You are on page 1of 10

International Journal of Solids and Structures 48 (2011) 2373–2382

Contents lists available at ScienceDirect

International Journal of Solids and Structures


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

A Timoshenko beam theory with pressure corrections for layered orthotropic beams
Graeme J. Kennedy a,⇑, Jorn S. Hansen a, Joaquim R.R.A. Martins b
a
University of Toronto Institute for Aerospace Studies, 4925 Dufferin Street, Toronto, Canada M3H 5T6
b
Department of Aerospace Engineering, University of Michigan, Ann Arbor, MI 48109, USA

a r t i c l e i n f o a b s t r a c t

Article history: A Timoshenko beam theory for layered orthotropic beams is presented. The theory consists of a novel
Received 1 October 2010 combination of three key components: average displacement and rotation variables that provide the
Received in revised form 30 March 2011 kinematic description of the beam, stress and strain moments used to represent the average stress and
Available online 24 April 2011
strain state in the beam, and the use of exact axially-invariant plane stress solutions to calibrate the rela-
tionships between all these quantities. These axially-invariant solutions, which we call the fundamental
Keywords: states, are also used to determine a shear strain correction factor as well as corrections to account for
Timoshenko beam theory
effects produced by externally-applied loads. The shear strain correction factor and the external load cor-
Shear correction factor
rections are computed for a beam composed of isotropic layers. The proposed theory yields Cowper’s
shear correction for a single isotropic layer, while for multiple layers new expressions for the shear
correction factor are obtained. A body-force correction is shown to account for the difference between
Cowper’s shear correction and the factor originally proposed by Timoshenko. Numerical comparisons
between the theory and finite-elements results show good agreement.
Crown Copyright Ó 2011 Published by Elsevier Ltd. All rights reserved.

1. Introduction displacement representation. These residual displacements ac-


count for the difference between the average shear strain and the
The equations of motion for a deep beam that include the shear strain distribution. Cowper introduced a correction factor
effects of shear deformation and rotary inertia were first to account for this difference and computed its value based on
derived by Timoshenko (1921, 1922). Two essential aspects of the three-dimensional solution of a cantilever beam subjected to
Timoshenko’s beam theory are the treatment of shear deformation a tip load.
by the introduction of a mid-plane rotation variable, and the use of Stephen and Levinson (1979) developed a beam theory along
a shear correction factor. The definition and value of the shear the lines of Cowper’s, but recognized that the variation in shear
correction factor have been the subject of numerous research along the length of the beam would lead to a modification of the
papers, some of which are discussed below. Shames and Dym relationship between bending moment and rotation. This variation
(1985, Ch. 4, p. 197) provide an excellent overview of the classical had been neglected by Cowper.
approach to Timoshenko beam theory. This paper however, draws Following the work of Cowper (1966) and Stephen and Levinson
primarily from research and theories which refine Timoshenko’s (1979), in this paper we seek a solution to a beam problem based
original approximations. on average through-thickness displacement and rotation variables.
Prescott (1942) derived the equations of vibration for thin rods In a departure from previous work, we introduce strain moments,
using average through-thickness displacement and average rota- which are analogous to the stress moments used in the equilibrium
tion variables. He introduced a shear correction factor to account equations. These strain moments remove the restriction of working
for the difference between the average shear on a cross section with an isotropic, homogeneous beam. This is an essential compo-
and the expected quadratic distribution of shear. nent of the present approach, as sandwich and layered orthotropic
Cowper (1966) presented a revised derivation of Timoshenko’s beams are often used for high-performance, aerospace applications
beam theory starting from the equations of linear elasticity for a (Flower and Soutis, 2003).
prismatic, isotropic beam in static equilibrium. Cowper introduced Another important feature of our theory is the use of certain
residual displacement terms that he defined as the difference statically determinate beam problems that we use to construct
between the actual displacement in the beam and the average the relationship between stress and strain moments, and to recon-
struct the stress and strain solution in a post-processing step. We
⇑ Corresponding author. call these solutions the fundamental states of the beam. The present
E-mail addresses: graeme.kennedy@utoronto.ca (G.J. Kennedy), hansen@utias.u- theory was first pursued by Hansen and Almeida (2001) and
toronto.ca (J.S. Hansen), jrram@umich.edu (J.R.R.A. Martins). Hansen et al. (2005), and an extension of this theory to the analysis

0020-7683/$ - see front matter Crown Copyright Ó 2011 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijsolstr.2011.04.009
2374 G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382

of plates was presented by Guiamatsia and Hansen (2004), beam. For a beam with a rectangular cross section, Hutchinson ob-
Tafeuvoukeng (2007) and Guiamatsia (2010). tained a shear correction factor that depends on the Poisson ratio
This paper begins with a brief discussion of two classical meth- and the width to depth ratio. In a later discussion of the paper, Ste-
ods used to calculate the shear correction factor in Section 2. phen (2001) showed that the shear correction factors he obtained
Section 3 describes the proposed theory and Section 3.2 introduces in Stephen (1980) were equivalent.
the fundamental states. In Section 4, calculations are presented for More recently Dong et al. (2010), presented a semi-analytic fi-
a beam composed of multiple isotropic layers. Section 5 briefly pre- nite-element technique for calculating the shear correction factor
sents the modified equations of motion for an isotropic beam. In based either on the Saint–Venant warping function or the free
Section 6, comparisons are made with finite-element calculations. vibration of a beam.
Section 7 outlines conclusions based on the theory presented Some experimental studies have been performed to try and
herein. measure the shear correction factor based on the original equa-
tions proposed by Timoshenko. Spence and Seldin (1970) obtained
2. The shear correction factor experimental values of the shear correction factor for a series of
square and circular beams composed of both isotropic and aniso-
One of the main difficulties in using Timoshenko beam theory is tropic materials by determining their natural frequencies. Kaneko
the proper selection of the shear correction factor. Many authors (1975) performed an extensive review of the shear correction fac-
have published definitions of the shear correction factor and have tors for rectangular and circular cross sections obtained by various
proposed various methods to calculate it. Most of these approaches authors using either experimental techniques or analysis. Experi-
fall into one of two categories. The first approach is to use the shear mental studies have generally used a natural frequency approach
correction factor to match the frequencies of vibration of various to determine the shear correction factor and have generally found
beam constructions with exact solutions to the theory of elasticity. that Timoshenko’s value is superior to Cowper’s. This is perhaps
The second approach is to use the shear correction factor to ac- not surprising, since Timoshenko’s correction was obtained by
count for the difference between the average shear or shear strain matching frequencies in the same manner in which the experi-
and the actual shear or shear strain using exact solutions to the ments are performed. However, the frequency matching approach
theory of elasticity. fails to provide a theoretical explanation as to why the value of a
Timoshenko (1922) developed the frequency-matching ap- factor that modifies the relationship between the shear resultant
proach. He calculated the shear correction factor by equating the and the average shear strain should be determined by the natural
frequency of vibration determined using the plane stress equations frequency of vibration. It is this deficiency that motivates the work
of elasticity to those computed using his beam theory. Although presented here.
not explicitly written in the paper, the shear correction factor ob-
tained in this manner for a rectangular beam is
3. The theory
5ð1 þ mÞ
kxy ¼ : ð1Þ
6 þ 5m The geometry of the beam under consideration is shown in
Fig. 1. The beam extends along the x-direction subject to forces
Cowper (1966) calculated the shear correction factor using an
on the top and bottom surfaces in the y-direction. The reference
approach from the second category described above. Using residual
axis is placed at the centroid of the cross-section. The half-thick-
displacements designed to take into account the distortion of the
ness in the y-direction is c, while the length of the beam in the
cross sections under shear loads, Cowper was able to derive a for-
x-direction is L. The beam is of uniform composition in both the
mula for the shear correction factor based on solutions of a canti-
x and z-directions and so consists of a series of layers with different
lever beam subjected to a tip load. For a rectangular isotropic
material properties. We assume that each layer is composed of an
homogeneous beam, Cowper found the following shear correction
orthotropic material, with material properties aligned with the
factor:
coordinate axes. These assumptions eliminate the possibility of
10ð1 þ mÞ twisting and allow the beam to be modeled using a plane stress
kxy ¼ : ð2Þ
12 þ 11m assumption in the z plane. In each layer k, numbered from the bot-
Following Cowper’s approach, Stephen (1980) computed the tom to the top of the beam, the following constitutive law holds:
shear correction factor for beams of various cross sections by using
rðkÞ ¼ C ðkÞ ðkÞ ;
the exact solutions for a beam subject to a uniform gravity load. He
employed a modified form of the Kennard–Leibowitz method where rðkÞ ¼ ½ rx ry rxy TðkÞ and ðkÞ ¼ ½ x y cxy TðkÞ . Since the
(Leibowitz and Kennard, 1961), to obtain the shear correction fac- beam is composed of an orthotropic material, there is no coupling
tor by equating the average centerline curvature of the exact result between shear and normal stresses. Although variation in the Pois-
with the Timoshenko solution. He obtained a modified form of son’s ratio between layers would lead to a violation of the plane
Timoshenko’s shear correction factor for rectangular sections that stress assumption, we include this possibility and ignore the edge
approached Eq. (1) for thin cross-sections. effects in the cross-section in such situations.
Using the frequency matching approach, Hutchinson (1981) These assumptions are an extension of the conditions originally
computed the shear correction factor by performing a comparison used by Timoshenko, who limited his analysis to plane stress
between Timoshenko beam theory and three solutions from the beams composed of a single isotropic material (Timoshenko, 1922).
theory of elasticity, the Pochhammer–Chree solution in Love
(1920), a Fourier solution due to Pickett (1944) and a series solu-
tion computed by Hutchinson (1980). Hutchinson found that the y y
best shear correction factor was dependent on the frequency and
z x
Poisson’s ratio of the beam, but that Timoshenko’s value was better 2c
than Cowper’s.
Later, Hutchinson (2001) introduced a new Timoshenko beam 1 L
formulation and computed the shear correction factor for various
cross sections based on a comparison with a tip-loaded cantilever Fig. 1. The geometry of the beam composed of layers of different materials.
G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382 2375

3.1. The displacement representation ments can be used to capture an arbitrary displacement field. Next,
we examine the state of stress within the beam.
Following Prescott (1942) and Cowper (1966), the average
through-thickness displacements and average rotation are defined 3.2. The fundamental states
as follows:
The basic assumption made in the development of the present
Z c
1 beam theory is that the stress and strain state in the beam can
u0 ðx; tÞ ¼ uðx; y; tÞdy;
2c c be approximated using a linear combination of axially-invariant
Z c solutions. We call these axially-invariant solutions the fundamental
3
u1 ðx; tÞ ¼ 3 yuðx; y; tÞdy; ð3Þ states. The fundamental states can be used to capture the complex
2c c
Z c interaction between the stresses in layered orthotropic beams,
1
v 0 ðx; tÞ ¼ v ðx; y; tÞdy; away from the ends of the beam. As is the case in many beam the-
2c c
ories, end effects cannot be captured using this approach. In this
section we address how to determine the fundamental states.
where u and v are the displacements in the x and y directions, The fundamental states are determined from a hierarchy of stat-
respectively. The average displacements and rotation are defined ically determinate beam problems. These beam problems are for-
regardless of the through-thickness behavior of u and v, which are mulated using a series of self-equilibrating loads applied to a
piecewise continuous through the thickness of the beam in this beam with the same sectional properties as the beam under con-
problem. The average displacements are an incomplete representa- sideration. Rigid body translation and rotation modes are removed
tion of the total displacement field in the beam, in the sense that the from the solution by imposing three displacement constraints so
average quantities do not capture the point-wise behavior of the ex- that no stress concentrations are present. The first four loading
act displacement. In order to capture this behavior, it is necessary to conditions leading to the first four fundamental states are shown
introduce residual displacements that account for the difference be- in Fig. 2. N, M and Q are the axial, bending, and shear resultants, de-
tween the average and point-wise quantities in the following fined as follows:
manner: Z c
Nðx; tÞ ¼ rx ðx; y; tÞdy;
uðx; y; tÞ ¼ u0 ðx; tÞ þ yu1 ðx; tÞ þ u ~ ðx; y; tÞ; c
ð4Þ Z c
v ðx; y; tÞ ¼ v 0 ðx; tÞ þ v~ ðx; y; tÞ; Mðx; tÞ ¼ yrx ðx; y; tÞdy; ð6Þ
Z cc
where u ~ and v~ are the residual displacements in the x and y direc-
tions, as introduced by Cowper (1966). Given the definitions of the
Q ðx; tÞ ¼ rxy ðx; y; tÞdy:
c
average displacements (3), the zeroth and first moments of u ~ , and
the zeroth moment of v~ through the thickness, must be zero: We also refer to these as the stress moments.
The beam in Fig. 2 has the same cross-sectional properties as
Z c
the beam under consideration, but is extended between the coor-
~ ðx; y; tÞdy ¼ 0;
u
dinates x = Lf to x = Lf. Lf is the half-length of the beam for the fun-
Zcc damental state analysis, and must be sufficiently large such that
~ ðx; y; tÞdy ¼ 0;
yu the end effects do not influence the state of stress or strain at the
c
Z c middle of the beam.
v~ ðx; y; tÞdy ¼ 0: The fundamental states are obtained from the solution of the
c
beam problems illustrated in Fig. 2 by taking the through-thick-
The average displacements and displacement residuals may be ness stress and strain distribution at x = 0. As a result, the funda-
used to determine the strain at any point in the beam. In this ap- mental states represent a distribution of stress and strain only in
proach however, we are interested in the average through-thick- the y direction. The loading conditions are constructed such that
ness strain. To this end, we introduce the following strain
moments:
y
Z c
@u @u0
0 ðx; tÞ ¼ dy ¼ 2c ; ð5aÞ Fundamental states 2c x
c @x @x
Z c 2L f
@u 2c3 @u1
jðx; tÞ ¼ y dy ¼ ; ð5bÞ
c @x 3 @x First N =1 N =1
Z c    Z c
@u @ v @v 0 ~
@u
cðx; tÞ ¼ þ dy ¼ 2c u1 þ þ dy; ð5cÞ
c @y @x @x c @y Second M = 1 M =1

that represent the axial, bending and shear strain moments, respec- Q=1
tively. These strain moments are analogous to the stress moments Third QL f QL f
that are used to define the equilibrium equations for a beam. Note Q=1
that these strain moments are not normalized and as a result have P =1
different dimensions than the point-wise strain. The main advan-
tage of using the strain moments (5) over point-wise strain vari- P L 2f P L 2f
Fourth
2 2
ables is that they are always defined, regardless of the through- P Lf P Lf
thickness distribution of the strain. This is an important property,
since the point-wise shear strain can be discontinuous at material Fig. 2. An illustration of the loading conditions used to obtain the first four
fundamental states. The states are: axial loading, bending moment, shear and
interfaces. pressure load. The fundamental states are extracted from the solution at the x = 0
Thus far, no assumptions beyond those of linear elasticity have plane. Lf, the half-length of the beam used to calculate the fundamental states, must
been made. The combination of the average and residual displace- be large enough that the end effects do not influence stress distribution at x = 0.
2376 G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382

only one stress resultant or load is non-zero at x = 0. For instance, states. On the other hand, the strain moments may have contribu-
in the third fundamental state, which corresponds to a shear load, tions from all fundamental states and the strain residuals. Using
the bending resultant is zero at the mid-section and the shear Eq. (7b), the required moments of strain are,
resultant is unity, while in the fourth fundamental state, which 2 3 2 32 3 2 3 2 3
corresponds to a pressure load, the shear resultant and bending
0 N0 M0 Q0 N P0 ~0
6 7 6 N Q 76 M 7 þ P 6 jP 7 þ 6 j 7
moment are zero and the pressure load is unity. 4 j 5 ¼ 4j M
j j 54 5 4 5 4 ~ 5; ð8Þ
We label the fundamental states with a superscript for the cor- c cN cM cQ Q cP c~
responding condition: N, M and Q for the axial resultant, bending
moment, and shear resultant, and P for any externally applied load. where  ~0 ; j
~ and c ~ are the moments of the strain residuals. Note that
For instance, rM(y) and M(y) is the fundamental state correspond- the left-hand side of Eq. (8) is equal to the moments from Eq. (5).
ing to bending with strain moments M M
and cM. Note that the Recall also that the moments of the fundamental states are
0 , j
strain moments of the fundamental states are scalar values inde- constant.
pendent of any coordinate. It is important to distinguish between the three different terms
In the present work, we obtain the fundamental states through in the expression for the strain moments (8). The first term is due
a series of analytic calculations presented below. In these calcula- to the stress resultants, the second term is due to the applied loads,
tions the beam used to calculate the fundamental states is essen- and the remaining term is due to the strain residuals. The final
tially of infinite length since the stress resultants satisfy the term is neglected based on the assumption that its contribution
loading conditions illustrated in Fig. 2 in an average sense for will be small.
any length Lf. The fundamental states could also be obtained Setting  ~0 ; j
~ and c
~ to zero, and re-arranging Eq. (8) results in the
approximately using a finite-element approach. following constitutive relation:
The stress and strain state in the beam can be written as a linear 2 82 3
3 2 P 39
combination of the fundamental states and stress and strain resid- N < 0
> 0 >=
6 7 6 7 6 7
uals, r~ and ~ , respectively. Using this linear superposition, the 4 M 5 ¼ D 4 j 5  P4 jP 5 ; ð9Þ
>
: >
;
stress and strain state in the beam is given by: Q c cP
rðx; y; tÞ ¼ N rN þ MrM þ Q rQ þ PrP þ r~ ðx; y; tÞ; ð7aÞ
where the components of the constitutive matrix D can be found as
ðx; y; tÞ ¼ NN þ MM þ Q Q þ PP þ ~ ðx; y; tÞ; ð7bÞ follows:

where the magnitudes of the fundamental states—the axial, bend-


2 3 2 31
D11 D12 D13 N0 M0 Q0
ing, and shear resultants and the pressure load—are functions of x 6 7 6 7 ð10Þ
D ¼ 4 D21 D22 D23 5 ¼ 4 jN jM jQ 5 :
and t while the fundamental states are functions only of the
through-thickness coordinate y. The stress and strain residuals r ~ D31 D32 D33 cN M
c c Q

and ~ represent deviations due to end effects and higher-order fun-


Note that this matrix is not necessarily symmetric. Due to the
damental states. For instance, a linear or quadratic pressure load orthotropic construction of the beam, cN ; cM ; Q0 , and jQ are zero.
would induce stresses and strains not captured by the first four As a result D13, D23, D31, and D32 are also zero.
states discussed here. It is important to recognize that as a result If only axial, bending and shear loads are applied to the beam,
of Eq. (6), the stress residuals r
~ do not contribute to the axial, bend-
then there is no load-dependent strain moment contribution. How-
ing, or shear resultants. ever, when external loads are applied to the beam, the relationship
The assumption that the stress and strain state in the beam can between the strain moments and stress moments is modified as
be approximated by a linear combination of the fundamental follows:
states is equivalent to assuming that the terms r ~ and  ~ may be
2 3 2 3 2 P3
omitted in the analysis. As a result, end effects are not captured N 0 N
within the theory. Furthermore, rapidly varying loads produce sim- 6 7 6 7 6 7
4 M 5 ¼ D4 j 5  P4 MP 5; ð11Þ
ilar terms from linear, quadratic, and higher-order polynomial Q c QP
loading fundamental states. If these higher-order fundamental
states are not included in the analysis, they will essentially pro- where NP, MP and QP are the product of the strain moments, P0 ; jP
duce additional  ~ terms. and cP, and the constitutive matrix D. NP, MP and QP represent a
The fundamental states also provide a self-consistent method load-dependent pressure correction to the constitutive equations.
for reconstructing the stress and strain distribution within the Note that the constitutive matrix D is derived using the strain
beam in a post-processing step using Eq. (7). This reconstruction moments from the first three fundamental states. The only
includes stress and strain components that are not normally con- assumption used to derive this relationship is that the moments
sidered in classical approaches without recourse to a post-analysis of the strain residuals are small. The influence of externally applied
integration of the equilibrium equations through the thickness. loads can be accounted for by including strain moment terms from
However, as is well known, this integration procedure can intro- the fundamental state corresponding to pressure loading. Higher-
duce compatibility problems, whereas Eq. (7) does not suffer from order loading effects could be included by taking into account
this issue. the strain moments due to linear, quadratic, and polynomial pres-
sure distributions in general. Neglecting these effects is equivalent
3.3. The constitutive relation and pressure correction to introducing a non-zero strain residual moment.

We now develop a constitutive relationship between moments 3.4. The shear strain correction
of stress and moments of strain. A pressure correction is also intro-
duced to account for the influence of externally applied loads. To The additional integral in the expression for the shear strain
develop these relationships it is necessary to examine the stress moment in Eq. (5c), involves a correction from the residual dis-
and strain moments in the context of the stress and strain decom- placements. The value of this integral depends on the distribution
position in Eq. (7). By construction, the stress resultants found in of the shear strain through the thickness. Several authors have sug-
Eq. (6) are always equal to the magnitudes of the fundamental gested that this shear strain correction should be computed under
G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382 2377

different loading conditions. For example, Cowper (1966) com- ness stress and strain distributions. Finally, the theory contains a
putes his value of the shear correction factor for a beam subject consistent method for predicting the shear strain correction (12),
to a constant shear load, while Stephen (1980) and Hutchinson the pressure correction (11), and the stiffness (10) and using the
(2001) compute the correction for a beam subject to a gravity load. fundamental states. These additions enhance the capabilities of
In a similar approach to Cowper, we set the shear strain correc- classical Timoshenko beam theory.
tion equal to the ratio of the shear strain moment over the average
shear strain computed using the fundamental state corresponding 4. Isotropic layered beam
to shear:
Rc  In this section we derive the fundamental states, the stress–
~
@u 
dy
cQ c @y
Q strain moment constitutive equation, the shear correction factor,
kxy ¼  v0 ¼ 1 þ  v0 : ð12Þ
2c u1 þ @@x 2c u1 þ @@x and the pressure strain moment corrections for a beam composed
Q Q
of K isotropic layers. Each layer has Young’s modulus Ek, Poisson’s
The subscript Q is used to denote that the expression is evaluated ratio mk, and is situated between y = hk and y = hk+1, where hk is de-
using the fundamental state corresponding to shear. fined relative to the centroid of the cross section. It is often conve-
The corrected shear strain moment is therefore: nient to use the ratio of the Young’s moduli ak, defined such that
  Ek = Eak, where E may be chosen as the Young’s modulus in any
@v 0 convenient layer. Furthermore, we use the non-dimensional ratio
c ¼ 2ckxy u1 þ :
@x of the stations, nk = hk/c. For convenience in presenting various for-
n n
mula, we define Dnk ¼ hkþ1  hk and dnk ¼ nnkþ1  nnk . The weighted
It is important to realize that this is not a correction for the shear area A, the weighted second moment of area I, and a stretching-
stiffness of the beam, but rather a correction of the discrepancy be- bending parameter tb, are defined as follows:
tween the average shear strain and the displacement representa-
X
K XK
ai 3 1X K
ai 2
tion. It is therefore more correct to refer to it as a shear strain A ai Dk ; I  D ; tb  D :
correction. i¼1 i¼1
3 k A i¼1 2 k

In the following formula, a subscript k is used to represent the


3.5. Equilibrium equations
stress or strain distribution in the kth layer, lying between
hk 6 y 6 hk+1.
The equilibrium equations for the stress resultants are obtained
by the standard approach of integrating the two-dimensional
4.1. Axial and bending states
momentum equations. When the density of the material q is con-
stant, these equations are:
The first fundamental state solution corresponds to a beam sub-
ject to a unit axial load that results in the following stress:
@N @ 2 u0
¼ 2cq 2 ; ð13aÞ I 1
@x @t rNx ðkÞ ¼ ak ð1  ryÞ;
@M 2c3 @ 2 u1 A I  At2b
Q ¼ q 2 ; ð13bÞ
@x 3 @t where r = tbA/I. The strain moments in this fundamental state are:
@Q @2v 0
þ P ¼ 2cq 2 : ð13cÞ 2cI 1 2c3 t b 1
@x @t N0 ¼ ; jN ¼  ;
A EðI  At2b Þ 3 EðI  At2b Þ
If the density of the material varies in the through-thickness direc-
and cN = 0.
tion, these equations would involve integrals of the residual
The second fundamental state solution corresponds to a unit
displacements.
bending moment that results in the following stress:

3.6. Discussion 1
rMx ðkÞ ¼ ak ðy  tb Þ:
I  At2b
Our proposed theory fits almost entirely within Timoshenko’s
The strain moments in this fundamental state are:
original beam theory (Timoshenko, 1921, 1922). While the dis-
placement variables involved have a different interpretation, the 1 2c3 1
equations themselves take essentially the same form, except for M0 ¼ 2ctb ; jM ¼ ;
EðI  At 2b Þ 3 EðI  At2b Þ
the pressure correction. The pressure correction can be treated as
M
an additional force arising from the application of a pressure load. and c = 0. Using Eq. (10), the relationship between the strain mo-
As a result, beyond the calculation of the fundamental states, the ments and the stress moments, can be determined as follows:
theory does not require much more computational effort than clas-    
D11 D12 2cI=A 2ctb 1
sical Timoshenko beam theory. In addition, the proposed theory ¼ EðI  At 2b Þ
can handle any combination of boundary conditions typically im- D21 D22 2c3 t b =3 2c3 =3
" #
posed for classical Timoshenko beam problems. Within the context 3EA 2c3 =3 2ctb
of our theory, different boundary conditions result in additional ¼ 4 :
4c 2c3 t b =3 2cI=A
strain residual moment terms in Eq. (8).
Not only does our proposed theory take a similar form to Tim- This equation defines the constitutive relationship for the first two
oshenko’s beam theory, but the additional modifications proposed fundamental states.
above have several important benefits. As with Cowper’s theory,
the proposed approach has a completely general displacement rep- 4.2. Shear state and shear strain correction
resentation (4). We have introduced a stress and strain decompo-
sition (7), based on the fundamental states, that also provides a The third fundamental state corresponds to a constant unit
self-consistent method for the reconstruction of the through-thick- shear load. The stresses in the beam corresponding to this case are:
2378 G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382

2 3 2 3
rx 2ak xðy  tb Þ 4.3. Pressure state and pressure strain correction
1
rðkÞ ¼ 6 7
4 ry 5 ¼
6
4 0
7
5; ð14Þ
2ðI  At2b Þ The fourth fundamental state corresponds to a pressure load ap-
rxy ðkÞ 2
ak ðck þ 2tb y  y Þ
plied to the beam. The total force in the y-direction per unit length
where the ck terms are determined to ensure a continuous variation of the beam is distributed between a traction on the top surface Pt,
of the shear stress through the thickness. The ck coefficient in the and a traction on the bottom surface Pb. Both tractions act in the
2
first layer is c1 ¼ h1  2t b h1 , and can be obtained for subsequent lay- positive y direction. The total force is such that the contributions
ers using the following formula: sum to unity Pt + Pb = 1.
2
The pressure load causes a linearly varying shear and quadrat-
ck ¼ ððak1  ak Þð2t b hk  hk Þ þ ak1 ck1 Þ=ak : ically varying moment in the beam, resulting in the following state
The fundamental state consists of only the stresses correspond- of stress:
ing to the axial-invariant components of the solution. These are ob-
2 3
rx
tained from Eq. (14) by setting rQ(y) = r(x = 0,y). 6
rðkÞ ¼4 ry 7
5
The shear strain moment is determined by integrating the shear
strain through the thickness: rxy ðkÞ
2 3
  ak ðx2 y  t b x2  2y3 =3 þ 2t b y2 þ ek y þ fk Þ
XK
ð1 þ mk Þ 2 1 3 1 6 7
cQ ¼ c D
k k þ t D
b k  D : ¼ 4 ak ðdk þ ck y þ tb y2  y3 =3Þ 5:
2
k¼1 EðI  At b Þ
3 k 2ðI  At2 Þ
ak xðck þ 2tb y  y2 Þ
The relationship between the shear stress resultant and the shear ð17Þ
strain resultant is, Q = D33c, where
The fundamental state is determined by taking only the axially-
D33 ¼ 1=cQ : ð15Þ invariant components of the stress state given in Eq. (17):
This is not a simple average of the shear-modulus through the rP(y) = r(x = 0, y).
thickness, which is often used in beam theories. Eq. (15) is a The coefficients dk are determined from the inter-layer continu-
weighted average dependent on the relative distribution of shear ity of ry, while the coefficients ek and fk are used to satisfy two
Rc Rc
through the thickness. equilibrium equations: c yrx dy ¼ x2 =2 and c rx dy ¼ 0, as well
The shear correction factor for the multi-layer beam kxy, is as K  2 inter-layer displacement continuity constraints. The dk
determined from Eq. (12). It is a dimensionless quantity that de- coefficients can be determined using the following relationship,
2 3
pends only on the Poisson ratio, the relative position of the layers, d1 ¼ 2ðI  At 2b ÞPb =a1  ðc1 h1 þ th1  h1 =3Þ for the first layer, and
and the relative magnitudes of the stiffnesses of each layer. As in all subsequent layers using
h i
2 3
such, it is expressed using dimensionless quantities. dk ¼ 1=ak ðak1  ak Þðtb hk  hk =3Þ þ hk ðak1 ck1  ak ck Þ þ ak1 dk1 :
The dimensionless bending-stretching coupling constant s is gi-
ven by The additional equations for the inter-layer continuity of the
, displacements are
1X K

X K  
s¼ ak n2kþ1  n2k ak ðnkþ1  nk Þ: 2 3
ðek  ek1 Þhk  ðfk  fk1 Þ ¼ ðmk1  mk Þ tb hk  hk =3
2 k¼1 k¼1
 mk ðdk þ ck hk Þ þ mk1 ðdk1 þ ck1 hk Þ;
We next introduce the constants Ck, Bk, and Ak, which are defined
sequentially for each layer. For k = 1, C 1 ¼ n21  2sn1 ; B1 ¼ ek  ek1 ¼ ck ð2 þ mk Þ  ck1 ð2 þ mk1 Þ þ ðmk1  mk Þðhk  2tb hk Þ;
2

2ð1 þ m1 ÞC 1 and A1 = 0. For each subsequent layer,


for k = 2, . . . , K. The two additional equilibrium equations are


C k ¼ ðak1  ak Þð2snk  n2k Þ þ ak1 C k1 ak ; X
K X K
ek 3 fk 2 2 5 tb 4
Bk ¼ ðmk1  mk Þðn2k  2snk Þ þ Bk1 ; ak Dk þ Dk ¼ ak Dk  Dk ;
i¼1
3 2 i¼1
15 2
 
1
Ak ¼ 2nk ð1 þ mk1 ÞC k1  ð1 þ mk ÞC k þ ðBk1  Bk Þ
2 X
K ne o X K
k 2 1 2t
þ ðmk1  mk Þðsn2k  n3k =3Þ þ Ak1 : ak Dk þ fk Dk ¼ ak D4k  b D3k :
i¼1
2 i¼1
6 3
The shear correction factor for the layered, isotropic beam is
Using the values obtained by solving these for ek and fk with the
kxy ¼ D=F; ð16Þ above 2K equations, the strain moments for this fundamental state
can be written as
(
where K 
P 1 4 3 X ek 2
0 ¼ tb c þ D þ f k Dk
XK
1
2
2EðI  At b Þ 3 k¼1
2 k
D¼ ð1 þ mk Þ C k dk þ sd2k  d3k ;  
3 ck tb 1 4
k¼1 þ mk dk Dk þ D2k þ D3k  Dk ; ð18aÞ
2 3 12
XK
1 3 1
(
F¼ dk ð2ð1 þ mk ÞC k þ Bk Þ þ ð2 þ mk Þ 15sd4k  4d5k K 
2 40 P 1 4 5 X ek 3 fk 2
k¼1
  j ¼ cþ D þ D
2
2EðI  Atb Þ 15 3 k 2 k
3 1 mk 1 k¼1
þ Ak d2k  Bk dk þ sd2k  d3k :  
4 2 2 3 dk 2 ck 3 t b 4 1
þ mk Dk þ Dk þ Dk  D5k : ð18bÞ
2 3 4 15
For a single-layer beam, this expression simplifies to Cowper’s shear
correction factor (2). The shear strain moment for this fundamental state is zero, cP = 0.
G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382 2379

5. Equations of motion for an isotropic beam 6. Results

Before examining several static cases using the shear and pres- In this section, we examine the shear strain and pressure cor-
sure corrections derived above, we will briefly examine the natural rections, and the constitutive relationship obtained above, for
frequency of vibration of an isotropic beam with a body-force cor- two cases: a three-layer symmetric beam, and a multi-layer beam
rection. For this isotropic case, I = 2c3/3, A = 2c and tb = 0. composed of alternating materials. Results from a finite-element
Under a constant body load with a value of 1/2c, the stress state analysis are used to compare with the formulas derived above.
in an isotropic beam is, The first beam considered is composed of three layers, where
the middle layer is made of a material that has a lower Young’s
2 3 2 3
rx x2 y þ 2y3 =3  2c2 y=5 modulus than the outer layers. This problem is designed to model
6 7 16 7 a sandwich structure in which the inner core material is less stiff
r ¼ 4 ry 5 ¼ 4 ðc2 y  y3 Þ=3 5: ð19Þ
2I than the outer material. The outer two layers have Young’s modu-
rxy 2
xy  xc 2
lus E and Poisson’s ratio m, while the inner core has Young’s mod-
ulus aE and Poisson ratio m. The depth of the beam is 2c and the
The stresses have a linear varying shear and a quadratically varying inner core extends from y = rc to y = rc, where r is the fraction
bending moment, as in the pressure state described above. From of the beam that is composed of the core material. For this beam,
Eq. (19), the fundamental state corresponding to a body load is simplifications from the general formulas above are possible. The
rB(y) = r(x = 0, y). The strain moments corresponding to this funda- average shear stiffness (15) simplifies to
mental state are B0 ¼ 0, cB = 0, and
3EI
  D33 ¼ ; ð25Þ
1 m 2c 5
2c 5
mc 2 2ð1 þ mÞð2c3  3c3 rð1  sÞÞ
jB ¼   ¼ : ð20Þ
2EI 3 3 5 15E
where s = (1  (1  a)r2)/a and the shear correction factor (16)
B 2 becomes
The bending moment correction is M = mc /15. Under conditions
of free-vibration, the magnitude of this body-force fundamental
ð1 þ mÞð30rðs  1Þ þ 20Þ
state is equal to the inertial force per unit span. As a result, Eq. kxy ¼ : ð26Þ
30ð1 þ mÞs  ð6 þ 8mÞ þ 15ð1 þ mÞð1  sÞð2 þ r 3  3rÞ
(11) becomes
As the ratio of the Young’s modulus of the core decreases, it is
@u1 @2v 0 interesting to note that a limiting case is reached that is indepen-
M ¼ EI þ qAM B : ð21Þ dent of the Poisson’s ratio. This limit as a ? 0 is,
@x @t 2
2
kxy ¼ : ð27Þ
Using this relationship, the equation of motion for a freely vibrating 3  r2
beam is The second beam considered is composed of alternating
" # isotropic layers that have relative Young’s modulus E1/E2 = 10 and
4 2 B 4 4
@ v0 @ v0 E AM @ v0 q I @ v0
2
Poisson’s ratios of m1 = 0.2 and m2 = 0.4. For this case, we vary the
EI þ qA 2  qI 1 þ þ þ ¼ 0:
@x4 @t kxy G I @t2 @x2 kxy G @t4 number of layers, keeping the depth of the beam constant, c = 1/2
ð22Þ while altering the thickness of the layers to match. As a result
hk = c + 2c(k  1)/K. The plies are composed of alternating material
The classical equation of motion may be obtained by setting starting from the bottom layer. The beam is symmetric for odd K.
MB = 0. The equation of motion for an isotropic beam, using the For finite-element calculations, we use bi-cubic Lagrange plane
body-force correction (20) and Cowper’s shear correction factor stress elements with a standard formulation. We choose these
(2), is high-order elements because they capture the piecewise parabolic
 2 shear stress accurately through the thickness of the beam.
@4v 0 @ 2 v 0 17 þ 10m @4v 0 12 þ 11m qI @ 4 v 0 The finite-element model is constructed with L/2c = 10 with 50
EI þ qA  q I þ ¼ 0:
@x4 @t 2 5 @t2 @x2 5 A @t 4 elements along the length of the beam. For the three-layer beam,
ð23Þ we take 20 elements through the thickness resulting in 18,422 de-
grees of freedom. For the multi-layer beam the number of through-
While for the classical equation, with Timoshenko’s shear correc- thickness elements varies so that the number of elements in each
tion factor (1), the equation of motion is layer is the same, while the total number of elements through
the thickness does not fall below 20. The number of elements
 2 through the thickness is Kd20/K e.
@4v 0 @ 2 v 0 17 þ 10m @4v 0 12 þ 10m qI @ 4 v 0 In order to compare the value of the shear correction factor de-
EI þ qA  q I þ ¼ 0:
@x4 @t 2 5 @t2 @x2 5 A @t 4 rived above with finite-element results, we use results from a
ð24Þ beam subject to a shear load at the tip, with the root fully fixed.
An approximate shear correction factor is computed from the fi-
Eqs. (23) and (24) differ only in the coefficient of the fourth
nite-element solution based on Eq. (12). This approximate shear
term by 1/5m(qI/A)2. The relative difference between these terms FE
correction factor, kxy , is computed as follows:
is 2% for m = 0.3. This suggests that for vibration problems, using
Rc
the proposed theory with Cowper’s shear correction factor and a FE cxy dy
body-force correction is essentially equivalent to using Timo- kxy ðxÞ ¼ c v0 ; ð28Þ
2c u1 þ @@x
shenko’s shear correction factor and the equations of motion he
originally derived. This agreement should be expected, as experi- where numerical integration is used to evaluate u1 and v0 from the
ments based on the natural frequencies of vibration have typically finite-element results based on Eq. (3), and the derivative is per-
FE
demonstrated that Timoshenko’s shear correction factor is superior formed using a central-difference calculation with Dx = 105. kxy is
(Spence and Seldin, 1970; Kaneko, 1975). calculated at every Gauss point along the x-direction.
2380 G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382

For comparison with the pressure corrections, we calculate a shear stiffness computed using Eq. (15) for the multi-layer beam.
solution of a cantilevered beam subject to a pressure load distrib- Finite-element calculations were performed for the first 10 beams
uted on the top and bottom with Pb = 1/5 and Pt = 4/5. The pressure with K = 1, . . . , 10, while the analytic formulas are used up to K = 50
correction is evaluated using a combination of finite-element and to show the trend. As previously mentioned, for odd K the beams
beam theory values where the total strain and stress moments are symmetric, and for even K the beams exhibit bending-stretch-
are computed from the finite-element method, while the constitu- ing coupling. As K becomes larger, the coefficients tend towards a
tive relation is used from Eq. (9). This gives the following equation limiting case. Excellent agreement is obtained. The average relative
for the pressure correction to the axial resultant: error for K = 1, . . . , 10 is 3.3  106 and 1.7  105 for the shear
  strain correction and shear stiffness respectively, while the maxi-
@u0  2c3 @u1  mum errors are 1.0  105 and 9.3  105, respectively.
NPFE ¼ 2cD11 þ D  NFE : ð29Þ
@x FE @x FE
12
3 Fig. 6 shows the pressure corrections for the axial resultant and
bending moment for the multi-layer beam. The theoretical results
Similar expressions are used for the bending and shear corrections.
were computed by first finding the strain moment corrections from
Typical results for the variation of the approximate shear
Eqs. (18a) and (18b) and multiplying by the average constitutive
correction factor, shear stiffness and approximate pressure correc-
relation from Eq. (9). The average relative error for the pressure
tions with axial direction are plotted in Fig. 3 for the multi-layer
corrections are 3.6  105 and 4.6  105 for NP and MP, respec-
beam with K = 5. These show that there is a strong variation of
tively, while the maximum errors are, 1.8  104 and 1.2  104.
these approximations close to the ends of the beam but that these
These results demonstrate that the constitutive equation is modi-
variations quickly settle to a constant value over most of the length
fied by the presence of an externally applied pressure load, other-
of the beam. For all comparisons that follow, we average the
wise the predicted correction would be zero. In addition, these
approximate shear correction factor (28), the shear stiffness and
results show that these corrections are correctly predicted by Eq.
the approximate pressure corrections (29) obtained from the fi-
(18).
nite-element method over the span x = 4 to x = 6.
Fig. 4 shows the variation of the shear correction factor and the
average shear stiffness computed using Eqs. (26) and (25) respec- 6.1. Impact of the corrections
tively. The finite-element calculations were performed at a core ra-
tios of r = 0.2, 0.5, 0.8, 0.9, 0.95, 0.98 and at relative stiffness ratios We have demonstrated good agreement between the shear
of a = 1, 0.5, 0.1, 0.01. Good agreement is obtained at all values. strain correction factor and the pressure correction when com-
Fig. 4 shows the limiting case from Eq. (27) for zero core stiffness. pared with finite-element computations. To put these results in
Fig. 5 shows the variation of the shear correction factor com- perspective, it is necessary to assess the relative importance of
puted using the general form from Eq. (18) and the homogenized these values in predicting the stress or strain distribution and the

0.82
1 0.1 3
0.8 NP
0.05 M
P
2.5
0.8
NPFE
0.78 0
kxy P
M FE 2
D 33 /G 1

D 33 /G 1 P
-0.05
P
k xy

0.6 M
N

FE
0.76 kxy 1.5
D 33FE /G 1 -0.1
0.74 1
0.4
-0.15
0.5
0.72
-0.2
0.2 0
0.7 -0.25
0 2 4 6 8 10 0 2 4 6 8 10
X X

~xy , homogenized shear stiffness D and the pressure corrections NP and MP per unit length of the multi-layer
Fig. 3. The variation of the approximate shear correction factor, k 33
beam with K = 5. These results clearly show the end effects.

1 FE 1
α=1
FE
α = 0.5
0.95 α=1
Shear correction factor (k xy )

α = 0.1 0.8 α = 0.5


α = 0.01
0.9 α = 0.1
Limit α = 0
α = 0.01
0.6
D 33/G

0.85

0.8 0.4

0.75
0.2
0.7
0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Core fraction (r) Core fraction (r)

Fig. 4. A comparison between the shear correction factor kxy and homogenized shear stiffness D33 computed using the theory and the finite-element method for the three-
layer beam.
G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382 2381

0.168
0.84

0.166 FE
0.82
Theory
0.8 0.164

D 33/G
kxy
0.78
0.162
0.76

0.74 0.16
FE
Theory
0.72
0.158
0 10 20 30 40 50 10 20 30 40
K K

Fig. 5. A comparison between the shear correction factor kxy and homogenized shear stiffness D33 computed by theory and the finite-element method for the multi-layer
beam.

0.1

0.05 FE FE
Theory 0.05 Theory

MP
P

-0.05
N

-0.1
-0.05

-0.15

-0.1
-0.2
0 10 20 30 40 50 0 10 20 30 40 50
K K

Fig. 6. A comparison between the pressure corrections NP and MP computed using the theory K = 1. . .50 and the finite-element method K = 1. . .10.

deflection of a beam. This is a complex task that is highly problem- Pb = 1/5. The beam is composed of alternating layers as described
dependent. To make a concrete comparison, we examine two above for the K = 5 case. The dimensions of the beam are L = 10
cases: the deflection of a tip-loaded cantilever beam and the stress and c = 1/2.
distribution in a clamped–clamped pressure loaded beam. The pressure correction causes two effects: a modification of
For the case of the tip-loaded beam, we assess the importance of the constitutive relation, and additional contributions to the stress
the shear correction factor and homogenized shear stiffness. With reconstruction (7). Using the constitutive Eq. (9) and the force
no stretching-bending coupling, D12 and D21 are zero and the tip method, the stress resultants can be determined:
deflection is
" # NðxÞ ¼ PN P ;
3   !
L 4 L 1 x L2
v 0 ðLÞ ¼ Q þ : MðxÞ ¼ P ðL  xÞ   MP ;
2c D22 2c D33 kxy 2 12
 
The two terms in this expression represent a contribution to the L
Q ðxÞ ¼ P x :
deflection from the bending stiffness and a contribution from the 2
shear stiffness. The ratio of these two terms is,
 2
2c D22
r sb ¼ ;
L 4D33 kxy 2
4
where rsb is the shear to bending displacement ratio. Clearly the
slenderness ratio, Sr = L/2c, is the most important single factor. For 1
2
an isotropic beam, D22 = E and D33 = G, with kxy equal to Cowper’s
shear correction factor (2). For a Poisson ratio of m = 0.3, this results
σxy

0
σy

0
in r sb ¼ 0:765S2
r . For a reasonable slenderness ratio of Sr > 10, the
shear contributes very little to the deflection. On the other hand, σy
-2
for the three-layer symmetric beam discussed above, with m = 0.3, σxy -1
σy
FE
a = 0.01, and a core ratio r = 0.95, the shear to bending displacement
-4 σxyFE
ratio is r sb ¼ 10:14S2
r . This suggests that the shear stiffness plays a -2
much more important role in beams of this construction. Correct
determination of the shear strain correction factor and homoge- 0 2 4 6 8 10
nized shear stiffness is much more important for beams that have X
low shear stiffness such as sandwich beams. Fig. 7. A slice of the stress distribution at y = 0.6c for a clamped–clamped beam
We now examine a clamped–clamped beam subject to a distrib- with distributed pressure load on the top and bottom surface, Pt = 4/5 and Pb = 1/5.
uted pressure load on the top and the bottom surfaces, Pt = 4/5 and This location corresponds to an interface between the top and next lowest layers.
2382 G.J. Kennedy et al. / International Journal of Solids and Structures 48 (2011) 2373–2382

Note that even though the beam is symmetric, there is a non-zero References
axial compressive force and moment offset. This is due to the strain
moments caused by the pressure on the top and bottom surface of Cowper, G., 1966. The shear coefficient in Timoshenko’s beam theory. Journal of
Applied Mechanics 33 (5), 335–340.
the beam. Dong, S., Alpdogan, C., Taciroglu, E., 2010. Much ado about shear correction factors
Fig. 7 shows a comparison of ry and rxy predicted by the stress in Timoshenko beam theory. International Journal of Solids and Structures 47
reconstruction and the finite-element method over the length of (13), 1651–1665 <http://www.sciencedirect.com/science/article/B6VJS-
4YH56DC-2/2/21f0dd8f7d8e2e484f97b062d810c136>.
beam at a location y = 0.6c. These results show very good agree-
Flower, H., Soutis, C., 2003. Materials for airframes. The Aeronautical Journal 107
ment between the stress reconstruction and the finite-element re- (1072), 331–341.
sults. Neglecting the fundamental state corresponding to pressure Guiamatsia, I., 2010. A new approach to plate theory based on through-thickness
homogenization. International Journal for Numerical Methods in Engineering.
would result in ry = 0.
<http://dx.doi.org/10.1002/nme.2934>.
Guiamatsia, I., Hansen, J.S., 2004. A homogenization-based laminated plate theory.
In: ASME Conference Proceedings, vol. 47004, 2004, pp. 421–436. <http://
link.aip.org/link/abstract/ASMECP/v2004/i47004/p421/s1>
7. Conclusions Hansen, J., Kennedy, G.J., de Almeida, S.F.M., 2005. A homogenization-based theory
for laminated and sandwich beams. In: 7th International Conference on
A Timoshenko beam theory for layered orthotropic beams has Sandwich Structures, Aalborg University, Aalborg, Denmark, pp. 221–230.
Hansen, J.S., Almeida, S.D., 2001. A Theory for Laminated Composite Beams. Tech.
been presented in this paper. Following the work of Prescott Rep., Instituto Tecnológico de Aeronáutica.
(1942) and Cowper (1966), the beam kinematics are developed Hutchinson, J.R., 1980. Vibration of solid cylinders. Journal of Applied Mechanics 47
in terms of average through-thickness displacement and rotation (12), 901–907.
Hutchinson, J.R., 1981. Transverse vibration of beams, exact versus approximate
variables. The proposed theory includes a consistent method for solutions. Journal of Applied Mechanics 48 (12), 923–928.
calculating the stiffness of the beam, the shear strain correction Hutchinson, J.R., 2001. Shear coefficients for Timoshenko beam theory. Journal of
factor, and the strain-moment corrections for externally applied Applied Mechanics 68 (1), 87–92 <http://link.aip.org/link/?AMJ/68/87/1>.
Kaneko, T., 1975. On Timoshenko’s correction for shear in vibrating beams. Journal
loads. These values are based on the axially-invariant fundamental
of Physics D: Applied Physics 8 (16), 1927–1936 <http://stacks.iop.org/0022-
state solutions. We have demonstrated that the present approach 3727/8/i=16/a=003>.
easily handles layered beam constructions through the use of both Leibowitz, R.C., Kennard, E.H., 1961. Theory of freely vibrating nonuniform beams,
stress and strain moments that admit solutions where components including methods of solution and application to ships. David Taylor Model
Basin, Report 1317, p. 180–181.
of stress and strain may be discontinuous across interfaces. The Love, A.E.H., 1920. A Treatise on the Mathematical Theory of Elasticity, third ed.
external load corrections proposed in the theory modify the consti- Cambridge University Press.
tutive relationship and the equations of motion. The analysis pre- Pickett, G., 1944. Application of the Fourier method to the solution of certain
boundary problems in the theory of elasticity. Journal of Applied Mechanics 11,
sented suggests that for vibration problems, using the proposed 176–182.
theory with Cowper’s shear correction factor and a body-force Prescott, J., 1942. Elastic waves and vibrations of thin rods. Philosophical Magazine
correction is essentially equivalent to using Timoshenko’s shear 33, 703–754 <http://www.informaworld.com/10.1080/14786444208521261>.
Shames, I.H., Dym, C.L., 1985. Energy and Finite Element Methods in Structural
correction factor with the original equations of motion he derived. Mechanics. McGraw Hill Higher Education.
On the other hand, numerical comparisons using static analysis Spence, G.B., Seldin, E.J., 1970. Sonic resonances of a bar and compound torsion
demonstrated the accuracy and the consistency of the definitions oscillator. Journal of Applied Physics 41 (8), 3383–3389 <http://link.aip.org/
link/?JAP/41/3383/1>.
of the shear strain correction factor and the external load correc- Stephen, N.G., 1980. Timoshenko’s shear coefficient from a beam subjected to
tions. Both static and dynamic situations are treated by the theory gravity loading. Journal of Applied Mechanics 47 (1), 121–127 <http://
without inconsistency, as a result of the external load correction link.aip.org/link/?AMJ/47/121/1>.
Stephen, N.G., 2001. Discussion: shear coefficients for Timoshenko beam theory.
terms.
Journal of Applied Mechanics 68 (11), 959–960.
Stephen, N.G., Levinson, M., 1979. A second order beam theory. Journal of Sound and
Vibration 67 (3), 293–305.
Tafeuvoukeng, I.G., 2007. A Unified Theory for Laminated Plates. Ph.D. Thesis,
Acknowledgment University of Toronto.
Timoshenko, S.P., 1921. On the correction for shear of the differential equation for
transverse vibrations of prismatic bars. Philosophical Magazine 41, 744–746.
This research was funded by the Natural Sciences and Engineer- Timoshenko, S.P., 1922. On the transverse vibrations of bars of uniform cross-
ing Research Council (NSERC). section. Philosophical Magazine 43, 125–131.

You might also like