Sunteți pe pagina 1din 38

Author’s Accepted Manuscript

Surface effect on the large amplitude periodic


forced vibration of first-order shear deformable
rectangular nanoplates with various edge supports

R. Ansari, R. Gholami

www.elsevier.com

PII: S0094-5765(15)00362-8
DOI: http://dx.doi.org/10.1016/j.actaastro.2015.09.020
Reference: AA5563
To appear in: Acta Astronautica
Received date: 29 April 2015
Revised date: 21 September 2015
Accepted date: 30 September 2015
Cite this article as: R. Ansari and R. Gholami, Surface effect on the large
amplitude periodic forced vibration of first-order shear deformable rectangular
nanoplates with various edge supports, Acta Astronautica,
http://dx.doi.org/10.1016/j.actaastro.2015.09.020
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Surface effect on the large amplitude periodic forced vibration of first-order shear
deformable rectangular nanoplates with various edge supports

*,b
R. Ansaria, R. Gholami
a
Department of Mechanical Engineering, University of Guilan, P.O. Box 3756, Rasht, Iran
b
Department of Mechanical Engineering, Lahijan Branch, Islamic Azad University, P.O. Box 1616, Lahijan, Iran

Abstract
Surface stress and surface inertia effects may play a significant role in the mechanical characteristics
of nanostructures with a high surface to volume ratio. The objective of this study is to present a
comprehensive study on the surface stress and surface inertia effects on the large amplitude periodic
forced vibration of first-order shear deformable rectangular nanoplates. To this end, the Gurtin–Murdoch
theory, first-order shear deformation theory (FSDT) and Hamilton’s principle are employed to develop a
non-classical continuum plate model capable of taking the surface stress and surface inertia effects and
also the rotary and in-plane inertias into account. To solve numerically the geometrically nonlinear forced
vibration of nanoplates with different boundary conditions, the generalized differential quadrature (GDQ)
method, numerical Galerkin scheme, periodic time differential operators and pseudo arc-length
continuation method are employed. The effects of parameters such as thickness, surface residual stress,
surface elasticity, surface mass density, length-to-thickness ratio, width-to-thickness ratio and boundary
conditions on the nonlinear forced vibration of rectangular nanoplates are fully investigated. The results
demonstrate that surface effects on the nonlinear frequency response of aluminum (Al) nanoplate are
more prominent in comparison with the silicon (Si) nanoplate.

Keywords: Rectangular nanoplates; large amplitude periodic forced vibration; Surface effect; GDQ
method; pseudo arc-length continuation.

1. Introduction
The rapid advances in nanoscience and nanotechnology have led to the rapidly developments in
fabrication of Nano- and Micro- Electro-Mechanical Systems (NEMS and MEMS) in recent years, due to
their superior mechanical and physical properties. Such small-size structures have been widely utilized in
many fields, namely optics, aerospace technology, electronics, chemistry, mechanical engineering and

*
Corresponding author. Tel. /fax: +98 141 2222906.
E-mail address: gholami_r@liau.ac.ir (R. Gholami).

1
biomedical engineering [1-3]. Nanowire, nanobeam, nanoplate and nanoshell are the elementary building
blocks in MEMS and NEMS. In order to design, fabricate and develop such nanostructures, it is necessary
to study all crucial characteristics of their mechanical behaviors. Therefore, a variety of studies have been
carried out on the prediction of mechanical characteristics of nanostructures [4-13]. For example, Ansari
et al. [14] presented a size-dependent Timoshenko beam model on the basis of surface stress elasticity
theory to study the surface effects on the geometrically nonlinear forced vibration characteristics of
nanobeams with various edge conditions. Setoodeh et al.[15] developed a nonlocal Mindlin plate model to
investigate the geometrically nonlinear vibration of orthotropic graphene sheets. Moreover, Shen et
al.[16] examined the nonlinear vibration of bilayer graphene sheets in thermal environments employing
the molecular dynamics simulations and nonlocal elasticity.
Some atomistic and molecular dynamics simulations and experimental studies have demonstrated that
the mechanical characteristics of nano- and micro- structures are size-dependent and behave in a different
way from their macroscale counterparts [17-20]. On the other hand, the classical continuum mechanic is
not able to predict and explore the size-dependent mechanical characteristics of structures at micro- and
nano-scales. Therefore, several investigations have been performed to develop the non-classical
continuum theories such as nonlocal elasticity [21], strain gradient elasticity theory [22], modified strain
gradient elasticity [19], modified couple stress theory [23] and surface stress elasticity theory [24, 25]
which are capable of incorporating the size-effects into account. Among different non-classical
continuum theories, the surface stress elasticity theory has been adopted in many investigations to
investigate the effects of surface stress and surface inertias on the mechanical behavior of nanostructures.
The surface stress effect is particularly important in nano-scaled solids or structures with high surface-to-
volume ratios. Moreover, the positive/negative surface stresses lead to inducing a compressive/tension
residual stress fields in bulk part of nanostructures, respectively [25-27]. The compressive residual stress
fields may be leads to a self-instability in the nanostructures even in the absence of external mechanical
loadings [28]. Therefore, to avoid the self-instability of nanostructures, the critical size of nanostructures
should be determined.
A very elegant mathematical formulation within the framework of continuum mechanics was
developed by Gurtin and Murdoch [24, 25] to include the surface stress and interfacial energy into
classical continuum theories. On the basis of proposed model, the surface surfaces are simulated as layers
with zero thickness and different material properties from the bulk layer. Later, various researchers
developed the size-dependent nanobeam, nanoplate and nanoshell models incorporating the surface stress
effects to predict the static and dynamic mechanical behaviors of nanostructures [28-33]. For example,
Wang and Feng [34] presented a size-dependent Timoshenko beam model to study the surface effects on
the axial buckling and the transverse vibration of nanowires. They included that the positive surface

2
elastic constants lead to an increase in the critical buckling loads and natural frequencies. Assadi and
Farshi [35] modified the classical Kirchhoff’s circular plate model to include the effects of surface
properties on the vibration characteristics of circular nanoplates. It observed that surface stress effect on
the natural frequencies and mode shapes is more prominent in larger and thinner circular nanoplates.
Ansari et al.[36] developed a non-classical circular plate model to investigate the vibrational response of
circular nanoplates considering surface energies. Based on the Kirchhoff plate theory, Hasheminejad and
Gheshlaghi [37] adopted a dissipative surface stress model to illustrate the surface dissipation effect on
the quality factor and natural frequencies of elastic nanofilms. Utilizing a nonclassical geometrically
nonlinear beam model on the basis of the Euler-Bernoulli theory, Wang and Wang [38] performed a study
on the non-linear pull-in instability of nano-switches and discovered that surface energy effects on the
pull-in voltage depends on the geometric parameters such as length, height and initial gap of the nano-
switch. Ansari et al. [39] examined the surface effects on the free vibration characteristics of circular
nanoplates in the vicinity of postbuckling domain based on a newly developed nonlinear circular Mindlin
nanoplate model and numerical solution procedure.
To the authors’ knowledge, the surface stress and surface inertia effects on the geometrically
nonlinear forced vibration of nanoplates have not been investigated. Moreover, the influences of the
transverse shear deformation and rotary inertia become more prominent for the thick and moderately thick
nanoplates. Therefore, in this paper, a non-classical first-order shear deformable rectangular plate model
is developed on the basis of Gurtin–Murdoch theory using a variational procedure. The new developed
plate model incorporates the surface stress and surface inertia effects and can capture the small-scale
effect, unlike the classical first-order shear deformable plate model. The newly developed non-classical
plate model is employed to investigate the nonlinear forced vibration of nanoplates with various boundary
conditions employing the generalized differential quadrature (GDQ) method, numerical Galerkin scheme,
periodic time differential operators and pseudo arc-length continuation method.
The rest of the paper is organized as follows. In Sect. 2, a new non-classical model for a first-order
shear deformable plate is developed on the basis of Gurtin–Murdoch theory, Hamilton’s principle and a
variational approach. Also, by defining the appropriate non-dimensional parameters the governing
equations are expressed in the non-dimensional form. In Sect. 3, the geometrically nonlinear forced
vibration problem of a rectangular nanoplate is numerically solved by means of the GDQ method,
numerical Galerkin scheme, periodic time differential operators and pseudo arc-length continuation
method. In Sec. 4, the numerical results are presented to quantitatively show the effect of parameters such
as thickness, surface residual stress, surface mass density and boundary condition on the large amplitude
periodic forced vibration of rectangular nanoplates. Also, the differences between the results given by the

3
current non-classical plate model and classical counterpart are shown. The concluding remarks are given
in Sect. 5.

2. Mathematical Formulation of governing equations and boundary conditions


As shown in Fig. 1, consider a uniform rectangular nanoplate with length , width and thickness
subjected to a harmonic excitation transverse force . Introducing the Cartesian coordinate system on
the middle-plane of nanoplate, the upper ⁄ and lower ⁄ surface layers are
symbolized by and , respectively. According to the first-order shear deformation theory (FSDT),
the components of displacement field of nanoplate along the axes can be expressed
as
(1)
in which , and are, respectively, the displacement components of a point
laying on the middle plane of nanoplate associated with the x-, y- and z-axes. Moreover, and
represent the rotations of a transverse normal about the positive x- and y-axes,
correspondingly.
On the basis of FSDT and referring the von Kármán-type of kinematic nonlinearity, the nonlinear
strain-displacement relations can be expressed as follows

(2)

where

( ) ( )
(3)

In Eq. (2), , and are the in-plane strain components, , and are the analogous
curvature changes and and denote the transverse shearing strains.
For the surface nanoplate model adopted in this work, the constitutive relations of the surface and
bulk layers are different. The conventional constitutive equations are appropriate to apply on the interior
bulk material as those for the conventional materials. Thus, according to the linear elasticity, the
constitutive relations can be written as

[ ( ) ] [ ( ) ] (4)

4
[ ( ) ] [ ( ) ]

( ) ( )

( ) ( )

where ⁄ and ⁄ are the classical Lamé’s constants and and represent
the Young’s modulus and Poisson’s ratio, respectively. Moreover, denotes the shear correction factor,
which is usually considered as ⁄ , such as in the present work.
The Gurtin-Murdoch’s surface stress elasticity theory provides an efficient technique to modify the
classical continuum models and develop size-dependent mathematical formulation including surface
stress and surface inertia effects. On the basis of the Gurtin-Murdoch theory, the surface constitutive
equations in the surface layers can be determined by

(5)

in which and denote the surface Lame constants; is the surface residual stress parameter and the
notation is the Kronecker delta. Assuming that the top and bottom layers have the same material
properties, substituting the appropriate relations of Eqs. (2) and (3) into Eq. (5) yields the following
surface stress components at the upper and lower surfaces of nanoplate

[ ( ) ] [ ( ) ] ( )

[ ( ) ] [ ( ) ] ( )

(6)
[ ( )] [ ]

[ ( )] [ ]

For the very thin plate, the normal stress component in the bulk layer is negligible compared to
other component and is assumed to be zero in classical plate theories. Since this assumption is
inconsistent with the equilibrium of surface layers, it is assumed that the stress component is linearly
variable through the plate thickness to satisfy the equilibrium conditions on the surfaces layers [40]. With
this assumption, can be expressed as

5
( )( ) ( )( ) (7)

Inserting Eq. (6) into (7) yields

( ) (8)

Therefore, the modified components of normal in-plane stresses for the bulk layer of the nanoplate can be
obtained by substituting into Eq. (4) according to the following constitutive relations

[ ( ) ] [ ( ) ]
(9a)
[ ( ) ]

[ ( ) ] [ ( ) ]
(9b)
[ ( ) ]

Based on the continuum surface elasticity theory [15], the total strain energy of the nanoplate
including the bulk and the surface can be achieved as follows

∫∫ (∫ ∫ )

∫ {̅ [ ( ) ] ̅ ̅ [ ( ) ] ̅
(10)
̅ ( ) ̅ ( ) ( )

( ) }
where the in-plane force resultants, bending moments and shear forces can be calculated as

̅ [ ( ) ] [ ( ) ] ( )

̅ [ ( ) ] [ ( ) ] ( )

̅ ( ) ( ) (11)

̅ ( )

̅ ( ) ( )

6
̅ ( ) ( )

( ) ( )

The constants appeared in Eq. (11) are defied as

(12)

Also, the kinetic energy of nanoplate including the surface inertia effect can be expressed as

∫ { [( ) ( ) ( ) ] [( ) ( ) ]} (13)

where

( ) (14)

Moreover, the potential energy resulting from the harmonic excitation transverse load can be
obtained as

∫ (15)

Herein, Hamilton’s principle ( ∫ ) and a variational approach are employed

to derive the governing equations of motion and all possible boundary conditions. After obtaining the
appropriate form of and and by means of the fundamental lemma of the calculus of
variation, the governing equations of motion will be derived as
̅ ̅
(16a)

̅ ̅
(16b)

( )
(̅ ) (̅ ) (̅ ) (̅ )
(16c)

7
̅ ̅
(16d)

̅ ̅
(16e)

Also, the boundary conditions associated with each deformation parameter , , , and will
be achieved as
̅ ̅ (17a)
̅ ̅ (17b)

{ ̅ ̅ } { ̅ ̅ } (17c)

̅ ̅ (17d)
̅ ̅ (17e)
Substituting Eq. (11) into Eqs. (16a-16e) gives the following five governing equations of motion in
terms of the components of displacement , , and as follows

( ) ( ) ( )

(18a)

( ) ( ) ( )

(18b)

( ) ( ) (̅ ) (̅ )
(18c)
(̅ ) (̅ )

( ) ( )
(18d)

( ) ( )
(18e)

Similarly, one can express the boundary conditions in terms of the components of displacement. Now,
introducing the following non-dimensional parameters as

8
( )

( ) (19)

√ ( )

where and , the five geometrically nonlinear governing equations of motion


can be expressed in the non-dimensional form as follows

( ) ( )
(20a)
( )

( ) ( )
(20b)
( )

( ) ( ) (̃ )
(20c)
(̃ ) (̃ ) (̃ )

( )
(20d)
( )

( )
(20e)
( )

where

̃ [ ( ) ] [ ( ) ] ( )

̃ [ ( ) ] [ ( ) ] ( ) (21)

̃ ( )

9
Eq. (17) can provide the mathematical expressions associated with all types of boundary conditions
for a rectangular nanoplate. In present work, the rectangular nanoplates with the following boundary
conditions are considered.
a. All edges simply supported (SSSS)
̃
(22)
̃

b. All edges clamped (CCCC)

(23)
c. Two opposite edges clamped and the remaining edges simply supported (CSCS)

(24)
̃

where

̃ ( )

̃ ( ) (25)

̃ ( )

3. Nonlinear Forced Vibration Problem and a Numerical Solution Procedure


In order to study the effects of surface stress and surface inertia on the geometrically nonlinear force
vibration of nanoplates, firstly, GDQ method [41] is used to discretize the nonlinear governing equations
and corresponding boundary conditions. Also, the transverse load is assumed to be of the form of a
harmonic excitation force as in which and denote the force amplitude and excitation
frequency, respectively. Afterward, a numerical Galerkin scheme is employed to convert the discretized
partial differential governing equations into Duffing-type ordinary differential equations. Then, the
obtained equations are changed to a set of nonlinear algebraic parameterized equations by means of the
periodic time differential operators. As least, the pseudo arc-length continuation method is utilized to
solve the set of nonlinear algebraic equations and determine the nonlinear forced vibration characteristics
of nanoplates.

10
3.1. Discretization on Space Domain
The rth-order derivative of a variable function at a given point on the domain can
be approximated according to the GDQ method [33] as follows

∑ ( ) (26)

where denotes the weighting coefficients of rth-order derivative and can be calculated in the form of
a recursive formula as


( ) ( )

[ ] (27)


{{
Similarly, by means of the Kronecker tensor product indicated by , the partial derivative of a two
variable function introduced on intervals and can be approximately
calculated using the GDQ method. For example, the second order partial derivative of with
respect to and at a given point ( ) can be obtained as

( ) ̅ (28)

where ̅ is a column vector expressed as


̅ (29)
in which and are the number grid notes in the x- and y-directions, respectively.
The two-dimensional functions , , , and are defined on the
intervals and . The shifted Chebyshev–Gauss–Lobatto grid distribution are
employed as follows

( )
(30)
( )

Applying the GDQ methodology, the discretized form of the governing equations (20) are given as
̈ (31)
in which the over dot denotes the differentiation with respect to . Moreover, the parameters , , ,
and represent the field variables vector, stiffness matrix, mass matrix, forcing amplitude vector
and nonlinear stiffness vector, respectively and are defined as

11
[ ] (32a)

(32b)

[ ]
̅( )
( )
̅( )
(32c)
( ) ̅( )

( ) ̅( )
[ ]

[ ] (32d)

[ ] (32e)

The components of matrix are defined in Appendix A and the nonlinear terms
and are given in Appendix B. Also, denotes the
Hadamard product defined in Appendix C.

3.2. Derivation of Duffing-Type Equations


In this section, the efficient numerical Galerkin-based approach is utilized to reduce Eq. (31) into
Duffing-type ordinary differential equations. By neglecting the nonlinear term and external excitation
force, applying the corresponding boundary conditions to the stiffness and inertia matrices and
considering the harmonic solution of system as ̃ , the following generalized eigenvalue
problem is achieved
̃ ̃ ̃ [̃ ̃ ̃ ̃ ̃ ] (33)
in which denotes the non-dimensional natural frequency. Eq. (33) is solved to obtain the linear
frequencies and corresponding mode shapes. The solution of Eq. (33) can be written as
(34)
in which and are the reduced generalized coordinates and numerical Galerkin base function (a sparse
matrix containing the first mode shapes), respectively and are defined as

[ (35a)

12
(35b)

[ ]
where

[̃ ̃ ] [̃ ̃ ] [̃ ̃ ]
(36)
[̃ ̃ ] [̃ ̃ ]

After that, inserting Eq. (34) into (31) gives the residual as
̈ (37)
In the present numerical Galerkin-based approach, a matrix operator is introduced to multiply each
equation by related eigenvectors and integrate over the domain, simultaneously. This can be made easily
by means of the following matrix operator

(38)

[ ]
where and denote the integral matrix operators which are calculated based on the differential
quadrature and Taylor series in the domain and , respectively represented in the
Appendix C. Multiplying Eq. (38) by the residual vector Eq. (37) gives the following Duffing-type
ordinary differential equation
̃ ̈ ̃ ̃ ̃ (39)
where
̃ ̃ ̃ ̃ (40)
In fact, applying the numerical technique Galerkin-based leads to reduce the general coordinates from
discrete points to ones. Moreover, since the linear mode shapes employed in the Galerkin
method are numerically achieved, both essential and natural boundary conditions are satisfied in the
present method. Therefore, it can be expected to achieve the desired accuracy using a small number of
mode shape and reduce the computational effort, significantly.

3.3. Solution in the Time Domain


To discretize Eq. (39) over the tome domain, the time differentiation matrix operators are used [42].
The derivatives of periodic sinc function as base function in the spectral collocation approach are

13
employed to calculate these operators [43]. By defining the non-dimensional parameters ̃ ⁄ and
⁄ and considering the damping effects, Eq. (39) is change to

( ) ̃ ̈ ( )̃ ̇ ̃ ̃ ̃ ̃ (41)

in which the overdot stands for differentiation with respect to ̃ and ̃ is the proportional damping which
is described as ⁄ ̃ where is a small number. Also, the discretized counterpart of on the time
domain is expressed as

[
(42)
]

in which denotes the number of discrete points in the time domain; ̃ represents the nodal value of
̃ at grid point ̃ given by the following relation

̃ ̃ (43)

Considering Eq. (42), Eq. (41) can be discretized as follows

( ) ̃ ( )̃ ̃
̃ ̃ ̃
(44)
[ ̃ ( ̃ )]

here denotes the time differentiation matrix operator defined as follows

(45a)

(45b)

{
where and stand for Teoplitz matrices. Consider ( ) ; where and
are constant matrices, is an unknown matrix and represents the vectorization of matrix (see
Appendix C). Therefore, the vectorized form of Eq. (44) can be expressed as

14
(( ) ( ̃) ( )( ̃) ( ̃ )) (̃ ) ( ̃)
̃
(46)

where is an identity tensor. Eq. (46) is a set of nonlinear algebraic equations which can be
represented as
(47)
which can be solved by means of the pseudo-arc length continuation method [44]. The linear frequency
and linear solution of Eq. (47) are considered as the initial guess.

4. Results and Discussion


The geometrically nonlinear forced vibration characteristics of rectangular nanoplate including the
surface stress and surface inertia effects subjected to the uniformly distributed harmonic force with
frequency in the neighborhood of first free vibration mode are investigated.
In order to verify the validity and accuracy of the present results, the nondimensional linear

frequency ( ̅ ̃ √ ⁄ ; ̃ denotes the dimensional linear frequency and ⁄ )

of an SSSS nanoplate obtained from the present analysis corresponding to various total numbers of the
grid points are presented and are compared with analytical results reported in [45], as shown in Table 1.
The parameters used in the verification are as and , , , ,
⁄ , ⁄ , ⁄ and ⁄ . It should be noted that the
effect of surface inertia is neglected in Ref.[45]. A reasonable agreement is achieved between the present
results and those of [45]. According to Table 1, is used for all of the following numerical
computations. It is worth noting that the nondimensional frequency in the following examples is
calculated as ̃ √ ⁄ .
The material properties of nanoplate for the bulk and surface layers considered in the present study
are taken as [17, 46, 47]
Material 1 (Si <100>):

⁄ ⁄ ⁄ ⁄
Material 2 (Al <111>):

⁄ ⁄ ⁄ ⁄
Nonlinear frequency response of Al and Si nanoplates with SSSS, CSCS and CCCC boundary
conditions are shown in Figs. 2 and 3, respectively, where the maximum vibrational amplitude of

15
nanoplate versus the frequency ratio (ratio of the external excitation frequency to the linear natural
frequency) is plotted for various values of the nanoplate thickness. In addition to the associated linear first
frequencies, the results for the classical nanoplate (without considering the surface effect) are also
provided for a direct comparison. For both Si and Al nanoplates, the decrease of nanoplate thickness leads
to the increase of the importance of surface effect and consequently increasing the stiffness of nanoplate
and the linear frequency. Also, the difference between the frequencies calculated by the nonclassical
nanoplate model and classical model increases significantly, especially for Al nanoplate. It indicates that
the surface effects is more prominent in the prediction of vibrational characteristics of nanoplates at low
thickness. Furthermore, it is found that when the thickness increases the forced vibration amplitude
increases until reaches the one associated with the classical plate model at the higher external excitation
frequencies. The reason is that the hardening effects become more pronounced for nanoplates having the
lower stiffness. Compared to the Si nanoplates, the surface effects have more pronounced influence on
the forced vibration behavior of Al nanoplates. Moreover, it can be seen that the surface effect on the
nonlinear frequency response in the case of SSSS end conditions is more significant in comparison with
CCCC and CSCS end supports.
Figs. 4 and 5 examine the effect of the surface residual stress on the nonlinear frequency response of
Al and Si nanoplates with SSSS, CSCS and CCCC boundary conditions. The positive values of surface
residual stress increase the linear frequency of nanoplates, while the negative values have a decreasing
effect. This is expected because the tensile and compressive in-plane forces are generated in the
nanoplates because of the positive and negative surface residual stresses, respectively. Moreover, Figs. 4
and 5 demonstrate that the effect of surface residual stress is more considerable for SSSS boundary
conditions in comparison with other ones.
Figs. 6-9 represent the effect of surface Lame constants ( ) on the nonlinear frequency
response curves of Al and Si nanoplates with SSSS, CSCS and CCCC edge supports. According to the
results of these figures, the increase of negative and positive surface Lame constants has the opposite
influence on the stiffness of nanoplate as well as the linear frequency and typical hardening behavior.
Increasing the negative surface Lame constants decreases the stiffness of nanoplates and leads to the
decrease of the linear frequencies and increasing the hardening-type response of nanoplate compared to
the classical nanoplate model . In contrast, the increase of the positive values of these
parameters increases the stiffness of Al and Si nanoplates. Therefore, the frequencies given by the
nonclassical surface elasticity nanoplate model are higher than those of the classical plate model.
Furthermore, with the increase of positive surface Lame constants, the typical hardening nonlinearity of
Al and Si nanoplates decreases. Moreover, the effect of surface Lame constants on the typical hardening

16
characteristics of Al nanoplates with three commonly used boundary conditions is more considerable
compared to the Si nanoplates.
Fig. 10 shows the nonlinear frequency response of Si nanoplates for various values of the surface
mass density. It is found that decreasing the surface density enhance the natural frequencies, but has no
effect on the hardening effect of nonlinearity.
Fig. 11 shows the effect of length-to-thickness ratio ( ⁄ ) on the nonlinear frequency response curve
of Si nanoplate with SSSS, CSCS and CCCC edge supports. For all considered boundary conditions, the
increase of length-to-thickness ratio leads to the decrease of linear fundamental frequencies and typical
hardening nonlinearity of Si nanoplates. At low length-to-thickness ratios, the present FSDT plate model
is advantageous over the Kirchhoff plate model in which the shear deformation and rotary inertia are
neglected [48-50].
Fig. 12 is given to compare the nonlinear frequency response of Si and Al nanoplates with different
boundary conditions. It can be inferred from this figure that the deviation of the nonlinear frequency
response curve from the straight line ⁄ is more considerable for the SSSS edge condition
compared to the CSCS and CCCC boundary conditions, especially in the case of Si nanoplates.
For a given value of length-to-thickness ratio, the influence of width-to-thickness ratio ⁄ on the
nonlinear frequency response of Si nanoplate is illustrated in Fig. 13. For all types of boundary
conditions, the decrease of width-to-thickness ratio increases the linear frequency of nanoplates and
typical hardening nonlinearity and decreases the vibration amplitude peak. It should be mentioned that
when the nanoplate has a very small width, the nanoplate seems to behave like a nanobeam.

5. Concluding Remarks
In this study, the surface stress and inertia effects on the large amplitude periodic forced vibration of
first-order shear deformable rectangular nanoplates with various edge supports were investigated. On the
basis of the FSDT and Gurtin–Murdoch surface elasticity theory, the nonlinear governing equations of
motion and corresponding boundary conditions were derived using Hamilton’s principle and the
fundamental lemma of the calculus of variation. The equations of motion were then numerically solved
for the rectangular Si and Al nanoplates with various edge supports by means of the GDQ method,
numerical Galerkin scheme, periodic time differential operators and pseudo arc-length continuation
method. The influences of different parameters such as the nanoplate thickness, surface residual stress,
surface Lame constants, surface density, length-to-thickness ratio and boundary condition on the
amplitude-actuated frequency curves of Si and Al nanoplates were examined. It was inferred that the
influence of hardening nonlinearity and surface effect for Al nanoplates is more considerable compared to
that for Si nanoplates.

17
Appendix A

( ) ( ) ( )

( ) ( ) (( ) ( ))

( ) ( ),

( ) (( ) ( ))

( ) ( ) ( )

( )

( ) (( ) ( ))

( ) ( ) ( )

Appendix B
Applying the GDQ methodology, the nonlinear terms will be discretized as follows

(( ) ) (( ) ) (( ) )

(( ) ) (( ) ) (( ) )

(( ) ) (( ) ) (( ) )

(( ) ) (( ) ) (( ) )

{ [(( ) ) (( ) ) (( ) )]

[(( ) ) (( ) ) (( ) )]

(( ) ) (( ) )} (( ) )

{ [(( ) ) (( ) ) (( ) )]

[(( ) ) (( ) ) (( ) )]

18
(( ) ) (( ) )} (( ) )

{ ( (( ) ) (( ) ) (( ) ) (( ) ))}

(( ) )

Appendix C
Hadamard and Kronecker Products
Definition 1: Let [ ] and [ ] , then the Hadamard product of these matrices takes

the form as [ ] .

Definition 2: If is an m-by-n matrix and is a p-by-q matrix, then the Kronecker product is an
mp-by-nq block matrix and express as

[ ]

Integral Matrix Operators

∫ (∑ ̃ )

where is the GDQ differential operator, and

̃ [ ]

19
References
[1] Y. Cui, Z. Zhong, D. Wang, W. U. Wang, and C. M. Lieber, "High performance silicon nanowire
field effect transistors," Nano letters, vol. 3, pp. 149-152, 2003.
[2] Z. L. Wang and J. Song, "Piezoelectric nanogenerators based on zinc oxide nanowire arrays,"
Science, vol. 312, pp. 242-246, 2006.
[3] X. Feng, R. He, P. Yang, and M. Roukes, "Very high frequency silicon nanowire electromechanical
resonators," Nano Letters, vol. 7, pp. 1953-1959, 2007.
[4] M. Shaat, F. Mahmoud, X.-L. Gao, and A. F. Faheem, "Size-dependent bending analysis of
Kirchhoff nano-plates based on a modified couple-stress theory including surface effects,"
International Journal of Mechanical Sciences, vol. 79, pp. 31-37, 2014.
[5] R. Ansari, V. Mohammadi, M. F. Shojaei, R. Gholami, and M. Darabi, "A geometrically non-linear
plate model including surface stress effect for the pull-in instability analysis of rectangular
nanoplates under hydrostatic and electrostatic actuations," International Journal of Non-Linear
Mechanics, vol. 67, pp. 16-26, 2014.
[6] L. Zhang, J. Liu, X. Fang, and G. Nie, "Size-dependent dispersion characteristics in piezoelectric
nanoplates with surface effects," Physica E: Low-dimensional Systems and Nanostructures, vol. 57,
pp. 169-174, 2014.
[7] S. Sang, Y. Zhao, W. Zhang, P. Li, J. Hu, and G. Li, "Surface stress-based biosensors," Biosensors
and Bioelectronics, vol. 51, pp. 124-135, 2014.
[8] R. Ansari, R. Gholami, M. F. Shojaei, V. Mohammadi, and M. Darabi, "Surface stress effect on the
pull-in instability of hydrostatically and electrostatically actuated rectangular nanoplates with
various edge supports," Journal of Engineering Materials and Technology, vol. 134, p. 041013,
2012.
[9] L.-L. Ke, Y.-S. Wang, J. Yang, and S. Kitipornchai, "Free vibration of size-dependent magneto-
electro-elastic nanoplates based on the nonlocal theory," Acta Mechanica Sinica, vol. 30, pp. 516-
525, 2014.
[10] R. Ansari, M. F. Shojaei, V. Mohammadi, R. Gholami, and F. Sadeghi, "Nonlinear forced vibration
analysis of functionally graded carbon nanotube-reinforced composite Timoshenko beams,"
Composite Structures, vol. 113, pp. 316-327, 2014.
[11] H. Askari, E. Esmailzadeh, and D. Zhang, "Nonlinear vibration analysis of nonlocal nanowires,"
Composites Part B: Engineering, vol. 67, pp. 607-613, 2014.
[12] R. Ansari, E. Hasrati, M. F. Shojaei, R. Gholami, and A. Shahabodini, "Forced vibration analysis of
functionally graded carbon nanotube-reinforced composite plates using a numerical strategy,"
Physica E: Low-dimensional Systems and Nanostructures, 2015.
[13] J. Vila, R. Zaera, and J. Fernández-Sáez, "Axisymmetric free vibration of closed thin spherical
nanoshells with bending effects," Journal of Vibration and Control, p. 1077546314565808, 2015.
[14] R. Ansari, V. Mohammadi, M. F. Shojaei, R. Gholami, and S. Sahmani, "On the forced vibration
analysis of Timoshenko nanobeams based on the surface stress elasticity theory," Composites Part
B: Engineering, vol. 60, pp. 158-166, 2014.
[15] A. Setoodeh, P. Malekzadeh, and A. Vosoughi, "Nonlinear free vibration of orthotropic graphene
sheets using nonlocal Mindlin plate theory," Proceedings of the Institution of Mechanical Engineers,
Part C: Journal of Mechanical Engineering Science, vol. 226, pp. 1896-1906, 2012.
[16] H.-S. Shen, Y.-M. Xu, and C.-L. Zhang, "Prediction of nonlinear vibration of bilayer graphene
sheets in thermal environments via molecular dynamics simulations and nonlocal elasticity,"
Computer Methods in Applied Mechanics and Engineering, vol. 267, pp. 458-470, 2013.
[17] R. E. Miller and V. B. Shenoy, "Size-dependent elastic properties of nanosized structural elements,"
Nanotechnology, vol. 11, p. 139, 2000.
[18] F. Xu, Q. Qin, A. Mishra, Y. Gu, and Y. Zhu, "Mechanical properties of ZnO nanowires under
different loading modes," Nano Research, vol. 3, pp. 271-280, 2010.

20
[19] D. Lam, F. Yang, A. Chong, J. Wang, and P. Tong, "Experiments and theory in strain gradient
elasticity," Journal of the Mechanics and Physics of Solids, vol. 51, pp. 1477-1508, 2003.
[20] C. Chen, Y. Shi, Y. Zhang, J. Zhu, and Y. Yan, "Size dependence of Young’s modulus in ZnO
nanowires," Physical review letters, vol. 96, p. 075505, 2006.
[21] A. C. Eringen, "Linear theory of nonlocal elasticity and dispersion of plane waves," International
Journal of Engineering Science, vol. 10, pp. 425-435, 1972.
[22] R. D. Mindlin, "Micro-structure in linear elasticity," Archive for Rational Mechanics and Analysis,
vol. 16, pp. 51-78, 1964.
[23] F. Yang, A. Chong, D. Lam, and P. Tong, "Couple stress based strain gradient theory for elasticity,"
International Journal of Solids and Structures, vol. 39, pp. 2731-2743, 2002.
[24] M. E. Gurtin and A. I. Murdoch, "A continuum theory of elastic material surfaces," Archive for
Rational Mechanics and Analysis, vol. 57, pp. 291-323, 1975.
[25] M. E. Gurtin and A. I. Murdoch, "Surface stress in solids," International Journal of Solids and
Structures, vol. 14, pp. 431-440, 1978.
[26] Z. Huang and J. Wang, "A theory of hyperelasticity of multi-phase media with surface/interface
energy effect," Acta Mechanica, vol. 182, pp. 195-210, 2006.
[27] Z.-Q. Wang, Y.-P. Zhao, and Z.-P. Huang, "The effects of surface tension on the elastic properties of
nano structures," International journal of engineering science, vol. 48, pp. 140-150, 2010.
[28] Z. Wang and Y. Zhao, "Self-instability and bending behaviors of nano plates," Acta Mechanica
Solida Sinica, vol. 22, pp. 630-643, 2009.
[29] S. M. Hasheminejad and B. Gheshlaghi, "Dissipative surface stress effects on free vibrations of
nanowires," Applied physics letters, vol. 97, p. 253103, 2010.
[30] D. Huang, "Size-dependent response of ultra-thin films with surface effects," International Journal
of Solids and Structures, vol. 45, pp. 568-579, 2008.
[31] R. Ansari, A. Shahabodini, M. F. Shojaei, V. Mohammadi, and R. Gholami, "On the bending and
buckling behaviors of Mindlin nanoplates considering surface energies," Physica E: Low-
dimensional Systems and Nanostructures, vol. 57, pp. 126-137, 2014.
[32] H. Altenbach and V. A. Eremeyev, "On the shell theory on the nanoscale with surface stresses,"
International Journal of Engineering Science, vol. 49, pp. 1294-1301, 2011.
[33] R. Ansari, R. Gholami, M. F. Shojaei, V. Mohammadi, and S. Sahmani, "Surface stress effect on the
pull-in instability of circular nanoplates," Acta Astronautica, vol. 102, pp. 140-150, 2014.
[34] G.-F. Wang and X.-Q. Feng, "Timoshenko beam model for buckling and vibration of nanowires with
surface effects," Journal of physics D: applied physics, vol. 42, p. 155411, 2009.
[35] A. Assadi and B. Farshi, "Vibration characteristics of circular nanoplates," Journal of Applied
Physics, vol. 108, p. 074312, 2010.
[36] R. Ansari, R. Gholami, M. F. Shojaei, V. Mohammadi, and S. Sahmani, "Surface stress effect on the
vibrational response of circular nanoplates with various edge supports," Journal of Applied
Mechanics, vol. 80, p. 021021, 2013.
[37] S. M. Hasheminejad and B. Gheshlaghi, "Eigenfrequencies and quality factors of nanofilm
resonators with dissipative surface stress effects," Wave Motion, vol. 50, pp. 94-100, 2013.
[38] K. Wang and B. Wang, "Influence of surface energy on the non-linear pull-in instability of nano-
switches," International Journal of Non-Linear Mechanics, vol. 59, pp. 69-75, 2014.
[39] R. Ansari, V. Mohammadi, M. F. Shojaei, R. Gholami, and S. Sahmani, "Surface stress effect on the
postbuckling and free vibrations of axisymmetric circular Mindlin nanoplates subject to various edge
supports," Composite Structures, vol. 112, pp. 358-367, 2014.
[40] P. Lu, L. He, H. Lee, and C. Lu, "Thin plate theory including surface effects," International Journal
of Solids and Structures, vol. 43, pp. 4631-4647, 2006.
[41] C. Shu, Differential quadrature and its application in engineering: Springer Science & Business
Media, 2000.
[42] M. F. Shojaei, R. Ansari, V. Mohammadi, and H. Rouhi, "Nonlinear forced vibration analysis of
postbuckled beams," Archive of Applied Mechanics, vol. 84, pp. 421-440, 2014.

21
[43] L. N. Trefethen, Spectral methods in MATLAB vol. 10: Siam, 2000.
[44] H. B. Keller, "Numerical solution of bifurcation and nonlinear eigenvalue problems," Applications of
bifurcation theory, pp. 359-384, 1977.
[45] K. Wang and B. Wang, "Vibration of nanoscale plates with surface energy via nonlocal elasticity,"
Physica E: Low-dimensional Systems and Nanostructures, vol. 44, pp. 448-453, 2011.
[46] S. Ogata, J. Li, and S. Yip, "Ideal pure shear strength of aluminum and copper," Science, vol. 298,
pp. 807-811, 2002.
[47] R. Zhu, E. Pan, P. W. Chung, X. Cai, K. M. Liew, and A. Buldum, "Atomistic calculation of elastic
moduli in strained silicon," Semiconductor science and technology, vol. 21, p. 906, 2006.
[48] M. Amabili, Nonlinear vibrations and stability of shells and plates: Cambridge University Press,
2008.
[49] J. N. Reddy, Mechanics of laminated composite plates and shells: theory and analysis: CRC press,
2004.
[50] L.-L. Ke, Y.-S. Wang, J. Yang, and S. Kitipornchai, "Free vibration of size-dependent Mindlin
microplates based on the modified couple stress theory," Journal of Sound and Vibration, vol. 331,
pp. 94-106, 2012.

22
Figure 1: Schematic view of a nanoplate, coordinate system of the plate and geometry
Figure 2: Effect of nanoplate thickness on the frequency response curves of Al nanoplates with various
edge supports ( ⁄ ⁄ )
Figure 3: Effect of nanoplate thickness on the frequency response curves of Si nanoplates with various
edge supports ( ⁄ ⁄ )
Figure 4: Effect of the surface residual stress on the frequency response curves of Al nanoplates with
various edge supports ( ⁄ ⁄ )
Figure 5: Effect of the surface residual stress on the frequency response curves of Si nanoplates with
various edge supports ( ⁄ ⁄ )
Figure 6: Effect of the surface elasticity constant on the frequency response curves of Al nanoplates
with various edge supports ( ⁄ ⁄ )
Figure 7: Effect of the surface elasticity constant on the frequency response curves of Si nanoplates
with various edge supports ( ⁄ ⁄ )
Figure 8: Effect of the surface elasticity constant on the frequency response curves of Al nanoplates
with various edge supports ( ⁄ ⁄ )
Figure 9: Effect of the surface elasticity constant on the frequency response curves of Si nanoplates
with various edge supports ( ⁄ ⁄ )
Figure 10: Effect of the surface mass density on the frequency response curves of Si nanoplates with
clamped boundary conditions ( ⁄ ⁄ )
Figure 11: Effect of the length-to-thickness ratio ⁄ on the frequency response curves of Si nanoplates
with various boundary conditions ( ⁄ )
Figure 12: Nonlinear frequency response of Al and Si nanoplates with different boundary conditions (
⁄ ⁄ )
Figure 13: Effect of the width-to-thickness ratio ⁄ on the frequency response curves of Si nanoplates
with various boundary conditions ( ⁄ )

23
Tables:

Table 1: Convergence and accuracy of the nondimensional linear frequency of SSSS nanoplates including
the surface stress effects
Numbers of grid points
Ref. [45]
5 7 9 11 13 15 17
24.9480 25.1613 25.1556 25.1536 25.1525 25.1518 25.1514 25.6736

24
Figures:

𝑥
𝑧𝑤

𝑦
Surface layers
Bulk layer

Figure 1: Schematic view of a nanoplate, coordinate system of the plate and geometry

25
(a) SSSS (b) CSCS
1.6 1.2
1.1 L = 0.974
1.4
1 L = 0.816

1.2 0.9 L = 0.739


0.8  = 0.663

Amplitude ( wmax )
Amplitude ( wmax )

L
1 h = 1 nm, L = 0.883 0.7  = 0.592
L
h = 2 nm, L = 0.712 0.6
0.8 L = 0.565
h = 3 nm, L = 0.628 0.5 L = 0.504
0.6 h = 5 nm, L = 0.542
0.4
h = 10 nm, L = 0.46
0.4 0.3
h = 15 nm, L = 0.428
0.2
0.2 Classic, L = 0.351
0.1
0 0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6
Frequency ratio (  /L ) Frequency ratio (  /L )

(c) CCCC
0.9
L = 1.058
0.8 L = 0.908
0.7 L = 0.836
 = 0.764
Amplitude ( wmax )

0.6 L
 = 0.699
L
0.5
L = 0.675
0.4
L = 0.62
0.3

0.2

0.1

0.7 0.8 0.9 1 1.1 1.2 1.3


Frequency ratio (  /L )

Figure 2: Effect of nanoplate thickness on the frequency response curves of Al nanoplates with various
edge supports ( ⁄ ⁄ )

26
(a) SSSS (b) CSCS
1.6
h = 1 nm, L = 0.4888 1.1 L = 0.5754
1.4 h = 2 nm, L = 0.4339 1 L = 0.5466

1.2 h = 3 nm, L = 0.4105 0.9 L = 0.5346


h = 5 nm, L = 0.3892 0.8 L = 0.524

Amplitude ( wmax )
Amplitude ( wmax )

1 h = 10 nm, L = 0.3713 L = 0.5153


0.7
0.8 h = 15 nm, L = 0.365 0.6 L = 0.5123
Classic, L = 0.3515 0.5 L = 0.5059
0.6
0.4
0.4 0.3
0.2
0.2
0.1
0
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5
Frequency ratio (  /L ) Frequency ratio (  /L )

(c) CCCC
0.9
L = 0.6507
0.8 L = 0.64

0.7 L = 0.6353
L = 0.631
Amplitude ( wmax )

0.6
L = 0.6275
0.5 L = 0.6263

0.4 L = 0.6237

0.3

0.2

0.1

0.7 0.8 0.9 1 1.1 1.2 1.3 1.4


Frequency ratio (  /L )

Figure 3: Effect of nanoplate thickness on the frequency response curves of Si nanoplates with various
edge supports ( ⁄ ⁄ )

27
(a) SSSS (b) CSCS
1.2
1.6 L = 0.8548
1.1
1.4 1 L = 0.8116
0.9 L = 0.7659
1.2
0.8  = 0.7171

Amplitude ( wmax )
Amplitude ( wmax )

L
1 0.7  = 0.6644
L
0.6
0.8  = 0.4 N/m, L = 0.6604
s

0.5
0.6 s = 0.2 N/m, L = 0.6074
0.4
s = 0 N/m, L = 0.5493
0.4 0.3
s = -0.2 N/m, L = 0.4843 0.2
0.2  = -0.4 N/m, L = 0.4091
s
0.1

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6
Frequency ratio (  /L ) Frequency ratio (  /L )

(c) CCCC
0.9 L = 1.0121
0.8 L = 0.9735

0.7 L = 0.933
 = 0.8904
Amplitude ( wmax )

0.6 L
 = 0.8453
L
0.5

0.4

0.3

0.2

0.1

0.8 0.85 0.9 0.95 1 1.05 1.1 1.15 1.2 1.25 1.3
Frequency ratio (  /L )

Figure 4: Effect of the surface residual stress on the frequency response curves of Al nanoplates with
various edge supports ( ⁄ ⁄ )

28
(a) SSSS (b) CSCS
1.1
1.4 s = 0.4 N/m, L = 0.5927 L = 0.8062
1
 = 0.2 N/m,
s
L = 0.5725 L = 0.7903
1.2 0.9
 = 0 N/m,
s
L = 0.5515 L = 0.774
0.8
1  = 0.7574

Amplitude ( wmax )
Amplitude ( wmax )

s = -0.2 N/m, L = 0.5298 0.7 L

 = -0.4 N/m, L = 0.5071


s  = 0.7404
0.8 0.6 L

0.5
0.6
0.4
0.4 0.3
0.2
0.2
0.1
0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 0.8 0.9 1 1.1 1.2 1.3 1.4
Frequency ratio (  /L ) Frequency ratio (  /L )

(c) CCCC

0.8 L = 0.9735
L = 0.9594
0.7
 = 0.9451
L
0.6
 = 0.9305
Amplitude ( wmax )

L
0.5  = 0.9157
L

0.4

0.3

0.2

0.1

0.85 0.9 0.95 1 1.05 1.1 1.15 1.2


Frequency ratio (  /L )

Figure 5: Effect of the surface residual stress on the frequency response curves of Si nanoplates with
various edge supports ( ⁄ ⁄ )

29
(a) SSSS (b) CSCS
1.2
s = -10 N/m, L = 0.477 L = 0.673
1.4
 = -5 N/m, L = 0.515
s
1 L = 0.722
1.2
 = 0 N/m, L = 0.549
s
L = 0.766
0.8  = 0.806
Amplitude ( wmax )

s = 5 N/m, L = 0.581

Amplitude ( wmax )
1 L
 = 10 N/m, L = 0.611
s
 = 0.843
L
0.8
0.6
0.6
0.4
0.4

0.2 0.2

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 0


0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6
Frequency ratio (  /L )
Frequency ratio (  /L )

(c) CCCC

0.9 L = 0.823
L = 0.881
0.8
L = 0.933
0.7
 = 0.98
Amplitude ( wmax )

L
0.6
 = 1.023
L
0.5
0.4
0.3
0.2
0.1

0.8 0.9 1 1.1 1.2 1.3


Frequency ratio (  /L )

Figure 6: Effect of the surface elasticity constant on the frequency response curves of Al nanoplates
with various edge supports ( ⁄ ⁄ )

30
(b) CSCS
(a) SSSS
1.2
1.4 L = 0.743
s = -10 N/m, L = 0.527
1 L = 0.759
1.2 s = -5 N/m, L = 0.54
L = 0.774
s = 0 N/m, L = 0.552
1 0.8  = 0.789

Amplitude ( wmax )
L
Amplitude ( wmax )

s = 5 N/m, L = 0.563
 = 0.803
L
0.8 s = 10 N/m, L = 0.575 0.6

0.6
0.4

0.4
0.2
0.2
0
0.8 0.9 1 1.1 1.2 1.3 1.4
0.6 0.8 1 1.2 1.4 1.6 1.8 2 Frequency ratio (  /L )
Frequency ratio (  /L )

(c) CCCC

0.8 L = 0.908
L = 0.927
0.7
 = 0.945
L
0.6  = 0.963
Amplitude ( wmax )

L
0.5  = 0.98
L

0.4

0.3

0.2

0.1

0.8 0.85 0.9 0.95 1 1.05 1.1 1.15 1.2


Frequency ratio (  /L )

Figure 7: Effect of the surface elasticity constant on the frequency response curves of Si nanoplates
with various edge supports ( ⁄ ⁄ )

31
(a) SSSS (b) CSCS
1.6 1.2
s = -5 N/m, L = 0.477 L = 0.673
1.4 L = 0.731
 = -2 N/m, L = 0.522
s
1
1.2 s = 0 N/m, L = 0.549 L = 0.766
0.8  = 0.798

Amplitude ( wmax )
Amplitude ( wmax )

s = 2 N/m, L = 0.575 L
1
 = 0.843
 = 5 N/m, L = 0.611
s
L
0.8 0.6

0.6
0.4
0.4
0.2
0.2

0 0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6
Frequency ratio (  /L ) Frequency ratio (  /L )

(c) CCCC
1
L = 0.823
0.9
L = 0.892
0.8
L = 0.933
0.7
 = 0.971
Amplitude ( wmax )

L
0.6
 = 1.024
L
0.5
0.4
0.3
0.2
0.1
0
0.8 0.85 0.9 0.95 1 1.05 1.1 1.15 1.2 1.25 1.3
Frequency ratio (  /L )

Figure 8: Effect of the surface elasticity constant on the frequency response curves of Al nanoplates
with various edge supports ( ⁄ ⁄ )

32
(a) SSSS (b) CSCS
1.4
s = -5 N/m, L = 0.527 1 L = 0.743

1.2  = -2 N/m, L = 0.542


s L = 0.762

s = 0 N/m, L = 0.552 0.8 L = 0.774


1
L = 0.786

Amplitude ( wmax )
s = 2 N/m, L = 0.561
Amplitude ( wmax )

0.8 s = 5 N/m, L = 0.575 0.6 L = 0.804

0.6
0.4
0.4
0.2
0.2

0
0.6 0.8 1 1.2 1.4 1.6 1.8 2 0.8 0.9 1 1.1 1.2 1.3 1.4
Frequency ratio (  /L ) Frequency ratio (  /L )

(c) CCCC

0.8 L = 0.908
L = 0.93
0.7
L = 0.945
0.6
 = 0.959
Amplitude ( wmax )

L
0.5  = 0.98
L

0.4

0.3

0.2

0.1

0.8 0.85 0.9 0.95 1 1.05 1.1 1.15 1.2


Frequency ratio (  /L )

Figure 9: Effect of the surface elasticity constant on the frequency response curves of Si nanoplates
with various edge supports ( ⁄ ⁄ )

33
Figure 10: Effect of the surface mass density on the frequency response curves of Si nanoplates with
clamped boundary conditions ( ⁄ ⁄ )

34
(a) SSSS (b) CSCS
1.6
a/h = 10, L = 0.565 L = 0.758
1.2
1.4 a/h = 12, L = 0.495 L = 0.66
a/h = 14, L = 0.446 1 L = 0.589
1.2
a/h = 16, L = 0.411  = 0.535
Amplitude ( wmax )

Amplitude ( wmax )
L
1 0.8
a/h = 20, L = 0.363  = 0.46
L
0.8
0.6
0.6
0.4
0.4

0.2 0.2

0.6 0.8 1 1.2 1.4 1.6 1.8 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
Frequency ratio (  /L ) Frequency ratio (  /L )

(c) CCCC

L = 0.911
1
L = 0.791
L = 0.703
0.8
 = 0.635
Amplitude ( wmax )

L
 = 0.54
0.6 L

0.4

0.2

0.7 0.8 0.9 1 1.1 1.2 1.3


Frequency ratio (  /L )

Figure 11: Effect of the length-to-thickness ratio ⁄ on the frequency response curves of Si nanoplates
with various boundary conditions ( ⁄ )

35
(a) Al nanoplate (b) Si nanoplate
0.9 1.4
L = 0.628, SSSS L = 0.4105, SSSS
0.8
L = 0.739, CSCS 1.2 L = 0.5346, CSCS
0.7 L = 0.836, CCCC L = 0.6353, CCCC
1
0.6
Amplitude ( wmax )

Amplitude ( wmax )
0.5 0.8

0.4 0.6
0.3
0.4
0.2
0.2
0.1

0 0
0.8 0.85 0.9 0.95 1 1.05 1.1 1.15 1.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
Frequency ratio (  /L ) Frequency ratio (  /L )

Figure 12: Nonlinear frequency response of Si and Al nanoplates with different boundary conditions (
⁄ ⁄ )

36
(a) SSSS (b) CSCS
1
1.2 b/h = 2, L = 4.9856 L = 5.0023
b/h = 3, L = 2.6748 L = 2.7149
1 0.8
b/h = 5, L = 1.2508 L = 1.3433
b/h = 10, L = 0.5709  = 0.7507
Amplitude ( wmax )

Amplitude ( wmax )
0.8 L
0.6

0.6
0.4
0.4

0.2
0.2

0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
Frequency ratio (  /L ) Frequency ratio (  /L )

(c) CCCC
0.8
L = 7.2328
0.7 L = 4.2645

0.6 L = 2.0701
L = 0.8947
Amplitude ( wmax )

0.5

0.4

0.3

0.2

0.1

0.6 0.7 0.8 0.9 1 1.1 1.2 1.3


Frequency ratio (  /L )

Figure 13: Effect of the width-to-thickness ratio ⁄ on the frequency response curves of Si nanoplates
with various boundary conditions ( ⁄ )

37

S-ar putea să vă placă și