory forces; f(x, t) is an external disturbance; t is a temporal parameter; and x is a nondimensional, local spatial coordinate with x = 0 and 1 designating the two ends (nodes) of the subsystem. In a heat conduction problem, all coefficients ak that are related to the temporal operator ∂2/∂t2 are zeros. If the dynamic response of the subsystem involves multiple variables, such as the displacement components of a three-dimensional beam in vibration, Eq. (1) becomes a set of partial differential equations of similar format, with ak, bk, ck being matrices and w(x, t), f(x, t) being vectors.

Taking Laplace transform of Eq. (1) with respect to t and casting the resulting equation into a spatial state form, yields [11]
xη(x,s)=F(s)η(x,s)+q(s),0<x<1
(2)
where s is the complex Laplace transform parameter, η(x, s) is the state vector of the form
η(x,s)={w¯(x,s)w¯(x,s)xn-1w¯(x,s)xn-1}T
(3)
with w¯(x,s) being the Laplace transform of w(x, t), F(s) is an n-by-n matrix consisting of coefficients ak, bk, ck, and vector q(s) contains the Laplace transform of f(x, t) and the initial disturbances of the subsystem. The boundary and matching conditions of the subsystem at its two nodes can be expressed as
Mb(s)η(0,s)+Nb(s)η(1,s)=γb(s)
(4)

where Mb(s), Nb(s) are n-by-n matrices, and γb(s) is a vector consisting of the unknown nodal displacements of the subsystem and/or boundary disturbances.

The s-domain response of the distributed subsystem is obtained by solving the state equation (2) with the boundary condition (4) and it takes the form [11]
η(x,s)=01G(x,ξ,s)q(ξ)dξ+H(x,s)γb(s),0<x<1
(5)
Here matrices G(x, ξ, s) and H(x, s) are the distributed transfer functions of the subsystem, and they are in the exact and closed form
G(x,ξ,s)={H(x,s)Mb(s)e-F(s)ξ,ξx-H(x,s)Nb(s)eF(s)(1-ξ),ξ>xH(x,s)=eF(s)x[Mb(s)+Nb(s)eF(s)]-1
(6)
With the transfer functions of the distributed subsystems given in Eq. (6), imposing “force balance” and “displacement compatibility” at the nodes yields a global dynamic equilibrium equation of the multibody system in s domain [12]
K(s)U(s)=Q(s)
(7)

where K(s) is the global dynamic stiffness matrix that is assembled from the elements of the subsystem transfer functions, and U(s) and Q(s) are the global nodal displacement and nodal force vectors, respectively. In the formation of Eq. (7), which is called the distributed transfer function synthesis [12,13], no discretization or series truncation has been made.

Equation (7) can be applied to obtain solutions for various static and dynamic problems of multibody distributed systems. For instance, the eigenequation of a multibody distributed system is K(s) U(s) = 0, by which the system eigenvalues are the roots of the transcendental equation detK(λk)=0, k=1,2,3,.... The frequency response of the system subject to an excitation of frequency ω can be estimated by u(t)=K-1(Jω)q0eJωt, where J=-1, u(t) is the vector of nodal displacements, and q0 is a vector of force amplitudes. The solution of a static problem is found to be U(0)=K-1(0)Q(0), where s = 0 indicating that the temporal operators have been removed. In the above static and dynamic problems, substitution of the nodal displacements into Eq. (5) eventually gives the response of each distributed subsystem. Furthermore, with the transfer function formulation provided by Eqs. (5) and (7), stability analysis and design of distributed dynamic systems with feedback controllers can be performed [16].

Although the DTFM has found success in modeling, analysis, and control of distributed dynamic systems, it has two mathematical issues in dealing with multibody systems. The first issue is that the DTFM is inconvenient in transient analysis of a multibody system. According to Eq. (7), to estimate the system transient response, one needs to determine the nodal displacements via inverse Laplace transform of U(s)=K-1(s)Q(s). Because K(s) is assembled from the elements of G(x, ξ, s) and H(x, s) of the subsystems, the analytical expressions of inversed matrix K-1(s) in general are not available, which renders the inverse Laplace transform of K-1(s)Q(s) extremely difficult. Without the knowledge of the nodal displacements in time domain, determination of the transient response of the distributed subsystems by Eq. (5) is impossible. One may think about numerical evaluation of inverse Laplace transform, which, however, is impractical due to the second issue with the DTFM as follows.

The second issue with the DTFM is the singularities of G(x, ξ, s) and H(x, s), which are characterized by the poles of the transfer functions or the roots of the transcendental equation det[Mb(s)+Nb(s)eF(s)]=0; see Eq. (6). These roots are the eigenvalues of a “fully confined” subsystem, which in general are not the eigenvalues of the multibody system. The singularities of the transfer functions of distributed subsystems can cause the following three problems: it makes root finding for the characteristic equation detK(λk)=0 difficult and inefficient, and in some cases even misses roots; it may yield unbounded frequency response at a nonresonant excitation frequency, which is physically incorrect; and it renders numerical inverse Laplace transform of K-1(s)Q(s) erroneous and unreliable.

The above-mentioned issues have limited the utility of the DTFM for branched multibody distributed systems, especially in transient response analysis. It is upon this understanding that the current investigation introduces an augmented state formulation that can resolve these issues, as shown in the subsequent sections.

Augmented State Formulation

Assume that the multibody system in consideration has N uniformly distributed subsystems. The s-domain response of the subsystems is governed by the state equation
xη^i(x,s)=Fi(s)η^i(x,s)+pi(x,s),0<x<1,i=1,2,,N
(8)
where η^i,Fi(s),pi(x,s) are similarly defined as η(x,s),F(s),q(s) in Eq. (2). In the proposed method, the state equations for the N subsystems are cast into an augmented form
xη^(x,s)=F(s)η^(x,s)+p(x,s),0<x<1
(9)
where
η^(x,s)=(η^1(x,s)η^2(x,s):η^N(x,s)),p(x,s)=(p1(x,s)p2(x,s):pN(x,s)),F(s)=diag0kN{Fk(s)}
(10)

The dimension of the state vector η^(x,s) is given by n1+n2++nN, where ni is the dimension of η^i(x,s) for the ith distributed subsystem.

Now consider the boundary and matching conditions of the multibody system at its nodes. Denote a node of the ith subsystem by ξi*, the value of which can be either 0 or 1, depending on the chosen coordinate system. If ξi* is a boundary node of the multibody system, like node 19 in Fig. 1, the boundary conditions at the node are of the form
Γi(s)η^i(ξi*,s)=γ^i(s)
(11)
where Γi(s) is a matrix of proper dimensions, and γ^i(s) is a vector of boundary disturbances. If ξi* is a point where subsystems i,j,,k are interconnected, like node 12 in Fig. 1, the matching conditions at the node can be written as
Cii(s)η^i(ξi*,s)+Cij(s)η^j(ξj*,s)++Cik(s)η^k(ξk*,s)=γ^i(s)
(12)
where ξi*,ξj*,ξk* may take different values (either 0 or 1) but they all refer to the same node; matrices Cii(s),Cij(s),,Cik(s) describe the force balance and displacement compatibility at the node; and γ^i(s) is a vector of external disturbances at the node. When a feedback controller with an actuator at node ξi* of the ith subsystem and a sensor at node ξl* of the lth subsystem is implemented, like nodes 5 and 6 in Fig. 1, the actuation action at the ξi* is described through modification of Eq. (12) by
Cii(s)η^i(ξi*,s)+Cij(s)η^j(ξj*,s)++Cik(s)η^k(ξk*,s)+Gil(s)η^l(ξl*,s)=γ^i(s)
(13)
where the added matrix Gil(s) contains the transfer functions of the sensor, actuator, and controller. Here ξi* and ξl* may refer to two different nodes (for a noncollocated control system) or the same node (for a collocated control system). A feedback control system with multiple sensors and actuators can be treated similarly. Assembly of the boundary and matching conditions for all the subsystems yields the boundary conditions for the state equation (9)
MG(s)η^(0,s)+NG(s)η^(1,s)=γ^G(s)
(14)

where matrices MG(s),NG(s) are composed of those matrices in Eqs. (11) to (13), and vector γ^G(s) consists of the disturbances at the related nodes. Examples on boundary and matching conditions via this augmented form are given in Sec. 6.

It is easy to see that the augmented state equation (9) and boundary condition (14) for a multibody distributed system have the same form as Eqs. (2) and (4). Consequently, the s-domain response of the multibody distributed system is given by
η(x,s)=01G(x,ξ,s)p(ξ,s)dξ+H(s)γ^G(s),0<x<1
(15)
where G and H, the global distributed transfer functions of the multibody system, are given by
G(x,ξ,s)={H(x,s)MG(s)e-F(s)ξ,ξx-H(x,s)NG(s)eF(s)(1-ξ),ξ>xH(x,s)=eF(s)x[MG(s)+NG(s)eF(s)]-1
(16)

The state equation (9) subject to the boundary condition (14) and the solution (15) are the core of the augmented DTFM, which lays out a new foundation for modeling and analysis of branched multibody distributed dynamic systems. The augmented DTFM does not make any approximation or discretization. Neither does it rely on the orthogonality of system eigenfunctions. A comparison of the augmented DTFM with the previous DTFM [12] is given in Table 1. Different from the previous DTFM, the augmented DTFM does not need to form a global dynamic equilibrium equation like Eq. (7). Because of this, the singularities with the previous DTFM are completely avoided and inverse Laplace transform of Eq. (15) can be systematically and accurately performed without any difficulty [14].

Table 1

Comparison of the previous DTFM and the augmented DTFM

Previous DTFM [12]Proposed method (augmented DTFM)
Global (system-level) equationDynamic equilibrium equation K(s)U(s)=Q(s)Augmented state equation (9) with boundary condition (14)
Singularities in global equationYes, due to the singularities of K(s) that come from subsystem transfer functionsNo
Coordinates used to describe multibody interactionsGlobal coordinatesLocal coordinate (x) of distributed subsystems
Transfer functions of distributed subsystemsYes, given in Eq. (6)No
Transfer functions for the entire multibody systemNoYes, given in Eq. (15)
Characteristic equation for system eigenvaluesdetK(s)=0 with singularitiesdet[MG(s)+NG(s)e2F(s)]=0 continuous and smooth, without singularities
Root search for eigenvaluesLess efficient and possible to miss rootsHighly efficient and accurate
Transient analysis via inverse Laplace transformDifficult and erroneous due to the singularities in K(s)Yes, it can be performed systematically and accurately
Previous DTFM [12]Proposed method (augmented DTFM)
Global (system-level) equationDynamic equilibrium equation K(s)U(s)=Q(s)Augmented state equation (9) with boundary condition (14)
Singularities in global equationYes, due to the singularities of K(s) that come from subsystem transfer functionsNo
Coordinates used to describe multibody interactionsGlobal coordinatesLocal coordinate (x) of distributed subsystems
Transfer functions of distributed subsystemsYes, given in Eq. (6)No
Transfer functions for the entire multibody systemNoYes, given in Eq. (15)
Characteristic equation for system eigenvaluesdetK(s)=0 with singularitiesdet[MG(s)+NG(s)e2F(s)]=0 continuous and smooth, without singularities
Root search for eigenvaluesLess efficient and possible to miss rootsHighly efficient and accurate
Transient analysis via inverse Laplace transformDifficult and erroneous due to the singularities in K(s)Yes, it can be performed systematically and accurately

Solution of Dynamic Problems Via the Augmented DTFM

The augmented DTFM presented in Sec. 3 is applicable to various problems of branched multibody distributed systems. As pointed out previously, with similar formats, the results in DTFM analysis of single-body systems can be conveniently borrowed by the augmented DTFM for multibody systems.

Eigenvalue Problem.

The eigenvalue problem of a multibody distributed system by the augmented state formulation is described by the homogeneous equation [14]
xψ(x)=F(s)ψ(x),0<x<1
(17)
subject to the boundary condition
MG(s)ψ(0)+NG(s)ψ(1)=0
(18)
where s is an eigenvalue and ψ(x) is the associate eigenfunction. The system eigenvalues are the roots of the transcendental characteristic equation
det[MG(s)+NG(s)eF(s)]=0
(19)
The associate eigenfunctions have the form
ψ(x)=eF(s)xa
(20)
where a is a nonzero constant vector that is a solution of the homogeneous equation
[MG(s)+NG(s)eF(s)]a=0
(21)

Unlike detK(λk)=0 in the previous DTFM, the characteristic function in Eq. (19) is smooth and well-behaved, and does not contain any singularity at all. This shall also be demonstrated in a numerical example in Sec. 6.

Frequency Response.

For a multibody distributed system subject to a harmonic excitation p0(x)eJωt, where J=-1, ω is the excitation frequency, and vector p0(x) gives a spatial distribution of the excitation, its steady-state response by Eq. (15) is
ηss(x,t)=01G(x,ξ,Jω)p0(ξ)dξeJωt
(22)
If a boundary excitation γ0eJωt is applied, the steady-state response, again by Eq. (15), can be written as
ηss(x,t)=H(x,Jω)γ0eJωt
(23)

The G(x,ξ,Jω) and H(x,Jω) are the frequency response functions of the multibody system.

Transient Response.

The transient response of a branched multibody distributed dynamic system can be obtained via inverse Laplace transform of Eq. (15). This yields the following convolution integral:
η(x,t)=0t{01G(x,ξ,t-τ)p(ξ,τ)dξ+H(x,t-τ)γG(τ)}dτ,0<x<1
(24)
where the time-domain quantities η(x,t),p(ξ,t),γG(t) are the inverse Laplace transforms of η(x,s),p(ξ,s),γ^G(s), respectively; G(x,ξ,t),H(x,t) are the Green's functions of the multibody distributed system, and they are the inverse Laplace transforms of G(x,ξ,s),H(x,s). Following Ref. [14], the Green's functions can be expressed as
G(x,ξ,t)=k=1eF(sk)xRkD(x,ξ,sk)esktH(x,t)=k=1eF(sk)xRkeskt
(25)
where sk is the kth eigenvalue of the multibody system that is a root of the characteristic equation (19); Rk is the residue of [MG(s)+NG(s)eF(s)]-1 at sk; and
D(x,ξ,s)={Mbe-F(sk)ξ,ξx-NbeF(sk)(1-ξ),ξ>x
(26)

According to Ref. [14], an analytical formula for determination of the transfer function residues Rk is obtainable. Limited by space, the derivation of this formula is neglected. Note that the transient solution given herein avoids dealing with the inverse Laplace transform of K-1(s)Q(s) as in the previous DTFM (see Sec. 2), which is difficult to compute, and erroneous due to the singularities of K(s).

Treatment of Certain Types of Connection of Distributed Subsystems

Section 3 gives the general guidelines of the augmented DTFM. This section presents the treatment of three types of connection of distributed subsystems, which requires proper modification of the state equation (9) or the boundary condition (14).

Serially Connected Distributed Systems.

Consider an N-body serially connected system in Fig. 2, where X0,XN are the boundary nodes, and X1,X2,,XN-1 are the interior nodes with constraints represented by C1,C2,,CN-1. Distributed systems like this are also referred as stepped systems. The state equation of this type of systems is the same as Eq. (9). The boundary conditions at X0,XN are described by

Fig. 2
Distributed subsystems serially connected
Fig. 2
Distributed subsystems serially connected
Close modal
A0η^1(0,s)+ANη^N(1,s)=γ^b(s)
(27)
and the matching conditions at X1,X2,,XN-1 can be written as
Aiη^i(0,s)+Bi-1η^i-1(1,s)=νi(s),i=2,3,,N
(28)
where Ai and Bi are matrices describing the subsystem interactions and constraints, and γ^b(s),ν1(s),ν1(s),,νN(s) are the vectors of disturbances at the nodes. Equations (27) and (28) can be cast into the state form of Eq. (14), with
MG=[A0000A1::000AN-1],NG=[00ANB1000:00BN-10],γG=(γ^b(s)ν1(s):νN-1(s))
(29)

Pointwise Connection of Distributed Subsystems.

In Fig. 3, two distributed subsystems are connected at their nodes by some mechanism C (e.g., the spring connection between nodes 12 and 18 in Fig. 1). The state equation of this type of systems is the same as Eq. (9). The boundary and matching conditions of the subsystems are of the form

Fig. 3
Pointwise connection of two distributed subsystems
Fig. 3
Pointwise connection of two distributed subsystems
Close modal
Mb(s)η^1(0,s)+C11(s)η^1(1,s)+C12(s)η^2(0,s)=γb1(s)C21(s)η^1(1,s)+C22(s)η^2(0,s)+Nb(s)η^2(1,s)=γb2(s)
(30)

where Mb(s)andNb(s) specify the boundary conditions at nodes X0andX3, Cij(s) describe the connection mechanism C, and γb1(s) and γb2(s) represent prescribed boundary disturbances. Because Eq. (30) is a special case of Eq. (12), it can be cast in the form of boundary condition (14).

Distributed Connection of Two Distributed Subsystems.

In Fig. 4(a) two distributed subsystems (I) and (II) are connected throughout their domains. One example of such connection is seen in Fig. 1, where a viscoelastic bounding is reinforced between the subsystem of nodes 10 and 11 and the subsystem of nodes 16 and 17. This type of connection can also be used to model the vibration of carbon nanotubes [9]. Because the connection is imposed at the interior points of the subsystems, the state equation (9) has to be modified. (The boundary condition (14) remains the same.) Following a setting in Ref. [12], the state equations of the subsystems can be written as

Fig. 4
Distributed connection of two distributed systems
Fig. 4
Distributed connection of two distributed systems
Close modal
xη^i(x,s)=Fi(x,s)η^i(x,s)+pi(x,s)+Qci(x,s),0<x<1,i=1,2
(31)
where i=1 and 2 refers to subsystems (I) and (II), respectively; Qc1(x,s)andQc2(x,s) denote the constraint forces applied at the subsystems, in order to maintain the distributed connection, as shown in Fig. 4(b). Assume that the constraint forces are given by
(Qc1(x1,s)Qc2(x2,s))=[C11(s)C12(s)C21(s)C22(s)](η^1(x1,s)η^2(x2,s))
(32)
with matrices Cij(s) describing the connection. Substitution of Eq. (32) into Eq. (31) yields a coupled state equation
xη^c(x,s)=Fc(s)η^c(x,s)+pc(x,s)
(33)
where
Fc(s)=[F1(s)+C11(s)C12(s)C21(s)F2(s)+C22(s)]η^c(x,s)=(η^1(x,s)η^2(x,s)),pc(x,s)=(p1(x,s)p2(x,s))
(34)

Thus, to implement the distributed connection, Fc(s) is used to modify matrix F(s) in the state equation (9). Note that F(s) after the modification may not be a block-diagonal matrix.

Numerical Examples

The augmented DTFM is applied to compute the eigensolutions of two flexible structures in vibration: a coupled beam structure and a frame structure.

Example 1: A Coupled Beam Structure.

In Fig. 5, two cantilever Euler–Bernoulli beams of unity length are coupled by a spring of coefficient k at their tips (nodes X1 and X2). The governing equations of the beams are given by

Fig. 5
A coupled beam structure
Fig. 5
A coupled beam structure
Close modal
ρi2t2wi(x,t)+EIi4x4wi(x,t)=fi(x,t),0<x<1
(35)
where i=1and2,EIi and ρi are the bending stiffness and linear density of the beams, respectively, and the local coordinate x=xi has been used. The boundary and matching conditions of the system are
w1(0,t)=0,xw1(0,t)=0,w2(1,t)=0,xw2(1,t)=0,EI12x2w1(1,t)=0-EI13x3w1(1,t)=kw2(0,t)-kw1(1,t),-EI22x2w2(0,t)=0EI23x3w2(0,t)=kw1(1,t)-kw2(0,t)
(36)
Define the state vectors of the beam elements as
η^i(x,s)={w¯i(x,s)w¯i(x,s)x2w¯i(x,s)x23w¯i(x,s)x3}T
(37)
where w¯i(x,s) is the Laplace transform of wi(x,t) with respect to time t. Equations (35) and (36) are cast into the augmented state equations (9) and (14), in which
F(s)=[F1(s)00F2(s)],MG=[A000A1],NG=[0A2B10]
(38)
with
Fi(s)=[010000100001-ρis2EIi000]for   i=1,2A0=[1000010000000000],A1=[0000-k00000-EI20k00EI2]A2=[0000000010000100],B1=[00EI10k00-EI10000-k000]
(39)
By Eq. (19), the natural frequencies of the beam coupled structure are the roots of the characteristic equation
Δ(ω)det(MG+NGeF(Jω))=0,ω0,J=-1
(40)

With the nondimensional values of the system parameters given in Fig. 5, the first 10 natural frequencies of the coupled beam structure are computed by the proposed method and the finite element method. The results obtained are listed in Table 2. As can be seen, the results obtained by the augmented DTFM are the limits of the finite element result as the number N of elements increases.

Table 2

The natural frequencies of the coupled beams (rad/s)


Finite element method
kAugmented DTFMN = 8N = 16N = 20
127.899927.901127.900027.8999
249.402949.407449.403249.4030
3142.3247142.4970142.3365142.3296
4222.2395222.5032222.2576222.2470
5391.2485394.2897391.4877391.3486
6617.6240622.4137618.0005617.7816
7765.1783776.2695766.8945765.9085
81209.34911226.89961212.05961210.5022
91264.34181443.17741271.66011267.5303
101888.44282281.56821910.89781898.6379

Finite element method
kAugmented DTFMN = 8N = 16N = 20
127.899927.901127.900027.8999
249.402949.407449.403249.4030
3142.3247142.4970142.3365142.3296
4222.2395222.5032222.2576222.2470
5391.2485394.2897391.4877391.3486
6617.6240622.4137618.0005617.7816
7765.1783776.2695766.8945765.9085
81209.34911226.89961212.05961210.5022
91264.34181443.17741271.66011267.5303
101888.44282281.56821910.89781898.6379

For comparison purposes, the characteristic equation detK(Jω)=0 in the previous DTFM [12] is also considered, where K(s) is given in Eq. (7). In Fig. 6, the characteristic functions detK(Jω) and Δ(ω) are plotted against ω, where the zeros of the functions are the natural frequencies of the structure. As can be seen, detK(Jω) in Fig. 6(a) is piecewise continuous, and becomes unbounded at the certain frequencies. These frequencies are not the natural frequencies of the beam structure, but those of the beam elements that are “fully constrained,” that is, the beam elements with clamped-clamped boundaries. The singularities of detK(Jω)=0 can cause problems in root finding, as discussed in Sec. 2. On the other hand, Δ(ω) in Fig. 6(b) is continuous and smooth in the entire frequency region (0ω<+). A well-behaved Δ(ω) renders root search more efficient, and avoids the possibility of missing roots as with the previous DTFM. Moreover, due to its smoothness, Δ(ω) can be scaled to maintain relatively small values in computation. For instance, one may use Δ(ω)/(1+ωμ), with μ being a properly selected positive number.

Fig. 6
The characteristic functions of the coupled beam structure: (a) detK(Jω) in Ref. [12]; and (b) Δ(ω) in Eq. (40)
Fig. 6
The characteristic functions of the coupled beam structure: (a) detK(Jω) in Ref. [12]; and (b) Δ(ω) in Eq. (40)
Close modal

Example 2: A Frame Structure.

A frame of three beam elements is shown in Fig. 7, where the numbers in parentheses are the element numbers, X0,X1,X2,X3 are the nodes of the frame, and x is the nondimensional local coordinate of the elements. Let the ρi,EAi,EIi,Li (i = 1, 2, 3) be the linear density, longitudinal stiffness, bending stiffness, and length of the beam elements, respectively. The vibrations of the beam elements are governed by the partial differential equations

Fig. 7
A frame structure
ρi2t2ui(x,t)-EAiLi22x2ui(x,t)=fA,i(x,t)ρi2t2wi(x,t)+EIiLi44x4wi(x,t)=fT,i(x,t)
(41)
for 0<x<1, where ui(x,t) and wi(x,t) are the longitudinal and transverse displacements of the ith beam, respectively, fA,i(x,t)andfT,i(x,t) are external forces, and the 1/Li4 is associated with the spatial derivatives because the local coordinate x is nondimensional. Nine boundary conditions of the frame structure are prescribed at nodes X0,X2andX3
u1(0,t)=0,w1(0,t)=0,1L1xw1(0,t)=0u2(1,t)=0,w2(1,t)=0,1L2xw2(1,t)=0u3(1,t)=0,w3(1,t)=0,1L3xw3(1,t)=0
(42a)

With the assumption of rigid connection of the beam elements, there are nine matching conditions imposed at X1:

  • (i)
    Six conditions of displacement compatibility
    u1(1,t)=u2(0,t)cosα+w2(0,t)sinαw1(1,t)=w2(0,t)cosα-u2(0,t)sinαu1(1,t)=w3(0,t),w1(1,t)=-u3(0,t)1L1w1(1,t)x=1L2w2(0,t)x=1L2w3(0,t)x
    (42b)
  • (ii)
    Three conditions of force balance
    -EI1L122w1(1,t)x2+EI2L222w2(0,t)x2+EI3L322w3(0,t)x2=0-EA1L1u1(1,t)x+EA2L2u2(0,t)xcosα-EI2L233w2(0,t)x3sinα-EI3L333w3(0,t)x3=0EI1L133w1(1,t)x13-EA2L2u2(0,t)xsinα+EI2L233w2(0,t)x3cosα-EA3L1u3(0,t)x=0
    (42c)

Again, factors 1/Li,1/Li2,1/Li3 in the previous equations are related to the utility of the nondimensional coordinate x.

Define the state vectors of the beam elements as
η^i(x,s)={u¯i(x)u¯i(x)xw¯i(x,s)w¯i(x,s)x2w¯i(x,s)x23w¯i(x,s)x3}T
(43)
for i=1,2,3. With Eqs. (41) to (43), the state equation (9) and the boundary condition (14) can be easily formed. By the augmented state formulation, the eigensolutions of the frame structure can be determined through use of the formulas in Sec. 4.1. In numerical simulation, the following nondimensional values of the frame parameters are assigned:
Element(1):ρ1=78,EA1=2×107,EI1=2000,L1=1Element(2):ρ2=27,EA2=7×106,EI2=700,L2=1,α=π/4Element(3):ρ3=27,EA3=7×106,EI3=700,L3=2/2
(44)

The first eight natural frequencies of the frame structure are computed by the augmented DTFM, and tabulated in Table 3. For comparison, the results computed by the finite element method are also included in the table. It is seen that as the number of elements increases, the finite element predictions converge to the results given by the proposed method. Additionally, the first four mode shapes of the frame structure are computed by Eqs. (20) and (21), and they are plotted in Fig. 8.

Fig. 8
The first four mode shapes of the frame structure: (a) mode 1: ω1 = 83.524, (b) mode 2: ω2 = 113.05, (c) mode 3: ω3 = 206.44, and (d) mode 4: ω4 = 271.71
Fig. 8
The first four mode shapes of the frame structure: (a) mode 1: ω1 = 83.524, (b) mode 2: ω2 = 113.05, (c) mode 3: ω3 = 206.44, and (d) mode 4: ω4 = 271.71
Close modal
Table 3

The natural frequencies of the frame structure (rad/s)


Finite element method
kThe proposed methodN = 9N = 18N = 27
183.523683.706783.535483.5259
2113.0478113.5050113.0774113.0537
3206.4408207.4903206.5098206.4546
4271.7135276.2557272.0697271.7852
5307.5947313.7979308.1738307.7115
6519.4175562.7119521.7580519.9003
7570.5345628.6682573.3994571.1386
8620.1257700.5947622.6839620.6544

Finite element method
kThe proposed methodN = 9N = 18N = 27
183.523683.706783.535483.5259
2113.0478113.5050113.0774113.0537
3206.4408207.4903206.5098206.4546
4271.7135276.2557272.0697271.7852
5307.5947313.7979308.1738307.7115
6519.4175562.7119521.7580519.9003
7570.5345628.6682573.3994571.1386
8620.1257700.5947622.6839620.6544

Conclusions

This paper presents a new analytical method, namely the augmented distributed transfer function method, for modeling and analysis of one-dimensional multibody distributed dynamic systems. The proposed method has the following special features.

  1. (a)

    By using an augmented state formulation, Eqs. (9), (14), and (15), the proposed method is flexible and convenient in description of interactions of multiple distributed and lumped bodies that are governed by a mixture of partial and ordinary differential equations. With this setting, various dynamic problems of branched multibody systems can be systematically solved.

  2. (b)

    The augmented DTFM does not use dynamic equilibrium equations that are assembled from the transfer functions of distributed subsystems. Because of this, the singularity problems with the previous DTFM and the transfer matrix method are completely eliminated. Indeed, the characteristic function in Eq. (19) is smooth and well-behaved for any value of s, which renders root finding efficient, and avoids the problem of missing roots in computation. Due to this advantage of the augmented DTFM over conventional transfer function methods, inverse Laplace transform for time–domain solutions can be accurately and easily performed.

  3. (c)

    In modeling and analysis, the augmented DTFM does not make any approximation or discretization; neither does it require orthogonal system eigenfunctions. To the authors' knowledge, the augmented DTFM is the first solution technique that delivers exact and closed form solutions for a wide class of branched multibody distributed systems, in both frequency and time domains.

  4. (d)

    In the proposed method, modeling of different systems with different boundary and matching conditions just needs simple formation of the state matrix F(s) in Eq. (9) and the boundary matrices MG(s)andNG(s) in Eq. (14); solution procedures and algorithms are essentially same. This feature of symbolic manipulation renders the proposed method easy in use and convenient for computer coding, as has been shown in the numerical examples.

References

1.
Prestel
,
E. C.
, and
Leckie
,
F. A.
,
1963
,
Matrix Methods in Elastomechanics
,
McGraw-Hill
,
New York
.
2.
Dowell
,
E. H.
,
1972
, “
Free Vibration of an Arbitrary Structure in Terms of Component Modes
,”
ASME J. Appl. Mech.
,
39
, pp.
727
732
.10.1115/1.3422780
3.
Hallquist
,
J.
, and
Snyder
,
V. W.
,
1973
, “
Linear Damped Vibratory Structures With Arbitrary Support Conditions
,”
ASME J. Appl. Mech.
,
40
, pp.
312
313
.10.1115/1.3422956
4.
Mead
,
D. J.
,
1971
, “
Vibration Response and Wave Propagation in Periodic Structures
,”
ASME J. Eng. Ind.
,
93
(
3
), pp.
783
792
.10.1115/1.3428014
5.
Mead
,
D. J.
, and
Yaman
,
Y.
,
1990
, “
The Harmonic Response of Uniform Beams on Multiple Linear Supports: A Flexural Wave Analysis
,”
J. Sound Vib.
,
141
, pp.
465
484
.10.1016/0022-460X(90)90639-H
6.
Bergman
,
L. A.
, and
Nicholson
,
J. W.
,
1985
, “
Forced Vibration of a Damped Combined Linear System
,”
ASME J. Vib. Acoust. Stress Reliability
,
107
(
3
), pp.
275
281
.10.1115/1.3269257
7.
Doyle
,
J. F.
,
1997
,
Wave Propagation in Structures: Spectral Analysis Using Fast Discrete Fourier Transforms
, 2nd ed.,
Springer
,
New York
.
8.
Lee
,
U.
,
2009
,
Spectral Element Method in Structural Dynamics
,
John Wiley and Sons
,
Singapore
.
9.
Kelly
,
S. G.
, and
Srinivas
,
S.
,
2009
, “
Free Vibrations of Elastically Connected Stretched Beams
,”
J. Sound Vib.
,
326
, pp.
883
893
.10.1016/j.jsv.2009.06.004
10.
Butkoviskiy
,
A. G.
,
1983
,
Structural Theory of Distributed Systems
,
Halsted, John Wiley and Sons
,
New York
.
11.
Yang
,
B.
, and
Tan
,
C. A.
,
1992
, “
Transfer Functions of One-Dimensional Distributed Parameter Systems
,”
ASME J. Appl. Mech.
,
59
, pp.
1009
1014
.10.1115/1.2894015
12.
Yang
,
B.
,
1994
, “
Distributed Transfer Function Analysis of Complex Distributed Parameter Systems
,”
ASME J. Appl. Mech.
,
61
, pp.
84
92
.10.1115/1.2901426
13.
Yang
,
B.
,
2005
,
Stress, Strain, and Structural Dynamics: An Interactive Handbook of Formulas, Solutions, and MATLAB Toolboxes
,
Elsevier Science
,
Boston
.
14.
Yang
,
B.
,
2010
, “
Exact Transient Vibration of Stepped Bars, Shafts and Strings Carrying Lumped Masses
,”
J. Sound Vib.
,
329
, pp.
1191
1207
.10.1016/j.jsv.2009.10.035
15.
Yang
,
B.
, and
Noh
,
K.
,
2012
, “
Exact Transient Vibration of Non-Uniform Bars, Shafts and Strings Governed by Wave Equations
”,
Int. J. Struct. Stability Dyn.
,
12
(
4
), p.
1250022
.10.1142/S0219455412500228
16.
Yang
,
B.
, and
Mote
,
C.D.
, Jr.
,
1991
, “
Frequency–Domain Vibration Control of Distributed Gyroscopic Systems
,”
ASME J. Dyn. Syst. Measure. Control
,
113
(
1
), pp.
18
25
.10.1115/1.2896350