Você está na página 1de 6

Physica E 67 (2015) 12–17

Contents lists available at ScienceDirect

Physica E
journal homepage: www.elsevier.com/locate/physe

Size-dependent resonance and buckling behavior of nanoplates with


high-order surface stress effects
Chih-Hao Cheng, Tungyang Chen n
Department of Civil Engineering, National Cheng Kung University, Tainan 70101, Taiwan

H I G H L I G H T S G R A P H I C A L A B S T R A C T

 We derive the resonance frequency This figure presents the compressive buckling force for a simply-supported circular nanoplate. Our
and buckling load for nanoplates model demonstrates that the high-order surface effect can be significant.
with a high-order surface stress
model.
 Compared to conventional surface
stress model, our model shows that
the high-order surface stress effect
could be significant.

art ic l e i nf o a b s t r a c t

Article history: This work presents a theoretical study of the resonance frequency and buckling load of nanoplates with
Received 12 September 2014 high-order surface stress model. A classical thin plate theory based on Kirchhoff–Love assumption is
Accepted 31 October 2014 implemented with surface effects. Circular and rectangular nanoplates with simply supported end
Available online 6 November 2014
conditions are exemplified. The size-dependent solutions are compared with the simplified solutions
Keywords: based on simple surface stress model, and also on the classical theory of elasticity. We aim to explore the
High-order surface stress scope of applicability so that the modified continuum mechanics model could serve as a refined ap-
Size-dependent behavior proach in the prediction of mechanical behavior of nanoplates.
Nanoplates & 2014 Elsevier B.V. All rights reserved.

1. Introduction [2–6]. The approach is mainly based on the classical continuum


mechanics incorporated with surface effects to simulate the size-
Nanoplates (NPs) and nanowires (NWs) are typical nanos- dependent behavior in nanoscaled solids. Specifically, the surface
tructures in nanoelectromechanical systems. From mechanics effect is often modeled by a thin membrane assumption [7–9],
viewpoints, NPs can be viewed as a two-dimensional analog of assuming that the surface effect is acting like a membrane sub-
NWs, as the former can be modeled as a thin plate, while the latter jected to in-plane stress. The jump condition along the interface
is often modeled as a one-dimensional beam. Continuum me- was known as the generalized Young–Laplace equation. In many
chanics approaches [1], with suitable implementations, have been cases the prediction is fairly satisfactory and efficient compared
demonstrated as useful tools to model the mechanical behavior with the atomistic analysis or molecular dynamics calculation [10].
of nanostructures. In the last years, quite a few studies have been In our previous study [25] on buckling and resonance behavior of
focused on the mechanical behavior of NWs, see for example NWs, however, we found that there are situations where the nu-
merical prediction by the pure surface stress model is not able to
n
Corresponding author. capture the general trend of the experimental data, especially for
E-mail address: tchen@mail.ncku.edu.tw (T. Chen).

http://dx.doi.org/10.1016/j.physe.2014.10.040
1386-9477/& 2014 Elsevier B.V. All rights reserved.
C.-H. Cheng, T. Chen / Physica E 67 (2015) 12–17 13

small sizes of NWs. In contrast, the high-order surface effect model is the normal direction measured linearly from the surface. The
will give to a good simulation with the experimental results. coefficients h1, h2 , h3, with h3 = 1, are the metric coefficients of the
Motivated by the lack of sufficient accuracy in NWs, here we coordinate system.
implement the high-order surface model in the modeling of We first consider the resonance frequency of NPs with the
buckling load and resonance frequency of NPs. High-order surface high-order surface stress model. For a circular NP, letting
stress model [11,12] is a refined model of the membrane surface (v1, v2, v3 ) → (r , θ , z) and (h1, h2, h3 ) → (1, r , 1) in (1)–(3), the
stress model, accounting into the flexural rigidity of the mem- traction jump condition (3) in the normal direction (z-direction)
brane. The main difference is that the in-plane surface stress could can be written as
be varying across the thin layer thickness and thus, equivalently, it
is equivalent to a resultant in-plane surface force together with a ⎛σ s σ s ⎞ ⎛ 2 ∂mrs ∂ 2mrs 2 ∂ 2mrsθ 1 ∂mθs ⎞
σ zz(m) − σ zz(i) = − ⎜ r + θ ⎟ − ⎜⎜ + + − ⎟⎟
resultant surface moment. As such the model that incorporates ⎝ Rr R θ ⎠ ⎝ r ∂r ∂r 2 r ∂r ∂θ r ∂r ⎠
surface moment is referred to as high-order surface stresses
⎛ 2 ∂m s 1 ∂ 2mθs ⎞
model, as it includes microstrain as well as curvature along the − ⎜⎜ rθ
+ ⎟⎟,
interface. We mentioned that the conventional surface stress ⎝ r ∂θ
2 r 2 ∂θ 2 ⎠ (4)
theory, proposed by Gurtin and Murdoch [7,8], postulated that the
where Rr and R θ are respectively the radii of principal curvatures
surface or interface can be modeled as a thin membrane that can
along the radial and tangential directions. To consider the surface
sustain only in-plane stresses, and thus can be viewed as a sim-
stress effects with residual surface tension, we introduce the sur-
plified model of the high-order one. We find interestingly that the
face constitutive relation [8]
governing frameworks of both models do not differ much. Yet the
s
numerical calculations demonstrate that, depending on the re- σαβ = τ0 δ αβ + (μ s − τ0 )(uαs, β + u βs, α ) + (λ s + τ0 ) u γs, γ δ αβ + τ0 uαs, β , (5)
lative size of the plate, the high-order effect can be significant and
should not be ignored in certain cases. In illustration, circular and in which λ s and μs are surface Lame constants, τ0 is the residual
rectangular nanoplates with simply supported are exemplified. surface tension and uαs is the surface displacement component. The
Analytic and numerical solutions of the derived results are com- Greek indices here take the numbers of 1–2 to describe the in-
pared with the simplified solutions based on conventional surface plane coordinates. As in [11], the surface stress and stress couples
stress model and on the classical elasticity solutions. It is our aim can be integrated from (5) through the thickness t of the isotropic
to explore the scope of applicability to the analysis of nanos- interphase layer (with Young's modulus Ec and Poisson's ratio νc ).
tructures that the refined continuum mechanics model could be a As such the resultant force and couple (surface stress and surface
feasible approach in the estimate of mechanical behavior of na- moment) can be integrated as
noplates. In the literature the related studies on nanoplates in- ⎧σ s ⎫ ⎡ 1 vc ⎧ 0⎫
0 ⎤ ⎪ εr ⎪
clude the works [13–23], in which the formulation is mainly based ⎪ r ⎪ ⎢ ⎥⎪ ⎪
on simple surface stress model. We mention that the high-order ⎨ σθ ⎬ = E s ⎢ vc 1
s
0 ⎥ ⎨ σθ0 ⎬ ,
⎪ s⎪ ⎢ 0 0 1 − v ⎥⎪ ⎪
surface stresses model has been utilized in the modeling of NWs ⎩ σ rθ ⎭ ⎣ c ⎦⎪
⎩ σ rθ ⎪
0

[24,25] and also in the bending behavior of NPs [11].
⎧ms ⎫ ⎡ 1 vc ⎧ ⎫
0 ⎤ ⎪ κ r0 ⎪
⎪ r ⎪ ⎢ ⎥⎪ ⎪
⎨ mθs ⎬ = − Ds ⎢ vc 1 0 ⎥ ⎨ κ θ0 ⎬ ,
⎪ s⎪ ⎢ 0 0 1 − v ⎥⎪ ⎪
c ⎦⎪
⎩ κ rθ ⎪
2. Resonance frequency of nanoplates ⎩ m rθ ⎭ ⎣ 0
⎭ (6)
The high-order surface stress model incorporates the effect of where we have defined
in-plane surface stress σαjs as well as surface moment mαβ
s
along the
interface. By the balance of force and moment, the traction jumps Ec t 1 Ec t 3
Es = , Ds = .
across the interface, between two different regions denoted by 1 − νc2 12 1 − νc2 (7)
superscripts (i) and (m), is given by [11]
0
In the formulation, the mid-plane curvature of thin layer καβ and
1 ⎡ ∂ (h 2 σ12
s
) s
∂ (h1σ22 ) s ∂h1 s ∂h 2

0
(m)
σ32 − (i )
σ32 =− ⎢ + − σ11 + σ21 ⎥ the mid-plane strain εαβ are taken to be the same as those of the
h1h 2 ⎣ ∂v1 ∂v2 ∂v2 ∂v1 ⎦
nanoplate.
1 ⎡ ∂ (h 2 m12
s
) ∂ (h1m22s
) s ∂h1 s ∂h 2

To proceed, substituting (6) into (4), invoking that (1/Rr ) = κr ,
+ ⎢ + − m11 + m21 ⎥,
R 2 h1h 2 ⎣ ∂v1 ∂v2 ∂v2 ∂v1 ⎦ (1) (1/R θ ) = κθ , the jump condition (4) is equivalent to a transverse
distributed load, denoted by q⁎ (r , θ)
1 ⎡ ∂ (h 2 σ11
s
) s
∂ (h1σ21 ) s ∂h1 s ∂h 2

(m)
σ31 (i )
− σ31 =− ⎢ + + σ12 − σ22 ⎥
q (r , θ) = 2 ⎡⎣σ zz(m) − σ zz(i) ⎤⎦ = 2τ0 ∇2w − 2Ds ∇2 (∇2w).
h1h 2 ⎣ ∂v1 ∂v2 ∂v2 ∂v1 ⎦ ⁎
(8)
1 ⎡ ∂ (h 2 m11
s
) ∂ (h1m21s
) s ∂h1 s ∂h 2

+ ⎢ + + m12 − m22 ⎥,
R1h1h 2 ⎣ ∂v1 ∂v2 ∂v2 ∂v1 ⎦ (2) Here the factor ‘2’ comes from the parts of top and bottom surface
layers.
⎛σ s σs ⎞ We find that the NP, incorporating the high-order surface stress
(m)
σ33 (i)
− σ33 = − ⎜ 11 + 22 ⎟ and residual surface stress, under the free vibration is governed by
⎝ R1 R2 ⎠
⎧ ∂ ⎧ 1 ⎡ ∂ (h m s ) ∂ (h1m21s ⎤⎫ ⎫ ∂ 2w
1 ⎪ ) s ∂h1 s ∂h2 ⎬ ⎬
2 11 ⎪ ⁎
− ⎨ ⎨ ⎢ + + m12 − m22 ⎥ ⎪ (D + 2Ds ) ∇4 w − 2τ0 ∇2w + ρh = 0.
h1h2 ⎩ ∂v1 ⎩ h1 ⎣ ∂v1
⎪ ⎪
∂v2 ∂v2 ∂v1 ⎦ ⎭ ⎭ ⎪
∂t 2 (9)

1 ⎧ ⎪ ∂
⎧ 1 ⎡ ∂ (h2 m s ) ∂ (h1m22s
) s ∂h1
⎤⎫ ⎫
s ∂h2 ⎬ ⎬

Here D⁎ is the effective flexural rigidity covering the effect of the
− ⎨ ⎨ ⎢ 12
+ − m11 + m21 ⎥ ⎪. bulk and two thin surface layers based on classical lamination
h1h2 ⎩ ∂v2 ⎩ h2 ⎣
⎪ ⎪
∂v1 ∂v2 ∂v2 ∂v1 ⎦ ⎭ ⎭ (3)

theory [26]
Here an orthogonal curvilinear coordinate (ν1, ν2, ν3 ) is used
⁎ E s h2 Ec h3
to describe the curved interface, in which ν1 and ν2 are two para- D =D+ , D= .
2(1 − vs 2) 12(1 − vc 2)
metric curves describing the coordinate lines on the surface, and ν3 (10)
14 C.-H. Cheng, T. Chen / Physica E 67 (2015) 12–17

where vs is Poisson ratio of surface. When high-order effect is where a and b are the side lengths of the plate. This together with
neglected, Eq. (9) will reduce to the one formulated by the the simply-supported end conditions
generalized Young–Laplace equation [16] and, if the surface term
is further neglected, the corresponding equation of classical W (x, y) x = 0, x = a = 0,
elasticity ([27, pp. 307–308]) will be recovered. W (x , y ) = 0,
y = 0, x = b
To solve (9), by letting w = W (r , θ) eiωt and using the method of
separation of variables, the general solution of the deflection W ∂ 2W (x, y)
= 0,
can be expressed as ∂x2 x = 0, x = a

∞ ∂ 2W (x, y)
⎡⎣A m J (ζm r) + Cm Im (η r) ⎤⎦ cos mθ = 0,
W (r , θ ) = ∑ m m ∂y2 x = 0, x = b (18)
m=0

⎡⎣A ⁎ J (ζm r) + C ⁎ Im (η r) ⎤⎦ sin mθ , will give the natural frequency ω with double suffix m and n as
+ ∑ m m m m
m=0 (11) ⎡ ⎤
D + 2Ds ⎢ 4 ⎛ m2 n2 ⎞ ⎛ m2 n2 ⎞ ⎥
⁎ 2
ωmn = π ⎜ + ⎟ + π 2β 2 ⎜ + ⎟ .
where ρh ⎢ ⎝ a2 b2 ⎠ ⎝ a2 b2 ⎠ ⎥⎦
⎣ (19)

β2 − β 4 + 4λ m4 β2 + β 4 + 4λ m4 Again, we can reduce to the simplified solutions for the cases of


ζm2 = − , ηm2 = , conventional surface stress model based on Young–Laplace equa-
2 2 (12)
tions and the classical elasticity solutions (without surface effect)
and ([28, pp.334–335]).

2τ0 ρhωm2
β2 = ⁎ , λ m4 = .
D + 2D s D⁎ + 2D s (13)
3. Buckling of nanoplates
Here Jm (x) and Im (x) are respectively the Bessel function and
modified Bessel function of the first kind of order m. For a simply- We now consider the buckling of NP with the high-order sur-
supported circular plate, the end conditions face stress model. The equilibrium equations of NP under buckling
load based on Kirchhoff–Love thin plate theory are given as [29]
W (r , θ) r = R = Mr (r , θ) r = R = 0, (14) d
(rMr ) − Mθ − rQ r + rq⁎ = 0,
dr
together with (11) will give
1 d 1 d ⎛ dw ⎞
(rQ r ) + N ⎜r ⎟ = 0.
2
r dr r dr ⎝ dr ⎠ (20)
ζmn Jm − 1 (ζmn R) Im − 1 ( β 2 + ζmn R)
− +
2
β 2 + ζmn Jm (ζmn R) 2
Im ( β 2 + ζmn R) To proceed, one can write the bending moments and shearing
forces in terms of w [26,11, Eqs. (60) and (61)]. The circular na-
2
β 4 + 4λ mn R noplate incorporating the effects of residual surface stresses and
=
β2 + 2
ζmn 1 − vc (15) high-order surface stresses under buckling load can be formulated
as
Eq. (15) is a transcendental equation for ζmn that may have an
(D⁎ + 2Ds ) ∇4 w − (N + 2τ0 ) ∇2w = 0. (21)
infinite number of solutions. Note that in (15) we have used double
subscripts to designate the values of ζmn and λ mn , where the first We mention that when the high-order surface effect are ne-
index m refers to the summation term m in (11), while the second glected, i.e. Ds = 0, Eq. (21) will reduce to the one based on the
index n designates the nth solution corresponding to the first in- conventional surface stress model. If both the residual surface
dex m. The unknown ζmn , governed by (15), can be solved from stresses effect and the high-order surface stresses effect are ne-
numerical methods. The corresponding natural frequency ωmn can glected, i.e. D⁎ = D and τ0 = Ds = 0 , the governing equation in (21)
then be obtained from (12) and (13) as will recover the corresponding elasticity equation [29, pp. 389–
391]. We mention that a similar study with the simple surface

(D + 2Ds )(ζmn2 + β 2) model, incorporating the piezoelectric effect, was considered in
ωmn = ζmn.
ρh (16) [31]. Our simplified result without the high-order effect will agree
with the solution of [31] without the piezoelectric effect.
We mention that when the high-order surface effect is ne- To solve the field equation, we can set
glected, i.e.Ds = 0, the transcendental Eq. (15) and natural fre-
∇2w = V (r), (22)
quency (16) will reduce to these by the conventional surface stress
model. In addition, if the surface stress effects are further ignored, so that (21) becomes
they will recover the classical elasticity solution [27].
d2V 1 dV
For a rectangular plate, the deflection function W can be writ- + + ψV = 0
dr 2 r dr (23)
ten as
where
∞ ∞
mπx nπ y
W (x , y ) = ∑ ∑ Emn sin sin , N + 2τ0
.
m= 0 n= 0
a b (17) ψ=− ⁎
D + 2D s (24)
C.-H. Cheng, T. Chen / Physica E 67 (2015) 12–17 15

The solution of w can be written in terms of Bessel functions force (36) for a square nanoplate will reduce to the one based on
the classical elasticity theory [29, pp. 348–353].
V = AJ0 (r ψ ) + BY0 (r ψ ). (25)

Upon substitution back to (22), we find that


4. Numerical results
w = EJ0 (r ψ ) + FY0 (r ψ ) + c1 ln r + c2. (26)

Since the functions Y0 (r ψ ) and ln r are both singular at r = 0, We first present numerical solutions for the resonance fre-
we can let F = c1 = 0. Thus the displacement of circular nanoplate quency ωmn given in (16). The material parameters for anodic
under buckling load can be obtained as alumina are selected in our numerical calculations [16], in which
Ec = 70 GPa , Gc = 27 GPa and ρ = 2700 kg/m3. The surface proper-
w = EJ0 (r ψ ) + c2. (27) ties are Es = 5.1882 N /m and τ0 = 0.9108 N /m [17]. For the bending
For a simply-supported circular plate, the use of the end con- rigidity parameter Ds of the surface layer, a quantity varying from
ditions 103 × (10−18 Nm) to 105 × (10−18 Nm) is used in our numerical
calculations. In Table 1, we present numerical calculations of ωmn
W (r , θ) r = R = Mr (r , θ) r = R = 0 (28) associated with m = 0 for a circular nanoplate with radius
R = 100 nm based on the present high-order surface stress model,
in (27) will give and those of the simple surface stress model, based on generalized
EJ0 (R ψ ) + c2 = 0, (29) YL equation, and classical elasticity solutions. Three different
thicknesses of circular nanoplates: (a) h = 5 nm , (b) h = 7 nm and
⎡ ψ ⎤ (c) h = 10 nm are numerically calculated. Here HO–A and HO–B
E ⎢−ψJ0 (R ψ ) + (1 − νc ) J1 (R ψ ) ⎥ = 0. respectively indicate the present estimates with the bending ri-
⎣ R ⎦ (30)
gidity parameter Ds = 103 × (10−18 nm) and Ds = 104 × (10−18 nm).
The nontrivial solution is E ≠ 0. We need to find out the roots It is seen that the resonance frequency of nanoplate depends on
R ψ of the transcendental Eq. (30). The smallest root can be found the thickness of the nanoplate and is thus size-dependent, for both
as R ψ = 2.05 for νc = 0.3. After substituting the root R ψ = 2.05 the high-order surface stress model and the simple surface stress
into Eq. (24), we have model based on the generalized YL equation. Specifically when the
thickness of nanoplate is sufficiently thin, the resonance frequency
2.052 ⁎
Ncr = − (D + 2Ds ) − 2τ0. of the nanoplate exhibits significant size dependence with the
R2 (31)
decrease of the thickness, for both models. In addition, when the
Now letting Ds = 0, the solution will reduce to the corre- value of Ds is greater, the resonance frequency becomes higher.
sponding solution of surface stress. When both the residual sur- Table 2 shows the resonance frequency ω11 for a rectangular na-
face stresses effect and the high-order surface stresses effect are noplate with edge length b = 100 nm versus the different edge
neglected, i.e. D⁎ = D and τ0 = Ds = 0 , the buckling load in (31) will length ratios: (a) a/b = 1, (b) a/b = 2 and (c) a/b = 3.
reduce to the corresponding classical elasticity solution [29, pp. In Figs. 1 and 2, we present the critical buckling force derived in
389–391] (31) for a circular NP with high-order surface stress model for two
values of surface bending stiffness Ds . The critical buckling force Ncr
2.052
Ncr = − D. is presented versus the normalized radius h/R (Fig. 1). For con-
R2 (32)
venience, to compare the difference, in Figs. 1 and 2, we mark the
Now we consider the rectangular nanoplate subjected to a values of Ncr for h/R ¼0.04. We see in Fig. 1(a) that the calculated
uniaxial in-plane compression force Nx = − N and Nxy = Ny = 0. value of Ncr based on the simple surface stress model (generalized
The field equation, incorporating high-order surface stress effects Young–Laplace equation) is significantly greater than that of
and the residual surface stresses, can be derived as
Table 1
∂ 2w Comparison of the resonance frequency ωmn of a circular nanoplate associated with
(D⁎ + 2Ds ) ∇4 w − 2τ0 ∇2w + N = 0.
∂x2 (33) m = 0 and radius R = 100 nm for classical elasticity solution, generalized YL equa-
tion and high-order surface stress model versus the different thickness: (a)
The deflection can assume the form [30] h = 5 nm , (b) h = 7 nm and (c) h = 10 nm .
∞ ∞
⎛ mπx ⎞ ⎛ nπ y ⎞ ωmn (×109 ) m=0
w (x , y ) = ∑ ∑ a mn sin ⎜⎝ ⎟ sin ⎜ ⎟,

m= 1 n= 1
a ⎠ ⎝ b ⎠ (34)
n 1 2 3 4 5
where amn are unknown coefficients. This together with the end
conditions (18) will give the buckling load N: (a) h = 5 nm
Classical elasticity 3.802 22.900 57.131 106.563 171.198

⎛ m2 n2 n 4 a2 ⎞ ⎛ n2a2 ⎞ Young–Laplace equation 9.692 31.339 67.563 119.339 186.879
N = (D + 2D s ) π 2 ⎜ +2 + ⎟ + 2τ0 ⎜1 + ⎟. HO–A 11.404 47.861 112.746 206.362 328.753
⎝a2 b2 b4 m2 ⎠ ⎝ m2b2 ⎠ (35) HO–B 21.327 118.607 293.314 545.598 875.484

It is seen that the minimum value of force N occurs at n = 1 and (b) h = 7 nm


m = 1. Thus, the critical buckling force Ncr for the square plate a = b Classical elasticity 5.323 32.056 79.984 149.188 239.678
can be expressed as Young–Laplace equation 9.274 37.237 86.750 158.149 251.482
HO–A 10.573 48.179 115.519 212.736 339.848
4π 2 ⁎ HO–B 18.541 103.602 256.354 476.933 765.361
Ncr = (D + 2Ds ) + 4τ0.
b2 (36)
(c) h = 10 nm
The corresponding solution based on the generalized Young– Classical elasticity 7.604 45.794 114.262 213.125 342.396
Young–Laplace equation 9.976 48.948 118.919 219.955 352.069
Laplace equation can be directly deduced from (36) by letting
HO–A 10.843 55.229 134.963 250.104 400.662
τ0 = Ds = 0 . To our knowledge, the latter solution is not known in HO–B 16.731 94.545 234.256 436.002 699.805

the literature. When D = D and τ0 = Ds = 0 , the critical buckling
16 C.-H. Cheng, T. Chen / Physica E 67 (2015) 12–17

Table 2
Comparison of the resonance frequency ω11 of a rectangular nanoplate with edge
a 90
length b = 100 nm versus different edge length ratios for classical elasticity solu- 84.4
80
tion, generalized YL equation and high-order surface stress model: (a) a/b = 1, (b)
a/b = 2 and (c) a/b = 3. 70 classical sol.
generalized YL eq.
ω11(×109 ) 60 Ds=10-15 N.m

Ncr (N/m)
h 5 6 7 8 9 10 50 Ds=10-14 N.m

(a) a/b = 1 40
Classical elasticity 15.207 18.249 21.290 24.332 27.373 30.415
Young–Laplace equation 22.764 24.076 25.929 28.132 30.566 33.157 30
HO–A 33.097 32.568 32.934 33.944 35.426 37.256
HO–B 79.313 73.417 69.249 66.326 64.352 63.131 20
13.3
(b) a/b = 2 10
Classical elasticity 9.505 11.406 13.306 15.207 17.108 19.009
5.44
1.62
Young–Laplace equation 16.274 16.687 17.528 18.659 19.991 21.463 0
HO–A 22.143 21.595 21.640 22.115 22.911 23.946 0 0.02 0.04 0.06 0.08 0.1
HO–B 50.196 46.449 43.793 41.922 40.649 39.851
h/b
(c) a/b = 3 b 90
Classical elasticity 8.449 10.138 11.828 13.518 15.207 16.897 classical sol.
Young–Laplace equation 15.023 15.288 15.938 16.893 18.025 19.292 80 generalized YL eq.
HO–A 20.096 19.549 19.538 19.918 20.588 21.477
HO–B 44.803 41.453 39.078 37.401 36.258 35.539 70 Ds=10-15 N.m
Ds=10-14 N.m
60

Ncr (N/m) 50
a 14 40

30 26.8
12
10.4 20
10 9.03
classical sol.
Ncr (N/m)

10
generalized YL eq.
8 7.063
3.23
Ds=10-15 N.m 0
0 0.02 0.04 0.06 0.08 0.1
6 Ds=10-14 N.m
h/b
4 Fig. 2. The compressive critical buckling force of simply-supported square nano-
2.85 plates with respect to the ratio h/b for different edge lengths: (a) b = 100 nm , (b)
2.01 b = 200 nm .
2
0.17 classical elasticity (2.013 N/m versus 0.1724 N/m). The effect of
0 high-order surface stress, depending on the value of the parameter
0 0.02 0.04 0.06 0.08 0.1
h/R Ds , could be 1.5–5 times compared with that estimated by the
simple surface stress model. This reflects the significance of high-
b order effect and thus, in reference to real experimental data, the
14 classical sol. present simulation provides a freedom to characterize the
generalized YL eq. significant variation based on different continuum mechanics
12 Ds=10-15 N.m models. It is seen that high-order surface stress effects become
Ds=10-14 N.m significant when the NP is increasingly thin. It is also seen that the
10 critical buckling force exhibits size dependent behavior with
Ncr (N/m)

incorporation of residual surface tension and high-order surface


8 stresses here. In Fig. 2 we illustrate the buckling force versus dif-
ferent ratios, h/b, for a simply supported rectangular NP with the
6 effect of high-order surface stress. It is seen that the results for
4.28 circular and rectangular plates have a similar trend.
4
2.39
2
2.18 5. Conclusions
0.34
0
0 0.02 0.04 0.06 0.08 0.1 We have derived the resonance frequency and buckling load for
h/R nanoplates based on the thin plate theory implemented with high-
order surface stress effects. Specific results are derived for circular
Fig. 1. The compressive critical buckling force of simply-supported circular nano-
plates with respect to the ratio h/R for different radius, (a) R1 = 100 nm , (b) and rectangular plates with simply supported end conditions. It is
R2 = 200 nm . seen that the numerical solutions are size-dependent. Our
C.-H. Cheng, T. Chen / Physica E 67 (2015) 12–17 17

solutions are compared with the simplified solutions based on [10] V.B. Shenoy, Phys. Rev. B 71 (2005) 094104.
simple surface stress model and on the classical elasticity theory. [11] T. Chen, M.S. Chiu, Acta Mech. 225 (2014) 1043.
[12] T. Chen, M.S. Chiu, Mech. Mater. 43 (2011) 212.
The approach provides a simple continuum mechanics approach, [13] C.T. Sun, H. Zhang, J. Appl. Phys. 93 (2003) 1212.
in place of atomistic analysis or experiments, to analyze the me- [14] L.H. He, C.W. Lim, B.S. Wu, Int. J. Solids Struct. 41 (2004) 847.
chanical behavior of nanoplates in a refined manner. [15] D.W. Huang, Int. J. Solids Struct. 45 (2008) 568.
[16] A. Assadi, B. Farshi, J. Appl. Phys. 108 (2010) 074312.
[17] A. Assadi, B. Farshi, A. Alinia-Ziazi, J. Appl. Phys. 107 (2010) 124310.
[18] E. Jomehzadeh, H.R. Noori, A.R. Saidi, Physica E 43 (2011) 877.
Acknowledgments [19] R. Ansari, S. Sahmani, Int. J. Eng. Sci. 49 (2011) 1204.
[20] R. Ansari, R. Gholami, M.F. Shojaei, V. Mohammadi, S. Sahmani, J. Appl. Mech.
80 (2013) 021021.
This work was supported by the Ministry of Science and [21] K.F. Wang, B.L. Wang, Finite Elem. Anal. Des. 74 (2013) 22.
Technology, Taiwan, under Contract NSC 102-2221-E-006-151- [22] L.L. Zhang, J.X. Liu, X.Q. Fang, G.Q. Nie, Eur. J. Mech.—A/Solid 46 (2014) 22.
[23] L.L. Zhang, J.X. Liu, X.Q. Fang, G.Q. Nie, Physica E 57 (2014) 169.
MY3. [24] M.S. Chiu, T. Chen, Physica E 44 (2011) 714.
[25] M.S. Chiu, T. Chen, Acta Mech. 223 (2012) 1473.
[26] J.N. Reddy, Theory and Analysis of Elastic Plates and Shells, CRC Press, Florida,
2006.
References
[27] R. Courant, D. Hilber, Methods of Mathematical Physics, Interscience Pub-
lishers, New York, 1953.
[1] R.E. Miller, V.B. Shenoy, Nanotechnology 11 (2000) 139. [28] S.P. Timoshenko, S. Woinowsky-Krieger, Theory of Plates and Shells, McGraw-
[2] G.F. Wang, X.Q. Feng, Appl. Phys. Lett. 94 (2007) 231904. Hill, New York, 1959.
[3] J. He, C.M. Lilley, Nano Lett. 8 (2008) 1798. [29] S.P. Timoshenko, J.M. Gere, Theory of Elastic Stability, McGraw-Hill, New York,
[4] J. He, C.M. Lilley, Appl. Phys. Lett. 93 (2008) 263108. 1961.
[5] L.Y. Jiang, Z. Yan, Physica E: Low-Dimens. Syst. Nanostruct. 42 (2010) 2274. [30] A.C. Ugural, Stress in Plates and Shells, McGraw-Hill, New York, 1999.
[6] G.F. Wang, F. Yang, J. Appl. Phys. 109 (2011) 063535. [31] A. Farajpour, M. Dehghany, A.R. Shahidi, Compos. Eng. 50 (2013) 333.
[7] M.E. Gurtin, A.I. Murdoch, Arch. Ration. Mech. Anal. 57 (1975) 291.
[8] M.E. Gurtin, A.I. Murdoch, Int. J. Solids Struct. 14 (1978) 431.
[9] T. Chen, M.S. Chiu, C.N. Weng, J. Appl. Phys. 100 (2006) 074308.

Você também pode gostar