Analysis of The Effects of Laminated Depth and Material Properties On The Damping Associated With Layered Structures in A Pressurized Environment

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/275207660

Analysis of the Effects of Laminated depth and Material Properties on the


Damping Associated with Layered Structures in a Pressurized Environment

Article in Transactions of the Canadian Society for Mechanical Engineering · June 2010
DOI: 10.1139/tcsme-2010-0011

CITATIONS READS

7 152

4 authors:

Vincent O. S. Olunloyo Olatunde Damisa


University of Lagos Federal University of Petroleum Resources
55 PUBLICATIONS 425 CITATIONS 21 PUBLICATIONS 192 CITATIONS

SEE PROFILE SEE PROFILE

Charles Osheku A.A. Oyediran


National Space Research and Development Agency University of Lagos
71 PUBLICATIONS 330 CITATIONS 48 PUBLICATIONS 304 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Charles Osheku on 22 April 2015.

The user has requested enhancement of the downloaded file.


ANALYSIS OF THE EFFECTS OF LAMINATE DEPTH AND MATERIAL
PROPERTIES ON THE DAMPING ASSOCIATED WITH LAYERED STRUCTURES
IN A PRESSURIZED ENVIRONMENT
Vincent O. S. Olunloyo1, Olatunde Damisa2, Charles A. Osheku2, Ayo A. Oyediran3
1
Department of Systems Engineering, Faculty of Engineering, University of Lagos
2
Department of Mechanical Engineering, Faculty of Engineering, University of Lagos
3
AYT Research Corp., McLean VA. 22102, USA
E-mail: vosolunloyo@hotmail.com
Received September 2008, Accepted February 14, 2009
No. 06-CSME-10, E.I.C. Accession 2929

ABSTRACT
In aerodynamic and machine structures, one of the effective ways of dissipating unwanted
vibration or noise is to exploit the occurrence of slip at the interface of structural laminates where
such members are held together in a pressurised environment. The analysis and experimental
investigation of such laminates have established that when subjected to either static or dynamic
loading, non-uniformity in interface pressure can have significant effect on both the energy
dissipation and the logarithmic damping decrement associated with the mechanism of slip damping.
Such behaviour can in fact be effectively exploited to increase the level of damping available in such
a mechanism. What has however not been examined is to what extent is the energy dissipation
affected by the relative sizes or the material properties of the upper and lower laminates? In this
paper the analysis is extended to incorporate such effects. In particular, by invoking operational
methods, it is shown that variation in laminate thickness may provide less efficacious means of
maximizing energy dissipation than varying the choice of laminate materials but that either of these
effects can in fact dwarf those associated with non-uniformity in interface pressure. To achieve this,
a special configuration is required for the relative sizes and layering of the laminates. In particular, it
is shown that for the case of two laminates, the upper laminate is required to be thinner and harder
than the lower one. These results provide a basis for the design of such structures.

ANALYSE DES EFFETS DES PROFONDEURS LAMINÉES ET DES PROPRIÉTÉS


MATÉRIELLES SUR LE DÉCRÉMENT ASSOCIÉS À DES COUCHES
STRUCTURELLES DANS UN ENVIRONNEMENT PRESSURISÉ
RÉSUMÉ
Dans l’aérodynamique et la structure des machines, une des façons efficaces de dissiper les
vibrations ou les bruits non désirés est d’exploiter la présence du glissement au niveau de l’interface
laminée où de tels éléments sont tenus ensemble dans un environnement pressurisé. L’analyse et la
recherche expérimentale de cette lamination a établi que, assujettie à un chargement statique
ou dynamique, la non uniformité dans l’interface pressurisée peut avoir un effet significatif, à la fois,
sur la dissipation de l’énergie et le décrément d’amortissement logarithmique associé avec le
mécanisme de glissement du décrément. Un tel événement peut être utilisé efficacement pour
augmenter le niveau de décrément disponible dans un tel mécanisme. Ce qui n’a pas été étudié est
l’ampleur de la dissipation de l’énergie utilisée par les parties supérieures ou inférieures laminées.
Dans cet article, l’analyse s’est appliquée à introduire de tels effets. En particulier, en utilisant ces
méthodes, on démontre que les variations dans les épaisseurs laminées pourraient donner des
moyens moins efficaces dans la maximisation de la dissipation de l’énergie. Une configuration
spéciale est requise pour des tailles spécifiques et les couches laminées. En particulier, il est démontré
que dans le cas de deux couches laminées, la couche supérieure doit être plus mince et dure que la
couche inférieure. Ces résultats nous procurent une base pour le design de telles structures.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 165
Nomenclature
b width of laminated beams x space coordinate along the beam interface
differential operator y space coordinate perpendicular to the
E1 modulus of rigidity of lower laminate beam interface
E2 modulus of rigidity of upper laminate Greek letters
F applied end force amplitude ratio of Young’s moduli of laminates
h1 depth of lower laminate beam m coefficient of friction at the interface
h2 depth of upper laminate beam of the laminated beams
I1 moment of inertia for lower laminate j dummy variable
I2 moment of inertia for upper laminate r1 density of lower laminate material
L length of Laminated beams r2 density of upper laminate material
P clamping pressure at the interface of (sx)1 bending stress at the lower half of the
the laminated beams laminates
t time coordinate (sx)2 bending stress at the upper half of the
u1 displacement of the lower laminate laminates
u2 displacement of the upper laminate txz shear stress at the interface of the
W dynamic response laminated beams
WF transverse response in Fourier plane Y ratio of the laminate thicknesses

1. INTRODUCTION
The mechanism of damping as a means of controlling undesirable effects of vibration has
received considerable attention in the literature over the years. Within this context, slip damping
is a mechanism exploited for dissipating noise and vibration energy in aerodynamic and
machine structures. There are in fact, several ways of effecting such damping; including the
introduction of either constrained, unconstrained and even viscoelastic layers. One of such
techniques is layered construction made possible by externally applied pressure that holds the
members together at the interface. Such layers could also either be jointed or held together by
appropriately spaced bolts. Under such circumstances, the profile of the interface pressure
assumes a significant role, especially in the presence of slip, to dissipate the vibration energy. A
full account of the nature of the pressure profile can be found in [1–3] where the work of Gould
and Mikic [4] and Ziada and Abd [5] as well as Nanda and Behera [6] are discussed in further
details.
Right from the time of Goodman and Klumpp [7] who were credited with one of the earliest
works on slip damping, the emphasis has usually revolved around the maximum amount of
energy dissipation that could be arranged and for the case of dynamic loading, what level of
logarithmic damping decrement could be achieved. However all the early workers including
Masuko et al [8], Nishiwaki et al [9, 10] and Motosh [11] limited their investigations to the case
of uniform or constant intensity of pressure distribution at the interface.
More recently, there have been attempts to relax the restriction of uniform interface pressure
to allow for more realistic pressure profiles that are encountered in practice. Such attempts
include both experimental and numerical treatments such as the work of Shin et al [12], Song
et al [13], Nanda [14, 15] as well as Nanda and Behera [16, 17].
The analytical analysis of the effect of non-uniform interface pressure distribution on the
mechanism of slip damping for layered beams was also recently examined for both static and
dynamic loads. In particular, whereas the investigation in Damisa et al [1, 3] was limited to the

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 166
case of linear pressure profile, the static analysis in Olunloyo et al [2] included other forms of
interfacial pressure distributions such as polynomial or hyperbolic representations and the
results obtained demonstrated that the effects of such distributions in comparison with the
linear profile were largely incremental in nature and no fundamental differences were found.
This provides additional justification for the linear pressure profile selected for the cantilever
architecture used in our present investigation.
The results of the analysis of the cantibeam in [1–3] revealed that when the beam laminates
are of the same material and thickness, non-uniformity in interface pressure can for example
have significant effect on the mechanism of slip damping for static load while the energy
dissipation and the logarithmic damping decrement associated with dynamic loads are
significantly influenced by the nature of the interfacial pressure profile between the laminates.
What has not been studied is the effect of asymmetry either in the dimensions of the laminates
or in the choice of materials for the upper and lower laminates.
The aim of the present work is to extend earlier dynamic analysis to cover the case where the
upper and lower laminates need not be of the same dimensions, neither do they need to be of the
same material so as to accommodate the use of composites and study the additional effects that
might arise in the context of energy dissipation and logarithmic damping decrement.

2.1. Governing Differential Equation


For the case of a layered beam of two dissimilar materials and laminate thicknesses, Osheku
[18] has shown that the governing equations of motion corresponding to microscopic slip at the
interface can be derived as:
  a za  dp
L4 W b1 zb2 L2 W 1 2
z ~ ð1Þ
Lx4 2 Lt2 2 dx

where the following have been defined


6m r bh1
a1 ~ 2
, b1 ~ 1
E1 h1 E 1 I1

6m r bh2
a2 ~ 2
, b2 ~ 2
E2 h2 E 2 I2

2.2. Problem Definition and Method of Analysis


One of the principal reasons for this investigation is to determine to what extent the structure
illustrated in Fig. 1 below can serve as an energy dissipating mechanism. In particular, the

Fig. 1. Coordinate axes and geometry for layered beam of dissimilar laminates under dynamic load.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 167
objective here, is to examine analytically the effect of the nature of load, laminate depth ratio
and material properties variation on:
(i) the dynamic response of a clamped layered beam made from dissimilar materials and held
together by some externally applied non-uniform force;
(ii) the profile of interfacial slip;
(iii) the slip energy dissipation under dynamic conditions;
(iv) logarithmic damping decrement associated with mechanism of slip damping in such
layered structures.
A general theory of the energy dissipation properties of press-fit-joints in the presence of
coulomb friction as originally developed by Goodman and Klumpp provides the basis for the
physics of the problem. The contact conditions between the two layers are:
(i) there is continuity of stress distributions at the interface to sufficiently hold the two layers
together both in the pre- and post- slip conditions.
(ii) a stick elastic slip with presence of coulomb friction occurs at the interface of the sandwich
elastic beams to dissipate energy and does not remain constant as a function of some other
variable such as spatial distance, time or velocity.

3. ANALYSIS OF DYNAMIC RESPONSE FOR LINEAR INTERFACE PRESSURE


PROFILE
When we take the Laplace transform of the governing differential Eq. (1), we obtain
*    a za  dP
d 4 W ðx,sÞ b1 zb2 2*
. 1 2
z s W W ðx,sÞ{sW ð0Þ{ W ð0Þ ~ ð2Þ
dx4 2 2s dx

where the Laplace transform viz:

ð
? ð
gzi?
1
ð~.Þ ~ ð.Þe{st dt , ð.Þ ~ ð~.Þest ds ð2aÞ
2pi
0 g{i?

has been invoked.


If the corresponding analysis is also limited to the case of linear pressure variation at the
interface viz:
 e 
P ðx Þ ~ P0 1z x ð2bÞ
L

then substitution for the pressure in Eq. (2) gives

*    a za  P e
d 4W ðx,sÞ b1 zb2 2*
. 1 2 0
z s W ðx,sÞ{sW ð0Þ{ W ð0Þ ~ ð3Þ
dx4 2 2s L

the next step is to introduce the Fourier finite sine transform namely:

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 168
ðL
F npx 2X ?
npx
½. ~ ½.sin dx ; ½. ~ ½.F sin ð4Þ
L L n~1 L
0

from which the following relation can be inferred viz:

* F n 4 p 4 * F n 3 p 3 * 
nz1 *
W xxxx ðx,sÞ ~ W ð x,s Þ { W ð 0,s Þzð {1Þ W ð L,s Þ
L4 L3 ð5Þ

The operating boundary conditions at the ends of the beam in the Laplace transform plane are:

* d * d2 *
W ð0,sÞ ~ W ð0,sÞ ~ W ðL,sÞ ~ 0 ð6Þ
dx dx2

and use of the first and third conditions in the preceding equation reduce Eq. (5) to

* F n4 p4 * F
W xxxx ðx,sÞ ~ W ð x,s Þ
L4
ð7Þ
n3 p3 * np *
{ 3 ð{1Þnz1W ðL,sÞz W xx ð0,sÞ
L L

so that on assuming zero initial conditions for W, the Fourier sine transform of Eq. (3) gives the
result
  a za  P e  
n4 p4 *F b1 zb2 2 *F 1 2 0
W ðln ,sÞz s W ðln ,sÞ ~ 1zð{1Þnz1
L4 2 2s np
ð8Þ
n3 p3 * np *
z 3 ð{1Þnz1W ðL,sÞ{ W xx ð0,sÞ
L L *
To further simplify Eq. (8), one can proceed to evaluate the term Wxx ð0,sÞ by applying the
Goodman and Klumpp end condition in the spatial-state form as
0 1
ð0 hð2
B C F ðtÞ
@ tðxzÞ1 ðx,tÞz tðxzÞ2 ðx,tÞAdz ~ at x~L ð9Þ
b
{h1 0

so that using the out of plane shear stress relations namely

1  2  mp
tðxzÞ1 ðx,tÞ ~ E1 z zh1 z W1xxx ðx,tÞz ðzzh1 Þ ð10aÞ
2 h1

and

1  2  mp
tðxzÞ2 ðx,tÞ ~ E2 z {h2 z W2xxx ðx,tÞ{ ðz{h2 Þ ð10bÞ
2 h2

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 169
makes it possible to rewrite Eq. (9) in the Laplace transform plane as
ðL ð0 ðL ð0
1  2 * mp
E1 z zh1 z W xxx ðx,sÞdzdxz ðzzh1 Þdzdx
2 sh1 ð11Þ
0 {h1 0 {h1

Integration of this equation then reveals that the bending moment of the Euler-Bernoulli’s
clamped laminated beams admits the form

*   !
* 12 F ðsÞ 6mP h1 zh2
Wxx ð0,sÞ ~  { ð1zeÞ L ð12Þ
b E1 h1 3 zE2 h2 3 s E1 h1 3 zE2 h2 3

These results clearly indicate that the value for the expression (12) cannot be fully determined
until the forcing function F ðtÞ is fully specified. Consequently, further analysis is limited to the
following cases namely:
(a) F ðtÞ~F0 H ðt{t0 Þ where, H(t), is the Heaviside function and
(b) F ðtÞ~F0 eivt

3.1. Case of Heaviside Loading Function


For case (a) above, the forcing function is F ðtÞ~F0 H ðt{t0 Þ which gives the Laplace
* F0 *
transform as F ðsÞ~ e{t0 s :F0 H ðsÞ.
s d *
By recalling the only unutilized boundary condition in Eq. (6), viz W ð0,sÞ~0 and guided
dx
by the Laplace transform of the forcing function, the expression for the bending moment can be
rewritten as
*   !
* 12F 0 H ð s Þ 6mP 0 h1 zh2
Wxx ð0,sÞ ~  3 3
{ ð1zeÞ L ð13Þ
b E1 h1 zE2 h2 s E1 h1 zE2 h2 3
3

Hence, the corresponding response of the Euler-Bernoulli’s laminated beam in the Fourier-
Laplace transform plane as presented in Eq. (8) admits the form
0 * 1
n3 p3 nz1 * np 12F0 H ðsÞ
B L3 ð{1Þ sW 1 ðL,sÞ{ L  LC
B b E1 h1 3 zE2 h2 3 C
B C
B np 6mP0  h1 zh2  C
Bz ð 1ze ÞL C
B 3 3 C
B L s E1 h1 zE2 h2 C
B ! C
B a za  1z ð {1 Þ nz1 C
@ 1 2 A
z P0 e
*F 2 np
W ðln ,sÞ ~   ð14Þ
b1 zb2
sðs{ivn Þðszivn Þ
2
2n4 p4
where v2n ~ is the natural frequency of vibration of the clamped dissimilar layered
ðb1 zb2 ÞL4
beam as can be derived by setting the right hand side of Eq. (8) to zero.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 170
The next step is to evaluate the Fourier inversion of Eq. (14) as
8 P
? 9
> * nz1 sinnp x >
>
> 2sW ð L,sÞ ð {1 Þ >
>
>
> np >
>
>
>
n~1
! >
>
>
> *   >
>
>
> 12F 0 H ð s Þ 6mP 0 h 1 zh2 P
? sinnp x >
>
>
> 3
  >
3 p3 >
> {2L { >
< 3
b E1 h1 zE2 h2 3 s 3
E1 h1 zE2 h2 3
n~1 n =
 X ?
>
> 3 6mP0 e h1 zh2 sinnp x >
>
>
> z2L >
>
>
> s E h 3
zE h 3 n3 p 3 >
>
>
> 1 1 2 2 n~1 >
>
>
> >
>
>
> L 3 X?
sinnp x >
>
>
> z ða1 za2 ÞP0 e >
>
: 5 5 ;
* 32 n~1
n p
W ðx,sÞ~   ð15Þ
b1 zb2 L4
sðs{ivn Þðszivn Þ 4 4
2 np

where
x
x
~
L
To further simplify the series in Eq. (15), one can invoke the well known closed form Fourier
series representations namely:

1X ?
ð{1Þnz1
x
~ sinnp
x, V 0v
xv1 ð16aÞ
p n~1 n
X
?
sinn
x p2 x
 px2 x3
~ { z , V 0v
xv2 ð16bÞ
n~1
n3 6 4 12

and
X
?
sinn
x p4 x
 p x3 px4 5
x
~ { z { , V 0v
xv2 ð16cÞ
n~1
n5 90 36 48 240

Consequently, Eq. (15) can be expressed in the form


80   19
>
> * L3 x
 2 x3 x 4 2 x5 >
>
>
> B 2sW 1 ðL,sÞ xz ða1 za2 ÞP0 e
32 45
{
9
z {
3 15 C>>
>
>
>
> B C >
>
>
> B *   !   C >
>
>
> B 12F H ðs Þ 6mP h zh x
 x
 2
x
 3 C >
>
> B
> B {2L 3 0 0 1 2 C >
>
>
<B  3 3
 { 3 3
{ z C >
=
* b E 1 h1 zE 2 h2 s E 1 h 1 zE 2 h2 6 4 12 C
W ðx
,sÞ~ BB C
>
    C > ð17Þ
>
> @ 3 6mP0 e h1 zh2 x
 x 2 x 3 A>>
>
> z2L { z >
>
>
> s E h 3
zE h 3 6 4 12 >
>
>
> 1 1 22 >
>
>
> b zb L 4 >
>
>
> 1 2
sðs{ivn Þðszivn Þ 4 4 >
>
>
: 2 np >
;

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 171
*
dW ð0,sÞ
By imposing the boundary condition ~0 in Eq. (17) the deflection at the end of the
d
x
laminated Euler-Bernoulli’s cantilever beam can be evaluated in the Laplace domain as
8 0   19
>
> 2F0 h1 zh2 >
>
<L3 B bE h 3 zE h 3  {mP0 E h 3 zE h 3 ð1zeÞ C>
> =
* B 1 1 2 2 1 1 2 2 C
W ðL,sÞ~ B    C ð18Þ
>
> s @ 3 1 1 A>>
>
: { mP0 e z >
;
ð32Þð45Þ 2 2
E1 h1 E2 h 2

Thus, one can now express Eq. (17) in the form


0 ! 1
2 3
12F 0 6mP 0 ð h1 zh 2 Þ x
 x x

B 2L3  {  ð1zeÞ { z C
B b E1 h1 3 zE2 h2 3 E1 h1 3 zE2 h2 3 6 4 12 C
B C
B    C
@ 3 3 1 1 x
 2x 3
x
 4
2x5 A
zL mP0 e 2
z 2
{ z {
* ð16Þ Eh Eh 45 9 3 15
W ðx
,sÞ~ 1 1 2 2 4
ð19Þ
b1 zb2 L
sðs{ivn Þðszivn Þ 4 4
2 np

which may be rearranged as


0 ! 1
2F0 mP0 ðh1 zh2 Þ  2 3

B  {  3 x {
x C
B b E1 h1 3 zE2 h2 3 E1 h1 3 zE2 h2 3 C
B C
B C
B mP0 eðh1 zh2 Þ  2  C
L3 B z  3x {x 3 C
B 3
E1 h1 zE2 h2 3 C
B C
B    C
@ 1 1 3 x
x 4 x 5 A
zmP0 e z { z {
* E1 h1 2 E2 h2 2 24 16 40
W ðx
,sÞ ~   ð20Þ
b1 zb2 L4
sðs{ivn Þðszivn Þ 4 4
2 np

On the other hand by invoking Laplace inversion

ð
gzi?
1 *
,^tÞ
W ðx ~ ,sÞes^t ds
W ðx ð21Þ
2pi
g{i?

where

^t~t{t0

the dynamic response in state-space domain can be evaluated as

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 172
0 ! 1
2F0 mP0 ðh1 zh2 Þ  2 3

B  {  3 x {x C
B b E1 h1 3 zE2 h2 3 E h 3
zE h 3 C
B 1 1 2 2 C
B   C
B mP e ð h zh Þ C
,^tÞ
W ðx ~ L3 ð1{cosvn^tÞB z
0 1 2
 {3 x 2
z x 3 C ð22Þ
B E h 3
zE h 3 C
B 1 1 2 2 C
B    C
@ 1 1 3 x
x 4 x 5 A
zmP0 e 2
z 2
{ z {
E1 h 1 E2 h2 24 16 40

or as
0  2  1
D 1 {mP  2 3
 0D x3
x {
B    3   4 C
W ðx
,tÞ ~ F 1 ðt Þ@ D x
 5 A
x ð23Þ

zmP0 e {3D2 x 2 
 { D2 { 3 
 zD3
x {
12 8 20

where the following have been used


 {1 {2

 1~  2  ,  2 ~ 1zY ,  3~ 1zc Y
D D D , F1 ðtÞ~ð1{cos2ptÞ ð23aÞ
1zcY3 1zcY3 2

in conjunction with the non-dimensionalized parameters viz:

,tÞE1 bh1 3
W ðx  0 ~ P0 ; 2p E2 h2
W ðx
,tÞ~ ; P ^t~ t; c~ ; Y~ ð24Þ
L3 F0 F= vn E1 h1
bh1

3.2. Case of Harmonic Loading Function


For this case, the forcing function is F(t) 5 F0eivt. This gives its Laplace transform as

* F0
F ðsÞ~ ð25Þ
s{iv

where, v is the excitation frequency. On the other hand,


8 * 9
>
> 2sW ðL,sÞx >
>
>
> *  ! > >
>
> 12sF0 H ðsÞ 6mP0 ðh1 zh2 Þ >
>
< 3 =
{2L  3 3
 {ð 1ze Þ 3 3
L1
> ð s{ivÞb E h
1 1 zE h
2 2 E h
1 1 zE h
2 2 >
>
>   >
>
>
> 3 1 1 >
>
>
: z mP0 eL 3
z L >
;
2 2 2
* 16 E1 h1 E2 h2
W ðx,sÞ~   ð26Þ
b1 zb2 L4
sðs{ivn Þðszivn Þ 4 4
2 np

where

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 173
   
x 2 x
 x 3 x
 x3 x
2 4 2
x5
L1 ~ { z ; L2 ~ { z {
6 4 12 45 9 3 15
*
dW ð0,sÞ
By imposing the boundary condition ~0 in Eq. (17) one can evaluate the deflection
dx
at the end of the laminated Euler-Bernoulli’s cantilever beam in the Laplace domain as:
0 !1
2F0 6mP0 ðh1 zh2 Þ
B   {ð1zeÞ   C
B 3
3 B ðs{ivÞb E1 h1 zE2 h2
3
s E1 h1 3 zE2 h2 3 C
LB C
  C
@ mP0 e 1 1 A
z z
* 480 E1 h1 2 E2 h2 2
W ðL,sÞ~   ð27Þ
b1 zb2 L4
sðs{ivn Þðszivn Þ 4 4
2 np

Subsequent substitution into Eq. (26) and carrying out the Laplace inversion for the non-
dimensionalized variable gives the result
0 2 3 1
2F2 ðtÞ
B 6 1zcY3  7 C
B 6 7 C
B ð2
x{12L1 Þ6    7 C
B 4 m P ð 1zY Þ 5 C
W ðx
,tÞ ~ B {ð1zeÞ
0
F ð tÞ C ð28Þ
B 1 C
B 1zcY3 C
B   C
@ x
 3   A
{ 
{ L2 mP0 e 1zc Y {1 {2
F1 ðtÞ
240 16

which on introducing some of the non-dimensionalized parameters earlier used, can be re-
arranged as
0  2  1
D 1 F2 ðtÞ{mP  2 F1 ðtÞ 3
 0D x3
x {
B    3   4 C
W ðx
,tÞ ~ @ D x
 5 A
x ð29Þ

zmP0 eF1 ðtÞ {3D2 x  2
 { D2 { 3 
 zD3
x {
12 8 20

where D  1 ,D
 2 ,D
 3 are as previously defined.
It is also convenient to introduce the additional normalizations

v 2pt
~g; ~t ð30Þ
vn vn

where g can be regarded as the associated frequency ratio of the driving load.
This facilitates the rearrangement of the earlier expressions for F1(t) and F2(t) as

F1 ðtÞ~F1 ðtÞ~ð1{cos2ptÞ ð31Þ

and

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 174

1 ðcosð2pgtÞ{cos2ptÞ
F2 ðtÞ~F2 ðtÞ~ ð32Þ
ð1{g2 Þ ziðsinð2pgtÞzgsin2ptÞ

4. COMPUTATION OF DYNAMIC SLIP


In this section, we can proceed to compute the slip associated with the motion under study for
the two cases being considered.

4.1. Case of Heaviside Loading Function


The relative dynamic slip at the interface of the laminated beams is given by

Duðx,^tÞ ~ u2 ðx,^t,0{Þ{u1 ðx,^t,0zÞ ð33Þ

Following Goodman and Klumpp, this can also be written as

ðx ðx
{1 {1
Duðx,^tÞ ~ E1 ðsx Þ1 ðj,^t,0{Þ dj{E2 ðsx Þ2 ðj,^t,0zÞ dj ð34Þ
0 0

so that on substituting the relevant bending stress relations in state-space domain namely:

E1 ð2zzh1 Þ L2 W1 ðx,^tÞ mPav ðx{LÞ


ðsx Þ1 ðx,z,^tÞ~{ { ð35aÞ
2 Lx2 h1

and
E2 ð2z{h1 Þ L2 W2 ðx,^tÞ mPav ðx{LÞ
ðsx Þ2 ðx,z,^tÞ~{ z ð35bÞ
2 Lx2 h2

as listed in Eqs. (35a) and (35b) above, Eq. (34) can now be expressed in the form
8 2 2
9
> h L W ðj,^
t Þ h L W ð j,^t Þ >
ðx >
>
<
2 2
2
z
1 1
2
z >
>
=
2 Lj 2 Lj
Duðx,^tÞ ~ dj ð36Þ
>
> mP av  av
mP >
>
0 :> z ðj{LÞz >
ðj{LÞ ;
E2 h2 E1 h1

so that Eq. (36) is then integrated to give


8 9
>
> h2 L2 W2 ðx,^tÞ h1 L2 W1 ðx,^tÞ >
>
>
> z >
>
>
> 2 Lx 2 2 Lx2 >
>
>
<  2
  2
 >
=
mP0 x mP0 x
Duðx,^tÞ ~ z {Lx z {Lx ð37Þ
>
> E2 h2 2 E1 h1 2 >
>
>
>     >
>
>
> mP0 e x 3
x 2
mP0 e x 3
x 2 >
>
>
: z { L z { L > ;
E2 h2 L 3 2 E1 h1 L 3 2

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 175
which that on introducing the usual non-dimensionalised parameters, gives
    
LW  2  2 3
Duðx
,tÞ ~ 
D4  
zmP0 D5 x x ze x
 {2  {
x 2
ð38Þ
Lx 3

where the following parameters have been introduced


   
 1zY{1  1zc{1 Y{1
D4 ~ , D5 ~
2 2

Thus by using the results for W from Eq. (23), we can rewrite Eq. (38) as
2   
 5 {3
 2  3
mP0 3 k1 F1 ðtÞzD k3 F1 ðtÞ x {2 x
6 2    37
6 k
4 5 x 7
 06 6 {6 k1 F1 ðtÞxz 3 k1 { F1 ðtÞ{D 2 7 7
Duðx
,tÞ~F1 ðtÞmP 6 4 77 ð39Þ
 0 e6
6 zmP 6   77
4 4 k4 2 k
4 55
3 4
z F1 ðtÞz D5 x  { F1 ðtÞ x
2 3 2

Here the following have also been defined


 2
ð1zYÞ 1zY{1
1 ~
k   ,
4 1zcY3
 2   2
1zY{1 1zc{2 Y{2 1zY{1
3 ~ 
k , k4 ~
2 1zcY3 8

4.2. Case of Harmonic Forcing Function


Following the procedure introduced in [3], it is now possible to compute the slip under

harmonic load by recalling Eq. (38) and substituting for W from Eqs. (29), (31) and (32) to
obtain
8    
 4 {3L
 2
 1 F2 ðtÞ x
 9
> mP0 3L2 F1 ðtÞzL  {2 x >
>
>     > >
>
>  >
>
< 
 0 e {6L2 F1 ðtÞ  L 3  2 =
zmP xz 3L2 { F1 ðtÞ{D5 3
x
Du~ 4 ð40Þ
>
>   >
>
>
> F1 ðtÞ 2 3  F1 ðtÞ 4 >
>
>
: z D3 { x  {L3 x
 >
;
4 3 2

where the following have been defined

   {1
!   !
 1zY{1  2~ ð 1zY Þ 1zY  3~ 1zc{1 Y{2 1zY{1
L1 ~ , L   , L
1zcY3 2 1zcY3 4 ð40aÞ

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 176
5. ENERGY DISSIPATION
5.1. Case of Heaviside Forcing Function
It can be recalled from earlier work [3] that the energy dissipated per cycle, following
Goodman and Klumpp is given by the relation
p
=2vn L
ð ð
D ~ 4mb PðxÞDuðx,^tÞdxdt ð41Þ
0 0

which can also be expressed as

1= 1
ð4 ð
D ~ 4m P ðx
ÞD
uðx
,tÞd
xdt ð42Þ
0 0

whereby on recalling the linear pressure profile and substituting for Du from Eq. (39) we can
evaluate the right hand side of relation (42) to give.
 
D~D1 zD2 ð43Þ

with
   
 e 8  0 {m P
2  2 8 2 5
D1 ~ 1z 3 mP
k 0 1 z D
k ð43aÞ
2 11 11 3

and
8   9
> 12 5
D 3 >
>
> 2 2 >
>
< m P0 e { 11 k1 z z
6 110
k
4 =

D2 ~    ð43bÞ
> 4 >
>
>
: ze  3 mP
k 2 4 k
 0 {m2 P 1
1 z D 5
>
>
;
0
11 11 3

5.2. Case of Harmonic Forcing Function


Following the procedure outlined in the preceding section and substituting the corresponding
values for Du for the harmonic excitation from Eq. (40) into Eq. (42), we obtain the energy
dissipated through slip for this case as

D~D1 zD2 zD3 zD4 ð44Þ

where,
  
2 2 8 2
D1 ~ 8DðgÞmP0 L1 {m P0 L2 z L3 ð45aÞ
11 3

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 177
  
   2 2 4  1
D2 ~e 4DðgÞmP0 L1 {m P0 L2 z L3 ð45bÞ
11 3

 
2
 3 ~em2 P 12  3  1
D 0 { L2 z L3 z D5 ð45cÞ
11 110 6

 
2
 4 ~e2 m2 P 6  3  1 
D 0 { L 2z L3 z D 5 ð45dÞ
11 220 12

and where
20 p  1 0 p  13
1 4@ sin g 1 cos g {1 g
DðgÞ ~ 2 { A{i@ 2 z A5 ð45eÞ
ð1{g2 Þ 2pg 2p 2pg 2p

5.3. Analysis of Optimum Energy Dissipation for The Case of Harmonic Load Excitation
From Eqs. (44) and (45), it can be deduced that

 1 ð2zeÞDðgÞ
L
mPopt ~      
4  1 8  101  1  2 3  3  1 
L2 z L3 ze L2 z L3 { D5 ze L2 { L3 { D5
11 3 11 660 12 11 440 24

~Jd DðgÞ

where

Jd ~Jd ðe,Y,cÞ

 1 ð2zeÞ
L
~      
4  1 8  101  1  2 3  3  1 
L2 z L3 ze L2 z L3 { D5 ze L2 { L3 { D5
11 3 11 660 12 11 440 24

Thus, the optimal slip energy is computed as


X
4
 opt ~
D  nopt
D ð46Þ
n~1

where,
  
 1 {Jd
 1opt ~Jd D 8L 2 8 2
 2z L3
D L ð47aÞ
11 3

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 178
 
 2  Jd   

D2opt ~eJd D 4L1 { 12L2 z11L3 ð47bÞ
33

2 2
 3 opt ~ eJd D {360L
D  3 z55D
 2 z9L 5 ð47cÞ
330
2 2 2
 4opt ~ e Jd D {360L
D  3 z55D
 2 z9L 5 ~ e D
 3opt ð47dÞ
660 2

5.4. Analysis of Optimum Thicknesses Ratio of Laminate for Energy Dissipation for The Case
of Harmonic load Excitation
The optimum thicknesses ratio of laminate for maximum dissipation of energy can be
computed by employing term wise differentiation via Eq. (46), viz;
!
L X4
E2
Dnopt ðY,c,eÞ ~ðU1 ðY,c,eÞzU2 ðY,c,eÞÞ~0, V c~ ð48aÞ
LY n~1 E1

where;
8 2       39
>
> LJd 12 6 2  12 3 2  11 11 2  >
>
>
> 6 2Jd LY 11 ez 11 e L2 { 110 ez 220 e L3 { 65 ez 130 e L4 7>
>
>
> 6 7>
>
>
> 6   ! 7>
>
>
< 6 {2 3 2 7>
=
26 zJ 2
12 6 2 1{Y {2cY {4cY{6cY 7
U1 ðY,c,eÞ~ D 6 d ez e  2 7 ð48bÞ
(48b)
>
> 6 11 11 2 1zcY3 7>
>
>
> 6 7>
>
> 6   {2 {1 {3    7>
>
>
> 4 3 3 Y zc Y z3c {1 {4
Y 11 11 5>
>
>
>
: z ez e 2
z ez 2
e c Y{1 {2 >
;
110 220 4 65 130

and
8 2     39
  LJd
>
>
>  1 {Jd 8 L  2z 2 L
 3 ze 4L  1 {Jd 4 L  1 z L3 >
>
>
> 6 8L
11 3 11 3 LY 7>
>
>
>
> 6 7>
>
> 6 !   7>
>
>
> 6 7>
   7= (48c)
< 6 2 {2
3cY z4cYzY 8  2L3 4L2 e 2L3 e LJd
26 {Jd ð8z4eÞ   {J L z z z 7
U2 ðY,c,eÞ~ D 6 2 d
11
2
3 11 3 LY 7> ð48cÞ
>
> 6 1zcY3 >
>
> 6 " !   {2 {1 {3 #7
7>
>
>
> 6 {2 3 2 {1 {4  7>
>
>
> 4 {J 8 ð2zeÞ 1{Y {2cY {4cY{6cY { 2ze Y zc Y z3c Y 5>
>
>
> d   >
>
: 11 3 2 3 4 ;
2 1zcY

Here;
!
3cY2 z4cYzY{2  1 LH ðY,c,eÞ
{H ðY,c,eÞ   2
{L
LJd 1zcY3 LY
~ ð48dÞ
LY H 2 ðY,c,eÞ

while;

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 179
H ðY,c,eÞ
     
4  1 8  101  1  2 3  3  1 
~ L2 z L3 ze L2 z L3 { D5 ze L2 { L 3 { D5
11 3 11 660 12 11 440 24

and

LH ðY,c,eÞ
LY
8   !   {2 {1 {3 ! 9
>
> 4 3 2 1{Y{2 {2cY3 {4cY{6cY2 3 101 3 2 Y zc Y z3c{1 Y{4 >
>
>
> ð1zeÞz e   z z e{ e >
>
< 11 11 3 2 11 660 440 4 =
2 1zcY
~   
>
> >
>
>
> 1 1 >
>
: z ez e2 c{1 Y{2 ;
24 48

Consequently, the optimum thicknesses ratio Ymin is a solution to the transcendental


equation namely;
E2
U1 ðYmin ,c,eÞzU2 ðYmim ,c,eÞ~0 ; V c~ ð49Þ
E1

5.5. Analysis of Optimum Moduli Ratio at Optimum Thicknesses Ratio of Laminate for
Energy Dissipation for The Case of Harmonic load Excitation
The optimum moduli ratio at optimum thicknesses ratio of laminate for maximum
dissipation of energy can be computed by employing term wise differentiation via Eq. (48), viz;

Fig. 2. Energy dissipation as a function of interfacial pressure gradient.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 180
0 ! 1
X
4   
L@ L  A~ L ðU1 ðYmin ,c,eÞzU2 ðYmin ,c,eÞÞ ~0
Dnopt ðY,c,eÞ  ð50aÞ
Lc LY n~1  Lc
Ymin

where;
8 2 0  1 39

>
>
>  2 ðYmin ,cÞ LJd ðYmin, cÞ :LJd 
L
>
>
>
>
> 6 B Lc LY  C 7>
>
>
> 6 24 12 B Y~Ymin C 7>
>
>
> 6 ez e 2 B  ! C 7>
>
>
> 6 11 11 B  C 7>
>
>
> 6 @ L LJd  A 7>
>
>
> 6 zJd ðYmin ,cÞ  7>
>
>
> 6 Lc LY 7>
>
>
> 6  Y~Ymin >
7>
>
> 6    >
7>
>
> {2
3 2 c ð1zYmin Þ {1 >
7>
L < 6 6z
12
ez e z
11
ez
11 2 cYmin
e 7=
ðU1 ðYmin ,c,eÞÞ~ D2 6
6 110 220 4 65 130 2 7 (50b)
7> ð50bÞ
Lc >
> 0 17>
> 6
> 6 zJ 2
ð Y c ÞH ðY ,c Þ 7>
>
>
> 6 12  d min, 1 min
! 7>
>
>
> 6 6 B C7>
>
> 6
>
> ez e B 2
@ LJd ðYmin ,cÞ 1{Ymin {2cYmin {4cYmin {6cYmin C
{2 3 2
A7>
>
> 6 11
> 11 z2Jd ðYmin ,cÞ :   7>
>
>
>
> 6 Lc 2 7>
7>
3
>
> 6 2 1zcY >
>
>
> 6       7>
>
>
> 4 3 3 {c {2
Y {3
{3c {2
Y {4
11 11 5>
>
>
> z ez e 2 min min
{ ez 2 {2
e c Ymin {2 >
>
: ;
110 220 4 65 130

Here,

H1 ðYmin ,cÞ~
0 2     1
1zcYmin 3 {Ymin 3{2Ymin {3Ymin 2 { 1{Y{2 {2cY3 {4cY{6cY2 Y3 minzcY6 min
@   A
3 4
1zcYmin

Fig. 3. Energy dissipation as a function of interfacial pressure gradient.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 181
while

LU2 ðYmin ,c,eÞ


~
Lc
8 2  !     39
>
> Ymin 3 zYmin 2 LJd ðYmin ,cÞ 8  2 8 L   2L  >
>
>
>
> 6
> {8   { L2 ðYmin ,cÞz L3 ðYmin ,cÞ {Jd ðYmin ,cÞ L2 ðYmin ,cÞ z 
L3 ðYmin ,cÞ 7>
>
>
>
>
> 6 1zcYmin 3 2 Lc 11 3 11 Lc 3 Lc 7>
>
>
> 6 7>
>
> 6
> 6 0   !  !   !  {2  ! 1 7>
>
7
> 7>
3 2  3 2 {2 {3
>
> 6 Y zY LJ ð Y,c Þ 4 Y zY c Y zY >
7>
min min : d  min min min min
>
> 6 B {4   { J d ð Y min ,c Þ  {   { C >
>
>
> 6 B 3 2 LY 11 3 2 12 C 7>
>
>
> 6 B 1zcY min Y~Y min 1zcY min C 7>
>
>
> 6 ze B      ! C 7>
>
>
> 6 B    C 7>
>
>
> 6 @ 4   LJd ðYmin ,cÞ  4  1 L LJd ðY,cÞ A 7>
>
>
> 6 { L 1 ð Y min ,c Þz L 3 ðY min ,c Þ z 4 L 1 ð Y min ,c Þ{J d ð Y min ,c Þ L 1 ð Y min ,c Þz L 1 ð Y min ,c Þ  7>
>
>
> 6 11 Lc 11 3 Lc LY 7>
>
>
> 6 Y~Ymin 7>
>
>
> 6 !         ! !7>
>
>
> 6 2 {2 3 2 2 2 {2 3 6 7>
>
>
> 6 {ð8z4eÞ 3cY min z4cY min zY min LJ d ð Y min ,c Þ 1zcY min 3Y min z4Y min {2 3cY min z4cY min zY min Y min zcYmin 7>
>
>
> zJd ðYmin ,cÞ >
>
< 6 6
 3
 2 Lc ð1zcY 3 Þ
4 7
7=
26
D6 
1zcY

min
 !  ! !
min 7
7
(50c)
>
> 7>
>
> 6
> 6 { ð8z4eÞ L  2 z ð2zeÞL3
 LJd ðYmin,, cÞ :LJd ðY,cÞ L LJd ðY,cÞ 7>
>
>
>
> 6  zJd ðYmin,, cÞ:  7>
>
> 6
> 11 3 Lc LY Y~Ymin Lc LY Y~Ymin 7>
>
>
>
> 6 7>
>
> 6  !   !   {2  {2  ! 7>
>
>
> 6
> LJ ð Y,c Þ  2ze 2Y 3
zY 2
zY 4
2ze c Y min zY {3
min 7>
>
>
> 6 { J ð Y c Þ : d  min min min
{ 7>
>
>
> 6 6 d min,,
LY  11 3 2 3 4 7>
>
>
> Y~Y ð 1zcY min Þ 7>
>
>
> 6 6 "
min
!     # 7>
>
>
> 7>
>
>
> 6 8 1{Y 2
min {2cY 3
min {4cY min {6cY min
2
2ze Y {2
min zc {1 {3
Y min z3c {1 {4
Y min LJ d ð Y min ,c Þ 7>
>
>
>
>
> 6 { ð 2ze Þ { 7>
>
>
> 6 11 2 ð 1zcY 3 Þ
2 3 4 Lc 7>
>
>
> 6 6 min 7>
>
>
> "         ! #! 7>
>
>
> 4 6 3 2 3 2 3
1zcY min Y min z2Ymin z3Y min z Y min zcY min 1{Y min {2cY min {4cYmin {6cYmin 6 2 3 2 7>
>
>
> zJ ð Y ,c Þ 5>
>
>
: d min
3 4 >
;
ð1zcY min Þ

On the other hand;


 
3cY2 z4cYzY{2  1 LH ðY,c,eÞ
{H ðY,c,eÞ 2 {L LY
LJd ð1zcY3 Þ
~ ð50dÞ
LY H 2 ðY,c,eÞ

Fig. 4. Energy dissipation as a function of interfacial pressure gradient.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 182
Fig. 5. Energy dissipation as a function of interfacial pressure.
where;

H ðY,c,eÞ
     
4  1 8  101  1  2 3  3  1 
~ L2 z L3 ze L2 z L3 { D5 ze L2 { L3 { D5
11 3 11 660 12 11 440 24
and

Fig. 6. Energy dissipation as a function of interfacial pressure gradient.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 183
Fig. 7. Energy dissipation as a function of interfacial pressure gradient.

LH ðY,c,eÞ
LY
8  !   !9
>
> 4 3 2 1{Y{2 {2cY3 {4cY{6cY2 3 101 3 2 Y{2 zc{1 Y{3z3c{1 Y{4 >>
> >
< 11 ð1zeÞz 11 e
>  2 z z
11 660
e{
440
e
4
>
=
2 1zcY3
~   
>
> >
>
>
> 1 1 2 {1 {2 >
>
:z ez e c Y ;
24 48

Fig. 8. Energy dissipation as a function of c for g 5 0.1; and Y 5 1.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 184
Consequently, the optimum moduli ratio cmin is a solution of the transcendental equation
namely;
 
E2
Y1 ðYmin ,c,eÞzY2 ðYmim ,c,eÞ~0; V c~ ~1 ð51Þ
E1

6. LOGARITHMIC DAMPING DECREMENT FOR THE CASE OF HARMONIC


LOAD EXCITATION
Following similar procedural analysis introduced in an earlier paper [3], the logarithmic
damping for this generalised problem can be defined as
 
1 D
d~ Ln 1z ð52Þ
2 U 1 zU 2

where for this case, U1 is the energy introduced by the bending moment and can in this case be
calculated from the theorem of Castigliano as
n o
2
U 1 ~27 H1 ei4pg {2H2 mP0 ð1zeÞei2pg zH3 m2 P0 1z2eze2 ð53aÞ

On the other hand U2 is the energy stored from the deflection at the free end which can also
be computed in this case from the theory of strength of materials as
8  e   i2pg 9
>
> H ei4pg
{2mP H z 240 H zH e >
>
< 1 0 2
240
2 4 =
 2~
U       ð53bÞ
>
> 2 240H3 zH5 2402 H3 z480H5 zH6 2 >
: zm2 P0 H3 z ez e >;
120 2402

Fig. 9. Energy dissipation as a function of c for g 5 0.1; and Y 5 0.45.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 185
Fig. 10(ab) Energy dissipation as a function of Y for g 5 0.1; and c 5 1. Energy dissipation as a
function of Y for g 5 0.1; and c 5 2.0.

In writing the results in Eqs. (52), (53a) and (53b) we have made use of the non-
dimensionalized variables viz:

Un Ebh3
U n~ ; n ~ 1, 2
F02 L3

4 2ð1zYÞ ð1zYÞ2
H1 ~  2 , H2 ~  2 , H3 ~  2
1zcY3 1zcY3 1zcY3

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 186
Fig. 10(cd) Energy dissipation as a function of Y for g 5 0.1; and c 5 4. Energy dissipation as a
function of Y for g 5 0.1; and c 5 6.

     2
1zc{1 Y{2 1zc{1 Y{2 ð1zYÞ 1zc{1 Y{2
H4 ~   , H5 ~   , H6 ~
1zcY3 2 1zcY3 4

7. ANALYSIS OF RESULTS
In general, the results indicate that both the displacement and slip behave, by and large, as
was reported earlier in [3]. What is of interest here is to ascertain to what extent can we influence
the amount of energy dissipated either by our choice of different materials for the upper and
ower laminates or by varying the thicknesses of the upper and lower laminates? These effects are
best simulated by varying the values of c and Y these being the relative ratios of the Young’s moduli

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 187
Fig. 11. Energy dissipation as a function of c for g 5 0.1; and Y50.35.
E2 h2
and laminate thicknesses respectively (c~ ; Y~ ). In such circumstances, it stands to reason
E1 h1
that for the special case Y 5 c 5 1, we expect to recover the earlier results reported in [3].
In particular, as a check, the results reported in [3] for the slip energy dissipation profiles are
correctly recovered from the present work as the special case c 5 1, Y 5 1. In this respect, it can
be observed that Figs. 2–4 confirm the convergence of both solutions for the sample cases g 5
0.1, 0.5, and 0.85 respectively.
Figures. 5–7 on the other hand, amply demonstrate that given some fixed values of c and Y,
the amount of energy dissipated through slip progressively decreases as g increases as was
reported in [3], whenever g is restricted within the pre-resonance zone.
In this case, it is also confirmed that as the value of g increases the value of energy dissipation
drops for a prescribed value of the normalised pressure mP. More importantly, it is observed

Fig. 12. Energy dissipation as a function of Y for g 5 0.5; and c 5 1.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 188
Fig. 13. Energy dissipation as a function of c for g 5 0.5; and Y51.

that the relative ordering of the energy curves as a function of the pressure gradient e appears
reversed from the usual pattern where negative pressure gradient at the laminate interface
induces higher energy dissipation.
However, part of the exercise here is to know how the ratio of laminate thicknesses, Y and
modulus of rigidity of laminate materials c influence the level of dissipation of the vibration
energy. The effect of the ratio of modulus of rigidity of laminate materials c is shown in Figs. 8,
9, 11, 13 and 14 for different values of g and Y whilst the effect of laminate thicknesses Y is
shown in Figs. 10, 12, 15, and 16. We find that, more energy is dissipated in almost all cases
than the results earlier reported in [3] for which we had Y 5 c 5 1. The results also indicate that
in general, the modulating role of pressure gradient becomes relatively insignificant as
compared with some of the other parameters under consideration.

Fig. 14. Energy dissipation as a function of Y for g 5 0.5; and c 5 0.35.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 189
Fig. 15. Energy dissipation as a function of Y for g 5 0.5; and c 5 2.5.

In fact, by studying Figs. 8–13, it is clear that for energy dissipation, it is more efficacious to
play with either Y or c than to settle for the symmetric case Y 5 c 5 1 as this gives less energy
for a given value of g than either of the two other cases.
It is also clear that in comparison more energy can be dissipated by varying c as against Y
since for any given g, the maximum dissipated energy from the c curve is larger than that
obtained from the Y curve.
Moreover, by carefully selecting both Y or c much larger energy dissipation can in fact be
arranged than in any of the cases discussed above.
For example by choosing a value of c . 1, one can arrange a much higher value for energy
dissipation and vice-versa.

Fig. 16. Energy dissipation as a function of (Y,c) for g 5 0.1; and e 5 0.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 190
 opt versus Y as modulated by g , c for the case
Table (a). Optimum energy, derived from plot of D
e~0.
g c~0:075 c~0:25 c~0:35 c~1 c~2 c~4 c~6
0.1 D opt 50.0958  opt 50.1451 D
D  opt 5 0.1601 D
 opt 50.2160 D
 opt 50.2346  opt 50.2464
D  opt 50.2121
D
0.5 D opt 50.0936  opt 50.1418 D
D  opt 50.1565 D
 opt 50.2111 D
 opt 50.2293  opt 50.2408
D  opt 50.2072
D
 opt 50.0895
0.85 D   
Dopt 50.1356 Dopt 50.1495 Dopt 50.2018 D  opt 50.2191  opt 50.2301
D  opt 50.1981
D
2.5 D opt 50.0528  opt 50.0800 D
D  opt 50.0883 D
 opt 50.1191 D
 opt 50.1294  opt 50.1359
D  opt 50.1169
D

 opt versus Y as modulated by g , c for the case


Table (b). Optimum energy, derived from plot of D
e~{0:2.

g c~0:075 c~0:25 c~0:35 c~1 c~2 c~4 c~6


0.1 D opt 50.0854 D
 opt 50.1362 D
 opt 50.1533 D
 opt 50.2134  opt 50.2311
D  opt 50.2614
D  opt 50.2333
D
0.5 D  opt 50.1331 D
 opt 50.0834 D  opt 50.1498 D
 opt 50.2085  opt 50.2259
D  opt 50.2554
D  opt 50.2280
D
  
0.85 Dopt 50.0797 Dopt 50.1272 Dopt 50.1432 D  opt 50.1993  opt 50.2159
D  opt 50.2441
D  opt 50.2179
D
2.5 D  opt 50.0751 D
 opt 50.0471 D  opt 50.0846 D
 opt 50.1177  opt 50.1275
D  opt 50.1441
D  opt 50.1286
D

 opt versus Y as modulated by g , c for the case


Table (c). Optimum energy, derived from plot of D
e~0:2.

g c~0:075 c~0:25 c~0:35 c~1 c~2 c~4 c~6


0.1 D opt 50.1064 D
 opt 50.1529 D
 opt 50.1652 D
 opt 50.2145  opt 50.2336 D
D  opt 50.2226 D
 opt 50.1814
  
0.5 Dopt 50.1040 Dopt 50.1494 Dopt 50.1614 D  opt 50.2096  
Dopt 50.2283 Dopt 50.2175 D  opt 50.1773
0.85 D  opt 50.1428 D
 opt 50.0994 D  opt 50.1543 D
 opt 50.2003  opt 50.2182 D
D  opt 50.2079 D
 opt 50.1695
2.5 D  opt 50.0843 D
 opt 50.0587 D  opt 50.1241 D
 opt 50.1170  opt 50.0793 D
D  opt 50.0313 D
 opt 50.0175

 opt versus c as modulated by g , Y for the case


Table (d). Optimum energy, derived from plot of D
e~0.
g Y~0:075 Y~0:25 Y~0:35 Y~0:45 Y~1 Y~4 Y~6
0.1 D opt 50.2650 D
 opt 50.2485 D
 opt 50.2347  opt 50.2196
D  opt 50.1415
D  opt 50.0426 D
D  opt 50.0220
0.5 D  opt 50.2429 D
 opt 50.2590 D  opt 50.2293  opt 50.2146
D  opt 50.1383
D  opt 50.0417 D
D  opt 50.0215
 
0.85 Dopt 50.2475 Dopt 50.2321 D  opt 50.2192  opt 50.2051
D  opt 50.1322
D 
Dopt 50.0398 D opt 50.0206
2.5 D  opt 50.1371 D
 opt 50.1461 D  opt 50.1294  opt 50.1211
D  opt 50.0780
D  opt 50.0235 D
D  opt 50.0122

From the engineering and economic points of view, it is much easier to have a hold on Y than
c since the former requires the preparation of laminates of certain thickness ratios which is a
straightforward matter whereas it is sometimes a formidable problem to find appropriate
laminate materials that will give a prescribed c ratio knowing that there might be other
engineering, economic or environmental requirements to be met in the use of such materials.
The pattern of results for energy dissipation also suggests that for effective energy dissipation
there is a preferred arrangement of the laminates. For example, it is observed from Figs. 10, 12,

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 191
 opt versus c as modulated by g , Y for the case
Table (e). Optimum energy, derived from plot of D
e~{0:2.

g Y~0:075 Y~0:25 Y~0:35 Y~0:45 Y~1 Y~4 Y~6


0.1 D opt 50.2762D
 opt 50.2578D
 opt 50.2419  opt 50.2242
D  opt 50.1377 D
D  opt 50.0304 D
 opt 50.0138
 
0.5 Dopt 50.2699Dopt 50.2520D  opt 50.2364  opt 50.2191
D  
Dopt 50.1346 Dopt 50.0297 D  opt 50.0134
0.85 D  opt 50.2408D
 opt 50.2580D  opt 50.2259  opt 50.2095
D  opt 50.1286 D
D  opt 50.0284 D
 opt 50.0128
 
2.5 Dopt 50.1523Dopt 50.1422D  opt 50.1334  opt 50.1237
D  
Dopt 50.0759 Dopt 50.0168 D  opt 50.0076

 opt versus c as modulated by g , Y for the case


Table (f). Optimum energy, derived from plot of D
e~0:2.

g Y~0:075 Y~0:25 Y~0:35 Y~0:45 Y~1 Y~4 Y~6


0.1 D opt 50.2517D
 opt 50.2365D
 opt 50.2250  opt 50.2122
D  opt 50.1438 D
D  opt 50.0567 D
 opt 50.0317
 
0.5 Dopt 50.2460Dopt 50.2311D  opt 50.2199  opt 50.2074
D  
Dopt 50.1406 Dopt 50.0554 D  opt 50.0310
0.85 D  opt 50.2209D
 opt 50.2351D  opt 50.2101  opt 50.1982
D  opt 50.1343 D
D  opt 50.0530 D
 opt 50.0296
 
2.5 Dopt 50.1388Dopt 50.1304D  opt 50.1241  opt 50.1170
D  
Dopt 50.0780 Dopt 50.0313 D  opt 50.0175

15 and 16 that the optimum value of Y for which you have maximum dissipation usually occurs
somewhere between 0.35 and 0.55 and certainly below the value Y 5 1. Since Y is the ratio h2/
h1, this means that the thickness of the upper laminate h2 should always be less than that of the
lower laminate as illustrated in Tables (a–f). Similarly, if the ratio of Young’s Moduli of the
laminates c 5 E2/E1 has to be greater than 1, the harder laminate is to be placed on top of the
softer one. Thus when dealing with composite laminates, for optimal energy dissipation, it is
imperative that the softer material (i.e. with Youngs Modulus E1) be placed under the harder
laminate.

Fig. 17. Energy dissipation as a function of (Y,c) for g 5 0.1; and e 5 20.2.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 192
Fig. 18. Energy dissipation as a function of (Y,c) for g 5 0.5; and e 5 20.2.

However there is the separate but related issue of the global ordering of these effects viz: of
the two additional options available for the dissipation of unwanted energy which is more
efficacious; is it by varying the laminate materials, or by varying the thicknesses of the
laminated slabs?
This question is best answered by studying the 3-D plots in Figs. 17–19 which show that whereas
in any given case, there is always an optimum laminate thickness ratio Y below 1 for energy
dissipation the same cannot be said for c as the amount of energy dissipated seems to increase
monotonically well beyond 1 before the dissipated energy peaks. Thus if one is restricted to the use

Fig. 19. Plot of Y1 zY2 ~0 as a function of Y for g 5 5.5; and c 50.85.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 193
Fig. 20. Plot of Y1 zY2 ~0 as a function of Y for g 5 0.5; and c 52.

of a single material for the laminates one can maximise energy dissipation by keeping the ratio of the
laminate thicknesses within the neighbourhood of the optimal
 valueas illustrated in Figs. (19–20)
E2
via the relation U1 ðYmin ,c,eÞzU2 ðYmim ,c,eÞ~0 ; V c~ ~1 for any given g. In the
E1
alternative, for a desired level of energy dissipation achievable from uniform laminate sizes one
can conceivably obtain the same level of energy dissipation by using a cheaper material but now
varying the thickness ratio of the upper to lower laminate to advantage.

8. SUMMARY AND CONCLUSION


In this paper we have revisited the problem of using a layered structural member as a
mechanism for dissipating unwanted vibration or noise, be it in an aerodynamic or machine
structure.
Earlier work had established that some of the factors influencing the level of energy
dissipation include the nature of the pressure distribution profile at the interface of the
laminates as well as the nature of the external force to which the structure is subjected.
In fact, prior to this paper it was well known that a negative pressure gradient in a cantibeam
tends to increase the level of energy dissipation whereas an enhanced frequency ratio of the
driving load tends to reduce the amount of energy dissipation that can be arranged via slip at
the laminate interface. These observations however presume that both upper and lower
laminates are of the same thickness and are made from the same material.
When such restrictions are removed, two new effects arise and are the subject of this paper.
Our findings in fact confirm that each of these factors can independently be exploited to
enhance the level of energy dissipation that can be arranged. In other words such increases can
be arranged either by using different materials for the upper and lower laminates in a prescribed
fashion or by retaining the same material for both laminates but varying the individual ratios of
the laminate thicknesses in a defined manner.
Another conclusion from the present work is that for effective energy dissipation, it is better to
simultaneously play with choice of the laminate materials and their thickness ratios rather than tinker
with any one of them by itself. In fact there are many instances when choice of material alone eclipses

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 194
whatever gains can be made from playing with interfacial pressure gradient. This underscores the
wisdom in the search for composites in the construction of such laminates.
The strategy here is to exploit the advantage of composite structures to dissipate vibration
energy via slip damping especially in aerodynamic structures where the effect of weight of
structural member becomes significant.
The conclusion, then, is that for maximum energy dissipation, you need laminates of different
materials and of different thicknesses. This makes the use of composites inevitable.
These results can be positively exploited in the design of aerodynamic and machine
structures.

9. REFERENCES

1. Damisa, O., Olunloyo, V.O.S., Osheku, C.A. and Oyediran, A.A., ‘‘Static analysis of slip
damping with clamped laminated beams.’’ European Journal of Scientific Research, ISSN
1450–216X, Vol. 17, No. 4, pp. 455–476, 2007.
2. Olunloyo, V.O.S., Damisa, O., Osheku, C.A. and Oyediran, A.A., ‘‘Further results on static
analysis of slip damping with clamped laminated beams.’’ European Journal of Scientific
Research, ISSN 1450–216X, Vol. 17, No. 4, pp. 491–508, 2007.
3. Damisa, O., Olunloyo, V.O.S., Osheku, C.A. and Oyediran, A.A., ‘‘Dynamic analysis of slip
damping in clamped layered beams with Non-Uniform pressure distribution at the interface,’’
Journalof Sound and Vibration, Vol. 309, pp. 349–374, 2008.
4. Gould, H.H. and Mikic, B.B., ‘‘Areas of contact and pressure distribution in bolted joints.
Transactions of ASME,’’ Journal of Engineering for Industry, pp. 864–870, 1972.
5. Ziada, H.H. and Abd, A.K., ‘‘Load pressure distribution and contact areas in bolted joints.’’
Inst. of Engineers (India), Vol. 61, pp. 93–100, 1980.
6. Nanda, B.K. and Behera, A.K., ‘‘Study on damping in layered and jointed structures with
uniform pressure distribution at the interfaces.’’ Journal of Sound and Vibration, Vol. 226, No. 4,
pp. 607–624, 1999.
7. Goodman, L.E. and Klumpp, J.H., ‘‘Analysis of slip damping with reference to turbine blade
vibration.’’ Journal of Applied Mechanics, Vol. 23, pp. 421, 1956.
8. Masuko, M., Ito, Y. and Yoshida, K., ‘‘Theoretical analysis for a damping ratio of jointed
cantibeam.’’ Bulletin of JSME, Vol. 16, pp. 1421–1432, 1973.
9. Nishiwaki, N., Masuko, M., Ito, Y. and Okumura, I., ‘‘Astudy on damping capacity of a jointed
cantilever beam (1st Report; Experimental Results).’’ Bulletin of JSME, Vol. 21, pp. 524–531, 1978.
10. Nishiwaki, N., Masuko, M., Ito, Y. and Okumura, I., ‘‘Astudy on damping capacity of a jointed
cantilever beam (2nd Report; comparison between theoretical and experimental results).’’
Bulletin of JSME, Vol. 23, pp. 469–475, 1980.
11. Motosh, M., ‘‘Stress distribution in joints of bolted or riveted connections. Transactions of
ASME.’’ Journal of Engineering for Industry, 1975.
12. Shin, Y.S., Inverson, J.C. and Kim, K.S., ‘‘Experimental studies on damping characteristics of
bolted joints forplates and shells.’’ Trans ASME, J. Pressure Vessel Technology, Vol. 113, pp.
402–408, 1991.
13. Song, S., Park, C., Moran, K.P. and Lee, S., ‘‘Contact area of bolted joints interface; Analytical,
finite element modelling and experimental study.’’ ASME, EEP, Vol. 3, pp. 73–81, 1992.
14. Nanda, B.K., ‘‘Study of damping in structural members under controlled dynamic slip’’ Ph.D
Thesis, Sambalpur University, 1992.
15. Nanda, B.K., ‘‘Study of the effect of bolt diameter and washer on damping in layered and
jointed structures,’’ Journal of Sound andVibration, Vol. 290, pp. 1290–1314, 2006.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 195
16. Nanda, B.K. and Behera, A.K., ‘‘Dampingin layered and jointed structures,’’ International
Journal of Acoustics and Vibration, Vol. 5, No. 2, pp. 89–95, 2000.
17. Nanda, B.K. and Behera, A.K., ‘‘Improvementof damping capacity of structured members
using layered construction,’’ Seventh International Congress on Sound and Vibration, Garmisch-
Partenkirchen, Germany, pp. 3059–3066, 2000.
18. Osheku, C.A., ‘‘Application of beam theory to machine, aero and hydro-dynamic structures in
pressurized environment’’ Ph.D Thesis University of Lagos, 2005.
19. Hansen, S.W. and Spies, R., ‘‘Structural damping in laminatedbeams due to interfacial slip.’’
Journal of Sound and Vibration, Vol. 204, pp. 183–202, 1997.

Transactions of the Canadian Society for Mechanical Engineering, Vol. 34, No. 2, 2010 196

View publication stats

You might also like