You are on page 1of 13

RSC Advances

View Article Online


PAPER View Journal | View Issue

System-dependent melting behavior of icosahedral


anti-Mackay nanoalloys
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

Cite this: RSC Adv., 2013, 3, 21981

Hassan Yousefi Oderji,ab Hassan Behnejad,b Riccardo Ferrandoc and Hongbin Ding*a

Anti-Mackay icosahedral clusters of composition Ag32M13, where M is either Cu, Ni, or Co, have been
recently shown to possess special structural stability both by calculations and experiments. These
nanoalloys assume a core–shell arrangement, with an icosahedral core of 13 M atoms surrounded by an
Ag shell of anti-Mackay structure. In this paper we study the melting of these three nanoalloys, showing
that, despite the close similarity of the structures, melting takes place through quite different
mechanisms. In particular, we find that Ag32Co13 and Ag32Ni13 present a premelting phenomenon
which involves only the shell of the cluster while the core melts at higher temperatures, in agreement
with previous calculations. On the contrary, in Ag32Cu13, melting occurs through stages that involve the
shell and the core at the same time. These findings are rationalized in terms of the different features of
Received 4th July 2013
Accepted 10th September 2013
the energy landscape of these nanoalloys. Our simulations, in which special care has been devoted to
avoid non-ergodicity problems, show also that the particles keep their core–shell structures even in the
DOI: 10.1039/c3ra43401j
liquid phase, indicating an incomplete miscibility of Ag with Ni, Co or Cu at the nanoscale up to quite
www.rsc.org/advances high temperatures.

1 Introduction catalytic effects, it is useful to know the highest temperature at


which the structure is stable and the active atoms are located on
Nanoalloys are recently attracting notable interest because of the surface.
their optical, mechanical and magnetic behaviors that may Nanoalloys have more complex phase transition mecha-
differ from those of both the corresponding bulk alloys and nisms than pure nanoparticles. Single-component nanoparticle
pure nanometals. These differences arise from the interplay already show rather complicated melting behaviors. The
between nanoscale and alloying effects, which in principle occurrence of melting is not at a distinct melting point, but
depend on size and composition. In fact, nanoalloy properties rather in a temperature range, which is strongly dependent on
can be modied and tuned not only by varying size and nanoparticle size. In addition to that, nanoalloy melting has
geometric structure, but also by varying composition and shown a strong dependence on composition. For example, a
atomic ordering within the nanoparticle.1 These features make single Ni or Cu impurity can considerably change the melting
nanoalloys even more promising than pure clusters to develop behavior of pure Ag icosahedra of sizes of several hundred
applications in several biomedical, catalytic, optical and elec- atoms.11
tronic technologies.1–10 However, there are still some open questions about the
A crucial issue concerning the possibility of applications is effects of geometric structure and composition on phase tran-
related to the stability of nanoalloy structures to changes in sition mechanisms in nanoalloys; e.g., whether the particles
their environment and external conditions. In particular, it is with similar structures can exhibit very different melting
very important whether a given nanoalloy structure is stable behaviors, and whether the melting of the shell is prior to the
with respect to changes in temperature. For this reason, the melting of the core in core–shell nanoalloys. The latter is
study of thermal stability and phase transitions is important to particularly interesting for nanoparticles whose shell thickness
nd and design new applications for nanoalloys. For example, is a single atomic layer. In this case, an interesting question
whenever only the shell atoms of a core–shell nanoalloy have arises about the possibility of nding a temperature range in
which a single-layer liquid is oating above an otherwise solid
a
School of Physics and Optical Electronic Technology, Dalian University of Technology, core.
Dalian, Liaoning 116024, P. R. China. E-mail: hding@dlut.edu.cn; Fax: +86 411 Core–shell nanoparticles with shells of monatomic thickness
84706730; Tel: +86 411 84706730
b
have been the subject of several computational studies in the
Department of Physical Chemistry, School of Chemistry, University College of Science,
last decade. Most of these studies focused on nanoparticle
University of Tehran, Tehran 14155, Iran
c
Dipartimento di Fisica, dell'Università di Genova, via Dodecaneso 33, 16146, Genova,
structures.9,12–16 In particular, it was computationally predicted
Italy. E-mail: ferrando@sica.unige.it; Fax: +39 010 311066; Tel: +39 010 3536214 in ref. 9 that core–shell polyicosahedral nanoparticles are

This journal is ª The Royal Society of Chemistry 2013 RSC Adv., 2013, 3, 21981–21993 | 21981
View Article Online

RSC Advances Paper

especially stable for small sizes. In that study, several magic using long-length simulations and from the multiple histogram
sizes and compositions were singled out for AgCu and AgNi method, which is used to remove the nonergodicity problem,
clusters, among which the highly symmetric Ag32M13 anti- and thus correctly accounting for all possible congurations
Mackay icosahedron.17,18 and their occurrence probabilities. This allows a much more
The anti-Mackay icosahedron has a core of 13 M atoms precise interpretation of the phenomena occurring at
covered by a faulted Ag icosahedral shell, which is a twin-like increasing temperatures, and of their relationship with the
face capping of the underlying triangles. In this structure type, features of the heat capacity behavior.
one of the core atoms is covered by 12 atoms at icosahedral Moreover, we present a more detailed description about the
locations (Fig. 1(a) and (b)) while the arrangement of shell phase transition mechanism by analyzing several quantities.
atoms on top of the second layer in every cluster facet is as We have quenched the instantaneous congurations to analyze
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

shown in Fig. 1(c). Fig. 1(d) shows the complete anti-Mackay the occupation of different local minima and used the distance
core–shell icosahedral structure. The calculations predicted this uctuations and the correlation functions to analyze simulated
structure to be the lowest in energy also for AgCo.19 trajectories. The latter quantities are especially important due
Very recently, the possible magic character of this cluster to the dynamical nature of phase coexistence in nanoparticles.22
structure has been supported by experimental results for AgCu.20 While only the most thermodynamically stable phase is
In fact, the abundance spectra of AgCu clusters produced in gas observable in bulk materials, phase coexistence in nano-
phase have singled out a prominent peak corresponding to the particles might be established only if there is a local minimum
composition Ag32Cu13, which is thus a magic composition. in the free energy for every phase-like form.23 We also developed
Therefore the study of the thermal behavior of nanoparticles at the self space–time correlation method to discriminate between
this specic composition is now of special interest. the diffusive and non-diffusive motion of each atom and pair
In general, the melting behavior of core–shell poly- distribution functions to examine the nanoscale miscibility of
icosahedral is quite interesting because these systems can Ag and M atoms, which are not miscible in bulk. Our results will
present a rich phenomenology at increasing temperature,9,19,21 show that the sole analysis of the heat capacity curve is not at all
being prototypical systems for studying melting in clusters of sufficient for understanding what are the crucial steps in the
weakly miscible metals. Concerning composition Ag32M13, the melting of these nanoparticles.
melting behavior has been studied for M ¼ Ni and Co, whereas Our results will show also that systems with the same
results related to AgCu are still lacking. geometric structure and chemical ordering can indeed present
In fact, Kuntová et al.19 computationally studied the melting quite different melting behaviors. In particular, the melting
of Ag32Ni13 and Ag32Co13. Their work was focused especially on behavior of AgCu will be signicantly different from that of
the rst stages of the melting phenomenon, which involve AgNi, with AgCo presenting intermediate characteristics.
mostly the behavior of the shell. Moreover, their calculated heat This paper is organized as follows. Simulation and analysis
capacity curves suffered from a limited sampling, so that they methodology are explained in Section 2. The results are pre-
were affected by considerable noise, which rendered their sented in Section 3. Finally, Section 4 contains the discussion of
interpretation rather difficult. the different results and the conclusions.
In the present work we consider Ag32M13 for M ¼ Cu, Ni and
Co. We demonstrate that accurate heat capacity curves can be 2 Simulation and analysis methodology
obtained from the energy uctuations in the NVT ensemble
2.1 Simulation details
The simulations have been accomplished using the DL_POLY
molecular dynamics parallel simulation package24 under two
different procedures. The rst procedure was composed of two
stages: the rst stage for melting the particles, by increasing the
temperature at a constant rate of 1 K ns1 up to vaporization,
and the second stage for sampling the phase space trajectories,
by simulating the systems during 500 ns at temperature inter-
vals of 10 K using the initial conditions obtained from the rst
stage. In the second procedure, all simulations, regardless to
the temperature, were started from the global minimum
structure, but with longer time devoted to the equilibration part
of simulations (50 ns). This makes every simulation to be
independent from the other simulations. The Gupta potential
was used to model the interaction between atoms and the
velocity-Verlet algorithm was used to integrate the equations of
motion in the NVT ensemble. The length of time-step was
Fig. 1 The icosahedron structure [CPK model (a), VDW model (b)]; the anti-
adjusted to sufficiently small value of 1 fs to trace the atomic
Mackay growth of the second layer on the icosahedron core (c); the anti-Mackay trajectories during the fast conguration transitions occurring
core–shell VDW model (d). in liquids. The Evans thermostat25 was used to conserve the

21982 | RSC Adv., 2013, 3, 21981–21993 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper RSC Advances

temperature. Simulated positions and velocities of the atoms 3 hE 2 it  hEi2t


were saved every 5 fs for further analysis. The Gupta model and Cv ðTÞ ¼ NkB þ ; (1)
2 kB T 2
the multiple histogram method have been explained in detail in
our previous work26 and in ref. 27–30. where E is the potential energy, N is the number of atoms, and
The parameters of the Gupta potential model used in this kB is the Boltzmann constant. However, to calculate the heat
study are from ref. 19, 31 and 32, which have been favorably capacity for the shell and core separately, none of them can
checked against a density-functional calculations.14 It should be individually represent a canonical ensemble and thus eqn (1) is
noted that the A parameter as dened in ref. 26 or in DL_POLY not established. In this case the heat capacities can be obtained
codes is twice in magnitude of the A parameter in ref. 19. directly from the derivation of the energy with respect to
We note that the Gupta potential produces lowest-energy temperature. We used Bézier interpolation35 to smooth the
results because the curve is guaranteed to pass trough the rst
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

structures of AgCu nanoclusters in gas phase9 that agree well


with the magic compositions observed in recent experiments.20 and the last data points, but in general will not pass through any
The Gupta model has been employed to predict the melting of the internal points that in some intermediate temperatures
temperature of bulk Ag, Cu, Co and Ni materials by Cleri et al.,27 in our study possess lower statistics precisions than two limits
and it has been found that the Gupta model well reproduces the due to the nonergodicity.
melting temperature of transition metals and alloys. This 2.2.2 Steepest descent quenching method. According to
potential model has also been employed in several simulations Berry and Smirnov22 the melting in nanoparticles may be
studying on the melting of nanometals and nanoalloys.32,33 considered as a result of the conguration transitions. While
the global minimum and at most some of the local minima are
the exclusive congurations in the solid phase, the fast transi-
2.2 Analysis methods tions between many high energetic congurations occur in the
The details of the melting mechanism in nanoparticles may be liquid phase. Hence, the overall thermal behavior of the nano-
obtained by several methods through the simulation data. particles can be determined from histograms of the congura-
Caloric curves can be used to obtain an overall image in melting tional energy. Due to the atomic vibrations, however, the
processes showing the latent heat and the temperature region instantaneous positions of atoms in trajectories are different
in which the melting occurs. The melting point and the melting from the positions in equilibrium congurations. We used the
stages or probable pre- or post-melting phenomena can be quenching steepest descent method to remove these displace-
respectively attributed to the peak and the number of peaks or ments and thus to bring the atoms to their equilibrium posi-
shoulders in the heat capacity curves. The excited congura- tions. The coupling of this method to MD36 has been previously
tions occurring in the melting process can be identied by the used to investigate the physical mechanisms of transition from
steepest descent quenching method. The phase behaviors of the rigid to nonrigid behavior of the small Argon clusters.37
shells and cores can be individually determined by the uctu- In this method, the following differential equation,
ations in the distances between the same atoms. Atoms located
at similar positions in solids and atoms of the same kind in r_ ¼ VE, (2)
liquids should feel the same mean forces, so the environment of
atoms can be compared using the power spectra. Similarly, the is solved for some selected congurations on MD trajectories at
kind of motion for a typical atom, i.e. non-diffusive in solids and determined time intervals.37 In our simulations, the congura-
diffusive in liquids, can be deduced from the correlation of the tions were selected from time intervals of 1 ns.
position of an atom by itself at different times. Finally, the 2.2.3 The distance uctuation criterion. A commonly used
miscibility of two constituents of each nanoalloy in the liquid criterion for singling out the melting transition is based on
phase can be surveyed from the radial distribution of atoms analyzing the relative displacements of atoms with increasing
around each constituent. temperatures. This is based on the following line of reasoning.
2.2.1 Caloric and heat capacity curves. The caloric curve At temperatures below the melting point, the uctuations in the
and the heat capacity are two common thermodynamical mean distances between the pairs of atoms have relatively small
criteria which are oen used to identify the melting transi- values because atoms can only have vibrating motions in the
tion.1,19,23,30,34 The caloric curve reports the average energy of the solid phase. However, at temperatures higher than the melting
nanoparticle as a function of temperature. For the completely point, atoms make diffusive motions and the mean distances
solid or liquid phases, the caloric curve is a quite gradual between atomic pairs uctuate much more signicantly. Thus,
increasing function of temperature while for coexistence region the magnitude of distance uctuations can be used as a crite-
it usually presents a drastically change relating to the latent rion to determine the melting point. A well-known measure for
heat. In contrast, the heat capacity of a nanoparticle as a this purpose is the Lindemann parameter, d, which represents
function of temperature represents the phase transitions as the root-mean-square (rms) relative bond length uctuation:38–40
broad peaks, in which the number and the shape of peaks reveal rD
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
E  2
some features of the melting mechanism. rij ðtÞ2  rij ðtÞ t
1 X
N X
N
t
In the canonical ensemble, the constant-volume heat d¼   ; (3)
capacity, Cv, is related to the potential energy uctuations as NðN  1Þ i¼0 j¼0; jsi
rij ðtÞ t
follows:

This journal is ª The Royal Society of Chemistry 2013 RSC Adv., 2013, 3, 21981–21993 | 21983
View Article Online

RSC Advances Paper

where rij is the distance between ith and jth atoms and N is the 1 XM

number of atoms. In the present research, we calculate the dshell GsðiÞ ðr; tÞ ¼ d½r  jri ðtk Þ  ri ðtk þ tÞj; (6)
4pr2 M k¼1
and dcore values, instead of d for the whole of the nanoparticle,
to discriminate between the melting of shells and cores. where index k runs over a total of M selected time origins from a
2.2.4 Velocity autocorrelation function and power spec- system at equilibrium, and d is the Dirac delta symbol.
trum. All atoms located at similar positions in a crystalline solid 2.2.6 Pair distribution function. Pair distribution function
and all atoms of the same kind in a liquid system have similar (PDF) is dened as the probability of nding an atom at distance
environments and feel equal mean forces, so they are expected r from another atom.44 PDF is particularly important because it
to exhibit similar power spectra. Conversely, the phase transi- can be related to the results of neutron and X-ray scattering
tions and the changes in the structure of a system can be experiments.41,45 In present research, we used PDF to examine
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

determined by comparing the power spectra obtained at the miscibility of the nanoalloy constituents in the liquid phase.
different temperatures. There were several slightly different relations to calculate
The velocity autocorrelation function (VACF) shows how the PDF in nanoparticles and clusters.46 In this work, we used the
present velocity of a particle is related to its previous value and following relation in which PDF is averaged over all congura-
how it affects its subsequent velocities. For liquid systems, tions along a trajectory, and the sum extends to all mutual
VACF decays to zero aer a few collisions and for solid systems connections between atoms, namely N(N  1)/2, as follows:
it decays aer a few oscillations.41 Hence, the underlying * +
X
N X N
 
vibration frequencies for the individual atoms (power spectra) gðrÞ ¼ d r  rij (7)
can be obtained by the Fourier transform of the velocity auto- i¼0 j.i t
correlation functions:32,42
Consequently, the number of peaks represents the number of
ðN locations while positions and the area of peaks represent
UðiÞ ðuÞ ¼ 2 C ðiÞ ðtÞcosðutÞdt; (4) respectively the distances and ratio between locations in the
0
lattice of the solid structure. For example, the Cu–Cu pair distri-
where, C (i)(t) is the normalized velocity autocorrelation function bution functions at T ¼ 250 K (see red line in Fig. 13(a)) shows 3
of ith atom:43 peaks at 2.45, 4.0 and 4.75 Å with the areas of 42, 30 and 6, cor-
responding to the icosahedron structure with 13 similar atoms.
hvi ðt0 Þvi ðt þ t0 Þi
C ðiÞ ðtÞ ¼ ; (5)
hvi ðt0 Þvi ðt0 Þi
where, vi is the velocity of ith atom, and the angular brackets
indicate a time average over the simulations by different time
origins, t0. During the simulations, the particles' centers of
mass were xed and the rotational degrees of freedom were
restricted, so the velocities in eqn (5) are only related to the
internal degrees of freedom. In spite of the articial rescaling of
velocities in the NVT ensemble, VACFs can still be calculated
precisely by averaging over numerous samples while the simu-
lations in the NVE ensemble would suffer from the instability of
the temperature, especially in the coexistence phase region.
2.2.5 Self space–time correlation function. Although power
spectra show the environmental changes for each atom through
its instant velocity values, yet the status and the kind of motion
for individual atoms may remain unclear to some extent. In
order to overcome this problem, we recently developed a
method to discriminate between the diffusive and non-diffusive
motions of each atom using the self space–time correlation
function (SSTCF).26 SSTCF measures the probability that an
atom is at position r at time t given that the same atom was at
the origin r ¼ 0 at the initial time t ¼ 0.43 In this method, SSTCF
values are calculated at a close distance to each atom, namely
r ¼ 0.2 Å. For an atom with non-diffusive motions, as what
usually occurs in solids, SSTCF behaves as a constant function
aer a few oscillations, while for an atom with diffusive
motions, as what usually occurs in liquids, SSTCF behaves as a
decaying function aer few oscillations.
The SSTCF, Gs(r, t), can be evaluated from simulation data Fig. 2 Caloric curves (configuration energies per atom) for the core, shell and the
for each atom as follows: whole particles of Ag32Co13 (a), Ag32Ni13 (b), and Ag32Cu13 (c) anti-Mackay nanoalloys.

21984 | RSC Adv., 2013, 3, 21981–21993 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper RSC Advances

3 Results and discussion what happens rst, at least in AgNi and AgCo) modies the core
energy too, due to a change in the interaction between core and
We start off our results on investigation into the melting shell. The shell is one-layer thick, so that it is not possible to
mechanism of anti-Mackay nanoalloys with presenting caloric modify the shell without changing appreciably the energy of the
curves. Since the total potential energy of the cluster is written, core at the same time.
within the Gupta model, by a sum of individual atomic energies, Compared to the heat capacity curves reported in ref. 19, the
it is possible to associate energies to the core and to the shell present curves are obtained by means of a much richer statistics,
separately, by summing atomic energies on core and shell which is thus able to smooth the sharp changes in the thermal
atoms separately. However, it must be kept in mind that both evolution of the heat capacities, risen from nonergodicity.19
shell and core energies contain the interaction between the two The present calculations are thus able to improve the earlier
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

parts of the cluster. This core–shell interaction cannot be predictions stating two distinct stages for the melting of the
separated from the rest, due to the many-body nature of the core and shell.19 This is due to removing the nonergodicity at
potential, which cannot be written as a sum of pair interactions. some temperatures that would remain undetected in short-
Fig. 2 shows the variation of potential energy of core, shell length simulations of the anti-Mackay nanoalloys. For example,
and whole particle as functions of temperature. The caloric Fig. 4 that shows the quenched congurations of Ag32Ni13 along
curves show gradual increases, with smooth steps extending on the long length of simulation (1000 ns) at T ¼ 760 K clearly
wide temperature ranges. demonstrates that the system is trapped into a potential well
The study of the derivative of the caloric curve, which gives between t ¼ 780 and t ¼ 860 ns. The occurrence and the dura-
the heat capacity, shows more evident features, although it is tion of falling into the potential well is governed by probability
still quite smooth. Results about heat capacities for Ag32Co13, that is related to the temperature. Thus, the heat capacity value
Ag32Ni13, and Ag32Cu13 nanoalloys are shown as functions of obtained by this simulation (cross symbol in Fig. 3(b)) would
temperature in Fig. 3(a)–(c). It is evident that the separation into likely be different from those obtained by other independent
core and shell is not adding much information, because a simulations. This also can be considered as an origin for the
change in shape of the shell only (as we will see below this is dynamic phase-coexistence in nanoparticles. At a constant
temperature and pressure, it is possible to nd a nanoparticle in
a certain phase and then in another phase aer a while.
In Fig. 3 all heat capacity curves show qualitatively similar
behaviors, consisting in a main peak at high temperatures and a
shallow secondary peak at considerably lower temperatures.
However, as we will see in the following, the simple inspection
of the heat capacity curves is not sufficient to understand how
the melting phenomenon is taking place. For example, Fig. 3(c)
indicates that Ag32Cu13 melts during two distinct stages at T ¼
420 K and 840 K, respectively. It is tempting to attribute the
lower temperature to the melting point of the shell and the
higher one to the melting of the core but further investigation
will show that this is not the case. It should be noted that, since
the nonergodicity in simulations of Ag32Cu13 starts at low
temperatures, the smoothed results for the core and the shell at

Fig. 3 Heat capacity values per atom, shown in units of the Boltzmann constant,
for the core, shell and the whole particles of Ag32Co13 (a), Ag32Ni13 (b), and
Ag32Cu13 (c) anti-Mackay nanoalloys at different temperatures. The solid lines
with error bars are from the multiple histogram method and the cross symbol on Fig. 4 The energy of quenched configurations along very long-length simulated
the panel (b) is from very long-length simulations shown in Fig. 4. trajectory of Ag32Ni13 anti-Mackay nanoalloy at T ¼ 760 K.

This journal is ª The Royal Society of Chemistry 2013 RSC Adv., 2013, 3, 21981–21993 | 21985
View Article Online

RSC Advances Paper


Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

Fig. 6 Some of excited configurations for the anti-Mackay nanoalloys along the
melting reaction path obtained by quenching the simulated trajectories at
temperatures higher than T ¼ 300 K for Ag32Cu13, T ¼ 460 K for Ag32Co13, T ¼
480 K for Ag32Ni13 nanoalloys. The configurations of (a) to (f) are excited only due
to one or two Ag atom(s) while the configurations of (g) to (i) are excited due to
the core excitations. The excitation energies of these configurations are presented
in Table 1.

temperature of the shell. For Ag32Co13 (Fig. 5(b)) the core starts
to be excited just when the shell is already completely melted
(T x 610 K), while the core of Ag32Ni13 remains unchanged aer
Fig. 5 The mean distance fluctuation for the shell and core atoms of Ag32Cu13, the melting of the shell (occurring at T x 630 K) up to about
Ag32Ni13, and Ag32Co13 anti-Mackay nanoalloys as different temperatures. (The 720 K. Therefore, the rst melting stage of Ag32Cu13 that was
upper linear regimes are shown by solid lines and their starting points for shell predicted by the heat capacity criterion and the multiple
atoms are highlighted by vertical dotted lines.)
histogram to be at temperatures around 400 K (Fig. 3(c))
involves the participation of atoms from both the core and the
shell. Moreover, from Fig. 5(b) and (c) it follows that the pre-
temperatures less than 540 K would not be obtained by melting stages for the Ag32Co13 and Ag32Ni13 at temperatures
reasonable statistical precision and are excluded [see Fig. 3(c)]. lower than the T ¼ 610 K and 720 K, respectively, are merely
We also calculated the mean distance uctuations for the related to the shell atoms. This agrees with the conclusion that
core and shell atoms separately to survey the phase status of the in these two systems the shell melts before the core.19 The high
core while the shell is completely melted. Fig. 5 shows two values of dcore at the rst stage in the melting of Ag32Cu13
temperature intervals for each nanoalloy in which d linearly indicate the high level of participation in the rst stage by the
increases with temperature. For example, for the dshell values of core. Hence, two distinct stages in the melting of Ag32Cu13
Ag32Co13 (Fig. 5(b)), the rst linear regime occurs at tempera- should not be attributed exclusively to the melting of the core or
tures lower than T ¼ 430 K and the second one occurs at the shell. Later we will see this is due to the relative excitation
temperatures beyond T ¼ 650 K. If we suppose the rst linear energies corresponding to the core and shell atoms.
regime, occurring at lower temperatures, is related to the solid Quenching the congurations along the simulated trajecto-
and the second one, occurring at higher temperatures, is related ries of Ag32Ni13, Ag32Cu13 [Fig. 7 and 8] and Ag32Co13 at different
to the liquid phase, the relative status of the core and the shell temperatures indicate that they have similar excitation cong-
in the melting of core–shell nanoalloys can be deduced from the urations as shown in Fig. 6 and the corresponding energies as
dcore values just at temperatures in which the shells are presented in Table 1.
completely melted (marked by vertical dotted lines in Fig. 5). At low temperatures, all of the quenched congurations have
Fig. 5(a) demonstrates that the excitations in core of Ag32Cu13 the anti-Mackay structure (Fig. 1(d)), indicating the solidity of
occur beyond T x 360 K (note that the uctuations between T ¼ the nanoalloys. As temperature increases above 400 K, other
360 K and 400 K are due to the nonergodicity and in the real minima begin to be populated.19 The rst excited conguration
physical time scales excitation is expected to start beyond T x for Ag32Ni13 occurs approximately at T ¼ 480 K in which, over
360 K), a somewhat lower temperature than the melting one of the anti-Mackay faces, the central Ag atom is replaced

21986 | RSC Adv., 2013, 3, 21981–21993 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper RSC Advances

second Ag atom, the central atom of an anti-Mackay face, is also


replaced and forced by a vertex Ag atom to locate in a fourfold
position, creating the second hole (Fig. 6(e)). Again, due to the
relative positions of two holes, the higher excited congurations
occur at rather higher temperatures (Fig. 6(f)). Increasing the
temperature to T ¼ 760 K leads to the excitation congurations
involving the core atoms in which one of the Ni atoms relocates
from a vefold location to a threefold one above the surface of
core (Fig. 6(g)–(i)). The core excited congurations have lower
energies than some of the excitations occurring at lower
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

temperatures (see Table 1). This indicates the existence of


rather high energy barriers between the global minimum and
the core excited congurations. Such barriers hinder the reverse
path towards the global minimum and thus cause the cong-
urations to survive at the excited states for a longer while.
The similar excitations are also observed for Ag32Co13 at
somewhat different temperatures and for Ag32Cu13 at the
temperatures shown in Fig. 8. However, while the behavior of

Fig. 7 The energy of quenched configurations of Ag32Ni13 anti-Mackay nano-


alloy (a-panels) and their corresponding cores (b-panels) along the simulated
trajectories at different temperatures. The solid lines in a-panels are corresponding
to the energies of global minimum Fig. 1(d) [red] and the configurations shown in
Fig. 6(a) [green], (d) [blue] and the solid lines in b-panels are corresponding to the
cores in Fig. 1(d) [red] and Fig. 6(g) [red], and (h) [green] or (i) [blue].

and forced by a vertex Ag atom to locate in one of the opposite


fourfold positions above the particle surface (Fig. 6(a)). Further
excited congurations arise from moving this Ag atom to other
fourfold positions (Fig. 6(b) and (c)). The energy differences
between these congurations are only 0.0, 0.002 or 0.003 eV,
depending on the relative positions of the hole and fourfold
position, thus we consider all of them to be the rst excitation
with almost degenerate states. The second excitation is
observed rstly at T ¼ 530 K by relocating the Ag atom from
Fig. 8 The energy of quenched configurations of Ag32Cu13 anti-Mackay nano-
fourfold to one of the ve threefold positions around the hole alloy (a-panels) and their corresponding cores (b-panels) along the simulated
(Fig. 6(d)). The third excitation occurs at T ¼ 540 K when the trajectories at different temperatures. (Legends are as Fig. 7.)

Table 1 The excitation energies for the configurations shown in Fig. 6 for Ag32Ni13, Ag32Co13 and Ag32Cu13 nanoalloys in terms of eV

a b c d e f g h i

Ag32Ni13 0.468 0.466 0.465 0.601 0.968 0.843 0.864 0.830 0.676
Ag32Co13 0.436 0.431 0.431 0.569 0.900 0.772 0.646 0.661 0.507
Ag32Cu13 0.266 0.253 0.251 0.400 0.577 0.497 0.325 0.369 0.160

This journal is ª The Royal Society of Chemistry 2013 RSC Adv., 2013, 3, 21981–21993 | 21987
View Article Online

RSC Advances Paper

Ag32Co13 resembles that of Ag32Ni13, the silver–copper cluster different structures (anti-Mackay and truncated octahedron),
shows some signicant differences. What makes the melting of nonergodicity can be observed as a result of core excitation (in
Ag32Cu13 to be different from two other nanoalloys is the energy anti-Mackay) or core deformation (in truncated octahedron).
of the third excited conguration involving the shape change of This is due to the energy barriers that are signicantly higher for
the core (Fig. 6(i)) which is even lower than the energy of the the core than the shell atoms. The energy distribution or the
primarily excited conguration involving only the shell atoms uctuation in conguration energies of a nanoparticle at
(Fig. 6(a)). We note that global optimization searches show that constant temperatures are stochastic. Then, once the core is
there is an even lower excited isomer with deformed core, which accidentally excited or deformed, the reverse phenomena
is however not encountered in the MD simulations. requires more energy and then occurs by lower probabilities.
The effect of the low-energy excited isomers with deformed It is however important to note that although these
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

core that are found in Ag32Cu13 can be better understood by phenomena are the cause of signicant changes in the cluster
analyzing the energy histograms of the quenched congura- shape, they only produce shallow secondary features in the heat
tions, which are reported in Fig. 7 and 8 for Ag32Ni13 and capacity curves whose main peaks are indeed found at much
Ag32Cu13 at several temperatures. It is clear that the melting of higher temperatures. Therefore, these nanoalloys present a
Ag32Ni13 (and Ag32Co13) starts by the deformation of the shell quite peculiar behavior, since their shape changes signicantly
only, as it was demonstrated in ref. 19, while the core retains its (even though being still core–shell) and dynamically, but this
icosahedral shape. On the contrary, in Ag32Cu13, the melting happens already at a temperature which is much lower than the
involves from the beginning transitions to isomers in which the temperature of the main peak in the heat capacity. In the
core has changed its shape. following we exploit the power spectra of the individual atoms
In all cases, the probability of nding the particle in its to elucidate this peculiar behavior.
excited conguration increases with temperature. It is infor- Power spectra benet from two features to show the phase
mative to compare this mechanism with the phase transition behavior of a particle. If the particle is liquid (i) all the same
mechanism of the TO Pd24Pt14 nanoalloy in which the melting atoms show similar power spectra and furthermore (ii) the
occurs aer a solid–solid transition from the truncated octa- intensity at the zero frequency is signicant.
hedral to the icosahedral.26 However, for both of these two Let us consider the low-temperature situation. Fig. 9(a) (and
similarly Fig. 10(a) and 11(a)) presents four different power
spectra corresponding to the anti-Mackay structure of the solid

Fig. 9 Power spectra of all constituent atoms of Ag32Cu13 anti-Mackay nano-


alloy at different temperatures (black: central atom, red: 12 atoms in icosahedral
locations, green: 12 Ag atoms at the vertexes of triangle anti-Mackay faces, blue: Fig. 10 Power spectra of all constituent atoms of Ag32Co13 anti-Mackay nano-
20 Ag atoms at the center of triangle anti-Mackay faces). alloy at different temperatures (legends are same as Fig. 9).

21988 | RSC Adv., 2013, 3, 21981–21993 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper RSC Advances

spectra only, one for all Ag atoms and one for all Co atoms. On
the other hand, the intensity of the power spectra at zero
frequency is low, as in the solid case.
In Fig. 11(f) the temperature (760 K) is such that, for Ag32Ni13,
the shell is melted but the core is not melted, again according to
the criterion of the atomic relative displacements (see Fig. 5(c)).
In this case, the shell atoms present a single spectrum, as in the
liquid case, but the core atoms show different spectra for
different atoms, with the central atom clearly distinguishable.
This agrees with the previous observation of liquid-like character
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

of the shell and solid-like character of the core. However, also in


this case, the power spectra have low intensity at zero frequency.
These results about Ag32Co13 and Ag32Ni13 indicate that there
is a further step must be accomplished to achieve the complete
melting of the structures, as indicated also by the position of the
main peak of the heat capacity curve, which occurs at tempera-
tures signicantly higher than 760 K in both systems.
The essential difference between liquids and solids is in the
kind of atomic motions that is diffusive for liquids. The diffu-
sivity of atomic motions were examined by G(i) s (r, t) functions in
this work. Fig. 12 presents the results of Gs(i)(r, t) at a constant
distance (r ¼ 0.2 Å) for Ag32Cu13, Ag32Co13, and Ag32Ni13
nanoalloys for some selected temperatures. Fig. 12(a) shows the
typical behavior of atoms of the anti-Mackay solid structure at

Fig. 11 Power spectra of all constituent atoms of Ag32Ni13 anti-Mackay nano-


alloy at different temperatures (legends are same as Fig. 9).

phase: the black one corresponds to the central atom in the core
of the nanoalloy, the red one corresponds to the 12 icosahedral
core atoms, the green and blue ones correspond to the 20 and 12
shell atoms at the centers and vertexes of the triangle anti-
Mackay faces, respectively. The core atoms are compressed by
forces from the shell, so their corresponding frequencies are
higher than those of the shell atoms; the peak corresponding to
the central atom surrounded by two layers represents the highest
frequencies. Nonequivalent atoms in the anti-Mackay structure
present signicantly different power spectra. Ag vertex atoms
present quite different spectra from those of other shell atoms.
On the opposite high-temperature side, Fig. 9(j), 10(j) and
11(j) demonstrate that the nanoparticles are completely melted
because atoms of the same kind in the liquid phase have similar
power spectra and the gures only show two power spectra
corresponding to the Ag and M atoms of Ag32M13 nanoalloys.
Moreover, the relatively high intensity values of the power
spectra at zero frequency indicate the diffusive motions,
occurring commonly in liquids. Thus, the solid and liquid
phases are distinguishable using the power spectra.
For intermediate temperatures, the situation is more
complex, presenting some features of both limiting cases. For
example, in Fig. 10(f) the temperature (750 K) is such that, for
Ag32Co13, both core and shell are melted according to the
criterion of the atomic relative displacements (see Fig. 5(b)). Fig. 12 Self space–time correlation function of all constituent atoms of
Accordingly, the power spectra retain some features of the Ag32Co13, Ag32Cu13, and Ag32Ni13 anti-Mackay nanoalloys at different temper-
liquid high-temperature case, such as the presence of two power atures (legends are same as Fig. 9).

This journal is ª The Royal Society of Chemistry 2013 RSC Adv., 2013, 3, 21981–21993 | 21989
View Article Online

RSC Advances Paper

low temperatures. The black curve, corresponding to the central


atom, has the lowest values and the most oscillations. The red
curves are related to the icosahedral locations of the core and
the green and blue ones are related to the locations at the center
and vertexes of the triangular anti-Mackay faces, respectively.
The G(i)
s (r, t) values for all of the atoms tend towards constant
values at low temperatures. A comparison between Fig. 12(a)
and (b) (the latter showing results at 600 K) demonstrates that
the correlation orders are inverted for the central and the shell
atoms. This is due to the cage effect on the central atom for
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

which an increase in vibration frequencies due to an increase in


temperature enhances the probability of nding the atom at r ¼
0.2 Å around its equilibrium position. Note also that all shell
atoms present essentially the same SSTCF, an indication of a
liquid-like behavior, while unequal core atoms are still distin-
guishable. In fact, according to Fig. 5(b), the shell should be
liquid and the core still solid at that temperature. The shell
atom curves in Fig. 12(b) are also showing the hint of a slow
decay with a large time constant. However, the presence of some
short-time weak oscillations for all atoms is a feature that is
closer to a solid-like than to a liquid-like behavior. Increasing
temperature further, one nds the behavior of Fig. 12(c), which
clearly shows that both the core and shell atoms of Ag32Co13
exhibit diffusive motions at T ¼ 750 K. Similar behaviors can be
observed for Ag32Ni13 and Ag32Cu13 (Fig. 12(d)–(e)), even though
in the latter case, at T ¼ 440 K shell and core atoms behave in

Fig. 14 Pair distribution functions of Ag–Co pair atoms in Ag32Co13 anti-Mackay


nanoalloy at different temperatures (blue: Ag–Ag, green: Ag–Co, red: Co–Co).

the same way, so that a difference between core and shell


cannot be singled out. Finally, the typical SSTCF behaviors for
the fast diffusive motions is shown in Fig. 12(f) for Ag32Cu13 at
T ¼ 840 K. Similar SSTCFs for atoms of the same kind indicate
similar environments and behaviors of the atoms in liquids.
Fig. 13–15 show the A–B pair distribution functions of
Ag32Cu13, Ag32Co13 and Ag32Ni13 nanoalloys, respectively, at
some temperatures, where A and B represent the constituent
elements of the nanoalloys. At low temperatures, A–B curves
characterize the solid structure of anti-Mackay (panel (a) in
Fig. 13–15), while at higher temperatures they present the liquid
behavior (panel (j) in Fig. 13–15). The only schematic difference
between the PDFs of liquid nano-droplets and bulk liquids is
due to the nite size for which the PDF values must be zero at
distances greater than the size of the particles. Panel (j) in
Fig. 13–15 demonstrate that Ag is not completely miscible with
Cu, Co and Ni. If they were miscible, the pair distributions of
Cu–Cu (Co–Co or Ni–Ni) would widen to the particle diameter.
The probability of nding two Cu (Co or Ni) atoms at the
opposite sides of the nanoparticle is not signicant, indicating
the inability of Ag atoms to be dissolved into M (M ¼ Cu, Ni, Co)
atoms to intersperse them. It also indicates that the melting of
the nanoalloys is not homogeneous as it can be easily deduced
from the snapshots taken from the nanoalloys at temperatures
Fig. 13 Pair distribution functions of Ag–Cu pair atoms in Ag32Cu13 anti-Mackay beyond the melting point shown in Fig. 16. These snapshots
nanoalloy at different temperatures (blue: Ag–Ag, green: Ag–Cu, red: Cu–Cu). show that even at the melting states, the nanoalloys generally

21990 | RSC Adv., 2013, 3, 21981–21993 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper RSC Advances

keep their core–shell structures with the segregation of Ag


atoms on surface.

4 Conclusion
In this work, several complementary methods were used to
investigate the melting mechanisms of Ag32Cu13, Ag32Co13 and
Ag32Ni13 core–shell nanoalloys. These systems are quite inter-
esting from this point of view, because they present a rather
complex melting behavior, which occurs through different
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

stages. In principle, these systems have several common features,


the most important being that they share the same type of global
minimum structure, a core–shell anti-Mackay icosahedron. The
main ndings of our investigation can be summarized as follows.
First of all, we note that reliable and smooth heat capacity
curves require very long simulations in order to collect a suffi-
cient statistics. This is especially true in the zone at which
premelting phenomena start because short simulations may
reveal problems of nonergodicity.
All the systems investigated here present qualitatively similar
plots of the heat capacity, whose most prominent feature is a
wide main peak at relatively high temperatures (maximum at
920, 1100 and 840 K for Ag32Co13, Ag32Ni13 and Ag32Cu13,
respectively). In addition, there is a shallow bump at much lower
temperatures (650, 800 and 400 K for Ag32Co13, Ag32Ni13 and
Ag32Cu13, respectively). From the inspection of the heat capacity
curves alone one might conclude that these systems present the
Fig. 15 Pair distribution functions of Ag–Ni pair atoms in Ag32Ni13 anti-Mackay same kind of melting behavior, with the main transformations
nanoalloy at different temperatures (blue: Ag–Ag, green: Ag–Ni, red: Ni–Ni).
occurring in correspondence with the main peak.

Fig. 16 Some snapshots taken from Ag32Ni13 at T ¼ 1390 K (a–c), Ag32Co13 at T ¼ 1480 K (d–f), and Ag32Cu13 at T ¼ 1450 K (g–i). These snapshots show that even at
temperatures beyond the melting point, Ag atoms are mostly segregated on surface.

This journal is ª The Royal Society of Chemistry 2013 RSC Adv., 2013, 3, 21981–21993 | 21991
View Article Online

RSC Advances Paper

A deeper analysis based on several different indicators References


(occupation probabilities of the minima, relative displacements
of atoms, velocity power spectra, space–time correlation func- 1 R. Ferrando, J. Jellinek and R. L. Johnston, Chem. Rev., 2008,
tions) reveal that this is not the case. In fact: 108, 845–910.
(i) All clusters undergo important changes well below the 2 U. Kreibig and M. Vollmer, Optical Properties of Metal
temperature of the main peak, i.e. in the temperature ranges of Clusters, Springer Series in Material Science 25, Springer,
the shallow bumps. These changes cannot be ascribed to solid– Berlin, 1995.
solid morphology transitions because they are associate to steep 3 E. Cottancin, J. Lermé, M. Gaudry, M. Pellarin, J.-L. Vialle,
increases of the atomic mobility. M. Broyer, B. Prével, M. Treilleux and P. Mélinon, Phys.
(ii) The behavior of Ag32Co13 and Ag32Ni13 on one side, and Rev. B: Condens. Matter Mater. Phys., 2000, 62, 5179–5185.
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

that of Ag32Cu13 on the other side are quite different. In fact, in 4 M. Haruta, Catal. Today, 1997, 36, 153–166.
the temperature ranges of the shallow bump, Ag32Co13 and 5 S. Lee, C. Fan, T. Wu and S. L. Anderson, J. Chem. Phys., 2005,
Ag32Ni13 present a premelting phenomenon which involves the 123, 124710.
external shell rst, and then the core at somewhat higher 6 A. M. Molenbroek, S. Haukka and B. S. Clausen, J. Phys.
temperatures.19 This is especially evident for Ag32Ni13, in which Chem. B, 1998, 102, 10680–10689.
the shell starts premelting at considerably lower temperatures 7 L. M. Molina and B. Hammer, Phys. Rev. Lett., 2003, 90,
than the core. On the contrary, in the temperature range of the 206102.
shallow bump, Ag32Cu13 starts a premelting phenomenon that 8 H. Portales, L. Saviot, E. Duval, M. Gaudry, E. Cottancin,
involves both the core and the shell at the same time. This M. Pellarin, J. Lermé and M. Broyer, Phys. Rev. B: Condens.
difference between the systems is due to specic features of the Matter Mater. Phys., 2002, 65, 165422.
energy landscapes of these nanoalloys. In fact, the rst excited 9 G. Rossi, A. Rapallo, C. Mottet, A. Fortunelli, F. Baletto and
isomers of Ag32Cu13 involve a shape change of the core, while R. Ferrando, Phys. Rev. Lett., 2004, 93, 105503.
the rst excited isomers of Ag32Co13 and Ag32Ni13 involve a 10 M. Valden, X. Lai and D. W. Goodman, Science, 1998, 281,
shape change of the shell alone. 1647–1650.
(iii) Even below the temperature of the maximum of the heat 11 C. Mottet, G. Rossi, F. Baletto and R. Ferrando, Phys. Rev.
capacity, the clusters do not preserve practically any memory of Lett., 2005, 95, 035501.
their low-temperature congurations. For example, the proba- 12 M. A. Ortigoza and T. S. Rahman, Phys. Rev. B: Condens.
bility of nding Ag32Ni13 in its global minimum conguration is Matter Mater. Phys., 2008, 77, 195404.
practically negligible at 1000 K, well below the temperature of 13 F. Delogu, E. Arca, G. Mulas, G. Manai and I. Shvets, Phys.
the maximum, which is 1100 K. At 1000 K, Ag32Ni13 is uctu- Rev. B: Condens. Matter Mater. Phys., 2008, 87, 024103.
ating among a large number of different congurations 14 D. Bochicchio and R. Ferrando, Nano Lett., 2010, 10, 4211–
belonging to basins of different local minima. In practice, the 4216.
most signicant transformations occur well below the temper- 15 M. Molayem, V. G. Grigoryan and M. Springborg, J. Phys.
ature of the main maximum of the heat capacity curve. Chem. C, 2011, 115, 7179–7192.
(iv) The analysis of the pair correlation function at high 16 H. Yildirim, A. Kara and T. S. Rahman, J. Phys. Chem. C, 2012,
temperature show that complete mixing of the two metals is not 116, 281–291.
even achieved in the liquid state. This agrees with the bulk 17 A. L. Mackay, Acta Crystallogr., 1962, 15, 916–918.
behavior in AgCo and AgNi. In AgCu bulk systems, the two 18 K. H. Kuo, Struct. Chem., 2002, 13, 221–230.
metals are miscible in the liquid phase, but well above 1000 K.47 19 Z. Kuntová, G. Rossi and R. Ferrando, Phys. Rev. B: Condens.
In conclusion, we have shown that the study of the melting Matter Mater. Phys., 2008, 77, 205431.
process of nanoalloys needs the analysis of several different 20 T. Momin and A. Bhowmick, J. Alloys Compd., 2013, 559, 24–
quantities in order to display all the complex phenomena that 33.
take place when temperature increases. 21 F. Baletto, C. Mottet, A. Rapallo, G. Rossi and R. Ferrando,
Surf. Sci., 2004, 566, 192–196.
22 R. S. Berry and B. M. Smirnov, Low Temp. Phys., 2009, 35,
256–264.
Acknowledgements
23 A. Proykova and R. S. Berry, J. Phys. B: At., Mol. Opt. Phys.,
This work was supported by the National Science Foundation of 2006, 39, R167.
China (nos 11175035, 10875023), the National Magnetic 24 W. Smith and T. R. Forester, J. Mol. Graphics, 1996, 14, 136–
Connement Fusion Science Program (no. 2013GB109005), 141.
Chinesisch-Deutsches Forschungsprojekt (GZ768), the Funda- 25 D. J. Evans and O. P. Morriss, Comput. Phys. Rep., 1984, 1,
mental Research Funds for the Central Universities (no. 297–343.
DUT12ZD(G)01). H. Y. Oderji gratefully acknowledge the 26 H. Y. Oderji and H. Ding, Chem. Phys., 2011, 388, 23–30.
support from the Research Council of Tehran University, the 27 F. Cleri and V. Rosato, Phys. Rev. B: Condens. Matter Mater.
Chinese Scholarship Council and Iran Nanotechnology Initia- Phys., 1993, 48, 22–33.
tive Council. Support is acknowledged from the COST Action 28 A. M. Ferrenberg and R. H. Swendsen, Phys. Rev. Lett., 1989,
MP0903 NANOALLOY. 63, 1195–1198.

21992 | RSC Adv., 2013, 3, 21981–21993 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper RSC Advances

29 T. Bereau and R. H. Swendsen, J. Comput. Phys., 2009, 228, 39 F. Lindemann, Phys. Z., 1910, 11, 609.
6119–6129. 40 S. K. Lai, W. D. Lin, K. L. Wu, W. H. Li and K. C. Lee, J. Chem.
30 A. Lyalin, A. Hussien, A. V. Solov'yov and W. Greiner, Phys. Phys., 2004, 121, 1487–1498.
Rev. B: Condens. Matter Mater. Phys., 2009, 79, 165403. 41 D. McQuarrie, Statistical mechanics, Harper and Row, New
31 C. Mottet, G. Tréglia and B. Legrand, Phys. Rev. B: Condens. York, 1976.
Matter Mater. Phys., 1992, 46, 16018–16030. 42 J. Jellinek, T. L. Beck and R. S. Berry, J. Chem. Phys., 1986, 84,
32 P. J. Hsu, J. S. Luo, S. K. Lai, J. F. Wax and J.-L. Bretonnet, 2783–2794.
J. Chem. Phys., 2008, 129, 194302. 43 J. M. Haile, Molecular Dynamics Simulation: Elementary
33 S. Gafner, L. Redel and Y. Gafner, J. Exp. Theor. Phys., 2012, Methods, John Wiley & Sons, Inc., New York, USA, 1st edn,
114, 428–439. 1992.
Published on 13 September 2013. Downloaded by Trent University on 06/04/2017 06:40:43.

34 M. F. Aguado and A. Jarrold, Annu. Rev. Phys. Chem., 2011, 62, 44 M. P. Allen and D. J. Tildesley, Computer Simulation of
151. Liquids, Oxford University Press, USA, 1989.
35 P. K. Janert, Gnuplot in action: understanding data with 45 L. C. Bourne, S. C. Rowland and A. Bienenstock, J. Phys.,
graphs, Manning Publications, Greenwich, CT 06830, 2010. Colloq., 1981, 42, C4-951–C4-954.
36 F. H. Stillinger and T. A. Weber, Phys. Rev. A, 1982, 25, 978– 46 M. M. Mariscal, N. A. Oldani, S. A. Dassie and E. P. M. Leiva,
989. Faraday Discuss., 2008, 138, 89–104.
37 T. L. Beck and R. S. Berry, J. Chem. Phys., 1988, 88, 3910– 47 R. Hultgren, P. D. Desai, D. T. Hawkins, M. Gleiser and
3922. K. K. Kelley, Values of the Thermodynamic Properties of
38 I. Fisher, Statistical Theory of Liquids, Univ. of Chicago Press, Binary Alloys, American Society for Metals, Berkeley,
Chicago, 1966. 1981.

This journal is ª The Royal Society of Chemistry 2013 RSC Adv., 2013, 3, 21981–21993 | 21993

You might also like