You are on page 1of 11

International Journal of Mechanical Sciences 150 (2019) 337–347

Contents lists available at ScienceDirect

International Journal of Mechanical Sciences


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

An exact dynamic stiffness matrix for a beam incorporating Rayleigh–Love


and Timoshenko theories
J.R. Banerjee∗, A. Ananthapuvirajah
Department of Mechanical Engineering and Aeronautics, School of Mathematics, Computer Science and Engineering, City, University of London, Northampton Square,
London EC1V 0HB, United Kingdom

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

Keywords: An exact dynamic stiffness matrix for a beam is developed by integrating the Rayleigh–Love theory for longitu-
Rayleigh–Love bar dinal vibration into the Timoshenko theory for bending vibration. In the formulation, the Rayleigh–Love theory
Timoshenko beam accounted for the transverse inertia in longitudinal vibration whereas the Timoshenko beam theory accounted
Dynamic stiffness method
for the effects of shear deformation and rotating inertia in bending vibration. The dynamic stiffness matrix is
Wittrick–Williams algorithm
developed by solving the governing differential equations of motion in free vibration of a Rayleigh–Love bar
and a Timoshenko beam and then imposing the boundary conditions for displacements and forces. Next the two
dynamic stiffness theories are combined using a unified notation. The ensuing dynamic stiffness matrix is subse-
quently used for free vibration analysis of uniform and stepped bars as well as frameworks through the application
of the Wittrick–Williams algorithm as solution technique. Illustrative examples are given to demonstrate the use-
fulness of the theory and some of the computed results are compared with published ones. The paper closes with
some concluding remarks.

1. Introduction dergraduate engineering curriculum across the globe. As the classical


results of the free vibration analysis come from the solution of the gov-
Free vibration analysis in the high frequency range is of great im- erning differential equation, they are generally considered to be exact.
portance to assess the flow of vibrational energy in structures, particu- Similar and comparable, but somehow approximate results for beam
larly when the widely accepted Statistical Energy Analysis (SEA) method vibration problems can also be obtained by applying the FEM which re-
[1,2] is used. Research in this area is further motivated by the fact that quires discretisation of the beam into several elements and assembling
the modal density required for the energy flow analysis in structures the element stiffness and mass matrices which ultimately lead to a lin-
is generally very high in the high frequency range. To this end there ear eigenvalue formulation. Clearly the order of the mass and stiffness
are several research papers in the published literature on the energy matrices in the FEM decides the number of natural frequencies that can
flow analysis in classical structures such as bars [3], beams [4], mem- be meaningfully computed. The higher order natural frequencies and
branes [5] and plates [6] which emphasize the need for high frequency mode shapes will, of course, be considerably less accurate. At this point
vibration analysis. For accurate and efficient high frequency vibration it should be noted that there is a powerful alternative to FEM as well
analysis, these publications highlight the inadequacy of the traditional the classical method, which has no restriction on higher order natural
finite element method (FEM) which is somehow limited to low and per- frequency computation and yet it retains the exactness of results. The
haps medium frequency range unless high-precision, good quality finite alternative is that of the dynamic stiffness method (DSM) which is ele-
elements are used which may become computationally very expensive. gant and versatile and hence can be used in a much broader context to
In the particular context of free vibration of beams, there are numerous analyse the free vibration behaviour of complex structures. The DSM is
books on mechanical vibration [7–11] which give the natural frequen- different, but in many ways similar to the FEM in that it has analogous
cies and mode shapes in longitudinal, torsional and bending vibration procedure for assembling structural properties of individual structural
through the solution of the governing differential equations and impo- elements. However, a major difference exists between the DSM and the
sition of the boundary conditions to eliminate the integration constants FEM which is that the former is unaffected by the number of elements
which eventually lead to the frequency equation. This standard and rel- used in the analysis and always gives exact results whereas the latter is
atively simple procedure is straightforwardly taught in most of the un- mesh dependent and the accuracy of results depends on the number of
elements used in the analysis. For instance, one single structural element
can be used in the DSM to compute any number of natural frequencies

Corresponding author.
without any loss of accuracy which, of course, is impossible in the FEM.
E-mail address: j.r.banerjee@city.ac.uk (J.R. Banerjee).

https://doi.org/10.1016/j.ijmecsci.2018.10.012
Received 9 June 2018; Received in revised form 2 October 2018; Accepted 4 October 2018
Available online 10 October 2018
0020-7403/© 2018 Elsevier Ltd. All rights reserved.
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

The DSM was pioneered by Kolousek [12–14] in the early 1940s and it is solved and some of the computed results are compared with pub-
has since been used to investigate the free vibration behaviour of beams lished literature. The paper draws significant conclusions on the effects
and frameworks in an exact sense [15–17]. The uncompromising ac- of the inclusion of transverse inertia arising from the Rayleigh-Love the-
curacy of the DSM stems from the fact that the frequency dependent ory and shear deformation and rotatory inertia from the Timoshenko
shape function used to derive the element dynamic stiffness matrix of a theory when investigating the free vibration characteristics of individ-
structural element comes from the exact solution of the governing differ- ual members, stepped members as well as frameworks.
ential equation of motion of the element undergoing free natural vibra- It should be noted that there are no specific hard boundaries between
tion. The element dynamic stiffness matrix derived in this way contains the regimes of low, medium and high frequencies, but a useful descriptor
both the mass and stiffness properties of the element, unlike the FEM for which gives an indicative guidance to frequency range is the vibrational
which the mass and stiffness matrices are always separate and frequency wavelength when compared to the overall length of the structure. Thus
independent, and they are generally derived from assumed shape func- an engineering judgement can be reasonably made based on the prod-
tions. An outline for the procedure to derive the dynamic stiffness matrix uct of the wave number and a typical length of the structure, which is
of a structural element can be found in the work of Banerjee [18]. The essentially the Helmholtz number. Large values of this number repre-
overall frequency dependent dynamic stiffness matrix of the final struc- sent the high frequency range whereas lower values determine the low
ture is obtained by assembling the individual dynamic stiffness matrices to medium frequency range. For the type of problems investigated in
of all constituent elements in the structure, in the usual way as in the this paper, the low to medium range of frequencies is characterised to
case of the FEM, but the formulation leads to a non-linear eigenvalue be below 1500 Hz whereas frequencies above this value constitute the
problem and the natural frequencies are generally extracted by applying high frequency range.
the well-established algorithm of Wittrick and William [19]. Because of
the independency of the accuracy of results on the number of element 2. Dynamic stiffness formulation
used in the analysis, the DSM is ideally suited for free vibration analysis
in all frequency ranges. The dynamic stiffness matrix of a structural element essentially re-
Following the work of Kolousek [12–14], the DSM has been imple- lates the amplitudes of the forces to those of the corresponding displace-
mented in computer programs published by Akesson [15], Williams and ments at the nodes of the harmonically vibrating structural element. A
Howson [16] and Howson et al [17] to investigate the free vibration general procedure to formulate the dynamic stiffness matrix of a struc-
characteristics of plane frames, which required the dynamic stiffness tural element is briefly described in following steps:
matrices of both bar and beam elements as building blocks. The bar the-
(i) Derive the governing differential equation of motion in free vibration
ory accounts for the axial stiffness and the beam theory accounts for the
of the structural element for which the dynamic stiffness matrix is to
bending stiffness. As the coupling between them is generally ignored,
be developed. This can be achieved by applying Newton’s second law
the dynamic stiffness matrix of the element used to investigate the free
or Lagrange’s equation or Hamilton’s principle. However, Hamilton’s
vibration behaviour of plane frames [15–17] was obtained by separate
principle is preferred because unlike Newton’s second law and La-
consideration of axial and bending deformation and then combining the
grange’s equation, the variationally based Hamilton’s principle pro-
two together in matrix form. In these earlier works, when the axial stiff-
vides natural boundary conditions, giving the expressions for forces
nesses were incorporated into the bending stiffnesses to construct the
and moments which are required in the dynamic stiffness formula-
dynamic stiffness matrix of an individual element, only classical theory
tion.
for longitudinal free vibration of bars which ignores the transverse iner-
(ii) For harmonic oscillation, seek a closed form analytical solution of
tia effect was used. This is generally justified, particularly in the low and
the governing differential equation derived in (i) above, in terms of
probably in the medium frequency range, but for high frequency vibra-
the arbitrary integration constants. The number of constants in the
tion, the so-called Rayleigh–Love theory [20,21] which accounts for the
general solution will, of course, depend on the order of the differen-
effects of transverse inertia during longitudinal vibration and the Timo-
tial equation.
shenko theory [17] which accounts for the effects of shear deformation
(iii) Apply the boundary conditions in algebraic form. The number of
and rotatory inertia during bending vibration need to be considered.
boundary conditions is generally equal to twice the number of inte-
This is particularly important when applying the widely accepted SEA
gration constants. The boundary conditions are typically the nodal
technique for which the high frequency vibration problem must be mod-
displacements and forces.
elled properly [1,2]. In this respect, the traditional FEM may become
(iv) Eliminate the constants by relating the harmonically varying am-
inaccurate.
plitudes of nodal forces to the corresponding displacements at the
From a historical perspective, it was Lord Rayleigh [22] who first
nodes of the element. This will generate the frequency dependent
recognised the importance of transverse inertia on the longitudinal free
dynamic stiffness matrix connecting dynamically the amplitudes of
vibration of bars, particularly at high frequencies. Many years later,
the nodal forces to those of the nodal displacements.
Love [23] shed further lights on Lord Rayleigh’s work which eventually
took the name Rayleigh–Love theory and the research took significant The axial deformation of a Rayleigh–Love bar and the bending defor-
turn to wave propagation and vibrational energy analysis [24–26] of mations of a Timoshenko beam are considered uncoupled and treated
bars in longitudinal motion. No one appears to have made any attempt independently so that the derivation of the dynamic stiffness matrix for
to combine the Rayleigh–Love bar analysis with flexure, particularly each of them can be carried out separately, and later integrated.
when investigating the free vibration characteristics of frameworks. This
will be important within the high frequency range when using the SEA 2.1. Dynamic stiffness matrix of a Rayleigh–Love bar
technique. The purpose of this paper is to fill this gap in the literature.
First, the dynamic stiffness matrix of a Rayleigh–Love bar is developed A uniform Rayleigh–Love bar of length L is shown in Fig. 1 in a
from the fundamental equation of motion in longitudinal free vibration. rectangular right handed Cartesian co-ordinate system with the X-axis
Then the developed dynamic stiffness matrix of the Rayleigh–Love bar coinciding with the axis of the bar. Note that Fig. 1 can also be used
is integrated with the dynamic stiffness matrix of a Timoshenko beam to represent a beam which is also a two-noded line element like a bar
[27–30] which accounts for the effects of shear deformation and rota- element. The essential difference between a bar and a beam element
tory inertia to allow for the free vibration analysis of individual mem- is that the former can sustain only axial load whereas the latter can
bers and plane frames in the low, medium and high frequency range take bending and shear load, as well as the axial load. In other words,
through the application of the Wittrick–Williams algorithm [19] as so- in any local coordinate system such as the one shown in Fig. 1, a bar
lution technique. Using the developed theory, a wide range of problems element can undergo only axial deformation whereas a beam element

338
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

Fig. 2. Boundary conditions for displacements and forces in axial vibration for
a Rayleigh–Love bar.

where 𝜔 is the angular or circular frequency, and U(x) are the ampli-
tudes of u.
Substituting Eq. (6) into Eq. (4) gives
( ) 𝑑2𝑈
𝐸𝐴 − 𝜌𝐼𝑃 𝜈 2 𝜔2 + 𝜌𝐴𝜔2 𝑈 = 0 (7)
Fig. 1. Coordinate system and notation for a Rayleigh–Love bar and a Timo- 𝑑 𝑥2
shenko beam. As a result of the harmonic oscillation assumption, the amplitude
F(x) of the force f(x, t) in Eq. (5) becomes
( ) 𝑑𝑈
can undergo bending displacement, bending rotation as well as axial 𝐹 (𝑥) = − 𝐸𝐴 − 𝜌𝐼𝑃 𝜈 2 𝜔2 (8)
deformation. Now the governing differential equation of motion of the 𝑑𝑥
Rayleigh-Love bar in free axial (or longitudinal) vibration can be derived Introducing the differential operator D = d/d𝜉 and the non-
by using Hamilton’s principle as the first step towards the dynamic stiff- dimensional length 𝜉 as
ness formulation. The focus area of the derivation in this section is, of 𝑥
𝜉= (9)
course, on the axial stiffnesses only. 𝐿
Referring to Fig. 1 and noting that if u is the axial displacement at a Eq. (7) becomes
distance x from the origin, the kinetic and potential energies of the bar (𝐷 2 + 𝛾 2 )𝑈 = 0 (10)
Tbar and Vbar are respectively given by [7,21]
[ ( 2 )2 ] where
( )2
1 𝐿 𝜕𝑢 𝜕 𝑢 𝛼2
𝑇bar = ∫ 𝜌𝐴 + 𝜌𝐼𝑃 𝜈 2
𝑑𝑥 (1) 𝛾2 = (11)
20 𝜕𝑡 𝜕 𝑥𝜕 𝑡 1 − 𝛽2
and with
𝐿 ( )2 𝜌𝐴𝜔2 𝐿2 2 𝜌𝐼𝑃 𝜈 2 𝜔2
1 𝜕𝑢 𝛼2 = ;𝛽 = (12)
𝑉bar = ∫ 𝐸𝐴 𝑑x (2) 𝐸𝐴 𝐸𝐴
20 𝜕𝑥
The expression for the amplitude of the axial force in Eq. (8) using
where 𝜌 is the density of the bar material, A is the cross-sectional area
Eqs. (9) and (12) becomes
of the bar so that 𝜌A represents the mass per unit length, IP is the polar
second moment of area so that 𝜌IP represents the polar mass moment of 𝐸𝐴 ( ) 𝑑𝑈
𝐹 (𝜉) = − 1 − 𝛽2 (13)
inertia per unit length, E is the Young’s modulus of the bar material so 𝐿 𝑑𝜉
that EA represents the axial or extensional rigidity of the bar and 𝜈 is The solution of the differential equation, Eq. (10) is given by
the Poisson’s ratio of the bar material. 𝑈 (𝜉) = 𝐶1 sin 𝛾𝜉 + 𝐶2 cos 𝛾𝜉 (14)
Hamilton’s principle states
where C1 and C2 are constants.
𝑡2 ( )
𝛿 ∫ 𝑇bar − 𝑉bar 𝑑𝑡 = 0 (3) The expression for axial force F(𝜉) can now be expressed by substi-
𝑡1 tuting Eq. (14) into Eq. (13) to give
where t1 and t2 are the time interval in the dynamic trajectory, and 𝛿 is 𝐸𝐴 ( )
𝐹 (𝑥) = 𝐹 (𝜉) = − 1 − 𝛽 2 𝛾(𝐶1 cos 𝛾𝜉 − 𝐶2 sin 𝛾𝜉) (15)
the usual variational operator. 𝐿
The governing differential equations of motion of the Rayleigh–Love Now referring to Fig. 2, the boundary conditions for displacements
bar and the associated boundary condition in free vibration can now and forces can be applied as follows.
be derived by substituting the kinetic (Tbar ) and potential (Vbar ) energy
At 𝑥 = 0 (i.e. 𝜉 = 0), 𝑈 = Δ𝑥1 and 𝐹 = 𝐹𝑥1 (16)
expressions of Eqs. (1) and (2) into Eq. (3), using the 𝛿 operator, in-
tegrating by parts and then collecting terms. In an earlier publication,
the entire procedure to generate the governing differential equations of At 𝑥 = 𝐿(i.e. 𝜉 = 1), 𝑈 = Δ𝑥2 and 𝐹 = −𝐹𝑥2 (17)
motion and natural boundary conditions for bar or beam type structures
Substituting Eqs. (16) and (17) into Eqs. (14) and (15), the following
was automated by Banerjee et al [31] by applying symbolic computa-
matrix relationships can be obtained
tion. In this way, the governing differential equation of motion of the [ ] [ ][ ]
Rayleigh–Love bar is obtained as [7,21] Δ𝑥 1 0 1 𝐶1
= (18)
Δ𝑥2 sin 𝛾 cos 𝛾 𝐶2
𝜕2 𝑢 𝜕2 𝑢 𝜕4 𝑢
𝐸𝐴 − 𝜌𝐴 + 𝜌𝐼𝑃 𝜈 2 =0 (4) and
𝜕𝑥 2 𝜕𝑡 2 𝜕 𝑥2 𝜕 𝑡2 [ ] [ ][ ]
As a by-product of the Hamiltonian formulation, the expression for 𝐹𝑥1 𝐸𝐴 ( ) −1 0 𝐶1
= 𝛾 1 − 𝛽2 (19)
the axial force f(x, t) follows from the natural boundary condition to 𝐹𝑥2 𝐿 cos 𝛾 − sin 𝛾 𝐶2
give [7,21] The constants C1 and C2 can now be eliminated from Eqs. (18) and
𝜕𝑢 𝜕3 𝑢 (19) to give the dynamic stiffness matrix of an axially vibrating
𝑓 (𝑥, 𝑡) = −𝐸𝐴 − 𝜌𝐼𝑃 𝜈 2 (5) Rayleigh–Love bar relating amplitudes of the forces and displacements
𝜕𝑥 𝜕 𝑥𝜕 𝑡2
at its ends as follows:
If harmonic oscillation is assumed, then [ ] [ ][ ]
𝐹𝑥1 𝑎 𝑎2 Δ𝑥 1
𝑢(𝑥, 𝑡) = 𝑈 (𝑥)𝑒𝑖𝜔𝑡 = 1 (20)
(6) 𝐹𝑥2 𝑎2 𝑎1 Δ𝑥2

339
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

where the elements of the 2 × 2 dynamic stiffness matrix are given by


𝐸𝐴 ( ) 𝐸𝐴 ( ) 𝜃(𝑥, 𝑡) = Θ(𝜉)𝑒𝑖𝜔𝑡 (32)
𝑎1 = 𝛾 1 − 𝛽 2 cot 𝛾, 𝑎2 = − 𝛾 1 − 𝛽 2 cosec𝛾 (21)
𝐿 𝐿 where W(𝜉) and Θ(𝜉) are the amplitudes of the bending displacement
It should be noted that the Rayleigh–Love theory has a limitation and bending rotation of the harmonically vibrating Timoshenko beam.
that 𝛽 2 in Eqs. (11) and (12) must be less than one which is usually Eqs. (27) and (28) can now be combined to give a fourth order or-
the case, otherwise, the solution of Eq. (10) would not be harmonic and dinary differential equation as follows which is identically satisfied by
hence no oscillatory motion will take place. This limitation has been both W(𝜉) and Θ(𝜉)
pointed out in the literature, e.g. see Eq. (13) of [32]. [ 4 ( ) ( )]
𝐷 + 𝑏2 𝑟2 + 𝑠2 𝐷 2 − 𝑏2 1 − 𝑏2 𝑟2 𝑠2 𝐻 = 0 (33)
2.2. Dynamic stiffness matrix of a Timoshenko beam where
𝑑 𝑑
𝐷= =𝐿 (34)
The dynamic stiffness matrix of a Timoshenko beam has already been 𝑑𝜉 𝑑𝑥
published in the literature [27–30] in a rather longwinded and compli-
𝜌𝐴𝜔2 𝐿4 2 𝐼 𝐸𝐼
cated manner, the details of which are not repeated here. However, for 𝑏2 = ;𝑟 = ; 𝑠2 = (35)
𝐸𝐼 𝐴𝐿2 𝑘𝐴𝐺𝐿2
clarity, completeness and importantly to make this paper self-contained,
the existing literature is concisely congregated and simplified. The pro- and
cedure is briefly summarised below. 𝐻 = 𝑊 or Θ (36)
Considering Fig. 1 to be the Timoshenko beam under investigation
with bending rigidity EI, mass per unit length 𝜌A and length L, undergo- If a trial solution H = e𝜆𝜉
is assumed where 𝜆 is a constant, yet to
ing bending displacement w and bending rotation 𝜃, the expressions for be known, the auxiliary or characteristic equation of the differential
kinetic and potential energies Tbeam and Vbeam are respectively given by Eq. (33) is given by
( ) ( )
[33] 𝜆4 + 𝑏2 𝑟2 + 𝑠2 𝜆2 − 𝑏2 1 − 𝑏2 𝑟2 𝑠2 = 0 (37)
( ) ( )2
1 𝐿 𝜕𝑤 2 1 𝐿 𝜕𝜃 Eq. (37) is quartic in 𝜆, but quadratic in 𝜆2 so that
𝑇beam = ∫ 𝜌𝐴 𝑑 𝑥 + ∫ 𝜌𝐼 𝑑𝑥 (22)
20 𝜕𝑡 20 𝜕𝑡 √
( ) { ( )}2 ( )
−𝑏2 𝑟2 + 𝑠2 ± 𝑏2 𝑟2 + 𝑠2 + 4𝑏2 1 − 𝑏2 𝑟2 𝑠2
( )2
1 𝐿 𝜕𝜃 1 𝐿 𝜆 =
2
𝑉beam = ∫ 𝐸𝐼 𝑑 𝑥 + ∫ 𝑘𝐴𝐺𝛾 2 𝑑 𝑥 (23) 2
20 𝜕𝑥 20 √
( ) { ( )}2
−𝑏2 𝑟2 + 𝑠2 ± 𝑏2 𝑟2 − 𝑠2 + 4𝑏2
In Eqs. (22) and (23), 𝜌I is the rotatory inertia per unit length about = (38)
the bending axis, kAG is, the shear rigidity of the beam with k being the 2
shear correction (also known as the shape factor) and 𝛾 is the angle of Clearly 𝜆2 will be always real and for the negative value of the ex-
shear deformation which is essentially the shearing strain. It should be pression under the square root sign of Eq. (38), one of the two values of
noted that in the Timoshenko beam formulation the total slope 𝜕𝑤 𝜕𝑥
is the 𝜆2 will be always negative, resulting in two imaginary roots of 𝜆 which
sum of both bending slope 𝜃 and the slope due to shear 𝛾 [33] so that will lead to part of the solution of Eq. (33) in terms of trigonometric func-
𝜕𝑤 tions whereas the other value of 𝜆2 when using the positive value be-
=𝜃+𝛾 (24)
𝜕𝑥 fore the square root sign can be either positive or negative depending on
or whether the square root expression in Eq. (38) is bigger than or smaller
𝜕𝑤 than b2 (r2 + s2 ). If this second value of 𝜆2 is positive which is usually
𝛾= −𝜃 (25)
𝜕𝑥 the case, the two roots of 𝜆2 will be real, yielding the remaining solu-
Thus, the potential energy Vbeam of Eq. (23) becomes tion of Eq. (33) in terms of hyperbolic functions so that the two of the
( )2 ( )2 four integration constants in the solution will be connected to trigono-
1 𝐿 𝜕𝜃 1 𝐿 𝜕𝑤 metric functions and the other two to hyperbolic functions. However,
𝑉beam = ∫ 𝐸𝐼 𝑑 𝑥 + ∫ 𝑘𝐴𝐺 − 𝜃 𝑑𝑥 (26)
20 𝜕𝑥 20 𝜕𝑥 for exceptionally high frequencies or for exceptionally squat beams, the
Substituting the expressions for the kinetic and potential energies second value of 𝜆2 can be negative like the first one which will give the
Tbeam and Vbeam from Eqs. (22) and (26) into Hamilton’s principle and entire solution of Eq. (33) in terms of trigonometric functions only. The
then integrating by parts and collecting terms yield the governing dif- two sets of solutions and their conditionality are explained below.
ferential equations of motion and the associated boundary conditions The expression for 𝜆2 in Eq. (38) can be expressed in the following
providing the expressions for bending moment (M) and shear force (S) alternative form
{ √ }
as follows [33]. ( 2 ) ( )2
𝑏2 4( )
Governing differential equations 𝜆 =
2
− 𝑟 +𝑠 ± 2
𝑟 +𝑠
2 2 + 1−𝑏 𝑟 𝑠
2 2 2 (39)
2 𝑏2
( )
𝜕2 𝑤 𝜕 𝜕𝑤
−𝜌𝐴 + 𝑘𝐴𝐺 −𝜃 =0 (27) It is clear from Eq. (39) that if b2 r2 s2 < 1, one of the values of 𝜆2
𝜕𝑡 2 𝜕𝑥 𝜕𝑥
will be negative and the other value will be positive whereas if b2 r2 s2 >
1, they both will be negative. Thus the solutions for bending displace-
( )
𝜕2 𝜃 𝜕2 𝜃 𝜕𝑤 ment W and bending rotation Θ for these two cases resulting from the
−𝜌𝐼 + EI + kAG −𝜃 =0 (28)
𝜕𝑡 2 𝜕𝑥 2 𝜕𝑥 differential equation of Eq. (33) are given by
Natural boundary conditions (i) b2 r2 s2 < 1
( )
𝜕𝑤 𝜕2 𝜃 𝜕2 𝜃 𝑊 (𝜉) = 𝐴1 cos Φξ + 𝐴2 sin Φξ + 𝐴3 cosh Λξ + 𝐴4 sinh Λξ
Shear Force ∶ 𝑣 = −𝑘𝐴𝐺 − 𝜃 = 𝐸𝐼 − 𝜌𝐼 (29) (40)
𝜕𝑥 𝜕𝑥 2 𝜕 𝑡2
𝜕𝜃 Θ(𝜉) = 𝐵1 cos Φξ + 𝐵2 sin Φξ + 𝐵3 cosh Λξ + 𝐵4 sinh Λξ (41)
Bending Moment ∶ 𝑚 = −𝐸𝐼 (30)
𝜕𝑥 (ii) b2 r2 s2 >1
Introducing the non-dimensional length 𝜉 = x/L and assuming har-
𝑊 (𝜉) = 𝐴1 cos Φξ + 𝐴2 sin Φξ + 𝐴3 cos Λξ + 𝐴4 sin Λξ (42)
monic oscillation so that
𝑤(𝑥, 𝑡) = 𝑊 (𝜉)𝑒𝑖𝜔𝑡 (31) Θ(𝜉) = 𝐵1 cos Φξ + 𝐵2 sin Φξ + 𝐵3 cos Λξ + 𝐵4 sin Λξ (43)

340
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

where
( ) √
𝑏2 𝑟2 + 𝑠2 𝑏2 ( )2 4( )
2
Φ = + 𝑟2 + 𝑠2 + 1 − 𝑏2 𝑟2 𝑠2 (44)
2 2 𝑏2
and
( ) √
𝑏2 𝑟2 + 𝑠2 𝑏2 ( )2 4( )
𝑗Λ = −2
+ 𝑟2 + 𝑠2 + 1 − 𝑏2 𝑟2 𝑠2 (45)
2 2 𝑏2
with

𝑗 = 1 for 𝑏2 𝑟2 𝑠2 < 1; 𝑗 = −1 for 𝑏2 𝑟2 𝑠2 > 1 (46)

and A1 –A4 and B1 –B4 are two different sets of constants. Fig. 3. Boundary conditions for displacements and forces for a Timoshenko
It should be noted from Eq. (35) that beam.
𝜌𝐼 𝜔2
𝑏2 𝑟2 𝑠2 = (47)
𝑘𝐴𝐺 and forces, respectively, in terms of the constants A1 −A4 .
For most of the practical problems, b2 r2 s2 will be less than one un-
⎡ Δ𝑦1 ⎤ ⎡ 1 0 1 0 ⎤⎡ 𝐴1 ⎤
less 𝜔 is exceptionally large. This is because the shear rigidity kAG is ⎢ ⎥ ⎢ ⎥⎢ ⎥
generally much bigger than the rotatory inertia per unit length 𝜌I for ⎢ Θ1 ⎥ ⎢ 0 𝑘Φ ∕𝐿 0 𝑘Λ ∕𝐿 ⎥⎢ 𝐴2 ⎥
any realistic cross-section and beam material, but nevertheless, the so- ⎢ ⎥=⎢ ⎥⎢ ⎥ (55)
⎢ Δ𝑦2 ⎥ ⎢ 𝐶 𝑆 𝐶 𝑆 ⎥⎢ 𝐴3 ⎥
lutions given by Eqs. (42) and (43) are included in the theory to cover ⎢ ⎥ ⎢ ⎥⎢ ⎥
⎣ Θ2 ⎦ ⎣ −𝑘Φ 𝑆∕𝐿 𝑘Φ 𝐶∕𝐿 𝑗𝑘Λ 𝑆 ∕𝐿 𝑘Λ 𝐶 ∕𝐿 ⎦⎣ 𝐴4 ⎦
the exceptional case when b2 r2 s2 is greater than one.
With the help of Eq. (27) or (28) and the solution given by Eqs. (40)– or
(43), it can be shown that the two sets of constants A1 –A4 and B1 –B4
𝚫 = 𝐐A (56)
are related. Using Eq. (27), the following relationships between B1 –B4
and A1 –A4 are obtained. and
𝑘Φ 𝑘 𝑘 𝑘 ⎡ 𝐹 𝑦 1 ⎤ ⎡0 −𝑊3 𝑒Φ 0 𝑊3 𝑒Λ ⎤⎡𝐴1 ⎤
𝐵1 = 𝐴 ; 𝐵 = − Φ 𝐴1 ; 𝐵3 = Λ 𝐴4 ; 𝐵4 = 𝑗 Λ 𝐴3 (48) ⎢ ⎥ ⎢ ⎥⎢ ⎥
𝐿 2 2 𝐿 𝐿 𝐿 ⎢𝑀1 ⎥ = ⎢𝑊2 Φ𝑘Φ 0 −𝑗 𝑊2 Λ𝑘Λ 0 ⎥⎢𝐴2 ⎥ (57)
where ⎢ 𝐹 𝑦 2 ⎥ ⎢− 𝑊 3 𝑒 Φ 𝑆 𝑊3 𝑒Φ 𝐶 −𝑗 𝑊3 𝑒Λ 𝑆̄ −𝑊3 𝑒Λ 𝐶̄ ⎥⎢𝐴3 ⎥
⎢𝑀 ⎥ ⎢−𝑊 Φ𝑘 𝐶 −𝑊2 Φ𝑘Φ 𝑆 𝑗 𝑊2 Λ𝑘Λ 𝐶̄ 𝑗 𝑊2 Λ𝑘Λ 𝑆̄ ⎥⎦⎢⎣𝐴4 ⎥⎦
( 2 ) ( 2 ) ⎣ 2⎦ ⎣ 2 Φ
Φ − 𝑏2 𝑠2 Λ + 𝑗 𝑏2 𝑠2
𝑘Φ = ; 𝑘Λ = (49) or
Φ Λ
Because of the harmonic oscillation hypothesis adopted for the freely 𝐅 = 𝐑𝐀 (58)
vibrating Timoshenko beam as indicated by Eqs. (31) and (32) and also where
by the introduction of the non-dimensional length 𝜉 = x/L, the expres-
𝑆 = sin Φ; 𝐶 = cos Φ (59)
sions for the amplitudes of the shear force (V) and bending moment (M)
arising from Eqs. (29), (30) and (48) will take the following form. }
𝑆̄ = sinh Λ; 𝐶̄ = cosh Λ 𝑏2 𝑟2 𝑠2 < 1(𝑗 = 1)
( ) (60)
EI 𝑑 2 Θ 𝑆̄ = sin Λ; 𝐶̄ = cos Λ 𝑏2 𝑟2 𝑠2 > 1(𝑗 = −1)
𝑉 = − 𝑏2 𝑟2 Θ
𝐿2 𝑑𝜉 2 and
EI ( )
= 𝐴1 𝑒Φ sin Φ𝜉 − 𝐴2 𝑒Φ cos Φξ + 𝑗𝐴3 𝑒Λ sin Λ𝜉 + 𝐴4 𝑒Λ cos Λ𝜉 (50) W1 , W2 and W3 are defined as follows
𝐿3
𝐸𝐼 𝐸𝐼 𝐸𝐼
𝑊1 = ; 𝑊2 = ; 𝑊3 = (61)
𝐿 𝐿2 𝐿3
EI 𝑑Θ EI ( The constants A1 –A4 can now be eliminated from Eqs. (55) and
𝑀 =− =− −𝐴1 Φ𝑘Φ cos Φ𝜉 − 𝐴2 Φ𝑘Φ sin Φ𝜉 + 𝑗𝐴3 Λ𝑘Λ cos Λ𝜉
𝐿 𝑑𝜉 𝐿2 (57) to give the 4 × 4 dynamic stiffness matrix of the Timoshenko beam.
)
+ 𝑗𝐴4 Λ𝑘Λ sin Λ𝜉 (51) This can be achieved by inverting the square matrix of Eq. (55), i.e.
Q matrix of Eq. (56) and pre-multiplying it with the square matrix of
where Eq. (57), i.e. R matrix and performing the matrix operation RQ−1 nu-
( ) ( ) merically to give the dynamic stiffness matrix. Alternatively, the matrix
𝑒 Φ = Φ2 − 𝑏 2 𝑟 2 𝑘 Φ ; 𝑒 Λ = 𝑗 Λ2 + 𝑗 𝑏 2 𝑟 2 𝑘 Λ (52)
inversion and matrix multiplication procedures can be carried out sym-
and j and kΦ , kΛ have already been defined in Eqs. (46) and (49), re- bolically (algebraically) to generate explicit expressions for each of the
spectively. stiffness elements of the dynamic stiffness matrix to give.
Now from the expressions for the amplitudes of displacements W and
Θ given by Eqs. (40)–(43) and the corresponding forces V and M given by ⎡ 𝐹𝑦1 ⎤ ⎡ 𝑑1 𝑑2 𝑑4 𝑑5 ⎤⎡ Δ𝑦1 ⎤
⎢ ⎥ ⎢ ⎥⎢ ⎥
Eqs. (50) and (51), the dynamic stiffness matrix of the Timoshenko beam ⎢ 𝑀1 ⎥ ⎢ 𝑑2 𝑑3 −𝑑5 𝑑6 ⎥⎢ Θ1 ⎥
can be formulated by applying the boundary conditions in algebraic ⎢ ⎥=⎢ ⎥⎢ ⎥ (62)
⎢ 𝐹𝑦2 ⎥ ⎢ 𝑑4 −𝑑5 𝑑1 −𝑑2 ⎥⎢ Δ𝑦2 ⎥
form relating the amplitudes of forces and displacements. ⎢ ⎥ ⎢ ⎥⎢ ⎥
Referring to Fig. 3, the boundary conditions for the displacements ⎣ 𝑀2 ⎦ ⎣ 𝑑5 𝑑6 −𝑑2 𝑑3 ⎦⎣ Θ1 ⎦
and forces can be applied as follows
where
At 𝑥 = 0 (i.e. 𝜉 = 0), 𝑊 = Δ𝑦1 , Θ = Θ1 , 𝑉 = 𝐹y1 and 𝑀 = 𝑀1 ( )
(53) 𝑑1 = 𝑊3 𝑏2 Γ 𝐶 𝑆̄ + 𝜂𝑆 𝐶̄ ∕(ΛΦ) ⎫
{ ( )} ⎪
𝑑2 = 𝑊2 𝑍 Γ (Φ + 𝑗𝜂Λ)𝑆 𝑆̄ − (Λ − 𝜂Φ) 1 − 𝐶 𝐶̄ ∕(Λ + 𝜂Φ)⎪
( )
𝑑3 ̄
= 𝑊1 Γ 𝑆 𝐶 − 𝑗𝜂𝐶 𝑆 ̄ ⎪
At 𝑥 = 𝐿(𝑖.𝑒. 𝜉 = 1), 𝑊 = Δ𝑦2 , Θ = Θ2 , 𝑉 = −𝐹y2 and 𝑀 = −𝑀2 (54) ( ) ⎬ (63)
𝑑4 = −𝑊3 𝑏2 Γ 𝑆̄ + 𝜂𝑆 ∕(ΛΦ) ⎪
( )
Substituting Eqs. (53) and (54) into Eqs. (40)–(43) and Eqs. (50) and 𝑑5 = 𝑊2 𝑍Γ 𝐶̄ − 𝐶 ⎪
( ) ⎪
(51), the following two matrix equations are obtained for displacements 𝑑6 = 𝑊1 Γ 𝑗𝜂 𝑆̄ − 𝑆 ⎭

341
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

Fig. 4. Amplitudes of displacements and forces at the ends of a combined Rayleigh–Love bar and a Timoshenko beam.

with analysed are missed. Essentially the algorithm gives the number of nat-
( ) } ural frequencies that lie below an arbitrarily chosen trial frequency in
𝑍 =Φ− 𝑏2 𝑠2 ∕Φ 𝜂
= Z∕ 𝑗Λ + 𝑏2 𝑠2 ∕Λ ;
{ ( ) ( ) } (64) a straightforward and computationally efficient manner. As successive
Γ = (Λ + ηΦ)∕ 2𝜂 1 − 𝐶 𝐶̄ + 1 − 𝑗 𝜂 2 𝑆 𝑆̄
trial frequencies can be chosen, it is possible to bracket any natural fre-
2.3. Combination of axial and bending stiffnesses quency within any desired accuracy.
Before applying the algorithm the dynamic stiffness matrices of all
A simple superposition is now possible to put the axial and bending individual elements in a structure are to be assembled to form the over-
dynamic stiffnesses together in order to express the force-displacement all dynamic stiffness matrix Kf of the final (complete) structure, which
relationship of the combination of a Rayleigh–Love bar and a Timo- may, of course, consist of a single element. The main features of the
shenko beam. Superposing Figs. 2 and 3 to give Fig. 4 and then using Wittrick–Williams algorithm and its basic working principles are briefly
Eqs. (20) and (62), one obtains the dynamic stiffness matrix of the com- summarised as follows.
bination of a Rayleigh–Love bar and a Timoshenko beam to enable the Suppose that 𝜔 denotes the circular (or angular) frequency of a vi-
free vibration analysis of plane frames to be made. brating structure, then according to the Wittrick–Williams algorithm
Referring to Fig. 4 and Eqs. (20) and (62) the resulting dynamic stiff- [19], j, the number of natural frequencies passed, as 𝜔 is increased from
ness matrix is given by zero to 𝜔 ∗ , is given by
{ }
𝑗 = 𝑗0 + s 𝑲 𝑓 (67)
⎡𝐹𝑥1 ⎤ ⎡𝑎1 0 0 𝑎2 0 0 ⎤⎡Δ𝑥1 ⎤
⎢𝐹𝑦1 ⎥ ⎢0 𝑑1 𝑑2 0 𝑑4 𝑑5 ⎥⎢Δ𝑦1 ⎥ where Kf , the overall dynamic stiffness matrix of the final structure
⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢𝑀1 ⎥ = ⎢0 𝑑2 𝑑3 0 −𝑑5 𝑑6 ⎥⎢𝜃1 ⎥
(65) whose elements all depend on 𝜔, is evaluated at 𝜔 = 𝜔∗ ; s{Kf } is the num-
⎢𝐹𝑥2 ⎥ ⎢𝑎2 0 0 𝑎1 0 0 ⎥ ⎢ Δ𝑥2 ⎥ ber of negative elements on the leading diagonal of Kf Δ , Kf Δ is the upper
⎢𝐹 ⎥ ⎢0 𝑑4 −𝑑5 0 𝑑1 −𝑑2 ⎥⎥⎢⎢Δ𝑦2 ⎥⎥
⎢ 𝑦2 ⎥ ⎢ triangular matrix obtained by applying the usual form of Gauss elimina-
⎣𝑀2 ⎦ ⎣0 𝑑5 𝑑6 0 −𝑑2 𝑑3 ⎦⎣𝜃2 ⎦ tion to Kf , and j0 is the number of natural frequencies of the structure
or still lying between 𝜔=0 and 𝜔 = 𝜔∗ when the displacement components
to which Kf corresponds are all zeros. (Note that the structure can still
𝐅 = 𝐊𝚫 (66)
have natural frequencies when all its nodes are clamped, because exact
where F and 𝚫 are respectively the force and displacement vectors and member equations allow each individual member to displace between
K is the frequency dependent 6 × 6 dynamic stiffness matrix whose el- nodes with an infinite number of degrees of freedom, and hence infinite
ements k(i, j) (i = 1,2…6; j = 1,2,…6) are given by a1 , a2 and d1 –d6 de- number of natural frequencies between nodes.) Thus
fined in Eqs. (21) and (63), respectively. Note that K is symmetric as ∑
𝑗0 = 𝑗𝑚 (68)
expected.
where jm is the number of natural frequencies between 𝜔=0 and 𝜔 = 𝜔∗
3. Application of the dynamic stiffness matrix for a component member with its ends fully clamped, while the summa-
tion extends over all members of the structure. Thus, with the knowledge
3.1. The Wittrick–Williams algorithm of Eqs. (67) and (68), it is possible to ascertain how many natural fre-
quencies of a structure lie below an arbitrarily chosen trial frequency.
The developed dynamic stiffness matrix can now be used to com- This simple feature of the algorithm (coupled with the fact that suc-
pute the natural frequencies and mode shapes of either an individ- cessive trial frequencies can be chosen by the user to bracket a natural
ual Rayleigh–Love bar or a Timoshenko beam or a framework com- frequency) can be used to converge on any required natural frequency
prising beam elements whose axial stiffnesses are characterized by the to any desired (or specified) accuracy.
Rayleigh–Love bar theory and the bending stiffnesses are characterised
by the Timoshenko beam theory. The assembly procedure to obtain the 3.2. The significance of the j0 count in the Wittrick–Williams algorithm
overall dynamic stiffness matrix of the final structure in the global or
datum coordinates is similar to the finite element method. For a frame, As explained in Section 3.1, one of the requirements for the appli-
the standard procedure to create the transformation matrix comprising cation of the Wittrick–Williams algorithm is to acquire the needed in-
the sine and cosine of the angle made by the global and local x-axis of formation about the clamped–clamped natural frequencies of individ-
each individual bar or beam element are used to assemble the overall ual elements in a structures (the so-called j0 count) so as to enable the
dynamic stiffness matrix of the frame. A reliable and accurate method free vibration analysis to be carried out in a flawless and robust man-
of solving the free vibration problem is to apply the Wittrick–Williams ner. However, the determination of the natural frequencies using the
algorithm [19] which is well suited for the dynamic stiffness method. Wittrick–Williams algorithm is predominantly based on the sign count
The algorithm uses the Sturm sequence property of the dynamic stiff- s{Kf } described in Section 3.1. The j0 count of Eq. (68) is not always
ness matrix and ensures that no natural frequencies of the structure needed, particularly if the clamped-clamped natural frequency of none

342
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

of the constituent members in the structure is exceeded within the fre- analysed using both the classical Bernoulli Euler and the Rayleigh–Love
quency range of interest. One way of avoiding the computation of j0 is theories. Finally example-3 demonstrates the free vibration character-
to split the structure into large number of elements so that the clamped- istics of a plane frame for which the dynamic stiffness matrix for each
clamped natural frequencies of all individual elements become excep- constituent element is based on both Rayleigh–Love and Timoshenko
tionally high and thus will not be exceeded by any frequency of practical theories as well as classical Bernoulli Euler theories.
interest. Nevertheless, j0 count of the algorithm is not really a peripheral
issue, particularly for achieving computational efficiency and avoiding 4.1. Free longitudinal vibration of a uniform bar
further unnecessary discretisation of the structure. The need to compute
j0 stems from the fact that the DSM allows infinite number of natural Using the notations given in Section 2.1, the natural frequencies
frequencies to be accounted for when all the nodes of the structure are of a Rayleigh–Love bar with both ends clamped can be obtained from
fully restrained and yet one or more structural members can vibrate Eq. (14) by substituting U(𝜉) to zero at both 𝜉 = 0 and 𝜉 = 1 and making
freely on their own between the nodes resulting in 𝜹 =0 modes in the appropriate substitution for 𝛾 to give the nth natural frequency 𝜔n as
eigenvalue equation [KD ]{𝜹} = 0. √ ( )

√ 𝑛2 𝜋 2 𝐸𝐴
𝜔𝑛 = √√( ) (75)
3.3. Clamped–clamped natural frequencies of a Rayleigh–Love bar √ 𝜈 2 𝐼 𝑛2 𝜋 2 𝜌𝐴𝐿2
1 + 𝐴𝑝𝐿2

The clamped–clamped natural frequencies of a Rayleigh–Love bar where n = 1, 2, 3, ….


can be obtained from Eq. (14) by substituting the boundary conditions The corresponding natural frequencies for the classical Bernoulli–
of displacements to zero at both ends or alternatively by putting the de- Euler bar with clamped–clamped boundary conditions can be found in
terminant of the square matrix of Eq. (18) to zero, yielding the frequency standard texts [7] given by
equation as √ ( )
𝜔𝑛0 = 𝑛𝜋 𝐸𝐴∕ 𝜌𝐴𝐿2 (76)
sin 𝛾 = 0 = sin 𝑛𝜋 (69)
The ratio between the natural frequencies for the clamped–clamped
Thus, proceeding in the same way as in the case of classical
bar obtained from the Rayleigh–Love and classical Bernoulli–Euler the-
Bernoulli–Euler bar [16] the number of clamped–clamped natural fre-
ories can be expressed with the help of Eqs. (75) and (76) to give
quencies jR of a Rayleigh–Love bar lying below an arbitrarily chosen
trial frequency 𝜔 ∗ is given by 𝜔𝑛 1
= √ (77)
𝛾 𝜔𝑛0
1 + 𝜈( 𝑛 )𝜋2
2 2 2
𝑗𝑅 = highest integer < (70) 𝐿
𝜋 𝑟

3.4. Clamped–clamped natural frequencies of a Timoshenko beam where r is defined as the radius of gyration expressed as

𝐼𝑝
For a Timoshenko beam, the number of clamped–clamped natural 𝑟= (78)
frequencies exceeded by the trial frequency 𝜔 ∗ can be established using 𝐴
the procedure described in [17] to give Proceeding in a similar way and imposing appropriate boundary
[ { }] conditions, the natural frequency ratio for a cantilever bar using the
{ } 𝑑2
𝑗𝑇 = 𝑗𝑐 − 2 − 𝑠𝑔 𝑑3 − 𝑠𝑔 𝑑3 − 6 ∕2 (71) Rayleigh–Love and classical Bernoulli–Euler theories can be expressed
𝑑3
as
where sg{ } is +1 or −1 depending on the sign of the quantity within 𝜔𝑛 1
= √ (79)
the curly bracket, d3 and d6 have already been defined in Eq. (63) and 𝜔𝑛0 (2𝑛−1)2 𝜋 2 𝜈 2
1+ ( )2
jc is given by 4 𝐿𝑟
}
𝑗c = 𝑗d for 𝑏2 𝑟2 𝑠2 < 1 The validity of the Eqs. (77) and (79) has been further confirmed
(72)
𝑗c = 𝑗d + 𝑗e for 𝑏 𝑟 𝑠 ≥ 1
2 2 2
by using the developed dynamic stiffness matrix of a Rayleigh–Love bar
with shown in Eq. (20).
} Clearly Eqs. (77) and (79) indicate that the natural frequency ratio
𝑗d = highest int eger < Φ 𝜔𝑛
𝜋 (73) 𝜔
is dependent on the Poisson’s ratio 𝜈 of the bar material as well as
𝑗e = highest int eger < Λ𝜋 + 1 𝑛0
the slenderness ratio L/r of the bar. The Poisson’s ratio 𝜈 for an isotropic
In Eq. (73), Φ and Λ have already been defined in Eqs. (44) and material is generally constant and maybe assumed to be 0.3 which is
(45) respectively. Thus the number of clamped–clamped natural fre- used here in the analysis.
quencies jm exceeded by an individual member by the trial frequency 𝜔∗ Figs. 5 and 6 show the variation of the ratio of the first five natural
with the inclusion of the Rayleigh–Love bar and the Timoshenko beam frequencies using the Rayleigh–Love and classical Bernoulli–Euler theo-
theories is given by ries against the slenderness ratio L/r for the clamped–clamped and can-
tilever bar respectively. Clearly for smaller values of slenderness ratios
𝑗m = 𝑗R + 𝑗T (74)
and for higher natural frequencies, the classical Bernoulli–Euler theory
Now the root count j0 of Eq. (67) can be computed using the is considerably less accurate. The errors incurred in the fifth natural
Eq. (68) where the summation Σ over m is extended to include all ele- frequency when using the classical Bernoulli–Euler theory are 27% and
ments in the structure. 24% for the clamped–clamped and cantilever bar respectively when the
slenderness ratio is 5. It should be noted that in the Statistical Energy
4. Results and discussion Analysis (SEA) for which modal density in the high frequency range is
required, the classical Bernoulli–Euler theory can be inadequate.
Numerical examples are given for three different types of problems.
Example-1 is focused on the natural frequencies of a freely vibrating uni- 4.2. Free longitudinal vibration of a stepped bar
form Rayleigh–Love bar in longitudinal motion with clamped–clamped
and cantilever boundary conditions. This is followed by example-2 A stepped bar (example-2) which is taken from [34] and shown in
which is that of a stepped bar taken from the literature. This problem is Fig. 7 is analysed for its free vibration characteristics in longitudinal

343
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

Fig. 5. The first five natural frequency ratios using the Rayleigh–Love and classical Bernoulli–Euler theories for a clamped-clamped bar. 𝜔n = natural frequency using
Rayleigh–Love theory; 𝜔n0 = natural frequency using classical Bernoulli–Euler theory.

Fig. 6. The first five natural frequency ratios using the Rayleigh–Love and classical Bernoulli–Euler theories for a cantilever bar. 𝜔n = natural frequency using
Rayleigh–Love theory; 𝜔n0 = natural frequency using classical Bernoulli–Euler theory.

Table 1
Natural frequencies of a stepped bar in longitudinal vi-
bration (results from the conventional classical theory
are shown in the parenthesis in column 2).

Frequency number Natural frequency (Hz)


Current theory Ref. [34]

1 1184.312 (1184.39) 1362.79


2 11,732.86 (12,509.42) 11,679.6
3 14,503.42 (15,002.56) 12,640.5
4 20,014.45 (24,187.29) 19,461.9

The numerical values for the data taken from [34] are:
Fig. 7. A three-stepped bar for free vibration analysis.
r1 = 0.05 m, r2 = 0.03 m, r3 = 0.075 m,
l1 = 0.05 m, l2 = 0.17 m, l3 = 0.13 m,
E1 = 200 × 109 Pa, E2 = 70 × 109 Pa, E3 = 100 × 109 Pa,
motion using the developed dynamic stiffness matrix. The stepped bar 𝜌1 = 7.85 × 103 kg/m3 , 𝜌2 = 2.7 × 103 kg/m3 , 𝜌3 = 8.4 × 103 kg/m3 ,
is cantilevered at the left hand end as shown and consists of three indi- 𝜈 1 = 0.30, 𝜈 2 = 0.33, 𝜈 3 = 0.34.
vidual bars of solid circular cross-section with different geometrical di-
mensions and material properties for each. The essential data required The first four natural frequencies computed using the Rayleigh–Love
for the analysis are: radius of cross-section (ri ), length (li ), Young’s mod- dynamic stiffness theory are shown in column 2 of Table 1 alongside the
ulus (Ei ), density (𝜌i ) and Poisson’s ratio (𝜈 i ) (i representing the segment results reported in [34] shown in column 3. The corresponding natural
or element number). frequencies computed using classical Bernoulli–Euler dynamic stiffness

344
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

[34] is different from the dynamic stiffness methodology developed in


this paper. It is to be noted that both the Rayleigh–Love and the classi-
cal Bernoulli–Euler theories give almost the same results for the funda-
mental natural frequency, but the differences in the second, third and
fourth frequencies are 7%, 4% and 21% respectively. Understandably,
the classical Bernoulli–Euler theory overestimates the natural frequen-
cies whereas the more refined Rayleigh–Love theory which accounts for
the added transverse and lateral inertia of the bar yields lower values of
the natural frequencies which is apparently contradicted by the result
for the fundamental natural frequency reported in [34].

4.3. Free vibration of a plane frame

The final set of results was obtained using example-3 which is that
of a plane frame shown in Fig. 9. Each element of the frame has the
same uniform geometrical, cross sectional and material properties and
the data used in the analysis are as follows:

𝐸𝐼 = 4.0 × 106 Nm2 , 𝐸𝐴 = 8.0 × 108 N, 𝑘𝐴𝐺 = 2.0 × 108 N,

𝜌𝐴 = 30 kg∕m, 𝜌𝐼𝑝 = 0.157 kgm, 𝜈 = 1∕3, 𝑘 = 2∕3

A wide range of the natural frequencies of the frame was computed


using the present theory as well as the classical Bernoulli–Euler theory.
Apart from the computation of the first five natural frequencies which
were sequentially chosen, the higher order natural frequencies were
sparingly and sparsely chosen so as to cover the order of the natural
Fig. 8. Natural frequencies and mode shapes of the three-stepped bar of Fig 7.
frequencies between 50th and 400th. The results are shown in Table 2.
Clearly, higher the order of the frequency, higher the incurred error due
to using the classical Bernoulli–Euler theory. The first five natural fre-
theory [16] are also shown in the parenthesis in column 2. Although quencies of the frame are virtually unaltered. As expected, the classical
the agreement of the results between the present theory and those of Bernoulli–Euler theory overestimates the natural frequencies.
[34] is good for the second and fourth natural frequencies (the differ- One of the potential application areas of the theory developed in this
ences are well within 3%), but for the first and third natural frequencies paper is the Statistical Energy Analysis (SEA) for which accurate natu-
there are some discrepancies which are around 13% and 15% respec- ral frequency predictions in the low, medium and high frequency range
tively. The fundamental natural frequency of the bar quoted in [34] is are essential. To this end, the uncompromising accuracy of the dynamic
well above the corresponding natural frequency obtained from the clas- stiffness method developed in this paper by applying the Rayleigh–
sical Bernoulli–Euler theory. This is surely in error because the effect Love and Timoshenko theories is further demonstrated by computing
of the transverse inertia presumably accounted for in [34] is expected the number of natural frequencies of the frame (see Fig. 10) which lies
to diminish the natural frequency and not increase it. The mode shapes within the frequency ranges of 0 < fi ≤ 1kHz, 0 < fi ≤ 2kHz, 0 < fi ≤
corresponding to the four natural frequencies using the present theory 3kHz and up to 0 < fi ≤ 10kHz which cover low, medium and high fre-
are shown in Fig. 8 by solid lines which are in broad agreement with quency bands. Fig. 10 shows the frequency distribution, i.e. the modal
the ones reported in [34]. The mode shapes shown by dash lines are density of the frame. It will be difficult to obtain these results with such
those computed using the classical Bernoulli–Euler theory. Clearly, the accuracy using conventional Finite Element Method.
first three mode shapes have undergone very little changes as a result
of using the present theory as opposed to the classical Bernoulli–Euler 5. Limitations and scope for further developments of the theory
theory, but the fourth mode being a higher order mode has turned out
to be significantly different, as expected. The authors were unable to The dynamic stiffness theory presented in this paper deals with
pinpoint the exact reason for the discrepancies in the first and third nat- Rayleigh–Love bars and Timoshenko beams made of homogenous and
ural frequencies, when they compared their results with those of [34], isotropic materials for which the essential properties are Young’s mod-
but it should be recognised that the series solution approach used in ulus (E), shear modulus (G), Poisson’s ratio (𝜈) and density (𝜌). Further

Fig. 9. A plane frame for free vibration analysis using Rayleigh–Love and Timoshenko theories.

345
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

Table 2
Natural frequencies of plane frame.

Frequency range Natural frequency number (i) Natural frequency fi (Hz)


Rayleigh–Love and Timoshenko theory Classical Bernoulli–Euler theory

Low 1 35.38 35.77


2 38.56 39.10
3 41.98 42.56
4 50.73 51.39
5 53.14 53.94
Medium 50 565.54 600.97
60 635.28 709.09
70 828.95 934.39
80 964.48 1136.00
90 1108.90 1301.20
100 1306.80 1521.40
High 150 2151.10 2732.30
200 3047.50 4089.50
250 3940.30 5585.00
300 4859.70 7152.20
350 5767.40 8744.80
400 6112.80 10,495.00

Fig. 10. Modal density of plane frame.

development of the theory and its future applications to include lami- both the Rayleigh–Love and Timoshenko theories has been developed.
nated composites will be a challenge, and indeed, a major task mainly With the help of the Wittrick–Williams algorithm as solution technique,
because of the fibrous nature of such anisotropic materials, which re- the theory is applied to investigate the free vibration behaviour of a
quire more elastic constants to define their properties. Also the introduc- uniform Rayleigh–Love bar, a stepped Rayleigh–Love bar, and a frame-
tion of a fictitious shape factor or shear correction factor as demanded work for which the modal density distribution is presented by capturing
by the assumption in the Timoshenko beam theory to account for the its natural frequencies in the low, medium and high frequency range.
non-uniform shear stress distribution through the thickness of the beam Some representative mode shapes of the stepped bar are also illustrated.
cross-section, is no-doubt a limitation. In this respect, the current the- The theory developed is particularly helpful when carrying out high fre-
ory can be extended by incorporating higher order shear deformation quency free vibration analysis of skeletal structures. A potential appli-
theories [35,36] in the analysis. To overcome the limitations of the the- cation of the research described in this paper falls within the area of
ory presented in this paper, particularly for its extension to anisotropic statistical energy analysis for which the knowledge of modal density
fibrous composites, interested readers are referred to the review papers distributions in the high frequency range is essential.
of Sayyad and Ghugal [37,38] for necessary background information.

Acknowledgements
6. Conclusions
The authors are grateful to Dr Alfonso Pagani of the Polytechnic Uni-
Starting from the derivations of the governing differential equations versity of Turin, Italy and Dr William Tiu of BA Systems, UK for many
of motion in free vibration, the dynamic stiffness matrix of a beam using stimulating email exchanges on the subject. Also the first author benefit-

346
J.R. Banerjee, A. Ananthapuvirajah International Journal of Mechanical Sciences 150 (2019) 337–347

ted from a related past EPSRC (UK) project (Grant Ref GR/R21875/01) [20] Han JB, Hong SY, Song JH, Kwon HW. Vibrational energy flow models for
to whom acknowledgement is hereby made. the Rayleigh–Love and Rayleigh–Bishop rods. J Sound Vib 2014;333:520–40.
doi:10.1016/j.jsv.2013.08.027.
[21] Shatalov M, Marais J, Fedotov I, Tenkam MJ. Longitudinal vibration of isotropic
References solid rods: from classical to modern theories. Adv Comput Sci Engg 2011;333.
doi:10.5772/15662.
[1] Keane AJ, Price WC. Statistical energy analysis: an overview, with applications in [22] Rayleigh Lord. The theory of sound: vol. 1-2. Cambridge University Press; 2011.
structural dynamics. Cambridge University Press; 1997. [23] Love AEH. A treatise on the mathematical theory of elasticity. 4th ed. Cambridge
[2] Lyon RH, DeJong RG. Theory and application of statistical energy analysis. 2nd ed. University Press; 2013.
London: Butterworth-Heinermann; 1995. [24] Belov VD, Rybak SA, Tartakovskii BD. Propagation of vibrational energy in absorb-
[3] Wohlever JC, Bernhard RJ. Mechanical energy flow models of rods and beams. J ing structures. J Soviet Phys Acoust 1977;23:115–19.
Sound Vib 1992;153:1–19. doi:10.1016/0022-460X(92)90623-6. [25] Nefske DJ, Sung SH. Power flow finite element analysis of dynamic systems: basic
[4] Lase Y, Ichchou MN, Jezequel L. Energy flow analysis of bars and beams: theoretical theory and application to beams. J Vib Acoust Stress Reliab Des 1989;111:94–100.
formulations. J Sound Vib 1996;192:281–305. doi:10.1006/jsvi.1996.0188. doi:10.1115/1.3269830.
[5] Bouthier OM, Bernhard RJ, Laboratories H, Lafayette W. Simple models [26] Lase Y, Ichchou MN, Jezequel L. Energy flow analysis of bars and beams: theoretical
of energy flow in vibrating membranes. J Sound Vib 1995;182:129–47. formulations. J Sound Vib 1996;192:281–305.
doi:10.1006/jsvi.1995.0186. [27] Cheng FY. Vibration of Timoshenko beams and frameworks. J Struct Div ASCE
[6] Bouthier OM, Bernhard RJ. Simple models of the energetics of transversely vibrating 1970;96:551–71.
plates. J Sound Vib 1995;182:149–64. doi:10.1006/jsvi.1995.0187. [28] Wang TM, Kinsman TA. Vibration of frame structures according to the Timoshenko
[7] Rao SS. Vibration of continuous systems. 1st ed. John Wiley & Sons; 2007. theory. J Sound Vib 1971;14:215–27.
[8] Leissa AW, Qatu MS. Vibration of continuous systems. 1st ed. McGraw-Hill Educa- [29] Howson WP, Williams FW. Natural frequencies of frames with axially loaded Timo-
tion; 2011. shenko members. J Sound Vib 1973;26:503–15.
[9] Meirovitch L. Fundamentals of vibrations. 1st ed. Waveland Pr Inc; 2010. [30] Cheng FY, Tseng WH. Dynamic matrix of Timoshenko beam columns. J Struct Div,
[10] Thomson WT. Theory of vibration with applications. 4th ed. CRC Press; 1996. ASCE 1973;99:527–49.
[11] Chopra AK. Dynamics of structures: theory and applications to earthquake engineer- [31] Banerjee JR, Sobey AJ, Su H, Fitch JP. Use of computer algebra in Hamiltonian
ing. 4th ed. Pearson; 2015. calculations. Adv Eng Softw 2008;39(6):521–5.
[12] Koloušek V. Anwendung des gesetzes der virtuellen verschiebungen und des [32] Predoi MV, Petre CC, Vasile O, Boiangiu M. High frequency longitudinal
reziprozitätssatzes in der stabwerksdynamik. Ingenieur-Archiv 1941;12:363–70. damped vibrations of a cylindrical ultrasonic transducer. Shock Vib 2014.
doi:10.1007/BF02089894. doi:10.1155/2014/105971.
[13] Koloušek V. Berechnung der schwingenden stockwerkrahmen nach der deformation- [33] Hurty WC, Rubinstein MF. Dynamics of structures. Prentice Hall; 1964.
smethode. Der Stahlbau 1943;16:11–13. [34] Fedotov I, Gai Y, Polyanin A, Shatalov M. Analysis for an N-stepped Rayleigh
[14] Koloušek V. Dynamics in engineering structures. Newnes-Butterworth; 1973. bar with sections of complex geometry. Appl Math Model 2008;32:1–11.
[15] Åkesson B. PFVIBAT—a computer program for plane frame vibration anal- doi:10.1016/j.apm.2006.10.028.
ysis by an exact method. Int J Numer Methods Eng 1976;10:1221–31. [35] Heyliger PR, Reddy JN. A higher order beam finite element for bending and vibration
doi:10.1002/nme.1620100603. problems. J Sound Vib 1988;126:309–26. doi:10.1016/0022-460X(88)90244-1.
[16] Williams FW, Howson WP. Compact computation of natural frequencies and [36] Eisenberger M. Dynamic stiffness vibration analysis using a high-order beam model.
buckling loads for plane frames. Int J Numer Methods Eng 1977;11:1067–81. Int J Numer Methods Eng 2003;57:1603–14. doi:10.1002/nme.736.
doi:10.1002/nme.1620110704. [37] Sayyad AS, Ghugal YM. On the free vibration analysis of laminated composite and
[17] Howson WP, Banerjee JR, Williams FW. Concise equations and program for sandwich plates: a review of recent literature with some numerical results. Compos
exact eigensolution of plane frames including member shear. Adv Eng Softw Struct 2015;129:177–201. doi:10.1016/j.compstruct.2015.04.007.
1983;5:137–41. [38] Sayyad AS, Ghugal YM. Bending, buckling and free vibration of laminated composite
[18] Banerjee JR. Dynamic stiffness formulation for structural elements: a general ap- and sandwich beams: a critical review of literature. Compos Struct 2017;171:486–
proach. Comput Struct 1997;63:101–3. doi:10.1016/S0045-7949(96)00326-4. 504. doi:10.1016/j.compstruct.2017.03.053.
[19] Wittrick WH, Williams FW. A general algorithm for computing natural frequencies
of elastic structures. Q J Mech Appl Math 1971;24:263–84.

347

You might also like