You are on page 1of 6

Size-dependent phase transition temperatures of dispersed systems

Yong-Qiang Xue
a,n
, Miao-Zhi Zhao
a
, Wei-Peng Lai
b
a
Department of Applied Chemistry, Taiyuan University of Technology, Taiyuan 030024, Shanxi, PR China
b
Xian Modern Chemistry Research Institute, Xian 710065, Shanxi, PR China
a r t i c l e i n f o
Article history:
Received 19 June 2012
Accepted 26 September 2012
Available online 3 October 2012
Keywords:
Phase transition temperatures
Additional pressure
Thermodynamic equation
Dispersed systems
a b s t r a c t
A phase transition equation on the basis of the additional pressure on the curved surfaces of dispersed
systems has been proposed, and the specic differential equations for various kinds of phase transitions
of dispersed systems have been derived by the phase transition equation. Applying the fusion transition
equations, the melting temperatures of Au and Sn nanoparticles have been calculated, and the
predicted melting temperatures are in good agreement with the available experimental data. The
results show that the phase transition equations can be used to predict the temperatures of phase
transitions of dispersed systems and to explain the phenomenon of metastable states; the size of the
dispersed phase has noticeable effect on the phase transition temperature; all temperatures of fusion,
solidication, condensation, vaporization, sublimation and desublimation decrease with decreasing
radius of the dispersed phase, but the bubble point temperature of a planar liquid increases with
decreasing absolute value of radius of the bubbles; the depression of melting temperature for a
nanowire is approximately half of that for a spherical nanoparticle with identical radius.
& 2012 Elsevier B.V. All rights reserved.
1. Introduction
Nanomaterials have attracted remarkable interest due to their
unique physical and chemical properties. One of the well-known
interests is the variation of the melting temperature with the
particle size, which is widely studied both theoretically and
experimentally. As early as 1909, Pawlow [1] dened the tem-
perature at which a solid sphere, a liquid sphere and their vapor
were in equilibrium as the melting point; then he developed a
quantitative relation between the melting temperature and the
particle size. Later the relation was improved by Hanszen [2].
Another phenomenological model, put forward by Rie [3], how-
ever, considered the equilibrium as a solid particle surrounded by
an innite liquid. Taking into account that the initial step was
equivalent to the reversible transfer of material from a solid to a
liquid sphere, Reiss and Wilson [4] proposed a model in which a
solid core and a thin liquid shell surrounding the core were
considered as in equilibrium. Afterward, Curzon [5] modied the
model. The said three models, phenomenological fusion models
for free-standing nanocrystals, have predicted and explained
lots of experimental results [615]. However, the equations
derived from the above models are not accurate thermodynamics
equations.
Based on the empirical relation that the melting temperature
of a solid is linear on its cohesive energy [16,17] and under the
condition of some approximation, models for size-dependent
melting and dimension-dependent melting were proposed, such
as Safaeis [1820], Lu et al.s [21,22] and Qi et al.s [23,24], in
which the expressions of cohesive energy show some differences.
According to the size-dependent amplitude of the atomic thermal
vibrations of nanocrystals, a model for melting temperature was
developed by Jiang et al. [25,26] and Zhu et al. [27]. Since the
model involves a microcosmic parameter (vibrational entropy),
wide application of the equation is difcult. The melting tem-
peratures predicted by these models are consistent with some
experimental results [8,12,2830]. However, the models are
applicable well only to metallic crystals.
The above models or theories can be applied only to fusion
transition of nanocrystals, and some of them fail to predict
melting temperature well when particle size is smaller (e.g.,
particle radius o5 nm) [6,10,15,23,3133]. In our previous work
[34], we proposed an equation for a phase transition in a
dispersed system according to the specic surface area of the
dispersed system. The phase transition equations are able to be
applied to predict the phase transition temperatures of dispersed
systems. In the present research, another equation of phase
transitions is developed on the basis of the additional pressure
on the curved surfaces of dispersed systems, by which some
specic equations for various kinds of phase transitions of
dispersed systems are derived. Then, the regularity and the extent
of effect of size on phase transition temperature are discussed,
Contents lists available at SciVerse ScienceDirect
journal homepage: www.elsevier.com/locate/physb
Physica B
0921-4526/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.physb.2012.09.053
n
Corresponding author. Tel./fax: 86 351 6014476.
E-mail address: xyqlw@126.com (Y.-Q. Xue).
Physica B 408 (2013) 134139
and the melting temperatures of Au and Sn nanoparticles are
calculated.
2. Thermodynamic equation for phase transitions
It is supposed that in a phase transition from phase a to b, at
least one of the two phases is a dispersed phase. The phases a and
b can coexist on the phase boundary. Let T, p
a
and p
b
be changed
innitesimally, where T is the temperature of phase transition, p
a
and p
b
are the pressures in phases a and b, respectively. The
chemical potentials of the two phases are initially equal (the two
phases are in equilibrium). They remain equal when the condi-
tions change to another point on the phase boundary, where the
two phases continue to be in equilibrium. Therefore, the changes
in the chemical potentials of the two phases must be equal,
dm
a
dm
b
, i.e.,
S
a
dT V
a
dp
a
S
b
dT V
b
dp
b
1
where S
a
and S
b
are the molar entropies of the two phases, and V
a
and V
b
are their molar volumes, respectively. Eq. (1) can be
changed to
V
b
dp
b
dT
V
a
dp
a
dT
D
b
a
S
m
2
where D
b
a
S
m
S
b
S
a
is the change in molar entropy from phase a
to b, and equal to D
b
a
H
m
=T when the two phases are in
equilibrium, where D
b
a
H
m
is the molar enthalpy change. Then
Eq. (2) becomes
V
b
dp
b
dT
V
a
dp
a
dT

D
b
a
H
m
T
: 3
Eq. (3) is a phase transition equation which can be applied to any
phase transition between two phases. It is implied in Eq. (3) that
the pressure exerted on a phase by the curved surfaces is
responsible for the change of the melting point. When the sizes
of dispersed phase(s) in dispersed systems tends(tend) to innity,
then p
a
p
b
p
ex
, where p
ex
is external pressure, and the funda-
mental equation will become the Clapeyron equation in classical
thermodynamics.
3. Application of the phase transition equation
3.1. Application to fusion
3.1.1. Fusion of nanoparticles
Let a free-standing nanoparticle be spherical, and the melting
begin on the surface [8]. The molten liquid shell with width t
surrounds uniformly the solid core. When the solidliquid phase
is in equilibrium, the radii of the solid core and the liquid shell are
r
s
and r
l
, respectively (see Fig. 1).
According to the Laplace equation, Dp s 1=r
1
_ _
1=r
2
_ _

_
,
where Dp refers to the difference between the internal pressures
of the material in bulk and in spherule form; the pressures on the
liquid shell and the solid core are
p
l
p
ex

2s
lv
r
l
4
and
p
s
p
ex

2s
lv
r
l

2s
sl
r
s
, 5
respectively, where s
lv
and s
sl
are the interface tensions of the
liquidvapor interface and the solidliquid interface, respectively.
Since the solid core is acted by two curved surfaces, the
additional pressure applied to the core is made up of two parts,
one is between the solid particle and the liquid shell and the other
is between the liquid shell and its vapor. The derivatives of Eqs.
(4) and (5) against T at constant external pressure are
dp
l
dT

2
r
l
@s
lv
@T
_ _
p
ex

2s
lv
r
2
l
@r
l
@T
_ _
p
ex
6
and
dp
s
dT

2
r
l
@s
lv
@T
_ _
p
ex

2s
lv
r
2
l
@r
l
@T
_ _
p
ex

2
r
s
@s
sl
@T
_ _
p
ex

2s
sl
r
2
s
@r
s
@T
_ _
p
ex
,
7
respectively. The total mass in fusion process is changeless,
therefore,
4
3
pr
3
s
r
s

4
3
p r
3
l
r
3
s
_ _
r
l

4
3
pr
3
r
s
8
where r
s
and r
l
are the densities of the solid core and the liquid
shell, respectively, r is the radius of the nanoparticle before
melting, and r
l
r
s
t.
Partial derivative of Eq. (8) with respect to T gives
@r
s
@T
_ _
p
ex

@r
l
@T
_ _
p
ex

r
2
r
s
r
2
l
r
l
r
l
t
2
r
s
r
l
_ _
@r
@T
_ _
p
ex
: 9
Applying Eqs. (6), (7) and (9) to Eq. (3), one can obtain
@T
@r
_ _
p
ex

_
2r
2
r
s
=r
2
l
r
l
r
l
t
2
r
s
r
l
_ _
s
lv
=r
2
l
_ _
V
s
V
l
s
sl
V
s
=r
2
s
_ _

D
l
s
H
m
=T
_ _
2=r
l
_ _
@s
lv
=@T
_ _
p
ex
V
s
V
l
2V
s
=r
s
_ _
@s
sl
=@T
_ _
p
ex
:
10
Eq. (10) is a precise differential equation to describe the relation
between the melting temperature and the size of nanoparticles,
which can be solved by numerical differentiation.
If one deals with the precise equation approximatively, the
accepted thermodynamic relation [7,8] can be derived as follows.
During the melting process, r
s
r
l
; thus one can obtain from
Eq. (8) that r
l
r. If one neglects the effects of T on s
sl
and s
lv
,
Eq. (10) can be simplied as
@T
@r
_ _
p
ex

2T
D
l
s
H
m
s
lv
V
s
V
l

r
2

s
sl
V
s
rt
2
_ _
: 11
If the effects of T and r on D
l
s
H
m
, s
lv
, s
sl
, V
s
and V
l
are
neglected, Eq. (11) can be integrated from the bulk state (r-1)
to the nanosized state (r):
_
T
T
0
dT
T

2
D
l
s
H
m
_
r
1
s
lv
V
s
V
l

r
2

s
sl
V
s
rt
2
_ _
dr, 12
where T is the melting temperature of a solid with radius r and T
0
is the melting temperature of the same crystal in block state (i.e.,
normal m.p.), the below equation can be obtained:
ln
T
T
0

2
D
l
s
H
m
s
lv
V
s
V
l

r

s
sl
V
s
rt
_ _
: 13
r
s
solid core
thickness of liquid shell t
liquid shell
r
l
r
l
r
s
= t
Fig. 1. Fusion model of a free-standing nanoparticle.
Y.-Q. Xue et al. / Physica B 408 (2013) 134139 135
If 9TT
0
95T
0
, thus ln T=T
0
_ _
TT
0
=T
0
; combining
V
s
V
l
M 1=r
s
1=r
l
_ _
, Eq. (13) becomes
T
T
0
1
2V
s
D
l
s
H
m
s
sl
rt

s
lv
r
1
r
s
r
l
_ _ _ _
: 14
It is clear that Eq. (14) is the same as the equation originated
by Reiss and Wilson [4] and modied by Curzon [5]. When the
size is small (e.g., ro5 nm), the effects of r on D
l
s
H
m
, s
lv
and s
sl
are notable [22,3537]. Thus, the approximate relation fails for
smaller nanocrystals. The conclusion has been conrmed by the
experiments [6,10,15,23,3133,38,39].
If the outer thin liquid shell is considered as a solid shell, the
said relations can also be applied to solidsolid phase transition of
isolated particles (i.e., crystal phase transformation).
Applying the fundamental phase transition equation to
Pawlows and Ries phenomenological models [13], one can
obtain the corresponding equations by an approximate treatment.
3.1.2. Fusion of nanowires
Nanowires exhibit a similar melting behavior, although with
signicant differences due to their cylindrical symmetry and
innite extension along the axis.
For the melting model of a nanowire, one supposes that the
length of a nanowire is L, the radius before melting is r and Lbr.
When the solidliquid phase is in equilibrium, the radii of the
solid nanowire and the liquid shell which surrounds the nanowire
are r
s
and r
l
, respectively. The thickness of the liquid shell is t,
thus r
l
r
s
t (see Fig. 2).
According to the Laplace equation, Dp s 1=r
1
_ _
1=r
2
_ _

_
;
for the nanowire, r
1
r
l
, r
2
1. So the internal pressures of the
liquid shell and the solid nanowire are
p
l
p
ex

s
lv
r
l
15
and
p
s
p
ex

s
lv
r
l

s
sl
r
s
, 16
respectively. The partial derivatives of Eq. (15) and (16) with
respect to T at constant pressure are
dp
l
dT

1
r
l
@s
lv
@T
_ _
p
ex

s
lv
r
2
l
@r
l
@T
_ _
p
ex
17
and
dp
s
dT

1
r
l
@s
lv
@T
_ _
p
ex

s
lv
r
2
l
@r
l
@T
_ _
p
ex

1
r
s
@s
sl
@T
_ _
p
ex

s
sl
r
2
s
@r
s
@T
_ _
p
ex
,
18
respectively.
The total mass in fusion process is changeless; therefore,
pr
2
Lr
s
pr
2
s
L2t r
s
pr
2
l
Lpr
2
s
L2t r
l
19
thus
@r
s
@T
_ _
p
ex

@r
l
@T
_ _
p
ex
B
@r
@T
_ _
p
ex
20
where
B
rL
r
s
L2t 1 r
l
=r
s
_ _ _ _
r
l
L r
l
=r
s
_ _ : 21
Applying Eqs. (17), (18), (20) and (21) to Eq. (3), the precise
relation between the melting temperature of the nanowire and
the radius is obtained:
For the nanowire, the length L is much larger than the radius;
at the beginning of the melting, t5r
s
and r
s
r
l
r, so Eq. (22)
can be written as
If one disregards the effects of T and r on D
l
s
H
m
, s
lv
, s
sl
, V
s
and
V
l
, Eq. (23) can be integrated from the bulk state (r-1) to the
nanosized state (r) and the following equation is obtained:
ln
T
T
0

V
s
s
sl
V
s
s
lv
V
l
s
lv
D
l
s
H
m
r
: 24
If 9TT
0
95T
0
, thus ln T=T
0
_ _
TT
0
=T
0
; combining V
s
V
l

M 1=r
s
1=r
l
_ _
, Eq. (24) becomes
T
T
0
1
V
s
D
l
s
H
m
r
s
sl
s
lv
1
r
s
r
l
_ _ _ _
: 25
Eq. (25) is similar to the equations for the fusion of nanowire
derived by Gulseren et al. [40] and Sankaranarayanan et al.
[41,42] with the assumption r
s
r
l
.
It can be seen from Eq. (25) that the melting temperature
drops with decreasing radius of nanowire, which has been
conrmed by large number of molecular dynamics simulations
and experiments [21,22,4147]. However, comparing Eq. (25)
with Eq. (14), the ratio of depression of melting temperature for
a spherical nanoparticle to that of a nanowire of the same radius
is 2:1; such ratio was also obtained by our previous research [34]
and other researchers equations [40,42]. The factor of 2 arises
because of the changes in curvature and the Laplace pressure
L (L >> r)
thickness of liquid t
r
l
r
s
t
liquid
solid
Fig. 2. Fusion model of a nanowire.
@T
@r
_ _
p
ex

V
s
s
sl
=r
2
s
_ _
V
s
s
lv
=r
2
l
_ _
V
l
s
lv
=r
2
l
_ _
B
D
l
s
H
m
=T
_ _
V
l
=r
l
_ _
@s
lv
=@T
_ _
_ _
p
ex
V
s
=r
l
_ _
@s
lv
=@T
_ _
p
ex
V
s
=r
s
_ _
@s
sl
=@T
_ _
p
ex
_ _ 22
@T
@r
_ _
p
ex

1=r
2
_ _
V
s
s
sl
V
s
s
lv
V
l
s
lv

D
l
s
H
m
=T
_ _
V
l
=r
l
_ _
@s
lv
=@T
_ _
p
ex
V
s
=r
l
_ _
@s
lv
=@T
_ _
p
ex
V
s
=r
s
_ _
@s
sl
=@T
_ _
p
ex
:
_ _ _ 23
Y.-Q. Xue et al. / Physica B 408 (2013) 134139 136
when the nanosized substance goes from a spherical cluster to a
wire. The approximate ratio 2 is in good agreement with mole-
cular dynamics simulations and/or experimental data for both
pure metallic and bimetallic nanocrystals [4043].
3.2. Application to solidication
Let a tiny solid particle crystallized from a large quantity of liquid
be spherical. The solid particle is surrounded by the liquid, so the
solid particle is a disperse phase and the liquid is a continuous phase.
Thus,
p
l
p
ex
26
and
p
s
p
ex
p
st

2s
sl
r
s
, 27
where p
st
is the static pressure on the particle.
The derivatives of Eqs. (26) and (27) against T at constant
p
ex
are
dp
l
dT
0 28
and
dp
s
dT

2s
sv
r
2
s
@r
s
@T
_ _
p
ex

2
r
s
@s
sv
@T
_ _
p
ex
, 29
respectively. Applying Eqs. (28) and (29) to Eq. (3), one obtains
@T
@r
s
_ _
p
ex

2s
sl
V
s
2V
s
r
s
@s
sl
=@T
_ _
p
ex
D
s
l
H
m
r
2
s
=T
_ _ : 30
Eq. (30) is an accurate differential equation to describe the relation
between the freezing temperature and the size of the crystal particles,
which can also be solved by numerical differentiation.
If one neglects the effect of T on s
sl
, Eq. (31) can be obtained by
integrating Eq. (30) from the bulk state (r-1) to the nanosized
state (r)
ln
T
T
0

2s
sl
V
s
D
s
l
H
m
r
s
31
If TT
0
j j 5T
0
, ln T=T
0
_ _
TT
0
=T
0
and Eq. (31) becomes
T
T
0
1
2s
sl
V
s
D
s
l
H
m
r
s
32
The heat of solidication D
s
l
H
m
o0, therefore, the freezing
temperature decreases with decreasing particle size, by which
one can explain the phenomenon of supercooled liquid.
3.3. Application to condensation
For the condensation transition between small liquid droplet and
its vapor, the vapor is innite and the radius of the droplet is r
l
at
condensation equilibrium. In a similar way to that of solidication,
one obtains
@T
@r
l
_ _
p
ex

2s
lv
V
l
2V
l
r
l
@s
lv
=@T
_ _
p
ex
D
l
g
H
m
r
2
l
=T
_ _ :33
Eq. (33) is an accurate differential equation to describe the relation
between the condensing temperature and the size of the droplets,
which can be simplied to
T
T
0
1
2s
lv
V
l
D
l
g
H
m
r
l
34
The heat of condensation D
l
g
H
m
o0, therefore; the condensing
temperature decreases with the reduction of the radius of the
droplet, by which one can explain the supersaturated vapor.
The model can also be applied to the evaporation of tiny
droplet, and the same conclusion will be obtained.
If one considers a liquid droplet as a small solid particle, one
can derive similar equations for phase transitions (sublimation
and desublimation) between the small solid particle and its vapor.
3.4. Application to vaporization
The initial stage of ebullience begins with the formation of tiny
bubbles in the liquid. When the liquid and the tiny bubbles are in
equilibrium, suppose the radius of the bubbles is r
g
; then, the
internal pressure of the bubbles is
p
g
p
ex
p
st

2s
lv
r
g
35
Based on the fundamental equation, one can obtain the
accurate relation between the bubble point temperature of a
planar liquid and the radius of the bubbles:
@T
@r
g
_ _
p
ex

2s
lv
V
g
r
2
g
2V
g
=r
g
_ _
@s
lv
=@T
_ _
p
ex
D
g
l
H
m
=T
_ _

36
By an approximate treatment of Eq. (36), one can obtain
T
T
0
1
2s
lv
V
g
D
g
l
H
m
r
g
: 37
For the vaporization, D
g
l
H
m
40 and r
g
o0, so one can see from
Eq. (37) that the bubble point temperature of a planar liquid
increases with the decrease of the absolute value of radius of the
bubbles, by which one can explain the phenomenon of super-
heated liquid at constant external pressure.
When the size of the dispersed phase is large, one can estimate the
order of magnitude for the change in phase transition temperature by
the above approximate relations. Take Eq. (37) for an example, s
lv
for
a general liquid is on the order of 10
2
J m
2
, V
g
is on the order of
10
2
m
3
mol
1
and D
g
l
H
m
is on the order of 10
4
J mol
1
. Thus, for
9r910
6
m, 10
7
m and 10
8
m, the ratios DT=T
0
(DT TT
0
) are
on the orders of 10
2
, 10
1
and 10
0
, respectively. It is obvious that the
effect of size on bubble point temperature is considerable when the
radius 9r9o10
7
m.
Table 1
Formulas and parameters used in the calculation of melting temperatures.
Calculated items Formulas for the calculation Parameters
Melting temperature T Eq. (10) at t 0
D
l
s
Hmr
l
, r
s
, V
l
, Vss
lv
, s
sl
Melting enthalpy D
l
s
Hm D
b
a
Hm D
b
a
Hm T
0

_
T
T0
Cp,mb
_ _
Cp,madTCp,m abT cT
2
dT
3
D
b
a
Hm T
0
[48](a); a, b, c, d [48](b)
Density of liquid r
l r
l
A
2
B
2
1T=Tc
n
2
A
2
, B
2
, n
2
, Tc [48](c)
Density of solid r
s
Obtained by tting the densities of the solid at different temperatures r [54]
Molar volume V
l
, Vs V
M
r
M [48](d)
Interfacial tensions
lv
, s
sl
s
lv
s
l,1
exp s
l1
s
l2
s
l3
[34]s
sl
0:25 s
lv
A
1
n
1
1
T
Tc
_ _
n11
TT0
Tc
_ _
exp s
l1
s
l2
s
l3
[34]
A
1
, n
1
, Tc [48](e)
Y.-Q. Xue et al. / Physica B 408 (2013) 134139 137
To sum up the above equations and conclusions, the phase
transition temperature decreases with decreasing radius of a
dispersed phase (the radius of a particle or a droplet is positive;
the radius of a bubble is negative).
4. Calculation results
Applying the above equations, one can predict the phase
transition temperature of dispersed systems. In the following
part, taking the melting transition of Au and Sn nanoparticles as
examples, we calculate the melting temperature of the nanopar-
ticles with different sizes and make comparison with available
experimental data.
4.1. Calculation details
Lots of experimental [15,4951] and molecular dynamics
simulation results [52] indicated that the melting of nanoparticles
begins on the surface. Therefore, we take the thickness of the
liquid shell t 0 when calculating the onset melting temperature;
such a treatment agrees with that of Zhang et al. [53]. Considering
the effects of temperature on melting enthalpy, melting entropy,
molar volume and density and the effects of temperature and size
on interfacial tensions, the melting temperatures of Au and Sn
nanoparticles with different sizes are calculated. The correspond-
ing formulas and parameters are listed in Table 1.
The calculated melting temperatures and the available experi-
mental data of Au and Sn nanoparticles are shown in Figs. 3 and 4,
respectively. They are indicated that the melting temperatures of
Au and Sn nanoparticles calculated by the precise equation are in
good consistency with the corresponding experimental data
[7,13,15,32]. Even at low particle size (radius smaller than
5 nm) of Au, the calculated values agree well with the experi-
mental data. Because the fusion model is proposed by considering
a free-standing nanoparticle, which is not applicable to the
melting experiment of Sn carried out on silicon monoxide sub-
strate and carbon substrate, the calculated melting temperatures
of Sn deviate from the experimental data at low particle size.
In addition, applying these equations, the other phase transi-
tion temperatures of dispersed systems can also be calculated.
5. Conclusions
The phase transition equation according to the additional
pressure on the curved surfaces of dispersed systems can be
applied to various kinds of phase transitions of dispersed systems,
and accurate differential equations for the phase transitions of
various kinds of dispersed systems can be obtained by the phase
transition equation. The phase transition equations can be used to
predict temperatures of phase transitions and to explain the
phenomenon of metastable states. The size of a dispersed phase
in a dispersed system has signicant effect on the phase transition
temperature. All temperatures of fusion, solidication (from a
large quantity of liquid to tiny crystal), condensation (from a large
quantity of vapor to tiny droplet), vaporization (from tiny droplet
to a large quantity of vapor), sublimation (from tiny crystal to
their vapor) and desublimation (from a large quantity of vapor to
tiny crystal) decrease with decreasing radius of the dispersed
phase. However, the bubble point temperature of a planar liquid
increases with decreasing absolute value of radius of the bubbles.
The depression of melting temperature for a nanowire is approxi-
mately half of that for a spherical nanoparticle with identical
radius. In summary, the phase transition temperature decreases
with decreasing radius of dispersed phase .
Acknowledgments
This work was supported by the National Natural Science
Foundation of China (No. 20773092) and by the Program for the
Top Science and Technology Innovation Teams of Higher Learning
Institutions of Shanxi Province of China.
References
[1] P. Pawlow, Z. Phys. Chem. 65 (1909) 545.
[2] K.J. Hanszen, Z. Phys. 157 (1960) 523.
[3] E. Rie, Z. Phys. Chem. 104 (1923) 354.
[4] H. Reiss, I.B. Wilson, J. Colloid Sci. 3 (1948) 551.
[5] A. Curzon, Ph.D. Thesis, University of London, London, 1959.
[6] J.P. Borel, Surf. Sci. 106 (1981) 1.
[7] P. Buffat, J.P. Borel, Phys. Rev. A 13 (1976) 2287.
[8] T.B. David, Y. Lereah, G. Deutscher, R. Kofman, P. Cheyssac, Philos. Mag. A 71
(1995) 1135.
700
800
900
1000
1100
1200
1300
1400
0 3 6 9 12 15 18 21
r/nm
T
/
K
Experimental data [7]
Experimental data [13]
Calculated by Eq. (10)
Fig. 3. Melting temperatures of Au nanoparticle with different radii.
280
310
340
370
400
430
460
490
520
0 5 10 15 20 25
r/nm
T
/
K
experimental data in silicon monoxide
substrate[32]
Experimental data in carbon substrate[32]
Experimental data [15]
Calculated by Eq. (10)
Fig. 4. Melting temperatures of Sn nanoparticle with different radii.
Y.-Q. Xue et al. / Physica B 408 (2013) 134139 138
[9] T. Castro, R. Reifenberger, Phys. Rev. B 42 (1990) 8548.
[10] M. Zhao, X.H. Zhou, Q. Jiang, J. Mater. Res. 16 (2001) 3304.
[11] M. Mitome, Surf. Sci. 442 (1999) L953.
[12] D. Kimberly, T. Dhanasekaran, Z.Y. Zhang, D. Meisel, J. Am. Chem. Soc. 124
(2002) 2312.
[13] J.R. Sambles, Proc. R. Soc. London A 324 (1971) 339.
[14] S.J. Zhao, S.Q. Wang, D.Y. Cheng, H.Q. Ye, J. Phys. Chem. B 105 (2001) 12857.
[15] S.L. Lai, J.Y. Guo, V. Petrova, G. Ramanath, L.H. Allen, Phys. Rev. Lett. 77 (1996) 99.
[16] J.H. Rose, J. Ferrante, J.R. Smith, Phys. Rev. Lett. 47 (1981) 675.
[17] F. Guinea, J.H. Rose, J.R. Smith, J. Ferrante, Appl. Phys. Lett. 44 (1984) 53.
[18] A. Safaei, J. Phys. Chem. C 114 (2010) 13482.
[19] A. Safaei, M.A. Shandiz, S. Sanjabi, Z.H. Barber, J. Phys. Chem. C 112 (2008) 99.
[20] M.A. Shandiz, A. Safaei, Mater. Lett. 62 (2008) 3954.
[21] H.M. Lu, P.Y. Li, Z.H. Cao, X.K. Meng, J. Phys. Chem. C 113 (2009) 7598.
[22] H.M. Lu, F.Q. Han, X.K. Meng, J. Phys. Chem. B 112 (2008) 9444.
[23] W.H. Qi, M.P. Wang, Mater. Chem. Phys. 88 (2004) 280.
[24] W.H. Qi, Physica B 368 (2005) 46.
[25] Q. Jiang, S. Zhang, M. Zhao, Mater. Chem. Phys. 82 (2003) 225.
[26] Q. Jiang, Z. Zhang, Y.W. Wang, Mater. Sci. Eng. A 286 (2000) 139.
[27] Y.F. Zhu, J.S. Lian, Q. Jiang, J. Phys. Chem. C 113 (2009) 16896.
[28] G.L. Allen, R.A. Bayles, W.W. Gile, W.A. Jesser, Thin Solid Films 144 (1986)
297.
[29] H.K. Kim, S.H. Huh, J.W. Park, J.W. Jeong, G.H. Lee, Chem. Phys. Lett. 354
(2002) 165.
[30] Y.J. Li, W.H. Qi, B.Y. Huang, M.P. Wang, S.Y. Xiong, J. Phys. Chem. Solids 71
(2010) 810.
[31] A.P. Chernyshev, Mater. Lett. 63 (2009) 1525.
[32] C.R. Wronski, J. Appl. Phys. 18 (1967) 1731.
[33] T. Bachels, H.J. Guntherodt, R. Schafer, Phys. Rev. Lett. 85 (2000) 1250.
[34] Z.X. Cui, M.Z. Zhao, W.P. Lai, Y.Q. Xue, J. Phys. Chem. C 115 (2011) 22796.
[35] G. Guisbiers, L. Buchaillot, J. Phys. Chem. C 113 (2009) 3566.
[36] M.F. Vladimir, D.Z. Edgar, J. Non-Cryst. Solids 265 (2000) 105.
[37] V.M. Samsonov, N.Y. Sdobnyakov, A.N. Bazulev, Colloids Surf. A 239 (2004)
113.
[38] C.J. Coombes, J. Phys. F 2 (1972) 441.
[39] S.F. Xiao, W.Y. Hu, J.Y. Yang, J. Chem. Phys. 125 (2006) 184504/1.
[40] O. Gulseren, F. Ercolessi, E. Tosatti, Phys. Rev. B 51 (1995) 7377.
[41] S.K.R.S. Sankaranarayanan, V.R. Bhethanabotla, Babu Joseph, Phys. Rev. B 74
(2006) 155441/1.
[42] S.K.R.S. Sankaranarayanan, V.R. Bhethanabotla, Babu Joseph, J. Phys. Chem. C
111 (2007) 2430.
[43] E.H. Kim, B.J. Lee, Met. Mater. Int. 15 (2009) 531.
[44] W.X. Zhang, C. He, J. Phys. Chem. C 114 (2010) 8717.
[45] B. Wang, G. Wang, X. Chen, J. Zhao, Phys. Rev. B 67 (2003) 193403.
[46] K.K. Nanda, S.N. Sahu, S.N. Behera, Phys. Rev. A 66 (2002) 013208/1.
[47] L. Miao, Venkat R. Bhethanabotla, B. Joseph, Phys. Rev. B 72 (2005) 134109/1.
[48] C.L. Yaws, Chemical Properties Handbook, McGraw-Hill Book Co, Beijing,
1999, P. (a) 154, 157; (b) 56, 83, 78, 104, 81, 107; (c) 185, 207, 210; (d) 154,
157; (e) 212, 234, 237.
[49] P.R. Couchman, W.A. Jesser, Nature 269 (1977) 481.
[50] Y.Q. Xue, Q.S. Zhao, C.H. Luan, J. Colloid Interface Sci. 243 (2001) 388.
[51] H.W. Sheng, K. Lu, E.M. Ma, Acta Mater. 46 (1998) 5195.
[52] S.J. Zhao, S.Q. Wang, H.Q. Ye, J. Phys. Soc. Jpn. 70 (2001) 2953.
[53] Y.N. Zhang, L. Wang, X.F. Bian, Acta Phys.-Chim. Sin. 19 (2003) 35.
[54] R.H. Perry, D.W. Green, Perrys Chemical Engineers Handbook, 6th, McGraw-
Hill, New York, 1984, pp. 3128. (Chinese version).
Y.-Q. Xue et al. / Physica B 408 (2013) 134139 139

You might also like