You are on page 1of 12

Scientia Iranica B (2012) 19 (4), 1105–1116

Sharif University of Technology


Scientia Iranica
Transactions B: Mechanical Engineering
www.sciencedirect.com

Vibration of Euler–Bernoulli and Timoshenko beams in large overall


motion on flying support using finite element method
H. Zohoor a,b,∗ , F. Kakavand c
a
Center of Excellence in Design, Robotics and Automation, Sharif University of Technology, Tehran, P.O. Box 11155-9567, Iran
b
The Academy of Sciences of IR Iran, Tehran, P.O. Box 19735-167, Iran
c
School of Mechanical Engineering, Sharif University of Technology, Tehran, Iran

Received 18 December 2011; revised 18 February 2012; accepted 15 May 2012

KEYWORDS Abstract The equations of motion for a beam on a flying support for Euler–Bernoulli and Timoshenko
Euler–Bernoulli; Beam Theories is derived. In modeling and attempting to have an accurate model at high speeds, a stretch
Timoshenko; variable instead of conventional axial deformation is used. For a planar rotating beam and a spatial rotating
Beam; beam, equations of motion are lineralized and verified. Finite element and Newmark direct integration
Large overall motion; methods are employed for numerical simulations.
Flying support; © 2012 Sharif University of Technology. Production and hosting by Elsevier B.V.
Finite element; Open access under CC BY-NC-ND license.
Stretch.

1. Introduction of flexible multibody systems by finite element or assumed


mode methods were based on the assumption that small
A beam with flying support is a general case of at least two deformations in the flexible bodies do not affect rigid body
types of problem in dynamics. The first is rotating beams and motion significantly [1]. There are many publications listed in
the second flexible manipulators. Rotating cantilever beams Ref. [2] that solved the rotating beam problem with several
are found in several practical engineering examples, such methods, such as the finite element or assumed mode. The
as turbine blades and aircraft rotary wings. For reliable and acceleration and reaction forces were obtained from rigid
body motion analysis, and introduced to the linear elasticity
economic design of the structures, it is necessary to estimate
problem as external forces for computing deflections. The
the dynamic characteristics of those structures accurately and
elastic deformation is then superimposed on the rigid body
efficiently. Similarly, flexible manipulators are found in robotic
motion. These dynamic models, however, do not yield accurate
systems design, as are flexible gyroscopes, in general, in flexible
results, since they do not provide for the coupling of the rigid
multibody systems. Therefore, having a simple and accurate
and elastic motion. A hybrid-coordinate formulation, based on
dynamic model for estimating their dynamical behavior is
identifying the configuration of each flexible body by means
necessary.
of two coordinate systems, is developed in [3]. They employed
Dynamic analysis of flexible multibody systems has gained a reference coordinate system for defining the body fixed
attention from researchers in the past decades. Earlier models frame (elastic coordinate). The deformations are described
in the body fixed frame and then the rigid body motion
∗ Corresponding author at: Center of Excellence in Design, Robotics and and deformations are solved simultaneously. Kane et al. [4]
Automation, Sharif University of Technology, Tehran, P.O. Box 11155-9567, Iran. and Yoo et al. [5] described a conventional hybrid-coordinate
E-mail address: hzohoor@ias.ac.ir (H. Zohoor). formulation in which the Cartesian deformation fails to describe
Peer review under responsibility of Sharif University of Technology. the motion-induced stiffness terms and provides erroneous
dynamic results in cases of high rotating speed (Large overall
motion). Yoo et al. [5] showed, in detail, the use of conventional
axial deformation, since, in linearization of potential strain
energy, in some terms of retaining force loss, it can cause a
1026-3098 © 2012 Sharif University of Technology. Production and hosting by Elsevier B.V. Open access under CC BY-NC-ND license.
doi:10.1016/j.scient.2012.06.019
1106 H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116

diverging solution at high speed. A further improvement in 2. Equations of motion for Euler–Bernoulli model
the formulation can be achieved by employing non-Cartesian
deformation variables to derive equations of motion for a In this section, the geometric nonlinear formulation of a
thin beam or a thin plate [3,5–7]. With the inclusion of the three-dimensional beam undergoing large overall motion is
foreshortening deformation, the motion-induced stiffness term established, based on the following assumptions. The beam has
is derived, which is the lost term in the previous modeling homogeneous and isotropic material properties, the elastic and
method. Using a stretch variable provides a simple expression centroidal axes in the cross section of a beam coincide, so that
of strain energy. Then, in the linearization of strain energy
the effects due to eccentricity are not considered.
there is no lost term, and the required retaining force is
A three-dimensional beam is shown in Figure 1. Two coor-
available. Therefore, at high speed, the model gives accurate
dinate systems are introduced to describe the motion of the
and converging solution. It has been proved that this method
is as efficient as the conventional linear modeling method, and beam: The global coordinate system, O0 − X0 Y0 Z0 , and the body-
as accurate as nonlinear modeling methods [5]. fixed coordinate system, Ob − Xb Yb Zb .
On the other hand, there are many publications on ro- The position vector of point k on the central line of the beam
tating beam dynamics. Due to the progress of computing can be defined with respect to the Ob − Xb Yb Zb as:
technologies, a large number of papers based on numerical −

approaches have been published [8]. For instance, in [8–10], ρ Ok = (x + u) î + ν ĵ + wk̂, (1)
approximation methods for the modal analysis of rotating −

beams were employed. More complex shapes, and the effects in which vector U = uî + ν ĵ + w k̂ is the deformation vector
of beams, were also considered. The effects of tip mass [11,12], of point k, with respect to the body-fixed coordinate. All vectors
elastic foundation and cross-sectional variation [13], shear de- are in terms of body fixed coordinate unit vectors.
formation [14,15], pre-twist and orientation of a blade [16], The absolute velocity of point k is given by:
and the gyroscopic damping effect [17] on the modal char- −
→ −
→ −→
acteristics of rotating cantilever beams, were studied. Survey V k = V o+ V r +−

ω ×−

ρ Ok . (2)
papers for the vibration analysis of rotating structures are avail-
able [1,18]. The most widely used modeling method for the Using the following relations:
transient analysis of structures is the classical linear model- −

ing method [19–21]. This modeling method employs Carte- V O = VOx î + VOy ĵ + VOz k̂, (3)
sian deformation variables and linear Cauchy strain measures. −

Similar to the aforementioned concept concerning high speeds, V r = u̇î + ν̇ ĵ + ẇ k̂, (4)
Yoo et al. [5,8] showed that for high speeds, the stretch vari- −

ω = ωx î + ωy ĵ + ωz k̂, (5)
able should be considered in modeling (for more information
see [5]). Chung and Yoo [8] used a finite element method for a −
→ −

where V O and ω are the velocity of the base point of the
planar rotating Euler–Bernoulli beam. They found the time re- beam and the angular velocity of the beam, respectively, and
sponse and stresses for a prescribed motion.
substituting Eqs. (1), (3), (4) and (5) into Eq. (2), the velocity of
Liu and Hong [1,22] have developed a matrix presentation of
point k leads to:
spatial Euler–Bernoulli and Planar beams based on the assumed
mode method. They employed a non-Cartesian deformation −

= VOx + u̇ + wωy − νωz î
 
V k
variable for taking into account the motion-induced stiffness,
+ VOy + ν̇ + (x + u) ωz − wωx ĵ
 
and used a forward recursive formulation for driving the
dynamic equations of a flexible link system.
+ VOz + ẇ + νωx − (x + u) ωy k̂.
 
If, in a certain application, the rotary inertia and shear (6)
deformation effects are not significant, an analysis based on
the Euler–Bernoulli beam theory is sufficient. However, the Therefore, the kinetic energy of the beam can be written as:
error of using this theory can be significant for thicker beams.

1 −
→ − →
The error may also be significant in the calculation of natural T = ρ V k . V k dV ,
2 V
frequencies of vibration at higher modes and time response. In
L
the present paper, to provide a more general derivation that is

1
valid for both thin and thick beams, and is also accurate enough T = ρA [(VOx + u̇ + wωy − νωz )2
2 0
at high frequencies and high speeds, the Timoshenko beam
model is used. Rao and Gupta [15] used the Timoshenko model
+ (VOy + ν̇ + (x + u)ωz − wωx )2
for a rotating beam. They solved a twisted tapered Timoshenko + (VOz + ẇ + νωx − (x + u)ωy )2 ]dx, (7)
beam. Kyung-Su Na, Ji-Hwan Kim [23] had solved a multilink
where A and ρ are the cross section area and material density
system using the Timoshenko theory, but they did not take into
of the beam, respectively.
account the stretch variable.
Zohoor and Khorsandijou [24,25] derived the enhanced Let s be the neutral axis stretch. The geometric relation
nonlinear 3D-Euler–Bernoulli beam upon an exact strain between u and s, known as the Von–Karman relation, can be
field, and obtained a nonlinear dynamic model of a flying written as:
manipulator with two revolute joints and two highly flexible u = s − hν − hw , (8)
links [26]. In the present paper, the equations of motion, for
a three-dimensional rotating beam on flying support, using a where:
stretch variable in a non-Cartesian coordinate system, for both
∂ν 2
  x
Euler–Bernoulli and Timoshenko beams, have been derived. 1
hν = dη, (9)
Numerical examples are presented to examine the validity of 2 0 ∂η
the equations of motion and a comparison of the Timoshenko
1 x ∂w
  2
and Euler–Bernoulli models. The finite element and Newmark
hw = dη, (10)
direct integration methods have been employed. 2 0 ∂η
H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116 1107


= −ρ A aOy + xαz + ωz VOx

− ωx VOz + xωy ωx + pv , (17)

ρ A ẅ − 2ωy ṡ − h˙ν − h˙w + (s − hν − hw ) −α y +ωz ωx
   

+ 2ν̇ωx + ναx − w ωx2 + ωy2 + νωy ωz
 

∂ ∂w x ∂ w
    4 
− ρA D (η, t ) dη + EIy
∂x ∂x L ∂ x4

= −ρ A aOz − x(αy

− ωz ωx ) + ωx VOy − ωy VOx + pw , (18)

where:
Figure 1: Three-dimensional beam [23].  
D = aOx + s̈ − ḧν − ḧw + ẇωy + wαy − ν̇ωz − ναz
where η is the dummy variable. Similarly, the time derivative − ωz VOy + ν̇ + (x + s − hν − hw ) ωz − wωx
 
of u is given by:

+ ωy VOz + ẇ + νωx − (x + s − hν − hw ) ωy .

(19)
u̇ = ṡ − ḣν − ḣw , (11)
where the superposed dots indicate the derivative, with respect D is the axial force density induced by the rotation of the beam.
to time and [8]; In the above equations, aox , aoy , aoz , αox , αoy and αoz are the
scalar derivatives of Vox , Voy , Voz , ωox , ωoy and ωoz , respectively.
x
∂ν ∂ ν̇

ḣν = dη, (12) After linearization, they will be as follows:
0 ∂η ∂η 
ρ A s̈ + 2ẇωy + wαy − 2ν̇ωz − ναz − s ωz2 + ωy2
 
x
∂w ∂ ẇ

ḣw = dη. (13)
0 ∂η ∂η  ∂ 2s
+ wωx ωz + νωx ωy − EA 2
Using Eqs. (8)–(13), the kinetic energy equation (7) leads to: ∂x
= −ρ A aOx − x ωz + ωy + ωy VOz − ωz VOy + ps ,
  2 2
 
1
 L
2 (20)
ρA VOx + ṡ − ḣν − ḣw + wωy − νωz

T = 
2 ρ A ν̈ + 2ωz ṡ + s αz + ωx ωy
 
0
2
+ VOy + ν̇ + (x + s − hν − hw ) ωz − wωx
 
− 2ẇωx − wαx − ν ωx2 + ωz2 + wωy ωz
 
 2 
+ VOz + ẇ + νωx − (x + s − hν − hw ) ωy dx. (14) ∂ ∂v x
  
∂ ν
 4 
− ρA D (η, t ) dη + EIz
After introducing stretch variable, s, the strain potential energy ∂x ∂x L ∂ x4
can be written as follows: = −ρ A aOy + xαz + ωz VOx − ωx VOz + xωy ωx + pv ,
 
(21)
 L  2 2 
∂s ∂ 2ν ∂ 2w
2   
1 ρ A ẅ − 2ωy ṡ −α y +ωz ωx
 
U = E A + Iz + Iy dx. (15)
2 0 ∂x ∂ x2 ∂ x2 
+ 2ν̇ωx + ναx − w ωx2 + ωy2 + νωy ωz
 
Using the Hamilton Principle and integrating by parts, and
collecting all the items of the integrand with respect to δ s, δv ∂ ∂w x ∂ w
    4 
and δw , the coefficient of δ s, δv and δw result in the following − ρA D (η, t ) dη + EIy
∂x ∂x L ∂ x4
equations of motion:
= −ρ A aOz − x(αy − ωz ωx ) + ωx VOy − ωy VOx + pw ,
 
 (22)
ρ A s̈ − ḧν + ḧw + 2ẇωy + wαy − 2ν̇ωz − ναz
 
where:
− (s − hν − hw ) ωz2 + ωy2
 
D = aOx − ωz VOy + xωz + ωy VOz − xωy .
    
(23)
 ∂ 2s
+ wωx ωz + νωx ωy − EA 2 ps , pv and pw are applied forces per unit length in s, v and w
∂x
directions, respectively.
= −ρ A aOx − x ωz2 + ωy2 + ωy VOz − ωz VOy + ps ,
   
(16)

ρ A ν̈ + 2ωz ṡ − h˙ν − h˙w + (s − hν − hw ) αz + ωx ωy
   
3. Equations of motion for Timoshenko beam model

− 2ẇωx − wαx − ν ωx2 + ωz2 + wωy ωz
 
In this section, the geometric nonlinear formulation of a
three-dimensional Timoshenko Beam undergoing large overall
∂ ∂v x ∂ ν
    4 
− ρA D (η, t ) dη + EIz motion is established, based on some assumptions similar to
∂x ∂x L ∂ x4 those of the Euler–Bernoulli beam in the previous section.
1108 H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116

For Timoshenko beam kinetic energy, the velocity of an



+ 2ν̇ωx + ναx − w ωx2 + ωy2 + νωy ωz
 
arbitrary point of the cross section of the beam can be written
as: ∂ ∂w x
  

→ − ρA D (η, t ) dη
= VOx + u̇ + wωy − νωz î ∂x ∂x L
 
V g

∂ w ∂ψw
 2 
+ VOy + ν̇ + (x + u) ωz − wωx ĵ
 
− µAG −
∂ x2 ∂x
+ VOz + ẇ + νωx − (x + u) ωy k̂
 

= −ρ A aOz − x(αy − ωz ωx )
+ −yψ̇ν + z ψ̇w î.
 
(24)

Then, the kinetic energy can be written as: + ωx VOy − ωy VOx + pw , (31)
 L
1 −
→ − → ∂ 2 ψw ∂ 2 ψw

∂w

T = ρ V g . V g dAdx, (25) ρ Iy − EI y − µ AG − ψw = 0, (32)
2 0 ∂t2 ∂ x2 ∂x
 L
1 ∂ 2 ψν ∂ 2 ψν

∂ν

T = ρ (Vkx2 + Vky
2
+ Vkz2 + y2 ψ̇ν2 + z 2 ψ̇w2 ρ Iz − EIz − µAG − ψν = 0. (33)
2 0 A ∂t2 ∂ x2 ∂x
− 2yz ψ̇ν ψ̇w + 2Vkx (−yψ̇ν + z ψ̇w ))dAdx. (26)
After linearization, the above equations lead to:
Assuming that the cross section of the beam is homogenous and 
ρ A s̈ + 2ẇωy + wαy − 2ν̇ωz − ναz − s ωz2 + ωy2
 
symmetric, Eq. (26) leads to:

1
 L  ∂ 2s
ρ Iz ψ̇v2 + Iy ψ̇w2 dx, + wωx ωz + νωx ωy − EA 2
 
T = TE + (27)
2 0
∂x
= −ρ A aOx − x ωz + ωy + ωy VOz − ωz VOy + ps,
  2 2
 
where TE are the kinetic energy terms of the Euler–Bernoulli (34)
theory. In the same manner, strain potential energy can be 
ρ A ν̈ + 2ωz ṡ + s αz + ωy ωx − 2ẇωx − wαx
 
written as follows:

1
 L 
∂s 2
 
∂ψν 2
 
∂ψw 2
   ∂
− ν ωx2 + ωz2 + wωy ωz − ρ A
 
U = E A + Iz + Iy dx ∂x
2 0 ∂x ∂x ∂x
∂v ∂ ν ∂ψν
  x   2 
 L  2  × D (η, t ) dη − µAG −
∂w ∂ν
2 
1 ∂x L ∂ x2 ∂x
+ µAG − ψw + − ψν dx, (28)
2 ∂x ∂x

0
= −ρ A aOy + xαz + ωz VOx − ωx VOz
where µ is the shear factor (for more information, see [27]). 
The use of the Hamilton principle and integrating by parts the + xωy ωx + pv , (35)
nonlinear equations of motion are given as follows: 
ρ A ẅ − 2ωy ṡ + s −α y +ωz ωx + 2ν̇ωx + ναx
  
ρ A s̈ − ḧν + ḧw + 2ẇωy + wαy − 2ν̇ωz − ναz
 
 ∂
− w ωx2 + ωy2 + νωy ωz − ρ A
  
− (s − hν − hw ) ωz2 + ωy2 + wωx ωz + νωx ωy − EA
 
∂x
∂w x ∂ w ∂ψw
    2 
∂ 2s 
× D (η, t ) dη − µAG −
× 2 = −ρ A aOx − x ωz2 + ωy2
 
∂x ∂x L ∂ x2 ∂x

= −ρ A aOz − x(αy − ωz ωx )

+ ωy VOz − ωz VOy + ps , (29)

+ ωx VOy − ωy VOx + pw ,

ρ A ν̈ + 2ωz ṡ − h˙ν − h˙w (36)
 

+ (s − hν − hw ) αz + ωy ωx ∂ 2 ψw ∂ 2 ψw ∂w
   
ρ Iy − EI y − µ AG − ψ w = 0, (37)
∂t2 ∂ x2 ∂x

− 2ẇωx − wαx − ν ωx2 + ωz2 + wωy ωz
 
∂ 2 ψν ∂ 2 ψν ∂ν
 
∂ ∂v x ρ Iz − µAG − ψν = 0.
  
− EIz (38)
− ρA D (η, t ) dη ∂t2 ∂ x2 ∂x
∂x ∂x L
In the next sections, two numerical examples are presented.
∂ ν ∂ψν
 2 
− µAG −
∂ x2 ∂x 4. A planar rotating beam

= −ρ A aOy + xαz + ωz VOx Consider a rotating beam that has been solved in [8], and
 shown in Figure 2. For this problem, the following relations are
− ωx VOz + xωy ωx + pv , (30) given:
 ωx = 0, ωy = 0,
ρ A ẅ − 2ωy ṡ − h˙ν − h˙w
 
ωz = Ω , αx = 0,
+ (s − hν − hw ) −α y +ωz ωx
 
αy = 0, αz = Ω̇ ,
H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116 1109

For the Timoshenko model, Eqs. (45) and (46) are uncoupled
from others.
Therefore, the following equations have been solved in a
planar sense:

∂ 2s
ρ A s̈ − 2ν̇ Ω − ν Ω̇ − Ω 2 s − EA 2 = ρ A (a + x) Ω 2 , (47)
 
∂x
ρ A ν̈ + 2Ω ṡ + Ω̇ s − Ω 2 ν − ρ AΩ 2
 

∂ ∂v ∂ ν
    4 
1 2
a (L − x) + L − x2

× + EIz
∂x ∂x 2 ∂ x4
= −ρ AΩ̇ (a + x) . (48)
Figure 2: Planar rotating beam.
For the Timoshenko model:

∂ 2s
VOx = 0, VOy = aΩ , ρ A s̈ − 2ν̇ Ω − ν Ω̇ − Ω 2 s − EA 2 = ρ A (a + x) Ω 2 ,
 
(49)
∂x
VOz = 0, aOx = 0,
ρ A ν̈ + 2Ω ṡ + Ω̇ s − Ω 2 ν − ρ AΩ 2
 
aOy = aΩ̇ , aOz = 0.
∂ ∂v
  
1 2
a (L − x) + L − x2

Then, the Euler–Bernoulli model of the linear equations of ×
∂x ∂x 2
motion will be as follows:
∂ v ∂ψv
 2 
∂ 2s − µAG − = −ρ AΩ̇ (a + x) , (50)
ρ A s̈ − 2ν̇ Ω − ν Ω̇ − Ω 2 s − EA 2 = ρ A (a + x) Ω 2 , ∂ x2 ∂x
 
(39)
∂x
∂ ∂ 2 ψν ∂ 2 ψν ∂ν
 
ρ A ν̈ + 2Ω ṡ + Ω̇ s − Ω ν − ρ AΩ 2
2
ρ Iz − µAG − ψν = 0.
 
− EIz (51)
∂x ∂t2 ∂ x2 ∂x
∂v ∂ ν
    4 
1 2 The boundary conditions for the Euler–Bernoulli model are [8]:
a ( L − x) + 2

× L −x + EIz
∂x 2 ∂ x4
∂v
= −ρ AΩ̇ (a + x) , (40) s=v= = 0, at x = 0, (52)
∂x
∂ ∂w
 
ρ Aẅ − ρ AΩ 2 a ( L − x) ∂s ∂ 2v ∂ 3v
∂x ∂x = 2 = 3 = 0 at x = L, (53)
∂x ∂x ∂x
∂ w
  4 
1 and for the Timoshenko model:
+ L2 − x 2 = 0.

+ EIy (41)
2 ∂ x4 s = v = ψ = 0, at x = 0, (54)
The Timoshenko model of the linear equations of motion will ∂s ∂ψ ∂v
be as follows: = = − ψ = 0 at x = L. (55)
∂x ∂x ∂x
∂ 2s
ρ A s̈ − 2ν̇ Ω − ν Ω̇ − Ω 2 s − EA 2 = ρ A (a + x) Ω 2 ,
 
(42)
∂x 4.1. Finite element model

ρ A ν̈ + 2Ω ṡ + Ω̇ s − Ω ν − ρ AΩ 2
2
 
∂x Assuming the deformation vector and shape functions of the
∂v Euler Bernoulli beam as follows [8]:
  
1 2
a ( L − x) + L − x2

×
∂x 2 dseν = {se , νe , θe , se+1 , νe+1 , θe+1 }T , (56)
∂ v ∂ψv
 2 
− µAG − = −ρ AΩ̇ (a + x) , (43) Ns = {(xe+1 − x)/he , 0, 0, (x − xe )/he , 0, 0} , T
(57)
∂ x2 ∂x 
Nν = 0, (x − xe+1 )2 (2x − 3xe + xe+1 ) /h3e ,
∂ ∂w
  
1 2
ρ Aẅ − ρ AΩ 2 a ( L − x) + L − x2

∂x ∂x 2 (x − xe ) (x − xe+1 )2 /h2e , 0, − (x − xe )2
T
∂ w ∂ψw
 2 
− µAG − = 0, (44)
× (2x + xe − 3xe+1 )/h3e , (x − xe )2 (x − xe+1 )/h2e , (58)
∂ x2 ∂x
he = xe+1 − xe. (59)
∂ 2 ψw ∂ 2 ψw ∂w
 
ρ Iy − EI y − µ AG − ψ w = 0, (45) For the Timoshenko Beam, considering static equations and
∂t2 ∂ x2 ∂x imposing end conditions for variables, we will have (see
Appendix):
∂ 2 ψν ∂ 2 ψν ∂ν
 
ρ Iz − EIz − µAG − ψν = 0. (46) 
(x − xe+1 )
∂t2 ∂ x2 ∂x Nν = 0 −
(xe − xe+1 )
It is interesting that Eqs. (39) and (40) are coupled to each other,
2α x2 − α xxe+1 − α x2e+1 + 12 − 3α xe (x − xe+1 )
 
while Eq. (41) is not coupled with the other equations. The ×
α x2e − 2α xe xe+1 + α x2e+1 − 12
 
solutions of Eqs. (39) and (40) are described in Ref. [8] in detail.
1110 H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116

(x − xe ) (x − xe+1 ) (α xxe+1 − α xxe − α x2e+1 + α xe xe+1 + 6) where:


− 0
(xe − xe+1 ) (α x2e − 2α xe xe+1 + α − 12) x2e+1 d = {s2 , ν2 , ψ2 , s3 , ν3 , ψ3 , . . . , sN +1 , νN +1 , ψN +1 }T (73)
(x − xe ) 2α x − α xxe − α xe + 12 − 3α xe+1 (x − xe )
 2 2

× M = N
e=1 Ame , C = N
e=1 Ac e ,
(xe − xe+1 ) α x2e − 2α xe xe+1 + α x2e+1 − 12
 
1 2
 K1 = N
e=1 Ake , K2 = N
e=1 Ake , (74)
(x − xe ) (x − xe+1 ) (α xxe − α x2e − α xxe+1 + α xe xe+1 + 6) 3
×
(xe − xe+1 ) α x2e − 2α xe xe+1 + α x2e+1 − 12
(60) K3 = N
e=1 Ake , F = N
e=1 Af e
 where d is the global deformation matrix, M , C , K and F are
6α (x − xe ) (x − xe+1 ) (x − xe+1 )
Nψ = 0 − global mass, gyroscopic, stiffness, and force matrixes, respec-
(xe − xe+1 ) (α x2e − 2α xe xe+1 + α x2e+1 − 12) (xe − xe+1 ) tively, and A is the assembly operator. For simplicity, in this
3α (x − xe ) (x − xe+1 ) section, we use ψ instead of ψv .
+ 0
(α x2e − 2α xe xe+1 + α x2e+1 − 12)  xe+1
me = ρ A Ns NsT + Nν NνT dx
 
6α (x − xe ) (x − xe+1 )
(x − xe )
× − xe
(xe − xe+1 ) (α x2e − 2α xe xe+1 + α x2e+1 − 12) (xe − xe+1 )  L
 + ρ Iz Nψ NψT dx, (75)
3α (x − xe ) (x − xe+1 ) 0
+ (61)
(α x2e − 2α xe xe+1 + α x2e+1 − 12)  xe+1
ce = 2ρ A Nν NsT − Ns NνT dx,
 
(76)
where: xe

µAG  xe+1

.

a= (62) k1e = −ρ A Ns NsT + Nν NνT − a (L − x)
EIz
xe
Expressing the weak form of the Euler–Bernoulli equation and 1  dN dN T 
ν ν
imposing a boundary condition as described in [8], the finite + (L − x2 ) 2
dx, (77)
2 dx dx
element equations of the Euler–Bernoulli beam will be as
 xe+1
follows [8]:
k2e = ρ A Nν NsT − Ns NνT dx,
 
(78)
Msν d̈sν + 2Ω Gsν ḋsν xe

+ Ksν + Ω 2 (Ssν − Msν ) + Ω̇ Gsν dsν = fsν , xe+1


dNs dNsT
     
(63) dNv
k3e = EA − Nψ + µAG
where dsν is the global deformation matrix, Msν , Gsν , Ksν , Ssν xe dx dx dx
T T 
and fsν are global mass, gyroscopic, stiffness, motion induced dN ψ dN ψ

dNv
stiffness and force matrixes, respectively, and A is the assembly × − Nψ + EIz dx, (79)
dx dx dx
operator.  xe+1
fe = ρ A Ω (a + x) Ns − Ω̇ (a + x) Nν dx.
 2 
dsν = {s2 , ν2 , θ2 s3 , ν3 , θ3 , . . . , sN +1 , νN +1 , θN +1 } T
(64) (80)
xe
sν sν
Msν = Ne=1 Ame , Gsν = Ne=1 Ag e ,
sν sν
Ksν = Ne=1 Ake , Ssν = Ne=1 Ase , (65) 4.2. Time response

fsν = N
e=1 Af e ,
In this section, the time response of a planar rotating beam
in which the element mass, gyroscopic, stiffness, motion described in the previous section has been solved using the
induced stiffness and force matrixes are as follows [8]: Newmark direct integration method, introducing the following
 xe+1 parameters [8]:
mseν = ρ A Ns NsT + Nν NνT dx,
 
(66) 
xe τ = t /T , γ = T Ω, α= AL2 /Iz , δ = a/L. (81)
 xe+1
Consider the given angular velocity the same as [8], as follows:
gesν = ρ A Nν NsT − Ns NνT dx,
 
(67)
xe
 5 πτ
τ − π sin 5 0 ≤ τ ≤ 10


xe+1
dN s dN Ts d2 Nν d2 NνT
  
kseν = EA + EIz dx (68) γ = 10 10 ≤ τ ≤ 40
dx dx dx2 dx2 50 − τ + 5 sin πτ
xe

40 ≤ τ ≤ 50.

 xe+1

1

dN ν dN Tν π 5
sseν = ρA a (L − x) + (L − x ) 2 2
dx (69) Figures 3–6 show the response of the system for δ = 0.1.
xe 2 dx dx
Equations have been solved with 100 elements. For comparison,
 xe+1
two cases have been considered; the first for a thin beam,
fesν = ρ AΩ 2 (a + x) Ns − ρ AΩ̇ (a + x) Nν dx.
 
(70)
xe
α = 70, and the second for a thick beam, α = 8.
We can define the axial and bending stresses for the Euler
Using a similar way for Timoshenko equations, the finite ele-
Bernoulli model as follows:
ment equations of motion for the Timoshenko beam model will
be expressed as: ∂s
σs = E , (82)
∂x
M d̈ + C ḋ + Kd = F , (71)
∂ 2v
K = Ω 2 K1 + Ω̇ K2 + K3 , (72) σv = −ER∗ . (83)
∂ x2
H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116 1111

Figure 3: Tip deflection of beam for thin beam with T = 47.6 ms.
Figure 6: Tip stretch of beam for thick beam with T = 33.3 ms and R∗ = 0.1.

Figure 4: Tip stretch of beam for thin beam with T = 47.6 ms.
Figure 7: Axial stress of beam for thick beam with T = 33.3 ms and R∗ = 0.1.

Figure 5: Tip deflection of beam for thick beam with T = 33.3 ms.

As it appears in Figures 3 and 4, the result of the Timoshenko Figure 8: Bending stress of beam for thick beam with T = 33.3 ms and
and Euler–Bernoulli beams for a thin beam are completely R∗ = 0.1.
coincidental. But, as it appears in Figures 5 and 6, the results for
the Timoshenko and Euler–Bernoulli beams for a thick beam are one half of the height of the cross section. The following figures
different. show a comparison of axial and bending stresses at x = 0.
Similarly, the bending stress for Timoshenko is as follows: It is clear that the response for thin beams for Timoshenko
∂ψ and Euler–Bernoulli beams are completely coincidental. But
σv = −ER∗ . (84) there are some differences for a thick beam. It is evident that the
∂x Timoshenko model for a thick beam is more accurate than that
Using Eqs. (57), (58) and (61), stresses at any point of the beam of an Euler–Bernoulli model. It is interesting that stresses for
can be calculated. In the above equation, R∗ is the radius or both beams are, approximately, identical (see Figures 7 and 8).
1112 H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116

∂ 2v ∂ψv
 
− µAG −
∂x 2 ∂x
= −ρ Axωy ωx, (90)
ρ A ẅ − 2ωy ṡ − sαy + 2ν̇ωx + ναx − w ωx2 + ωy2
  

∂ ∂w x
  
− ρA D (η, t ) dη
∂x ∂x L
∂ w ∂ψw
 2 
− µAG −
∂ x2 ∂x
= ρ Axαy, (91)
∂ ψw
2
∂ ψw ∂w
2
 
ρIy − EIy − µAG − ψw = 0, (92)
∂t2 ∂ x2 ∂x
∂ 2 ψν ∂ 2 ψν ∂ν
 
ρ Iz − EI z − µ AG − ψ ν = 0. (93)
∂t2 ∂ x2 ∂x
Figure 9: Spatial rotating beam [23].

5.1. Finite element model:


5. A spatial rotating beam
Using the shape function in Eqs. (57) and (58) for the
Consider a famous spatial rotating beam that has been
Euler–Bernoulli beam, and imposing boundary conditions of a
solved in the literature (e.g. [5,23]) and is shown in Figure 9.
cantilever beam, the FE model can be written as follows:
They solve this problem using the assumed mode method for
the Euler–Bernoulli beam theory. In this section, it has been M d̈ + C ḋ + Kd = F , (94)
solved for both Euler–Bernoulli and Timoshenko models using
where:
FEM.
s2 , ν2 , θv 2 , w2 , θw2 , . . . , sN +1 ,

Let R = 0 in Figure 9, then the following relations are given: d =
T
VOx = 0, VOy = 0, νN +1 , θvN +1 , wN +1 , θwN +1 , (95)
VOz = 0, aOx = 0, M = N
, C = N
,
e=1 Ame e=1 Ac e
aOy = 0, aOz = 0, ωz = 0, αz = 0. (96)
K = N
e=1 Ake , F = N
e=1 Af e ,
Therefore, equations of motion for Euler–Bernoulli and Timo-
where d is the global deformation matrix, M , C , K and F are
shenko beams are as follows:
global mass, gyroscopic, stiffness, and force matrixes, respec-
ρ A s̈ + 2ẇωy + wαy − sωy2 + νωx ωy
 
tively, and A is the assembly operator.

∂ 2s xe+1

− EA 2 = ρ Axωy2 , me = ρ A Ns NsT + Nv NvT + Nw Nw
T
dx,
 
(85) (97)
∂x xe
ρ A ν̈ + sωx ωy − 2ẇωx − wαx − νωx2
   xe+1 
ce = 2ρ A ωy Ns NwT − Nw NsT
 
∂ ∂v x ∂ ν
    4 
− ρA D (η, t ) dη + EIz x
e
∂x ∂x L ∂ x4

+ ωx Nw NvT − Nv NwT dx, (98)
= −ρ Axωy ωx , (86)  xe+1 
ρ A ẅ − 2ωy s −
˙ sαy + 2ν̇ωx + ναx − w ωx2 + ωy2 ke = ρ A αy Ns NwT − Nw NsT
    
xe
∂ ∂w x ∂ w
    4 
+ αx Nw NvT − Nv NwT − ωx2 Nv NvT + Nw NwT
   
− ρA D (η, t ) dη + EIy
∂x ∂x L ∂ x4 − ωy2 Ns NsT + Nv NvT
 

= ρ Axαy, dNv dNvT T


 
(87) dNw dNw
+B +
where: dx dx dx dx

+ ωx ωy Ns Nv + Nv NsT dx + E
T
 
D = −xωy2 . (88)
 xe+1 
And for Timoshenko: dNs dNsT d2 Nv d2 NvT
× A + Iz 2
ρ A s̈ + 2ẇωy + wαy − sωy2 + νωx ωy dx2
 
xe dx dx dx
d2 N w d2 N wT

∂ 2s + Iy 2 dx, (99)
− EA 2 = ρ Axωy2 , (89) 2
dx dx
∂x  xe+1
ρ A ν̈ + sωx ωy − 2ẇωx − wαx − νωx2
 
fe = ρ A xωy2 Ns − xωx ωy Nv + xαy Nw dx,
 
(100)
∂ ∂v
  x  xe
− ρA D (η, t ) dη
∂x ∂x L where B = − 12 ωy2 x2 − L2 .
 
H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116 1113

For Timoshenko, the FE model will be as follows:

M d̈ + C ḋ + Kd = F , (101)

where:

s2 , ν2 , ψv 2 , w2 , ψw2 , . . . , sN +1 ,

d =
T
νN +1 , ψvN +1 , wN +1 , ψwN +1 , (102)

M = Ne=1 Ame , C = Ne=1 Ac e ,


(103)
K = Ne=1 Ake , F = Ne=1 Af e ,

where d is the global deformation matrix, M , C , K and F


are global mass, gyroscopic, stiffness, and force matrixes,
respectively, and A is the assembly operator.
xe+1 Figure 10: Tip deflection in v direction for thin beam.


ke = ρ A αy Ns NwT − Nw NsT
 
xe
+ αx Nw NvT − Nv NwT − ωx2 Nv NvT + Nw NwT
   

dNv dNvT

− ωy2 Ns NsT + Nv NvT + B
 
dx dx
T
 
dNw dNw
+ ωx ωy Ns Nv + Nv Ns dx
T T
 
+
dx dx
 xe+1 
dNs dNsT
 
dNv
+ EA + µAG − Nψv
dx dx dx
xe T  
dNv dNw
× − Nψv + µAG − Nψw
dx dx
 T 
dNw
× − Nψw dx, (104)
dx
 xe+1 
ce = 2ρ A ωy Ns NwT − Nw NsT Figure 11: Tip deflection in w direction for thin beam.
 
x
e 
+ ωx Nw NvT − Nv NwT dx, (105)
 xe+1
fe = ρ A xωy Ns − xωx ωy Nv + xαy Nw dx,
 2 
(106)
xe
 xe+1
me = ρ A Ns NsT + Nv NvT + Nw Nw
T
 
dx
xe  xe+1  xe+1
+ ρ Iz Nψv Nψv dx + ρ Iy
T
Nψw NψT w dx. (107)
xe xe

5.2. Time response

Apply the following prescribed angular velocity:


Ωt Ω sin 2π t

∂= − (rad/s), 0<t ≤T
T 2π T
Ω (rad/s) t >T Figure 12: Tip stretch for thin beam.

then, the following relations for angular velocities can be


written: The Newmark direct integration method is employed for
the numerical simulation. The time histories of the tip lateral
ωx = ω cos α, ωy = ω sin α.
deformations for α = 45°, Ω = 3 rad/s and T = 15 s are given
Similar to the previous example, two cases will be considered; in Figures 10–12. In [5], the following results had been given for
first, a thin beam, and second a thick beam. For the thin beam, T = 1.5 s, and in [23] for T = 0.9258 s.
we have considered the problem that has been solved in [5,23]. For a thick beam, the properties of the beam are given as
The properties of the beam are given as follows: mass density, follows: mass density, ρ = 3000 kg/m3 , modulus of elasticity,
ρ = 3000 kg/m3 , modulus of elasticity, E = 70 Gpa, area E = 70 Gpa, area moment of inertia, Iy = 2 × 10−7 m4 , Iz =
moment of inertia, Iy = 2 × 10−7 m4 , Iz = 4 × 10−7 m4 , cross- 4 × 10−7 m4 , cross-section area, A = 1 × 10−4 m2 , and length,
section area, A = 4 × 10−4 m2 , and length, L = 10 m. L/ry = 8, where ry is the gyration radius.
1114 H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116

Figure 13: Tip deflection in v direction for thick beam. Figure 16: Axial stress of thick beam for R∗ = 0.0282.

Figure 17: Bending stress in v direction of thick beam for R∗ = 0.0282.

Figure 14: Tip deflection in w direction for thick beam.


It is obvious that there is a great difference in deflections
between the Timoshenko and Euler–Bernoulli beams for the
thick beam. As we know, the Euler beam is stiffer than the
Timoshenko beam. Similar to those of the planar case, the
stresses can be defined. Figures 16–18 show the axial and
bending stresses for a thick beam at x = 0.
The above results show different solutions for a thick beam
for Euler–Bernoulli and Timoshenko beams. It is interesting that
the stretches for both models are identical and the difference is
only in lateral deflection.
It is expected that there will be some differences in
the natural frequency of the rotating beam, based on the
Euler–Bernoulli or Timoshenko models. From the discretized
equations of motion given by Eqs. (94) and (101), the eigenvalue
problems are derived, in which the natural frequencies can be
computed. Assume the steady state solutions of Eqs. (94) and
(101) as:

Figure 15: Tip stretch for thick beam. d = Xeiωτ (108)



where i = −1, ω and X are the natural frequency and the
The Newmark direct integration method is employed for amplitudes of vibrations, respectively. Neglecting the applied
the numerical simulation. The time histories of the tip lateral forces and the rotating acceleration, substitution of Eq. (108)
deformations for α = 45°, Ω = 300 rad/s and T = 15 s are into Eqs. (94) and (101) leads to the eigenvalue problems
given in Figures 13–15. given by:
H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116 1115

Table 1: First third mode natural frequencies for thin beam.

First mode Second mode Third mode


Euler–Bernoulli Timoshenko Euler–Bernoulli Timoshenko Euler–Bernoulli Timoshenko

Ω=0 3.7977 3.7980 5.3708 5.3716 23.8000 23.8144


Exact 3.7973 3.7975 5.3702 5.3702 23.7990 23.7920
Ω=1 3.5353 3.5357 5.6975 5.6983 23.8052 23.8196
Ω=2 3.0259 3.0264 6.3936 6.3943 23.8226 23.8374
Ω=3 2.4720 2.4725 7.2364 7.2371 23.8574 23.8727

Table 2: First third mode natural frequencies for thick beam.

First mode Second mode Third mode


Euler–Bernoulli Timoshenko Euler–Bernoulli Timoshenko Euler–Bernoulli Timoshenko

Ω=0 1.1867 × 104 1.0471 × 104 1.6783 × 104 1.3438 × 104 7.4375 × 104 4.2427 × 104
Exact 1.1866 × 104 1.0455 × 104 1.6781 × 104 1.3421 × 104 7.4372 × 104 4.2430 × 104
Ω = 50 1.1867 × 104 1.0470 × 104 1.6784 × 104 1.3438 × 104 7.4375 × 104 4.2426 × 104
Ω = 100 1.1866 × 104 1.0469 × 104 1.6784 × 104 1.3440 × 104 7.4375 × 104 4.2425 × 104
Ω = 300 1.1858 × 104 1.0457 × 104 1.6794 × 104 1.3453 × 104 7.4375 × 104 4.2412 × 104

almost 34%. But, in axial stress and deformation the difference


is not considerable. It is found, from natural frequency analyses,
that the errors of the Euler–Bernoulli beam, with respect to
the Timoshenko beam, are almost 13%, 25% and 75%, for first,
second and third natural frequencies, respectively.

Appendix

For obtaining the shape functions of the Timoshenko beam,


consider the static equation of Timoshenko beam theory:
∂ 2v ∂ψ
 
µAG − =0 (A.1)
∂ x2 ∂x
∂ 2ψ ∂ν
 
EIz + µAG − ψ = 0. (A.2)
∂ x2 ∂x
From Eq. (A.1), we can write:
∂ν
 
Figure 18: Bending stress in w direction of thick beam for R∗ = 0.0282. −ψ = C1 . (A.3)
∂x
−ω M + iωC + K X = 0. Then, from Eq. (A.2), we can write:
 2 
(109)
∂ 2ψ
The results of the first three modes of the spatial beam EIz = −µAGC1 .
are presented in Tables 1 and 2 for thin and thick beams, ∂ x2
respectively. For verification of the MATLAB code, the natural Then, we have:
frequencies for a stationary Ω = 0 beam are compared with an ∂ 2ψ µAG
exact solution. =− C1 = −α C1 .
It is obvious that in the thin beam case, the results for
∂ x2 EIz
the Timoshenko model and the Euler–Bernoulli beam are Then:
coincidental, since the beam is slender enough. But, in the thick 1
ψ = − α C 1 x2 + C 2 x + C 3 , (A.4)
beam case, the difference in the results is remarkable. The errors 2
of the Euler–Bernoulli beam, with respect to the Timoshenko 1 1
beam, are almost 13%, 25% and 75%, for first, second and third v = − α C1 x3 + C x2 + (C1 + C3 ) x + C4 . (A.5)
6 2 2
modes, respectively.
Imposing the following end conditions:
6. Conclusion ψ = ψe , v = ve at x = xe ,
ψ = ψe+1 , v = ve+1 at x = xe+1 .
Equations of motion of Timoshenko and Euler–Bernoulli
We will have:
beams on a flying support are derived. The analysis for the
presented examples shows that in flexural deflections, there ψ = NψT d, (A.6)
is a remarkable difference in the solutions of the Timoshenko
Euler–Bernoulli beam theories in the thick beam case. In a
v = NvT d, (A.7)
spatial rotating beam, which has been solved, the maximum where d = {ve , ψe , ve+1 , ψe+1 } , and Nv and Nψ are defined at
T

difference in the steady state solution for flexural deflection is Eqs. (60) and (61).
1116 H. Zohoor, F. Kakavand / Scientia Iranica, Transactions B: Mechanical Engineering 19 (2012) 1105–1116

References [22] Liu, J.Y. and Hong, J.Z. ‘‘Dynamics of three-dimensional beams undergoing
large overall motion’’, European Journal of Mechanics. A. Solids, 23,
[1] Liu, J.Y. and Hong, J.Z. ‘‘Geometric stiffening of flexible link system with pp. 1051–1068 (2004).
large overall motion’’, Computers & Structures, 81, pp. 2829–2841 (2003). [23] Na, K.S. and Kim, J.H. ‘‘Deployment of a multi-link flexible structure’’,
[2] Dwivedy, S.K. and Eberhard, P. ‘‘Dynamic analysis of flexible manipulators, Journal of Sound and Vibration, 294, pp. 298–313 (2006).
a literature review’’, Mechanism and Machine Theory, 41, pp. 749–777 [24] Zohoor, H. and Khorsandijou, S.M. ‘‘Enhanced nonlinear 3D Euler–
(2006). Bernoulli beam with flying support’’, Nonlinear Dynamics, 51, pp. 217–230
[3] Agarwal, O.P. and Shabana, A.A. ‘‘Dynamic analysis of multibody systems (2008).
using component modes’’, Computers & Structures, 21, pp. 1303–1312 [25] Zohoor, H. and Khorsandijou, S.M. ‘‘Generalized nonlinear 3D Euler–
(1985). Bernoulli beam theory’’, Iranian Journal of Science & Technology, Transaction
[4] Kane, T.R., Ryan, R.R. and Banerjee, A.K. ‘‘Dynamics of a cantilever beam B, Engineering, 32(B1), pp. 1–12 (2008).
attached to a moving base’’, Journal of Guidance, Control, and Dynamics, 10, [26] Zohoor, H. and Khorsandijou, S.M. ‘‘Dynamic model of a flying manipulator
pp. 139–151 (1987). with two highly flexible links’’, Applied Mathematical Modelling, 32,
[5] Yoo, H.H., Ryan, R.R. and Scott, R.A. ‘‘Dynamics of flexible beams pp. 2117–2132 (2008).
undergoing overall motions’’, Journal of Sound and Vibration, 181, [27] Han, S.M., Benaroya, H. and Wei, T. ‘‘Dynamics of transversely vibrating
pp. 261–278 (1995). beams using four engineering theories’’, Journal of Sound and Vibration,
[6] Mayo, J. and Dominguez, J.A. ‘‘Finite element geometrically nonlinear dy- 225(5), pp. 935–988 (1999).
namic formulation of flexible multibody systems using a new displace-
ments representation’’, Journal of Vibration and Acoustics, 119, pp. 573–581
(1997). Hassan Zohoor was born in Esfahan, Iran, in 1945. He obtained his Ph.D.
[7] Boutaghou, Z.E., Erdman, A.G. and Stolarski, H.K. ‘‘Dynamics of flexible degree from Purdue University, USA, and is currently Professor of Mechanical
beams and plates in large overall motions’’, Journal of Applied Mechanics, Engineering at Sharif University of Technology, Tehran. He is also Fellow
59, pp. 991–999 (1992). (Academician) and Secretary of the Academy of Sciences of IR Iran (IAS). He is
[8] Chung, J. and Yoo, H.H. ‘‘Dynamic analysis of a rotating cantilever beam by author or co-author of over 350 scientific papers, two chapters of two books
using the finite element method’’, Journal of Sound and Vibration, 249(1), published by UNESCO, two chapters of two other books, and author of four
pp. 147–164 (2002). technical pamphlets. He was also coordinator for compiling one e-book and
[9] Putter, S. and Manor, H. ‘‘A natural frequency of radial rotating beams’’, four CDs for four courses. He has also supervised over 150 graduate theses. He
Journal of Sound and Vibration, 56, pp. 175–185 (1978). has conducted more than twenty research funded projects, including an Iranian
[10] Bauer, H. ‘‘Vibration of a rotating uniform beam’’, Journal of Sound and project in the area of energy, and holds one patent approved by the Office
Vibration, 72, pp. 177–189 (1980). of Patent Management, Purdue Research Foundation, USA. He was Founder,
[11] Hoa, S. ‘‘Vibration of a rotating beam with tip mass’’, Journal of Sound and President and Developer of the principal codes and regulations of the Payame
Vibration, 67, pp. 369–381 (1979). Noor University in Iran. He was also Head of the Department of Engineering
[12] Wright, A., Smith, C., Thresher, R. and Wang, J. ‘‘Vibration modes of Sciences at IAS; Deputy for Infrastructure Affairs at the Budget and Planning
centrifugally stiffened beams’’, Transactions of the ASME. Journal of Applied Organization, Iran; Head of the Institute of Research and Planning in Higher
Mechanics, 49, pp. 197–202 (1982). Education, Iran; Academic Vice-Minister at the Ministry of Science and Higher
[13] Kuo, Y., Wu, T. and Lee, S. ‘‘Bending vibration of a rotating non-uniform Education, Iran; Acting President of Alzahra University, Iran, and President of
beam with tip mass and an elastically restrained root’’, Computers & Shiraz University, Iran. He has received several honor plaques and awards,
Structures, 22, pp. 229–236 (1994). including the Top Student Award from Shiraz University; two Ross Ade Awards
[14] Yokoyama, T. ‘‘Free vibration characteristics of rotating Timoshenko from Purdue University; an Award for Distinguished Professorship, Iran; the
beams’’, International Journal of Mechanical Sciences, 30, pp. 743–755 Lasting Personalities Award, Iran; an Honor Plaque for the most competent
(1988). fellow from the IAS; Finalist in the Best Paper Award Competition from the
[15] Rao, S.S. and Gupta, R.S. ‘‘Finite element vibration analysis of rotating American Society of Mechanical Engineers (ASME), USA; Honor Award from
Timoshenko beams’’, Journal of Sound and Vibration, 242(1), pp. 103–124 the International Council for Open and Distance Education (ICDE) Conference,
(2001). for sustained contributions to Distance Education and Open Learning, India; A
[16] Subrahmanyam, K., Kaza, K., Brown, G. and Lawrance, C. ‘‘Nonlinear Golden Plaque from Payame Noor University for the best contribution to Open
vibration and stability of rotating pre-twisted preconed blades including and Distance Education, Iran; an Honor Plaque for Distinguished Professorship
Coriolis effects’’, Journal of Aircraft, 24, pp. 342–352 (1987). from Sharif University of Technology (on the occasion of its 40th Anniversary);
[17] Yoo, H.H. and Shin, S.H. ‘‘Vibration analysis of rotating cantilever beams’’, and an Honor Plaque for Distinguished Professorship in Mechanical Engineering
Journal of Sound and Vibration, 212, pp. 807–828 (1998). from the Iranian Society of Mechanical Engineers (ISME).
[18] Leissa, A. ‘‘Vibration aspects of rotating turbo machinery blades’’, Applied
Mechanics Reviews, 34, pp. 629–635 (1981).
[19] Frisch, H. ‘‘A vector-dyadic development of the equations of motion for
N-coupled flexible bodies and point masses’’, NASA N D-8047 (1975). Farshad Kakavand received a B.S. degree in Mechanical Engineering from
[20] Ho, J. ‘‘Direct path method for flexible multibody spacecraft dynamics’’, Guilan University, Iran, in 1998, and an M.S. degree in Mechanical Engineering
Journal of Spacecraft and Rockets, 14, pp. 102–110 (1977). from Sharif University of Technology, Tehran, Iran, in 2000, where he is
[21] Bodley, C., Devers, A., Park, A. and Frisch, H. ‘‘A digital computer program currently a Ph.D. degree student under the supervision of Dr. Zohoor. Since
for the dynamic interaction simulation of controls and structure’’, NASA 2005, he has been Lecturer at the Islamic Azad University, Takestan Branch,
P-1219 (1978). Iran.

You might also like