You are on page 1of 24

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

net/publication/322015816

Effect of surface elasticity on extensional and torsional stiffnesses of isotropic


circular nanorods

Article  in  Mathematics and Mechanics of Solids · December 2017


DOI: 10.1177/1081286517753719

CITATIONS READS
7 486

2 authors:

Prakhar Gupta Ajeet Kumar


Indian Institute of Technology Hyderabad Indian Institute of Technology Delhi
7 PUBLICATIONS   53 CITATIONS    28 PUBLICATIONS   377 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Large deformations in one-dimensional flexoelectric structures View project

Incorporating plasticity within the theory of rods View project

All content following this page was uploaded by Prakhar Gupta on 09 December 2018.

The user has requested enhancement of the downloaded file.


Effect of surface elasticity on extensional and torsional
stiffnesses of isotropic circular nanorods
Prakhar Gupta and Ajeet Kumar
Department of Applied Mechanics, IIT Delhi, Hauz Khas, New Delhi, India, 110016
Correspondence: ajeetk@am.iitd.ac.in

Abstract. We present a continuum formulation to obtain the effects of surface residual stress
and surface elastic constants on extensional and torsional stiffnesses of isotropic circular nanorods.
Analytical expressions of axial force, twisting moment, extensional and torsional stiffnesses are ob-
tained. Unlike the case of rectangular nanorods, we show that the stiffnesses of circular nanorods
also depend on surface residual stress components. This is attributed to non-zero surface curvature
inherent in circular nanorods. We further normalize these expressions and analyze their asymptotic
limits in the limit of the nanorod’s radius approaching both zero and infinity which correspond to
surface dominated and bulk dominated regimes respectively. Finally, we use the recently proposed
Helical Cauchy-Born rule and perform molecular statics calculations to obtain axial force, twisting
moment and stiffnesses of the tungsten nanorod. The tungsten material is selected since its bulk
crystal exhibits isotropy in the stress-free state. The results from molecular statics calculations are
shown to match accurately with the derived continuum formulae.

Keywords: surface residual stress, surface elasticity, nanorod, isotropy, Cauchy-Born rule.

1 Introduction
One-dimensional nanostructures have gained increased attention due to their wide applicability in
various nanodevices such as sensors, transistors, AFM probes, actuators, resonators in nanoelec-
tromechanical systems etc. (Cui et al., 2001, 2003). It is also well known that the mechanical
properties of nanorods change dramatically due to surface effects inherent in them (Wong et al.,
1997; Cuenot et al., 2004; Jing et al., 2006). Gurtin and Murdoch (1975) were one of the first
to propose a general continuum theory for three-dimensional solids which also incorporated sur-
face elasticity. They propose the surface energy to be dependent on in-plane stretch and shear.
Steignmann and Ogden (1997) propose a more general form for surface energy which also depends
on surface curvature in addition to in-plane stretch and shear. To account for surface effects in
bulk as well as one-dimensional nanostructures in a multiscale framework, Park et al. (2006) pro-
posed a surface Cauchy-Born rule where they consider the surface and bulk energies separately.
Recently, Kumar et al. (2016) proposed a Helical Cauchy-Born rule for modeling deformations of
one-dimensional nanostructures where they treat the bulk and surface parts of the nanostructure
as a single entity. This concept is applicable for not just large radii nanorods but also for small
radii nanorods for which the surface and bulk regions essentially intermingle. Several authors have
investigated the effect of surface elastic constants on resonance frequency as well as on bending,
twisting and extensional stiffnesses of nanorods (Miller and Shenoy, 2000; Shenoy, 2005; On et al.,

1
2010). Miller and Shenoy (2000) obtained simple continuum formulae for extensional and bend-
ing stiffnesses of square nanorods whereas Shenoy (2002) obtained formula for twisting stiffness of
square nanorods. They show these stiffnesses to be dependent on surface and bulk elastic constants
but independent of surface residual stress components. Later on, Wang et al. (2010) show the de-
pendence of the surface tension term too on the square nanorod’s bending stiffness. Many authors
(Li et al., 2011; Wang et al., 2013; Lu et al., 2014; Thongyothee and Chucheepsakul, 2013, 2015)
have incorrectly extended the results of Miller and Shenoy (2000) to circular nanorods. Zhang et
al. (2008) derived the formula for extensional stiffness of circular nanorods but their expression also
does not show any dependence on surface residual stress components. They incorrectly use infinites-
imal strain definition for the three-dimensional normal strain in axial direction (see equations (8a,b)
in their paper). We show that even when the imposed axial strain is zero, finite in-plane radial
displacement generates in the nanorod (see Section 3). Accordingly, one should use the Lagrangian
strain definition for three-dimensional strain measures. He (2015) also derived extensional stiffness
for circular nanorods but they assume the surface layer of the nanorod as a shell under plane stress
condition (see equation (2) in their paper). However, the plane stress assumption is not a good
approximation in case of circular nanorods. Due to the inherent surface curvature and surface
residual stress, the nanorod’s bulk exerts out-of-plane normal traction on the nanorod’s surface
from within (see eq. (13) below). Pahlevani and Shodja (2011) obtained the formula for twisting
stiffness of circular nanorods. However, they do not consider the in-plane radial displacement that
generates in nanorods due to surface residual stress alone. Accordingly, their formula also shows
no dependence on surface residual stress components. In this paper, we show that surface residual
stress components do affect both extensional and torsional stiffnesses of circular nanorods.
The paper is organized as follows. In Section 2, we present the kinematics of deformation for
combined extension-torsion of circular nanorods. We then define the total energy of the nanorod’s
cross-section as the sum of its bulk energy and surface energy; the latter also includes surface
residual stress term in it. Finally, we derive the nonlinear elasticity equation and the boundary
condition solving which we obtain the unknown radial displacement in arbitrarily stretched and
twisted nanorod. In Section 3, we then show that the radial displacement generates in a nanorod
even when the imposed axial strain and twist are both zero. In Section 4, we obtain analytical
formulae for effective Poisson’s ratio, axial force and extensional stiffness. We also obtain the
expression for relaxed axial strain, i.e., the strain at which the axial force vanishes in the nanorod.
In Section 4.2, we discuss in detail the asymptotic limits of these expressions in both surface
dominated and bulk dominated regimes. We then compare the different formulae for extensional
stiffness obtained by various authors in Section 4.3. In Section 5, we derive expressions for twisting
moment and twisting stiffness and further comment on their asymptotic limits. We then discuss in
Section 6 how to obtain the surface residual stress components as well as surface and bulk elastic
constants for a nanorod using the continuum formulae derived earlier and comparing them with the
nanorod’s molecular simulation data. In Section 7, we then perform molecular statics simulation
on tungsten nanorods of different radii and show that the simulation data for axial force, twisting
moment and stiffnesses match accurately with the derived continuum formulae. Finally, Section 8
concludes our paper.

2
2 Kinematics of deformation, cross-sectional strain energy
and the governing equation for radial displacement
A nanorod can be thought of as being formed by carving out a cylindrical region of specified radius
and cylindrical axis from the associated bulk crystal. We define the strain-free/reference/undeformed
state of the nanorod to be the state that the nanorod attains within the stress-free bulk crystal.
All displacements and strains in the nanorod will be measured relative to this state.

2.1 Kinematics
As the nanorod is isolated from the bulk crystal, surface residual stress generates in it causing its
cross-section to deform even if the nanorod is constrained axially. When the nanorod is further
axially strained () and twisted (κ) from this state, a typical cross-section of the nanorod undergoes
additional in-plane displacement to minimize its energy1 . We use the function u(R) to denote the
total radial displacement of any material point originally at (R, Θ, Z) in the nanorod’s reference
state. Here, (R, Θ, Z) denote the usual cylindrical coordinates. As we are interested in nanorods of
arbitrary radius, we normalize various quantities with the nanorod’s outer radius Ro as follows2 :
∗ ∗ ∗
∗ ∗ κ ∗
R = αR, Θ = Θ, Z = αZ,  = , κ = , u = αu. (1)
α
1
Here α = has the physical meaning of the nanorod’s undeformed surface curvature. The
Ro
deformation map can then be written as

x(R, Θ, Z) = (1 + ) ZeZ + R(Z) R + u(R) eR ,
∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ 
∗ ∗
or, x(R, Θ, Z) = (1 + ) ZeZ + R(Z) R + u(R) eR . (2)
Here (eR , eΘ , eZ ) denote the orthonormal basis vectors of the cylindrical coordinate system attached


with the nanorod’s reference state and R(Z) denotes rotation by an angle of κZ(=κZ) due to
twisting of the nanorod and is defined as
∗ ∗
∗ ∗
 

cos(κZ) − sin(κZ) 0
∗ ∗
R = R =  sin(κ∗ Z) cos(κ∗ Z) 0 . (3)
0 0 1
The deformation gradient tensor can then be written as
 ∗ 

∗0 u ∗ ∗


F = F = (1 + u ) er ⊗ eR + 1 + ∗ eθ ⊗ eΘ + (1 + )ez ⊗ eZ + κ(R + u) eθ ⊗ eZ . (4)
R

Here the superscripted prime ()0 denotes derivative with respect to R while (er , eθ , ez ) denote the
orthonormal basis vectors of the cylindrical coordinate system attached with the deformed nanorod.
For the type of deformation that we are considering, we have:
∗ ∗ ∗ ∗ ∗
∗ ∗ ∗ ∗ ∗
r = R + u, θ = Θ + κZ, z = (1 + )Z, er = ReR , eθ = ReΘ , ez = eZ . (5)
1
In case of circular nanorods, one can employ axisymmetry to show that no out-of-plane displacement generates
for such loadings.

2
From now on, the normalized quantities will be denoted with the overhead asterisk symbol ().

3
Finally, the matrix of Lagrangian strain tensor can be written as follows in (eR , eΘ , eZ ) coordinate
system:
   
1 ∗0 2
(1 + u ) − 1 0 0
2  2  2 
∗ ∗ ∗ ∗
  
∗ 1 u κR u
E=E=
 0 2
1 + ∗ − 1 2
1 + ∗
.

(6)
R R
  
∗ ∗  ∗
 2  ∗
 ∗
 2
κR ∗ ∗
1 + u∗ 1
(1 + )2 + κR 1 + u∗
 
0 2 2
−1
R R

2.2 Cross-sectional strain energy and its first variation


The nanorod’s cross-sectional energy (Φ: strain energy per unit undeformed length of the nanorod,

Φ = α2 Φ: normalized strain energy per unit undeformed length) can now be defined as
 Z Ro 

∗ ∗ 2 2 s s
Φ(, κ) = α Φ(, κ) = α 2π W (E)R dR + 2πRo φ (E )
0
Z 1 
∗ ∗ ∗ ∗
s s
= 2π W (E)R dR + α φ (E ) . (7)
0

Here W denotes a general hyperelastic model for the three-dimensional strain energy density whereas
φs denotes the surface energy per unit undeformed surface area of the nanorod. The surface energy
φs is assumed to only depend on the surface strain tensor (Es ), i.e., the restriction of the Lagrangian
strain tensor (E) in the tangent plane of the nanorod’s undeformed lateral surface3 . Thus, the
surface strain tensor becomes the following 2 × 2 matrix in the (eΘ , eZ ) coordinate system:
 

s s EΘΘ EΘZ
E =E = . (8)
EΘZ EZZ

We further assume the following form for the nanorod’s surface energy per unit undeformed area:

φs (Es ) = KΘ EΘΘ + KΨ EΘZ + KΨ EZΘ + KZ EZZ + φ̂s (Es ). (9)

The first four terms are linear in the components of surface strain tensor and model the presence of
surface residual stress in the nanorod’s lateral surface in its reference state. Here (KΘ , KZ ) denote
the normal components of surface residual stress tensor in circumferential and axial directions
respectively whereas Kψ denotes the shear component. The shear component Kψ is assumed to be
zero in this paper which could be significant in case of intrinsically twisted or non-circular nanorods.
For simplicity, we also assume the remaining part of the surface energy density (φ̂s ) to be isotropic
and quadratic in the components of surface strain tensor, i.e.,
1 2
φ̂s = λs tr(Es ) + µs tr Es Es .

(10)
2
3
Our assumption of the dependence of the surface energy on just the surface strain tensor is in line with the
theory of Gurtin and Murdoch (1975). A more general hypothesis would have been to also include the dependence
of surface curvature as proposed by Steignmann and Ogden (1997). The readers are referred to the work of Fried
and Todres (2005) and Chhapadia et al. (2011) for the application of the curvature dependent model.

4
This quadratic model works as long as surface strains are small enough4 . Similarly, we assume the
Saint-Venant’s isotropic model for bulk energy density, i.e.,
1 2
W (E) = λ tr(E) + µ tr E2 .

(11)
2
Here the Lame’s constants (λ, µ) must satisfy the following constraints: (i) λ + 23 µ >0, (ii) µ >0.
However, there are no inequality constraints on surface parameters (λs , µs ) since the nanorod’s

surface does not exist in isolation. Upon taking the first variation of the cross-sectional energy Φ
∗ ∗
with respect to the unknown u, we obtain the following Euler-Lagrange equation for u:

dPrR 1 ∗ ∗


∗ ∗
∗ + ∗ PrR − PθΘ − κPθZ = 0, R ∈ (0, 1) (12)
dR R

and subject to the following boundary condition at R = 1:
" ! !#

∗ ∂ φ̂s ∗ ∂ φ̂
s
∗2 ∂ φ̂s
PrR = −α(1 + u) KΘ + + 2κ + κ KZ + . (13)
∂EΘΘ ∂EΘZ ∂EZZ

Here P = P = ∂W ∂F
denotes the first Piola-Kirchoff stress tensor. The eq.(12) is the usual equation
of nonlinear elasticity which has reduced to an ODE in the radial coordinate due to the assumed
axisymmetry and axial homogeneity. Similarly, eq.(13) implies that the normal component of
traction on the radial plane at outer radius does not vanish essentially due to non-zero surface
curvature. We can also notice the presence of surface residual stress components in the RHS of
(13).

3 Radial displacement in the nanorod due to surface resid-


ual stress alone
As mentioned earlier, the nanorod’s cross-section undergoes radial displacement even when the
imposed axial strain and twist are zero. For this state, the kinematics (2) reduces to
∗ ∗ ∗ ∗ ∗ ∗


x(R, Θ, Z) = ZeZ + R + u(R) eR (14)

while the Euler-Lagrange equation and the boundary condition reduce to



dPrR 1 ∗ ∗

∗ + ∗ PrR − PθΘ = 0,
dR R !

∗ ∂ φ̂s
PrR = −α(1 + u) KΘ + . (15)
∂EΘΘ
4
With the quadratic model in (10), the energy form (9) becomes Taylor’s expansion of φs in the components of
surface strain tensor up to quadratic order. Zhang et al. (2008) also use the same energy form with an additional
assumption of KΘ = KZ (see equation (8) in their paper).

5


We note that u = β R automatically satisfies the nonlinear equation (15a). Upon further substituting
it in the boundary condition (15b), we obtain
s
2 (λ + µ + αµs ) + αλs − 2αKΘ
β = −1 + . (16)
2 (λ + µ + αµs ) + αλs

We clearly see that β depends on both the surface and bulk elastic constants. Furthermore, it
depends only on the circumferential component (KΘ ) of the surface residual stress tensor in the
absence of which β vanishes identically for any α.
Let us now discuss the asymptotic limits of β. In the large radius limit (continuum limit, α → 0),
the parameter β could be approximated linearly in α as

β≈− α. (17)
2(λ + µ)
Clearly, β vanishes in the large radius limit irrespective of KΘ . Furthermore, the sign of β is opposite
of KΘ since (λ + µ) > 0 always. Thus, a tensile KΘ would lead to shrinking of the nanorod’s cross-
section whereas a compressive KΘ would lead to elongation. However, in the small radius limit
when the surface effects dominate (α → ∞), β does not vanish and becomes
s
−2KΘ + λs + 2µs
β = −1 + . (18)
λs + 2µs

Notice that while the large radius limit depends only on the bulk elastic constants, the small radius
limit depends on just the surface elastic constants. Furthermore, the sign of β isn’t straightforward
anymore in this regime since (λs + 2µs ) could take any sign.

4 Expressions of axial force and extensional stiffness in


finitely stretched nanorods
We would now like to obtain expressions of axial force and extensional stiffness in axially stretched

nanorods at zero twist (κ = 0). The deformation map (2) in this case reduces to
∗ ∗ ∗ ∗ ∗ ∗
∗ ∗

x(R, Θ, Z) = (1 + ) ZeZ + R + u(R) eR . (19)

Upon setting κ = 0, the governing equation (12) and the boundary condition (13) also simplify
greatly. We now assume the following functional form for the unknown displacement:
∗ ∗ ∗
∗ ∗ ∗
u(R; , α) = β R + γ(; α)R. (20)

The function γ models additional radial displacement due to the imposed axial strain. This form for

u again trivially satisfies the reduced form of equation (12). Upon substituting (20) in the reduced
boundary condition, we obtain
s
∗ ∗
∗ 2 (λ + µ + αµs ) + αλs − 2αKΘ − 2 (λ + αλs ) − 2 (λ + αλs )
γ(; α) = −1 − β + . (21)
2 (λ + µ + αµs ) + αλs

6
It would be interesting to understand how does the effective Poisson’s ratio (νef f ) of the nanorod

change due to surface effect. One can obtain it by simply linearizing γ in the imposed , i.e.,

∂γ (λ + αλs )
νef f = − ∗ =p . (22)
∂  =0
∗ (2 (λ + µ + αµs ) + αλs ) (2 (−αKΘ + λ + µ + αµs ) + αλs )

We note from the above expression that the sign of νef f depends on the sign of (λ + αλs ) which
λ
could be opposite to that of λ; the latter dictates the sign of the bulk Poisson’s ratio 2(λ+µ) .
∗ ∗
Having obtained γ, we can then obtain E and Es using (6) and further use (7) to obtain the
∗ ∗
normalized strain energy per unit undeformed length (Φ). The first and second derivatives of Φ with


respect to  then give us the expressions for normalized axial force (F) and normalized extensional

stiffness (D) which we show below:


∗ ∂ Φ ∗ ∗





F() = ∗ = F(0) + D(0)  + D1 2 + D2 3 ,
∂  κ=0∗


∗ ∂ 2 Φ ∗ ∗



D() = ∗2 = D(0) + 2D1  + 3D2 2 . (23)
∂ ∗ κ=0
∗ ∗ 
Here, F(0), D(0) are the normalized axial force and extensional stiffness respectively at zero axial
strain. These quantities and others in (23) above are as follows:
 
∗ KΘ (λ + αλs )
F(0) = 2πα KZ − ,
2 (λ + µ + αµs ) + αλs
 
∗ 2 (λ + αλs ) (λ + αλs + αKΘ )
D(0) = π λ + 2µ + 2α(λs + 2µs + KZ ) − ,
2 (λ + µ + αµs ) + αλs
2 (λ + αλs ) 2
 
∗ 3π
D1 = λ + 2µ − + 2αλs + 4αµs ,
2 2 (λ + µ + αµs ) + αλs

∗ D1
D2 = . (24)
3
We can note that although the normalized axial force in the nanorod’s reference state would be
just 2παKz (force due to axial component of surface residual stress tensor along the cross-section’s
∗ ∗
periphery), F(0) deviates from it due to the induced radial displacement (β R) at zero axial strain.

4.1 Extensional stiffness in the nanorod’s relaxed state


As the nanorod has non-vanishing axial force at zero strain, we can use formula (23a) to obtain the
axial strain at which the axial force vanishes. We call this strain the nanorod’s relaxation strain

(rel ) which turns out to be
v  
u
u 4α λKΘ + α(KΘ − KZ )λs − 2KZ (λ + µ + αµs )
∗ t
rel = −1 + 1 + . (25)
2µ(3λ + 2µ) + 2αµs (5λ + 6µ + 4αµs ) + αλs (λ + 6µ + 8αµs )

7

It can be seen that the relaxation strain approaches zero as α → 0. Furthermore, rel vanishes
identically for all α if we neglect all the surface residual stress components. We now substitute (25)
in (23b) to obtain extensional stiffness in the relaxed state:
 
∗ 2 (λ + αλs ) (λ + αλs − 2αKΘ )
Drel = π λ + 2µ + 2α(λs + 2µs − 2Kz ) − . (26)
2 (λ + µ + αµs ) + αλs
∗ ∗
The formulas for Drel and D(0) are identical except for the sign of surface residual stress components.
In fact, we obtain the following relation:
∗ ∗
Drel − D(0)
∗ = −3. (27)
F(0)
∗ ∗
Thus, Drel softens or stiffens depending on the tensile or compressive nature of F(0).

4.2 The large and small radii asymptotic limits


We now present the large and small radii asymptotic limits of various quantities derived till now. The
large radius limit (α → 0) corresponds to the continuum limit which is obtained by simply linearizing
the corresponding quantities in α. These linearized expressions are presented in Table 1. Upon
plugging α = 0 in these expressions, we simply obtain expressions of the corresponding quantities
λ
for continuum rods having no surface effect. For example, νef f becomes 2(λ+µ) which by definition
is the Poisson’s ratio for isotropic bulk materials. Similarly, the normalized extensional stiffness

D(0) becomes πµ(3λ+2µ)
λ+µ
which again by definition is π times the Young’s modulus of elasticity. If

Formula α→0
λ α (λ (KΘ − 2µs ) + (λ + 2µ)λs )
Effective poisson’s ratio +
2(λ + µ) 4(λ + µ)2
(νef f )
∗ λKΘ − 2KZ (λ + µ)
Relaxation strain (rel ) α
µ(3λ + 2µ) !
λKΘ
Normalized Axial Force 2πα KZ −

2(λ + µ)
(F(0))
2λ2 µs + (λ + 2µ)2 λs
 
πµ(3λ + 2µ) λKΘ
Normalized∗ Extensional +πα − + 2KZ + + 4µs
λ+µ λ+µ 2(λ + µ)2
stiffness (D(0))

Table 1: Formulae in the large radius limit of the nanorod



we further set KΘ = KZ which is a good approximation for nanorods of large radii, F(0) and rel
become:
λ

 λ 
∗ 2(λ+µ)
−1
F(0) = 2παKΘ 1 − , rel = αKΘ . (28)
2(λ + µ) 2µ(λ + µ)(3λ + 2µ)

8
Formula α→∞
λs
Effective poisson’s ratio (νef f ) p
(λs
s + 2µs )(−2KΘ + λs + 2µs )

∗ (KΘ − KZ )λs − 2µs KZ


Relaxation strain (rel ) −1 + 1 +
2µs (λs + µs )
∗ !  
F(0) K Θ λs
Normalized Axial Force 2π KZ −
α λs + 2µs
∗ ! !
D(0) KΘ λs 4µs (λs + µs )
Normalized Extensional stiffness 2π KZ − +
α λs + 2µs λs + 2µs

Table 2: Formulae in the small radius limit of the nanorod


As the bulk Poisson’s ratio is always less than 0.5, we deduce that for large radii nanorods, F(0)

is of the same sign as KΘ whereas the sign of rel is opposite of KΘ . Thus, a tensile KΘ leads to
shortening of the nanorod whereas a compressive KΘ would generate elongation.
We similarly present in Table 2 the corresponding expressions in small radius limit. It may
be noted that the axial force and extensional stiffness are now being normalized by dividing them
with just the nanorod’s radius (instead of its square). Physically, this implies normalizing by the
nanorod’s circumference rather than its cross-sectional area. Indeed, in the small radius limit, the
surface energy term in (7) which is proportional to the nanorod’s circumference becomes more
dominant than the bulk energy term which is proportional to the nanorod’s cross-sectional area.
Accordingly, the expressions in Table 2 become independent of the bulk elastic constants.
We thus notice a dramatic change in the nanorod’s mechanical properties in the two asymptotic
limits. Later, we will see graphically how these mechanical constants change in the intermediate
range of radius.

4.3 Discussion on existing formulae for extensional stiffness


Miller and Shenoy (2000) obtained the following simple formula of extensional stiffness for square
nanobars of side length h (see equation (15) in their paper):
µ(3λ + 2µ)
D = Eh2 + 4Es h, E= . (29)
λ+µ
They define Es as the surface elastic modulus and derive it by assuming the Poisson contraction of
the surface to be the same as that of the bulk which yields
λs λ
Es = λs + 2µs − . (30)
2(λ + µ)
Although the formula (29) is correct for square cross-sections, several researchers have incorrectly
extended this formula to circular nanorods. For instance, Wang et al. (2013); Li et al. (2011);
Thongyothee and Chucheepsakul (2013, 2015); Lu et al. (2014) assume the normalized extensional
stiffness for circular nanorods to be:

D = π (E + 2αEs ) . (31)

9
They all use formula (30) for Es . However, He (2015) define Es to be

4µs (λs + µs )
Es = (32)
λs + 2µs
which they show as the “surface Young’s modulus” of an isolated surface when it is uniaxially
loaded under plane stress condition (see equation (2) in their paper). However, we pointed in the
Introduction section that the plane stress assumption is not valid for the circular nanorod’s surface.
Furthermore, the formula (31) depicts only a linear dependence in α whereas our formula shows
nonlinear dependence in α (see (24b)). In fact, even in the large radius limit when the dependence
on α becomes linear, the coefficient of α does not match either (30) or (32) (see Table 1). In the
small radius limit (α → ∞) however, our formula does show the presence of (32) together with
surface residual stress components (see Table 2).
Zhang et al. (2008) also derived the formula for extensional stiffness of circular nanorods. In
fact, their expression match with ours exactly except for the presence of surface residual stress
components in our expression. This difference arises because they use infinitesimal strain definition
for the three-dimensional normal strain in the axial direction (ZZ ). As we showed in Section 3,
the radial displacement in the nanorod is finite even when the imposed axial strain is zero, they
should therefore use the Lagrangian strain definition for ZZ in their expression for bulk and surface
energy (see equations (8a,b) in their paper) while their expression for work done by external load
(equation (8c) in their paper) should use the imposed axial strain () instead of ZZ . Once, this is
rectified in their formulation, their formulation gives exactly the same expression as ours with the
surface residual stress components also present.

5 Expressions of twisting moment and twisting stiffness for


circular nanorods

We now focus on twisting of nanorods. We assume the following series solution in κ for radial
displacement:
∗ ∗ ∗ ∗
∗ ∗ ∗ ∗
u(R; κ, α) = β R + κg1 (R; α) + κ2 g2 (R; α) + ... (33)

We note here that unlike the case of extension, the functional form for additional radial displacement

due to imposed twist cannot be written as being linear in R since such a functional form does not

satisfy the governing equation (12) for non-zero κ. In order to obtain twisting moment up to linear
∗ ∗ ∗ ∗
order in κ and twisting stiffness at ( = 0, κ = 0), one would require series solution up to κ2 term5 .
The relevant equations and boundary conditions for the unknowns (g1 , g2 ) in (33) can be obtained

by simply substituting the expansion (33) in equations (12-13) for  = 0 and further collecting the
∗ ∗
coefficients of κ and κ2 separately. The equations and boundary conditions along with solutions
of (g1 , g2 ) are presented in the Appendix for readers’ reference. Solving them, one simply obtains
g1 = 06 . In fact, all the coefficients in the series solution (33) corresponding to odd powers of
5 ∗
In case β were zero, the series expansion up to κ order only would have been needed.
6
One can show that g1 won’t vanish in case the surface residual stress tensor also has non-zero shear component
(KΘZ 6= 0) which could be the case for intrinsically-twisted circular nanorods.

10
∗ ∗

κ vanish. To obtain the normalized twisting moment (M) and normalized twisting stiffness (B),

we substitute (33) in the expression for Lagrangian strain matrix (6) at  = 0 which in turn is

substituted in the nanorod’s strain energy (7) and finally differentiated with respect to κ. We thus
obtain


∗ ∂ Φ ∗

M(κ) = ∗ = B(0) κ + ... (34)
∂ κ =0∗


∗ ∗
Here B(0) denotes the normalized twisting stiffness at ( = 0, κ = 0) and is given by
!
∗ 4λ + 4αKΘ − 4αKZ + 3µ + 2αλs
B(0) = π 2 (λ + µ + αµs − αKΘ ) + αλs 1−
4 (λ + µ + αµs ) + 2αλs
!
3αKΘ (λ + µ)
+ . (35)
(2 (λ + µ + αµs ) + αλs ) 2

In the continuum limit (α → 0), the above expression can be linearized in α to yield:
   
∗ πµ λ
B(0) = + πα 2µs + KΘ − 1 + 2KZ + ... (36)
2 2(λ + µ)
From the above formula, we can clearly see that both surface residual stress components and surface

elastic constants affect B. Upon ignoring the effect of surface residual stress components, the above
expression simplifies to
∗ π
B(0) = (µ + 4αµs ) . (37)
2
Pahlevani and Shodja (2011) report the same expression for twisting stiffness since they do not

consider the radial displacement (β R) due to surface residual stress. Finally, in the small radius
limit (α → ∞), we obtain the following expression for twisting stiffness:

B(0) 2π (−KΘ + KZ + µs ) (−2KΘ + λs + 2µs )
= . (38)
α λs + 2µs

6 Evaluation of bulk and surface elastic constants as well


as surface residual stress components
We now describe how the various constants (λ, µ, λs , µs , KΘ , KZ ) can be obtained from molecular
simulation data of isotropic circular nanorods. We can note from the formulae of normalized ex-
tensional and twisting stiffnesses that their continuum limits depend only on bulk elastic constants
(λ, µ). In particular, we have the following two equations:
∗ µ(3λ + 2µ)
lim D(0) = π
α→0 λ+µ
∗ π
lim B(0) = µ. (39)
α→0 2

11
Thus, we first obtain the extensional and twisting stiffnesses from molecular statics simulation of
∗ ∗
nanorods of increasing radii. The large radius limits of their normalized values (D(0) = D(0)
R 2 , B(0) =
o
B(0)
Ro4
are then compared with equation (39) to obtain the bulk elastic constants7 .
)
To obtain the remaining unknowns, we assume that KΘ and KZ are equal which is a good ap-
proximation for nanorods whose radii are not small enough. The three unknowns (λs , µs , KΘ ) can
then be obtained by matching expressions of the following three quantities with the corresponding
data from molecular calculation:

1. Matching the slope of F(0) vs. α at α = 0 :

 
dF(0) λ
= 2πKΘ 1 − . (40)
dα 2(λ + µ)

α=0

2. Matching the slope of B(0) vs. α at α = 0:

   
dB(0) λ
= π KΘ + 1 + 2µs . (41)
dα 2(λ + µ)

α=0

3. Matching the slope of D(0) vs. α at α = 0 :


2λ2 µs + (λ + 2µ)2 λs
   
dD(0) λ
= π 2KΘ 1 − + + 4µs . (42)
dα 2(λ + µ) 2(λ + µ)2

α=0

These three equations are linear in the remaining unknowns and can be used in the given order to
obtain KΘ , µs and λs one after another. We would also like to note that the nanorods of smaller radii
do not have perfect circular cross-section (see Fig.1). However, as the nanorod’s radius increases,
the cross-section becomes more circular. Keeping this in perspective, the use of equations (39-42)
in obtaining the unknown parameters would be desirable since they all correspond to large radius
limits.

7 Numerical results for the tungsten nanorod


In this section, we present molecular statics simulation data of the tungsten nanorod corresponding
to its extension-torsion deformation. We show that the simulation data match pretty well with
our continuum formulae. The tungsten nanorod is specifically chosen since the tungsten’s bulk
crystal exhibits isotropy in its stress free state. The nanorod is formed from the body-centered
cubic crystal of tungsten with the nanorod’s axis chosen along the crystal’s [1 0 0] direction. All the
atoms in the crystal which fall within the prescribed distance of the nanorod’s radius from the axis
make up the nanorod. To impose extension-torsion deformation in the nanorod and further obtain
axial force, twisting moment and stiffnesses, we use the recently proposed Helical Cauchy-Born rule
7
The two bulk constants (λ, µ) can also be obtained by applying periodic boundary conditions in all three directions
to the unit cell of the associated bulk crystal and then computing the second derivative of the minimized unit cell
energy with respect to the deformation gradient.

12
(a) (b)

Figure 1: A typical simulation box of the tungsten nanorod of radius 17.407 Å containing 185
atoms (a) cross-sectional view (b) transverse view

(Kumar et al., 2016; Gupta and Kumar, 2017). The EAM potential (Daw and Baskes, 1984) is
used to model the interaction between tungsten atoms for which the EAM parameters are taken
from Zhou et al. (2001). We show in Fig.1 a typical simulation box of the tungsten nanorod that
we use for molecular statics simulation. It is the smallest repeating entity of the nanorod having an
extent of one unit cell length (3.165 Å) along the axis. It accordingly contains two layers of atoms
each of which forms a typical cross-section of the nanorod. The simulation box is drawn for the
nanorod of radius 17.407 Å containing eleven unit cells along the diameter. We can notice that
the cross-section is not circular at this radius. However, as we increase the nanorod’s radius, the
cross-section becomes more circular.
To impose extension-torsion deformation in the nanorod, the simulation box is helically repeated
along the nanorod’s axis (see Kumar et al. (2016)) and its inter-atomic energy is further minimized.
The derivatives of the minimized inter-atomic energy with respect to imposed strains give us axial
force, twisting moment, extensional and twisting stiffnesses of the nanorod. In order to obtain the
unknowns (λ, µ, λs , µs , KΘ ), we first obtain normalized axial force, extensional stiffness and twisting
∗ ∗
stiffness for nanorods of different radii at ( = 0, κ = 0) and then get their limiting values as well
as their slopes in the large radius limit. Upon comparing these data with their continuum formulae
as mentioned in Section 6, we finally obtain the unknowns (λ, µ, λs , µs , KΘ ). We emphasize that
the simulation data corresponding to nanorods of large radii only (in the range of 200-280 Å) were
used to obtain these unknowns which turn out to be
3 3 2
λ = 1.275 eV /Å , µ = 1.003 eV /Å , KΘ = KZ = 0.347 eV /Å ,
2 2
λs = 2.624 eV /Å , µs = −1.059 eV /Å . (43)

We can now use these parameters and the continuum formulae derived in the preceding section to
obtain axial force, twisting moment, stiffnesses, Poisson’s ratio etc. even for nanorods of smaller
radii.

13
1 1
0.995
Normalized Extensional Stiffness

Normalized Twisting Stiffness


0.95

Present
Present
0.98 0.9 Pahlevani & Shodja (2011)
Zhang et al. (2008)
Atomistic
Miller & Shenoy (2000)
He (2015)
Atomistic 0.85
0.965
0.8

0.95 0.75
0 100 200 300 0 100 200 300
Radius(Å) Radius(Å)
(a) (b)
 
D(0)
Figure 2: (a) Normalized extensional stiffness πRo2 E
of tungsten nanorod (b) Normalized twisting
 
2B(0)
stiffness πR 4µ of tungsten nanorod
o

7.1 Comparison of extensional and twisting stiffnesses obtained using


various approaches
We now present in Fig.2(a) variation in normalized extensional stiffness of the tungsten nanorod
vs. its radius. The normalization is done by dividing the extensional stiffness with the extensional
D(0)
stiffness value in the large radius limit. Thus, the normalized extensional stiffness would be πR 2E
o

D(0)
or πE
.For this reason, all the curves in Fig.2(a) approach unity in the large radius limit. As we
can see, the red curve based on our continuum formula (24b) matches with the atomistic data most
accurately even at small radius. However, the black curve based on the formulation of He (2015)
exhibits maximum error. We have also shown curves corresponding to formulations of Miller and
Shenoy (2000) and Zhang et al. (2008). Similarly, we present in Fig.2(b) the variation in normalized
twisting stiffness vs. the nanorod’s radius. The normalization is again done by dividing the twisting

2B(0)
stiffness with its large radius limit, i.e., we are plotting πR 4 or 2B(0) . We can again notice that
oµ πµ
the red curve based on our formula (35) matches accurately with the atomistic data even at small
radius. However, the curve based on formula (37) by Pahlevani and Shodja (2011) exhibits slight
error.
For readers’ reference, we have also presented in Fig.3 the variation in normalized extensional
and twisting stiffnesses of aluminum nanorod obtained using various approaches. The parameters

14
1.25 1.4
Present Present
Normalized Extensional Stiffness

1.2 Zhang et al. (2008) Pahlevani & Shodja (2011)

Normalized Twisting Stiffness


Miller & Shenoy (2000) 1.2
1.15 He (2015)

1
1.1

1.05
0.8

1
0.6
0.95

0.9 0.4
0 20 40 60 0 20 40 60
Radius(Å) Radius(Å)
(a) (b)
 
D(0)
Figure 3: (a) Normalized extensional stiffness πRo2 E
of aluminum nanorod (b) Normalized twisting
 
2B(0)
stiffness πR 4µ of aluminum nanorod
o

for aluminum nanorod are taken from Zhang et al. (2008) which are also shown below:
3 3 2
λ = 0.369 eV /Å , µ = 0.158 eV /Å , KΘ = KZ = 0.078 eV /Å
2 2
λs = 0.427 eV /Å , µs = −0.02344 eV /Å . (44)

We can note from Fig.3(a) that the four approaches give different results. In fact, the trend of
He (2015) is different from all other approaches. Similarly in Fig.3(b), the trend for normalized
twisting stiffness given by Pahlevani and Shodja (2011) is different from our approach although both
the approaches gave qualitatively the same trend in case of tungsten nanorod. Finally, we show in
Fig.4 the plots of normalized stiffnesses obtained using our continuum formulas for a hypothetical
material whose material constants are mentioned in figure caption. The material constants were
particularly chosen to demonstrate that the variation in stiffnesses with respect to radius need not
be monotonous as in the case of tungsten and aluminum nanorods.

7.2 Comparison of axial force and twisting moment data


We now present in Fig.5 the axial force data obtained from atomistic simulation as well as using
our continuum formula. We can notice the exact match between the two approaches. The trend
in the variation of axial force can be understood by analyzing the formula for axial force in large

15
1.05
1.8
Present
Normalized Extensional Stiffness

Present

Normalized Twisting Stiffness


1
1.4

0.95
1

0.9 0.6
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Radius(Å) Radius(Å)
(a) (b)

Figure 4: (a) Normalized extensional and twisting stiffnesses for a hypothetical material (λ = 1.0,
µ = 5.8, λs = -135.0, µs = 27.0, KΘ = KZ = 85.0)

3000
Present 0.18
Present
ǫ = -0.003 (Atomistic)
ǫ = -0.003 (Atomistic)
Normalized Axial Force (eV/Å )
3

2000 ǫ = 0.000 (Atomistic)


0.14 ǫ = 0.000 (Atomistic)
ǫ = 0.003 (Atomistic)
Axial Force (eV/Å)

ǫ = 0.003 (Atomistic)
1000
0.1

0 0.06

-1000 0.02

-2000 -0.02
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Radius(Å) Radius(Å)
(a) (b)

Figure
  5: (a) Axial force (F) in tungsten nanorods of different radii (b) Normalized axial force
F
R2
in tungsten nanorods of different radii
o

16
radius limit. For  6= 0 (but very small), the axial force could be approximated as
∗ ∗

F() ≈ F(0) + D(0) Ro2
∗ ∗
!
dF(0) dD(0)
= + Ro + πERo2 (large radius limit). (45)
dα dα

α=0 α=0

At  = 0, the quadratic term in radius drops out. Therefore, the curve corresponding to  = 0 in
Fig.5(a) looks like a straight line. For  6= 0 however, the axial force vs. radius curve becomes a
parabola as in Fig.5(a). Similarly, the normalized axial force can be approximated as follows:
∗ ∗
!
∗ dF(0) dD(0) 1
F() ≈ + + πE (large radius limit). (46)
dα dα Ro

α=0 α=0

The above formula implies that the normalized axial force becomes πE (proportional to ) as the
radius becomes large. Due to the first term however, the curve of normalized axial force vs. radius
becomes a hyperbola as in Fig.5(b).
Similarly, the twisting moment can be approximated in the large radius limit as
∗ ∗
! !

∗ dB(0) 1 µ ∗ dB(0) 1 πµ
M(κ) ≈ +π κ= + κRo
dα Ro 2 dα Ro 2

α=0

! α=0

∗ dB(0)
πµRo
⇒ M(κ) = M(κ) Ro3 = + κRo3 . (47)
dα 2

α=0

The above formulas have been derived in terms of κ (not κ) since the simulation data in Fig.6 were
∗ ∗

generated by subjecting nanorods of different radii to the same κ. We can note from (47a) that M(κ)
Ro
which equals M(κ)
Ro4
saturates to πµκ
2
just as in Fig.6(a). Similarly, the curve of M(κ) in Fig.34(b)
can be explained using (47b).

7.3 Relaxation strain



Finally, we present in Fig.7 the variation in relaxation strain (rel ) with respect to the nanorod’s
radius. We again notice that the relaxation strain obtained using our formula (25) and atomistic
calculation match accurately. The trend for relaxation strain curve can be understood easily by

analyzing its large radius limit in eq.(28). We can note from (28) that rel is inversely proportional

to radius. Furthermore, its sign is opoosite of KΘ . As KΘ is tensile for tungsten, the sign of rel is
accordingly negative.

8 Conclusions
We derived analytical expressions for axial force, twisting moment, extensional and twisting stiff-
nesses in isotropic circular nanorods. We also derived their expressions in the limit corresponding
to bulk and surface dominated regimes. The expressions for stiffnesses show dependence on not just

17
×10-3
5 ×105
1.5
Present
κ = 0.003 (Atomistic)
3 Present 1 κ = -0.003 (Atomistic)
κ = 0.003 (Atomistic)
M(κ)/R4o (eV/Å )
4

κ = -0.003 (Atomistic)
πµκ 0.5
1 2

M(κ) (eV)
0
-1
-0.5

-3
-1

-5 -1.5
0 20 40 60 0 20 40 60
Radius(Å) Radius(Å)
(a) (b)

M(κ)
Figure 6: (a) in tungsten nanorods of different radii (b) M(κ) in tungsten nanorods of
Ro4
different radii

0
Relaxation Strain

-0.015

-0.03

Present
Atomistic

-0.045
0 20 40 60
Radius(Å)

Figure 7: Relaxation strain vs radius for tungsten nanorods

18
1 1
Normalized Extensional Stiffness

Normalized Twisting Stiffness


0.99 Present (exact) 0.95 Present (exact)
Present (linearized in α)
Present (linearized in α)
Atomistic
Atomistic
0.98 0.9

0.97 0.85

0.96 0.8

0.95 0.75
0 100 200 300 0 100 200 300
Radius(Å) Radius(Å)
(a) (b)
 
D(0)
Figure 8: (a) Normalized extensional stiffness πRo2 E
of tungsten nanorod (b) Normalized twisting
 
2B(0)
stiffness πR 4µ of tungsten nanorod
o

bulk and surface elastic constants but also on surface residual stress components. We show that
the dependence on surface residual stress arises due to the radial displacement that generates in
the circular nanorod even when the imposed axial strain and twist are both zero. Our expressions
also show nonlinear dependence of stiffnesses on α: the undeformed surface curvature. Even when
our expression is linearized in α, the coefficient of α in the linearized expression differs from those
of other authors (see Section 4.3 for detailed comparison). We would also like to note that the
results from expressions for stiffnesses linearized in α match very well with the exact nonlinear
expression for the case of tungsten nanorod (see Fig.8). However, this cannot be generalized for a
general material as the linearized expressions can only show monotonic dependence of stiffnesses on
nanorod’s radius (see Fig.4 for the case when the dependence is not monotonic).
The formulation proposed here can now be extended to obtain the effect of surface elasticity on
bending and shearing stiffnesses of nanorods. Furthermore, one can also employ the material model
corresponding to cubic symmetry instead of isotropy that we used here. Such a formulation would
be useful for most of the crystals since they exhibit cubic symmetry at atomic scale. Of course, the
assumption of axisymmetric deformation will no more hold and one would have to solve a PDE in
the nanorod’s full cross-section to obtain the warping function. It would also be interesting to work
with the curvature dependent surface energy model of Steignmann and Ogden (1997) and analyze
its effect on extensional and torsional stiffnesses.

19
9 Acknowledgments
P. Gupta acknowledges the financial support from DST-INSPIRE fellowship and A. Kumar ac-
knowledges the financial support from Science and Engineering Research Board, India through the
project grant YSS/2014/000023.

Appendix: Equations to obtain the functions (g1, g2) in the


series solution for the case of twisting
The equations and boundary conditions for the unknowns (g1 , g2 ) in the series solution (33) can

be obtained by simply substituting the expansion (33) in equations (12-13) for  = 0 and further
∗ ∗2
collecting the coefficients of κ and κ separately. The equation for g1 turns out to be
∗ ∗
R2 g100 + Rg10 − g1 = 0, (48)

subject to the following boundary condition at R = 1:
 
0
g1 (2β(β + 2) + 1)λ + (3β(β + 2) + 2)µ + (β + 1)2 g1 λ =
!
α  
− g1 2 (3β(β + 2) + 2)µs + KΘ + (3β(β + 2) + 2)λs . (49)
2

The general solution of (48) is:


∗ c2
g1 = c1 R + ∗ . (50)
R

For a solid circular rod, no displacement arises at R = 0 which implies c2 = 0. Thus,

g1 = c1 R. (51)

Finally, upon substituting (51) in the boundary condition (49), we simply obtain g1 = 0.
Similarly, the equation for g2 turns out to be
∗ ∗ ∗ (β + 1) (λ − 2(β + 1)2 µ)
R2 g200 + Rg20 − g2 = AR3 where A = − (52)
(2β(β + 2) + 1)λ + (3β(β + 2) + 2)µ

subject to the following boundary condition at R = 1:
 
2g20 (2β(β + 2) + 1)λ + (3β(β + 2) + 2)µ + (β + 1)2 λ(β + 2g2 + 1) =
 
2
−α β(2β + 3(β + 2)g2 + 6β + 5) + 2g2 + 1 λs
  !
+2 (3β(β + 2) + 2)g2 + 2(β + 1)3 µs + βKZ + g2 KΘ + KZ . (53)

20
The general solution of (52) is:
∗ c2 A ∗3
g2 = c1 R + ∗ + R . (54)
R 8

For the same reason as above, c2 = 0 implying


∗ A ∗3
g2 = c1 R + R . (55)
8
Finally, upon substituting (55) in the boundary condition (53), we obtain

B
c1 = −   where
2 (3β(β + 2) + 2)(λ + µ) + αKΘ + α(3β(β + 2) + 2) (λs + 2µs )
A   3A 3A A 
B = (7β(β + 2) + 4)λ + 9β(β + 2)µ + αKΘ + 6µ + α β(β(2β + + 6) + + 5) + + 1 λs
4 8 4 4
 A 
+2α (3β(β + 2) + 2) + 2(β + 1)3 µs + (β + 1)3 λ + 2α(β + 1)KZ .
8

References
Antman, S.S., 1995. Nonlinear problems of elasticity. Springer-Verlag, New York.

Chhapadia, P., Mohammadi, P., Sharma, P., 2011. Curvature-dependent surface energy and impli-
cations for nanostructures. J. Mech. Phys. Solids, 59(10), 2103-2115.

Cuenot, S., Frtigny, C., Demoustier-Champagne, S., and Nysten, B., 2004. Surface tension effect
on the mechanical properties of nanomaterials measured by atomic force microscopy. Phys. Rev.
B. 69, 165410.

Cui, Y., Wei, Q., Park, H. and Lieber, C.M., 2001. Nanowire nanosensors for highly sensitive and
selective detection of biological and chemical species. Science, 293(5533), pp.1289-1292.

Cui, Y., Zhong, Z., Wang, D., Wang, W.U. and Lieber, C.M., 2003. High performance silicon
nanowire field effect transistors. Nano letters, 3(2), pp.149-152.

Daw, M.S. and Baskes, M.I., 1984. Embedded-atom method: Derivation and application to impu-
rities, surfaces, and other defects in metals. Physical Review B, 29(12), p.6443.

Fried, E., Todres, R.E., 2005. Mind the Gap: The Shape of the Free Surface of a Rubber-Like
Material in Proximity to a Rigid Contactor. J. Elas. 80, 97-151.

Gupta, P. and Kumar, A., 2017. Effect of material nonlinearity on spatial buckling of nanorods and
nanotubes. Journal of Elasticity, 126(2), pp.155-171.

Gurtin, M. E., and Murdoch, A. I., 1975. A continuum theory of elastic material surfaces. Arch.
Ration. Mech. An. 57, 291-323.

21
He, J., 2015. Surface stress on the effective Youngs modulus and Poissons ratio of isotropic nanowires
under tensile load. AIP Advances, 5(11), p.117206.

Healey, T.J., 2002. Material symmetry and chirality in nonlinearly elastic rods. Math. Mech. Solids.
7, 405-420.

Jing, G.Y., Duan, H., Sun, X.M., Zhang, Z.S., Xu, J., Li, Y.D., Wang, J.X. and Yu, D.P., 2006.
Surface effects on elastic properties of silver nanowires: contact atomic-force microscopy. Phys.
Rev. B. 73, 235409.

Kumar, A., Kumar, S., Gupta, P., 2016. A helical Cauchy-Born rule for special Cosserat rod
modeling of nano and continuum rods. J. Elas., 124, 81-106.

Li, Y., Song, J., Fang, B. and Zhang, J., 2011. Surface effects on the postbuckling of nanowires.
Journal of Physics D: Applied Physics, 44(42), p.425304.

Limkatanyu, S., Damrongwiriyanupap, N., Prachasaree, W. and Sae-Long, W., 2013. Modeling
of axially loaded nanowires embedded in elastic substrate media with inclusion of nonlocal and
surface effects. Journal of Nanomaterials, 2013, p.104.

Liu, J.L., Mei, Y., Xia, R. and Zhu, W.L., 2012. Large displacement of a static bending nanowire
with surface effects. Physica E: Low-dimensional systems and Nanostructures, 44(10), pp.2050-
2055.

Lu, Z., Zhu, M. and Liu, Q., 2014. Size-dependent mechanical properties of 2D random nanofibre
networks. Journal of Physics D: Applied Physics, 47(6), p.065310.

Miller, R. and Shenoy, V.B., 2000. Size-dependent elastic properties of nano-sized structural ele-
ments. Nanotechnology, 11,139-147.

On, B. B., Altus, E., and Tadmor, E. B., 2010. Surface effects in non-uniform nanobeams: continuum
vs. atomistic modeling. Int. J. Solids Struct. 47, 1243-1252.

Pahlevani, L. and Shodja, H.M., 2011. Surface and interface effects on torsion of eccentrically
two-phase fcc circular nanorods: determination of the surface/interface elastic properties via an
atomistic approach. Journal of Applied Mechanics, 78(1), p.011011.

Park, H. S., Klein, P. A., and Wagner, G. J., 2006. A surface Cauchy-Born model for nanoscale
materials. Int. J. Numer. Meth. Engg. 68, 1072-1095.

Park, H. S., 2012. Surface stress effects on the critical buckling strains of silicon nanowires. Comp.
Mater. Sci. 51, 396-401.

Shenoy, V. B., 2002. Size-dependent rigidities of nanosized torsional elements. International Journal
of Solids and Structures, 39(15), 4039-4052.

Shenoy, V. B., 2005. Atomistic calculations of elastic properties of metallic fcc crystal surfaces.
Phys. Rev. B. 71, 094104.

22
Steigmann, D. J. and Ogden, R. W., 1997. Plane deformations of elastic solids with intrinsic bound-
ary elasticity. Proc. R. Soc. London, A. 453, 853-877.

Tang, W., Lagadec, P., Gould, D., Wan, T. R., Zhai, J., and How, T., 2010. A realistic elastic rod
model for real-time simulation of minimally invasive vascular interventions. Visual Comput. 26,
1157-1165.

Thongyothee, C. and Chucheepsakul, S., 2013. Postbuckling behaviors of nanorods including the
effects of nonlocal elasticity theory and surface stress. Journal of Applied Physics, 114(24),
p.243507.

Thongyothee, C. and Chucheepsakul, S., 2015. Postbuckling of unknown-length nanobeam con-


sidering the effects of nonlocal elasticity and surface stress. International Journal of Applied
Mechanics, 7(03), p.1550042.

Wang, Z., Zhao, Y., Huang, Z., 2010. The effects of surface tension on the elastic properties of nano
structures. Int. J. Engg. Sci., 48(2), 140-150.

Wang, Y., Song, J. and Xiao, J., 2013. Surface effects on in-plane buckling of nanowires on elas-
tomeric substrates. Journal of Physics D: Applied Physics, 46(12), p.125309.

Wong, E. W., Sheehan, P. E., and Lieber, C. M., 1997. Nanobeam mechanics: elasticity, strength,
and toughness of nanorods and nanotubes. Science. 277, 1971-1975.

Zhang, W., Wang, T. and Chen, X., 2008. Effect of surface stress on the asymmetric yield strength
of nanowires. Journal of Applied Physics, 103(12), p.123527.

Zhou, X.W., Wadley, H.N.G., Johnson, R.A., Larson, D.J., Tabat, N., Cerezo, A., Petford-Long,
A.K., Smith, G.D.W., Clifton, P.H., Martens, R.L. and Kelly, T.F., 2001. Atomic scale structure
of sputtered metal multilayers. Acta materialia, 49(19), pp.4005-4015.

23

View publication stats

You might also like