You are on page 1of 12

Journal of Sound and Vibration ] (]]]]) ]]]–]]]

Contents lists available at SciVerse ScienceDirect

Journal of Sound and Vibration


journal homepage: www.elsevier.com/locate/jsvi

Dynamic model for free vibration and response analysis


of rotating beams
Hyungrae Kim a, Hong Hee Yoo b, Jintai Chung a,n
a
Department of Mechanical Engineering, Hanyang University, 1271 Sa-3-dong, Sangnok-gu, Ansan, Kyeonggi-do 425-791,
Republic of Korea
b
School of Mechanical Engineering, Hanyang University, 17 Haengdang-dong, Seongdong-gu, Seoul 133-791, Republic of Korea

a r t i c l e i n f o abstract

Article history: A new modeling method for the vibration analysis of a rotating cantilever beam is proposed
Received 27 September 2012 and compared to the other modeling methods. Applying the Hamilton principle, the
Received in revised form equations for the axial, chordwise and flapwise motions are derived. During the derivation
11 March 2013
of these equations, the nonlinear von Karman strain and the corresponding linear stress are
Accepted 6 June 2013
used to consider the stiffening effect due to rotation. The derived equations of motion
Handling Editor: H. Ouyang
obtained from the proposed method are analytically and numerically compared to equations
from the previous modeling methods. Computations of the natural frequencies and time
responses show that the described equations of motion are more reliable than the equations
of the other modeling methods.
& 2013 Elsevier Ltd. All rights reserved.

1. Introduction

The dynamic problems of rotating cantilever beams have drawn the attention of many researchers because rotating
beams have wide practical applications such as aircraft rotary wings and turbine blades. Since rotating beam structures have
small elastic vibrations as well as overall motion of rotation, more reliable dynamic models of these structures are required
for precise operation or accurate speed control. The elastic vibrations are caused by axial, chordwise (or lagwise), flapwise
and torsional deformations. The axial and chordwise deformations, which are in the radial and tangential directions
respectively, are the in-plane deformations, while the flapwise motion, which is in the direction of the rotation axis, is the
out-of-plane deformation.
Modeling issues regarding rotating beams are interesting research topics. Many dynamic models have been presented which
include the stiffening and Coriolis effects due to beam rotation. These effects induce coupling between the axial, chordwise and
flapwise motions. The modeling methods may be classified into three categories. The modeling methods of the first category [1–4]
are based on the hybrid set of deformation variables, which consist of a non-Cartesian variable (stretch deformation) and two
Cartesian variables (chordwise and flapwise deformations) to describe the elastic deformation. In the modeling methods of the
second category [5–10], the stiffening effect of a beam was accounted for by adding the elastic potential energy resulting from
centrifugal force to the total potential energy. The modeling methods of the third category [11–13] use the nonlinear strain/stress to
consider the stiffening effect of rotating beams.
Yoo and Shin [1] proposed a modeling method of a rotating beam using a hybrid set of deformation variables. After
expressing the potential energy with a hybrid set of deformation variables, the linear discretized equations of motion were

n
Corresponding author. Tel.: +82 31 400 5287; fax: +82 31 406 6964.
E-mail address: jchung@hanyang.ac.kr (J. Chung).

0022-460X/$ - see front matter & 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jsv.2013.06.004

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
2 H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]]

derived by employing Kane's method. This modeling method was applied for the finite element analyses of rotating beams
[2,3]. Different from the studies with discretized ordinary equations of motion [1–3], Chung and Yoo [4] derived linear
partial differential equations for the stretch, chordwise and flapwise motions of a rotating beam.
Based on the elastic potential energy related to the centrifugal force, Hoa [5] derived the equation of flapwise motion for
a rotating beam and Zhu and Mote [6] derived the equations of flapwise and torsional motions. In a similar manner, the
equations of the axial, chordwise and torsional motions were introduced by Yang et al. [7], and the equations of the
chordwise and torsional motions were presented by Al-Qaisia and Al-Bedoor [8]. If the torsional motion is neglected from
the equations of Refs. [7,8], the equations of the axial and chordwise motions are the same as the corresponding equations
of Chung and Yoo [4]. Banerjee [9] developed the dynamic stiffness matrix of a centrifugally stiffened Timoshenko beam and
used the matrix to carry out a free vibration analysis. Banerjee [10] also used the dynamic stiffness method to analyze the
free vibration of centrifugally stiffened uniform and tapered beams. Even though the equations of Ref. [10] were derived
by using Newton's second law, these equations are equivalent to the equations for the modeling methods of the second
category.
In order to consider the stiffening effect of a rotating beam, Pesheck et al. [11] used nonlinear strain when driving the
equations of axial and flapwise motions. After obtaining the steady-state solution for axial deformation, this solution was
substituted into the equation of flapwise motion. The final equations of this study are linear equations for axial and flapwise
motions. Huang et al. [12] investigated the natural frequency of the axial, chordwise and flapwise vibrations for a rotating
beam. They proposed a method based on the power series solution to solve the natural frequency of a slender rotating beam
at a high angular velocity. Arvin and Bakhtiari-Nejad [13] also derived the nonlinear equations of the axial and flapwise
motions. However, these equations of motion are only for the free vibration analysis of a rotating beam.
In this study, a new dynamic modeling method for a rotating cantilever beam is presented and the equations of motion
derived from the presented method are compared to the equations of other models. Based on the nonlinear von Karman
strain and the associated linear stress, the equations of motion are derived by using the Hamilton principle. In order to show
the reliability of the proposed model, the equations of the present study are analytically and numerically compared to the
equations of other modeling methods.

2. Derivation of the equations of motion

A flexible uniform cantilever beam is shown in Fig. 1, which is fixed to a rotating rigid hub at point O. The beam has
length L, cross-sectional area A, mass density ρ, Young's modulus E, and area moments of inertia about the y and z axes Iy and
Iz, respectively. The rigid hub with radius a rotates about the vertical axis passing through point O′ and the rotating speed of
the cantilever beam and hub is Ω. It is assumed that the beam has a doubly symmetric cross section such as a circle or a
rectangle. Therefore, the coupling effect between the bending and torsional motions is neglected and the torsional behavior
of the beam is not discussed in this paper. The x axis is along the undeflected beam, the y axis is in the rotation plane of the
beam, and the z axis is parallel to the rotation axis of the beam. The xyz coordinate system rotates with the hub. When point
Pn on the beam moves to point P, the deformation of the beam is described by the axial deformation u, the chordwise
deformation v, and the flapwise deformation w. In this case, the velocity of point P is given by
   
∂u ∂v ∂w
v¼ −Ωv i þ þ Ωða þ x þ uÞ j þ k (1)
∂t ∂t ∂t

where i, j and k are the unit vectors along the x, y and z directions, respectively. Note that the modeling methods of the first
category [1–4] use the stretch coordinate s instead of the axial deformation u (Fig. 1).
Modeling the beam as the Euler–Bernoulli beam, the shear deformation is negligible so the displacements of point (x,y)
in the x, y and z directions, ux(x,y,t), uy(x,y,t) and uz(x,y,t), can be represented in terms of the displacements at the beam
centerline, u, v and w:

∂vðx; tÞ ∂wðx; tÞ
ux ðx; y; tÞ ¼ uðx; tÞ−y −z ; uy ðx; y; tÞ ¼ vðx; tÞ; uz ðx; y; tÞ ¼ wðx; tÞ (2)
∂x ∂x

P
y
x s
u ui vj wk
x
O O P
a x
L
Fig. 1. Rotating beam with rotating speed Ω.

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]] 3

to consider the nonlinearity of the beam, the von Karman strain theory [14] is adopted in the formulation. Using the von
Karman strain theory, the x-directional normal strain at point (x,y) given by
   
∂ux 1 ∂uy 2 1 ∂uz 2
εx ¼ þ þ (3)
∂x 2 ∂x 2 ∂x

substituting Eq. (2) into Eq. (3), the nonlinear strain εx and the corresponding linear stress sx can be expressed as
   
∂u 1 ∂v 2 1 ∂w 2 ∂2 v ∂2 w
εx ¼ þ þ −y 2 −z 2 (4)
∂x 2 ∂x 2 ∂x ∂x ∂x

 
∂u ∂2 v ∂2 w
sx ¼ E −y 2 −z 2 (5)
∂x ∂x ∂x

a beam shows nonlinear behavior when the chordwise and flapwise deformations are much greater than the axial
deformation. The equations of motion derived by using the nonlinear strain and linear stress are much simpler than the
equations derived by using the nonlinear strain and nonlinear stress. In addition, the responses computed from the both sets
of equations are nearly the same. For this reason, the linear stress is often used with the nonlinear strain. The use of
nonlinear strain and linear stress when deriving the equations of motion can be found in the literature on the vibrations of a
string [15], a beam [16] and a rotating disk [17].
The variations of the kinetic energy K, potential energy P and virtual work done by non-conservative forces W are given by
Z L Z LZ Z L
δK ¼ ρA v⋅δvdx; δP ¼ sx δεx dAdx; δW ¼ ðpv δv þ pw δwÞdx (6)
0 0 A 0

where δ is the variational notation and pv and pw represent the applied forces per unit length in the y and z directions, respectively.
Introducing Eq. (6) into the following Hamilton's principle:
Z t2
ðδK−δP þ δWÞdt ¼ 0 (7)
t1

the equations of the axial, chordwise and flapwise motions can be obtained as follows:
 2 
∂ u ∂v _ ∂2 u
ρA −2Ω −Ω2 u−Ωv −EA 2 ¼ ρAΩ2 ða þ xÞ (8)
∂t 2 ∂t ∂x

 2   
∂ v ∂u _ 2 ∂ ∂u ∂v ∂4 v _ þ xÞ
ρA þ 2Ω þ Ωu−Ω v −EA þ EI z 4 ¼ pv −ρAΩða (9)
∂t 2 ∂t ∂x ∂x ∂x ∂x

 
∂2 w ∂ ∂u ∂w ∂4 w
ρA −EA þ EI y 4 ¼ pw (10)
∂t 2 ∂x ∂x ∂x ∂x

where the superposed dot represents differentiation with respect to time. The associated boundary conditions are given by

∂v ∂w ∂u ∂2 v ∂2 w ∂3 v ∂3 w
u¼v¼w¼ ¼ ¼ 0 at x ¼ 0; ¼ 2 ¼ ¼ 3 ¼ ¼ 0 at x ¼ L (11)
∂x ∂x ∂x ∂x ∂x 2 ∂x ∂x3
As shown in Eqs. (8)–(10), the equations of the chordwise and flapwise motions are coupled with the equation of axial
motion. The Coliolis effect of a rotating beam is related to the term 2Ω ∂v=∂t of Eq. (8) and the term 2Ω ∂u=∂t of Eq. (9). On
the other hand, the term EA ∂u=∂x in Eqs. (9) and (10) represents the axial force of a rotating beam that results in a stiffening
effect in the chordwise and flapwise motions. This axial force is not constant when a rotating beam is in a transient state
_
(or when Ω≠0), because the axial deformation is a function of time. Therefore, Eqs. (8)–(10) should be solved simultaneously
to analyze the dynamic responses of a rotating beam when the beam has a variable rotating speed.
When the rotating beam is in a steady state ðΩ _ ¼ 0Þ and the applied forces are zero (pv ¼pw ¼0), the steady-state
equilibrium deformations and the equations linearized in the neighborhood of the equilibrium position are obtained by
applying the perturbation method to Eqs. (8)–(10). According to the perturbation method, the deformation of the beam
can be expressed as the sum of the steady-state equilibrium deformations and the perturbed deformations around the
equilibrium deformations. Therefore, the axial, chordwise and flapwise deformations may be represented by
u ¼ u0 þ Δu; v ¼ v0 þ Δv; w ¼ w0 þ Δw (12)

where u0, v0 and w0 are the deformations corresponding to the equilibrium deformations, and Δu, Δv and Δw are the
perturbed deformations of u, v and w around the equilibrium deformations, respectively. Because u0, v0 and w0 have
constant values at a steady state, their derivatives with respect to time are zero.
Hence, neglecting the rotating acceleration and the applied forces and substituting Eq. (12) into Eqs. (8)–(10), the steady-
state equilibrium equations and the linearized equations of motion can be derived. The derived equilibrium equations are

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
4 H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]]

given by
2
d u0
EA þ ρAΩ2 u0 ¼ −ρAΩ2 ða þ xÞ; v0 ¼ w0 ¼ 0 (13)
dx2
solving the first equation in Eq. (13) by applying the boundary conditions of Eq. (11), the exact solution of this equation is
given by
aβ cos βðL−xÞ þ sin βx
u0 ¼ −ða þ xÞ (14)
β cos βL
where
pffiffiffiffiffiffiffiffi
β ¼ Ω ρ=E (15)

the linearized equations for the axial, chordwise and flapwise motions, which describe the free vibrations around the
equilibrium deformation, are given by
 2 
∂ u ∂v 2 ∂2 u
ρA −2Ω −Ω u −EA 2 ¼ 0 (16)
∂t 2 ∂t ∂x

 2   
∂ v ∂u ∂ ∂v ∂4 v
ρA þ 2Ω −Ω2 v − Fx þ EI z 4 ¼ 0 (17)
∂t 2 ∂t ∂x ∂x ∂x

 
∂2 w ∂ ∂w ∂4 w
ρA − Fx þ EI y 4 ¼ 0 (18)
∂t 2 ∂x ∂x ∂x

where Δ is deleted from Δu, Δv and Δw for simplicity of notation and F x is the axial force due to the centrifugal effect,
given by
 
du0 aβ sin βðL−xÞ þ cos βx
F x ¼ EA ¼ EA −1 (19)
dx cos βL

The discretized equations of motion are obtained by using the Galerkin method. Before applying the Galerkin method,
the variational equations are derived from the equations of motion given by Eqs. (8)–(10). To derive the variational
equations, the weighting functions need to be defined. The weighting functions are defined as arbitrary functions to satisfy
the essential (or geometric) boundary conditions. The weighting functions for the axial, chordwise and flapwise functions
(u, v and w) are denoted by u, v and w. The variational equations for the axial, chordwise and flapwise motions may be
obtained by multiplying Eqs. (8), (9) and (10) by u, v and w, respectively, integrating these equations over the length L, and
applying integration by parts. The variational equations of motion are represented by
Z L 2  Z L Z L
∂ u ∂v _ ∂u ∂u
ρA u 2 −2Ωu −Ω2 uu−Ωuv dx þ EA dx ¼ ρAΩ2 ða þ xÞudx (20)
0 ∂t ∂t 0 ∂x ∂x 0

Z L   Z L Z L 2 2
∂2 v ∂u _ 2 ∂v ∂u ∂v ∂ v∂ v
ρA v þ 2Ωv þ Ωvu−Ω vv dx þ EA dx þ EI z dx
0 ∂t 2 ∂t 0 ∂x ∂x ∂x 0 ∂x ∂x
2 2
Z L Z L
¼ pv vdx−ρAΩ_ ða þ xÞvdx (21)
0 0

Z L Z L Z L Z L
∂2 w ∂w ∂u ∂w ∂2 w ∂2 w
ρA w dx þ EA dx þ EI y dx ¼ pw wdx (22)
0 ∂t 2 0 ∂x ∂x ∂x 0 ∂x2 ∂x2 0

In order to find approximate solutions in finite-dimensional function spaces by using the Galerkin method, the axial, chordwise
and flapwise deformations are approximated by linear combinations of the comparison functions. These approximated functions,
often called the trial functions, can be expressed as
N N N
uðx; tÞ ¼ ∑ T uj ðtÞU j ðxÞ; vðx; tÞ ¼ ∑ T vn ðtÞV n ðxÞ; wðx; tÞ ¼ ∑ T w
q ðtÞW q ðxÞ (23)
j¼1 n¼1 q¼1

where N is the total number of comparison functions; T uj , T vn and T w


q are the functions of time to be determined; and U j , V n and W q
are the comparison functions satisfying both the essential and natural boundary conditions. In this study, the comparison functions
for the axial deformation are selected as the mode functions for the longitudinal vibration of a cantilever beam while the
comparison functions for the chordwise and flapwise motions are selected as the mode functions for the transverse vibration of a
cantilever beam. These mode functions are given by
pffiffiffi πx
U j ðx; tÞ ¼ 2 sin ð2j−1Þ (24)
2L

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]] 5

sinhλn L− sin λn L
V n ðxÞ ¼ W n ðxÞ ¼ coshλn x− cos λn x− ðsinhλn x− sin λn xÞ (25)
coshλn L þ cos λn L
where λn is the nth root of
cos λn Lcoshλn L þ 1 ¼ 0 (26)
similarly, the weighting functions for the axial, chordwise and flapwise deformations can be approximated by
N u N v N w
uðx; tÞ ¼ ∑ T i ðtÞU i ðxÞ; vðx; tÞ ¼ ∑ T m ðtÞV m ðxÞ; wðx; tÞ ¼ ∑ T p ðtÞW p ðxÞ (27)
i¼1 m¼1 p¼1
u v w
where T i , T m and T p are the arbitrary functions of time.
u v
The discretized equations of motion are obtained by using the arbitrariness of the time-dependent functions, T i , T m and
w
T p , after introducing the trial and weighting functions into the variational equations. Substituting Eqs. (23) and (27) into
Eqs. (20)–(22), the discretized equations of the axial, chordwise and flapwise motions are obtained as follows:
N u
∑ ðmuij T€ j þ g uv _v u u uv v u
in T n þ kij T j þ kin T n Þ ¼ f i ; i ¼ 1; 2; …; N (28)
j¼1

N v
∑ ½mvmn T€ n þ g vu _ u u v u vu v v
mj T j þ kmj T j þ kmn ðT j ÞT n  ¼ f m ; m ¼ 1; 2; …; N (29)
n¼1

N
∑ ½mw €w w u w w
pq T q þ kpq ðT j ÞT q  ¼ f p ; p ¼ 1; 2; …; N (30)
q¼1

where
Z L Z L Z L Z L
U ′i U ′j dx
u
muij ¼ ρA U i U j dx; g uv
in ¼ −2ρAΩ U i V n dx; kij ¼ −ρAΩ2 U i U j dx þ EA
0 0 0 0
Z L Z L Z L
uv _
kin ¼ −ρAΩ U i V n dx;
u
f i ¼ ρAΩ2 ða þ xÞU i dx; mvmn ¼ ρA V m V n dx
0 0 0
Z L Z L
g vu
mj ¼ 2ρAΩ V m U j dx;
vu _
kmj ¼ ρAΩ V m U j dx
0 0
Z L Z L N
 Z L 
V ″m V ″n dx þ ∑ U ′j V ′m V ′n dx T uj
v
kmn ðT uj Þ ¼ −ρAΩ2 V m V n dx þ EI z EA
0 0 j¼1 0
Z L Z L
v
fm ¼ _ þ xÞV m dx;
½pv −ρAΩða mw
pq ¼ ρA W p W q dx
0 0
Z L N
 Z L  Z L
W ″p W ″q dx þ ∑ U ′j W ′p W ′q dx T uj ;
w w
kpq ðT uj Þ ¼ EI y EA fp ¼ pw W p dx (31)
0 j¼1 0 0

The discretized equations of the axial and chordwise motions can be represented by the following matrix-vector
equation:
" u #8 u 9 " #8 _ u 9 2 u 3( ) ( u )
½0 < fT€ j g = < fT j g = uv
½mij  in 
½0 ½g uv ½kij  ½kin  fT uj g ff i g
þ þ 4 5 ¼ (32)
½0 ½mvmn  : fT€ v g ; mj 
½g vu ½0 : fT_ v g ; vu v
½kmj  ½kmn ðT uj Þ fT vn g
v
ff m g
n n

in a similar manner, the discretized equations of the flapwise motion may be expressed as
w
½mw € u w w w
pq fT q g þ ½kpq ðT j ÞfT q g ¼ ff p g (33)

on the other hand, the discretized equations corresponding to the linearized equations of Eqs. (16)–(18) can be written in
the following matrix-vector form:
" u #8 u 9 " #8 _ u 9 " u #( u ) ( )
½mij  ½0 < fT€ j g = ½0 ½g uv
in 
< fT j g = ½kij  ½0 fT j g f0g
þ ½g vu þ ¼ (34)
½0 ½mvmn  : fT€ v g ; mj  ½0 : fT_ v g ; f0g
v
n
½0 ½kmn  fT vn g
n

½mw €w w w
pq fT q g þ ½kpq fT q g ¼ f0g (35)

where
Z L Z L Z L
V ″m V ″n dx þ F x V ′m V ′n dx
v
kmn ¼ −ρAΩ2 V m V n dx þ EI z
0 0 0

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
6 H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]]

Z L Z L
W ″p W ″q dx þ F x W ′p W ′n dx
w
kpq ¼ EI y (36)
0 0

3. Comparison of the modeling methods

It is valuable to compare the equations of motion of this study with the equations of the modeling methods of the
first category mentioned in Introduction [1–4]. The equations of this category are obtained by using the hybrid set of
deformation variables including a non-Cartesian variable and two Cartesian variables. Most studies using the hybrid set of
variables [1–3] presented the equations of motion in the form of discretized ordinary differential equations, not in the form
of partial differential equations. However, Chung and Yoo [4] derived the partial differential equations of motion. Replacing
the stretch deformation s by the axial deformation u (Fig. 1), the equations of motion reported by Chung and Yoo [4] can be
rewritten as
 2 
∂ u ∂v _ ∂2 u
ρA −2Ω −Ω2 u−Ωv −EA 2 ¼ ρAΩ2 ða þ xÞ (37)
∂t 2 ∂t ∂x

 2   
∂ v ∂u _ 2 ∂ ∂v ∂4 v _ þ xÞ
ρA þ 2Ω þ Ωu−Ω v − F s þ EI z 4 ¼ pv −ρAΩða (38)
∂t 2 ∂t ∂x ∂x ∂x

 
∂2 w ∂ ∂w ∂4 w
ρA − Fs þ EI y 4 ¼ pw (39)
∂t 2 ∂x ∂x ∂x

where

F s ¼ ρAΩ2 ½aðL−xÞ þ 12ðL2 −x2 Þ (40)

note that Eq. (37) becomes identical to Eq. (8). The main difference in the equations of motion between this study and
Ref. [4] is the axial force expression. This study uses a time-dependent axial force EA ∂u=∂x while Chung and Yoo [4] used a
_
time-independent axial force given by Eq. (40). If a rotating beam is not in a steady state ðΩ≠0Þ and the applied forces are
non-zero (pv ≠0 and pw ≠0), the time-independent force of Eq. (40) is not acceptable for dynamic analysis of a rotating beam.
Therefore, the governing equations of the present study (i.e., Eqs. (8)–(10)) are more reasonable compared to the equations
of Ref. [4] (i.e., Eqs. (37)–(39)).
In the modeling methods of the second category [5–10], in order to consider the stiffening effect due to centrifugal force,
the potential energy arising from the centrifugal force was added to the elastic potential energy. In these modeling methods,
the axial force of a rotating beam induced by the centrifugal force is expressed as
Z L
F ¼ ρA Ω2 ða þ xÞdx ¼ ρAΩ2 ½aðL−xÞ þ 12ðL2 −x2 Þ (41)
x

it is interesting that this axial force is identical to the axial force, given by Eq. (40), obtained in the modeling methods of the
first category. Considering the potential energy due to the centrifugal force, the total potential energy can be written as
Z (  2  2 2  2 2 "   2 #)
1 L ∂u ∂ v ∂ w ∂v 2 ∂w
U¼ EA þ EI z þ EI y þ F þ dx (42)
2 0 ∂x ∂x2 ∂x2 ∂x ∂x

the equations of motion derived with the potential energy of Eq. (42) have the same forms as Eqs. (37)–(39). This means that
the modeling methods of the first and second categories eventually are the same even though the approaches of these
methods are different from each other. Therefore, in further comparisons, only the modeling method of the first category is
considered and the methods of the second category are excluded.
The difference in the axial forces between the present study and Ref. [4] (i.e., Eqs. (19) and (40)) can be explained with a
free-body diagram for an infinitesimal element of the rotating beam with rotating speed Ω (Fig. 2). When the beam is in a
free steady state, the infinitesimal element has an axial equilibrium deformation u0 and the distance between the element
and the rotation axis is a þ x þ u0 . Therefore, the centripetal acceleration of the element is given by −ða þ x þ u0 ÞΩ2 . In this
case, the first equation in Eq. (13) is obtained by applying Newton's second law to the element of mass ρAdx. However, if the

u u u
EA EA EA x
O x x x

a x u x
Fig. 2. Free-body diagram for an infinitesimal element of the rotating beam with angular velocity Ω.

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]] 7

deformation u0 is neglected in the distance a þ x þ u0 , the following equation is obtained:


2
d u0
EA ¼ −ρAΩ2 ða þ xÞ (43)
dx2
the axial force derived from the solution of Eq. (43) which satisfies the associated boundary conditions is the same as the
axial force of Ref. [4] given in Eq. (40). Since the equilibrium equation of this study, given in Eq. (13), considers an extra term
ρAΩ2 u0 when compared to the equilibrium equation of Ref. [4], Eq. (43), the axial force of this study (i.e., Eq. (19)) is more
accurate than the axial force of Ref. [4] (i.e., Eq. (40)).
The modeling methods of the third category [11–13] are similar to the modeling method of the present study. All of these
methods use the nonlinear strain to account for the stiffening effect due to the centrifugal force. However, the equations of
the previous studies [11–13] cannot be used to compute dynamic responses in a transient state. Therefore, if the beam has a
rotating acceleration or has applied forces, the equations proposed in this study (i.e., Eqs. (8)–(10)) should be used to obtain
the time responses of the beam. On the other hand, the linear equations presented in [11–13] are identical to the linear
equations of this study, Eqs. (16)–(18). Therefore, this study and the previous studies [11–13] result in the same natural
frequencies of a rotating beam.
The natural frequencies and time responses of rotating beams are compared between the models of this study and Ref.
[4]. The modeling methods of the second category [5–10] do not need to be considered as part of the comparison because
the equations of these methods are the same as those in Ref. [4]. For convenience of comparison, the beam is assumed to
have equal area moments of inertia with respect to x and y axes (i.e., Iy ¼ Iz) and the following dimensionless parameters are
used for numerical computations:
sffiffiffiffiffiffiffiffi
t x a AL2 _
τ ¼ ; ξ ¼ ; δ ¼ ; γ ¼ TΩ; α ¼ ; λ ¼ T 2Ω (44)
T L L Iz

where
sffiffiffiffiffiffi
2ρA
T ¼L (45)
EI z

The natural frequencies of the rotating beam obtained from the present study and Ref. [4] are compared. For this
purpose, the natural frequencies of this study are computed with the discretized equations of Eqs. (34) and (35), while the
natural frequencies of Ref. [4] are computed with the equations obtained from Eqs. (34) and (35) by replacing Fx with Fs.
Recall that Eqs. (34) and (35) are obtained by discretizing Eqs. (16)–(18) with the trial functions given in Eq. (23). As shown
in Eqs. (34) and (35), the axial motion is coupled with the chordwise motion but the flapwise motion is not coupled
with other motions. Since it is difficult to distinguish the natural frequencies between the axial and chordwise motions, the
natural frequencies of the axial and chordwise motions, denoted by ωuv, are computed together from Eq. (34). Meanwhile,
the natural frequencies of the flapwise motion, denoted by ωw, are computed from Eq. (35).
Before comparison, the convergence characteristics of the natural frequencies computed with the model of this study are
investigated for the rotating beam in a free steady state. Table 1 presents the convergence results for the dimensionless
natural frequencies of the axial and chordwise motions, Tωuv, when δ¼0, γ¼10 and α¼70. As shown in this table, the five
lowest natural frequencies for the axial and chordwise motions converge as the total number of comparison functions
increases. Similarly, the convergence characteristics for the dimensionless natural frequencies of flapwise motion, Tωw, are
also demonstrated in Table 2. Tables 1 and 2 show that reasonably converged natural frequencies are obtained when N¼ 10.
Hence, ten comparison functions are used for further computations of this study.
The natural frequencies of rotating beams are compared between the present study and Ref. [4] when δ¼0 and α¼70.
To investigate the differences in the natural frequencies, for the variation of the dimensionless rotating speed, the
dimensionless natural frequencies of the axial and chordwise motions are plotted in Fig. 3 and the natural frequencies of the
flapwise motion are plotted in Fig. 4. In these figures, the solid lines represent the natural frequencies computed in this

Table 1
Convergence characteristics of the dimensionless natural frequencies of the axial and chordwise motions, Tωuv, when δ ¼0, γ ¼ 10 and α¼ 70.

N 1st Mode 2nd Mode 3rd Mode 4th Mode 5th Mode

1 5.6282 – – – –
2 5.2027 32.2271 – – –
3 5.0966 32.2220 74.1576 – –
4 5.0705 32.1455 74.0957 111.3161 –
5 5.0619 32.1357 74.0267 111.3161 134.6382
6 5.0587 32.1292 74.0206 111.3161 134.5875
7 5.0573 32.1269 74.0140 111.3161 134.5842
8 5.0567 32.1257 74.0117 111.3161 134.5788
9 5.0563 32.1251 74.0102 111.3161 134.5771
10 5.0562 32.1247 74.0094 111.3161 134.5756

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
8 H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]]

Table 2
Convergence characteristics of the dimensionless natural frequencies of the flapwise motion, Tωw, when δ ¼0, γ ¼ 10 and α¼ 70.

N 1st Mode 2nd Mode 3rd Mode 4th Mode 5th Mode

1 11.5174 – – – –
2 11.3107 33.8110 – – –
3 11.2608 33.8093 74.8876 – –
4 11.2487 33.7359 74.8301 135.1517 –
5 11.2447 33.7266 74.7617 135.0553 214.7337
6 11.2432 33.7204 74.7559 135.0047 214.6226
7 11.2426 33.7182 74.7493 135.0017 214.5856
8 11.2423 33.7170 74.7471 134.9963 214.5841
9 11.2421 33.7164 74.7455 134.9946 214.5799
10 11.2420 33.7161 74.7448 134.9932 214.5787
T uv

Fig. 3. Dimensionless natural frequencies of the axial and chordwise motions, Tωuv, for variation in the dimensionless rotating speed, γ, when δ ¼ 0 and
α ¼ 70 (solid line: present study; dotted line: Ref. [4]).
T w

Fig. 4. Dimensionless natural frequencies of the flapwise motion, Tωw, for variation in the dimensionless rotating speed, γ, when δ ¼ 0 and α ¼70 (solid line:
present study; dotted line: Ref. [4]).

study while the dotted lines represent the frequencies computed in Ref. [4]. As shown in Figs. 3 and 4, the natural
frequencies computed by the present study and Ref. [4] do not exhibit large differences in the low rotating speed range.
However, the differences become large in the high rotating speed range.
It is interesting to investigate the mode shapes of the rotating beam. When the dimensionless parameters are given by
δ¼0, γ¼10 and α¼70, the lowest six mode shapes of the axial and chordwise motions are presented in Fig. 5, where the
dotted lines represent the undeformed beams and the numbers in parentheses are the dimensionless natural frequencies.
Since the axial motion is coupled with the chordwise motion, the mode shapes presented in Fig. 5 have both the axial and
chordwise deformations. As shown in Fig. 5(a)–(c), (e) and (f), the first, second, third, fifth and sixth mode shapes have one,
two, three, four and five nodal points, respectively. However, the fourth mode shape, shown in Fig. 5(d), has only one nodal
point. Therefore, the lowest five modes except the forth mode have similar shapes to the lowest five modes for the
transverse vibration of a stationary cantilever beam. On the other hand, the flapwise motion is not coupled with the axial
motion, so the mode shapes of the flapwise motion, illustrated in Fig. 6, do not look very different from the mode shapes of a
stationary cantilever beam.

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]] 9

Fig. 5. Mode shapes of the axial and chordwise motions when δ ¼0, γ¼ 10 and α ¼ 70. (a), 1st mode (5.0562), (b) 2nd mode (32.1247), (c) 3rd mode
(74.0094), (d) 4th mode (111.3161), (e) 5th mode (134.5756) and (f) 6th mode (214.2895)

Fig. 6. Mode shapes of the flapwise motion when δ ¼ 0, γ¼ 10 and α ¼70. (a) 1st mode (11.2420), (b) 2nd mode (33.7161), (c) 3rd mode (74.7448), (d) 4th
mode (134.9932), (e) 5th mode (214.5787) and (f) 6th mode (313.6798).

The axial and chordwise motions exhibit the veering phenomenon of the natural frequency loci and the associated mode
exchange. As shown in Fig. 7, which is a magnified plot of Fig. 3, the natural frequency loci of the third and fourth modes
veer rather than cross. For the purpose of comparison, the effect of the rotating speed on the second mode shape is
examined first. Points A2, B2, C2 and D2 on the natural frequency locus of the second mode correspond to the pairs of
the dimensionless rotating speed and dimensionless natural frequency, (10, 32.1247), (20, 51.6935), (30, 73.9120) and (40,
97.7582), respectively. The second mode shapes for these points are shown in Fig. 8, where it is shown that these mode
shapes are similar to the second mode shape of the transverse beam vibration and the nodal point shifts to the right as the
rotating speed increases.
The third mode shapes for the axial and chordwise motions are obtained for the rotating speeds corresponding to points
A3, B3, C3 and D3 of Fig. 7. These mode shapes are presented in Fig. 9, which shows that the mode shapes for points A3 and B3
(when γ¼10 and 20) are quite a different from the mode shapes for points C3 and D3 (when γ¼30 and 40). This abrupt

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
10 H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]]

T uv

Fig. 7. Veering phenomenon in the dimensionless natural frequencies of the axial and chordwise motions.

Fig. 8. Second mode shapes of the axial and chordwise motions when δ ¼ 0 and α ¼70.

Fig. 9. Third mode shapes of the axial and chordwise motions when δ ¼0 and α¼ 70.

Fig. 10. Fourth mode shapes of the axial and chordwise motions when δ ¼ 0 and α ¼70.

change of the mode shape is caused by the mode exchange from the chordwise mode to the axial mode. The modes when
γ ¼10 and 20 are dominated by the chordwise motion while the modes when γ¼30 and 40 are dominated by the axial
motion. On the contrary to the third mode, the fourth mode has the mode exchange from the axial mode to the chordwise
mode when the rotating speed increases along the natural frequency locus A4–B4–C4–D4 of Fig. 7. As shown in Fig. 10, the
axial-dominated modes (when γ¼ 10 and 20) change to the chordwise-dominated modes (when γ ¼30 and 40) as the
rotating speed increases.

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]] 11

Fig. 11. Dimensionless rotating speed (γ) profile for the dimensionless time (τ).
u/L

Fig. 12. Time responses of the dimensionless axial deformation u/L at ξ¼ 1 (solid line: present study; dotted line: Ref. [4]).
v/L

Fig. 13. Time responses of the dimensionless chordwise deformation v/L at ξ ¼1 (solid line: present study; dotted line: Ref. [4]).

The time responses obtained from this study and Ref. [4] are compared when the rotating speed is prescribed. Using
the generalized-α time integration method [18] with the time step size Δt ¼ 10−4 , the time responses of a rotating beam of
δ¼0.01 and α ¼70 are computed with the equations of the present study (i.e., Eqs. (8)–(10)) and the equations of Ref. [4] (i.e.,
Eqs. (37)–(39)). The dimensionless rotating speed is prescribed by γ ¼ τ−ð5=πÞ sin ðπτ=5Þ for 0 ≤τ ≤10, γ ¼ 10 for 10 ≤τ ≤40,
and γ ¼ 50−τ þ ð5=πÞ sin ðπτ=5Þ for 40 ≤τ ≤50. This rotating speed profile is illustrated in Fig. 11. The zero initial conditions are
imposed on the axial, chordwise and flapwise deformations. No force is applied in the axial and chordwise directions but the
unit impulsive pressure is exerted in the flapwise direction. The time responses of the deformation are computed at ξ¼ 1
with the rotating speed, initial conditions and loading conditions prescribed above. The time responses obtained from the
present study and Ref. [4] are presented in Figs. 12–14, where the solid and dotted lines represent the responses computed
in the present study and Ref. [4], respectively. Comparing the responses of the two studies, the main differences are found in
the time responses of the axial and chordwise deformations, as shown in Figs. 12 and 13. The time responses of the present
study have more high-frequency vibrations in the axial and chordwise motions than the responses of Ref. [4]. Therefore, if
the responses are computed by using the equations of motion presented by Ref. [4], the high-frequency vibrations in the
axial and chordwise motions cannot be obtained.

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i
12 H. Kim et al. / Journal of Sound and Vibration ] (]]]]) ]]]–]]]

w/L

Fig. 14. Time responses of the dimensionless flapwise deformation w/L at ξ¼ 1 (solid line: present study; dotted line: Ref. [4]).

4. Conclusions

In this study, a vibration model of a rotating cantilever beam is presented and compared to other models. Depending on how
the stiffening effect of a rotating beam is considered, the modeling methods of previous studies are classified into three categories.
The methods of the first, second, and third categories are based on a hybrid set of deformation variables, the potential energy
including the work done by the centrifugal force, and the nonlinear strain/stress, respectively. The modeling methods of the first
and second categories are shown to result in the same equations of motion. Although the methods of the third category are similar
to the method of the present study, only the linear equations are considered in the modeling methods of the third category.
Therefore, the time responses cannot be obtained from the methods of the third category when a rotating beam has rotating
acceleration or applied forces. However, the proposed equations enable computation of the time responses in any circumstance.
Furthermore, it is analytically shown that the proposed equations of motion are more reliable than the equations for the
modeling methods of the first category. To support this claim, the natural frequencies and time responses computed from
this study and Ref. [4] are compared because the equations of Ref. [4] are representative equations for the modeling
methods of the first category. The differences in the natural frequencies computed in the present study and Ref. [4] are small
in the low rotating speed range, but the differences become large in the high speed range.

Acknowledgment

This work was supported by a National Research Foundation of Korea (NRF) grant funded by the Korean government
(MEST) (No. 2012-0005689).

References

[1] H.H. Yoo, S.H. Shin, Vibration analysis of rotating cantilever beams, Journal of Sound and Vibration 212 (1998) 807–828.
[2] S.M. Hamza-Cherif, Free vibration analysis of rotating flexible beams by using the Fourier p-version of the finite element method, International Journal
of Computational Methods 2 (2005) 255–269.
[3] G.P. Cai, J.Z. Hong, S.X. Yang, Model study and active control of a rotating flexible cantilever beam, International Journal of Mechanical Sciences 46 (2004)
871–889.
[4] J. Chung, H.H. Yoo, Dynamic analysis of a rotating cantilever beam by using the finite element method, Journal of Sound and Vibration 249 (2002)
147–164.
[5] S.V. Hoa, Vibration of a rotating beam with tip mass, Journal of Sound and Vibration 67 (1979) 369–381.
[6] W.D. Zhu, C.D. Mote, Dynamic modeling and optimal control of rotating Euler–Bernoulli beams, Journal of Dynamic Systems, Measurement and Control,
Transactions of the ASME 119 (1997) 802–808.
[7] J.B. Yang, L.J. Jiang, D.C.H. Chen, Dynamic modelling and control of a rotating Euler–Bernoulli beam, Journal of Sound and Vibration 274 (2004) 863–875.
[8] A.A. Al-Qaisia, B.O. Al-Bedoor, Evaluation of different methods for the consideration of the effect of rotation on the stiffening of rotating beams, Journal
of Sound and Vibration 280 (2005) 531–553.
[9] J.R. Banerjee, Dynamic stiffness formulation and free vibration analysis of centrifugally stiffened Timoshenko beams, Journal of Sound and Vibration 247
(2001) 97–115.
[10] J.R. Banerjee, Free vibration of centrifugally stiffened uniform and tapered beams using the dynamic stiffness method, Journal of Sound and Vibration
233 (2000) 857–875.
[11] E. Pesheck, C. Pierre, S.W. Shaw, Modal reduction of a nonlinear rotating beam through nonlinear normal modes, Journal of Vibration and Acoustics,
Transactions of the ASME 124 (2002) 229–236.
[12] C.L. Huang, W.Y. Lin, K.M. Hsiao, Free vibration analysis of rotating Euler beams at high angular velocity, Computers and Structures 88 (2010) 991–1001.
[13] H. Arvin, F. Bakhtiari-Nejad, Non-linear modal analysis of a rotating beam, International Journal of Non-linear Mechanics 46 (2011) 877–897.
[14] M.A. Crisfield, Non-linear Finite Element Analysis of Solids and Structures, vol. 1, John Wiley and Sons, Chichester, 1997.
[15] J. Chung, C.S. Han, K. Yi, Vibration of an axially moving string with geometric non-linearity and translating acceleration, Journal of Sound and Vibration
240 (2001) 733–746.
[16] K. Lee, Y. Cho, J. Chung, Dynamic contact analysis of a tensioned beam with a moving mass–spring system, Journal of Sound and Vibration 331 (2012)
2520–2531.
[17] J. Chung, J.E. Oh, H.H. Yoo, Non-linear vibration of a flexible spinning disc with angular acceleration, Journal of Sound and Vibration 231 (2000) 375–391.
[18] J. Chung, G.M. Hulbert, A time integration algorithm for structural dynamics with improved numerical dissipation: the generalized-α method, Journal
of Applied Mechanics—Transactions of the ASME 60 (1993) 371–375.

Please cite this article as: H. Kim, et al., Dynamic model for free vibration and response analysis of rotating beams, Journal
of Sound and Vibration (2013), http://dx.doi.org/10.1016/j.jsv.2013.06.004i

You might also like