Free Vibration of Functionally Graded Beams and Frameworks Using The Dynamic Stiffness Method
Free Vibration of Functionally Graded Beams and Frameworks Using The Dynamic Stiffness Method
This version of the publication may differ from the final published version.
Reuse: Copies of full items can be used for personal research or study,
educational, or not-for-profit purposes without prior permission or charge.
Provided that the authors, title and full bibliographic details are credited, a
hyperlink and/or URL is given for the original metadata page and the content is
not changed in any way.
City Research Online: http://openaccess.city.ac.uk/ [email protected]
FREE VIBRATION OF FUNCTIONALLY GRADED BEAMS AND
FRAMEWORKS USING THE DYNAMIC STIFFNESS METHOD
Abstract
The free vibration analysis of functionally graded beams (FGBs) and frameworks containing
FGBs is carried out by applying the dynamic stiffness method and deriving the elements of
the dynamic stiffness matrix in explicit algebraic form. The usually adopted rule that the
material properties of the FGB vary continuously through the thickness according to a power
law forms the fundamental basis of the governing differential equations of motion in free
vibration. The differential equations are solved in closed analytical form when the free
vibratory motion is harmonic. The dynamic stiffness matrix is then formulated by relating the
amplitudes of forces to those of the displacements at the two ends of the beam. Next, the
explicit algebraic expressions for the dynamic stiffness elements are derived with the help of
technique to solve the free vibration problems of FGBs with uniform cross-section, stepped
FGBs and frameworks consisting of FGBs. Some numerical results are validated against
published results, but in the absence of published results for frameworks containing FGBs,
consistency checks on the reliability of results are performed. The paper closes with discussion
Keywords: Free vibration, functionally graded beams, dynamic stiffness method, frameworks,
Wittrick-Williams algorithm.
*
Corresponding author.
Email address: [email protected]
1
1. Introduction
In recent years interest in functionally graded material (FGM) has grown enormously. The
progress made in understanding this material has been phenomenal. One great advantage of
FGM is that the properties vary gradually in a continuous manner within the material so that
there is no abrupt change or mismatch of the properties which can cause delamination or other
problems generally associated with fibre-reinforced composites. Thus, FGM can be designed
in a way to have the properties of ceramic at one end and those of metal at the other so that
the thermal resistance of ceramic and the mechanical behaviour of metal can be exploited to
continually motivated to use various techniques and methodologies to deal with this exciting
material in order to enhance its state-of-the-art. There are now excellent books [1-4] available
on the subject. As potential application of FGM, beams which are extensively used in civil,
structural members can be investigated for their free vibration characteristics. Investigators
have expended considerable efforts which have led to the insurgence of massive literature on
the free vibration behaviour of Functionally Graded Beams (FGBs). A number of theories and
methodologies have been proposed to study the free vibration characteristics of FGBs.
Foremost amongst these are the applications of direct analytical procedure using the governing
differential equations of motion [5-19], finite element [20-22], Rayleigh-Ritz [23], finite
volume [24-26], differential quadrature [27], differential transformation [27, 28] and transfer
function [29, 30] methods. Recently the dynamic stiffness method (DSM) has also been
proposed [31, 32]. The current paper stems from the previously published DSM theories. The
entire formulation using DSM in this paper is accomplished in the real domain as opposed to
previous formulations which used complex arithmetic when developing the element dynamic
2
stiffness matrices [31, 32]. Another important further development reported in this paper is
the derivation of explicit algebraic expressions for the dynamic stiffness elements using
symbolic computation [33-35]. The explicit expressions for the dynamic stiffness elements
are particularly useful in optimisation studies and also when some, but not all of the stiffnesses
are needed. Of particular significance of this investigation is the application of DSM to analyse
the free vibration characteristics of stepped FGBs and frameworks containing FGBs. The
substantial advantages of the DSM over the conventional finite element method (FEM) are
well known [36-38]. The DSM is often called an exact method because in sharp contrast to
chosen approximate shape function assumed in the FEM, the DSM uses exact shape function
obtained from the analytical solution of the governing differential equation of motion of the
structural element in free vibration. The uncompromising accuracy of the DSM in all
frequency ranges and its independency on the number of elements used in the analysis makes
it particularly appealing in free vibration analysis. Within this pretext, the application of the
DSM in the free vibration analysis of FGBs and frameworks containing FGBs is considered
to be a welcome development. The solution technique used in the DSM is robust, particularly
when the well-established algorithm of Wittrick and Williams [39], known as Wittrick-
Williams algorithm in the literature, is applied. The algorithm ensures that no natural
frequency of the structure is missed, and it has featured in literally hundreds of papers. It is
worth noting that earlier investigations on the free vibration of FGBs were focused on
individual FGBs except for a few isolated cases where stepped FGBs with collinear axes were
reported [40, 41]. Accordingly, the literature on the free vibration of frameworks containing
FGBs is virtually non-existent. One of the essential purposes of this paper is to fill this gap.
3
2. Theory
section FGB of length L, width b and thickness h. The beam material has Young’s Modulus
E and mass density ρ which can vary through the thickness direction (Z) of the cross-section
according to the following power law distribution [14, 17, 30, 32]:
k k
z 1 z 1
E ( z ) = ( Et − Eb ) + + Eb , ρ ( z ) = ( ρ t − ρ b ) + + ρ b (1)
h 2 h 2
where Et and Eb are the Young’s moduli, and ρ t and ρb are the densities at the top and
In Eq. (1), k (k ≥ 0) is the power law index parameter which indicates the material property
variation through the beam thickness. The parameter k has been extensively discussed in the
literature [14-19] and hence it is not elaborated here. However, three special cases maybe
observed. Clearly k = 1 indicates a linear variation of the composition between the top and
bottom surfaces of the beam, k = 0 represents the case when the beam is made of full material
of the top surface whereas infinite k represents the case when the beam is made of full
The classical Bernoulli-Euler theory is considered here so that the effects of shear
deformation and rotary inertia that are relevant to the Timoshenko beam theory are assumed
to be small and hence disregarded in the analysis. Referring to Fig. 1, the displacements u1 , v1
and w1 along the X, Y and Z directions of a point on the cross-section are given by [6, 30,
32]:
∂w( y, t )
u1 = 0 , v1 ( y, z , t ) = v( y, t ) − z , w1 ( y, z, t ) = w( y, t ) (2)
∂y
4
where v and w are the corresponding displacements of a point on the neutral axis of the
beam. It should be noted that due to the variation of the material properties through the
thickness, the neutral axis would no longer be at the central line of the beam cross-section [42,
43].
Using the displacement field given by Eq. (2) and through the application of Hamilton’s
principle, the governing differential equations of motion in free vibration of the FGB are given
by [30, 32]
where
The natural boundary conditions from the Hamiltonian formulation [30, 32] give the following
expressions for axial force F , shear force S and bending moment M as follows:
Clearly, due to the use of FGM, the axial ( v ) and bending motions ( w ) are coupled as evident
v ( y , t ) = V ( y ) e iω t , w( y, t ) = W ( y )e iω t (6)
where ω is the angular or circular frequency, V ( y ) and W ( y ) are the amplitudes of v and
w , respectively.
5
d
Introducing the differential operator D = and the non-dimensional length ξ as:
dξ
y
ξ= (7)
L
The differential equations of motion in Eqs. (3) can now be written as:
( B0ω 2 L3 + A0 L D 2 )V (ξ ) − ( B1ω 2 L2 D + A1 D 3 )W (ξ ) = 0
(8)
( B1ω 2 L3 D + A1L D 3 )V (ξ ) + ( B0ω 2 L4 − B2ω 2 L2 D 2 − A2 D 4 )W (ξ ) = 0
By combining the above two differential equations, it is possible to obtain a sixth order
(D 6 + a D 4 − b D 2 − c) H = 0 (9)
where
H = V (ξ ) or W (ξ ) (10)
and
By assuming the solution in the form H = eλξ the characteristic or auxiliary equation of the
µ 3 + aµ 2 − bµ − c = 0 (13)
where
µ = λ2 (14)
6
Equation (13) can now be solved analytically for µ using standard procedure [44]. In earlier
investigations [30, 32], it was assumed that the square root of the three roots of the cubic in
Eq. (13) could be either real or complex and thus the six roots r j ( j = 1, 2,, 6) of the
characteristic equation Eq. (12) give the solutions of the differential equation (Eq. (9)) leading
6 6
V (ξ ) = ∑ R j e j , W (ξ ) = ∑ Q j e
rξ r jξ
(15)
j =1 j =1
where R j and Q j ( j = 1, 2,,6) are two sets of constants which could be possibly complex.
Clearly a method utilising Eq. (15) as the solution to derive the dynamic stiffness matrix will
involve numerical operations using complex arithmetic. This cumbersome and somehow
Using an approach similar to the one described in Refs [45, 46], it can be shown that the
three roots of the cubic equation Eq. (13) are real and the solution for H in Eq. (9) can be
exponential functions of Eq. (15). This is advantageous when deriving the explicit expressions
for the dynamic stiffness elements of the FGB. Explicit expressions are particularly useful
when some, but not all of the stiffness elements are needed, e.g. sensitivity analysis in
optimisation studies. Thus if the roots [44] of Eq. (13) are α, β and γ, the solution for H is
given by
7
where
1
1
2
q 2 φ a
α = 2 cos −
3 3 3
1
1
2
q 2 π −φ a
β = 2 cos + (17)
3 3 3
1
1
2
q π +φ a
γ = 2 cos +
2
3 3 3
with
a2
q =b+ (18)
3
and
27c − 9ab − 2a 3
φ = cos−1
(19)
( )
3
2 a 2
+ 3b 2
𝐻𝐻(𝜉𝜉) of Eq. (16) represents the solution for both axial displacement V(ξ) and bending
displacement 𝑊𝑊(ξ), containing different sets of constants as follows
The two different sets of constants 𝑄𝑄1 − 𝑄𝑄6 and 𝑅𝑅1 − 𝑅𝑅6 can be related with the help of any
one of the two of Eqs. (8) to give
k kβ kγ
Q1 = α R2 , Q3 = R4 , Q5 = R6 ,
L L L
k kβ kγ
Q2 = α R1 , Q4 = − R3 , Q6 = − R5 (22)
L L L
where
8
α (A1α 2 + B1ω 2 L2 ) β (A1β 2 − B1ω 2 L2 ) γ (A1γ 2 − B1ω 2 L2 )
kα =
(A0α 2 + B0ω 2 L2 ) , k β = (A0 β 2 − B0ω 2 L2 ) , kγ = (A0γ 2 − B0ω 2 L2 ) (23)
With the help of Eqs. (5), (20) and (21), the expressions for bending rotation 𝜃𝜃(𝜉𝜉), axial force
F(ξ), bending moment 𝑀𝑀(𝜉𝜉), and shear force 𝑆𝑆(𝜉𝜉) can be obtained after some simplification,
as
W ' (ξ ) 1
θ (ξ ) = = {R1α sinh αξ + R2α cosh αξ − R3 β sin βξ + R4 β cos βξ
L L (24)
− R5γ sin γξ + R6γ cos γξ }
' ''
A e
F (ξ ) = − 0 V − 1 W = 0 {− R1 α cosh αξ − R2 α sinh αξ + R3 β cos βξ
A A e e
L A0 L L L L L
(25)
eβ e e
+ R4 sin βξ + R5 γ cos γξ + R6 γ sin γξ}
L L L
where
The dynamic stiffness matrix of the FGB can now be formulated by applying natural
boundary conditions for displacements and forces at the ends of the beam. Referring to the
9
sign convention for positive axial force, shear force and bending moment shown in Fig. 2, the
At ξ = 0 : V = V1 , W = W1 , θ = θ1 , F = F1 , S = S1 , M = M 1 (31)
At ξ = 1 : V = V2 , W = W2 , θ = θ 2 , F = − F2 , S = − S 2 , M = − M 2 (32)
The displacement vector δ and the force vector P can be expressed as:
δ = [V1 W1 θ1 V2 W2 θ2 ]T , P = [ F1 S1 M 1 F2 S 2 M 2 ] T (33)
The relationship between the displacement δ and the constant vector R can be obtained by
δ=BR (34)
where
kα kβ kγ
0 L
0
L
0
L
1 0 1 0 1 0
α β γ
0 0 0
B = k S
L L L
kα C hα kβ Sβ k β Cβ kγ S γ k γ Cγ (35)
α hα
L − −
L L L L L
C S hα Cβ Sβ Cγ Sγ
hα
α S hα α C hα βS β βC β γSγ γCγ
L − −
L L L L L
with
Similarly, the relationship between the force vector P and the constant vector R can be
P = AR (37)
10
where
By eliminating the constant vector R from Eqs. (34) and (37), P and δ can be related to give
P=Kδ (39)
where
K = A B −1 (40)
is the required dynamic stiffness matrix with elements kij (i = 1, 2, 3..6; j=1,2, 3,..6).
With the help of symbolic computation [33-35], the matrix B of Eq. (35) was inverted
algebraically and the inverted matrix was pre-multiplied by the matrix A of Eq. (38) in order
to generate the explicit expressions for each of the elements of the dynamic stiffness matrix
computations. They are not necessarily in the shortest form, but they are surprisingly concise.
The twelve independent terms of the dynamic stiffness matrix K are given by
11
A Φ
K1,1 = K 4, 4 = 0 1
L ∆
A Φ
K1, 2 = − K 4,5 = 22 2
L ∆
A Φ
K1,3 = K 4, 6 = 2 3
L ∆
Φ
A
K1,4 = 0 4
L ∆
A Φ
K1,5 = − K 2, 4 = 22 5
L ∆
A Φ
K1, 6 = K 3, 4 = 2 6 (41)
L ∆
A Φ
K 2, 2 = K 5,5 = 32 7
L ∆
A Φ
K 2,3 = − K 5, 6 = 22 8
L ∆
A Φ
K 2,5 = 32 9
L ∆
A Φ
K 2, 6 = − K 3,5 = 22 10
L ∆
A Φ
K 3,3 = K 6 ,6 = 2 11
L ∆
A Φ
K 3,6 = 2 12
L ∆
where
12
Φ1 = ν 1 S β Cγ S hα +ν 3C β S γ S hα −ν 2 S β S γ C hα + σ 3C β Cγ C hα +ν 3C β +ν 1Cγ −ν 2 C hα
Φ 2 = −τ 3 S β Sγ S hα − λ3 S hα (1 − Cβ Cγ ) − λ1S β (1 − Cγ Chα ) − λ2 Sγ (1 − Cβ Chα )
Φ 3 = ρ1S β Cγ S hα + ρ3Cβ Sγ S hα − ρ 2 S β Sγ Chα + σ 2Cβ Cγ Chα + ρ3Cβ + ρ1Cγ − ρ 2Chα
Φ 4 = −ν 1S β S hα − ν 3 Sγ S hα + ν 2 S β Sγ + ν 3Cγ Chα − ν 2C β Cγ + ν 1C β Chα + σ 3
Φ 5 = −τ 1{αS hα (Cβ − Cγ ) − βS β (Cγ − Chα ) + γSγ (Cβ − Chα )}
Φ 6 = − ρ1S β S hα − ρ3 Sγ S hα + ρ 2 S β Sγ + ρ3Cγ Chα − ρ 2Cβ Cγ + ρ1Cβ Chα + σ 2 (42)
(
Φ 7 = −τ1 µ1S β Cγ Shα − µ2 S β Sγ Chα + µ3Cβ Sγ Shα )
Φ 8 = τ 2 S β Sγ S hα + λ3 S hα (1 − Cβ Cγ ) + λ1S β (1 − Cγ Chα ) + λ2 Sγ (1 − Cβ Chα )
Φ 9 = τ 1 (µ1S β S hα − µ 2 S β Sγ + µ3 Sγ S hα )
Φ10 = −τ 1{kα S hα (Cβ − Cγ ) − k β S β (Cγ − Chα ) + kγ Sγ (Cβ − Chα )}
Φ11 = −ξ1S β Cγ S hα + ξ 2 S β Sγ Chα − ξ 3Cβ Sγ S hα + σ 1Cβ Cγ Chα − ξ 3Cβ − ξ1Cγ + ξ 2Chα
Φ12 = ξ1S β S hα − ξ 2 S β Sγ + ξ 3 Sγ S hα − ξ 3Cγ Chα + ξ 2Cβ Cγ − ξ1Cβ Chα + σ 1
with
and
13
( ) ( ) ( ) (
∆ = µ12 − µ 2 2 + µ32 S β Sγ S hα − 2 µ1µ 2 S β 1 − Cγ Chα − 2 µ 2 µ3Sγ 1 − C β Chα + 2 µ1µ3S hα 1 − C β Cγ ) (44)
The above 6×6 frequency dependent dynamic stiffness matrix K of Eq. (40) can now be used
to compute the natural frequencies and mode shapes of either an individual FGB or an
assembly of FGBs for different boundary conditions. A reliable and accurate method of
solving the problem is to apply the Wittrick-Williams algorithm [39] which is well suited for
the application of DSM. The algorithm uses the Sturm sequence property of the dynamic
stiffness matrix to ensure that no natural frequencies of the structure analysed are missed. The
Wittrick-Williams algorithm has featured in hundreds of papers in the literature and its details
are not repeated here. Basically the algorithm [39] gives the number of natural frequencies of
a structure that lie below an arbitrarily chosen trial frequency specified by the user. As
successive trial frequencies can be chosen by the user, this simple feature of the algorithm can
be exploited to bracket any natural frequency between its upper and lower bounds to any
desired accuracy. The results given in the next section were computed by applying the
The first set of results was obtained for a uniform FGB with different boundary conditions
with the letters C, F and S denoting clamped, free and simple-support at each end of the beam.
Four classical boundary conditions are investigated, namely, clamped-free (C-F), simply-
supported (S-S), clamped-simply support (C-S) and clamped-clamped (C-C). The simple
support (S) boundary condition is assumed to be equivalent to a pinned support which prevents
both flexural and axial displacements. A wide range of investigations was carried out by
varying the length to thickness ratio (L/h) and the power law index k of the FGB which
14
controls the material property distribution through thickness. However, the authors have been
highly selective when presenting the results because existing literature already covers a huge
amount of data for natural frequencies and mode shapes with the variations of L/h and k, see
for example Ref.[30]. For presentational purposes, only the results for L/h =10 and k = 0.5
have been used in this paper, but the theory has been extensively validated against published
results.
In order to make the results universal and also to be consistent with the published results, the
L2 ρb
λi = ωi (45)
h Eb
where ωi is the ith angular natural frequency in rad/s, ρ b and E b are density and Young’s
Table. 1 shows the first five natural frequencies of the FGB with L/h = 10 and k = 0.5 for C-
F, S-S, C-S and C-C boundary conditions alongside the results reported in a recently published
paper [30]. The close agreement between the results from the current investigation and the
published ones is clearly evident. The maximum discrepancy between the two sets of results
is less than 4%. The corresponding mode shapes shown in Fig. 4, reveal that for the C-F and
C-C boundary conditions, the first, second, fourth and fifth modes are essentially bending
modes whereas the third one is axial. By contrast, for the S-S and C-S boundary conditions,
the first, second, third and fifth modes are basically bending modes and the fourth one axial.
The next set of results was obtained for a stepped beam (see Fig. 5) made of FGM, for which
some comparative results are available in the literature. The DSM theory developed in this
paper can easily account for such problems with any step location, thickness variation and
boundary conditions. However, for brevity only the results for a cantilever (C-F) stepped FGB
15
with step locations L1 = 0.25L, L1 = 0.5L and L1 = 0.75L (see Fig. 5) and the power law index
k = 0.5 are presented in Table. 2 together with the published results [41]. Note that for
consistency, FGM type – II of Ref. [41] is used so that the results are directly comparable.
Clearly, the results from the current investigation are in close agreement with those of [41].
The final set of results was obtained for a portal frame consisting of three beam members
AB, BC and CD as shown in Fig. 6. The natural frequencies of this frame are available in the
literature [47] when all the three beam members of the frame are made of isotropic material
and the supports at both points A and D are either clamped (built-in) or pinned (simply-
supported). These results are based on exact analytical theory. Using the current theory,
results are obtained for both the boundary conditions C-C and S-S at A and D, respectively
and making (i) all the three members AB, BC and CD isotropic (which is achieved by
substituting the power-law index parameter k to zero and using both the top and bottom surface
material properties to be the same and isotropic), (ii) AB and CD isotropic, but BC made of
FGM, (iii) BC isotropic and AB and CD made of FGM and (iv) AB, BC and CD are all made
of FGM. When computing numerical results, all three members of the portal frame are
assumed to have the same rectangular cross-section and length. The width and depth (height)
of the cross-section are taken to be 0.04m and 0.02m, respectively and length of each member
is set to 1m. When any of the beam members is isotropic, it is considered to be made of steel
with Young’s modulus 200 GPa and density 7500 kg/m3 whereas if it is made of FGM, the
bottom surface is considered to be steel with the above properties and the top surface ceramic
with Young’s modulus 380 GPa and density 3960 kg/m3. The computed natural frequencies
ρAL4
λi = ωi (46)
EI
16
The results of the investigation when all members of the portal frame (see Fig. 6) are metallic,
are given in Table. 3 showing the first three non-dimensional natural frequencies of the frame
when the points A and D are clamped (C-C) or simply-supported (S-S), together with the
results reported in Ref. [47]. The agreement between the sets of results in Table. 3 is excellent.
Table. 4 shows the three non-dimensional natural frequencies when the vertical members AB
and CD and the horizontal member BC of the frame are made of either isotropic metal or FGM
in turn, as indicated, and the points A and D are clamped (C-C). Similar results are obtained
for the case when the points A and D of Fig. 6 are simply-supported (S-S). The results for the
S-S case are shown in Table. 5. Clearly the results shown in Tables. 4 and 5 when compared
to Table. 3 indicate that significant changes in natural frequencies can occur as a result of
using functionally graded beams. This can have profound influence in the design of fire-
For illustrative purposes, representative mode shapes for the portal frame of Fig. 6 are
presented in Figs 7 and 8 when the points A and D of the frame are built-in (clamped) and
simply-supported, respectively. The power law index parameter k is set to 0.5 when computing
the mode shapes. Figs 7(a) and 8(a) represent the mode shapes when all three members of the
portal frame are metallic whereas Figs. 7(b) and 8(b) shows the mode shapes when the
columns AB and CD are metallic, but the beam BC is made of FGM. By contrast Figs. 7(c)
and 8(c) show the mode shapes for the case when the beam BC of the portal frame is metallic,
but its columns AB and CD are made of FGM. Finally, Figs. 7(d) and 8(d) show the mode
shapes when all three members of the portal frame are made of FGM. Although the basic
nature of the mode shapes for the portal frame remains the same depending on the order of
the natural frequency on a case to case basis, significant changes in the natural frequencies are
found to occur when using FGM as evident from Figs. 7 and 8. As expected the first mode of
17
the portal frame in each case is a sway mode with the frame oscillating between left and right
with virtually no elastic displacement of the central beam. The second mode shows elastic
deformations of all three members with no nodal point or any point of inflection within any
member. By contrast, the third mode reveals somehow a different picture in that a node with
zero displacement appears towards the top end of each columns whereas a node for the beam
appears near its centre. The mode shapes shown in Figs. 7 and 8 are typical, as expected from
the modal analysis of a portal frame and they are in accord with similar mode shapes reported
4. Conclusions
The dynamic stiffness matrix of a functionally graded beam is developed by deriving explicit
expressions for the individual stiffness elements in explicit algebraic form. This is achieved
through the application of symbolic computation. The dynamic stiffness theory is applied by
gradient beams, for which the material properties are considered to vary continuously in the
functionally graded material is also investigated for its free vibration characteristics. The
results show good agreement with published results. Importantly, the theory has been applied
to study the free vibration behaviour of a portal frame with its constituent members made of
both isotropic and functionally graded material (FGM). The investigation has shown that
significant changes in the free vibration behaviour are possible when using FGM. The
developed theory can be applied to analyse high-rise building structures made of FGM which
ceramic and it is in this context, the investigation carried out is expected to be most useful.
18
References
[1] Rasheedat Modupe Mahamood, Esther Titilayo Akinlabi, Functionally Graded Materials,
Springer International Publishing AG, 2017.
[2] Snehashish Chakraverty, Karan Kumar Pradhan, Vibration of Functionally Graded Beams
and Plates, Academic Press, 2016.
[3] Isaac Elishakoff, Demetris Pentaras, Cristina Gentilini, Mechanics of Functionally Graded
Material Structures, World Scientific, Singapore, 2015.
[4] Zheng Zhong, Linzhi Wu, Weiqiu Chen, Mechanics of Functionally Graded Materials and
Structures, Nova Science Publishers Inc, 2012.
[5] A.Y. Tang, J.X. Wu, X.F. Li, K.Y. Lee, Exact frequency equations of free vibration of
exponentially non-uniform functionally graded Timoshenko beams, Int. J. Mech. Sci. 89
(2014) 1–11.
[6] S.R. Li, Z.Q. Wan, J.H. Zhang, Free vibration of functionally graded beams based on both
classical and first-order shear deformation beam theories, Appl. Math. Mech. (English Ed.) 35
(2014) 591–606.
[7] K. Sarkar, R. Ganguli, Closed-form solutions for axially functionally graded Timoshenko
beams having uniform cross-section and fixed-fixed boundary condition, Compos. Part B Eng.
58 (2014) 361–370.
[8] T.P. Vo, H.T. Thai, T.K. Nguyen, F. Inam, Static and vibration analysis of functionally
graded beams using refined shear deformation theory, Meccanica. 49 (2014) 155–168.
[9] T.K. Nguyen, T.P. Vo, H.T. Thai, Static and free vibration of axially loaded functionally
graded beams based on the first-order shear deformation theory, Compos. Part B Eng. 55
(2013) 147–157.
[10] X.F. Li, Y.A. Kang, J.X. Wu, Exact frequency equations of free vibration of exponentially
functionally graded beams, Appl. Acoust. 74 (2013) 413–420.
[11] H.T. Thai, T.P. Vo, Bending and free vibration of functionally graded beams using
various higher-order shear deformation beam theories, Int. J. Mech. Sci. 62 (2012) 57–66.
[12] G. Giunta, D. Crisafulli, S. Belouettar, E. Carrera, Hierarchical theories for the free
vibration analysis of functionally graded beams, Compos. Struct. 94 (2011) 68–74.
[13] J. Murín, M. Aminbaghai, V. Kutiš, Exact solution of the bending vibration problem of
FGM beams with variation of material properties, Eng. Struct. 32 (2010) 1631–1640.
19
[14] M. Şimşek, Fundamental frequency analysis of functionally graded beams by using
different higher-order beam theories, Nucl. Eng. Des. 240 (2010) 697–705.
[15] S.A. Sina, H.M. Navazi, H. Haddadpour, An analytical method for free vibration analysis
of functionally graded beams, Mater. Des. 30 (2009) 741-747.
[16] X.F. Li, A unified approach for analyzing static and dynamic behaviors of functionally
graded Timoshenko and Euler-Bernoulli beams, J. Sound Vib. 318 (2008) 1210-1229.
[17] M. Aydogdu, V. Taskin, Free vibration analysis of functionally graded beams with simply
supported edges, Mater. Des. 28 (2007) 1651-1656.
[18] Z. Zhong, T. Yu, Analytical solution of a cantilever functionally graded beam, Compos.
Sci. Technol. 67 (2007) 481–488.
[19] B. V. Sankar, An elasticity solution for functionally graded beams, Compos. Sci. Technol.
61 (2001) 689-696.
[20] T.P. Vo, H.T. Thai, T.K. Nguyen, A. Maheri, J. Lee, Finite element model for vibration
and buckling of functionally graded sandwich beams based on a refined shear deformation
theory, Eng. Struct. 64 (2014) 12–22.
[21] A.E. Alshorbagy, M.A. Eltaher, F.F. Mahmoud, Free vibration characteristics of a
functionally graded beam by finite element method, Appl. Math. Model. 35 (2011) 412–425.
[22] A. Chakraborty, S. Gopalakrishnan, J.N. Reddy, A new beam finite element for the
analysis of functionally graded materials, Int. J. Mech. Sci. 45 (2003) 519–539.
[23] K.K. Pradhan, S. Chakraverty, Free vibration of Euler and Timoshenko functionally
graded beams by Rayleigh-Ritz method, Compos. Part B Eng. 51 (2013) 175–184.
[24] L.L. Jing, P.J. Ming, W.P. Zhang, L.R. Fu, Y.P. Cao, Static and free vibration analysis of
functionally graded beams by combination Timoshenko theory and finite volume method,
Compos. Struct. 138 (2016) 192-213.
[25] J. Gong, L. Xuan, P. Ming, W. Zhang, Thermoelastic analysis of functionally graded
solids using a staggered finite volume method, Compos. Struct. 104 (2013) 134–143.
[26] J.F. Gong, P.J. Ming, L.K. Xuan, W.P. Zhang, Thermoelastic analysis of three-
dimensional functionally graded rotating disks based on finite volume method, Proc. Inst.
Mech. Eng. Part C J. Mech. Eng. Sci. 228 (2013) 583–598.
[27] S. Rajasekaran, Differential transformation and differential quadrature methods for
centrifugally stiffened axially functionally graded tapered beams, Int. J. Mech. Sci. 74 (2013)
15–31.
20
[28] S. Rajasekaran, Buckling and vibration of axially functionally graded nonuniform beams
using differential transformation based dynamic stiffness approach, Meccanica. 48 (2013)
1053–1070.
[29] Y.S. Al Rjoub, A.G. Hamad, Free vibration of functionally Euler-Bernoulli and
Timoshenko graded porous beams using the transfer matrix method, KSCE J. Civ. Eng. 21
(2017) 792–806.
[30] J.W. Lee, J.Y. Lee, Free vibration analysis of functionally graded Bernoulli-Euler beams
using an exact transfer matrix expression, Int. J. Mech. Sci. 122 (2017) 1-17.
[31] H. Su, J.R. Banerjee, Development of dynamic stiffness method for free vibration of
functionally graded Timoshenko beams, Comput. Struct. 147 (2015) 107–116.
[32] H. Su, J.R. Banerjee, C.W. Cheung, Dynamic stiffness formulation and free vibration
analysis of functionally graded beams, Compos. Struct. 106 (2013) 854-862.
[33] J. Fitch, Solving algebraic problems with REDUCE, J. Symb. Comput. 1 (1985) 211-227.
[34] A.C. Hearn, REDUCE User’s Manual, Version 3.5, RAND Publication, Santa Monica,
California, USA, 1993.
[35] J.R. Banerjee, A.J. Sobey, H. Su, J.P. Fitch, Use of computer algebra in Hamiltonian
calculations, Adv. Eng. Software, 39 (2008) 521-525.
[36] F.W. Williams, W.H. Wittrick, Exact buckling and frequency calculations surveyed, J.
Struct. Eng, ASCE, 109 (1983) 169-187.
[37] F.W. Williams, Review of exact buckling and frequency calculations with optional multi-
level substructuring, Comput. Struct., 48 (1993) 547-552.
[38] J.R. Banerjee, Dynamic stiffness formulation for structural elements: A general approach,
Comput. Struct. 63 (1997) 101-103.
[39] W.H. Wittrick, F.W. Williams, A general algorithm for computing natural frequencies of
elastic structures, Quart. J. Mech. Appl. Math. 24 (1971) 263-284.
[40] D.V. Bambill, C.A. Rossit, D.H. Felix, Free vibrations of stepped axially functionally
graded Timoshenko beams, Meccanica. 50 (2015) 1073-1087.
[41] N. Wattanasakulpong, J. Charoensuk, Vibration characteristics of stepped beams made
of FGM using differential transformation method, Meccanica. 50 (2015) 1089-1101.
[42] M.A. Eltaher, A.E. Alshorbagy, F.F. Mahmoud, Determination of neutral axis position
and its effect on natural frequencies of functionally graded macro/nanobeams, Compos. Struct.
99 (2013) 193-201.
21
[43] K.S. Al-Basyouni, A. Tounsi, S.R. Mahmoud, Size dependent bending and vibration
analysis of functionally graded micro beams based on modified couple stress theory and
neutral surface position, Compos. Struct. 125 (2015) 621-630.
[44] L.A. Pipes, L.R. Harvill, Applied mathematics for engineers and physicists, McGraw-
Hill, 1970.
[45] J.R. Banerjee, Coupled bending-torsional dynamic stiffness matrix for beam elements,
Int. J. Num. Meth. Eng. 28 (1989) 1283-1298.
[46] J.R. Banerjee, Explicit analytical expressions for frequency equation and mode shapes of
composite beams, Int. J. Solids Structs. 38 (2001) 2415-2426.
[47] C.P. Filipich, P.A.A. Laura, In-plane vibrations of portal frames with end supports
elastically restrained against rotation and translation, J. Sound Vib. 117 (1987) 467–473.
[48] R.O. Grossi, C.M. Albarracin, A variational approach to vibrating frames, Proc. Inst.
Mech. Eng., Part K: J. Multibody Dyn. 221 (2007) 247-259.
[49] I. Tatar, Vibration characteristics of portal frames, MSc Thesis, June 2013, Izmir Institute
of Technology, Graduate School of Engineering and Sciences, Izmir, Turkey.
22