You are on page 1of 20

Arch Appl Mech (2014) 84:421–440

DOI 10.1007/s00419-013-0809-7

O R I G I NA L

M. Faghih Shojaei · R. Ansari · V. Mohammadi · H. Rouhi

Nonlinear forced vibration analysis of postbuckled beams

Received: 25 November 2012 / Accepted: 27 November 2013 / Published online: 19 December 2013
© Springer-Verlag Berlin Heidelberg 2013

Abstract A numerical solution methodology is proposed herein to investigate the nonlinear forced vibrations
of Euler–Bernoulli beams with different boundary conditions around the buckled configurations. By introduc-
ing a set of differential and integral matrix operators, the nonlinear integro-differential equation that governs
the buckling of beams is discretized and then solved using the pseudo-arc-length method. The discretized
governing equation of free vibration around the buckled configurations is also solved as an eigenvalue prob-
lem after imposing the boundary conditions and some complicated matrix manipulations. To study forced and
nonlinear vibrations that take place around a buckled configuration, a Galerkin-based numerical method is
applied to reduce the partial integro-differential equation into a time-varying ordinary differential equation
of Duffing type. The Duffing equation is then discretized using time differential matrix operators, which are
defined based on the derivatives of a periodic base function. Finally, for any given magnitude of axial load, the
pseudo -arc-length method is used to obtain the nonlinear frequencies of buckled beams. The effects of axial
load on the free vibration, nonlinear, and forced vibrations of beams in both prebuckling and postbuckling
domains for the lowest three vibration modes are analyzed. This study shows that the nonlinear response
of beams subjected to periodic excitation is complex in the postbuckling domain. For example, the type of
boundary conditions significantly affects the nonlinear response of the postbuckled beams.

Keywords Beam · Nonlinear forced vibration · Buckled configuration · Numerical solution methodology

List of symbols

N Number of grid points in space domain


Nt Number of grid points in time domain
xi Chebyshev–Gauss–Lobatto grid points
(1) (2)
Dt and Dt Time differentiation matrix operators
(r )
Dx Space differentiation matrix operator
S̃x and Sx Integral matrix operator
 Length of beam
m Mass per unit length
A Cross-section area

M. Faghih Shojaei · R. Ansari (B) · V. Mohammadi · H. Rouhi


Department of Mechanical Engineering, University of Guilan, P.O. Box 41635-3756, Rasht, Iran
E-mail: r_ansari@guilan.ac.ir
Tel.: +98-131-6690276
Fax: +98-131-6690276
422 M. Faghih Shojaei et al.

E Young’s modulus
I Moment of inertia
ŵ and w Transverse displacement and its dimensionless form
μ̂ and μ Damping coefficient and its dimensionless form
P̂ and P Axial load and its dimensionless form
F̂(x̂) and F(x) Transverse load and its dimensionless form
ˆ and  Excitation frequency and its dimensionless form
ω Natural frequency
q(t) Vibration amplitude
v(x, t) Dynamic part of field variable w(x, t)
ψ(x) Static part of field variable w(x, t)
φ (x) Normalized vector of the linear vibration
mode shape around the postbuckling config-
uration
T Period of response

1 Introduction

The free vibration and buckling of beams have been studied by many investigators. Also, the study on the forced
vibrations of buckled beams has received considerable attention from the scientific community for nearly 50
years. Eisley [1] in 1964 investigated the problem of nonlinear forced vibrations of buckled beams. Since then,
many investigators have studied the problem of forced vibration of beams around the buckled configurations.
In 1970, Tseng [2] and Eisley and Bennett [3] analyzed the nonlinear forced vibrations of buckled beams.
Tseng and Dugundji [4] conducted both analytical and experimental investigations into the nonlinear vibrations
of a buckled clamped beam excited by the harmonic motion of its supporting base. They used the Galerkin
method to reduce the governing partial differential equation into a modified Duffing equation, which was then
solved by the harmonic balance method. Min and Eisley [5] studied the steady-state response and stability of
forced vibration of axially restrained buckled beams under simply supported boundary conditions. Holmes
[6] investigated the sinusoidally forced vibrations of a buckled beam. Pezeshki and Dowell [7] examined the
initial condition maps for the forced vibrations of buckled beams via the forced Holmes-Duffing’s equation.
The dynamics of a magnetically buckled beam excited by a two-frequency forcing function was investigated
by Pezeshki et al. [8]. Higuchi and Dowell [9] analyzed the chaotic oscillation of a buckled beam subjected
to sinusoidally varying and static constant transverse external forces. Afaneh and Ibrahim [10] analytically,
numerically, and experimentally studied the nonlinear modal interaction in an initially buckled beam with
clamped boundary conditions, under harmonic excitation in the vicinity of a one-to-one internal resonance.
The response of a parametrically excited buckled beam was investigated by Abou-Rayan et al. [11]. Ji and Han-
sen [12] carried out an experimental investigation into the nonlinear response of a clamped–sliding postbuckled
beam under a harmonic axial load for fundamental and subharmonic resonances. Emam [13] conducted both
experimental and theoretical investigations to study the nonlinear dynamics of buckled beams. Virgin and Plaut
[14] studied the effects of axial load on the frequency response of beams with fixed ends. Emam and Nayfeh
[15] investigated the nonlinear response of a clamped–clamped buckled beams using the Galerkin method in
order to discretize the partial differential equation and corresponding boundary conditions. El-Bassiouny [16]
presented an analytical investigation into the nonlinear vibrations of a beam subjected to harmonic excitations
around the postbuckling configurations. Nagai et al. [17] conducted experiments on chaotic vibrations of a
postbuckled beam subjected to periodic lateral acceleration. The dynamics of a buckled beam under high-fre-
quency excitation were studied by Yabuno and Okada [18]. Zhang et al. [19] analyzed the steady-state periodic
response of an axially moving viscoelastic beam in the supercritical speed range. Kazemirad and co-workers
[20] investigated the thermomechanical nonlinear vibrations and stability of a buckled axially moving beam
using the pseudo-arc-length continuation method. Recently, Ghayesh and Amabili [21] studied the nonlinear
forced dynamics of an axially moving viscoelastic beam in the supercritical speed domain.
In this article, the free, nonlinear, and forced vibrations of buckled Euler–Bernoulli beams subject to vari-
ous types of boundary conditions are numerically studied. The effect of axial load on the frequency response
(i.e., the variation in vibration amplitude against excitation frequency) in both prebuckling and postbuckling
domains with considering all nonlinear terms for the lowest three vibration mode shapes is comprehensively
Nonlinear forced vibration analysis of postbuckled beams 423

investigated. The numerical solution method developed in this work is quite general so that it can be applied
to other types of structures with arbitrary boundary conditions. It is briefly described below.
First, based on the generalized differential quadrature (GDQ) method and Taylor series, a set of differential
and integral matrix operators are introduced for discretizing the space domain. Also, based on the derivatives
of a periodic base function, spectral differentiation matrix operators are introduced in order to discretize the
time domain. Using the space differential and integral matrix operators, the nonlinear integro-differential gov-
erning equation and associated boundary conditions for the buckling of beams are discretized and then solved
via the pseudo-arc-length technique. The discretized linear governing equation of free vibration around the
buckled configurations is solved as an eigenvalue problem after imposing the end conditions and some lengthy
matrix manipulations. To investigate the forced and nonlinear vibrations of buckled beams, a Galerkin-based
numerical method is first used to reduce the time- and space-varying partial integro-differential governing
equation into a time-varying ordinary differential equation of Duffing type. Emam [13] reported that using
lower than four free vibration modes for the Galerkin method leads to considerable errors in the predicted
frequency-response curves. In this work, rather than the summation of free vibration modes, only one linear
vibration mode shape in the postbuckling state is used as a weight function for minimizing the residual result-
ing from the pseudo-Galerkin method, since the effect of axial load is directly incorporated into the vibration
mode shape of the postbuckled beam. Thus, the computational efficiency as well as the solution accuracy is
significantly increased in the present study by using only one vibration mode of postbuckled beam. The time
periodic differential matrix operators are employed so as to discretize the Duffing equation, which is then
solved by the pseudo-arc-length method for any given magnitude of applied axial load. It is worth mentioning
that for the forced vibration of buckled beams, the problem is directly solved, whereas a normalizing equation
must be added to the system for solving the problem of nonlinear vibration around the buckled configurations
as a nonlinear eigenvalue problem.

2 Differential matrix operators

In the GDQ method [22], the r th derivative of the function f (x) is approximated by a linear sum of the function
values,

∂ r f (x)  N
(r )
= W i j f (x j ), (1)
∂ xr  x=xi j=1

where r is the order of differentiation in the x direction; Wi(rj ) are the weighting coefficients; and N is the
number of total discrete grid points. By introducing a column vector F as
    T
F = Fj = f x j = { f (x1 ) , f (x2 ) , . . . , f (x N )}T (2)

in which f (x j ) is the nodal value of f (x) at x j , one can define the following differential matrix operator

∂r 
(F) = Drx F = Dxr {F j } (3)
∂x r

(r )
where the differential operator Dx is expressed as
(r )
D(r )
x = Wi j , i, j = 1, 2, . . . , N (4)

The weighting coefficients of the first derivative and higher-order derivatives are obtained by means of the
following relation

⎪ Ix , where Ix is a N × N identitymatrix, r = 0

⎪ N  

⎪ P (xi )
⎨ (xi −x j )P (x j ) , where P (xi ) = j=1; j=i xi − x j , i  = j and i, j = 1, . . . , N and

(r )
Wi j =   (5)
⎪ (1) (r −1) Wi(rj −1)

⎪ r Wi j Wii − , i  = j and i, j = 1, . . . , N and r = 2, 3, . . . N − 1

⎪ xi −x j

⎩ N (r )
− j=1; j=i Wi j , i = j and i, j = 1, . . . , N and r = 1, 2, 3, . . . N − 1
424 M. Faghih Shojaei et al.

Also, the following shifted Chebyshev–Gauss–Lobatto grid points are used to generate the mesh in the x
direction  
1 i −1
xi = 1 − cos π , i = 1, 2, . . . , N (6)
2 N −1

3 Time differential matrix operators on a periodic grid

The derivatives of periodic sinc function as base function in spectral collocation method can be employed
to define highly accurate time differentiation matrix operators in order to discretize the time domain [23].
It should be noted that such operators have excellent performance for periodic functions, whereas they are
inappropriate for nonperiodic ones. The reason is that these operators only operate on an unbounded periodic
grid on which only functions with fix periodicity are authorized. Owing to differentiation of a periodic base
function, these operators naturally satisfy the following conditions
 
d f  d f 
f (0) = f (T ) , = (7)
dt  dt  t=0 t=T
Note that T is the period of function f . The periodic grid points are generated as
i
ti = , 0 < ti ≤ 1, i = 1, 2, . . . , Nt = 2k (8)
Nt
in which Nt is the number of grid points on time domain and must be an even number. The explicit formulation
for differential matrix operators for 0 < t ≤ 1 is given by

⎪ a =0
⎪ 11
⎨ ai,1 = (−1)i−1 cot π(i−1) (1)
π(Nt − j+1) , i, j = 2, 3, 4, . . . , Nt , Dt = 2π[ai, j ]
Nt
Nt − j+1 (9a)

⎪ a = (−1) cot
⎩ 1, j Nt
ai+1, j+1 = ai, j

N2

⎪ b11 = − 12t − 6
1



⎨ bi,1 = (−1)
i−1

2 sin2 π(i−1) (2)


Nt , i, j = 2, 3, 4, . . . , Nt , Dt = (2π)2 [bi, j ] (9b)

⎪ = (−1) Nt − j+1

⎪ b 1, j π(N −
⎪ 2 t j+1)
2 sin
⎩ Nt
bi+1, j+1 = bi, j
Also, higher-order derivatives can be obtained using the following formula
(r ) (1) (r −1)
Dt = Dt Dt (10)
(1) (2)
where Dt and Dt are Toeplitz matrices.

4 Integral matrix operator

The Taylor series of function f (x) is written as


∞
f (r ) (xi )
f (x) = (x − xi )r (11)
r!
r =0
 
By integrating analytically from Eq. (11) over the subinterval xi−12+xi , xi +x2 i+1 , one can arrive at the
following formula for the integral of f (x) from x1 to x N
x N ∞
 −1
 (x2 − x1 )r +1 (r ) 
N
(xi+1 − xi )r +1 − (xi−1 − xi )r +1 (r )
f (x) dx = f (x 1 ) + f (xi )
2r +1 (r + 1)! 2r +1 (r + 1)!
x1 r =0 i=2

(x N −1 − x N )r +1 (r )
− r +1 f (x N ) (12)
2 (r + 1)!
Nonlinear forced vibration analysis of postbuckled beams 425

By introducing the row vector X̃(r ) and column vector F(r ) as


 
  (x − x )r +1
(x − x )r +1
− (x − x )r +1
(x − x )r +1
(r ) −1
X̃(r ) = X i
2 1 i+1 i i−1 i N N
= , , − r +1 ,
2r +1 (r + 1)! 2r +1 (r + 1)! 2 (r + 1)!
i = 2, 3, . . . , N − 1 (13a)
  T  (r ) T
F(r ) = f (r ) x j = f (x1 ) , f (r ) (x2 ) , . . . , f (r ) (x N ) (13b)

in which the f (r ) (xi ) is the nodal value of r th derivative of f (x) at xi , one can obtain
x N ∞

f (x) dx = X̃(r ) F(r ) (14)
x1 r =0

By substituting Eq. (5) into Eq. (14), one can define S̃x as a matrix operator for approximating the integral
value as follows
x N  N −1 

N −1   
(r ) (r ) (r ) (r )
f (x) dx = X̃ Dx F = X̃ Dx F = S̃x F, S̃x = S̃x (15)
1×N
x1 r =0 r =0

S̃x F gives the integral value. In order to discretize the integral value over the domain, a column vector
O = {O} N ×1 = {1, 1, . . . , 1}T is used as
⎛ x ⎞  N −1 
N 
O⎝ f (x) dx ⎠ = OS̃x F = O (r ) (r )
X̃ Dx F = Sx F, Sx = [Sx ] N ×N (16)
x1 r =0

5 Problem formulation

Figure 1 shows a beam with length , mass per unit length m, cross-section A, Young’s modulus E, and moment
ˆ For this beam, the trans-
of inertia I , under axial load P̂ and transverse load F̂(x̂) with the frequency of .
verse vibration taking into account mid-plane stretching is governed by the following nonlinear Fredholm-type
integro-differential equation [24]

  2
∂ 4 ŵ ∂ 2 ŵ ∂ ŵ ∂ 2 ŵ E A ∂ 2 ŵ ∂ ŵ
E I 4 + m 2 + μ̂ + P̂ 2 − ˆ tˆ
d x̂ = F̂(x̂) cos  (17)
∂ x̂ ∂ tˆ ∂ tˆ ∂ x̂ 2 ∂ x̂ 2 ∂ x̂
0

Fig. 1 Schematic view of a clamped–clamped beam subjected to axial and transverse loads
426 M. Faghih Shojaei et al.

 
where ŵ x̂, tˆ is the transverse displacement at position x̂ and time tˆ, and μ̂ is the damping coefficient. The
boundary conditions are expressed as

Simply supported–simply supported (SS):

∂ 2 ŵ
ŵ = = 0 at x̂ = 0,  (18)
∂ x̂ 2
Clamped–clamped (CC):
∂ ŵ
ŵ = = 0 at x̂ = 0,  (19)
∂ x̂
Simply supported-clamped (SC):

∂ 2 ŵ
ŵ = = 0 at x̂ = 0, (20a)
∂ x̂ 2
∂ ŵ
ŵ = = 0 at x̂ =  (20b)
∂ x̂
By introducing the following nondimensional variables
! "
x̂ ŵ m4 EI
x = , w= , = ˆ , t =t ˆ (21)
 r EI m4

in which r = I /A is the radius of gyration of the cross-section, Eq. (17) can be rewritten as

1
1 
ẅ + w + Pw + μẇ − w 
iv 
w 2 dx = F(x) cos t (22)
2
0

Also, the boundary conditions given by Eqs. (18)–(20), respectively, are rewritten as

w = w  = 0 at x = 0, 1 (23)
w = w  = 0 at x = 0, 1 (24)

w = w  = 0 at x = 0 (25a)
w = w  = 0 at x = 1 (25b)

where overdot shows derivative with respect to t and prime stands for derivative with respect to x. Moreover,
μ, P, and F in Eq. (22) are given by

μ̂2 P̂2 F̂4


μ= √ , P= , F= (26)
mEI EI E Ir

6 Postbuckling problem

To study the postbuckling configurations of beam, the time-dependent terms are neglected in Eq. (22) and w
is denoted by ψ. A column vector  is introduced as follows

 = {ψ1 , ψ2 , . . . , ψ N }T (27)

with N entries ψi = ψ (xi ). First, the governing equation and boundary conditions are discretized by the
differential and integral matrix operators introduced in Sects. 2 and 4. The discretized form of Eq. (22) is
  #
1  (1) $
D(4)
x  + P − S x (D x )◦(D (1)
x ) ◦ D(2)
x  =0 (28)
2
Nonlinear forced vibration analysis of postbuckled beams 427

where ◦ denotes the Hadamard product (see the “Appendix”). Also, considering Eq. (27), the SS, CC, and SC
boundary conditions are, respectively, rewritten as
ψ1 = ψ N = 0 and ψ1 = ψ N = 0 (29)
ψ1 = ψ N = 0 and ψ1 = ψ N = 0 (30)
ψ1 = ψ N = 0 and ψ1 = ψ N = 0 (31)

Note that   = D(1)  (2)


x  and  = Dx . The set of nonlinear equations of the domain, Eq. (28), can be
shown as
F : R N +1 → R N , F (, P) = 0 (32)
where P is the applied axial load. Eq. (32) is solved using the pseudo-arc-length continuation method [25].
It should be noted that with the purpose of satisfying the boundary conditions, the residual of equations of
boundaries must be substituted into the residual vector of domain during the nonlinear solution procedure.
To accomplish this aim, the elements of the residual vector of domain equivalent to the grid points at the
boundaries are replaced with the residual values of boundaries which are obtained from one of the equations
of (29), (30), or (31).

7 Free vibration around the buckled configurations

To investigate vibrations taking place in the vicinity of a buckled equilibrium position, a small disturbance
is considered around the buckled configurations of the beam. Accordingly, the field variable of governing
equation can be split into two static and dynamic parts as follows
w (x, t) = ψ (x) + v(x, t) (33)
where v(x, t) is a small dynamic disturbance around the buckled configuration ψ (x). Substituting Eq. (33)
into Eq. (22), one can obtain [24]

1 1 1 1
2   1 2 1 
v̈ + v + λ v + μv̇ = ψ
iv
ψ v dx + ψ 
 
v dx + v 
ψ v dx + v 
 
v 2 dx + F cos t (34)
2 2
0 0 0 0

where
1
1
λ =P−
2
ψ 2 dx (35)
2
0
Further details about the derivation of Eq. (34) can be found in [24]. Also, the boundary conditions can be
written as
v = v  = 0 at x = 0, 1 (36)
v = v  = 0 at x = 0, 1 (37)
v = v  = 0 at x = 0 (38a)
v = v  = 0 at x = 1 (38b)
for SS, CC, and SC beams, respectively. By neglecting the nonlinear, forcing and damping terms, and substi-
tuting v (x, t) = φ (x) eiωt into Eq. (34), one can arrive at

1
2  
φ iv
+λ φ −ω φ =ψ 2
ψ  φ  dx (39)
0

where ω is the natural frequency and φ (x) is the corresponding mode shape. A column vector  is introduced
as
 = {φ1 , φ2 , . . . , φ N }T (40)
428 M. Faghih Shojaei et al.

Table 1 Natural frequencies of free vibration of SS, CC, and SC beams at p = 6π 2 in the postbuckling domain for the first
vibration mode shape obtained by different grid numbers based on GDQ and HDQ methods

Grid number SS CC SC
GDQ HDQ GDQ HDQ GDQ HDQ
9 31.2109 31.2104 22.3846 22.3195 27.4283 27.5011
11 31.2104 31.2104 22.2517 22.2639 27.4615 27.4573
13 31.2104 31.2104 22.2559 22.2570 27.4593 27.4568
15 31.2104 31.2104 22.2558 22.2560 27.4593 27.4584
17 31.2104 31.2104 22.2558 22.2558 27.4593 27.4591
19 31.2104 31.2104 22.2558 22.2558 27.4593 27.4593

SS CC

N=9 1 N=9
0.6
N = 11 N = 11
Vibration amplitude

Vibration amplitude
0.5 N = 13 0.8 N = 13
N = 15 N = 15
0.4 N = 17 N = 17
0.6
N = 19 N = 19
0.3
0.4
0.2
0.2
0.1

25 30 35 15 20 25 30

SC
0.8
N=9
0.7
N = 11
0.6
Vibration amplitude

N = 13
N = 15
0.5
N = 17
0.4 N = 19
0.3

0.2
0.1

20 25 30 35

Fig. 2 Nonlinear frequency response of SS, CC, and SC beams at p = 6π 2 in the postbuckling domain for the first vibration
mode shape obtained by different grid numbers. (Color figure online)

with N entries φi = φ (xi ). After some appropriate matrix manipulations, the discretized form of Eq. (39) is
written as
   #  $
1 # (1) $ # (1) $ # $$ #
D(4)
x + D (2)
x P − S x D x  ◦ D x  − S x D(1)
x D(1)
x  D (2)
x   = ω2 
2
(41)
where is a matrix product whose definition is given in the “Appendix.”
After imposing the boundary conditions on the discretized governing equation, the natural frequencies
and corresponding vibration mode shapes of the beam around the buckled configurations can be extracted. It
should be noted that in order to apply boundary conditions, their equations are directly substituted into the
equations of system at boundaries.
Nonlinear forced vibration analysis of postbuckled beams 429

0.3
2
P = -1
2
P=0
2
0.25 P=1
2
P=2
2
P=3
Vibration amplitude /f 0
2
P=4
0.2

0.15

0.1

0.05

0
0 5 10 15 20 25 30

Fig. 3 Nonlinear frequency response of a CC beam under different prebuckled positions for the first vibration mode shape obtained
by the present method (shown by solid lines) and obtained in [14] (shown by solid circles). (Color figure online)

8 Forced and nonlinear vibrations around the buckled configurations

Here, a Galerkin-based numerical method is employed to reduce the nonlinear integro-differential equation
of (34) into a time-varying ordinary differential equation of Duffing type. Rather than the summation of free
vibration mode shapes, only one linear vibration mode shape of postbuckled beams can be used as a weight
function for minimizing the residual resulting from the pseudo-Galerkin approach. This idea considerably
increases the solution accuracy, since the effect of applied axial load is directly incorporated into the vibration
mode shape of the postbuckled beam. Substituting v (x, t) = φ (x) q(t) into Eq. (34) and then discretizing on
the x domain, the residual is given by
   #
1 # (1) $ # (1) $ $
R = q̈ + μq̇ + D(4)
x  + P − S x D x  ◦ D x  ◦ D(2)
x 
2
# $ # # $ # $$
− D(2)x  ◦ Sx D(1)
x  ◦ Dx 
(1)
q
 # $ # # $ # $$ # $ # # $ # $$
1 (2) (1) (1) (2) (1) (1)
− Dx  ◦ Sx Dx  ◦ Dx  + Dx  ◦ Sx Dx  ◦ Dx  q2
2
1 # (2) $ # # (1) $ # (1) $$ 3
− Dx  ◦ Sx Dx  ◦ Dx  q (42)
2

where overdot denotes derivative with respect to time, and  is the normalized vector of the linear vibration
mode shape around the postbuckling configuration. %
In the Galerkin method, the residual is orthogonal to mode shapes, i.e., φ (x) R(x, t) = 0 or in the
numerical form
# $
S̃x ( ◦ R) = S̃x T R = 0 (43)

By inserting the residual into Eq. (43), the following ordinary differential equation is obtained

q̈ + μq̇ + ωl2 q + αq 2 + βq 3 = f cos t (44)


430 M. Faghih Shojaei et al.

6
2 2 2
Deflection

P= 1 P= 4 P= 9
5

0
0 2 4 6 8 10 12 14 16 18 20
2
Axial Load, P/
100

80

60

40

20

0
0 2 4 6 8 10 12 14 16 18 20
2
Axial Load, P/
Fig. 4 Variations in the static deflection and natural frequencies of free vibration for a SS beam versus the applied axial load
corresponding to the lowest three buckled configurations. (Color figure online)

where

# $   #
(4) 1 # (1) $ # (1) $ $
ωl2= S̃x  Dx  + P − Sx Dx  ◦ Dx 
T
◦ D(2)
x 
2
# $ # # $ # $$$ & # $ 
− D(2)x  ◦ Sx D(1)
x  ◦ Dx 
(1)
S̃x T  (45)
# $ 1 # $ # # $ # $$
α = − S̃x T D(2)
x  ◦ S x D (1)
x  ◦ D (1)
x 
2
# $ # # $ # $$$ & # $ 
+ D(2)x  ◦ S x D (1)
x  ◦ D (1)
x  S̃ x  T
 (46)

1 ## $ # $ # # $ # $$$ & # $ 
β=− S̃x T D(2)
x  ◦ S x D(1)
x  ◦ D(1)
x  S̃x T  (47)
2
# $  & # $ 
f = S̃x T F S̃x T  (48)
Nonlinear forced vibration analysis of postbuckled beams 431

7
2
6 P=4
2 2
P = 8.18 P = 16
5
Deflection

0
0 2 4 6 8 10 12 14 16 18 20
2
Axial Load, P/

120

100

80

60

40

20

0
0 2 4 6 8 10 12 14 16 18 20
2
Axial Load, P/
Fig. 5 Variations in the static deflection and natural frequencies of free vibration for a CC beam versus the applied axial load
corresponding to the lowest three buckled configurations. (Color figure online)

By defining τ = Tt (where T is the period of response) and considering the relation of  = 2π


T for
frequency, Eq. (44) can be represented as follows
 2  
 d2 q μ dq
+ + ωl2 q + αq 2 + βq 3 = f cos 2πτ (49)
2π dτ 2 2π dτ

A column vector X is introduced as


  T
Q = q (τ1 ) , q (τ2 ) , . . . , q τ Nt (50)

where Nt is the number of discrete points in the time domain, and q (τi ) is the nodal value of q (τ ) at grid
point τi which is generated using Eq. (8). After substituting Eq. (9) into Eq. (49), the discretized form of Eq.
(49) can be written as
 2  
 (2) µ (1)
Dt Q + Dt Q + ωl2 Q + α (Q◦Q) + β (Q ◦ Q◦Q) = f A,
2π 2π
A = {cos (2πτi )}T , i = 1, 2, . . . , Nt (51)
432 M. Faghih Shojaei et al.

6
2 2 2
P = 2.05 P = 6.05 P = 12.05
Deflection

0
0 2 4 6 8 10 12 14 16 18 20
2
Axial Load, P/

100

80

60

40

20

0
0 2 4 6 8 10 12 14 16 18 20
2
Axial Load, P/
Fig. 6 Variations in the static deflection and natural frequencies of free vibration for a SC beam versus the applied axial load
corresponding to the lowest three buckled configurations. (Color figure online)

Also, Eq. (51) can be shown as

F : R Nt +1 → R Nt , F (Q, ) = 0 (52)

In the absence of damping (μ) and external force ( f ) terms (i.e., nonlinear vibration around a buckled
configuration), the set of nonlinear equations of (52) can be considered as an eigenvalue problem in which 
is the eigenvalue and Q is the corresponding eigenvector. Such problem cannot be solved using the pseudo-
arc-length continuation method, because  which is an eigenvalue cannot be considered as a parameter. Thus,
a new parameter is introduced into the system by adding a normalizing equation as follows
# $ ' F (Q, ) = 0  T
F̄ : R Nt +2 → R Nt +1 , F̄ Q̃, c = , Q̃ = QT ,  (53)
Q Q−c =0
T

Note that in Eq. (53) which is now an augmented system, both  and Q are related to the value of parameter
c. Now, it is possible to solve Eq. (53) using the pseudo-arc-length continuation method.
For the case of forced vibration around the buckled configurations, the frequency of exited system is equal
to the frequency of applied force and any change in the value of  directly leads to change in the amplitude
of time response. Thus,  is a parameter for the forced vibration problem, and Eq. (52) can be directly solved
using the pseudo-arc-length method.
Nonlinear forced vibration analysis of postbuckled beams 433

Mode 1 Mode 2
0.6
2
0.5
Vibration amplitude

Vibration amplitude
1.5 0.4

0.3
1
0.2
0.5
0.1

0 0
4 6 8 10 12 30 32 34 36 38 40 42

Mode 3

0.25
Vibration amplitude

0.2 2
P=0
2
0.15 P = 0.2
2
P = 0.41
2
0.1 P = 0.71
2
P = 0.81
0.05

0
82 84 86 88 90 92

Fig. 7 Nonlinear frequency response of a SS beam under different prebuckled positions for the lowest three vibration mode
shapes. (Color figure online)

9 Results and discussion

The convergence characteristics and accuracy of the present numerical solution are carefully assessed first.
It has been previously reported that the harmonic differential quadrature (HDQ) method is more reliable for
buckling and vibration problems [26]. Also, it has been shown that the effects of grid numbers are significant
for the convergence of results in linear and nonlinear regimes [27]. To investigate these issues, Table 1 and
Fig. 2 are given. Table 1 shows the free vibration results of an axially loaded beam ( p = 6π 2 ) with different
boundary conditions in the postbuckling domain for different numbers of grid points. It is observed that for
all types of boundary conditions, with increasing the number of grid points, the results tend to converge. As it
is observed, for the problem under study, the difference between the results of GDQ and HDQ is negligible.
Moreover, in Fig. 2, the convergence of results is controlled for the case of nonlinear frequency response in
the postbuckling domain. This figure also indicates quite clearly the converging trend of the present numerical
scheme. Also, from Table 1 and Fig. 2, one can find that 13 grid points are enough to obtain converged results
for both linear and nonlinear analyses.
For the validation purpose, the present results are compared with those of [14] in Fig. 3 for the nonlinear
frequency response of a CC beam under different prebuckled positions corresponding to the first vibration
mode shape. It has been seen that there is an excellent agreement between two sets of results.
The static bifurcation diagrams of the first three buckled configurations of beams with SS, CC, and SC
boundary conditions are shown in Figs. 4, 5, and 6. The variations in natural frequencies of free vibration
against the applied axial load around the first three buckled configurations are also shown in these figures
for both prebuckling and postbuckling domains. These graphs can be found in [24] only for the postbuckling
434 M. Faghih Shojaei et al.

Mode 1 Mode 2
0.5
2
0.4
Vibration amplitude

Vibration amplitude
1.5
0.3

1
0.2

0.5 0.1

0 0
5 10 15 20 25 45 50 55 60 65

Mode 3
0.25

0.2
Vibration amplitude

2
P=0
0.15 2
P = 0.91
2
P = 1.82
0.1 P = 2.74 2

2
P = 3.65
0.05

0
100 105 110 115 120 125

Fig. 8 Nonlinear frequency response of a CC beam under different prebuckled positions for the lowest three vibration mode
shapes. (Color figure online)

domain. One can find that the results given in Figs. 4, 5, and 6 are in very good agreement with those presented
in [24] obtained by an exact solution. This further indicates the reliability and accuracy of the present numeri-
cal solution methodology in predicting the postbuckling configurations and vibrations of buckled beams. The
variations in natural frequencies around the buckled configurations for all types of boundary conditions show
that with increasing the axial load, the natural frequency of beam decreases in the prebuckling domain, while
it increases in the postbuckling domain. In the transition area from the prebuckling domain to the postbuckling
domain (around the critical buckling load), the natural frequency significantly decreases and frequency curves
tend to zero. This is largely because of the severe instability of the structure under the applied axial load.
In order to study the effects of axial load on the forced and nonlinear vibrations of buckled beams, some
specific points shown by solid circles with different colors are selected in Figs. 4, 5, and 6 for the vibration
around the first buckled configuration. As it is observed, 5 and 10 points are chosen in the prebuckling and
postbuckling domains, respectively. The vibration amplitude of beams with different end conditions is plotted
against  in Figs. 7, 8, 9, 10, 11, and 12 for the lowest three vibrations mode shapes around the first buckled
configuration in both prebuckling and postbuckling domains. The nonlinear vibration response is shown by
dashed lines in these figures, whereas the forced vibration response is shown by solid lines. These figures are
given for different magnitudes of axial load corresponding to the positions specified by color circles of Figs. 4,
5, and 6 (the colors of the curves of Figs. 7, 8, 9, 10, 11, 12 are the same as those used in Figs. 4, 5, 6 for color
circles). Also, for Figs. 7, 8, 9, 10, 11, and 12, it is assumed that μ = 0.1 and f = 2.
From Figs. 7, 8, 9, 10, 11, and 12, one can see that a nonlinear frequency curve passes through the intersec-
tion point of the left and right frequency-response curves. This observation confirms the validity of the present
numerical method. In addition, these figures show that by decreasing the vibration amplitude, the nonlinear
Nonlinear forced vibration analysis of postbuckled beams 435

Mode 1 Mode 2

2 0.5
Vibration amplitude

Vibration amplitude
0.4
1.5

0.3
1
0.2

0.5
0.1

0 0
5 10 15 40 45 50 55

Mode 3

0.2
Vibration amplitude

2
0.15 P=0
2
P = 0.51
2
0.1 P = 0.91
2
P = 1.42
2
P = 1.82
0.05

0
95 100 105 110

Fig. 9 Nonlinear frequency response of a SC beam under different prebuckled positions for the lowest three vibration mode
shapes. (Color figure online)

frequencies tend to their linear counterparts. For example, Fig. 7 indicates that the minimum of nonlinear
frequency curve in the first vibration mode and in the absence of axial load (the black dashed curve) is at
 = 9.87. A quick inspection of Fig. 4 also shows that the corresponding solid circle (the black circle in the
first vibration mode and prior to buckling) is located at  = 9.87. Figures 7, 8, and 9 present the effect of axial
load on the nonlinear and forced vibration behaviors of SS, CC, and SC beams prior to buckling. As it was
discussed earlier, the linear frequencies decrease prior to buckling with increasing the magnitude of axial load.
Correspondingly, from Figs. 7, 8, and 9, it is observed that for all kinds of boundary conditions, the nonlinear
frequencies decrease with the increase in axial load in the prebuckling domain. Moreover, these figures show
that as the magnitude of axial load becomes larger and approaches that of critical buckling load, the vibration
amplitude increases and the nonlinear effects become more pronounced. Also, it is found that a hardening-type
response occurs for the forced vibration prior to buckling specially for the first vibration mode shape.
Figures 10, 11, and 12 illustrate the effects of axial load on the forced and nonlinear vibrations of beams
under SS, CC, and SC boundary conditions in the postbuckling domain. It is observed that for symmetric (SS
and CC) boundary conditions and for the first vibration mode, the beam has a softening-type behavior in the
postbuckling domain. One can find that by using only one vibration mode of postbuckled beam, the correct
softening-type behavior in the postbuckling domain reported by Emam [13] is detected. As compared to the
results of Emam [13], which were generated by four free vibration modes, the present solution method seems
more computationally effective. Furthermore, it is seen that the vibration amplitude in the postbuckling domain
decreases with the increase in axial load especially for the first vibration mode. Accordingly, the maximum of
vibration amplitude happens when the magnitude of applied load is around that of critical buckling load. For
example, Fig. 10 related to the SS beam shows that for P = 1.52π 2 which corresponds to the first selected
436 M. Faghih Shojaei et al.

Mode 1

1.2
Vibration amplitude

1 Mode
2
0.8

0.6

0.4

0.2

0
0 10 20 30 40 50 60

Mode 3 2
P = 1.52
2
P = 3.58
0.25
2
P = 5.63
Vibration amplitude

0.2 2
P = 7.69
2
P = 9.74
0.15
2
P = 11.79
0.1 P = 13.85 2

2
0.05 P = 15.9
2
P = 17.96
0 2
75 80 85 90 P = 20.01

Fig. 10 Nonlinear frequency response of a SS beam under different postbuckled positions for the lowest three vibration mode
shapes. (Color figure online)

point in the postbuckling domain, the vibration amplitude significantly increases in the first vibration mode
so that the left and right frequency-response curves do not meet each other in the graph. Therefore, it may be
concluded that for beams in the postbuckling state which are subjected to a periodic force whose frequency is
around the natural frequency, the vibration amplitude significantly increases and the structure becomes unsta-
ble. It should be noted that by increasing the axial load, the dynamic part of the beam deflection (v(x, t)) or
vibration amplitude decreases, while its static part (ψ (x)) increases. In other words, as the axial load increases,
the beam is deflected more, but its vibration amplitude decreases.
It is interesting to note that the nonlinear response of the beam is significantly influenced by its free vibra-
tion in the linear regime. For example, according to Fig. 4, one can find that around the first postbuckled
configuration, the natural frequencies of SS beam in the second and third vibration modes are independent
of applied axial load. From Fig. 5, a similar observation can be made for the case of the second vibration
mode of CC beam. Correspondingly, Figs. 10 and 11 exhibit that the nonlinear response of the postbuck-
led SS and CC beams is independent of axial load in the above-mentioned vibration modes. In addition,
Figs. 4 and 5 indicate that an internal resonance might be activated among the first and second vibration
modes of the SS and CC beams around the first buckled configuration. Likewise, from Figs. 10 and 11, one
can see that the frequency-response curve of the second vibration mode is located between those of the first
vibration mode and there is a specific frequency at which the curves of these two vibration modes inter-
sect each other. The important finding is that when the beam is periodically excited with such frequency,
the occurrence of both softening- or hardening-type responses is possible due to the existence of internal
resonance.
Nonlinear forced vibration analysis of postbuckled beams 437

Mode 1

1.2

1
Vibration amplitude

0.8 Mode 2

0.6

0.4

0.2

0
10 15 20 25 30 35 40 45 50 55

Mode 3
0.25
2
P = 6.03
2
0.2 P = 7.56
Vibration amplitude

2
P = 9.1
2
P = 10.7
0.15
2
P = 12.23
2
P = 13.76
0.1
2
P = 15.36
2
P = 16.9
0.05
2
P = 18.43
2
P = 20.03
0
105 110 115 120 125 130

Fig. 11 Nonlinear frequency response of a CC beam under different postbuckled positions for the lowest three vibration mode
shapes. (Color figure online)

Also, as expected from Fig. 6, Fig. 12 reveals that the forced vibration behavior of the SC beam in the
postbuckling domain depends on the axial load in all of three vibration modes. Another interesting finding
is that in the first vibration mode by increasing the magnitude of axial load, a switching from softening- to
hardening- type behavior occurs, whereas in the second vibration mode, there are hardening- or softening-type
behaviors, depending on the magnitude of axial load.

10 Conclusion

The free, forced, and nonlinear vibrations of buckled Euler–Bernoulli beams under different boundary con-
ditions were studied in this paper using a numerical approach. First, a set of differential and integral matrix
operators were introduced to discretize the space and time domains. The pseudo-arc-length method was used
to obtain the static bifurcation diagrams. The governing equation of free vibration taking place in the vicinity
of a buckled equilibrium position was solved as an eigenvalue problem after suitable matrix manipulations. To
address the problem of forced and nonlinear vibrations of buckled beams, a Galerkin-based numerical scheme
was used to reduce the governing equation into a time-varying ordinary differential equation of Duffing type.
With the aim of increasing the solution accuracy and decreasing the computational cost, a vibration mode
shape around the postbuckling configuration (instead of a summation of free vibration mode shapes) was
employed as a weight function for minimizing the residual resulting from the pseudo-Galerkin method. After
discretizing the Duffing equation by time differential matrix operators, the pseudo-arc-length method was
438 M. Faghih Shojaei et al.

Vibration amplitude Mode 1

0.5

0
5 10 15 20 25 30 35 40

Mode 2
Vibration amplitude

0.4

0.2

0
40 45 50 55 60 65 70
2
P = 3.06
2
P = 4.96
Mode 3
Vibration amplitude

2
P = 6.81
2
P = 8.71
0.2 2
P = 10.62
2
0.1 P = 12.46
2
P = 14.37
0 2
P = 16.21
93 94 95 96 97 98 99 100 101 102
2
P = 18.12
2
P = 20.02

Fig. 12 Nonlinear frequency response of a SC beam under different postbuckled positions for the lowest three vibration mode
shapes. (Color figure online)

used for any given magnitude of axial load in order to study the nonlinear frequency response of periodically
excited buckled beams. The effects of applied axial load on the free, forced, and nonlinear vibrations of beams
under SS, CC, and SC boundary conditions in both prebuckling and postbuckling domains for the lowest three
vibration modes around the first buckled configuration were studied. The most important conclusions drawn
from the present study are the following:

• The free vibration analysis around the buckled configurations showed that for all types of boundary condi-
tions as the magnitude of axial load increases, the natural frequency of beam, respectively, decreases and
increases in the prebuckling and postbuckling domains.
• The nonlinear response of the beam in both prebuckling and postbuckling domains is significantly affected
by its linear free vibration response.
• Prior to buckling and for all kinds of boundary conditions, the hardening-type behavior and nonlinear
effects are more pronounced for lower magnitudes of natural frequency.
• After buckling and for symmetric boundary conditions, the nonlinear response is independent of axial load
in some vibration modes.
Nonlinear forced vibration analysis of postbuckled beams 439

• After buckling and for all kinds of boundary conditions and in the first vibration mode, the vibration
amplitude decreases and the nonlinear effects diminish by increasing the magnitude of axial load.
• After buckling and in the first vibration mode, a softening-type behavior is observed for symmetric boundary
conditions, irrespective of the magnitude of axial load, whereas for SC boundary conditions by increasing
the magnitude of axial load, a switching from softening- to hardening-type behavior happens.
• After buckling, the nonlinear behavior of the periodically excited beam might be a combination of the
softening- and hardening-type responses due to the existence of internal resonance.

Appendix: Matrix products


 
Definition 1 Let A = Ai j N ×M and B = Bi j N ×M , then the Hadamard product of these matrices take the form as A ◦ B =

Ai j Bi j N ×M .
 
Definition 2 Let A = Ai j N ×M and vector U = {Ui } N ×1 , then A U = Ai j Ui N ×M indicates the postmultiplying product of
matrix A and vector U.
 
Definition 3 Let A = Ai j N ×M and vector V = {vi }1×M , then VT A = Ai j V j N ×M indicates the premultiplying product
of matrix A and vector V.

References

1. Eisley, J.G.: Large amplitude vibration of buckled beams and rectangular plates. AIAA J. 2, 2207–2209 (1964)
2. Tseng, W.-Y.: Nonlinear vibrations of straight and buckled beams under harmonic excitation. Ph.D. thesis, Massachusetts
Institute of Technology, Department of Aeronautics and Astronautics (1970)
3. Eisley, J.G., Bennett, J.A.: Stability of large amplitude forced motion of a simply supported beam. Int. J. Non-Linear
Mech. 5, 645–657 (1970)
4. Tseng, W.-Y., Dugundji, J.: Nonlinear vibrations of a buckled beam under harmonic excitation. ASME J. Appl.
Mech. 38, 467 (1971)
5. Min, G.B., Eisley, J.G.: Nonlinear vibration of buckled beams. ASME J. Eng. Ind. 94, 637 (1972)
6. Holmes, P.: Global bifurcations and chaos in the forced oscillations of buckled structures. In: IEEE Conference, Decision
and Control including the 17th Symposium on Adaptive Processes (1978)
7. Pezeshki, C., Dowell, E.H.: An examination of initial condition maps for the sinusoidally excited buckled beam modeled by
the Duffing’s equation. J. Sound Vib. 117, 219–232 (1987)
8. Pezeshki, C., Elgar, S., Krishna, R.C.: An examination of multi-frequency excitation of the buckled beam. J. Sound Vib. 148,
1–9 (1991)
9. Higuchi, K., Dowell, E.H.: Effect of constant transverse force on chaotic oscillations of sinusoidally excited buckled beam. Int.
J. Non-Linear Mech. 26, 419–426 (1991)
10. Afaneh, A.A., Ibrahim, R.A.: Nonlinear response of an initially buckled beam with 1:1 internal resonance to sinusoidal
excitation. Nonlinear Dyn. 4, 547–571 (1993)
11. Abou-Rayan, A.M., Nayfeh, A.H., Mook, D.T., Nayfeh, M.A.: Nonlinear response of a parametrically excited buckled
beam. Nonlinear Dyn. 4, 499–525 (1993)
12. Ji, J.-C., Hansen, C.H.: Non-linear response of a post-buckled beam subjected to a harmonic axial excitation. J. Sound
Vib. 237, 303–318 (2000)
13. Emam, S.A. (2002) A theoretical and experimental study of nonlinear dynamics of buckled beams. Ph.D. thesis, Virginia
Polytechnic Institute and State University, Blacksburg, VA
14. Virgin, L.N., Plaut, R.H.: Axial load effects on the frequency response of a clamped beam. In: IMAC-XXI: Conference and
Exposition on Structural Dynamics (2003)
15. Emam, S.A., Nayfeh, A.H.: On the nonlinear dynamics of a buckled beam subjected to a primary-resonance excitation. Non-
linear Dyn. 35, 1–17 (2004)
16. El-Bassiouny, A.F.: Nonlinear vibration of a post-buckled beam subjected to external and parametric excitations. Phys.
Scr. 74, 39 (2006)
17. Nagai, K., Maruyama, S., Sakaimoto, K., Yamaguchi, T.: Experiments on chaotic vibrations of a post-buckled beam with an
axial elastic constraint. J. Sound Vib. 304, 541–555 (2007)
18. Yabuno, H., Okada, J.: Stabilization of buckled beam by high-frequency axial excitation. In: IEEE Conference, ICCAS-SICE,
Fukuoka, pp. 283–286 (2009)
19. Zhang, G.-C., Ding, H., Chen, L.-Q., Yang, S.-P.: Galerkin method for steady-state response of nonlinear forced vibration
of axially moving beams at supercritical speeds. J. Sound Vib. 331, 1612–1623 (2012)
20. Kazemirad, S., Ghayesh, M.H., Amabili, M.: Thermo-mechanical nonlinear dynamics of a buckled axially moving beam.
Arch. Appl. Mech. (2012). doi:10.1007/s00419-012-0630-8
21. Ghayesh, M.H., Amabili, M.: Nonlinear dynamics of axially moving viscoelastic beams over the buckled state. Comput.
Struct. 112–113, 406–421 (2012)
22. Shu, C.: Differential Quadrature and Its Application in Engineering. Springer, London (2000)
23. Trefethen, L.N.: Spectral Methods in MATLAB. Oxford University, Oxford (2000)
440 M. Faghih Shojaei et al.

24. Nayfeh, A.H., Emam, S.A.: Exact solution and stability of postbuckling configurations of beams. Nonlinear Dyn. 54, 395–
408 (2008)
25. Keller, H.B.: Numerical solution of bifurcation and nonlinear eigenvalue problems, applications of bifurcation theory (Proc.
Advanced Sem., Univ. Wisconsin, Madison, Wis., 1976). Academic Press, New York, pp. 359–384 (1977)
26. Civalek, Ö.: Application of differential quadrature (DQ) and harmonic differential quadrature (HDQ) for buckling analysis
of thin isotropic plates and elastic columns. Eng. Struct. 26, 171–186 (2004)
27. Yuan, Z., Wang, X.: Buckling and post-buckling analysis of extensible beam-columns by using the differential quadrature
method. Comput. Math. Appl. 62, 4499–4513 (2011)

You might also like