Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

J Nanopart Res (2011) 13:803815

DOI 10.1007/s11051-010-0082-4

RESEARCH PAPER

Molecular dynamics simulation of iron nanoparticle


sintering during flame synthesis
Ngoc Ha Nguyen Richter Henning John Z. Wen

Received: 15 July 2010 / Accepted: 1 September 2010 / Published online: 17 September 2010
Springer Science+Business Media B.V. 2010

Abstract The sintering process of iron nanoparticles produced in a flame environment is investigated
by molecular dynamic (MD) simulations. The thermodynamic characteristics and restructuring pathways
are studied for two-body and three-body sintering
processes. The melting point, energy, and structures
are computed for nanoparticles before and after
sintering. A simplified model is proposed to predict
the equilibrium temperature of nanoparticles upon
sintering. The MD results are used to explain the
formation mechanisms of two size ranges of nanoparticles during the flame synthesis. The role of
sintering during nanoparticle growth is analyzed.
Keywords Molecular dynamics  Sintering 
Coalescence  Flame synthesis  Nanoscale modeling

N. H. Nguyen  J. Z. Wen (&)


Department of Mechanical and Mechatronics
Engineering, University of Waterloo,
Waterloo, ON N2L 3G1, Canada
e-mail: jzwen@uwaterloo.ca
N. H. Nguyen
Department of Chemistry, Center for Computational
Science, Hanoi National University of Education,
Hanoi, Vietnam
R. Henning
Nano-C, Inc, 33 Southwest Park, Westwood,
MA 02090, USA

Introduction
In addition to a variety of production methods for
single-walled carbon nanotubes (SWNT) such as,
laser vaporization, arc discharge, chemical vapor
deposition, and gas-phase synthesis, flame synthesis
has attracted increasing interests in recent years
(Merchan-Merchan et al. 2010). It can be categorized
into the gas-phase approaches and, particularly using
floating catalysts supplied continuously with the fresh
gas mixture, has shown significant potential to
produce high-quality SWNT at low cost and large
scale (Richter et al. 2008). While carbon-containing
molecules, which are suitable as supply for SWNT
growth (e.g., gas-phase light-molecular-weight hydrocarbons and carbon monoxide), are generated by
means of the exothermic combustion process, catalyst
necessary for SWNT nucleation is formed simultaneously in situ by decomposition of solid or liquid
precursors such as iron pentacarbonyl (Fe(CO)5),
discussed in more detail. The thermal-decomposition
of Fe(CO)5 occurs when the temperature in the burner
reaches about 700 K (Giesen et al. 2003) and
produces a vapor of iron atoms. Subsequently, when
a supersaturation of iron vapor is reached, the
nucleation of the smallest iron nanoparticles takes
place. This nucleation is followed by the nanoparticle
coagulation (or agglomeration) and surface growth
process resulting from the condensation of gaseous
species. When the evolution of nanoparticle structures occurs under flame conditions with variable

123

804

temperatures over a large range, nanoparticles can


experience the sintering (or coalescence) and solidification processes, which will significantly determine
their final size, morphology, and crystalline structures.
Because in flame synthesis the produced metallic
nanoparticles will act as the catalyst for growing
SWNT, the structural properties of these nanoparticles
greatly affect the size, geometry, and properties of
SWNT (Nasibulin et al. 2005). In order to investigate
the formation mechanism and further optimize the
growth process of SWNT, it is essential to describe
accurately the formation and growth processes of
catalyst nanoparticles in the flame.
The studies on the gas-phase synthesis of metallic
nanoparticles (with or without accompanying SWNT
growth) have drawn increasing attentions during the
past decade. Experimental investigations have been
conducted in a variety of chemical reactors such as
aerosol flow reactors (Kock et al. 2005; Anisimov et al.
2010; Moisala et al. 2005), shock tubes (Drakon et al.
2008; Giesen et al. 2004a), diffusion flames (Guo and
Kennedy 2007; Pikhitsa and Choi 2007; Zhao et al.
2009), and premixed flames (Vander Wal and Hall
2001; Hecht et al. 2009; Wen et al. 2008a). When the
size distribution and morphology of nanoparticles
were measured using multiple, often complementary,
analytic techniques such as mass spectrometry (MS),
differential mobility analysis (DMA), transmission
electron microscopy (TEM), and laser-induced incandescence (LII), some interesting and distinguishable
phenomena have been reported for iron nanoparticles.
For instance, it was found that, when the reaction
temperature keeps nearly constant and uniform in the
reactor (e.g., in a flow reactor and a shock tube), a
particle size distribution with one maxima is produced
and preserved. When a large range of temperatures
presents in the burner (e.g., for the diffusion and
premixed flames), however, two size ranges of nanoparticles were reported. Guo and Kennedy (2007)
studied the synthesis of iron-oxide nanoparticles in a
H2/air diffusion flame. The structural properties of
nanoparticles were assessed using TEM, X-ray diffraction (XRD), and inductively coupled plasma
mass spectrometry (ICP-MS). Brunauer-Emmett-Teller
(BET) surface areas were determined. They found two
size modes of produced iron-oxide nanoparticles. The
large size mode contained crystalline and nonagglomerated nanoparticles with a median diameter
of about 45 nm, while the small size mode contained

123

J Nanopart Res (2011) 13:803815

nanoparticles which were agglomerated and in the size


range between 3 and 8 nm. The study on the history of
nanoparticle evolution was conducted using the thermophoretic sampling followed by TEM. It was found
that at the early stage of the flame where the
temperature is relatively low (from 1100 to 2000 K),
nanoparticles with a mean diameter of about 20 nm
were formed. With increasing the height above burner
(HAB), the small size mode appeared when the
temperature reached high (about 2100 K). When the
HAB was further increased, the co-existence of large
and small modes remained and the temperature
dropped from 2100 to 1100 K. The authors explained
the formation mechanism of the small-mode nanoparticles as the vaporization of the large and liquid
nanoparticles at the maximum temperature, followed
by the second nucleation from the vapor phase when
the temperature decreased.
In comparison, a subset of the authors of this study
was involved in the investigation of iron nanoparticle
and SWNT formation in a premixed CH4/O2/Ar
flame (Wen et al. 2008a, b). As discussed previously
(Height et al. 2004), iron nanoparticles were produced in situ as the catalyst to grow SWNT in the
same flame. Probe sampling has been conducted at
different heights above the burner (i.e., residence
times) and the collected condensed materials were
characterized using Raman spectroscopy, XRD, SEM
(Scanning electron microscopy), STEM (Scanning
transmission electron microscopy), and TEM. Two
classes of iron/iron-oxide nanoparticles with distinguishable sizes were identified. It was found that,
iron-oxide nanoparticles had an average size of about
23 nm (i.e., the small-mode nanoparticles), while
elemental iron nanoparticles were coated with graphite layers and had an average size of about 20 nm
(i.e., the large-mode nanoparticles). A, yet unpublished, TEM image of such structures is shown in
Fig. 1. The co-existence of two size ranges was found
for all samples taken at heights beyond the maximum
flame temperature of about 1750 K and persisted in
the postflame zone at a temperature of 700 K. Based
on a thermodynamic analysis, i.e., equilibrium calculations using carefully selected thermodynamic
data of possibly pertinent species, our study suggested that the formation of both elemental iron and
iron-oxide nanoparticles were favorable for the
corresponding flame temperature (Wen et al.
2008a). Also, we proposed that the coating of

J Nanopart Res (2011) 13:803815

Fig. 1 The TEM image of flame synthesized nanoparticles

graphite layers hindered the oxidation of iron into


iron oxide and hence, protected the iron phase
originally formed in large-mode nanoparticles. The
comparison of Guos study and our previous study
suggests that, without considering the oxidation and
carbon coating on iron nanoparticles (resulting from
the differences in fuel compositions and flame
structures), two classes of nanoparticles with distinguishable sizes can be produced in flames where a
large temperature gradient exists. In addition, both
studies found that the small-mode nanoparticles were
agglomerated, while the large-mode nanoparticles
were less agglomerated. Please note that the degree of
agglomeration of iron nanoparticles is controlled
mainly by sintering processes at different temperatures and caused by the phase change of crystalline
structures. In order to reveal the formation mechanisms of two size ranges of nanoparticles measured in
flames, it is essential to investigate specifically to the
sintering process of nanoparticles with different sizes
and fractal dimensions.
The sintering processes of silver and copper nanoscale agglomerates produced by an evaporation
condensation method were experimentally investigated with evolving fractal dimensions (Weber and
Friedlander 1997). Three typical states of agglomerates were proposed, i.e., the dispersed state at which
the nanoparticles are fully dispersed within the carrier
gas; the intermediate state at which the aggregates
consist of a number of primary particles and have
distinguishable fractal dimensions (between 1 and 3),
and the condensed state at which the agglomerates

805

have close-packed structures. The sintering process


was described as an activated process which is driven
by the excess Gibbs free energy in the initial state
compared with the condensed state. Evolution of the
fractal dimension of agglomerates was investigated
through measuring the ratio of nanoparticles mobility
equivalent diameters over diameters of the corresponding condensed state after sintering. The coordination number of the agglomerates (number of nearest
neighboring primary particles) was then derived from
the fractal dimension. The changing rate of the
coordination number was finally used to determine
the activation energy of sintering processes. It was
found that, for silver and copper nanoparticles with
small diameters (less than 18 nm) the activation
energy, which is much smaller than the bonding
energy between primary particles, increases with
increasing particle diameter. The particles larger than
18 nm showed the decreasing activation energy with
increasing particle diameter. Please note that, in the
above study the sintering process was studied without
considering the effect of melting of agglomerates.
In addition to experimental studies, to theoretically
describe the nanoparticle formation in a reactive flow
environment, a set of particle population equations
need to be solved by accounting for the contributions
from individual processes such as nucleation, coagulation, surface growth, and sintering. When a variety
of computational approaches have been implemented
for solving the population equations, e.g., using the
sectional models (Pope and Howard 1997; Yu et al.
1998), the method of moments (McGraw 1997;
Kazakov and Frenklach 1998), and the stochastic
techniques (Debry et al. 2003; Zhao et al. 2003), the
theoretical and/or semi-empirical expressions of
individual rates corresponding to aforementioned
structurally evolving processes need to be determined
in a much accurate way. During the gas-phase
synthesis the nuclei are often formed from the vapor
of corresponding metal, the nucleation rate can be
therefore calculated using the classical nucleation
theory (Oxtoby 1992; Ford 1997). Recently the
nucleation process of iron nanoparticles from the
supersaturated iron vapor was investigated by molecular dynamics (MD) simulations (Lummen and
Kraska 2005a, 2006). The nucleation rates were
calculated using a method proposed in the third study
(Yasuoka and Matsumoto 1998) where a significantly
large nucleation rate rather than the one predicted by

123

806

the classical nucleation theory was suggested. The


computed MD results were consistent with experimental data of iron nucleation rates measured in a
shock tube. After the nuclei are formed through the
condensation of the metal vapor, they will grow
bigger and become fractal aggregates due to intensive
collisions. Experiencing the large temperature gradients existing in the flame, these fractal aggregates
will endure phase change processes such as solidification and sintering and form less agglomerated large
nanoparticles. The investigation of sintering is,
therefore, important to understand the formation
mechanisms of the two size ranges of nanoparticles
reported in experimental studies (Guo and Kennedy
2007; Wen et al. 2008a).
The sintering of iron nanoparticles has been
studied theoretically by means of different computational methods as well. Moniruzzaman et al. (2007)
investigated the growth of iron nanoparticles generated from the thermal-decomposition of Fe(CO)5 in
an aerosol reactor. The particle population equations
were solved using a discrete sectional model, when
the sintering process was modeled in the continuum
regime. The sintering time scale was calculated by
considering the size-dependency of both surface and
bulk diffusion coefficients (Knight et al. 2000).
Another study used the similar continuum description
for the sintering process and used the method of
moments to solve the particle population equations
(Giesen et al. 2004b). Besides the continuum models,
McCarthy and Brown (2009) used kinetic Monte
Carlo (MC) simulations to study the coalescence of
fcc nanoparticles (ranging in sizes from 28,000 to
68,000 atoms) and calculated the lattice-based diffusion of surface atoms. Most recently, Ding et al.
(2009) compared the predictions of the sintering
process between nanoparticles obtained from a molecular dynamics (MD) model and a continuum model,
respectively. They concluded that the continuum
model was unable to capture the sintering behavior of
nanoparticles, which behave in many different scenarios of the continuum theory. For example, they
observed the reorientation of the crystalline structures
of nanoparticles at the beginning of the sintering and
the formation of different types of necks between
nanoparticles, which could lead to different mechanisms of matter redistribution and are difficult to be
described by the continuum model. Allowing for a
more detailed approach assessing individual atoms

123

J Nanopart Res (2011) 13:803815

and molecules, density functional theory (DFT) was


utilized for predicting the coalescence of iron clusters
and nanoparticles. Involving a part of the authors of
this study, a detailed chemical kinetic model for gasphase synthesis of iron nanoparticles and validated
model against shock-tube studies has been developed (Wen et al. 2007). While electronic structures
and thermodynamic property data for Fen clusters
(2 B n B 8) were computed using DFT at the
B3PW91/6-311 ? G(d) level of theory, the reactions
(including the sintering) between larger iron nanoparticles needed to be solved by a sectional description
to limit the number of species and reactions to be
handled. In order to improve the model accuracy and
to address the time-dependent evolution of the
sintering process, the use of molecular dynamics
(MD) simulations is desired.
This study focuses on sintering processes of
agglomerates accompanying with melting of their
primary nanoparticles. The major objective of this
study is to investigate the structural transformation and
evolution of thermodynamic properties during sintering of iron nanoparticles at varying sizes and temperatures pertinent to the flame synthesis of SWNT. The
melting processes are computed using the classic
molecule dynamics simulation. The localized equilibrium temperatures after sintering, which result in
excess heat and could affect further restructuring of
agglomerates, will be used for exploring the kinetic
pathways of the sintering process involving two or
more nanoparticles. The role of nanoparticle sintering
in the formation of different size ranges of nanoparticles during the flame synthesis will be discussed.

Molecular dynamics simulation


A classic molecular dynamics approach was used in
this study. Since the principle of the classic MD
method is well known, only the critical parameters
are presented here.
Force field selection
In this study, the FinnisSinclair (FS) potential
(Finnis and Sinclair 1984) for a body-centered cubic
(bcc) iron (Marchese et al. 1988) was used to describe
the inter-atomic potential between iron atoms. It was
found that among MD and MC simulations, the FS

J Nanopart Res (2011) 13:803815

807

potential provides a satisfying description of the


properties of transition metal clusters (Elliott et al.
2009). This is partially due to its analytical nature,
which differs from the potential derived from
numerical embedded atom models (EAM), and the
accuracy in computing high-order derivatives which
is important for calculating the shear elastic modulus.
Because the phase transformation between bcc and
fcc (face-centered cubic) phases was not the focus of
this study and the melting point is not expected to
depend significantly on the crystal structure (Engin
et al. 2008), the FS potential for the bcc iron was
chosen to study the melting point of iron nanoparticles (Shibuta and Suzuki 2008; Shibuta et al. 2009).
The FS potential U, which includes the pair-wise
and many-body parts, is expressed as:
U U1 U2

N
N
X
1X
V1 ri;j
V2 ni
2 i;j:i6j
i1

where the pair-wise part V1(ri,j) is represented by the


form:

 
r  c2 c0 c1 r c2 r 2 r  c
V1 ri;j
2
r [ c
0
where r is the distance; c0, c1, and c2 are free
parameters; and c is the cutoff distance. The manybody part V2(ni) is given by:
p
3
V2 ni f ni A ni
where ni is the local electronic charge density around
the atom i and A is the fitting parameters.

Bulk melting point


To determine the bulk melting point, MD simulations
were conducted on a cell with periodic boundary
conditions and under isothermal and isobaric conditions (i.e., the number of nanoparticles, pressure, and
temperature of the system were conserved). The
external hydrostatic pressure was set to 0.0 Pa on the
cell. To reduce the error caused by the eggbox effect, a
3 9 3 9 3 supercell size was used in this study as the
initial structure. This treatment, utilizing a larger
computational domain, is better than the calculation of
bulk melting points using the primitive unit cell. In
order to find the melting point, the system temperature
was increased from 1700 to 1900 K with an interval of
5 K. The simulation time for each temperature was set
to 500 ps. The typical structures from these MD
simulations are shown in Fig. 2. It shows clearly that
when the temperature was increased from 1825 to
1830 K, a dramatic change in the crystalline structure
occurs. The qualitative data on the cell structure
computed at different temperatures are shown in
Table 2. Both Fig. 2 and Table 2 confirm the computed melting point of the bulk iron is 1830 K, which
is in a very good agreement with the measured data of
1811 K (Lide 2010). Please note that the discrepancy
in predicted melting points between this study and a
previous study (Duan et al. 2007) results from the
different potential models used in two studies. And for
nanoparticles, the melting temperature varies with the
particle size as well.
Simulation method for nanoparticle sintering

Bulk modulus
The aforementioned FS potential was used to compute the bulk properties of iron. The calculated bcc
cell parameters, bulk modulus, and shear modulus are
shown in Table 1. These data show very good
agreements with experimental data.
Table 1 Properties of bulk iron calculated using FS potential
Cell
parameters
)
ao (A

Bulk
modulus (GPa)

Shear
modulus
(GPa)

Present work

2.869

172

79

Experimentala

2.867

170

82

From reference (Lide 2010)

In order to study their sintering, iron nanoparticles


were constructed being a sphere of 2 nm in diameter
and with a bcc lattice. Each of these nanoparticles
consists of 342 iron atoms and is initially placed in
vacuum. The nanoparticles were then heated to the
desired temperature T1. For each temperature, the
system was equilibrated and monitored until there
was no change in its average temperature in the
canonical ensemble (i.e., the NVT property). It was
found that, depending on the simulation cases, the
time to reach the system equilibrium at 0 K, which
was used to specify the initial structure, was about
5 ps when the time step of 1 fs was used. When the
FS potential was used, the bcc structures were
observed after the initial geometries at 0 K were

123

808

J Nanopart Res (2011) 13:803815

Fig. 2 Typical structures


computed on a 3 9 3 9 3
cell using periodical
boundary conditions

Table 2 Geometries of the iron 3 9 3 9 3 cell computed for different temperatures


Cell parameter

At 298 K

8.60

At 1825 K after 500 ps

8.78

At 1830 K after 200 ps

7.56

8.60

8.60

90.0

90.0

90.0

8.71

8.36

91.8

89.9

90.0

10.33

8.76

98.8

77.4

94.4

and the cell angles (a, b and c) are in degrees


The cell vectors (a, b and c) are in A

optimized. For the FS potential, the cut-off distance


for the many-body part V2(ni) and
was set to 3.57 A

3.4 A for the pair-wise part V1(ri,j), respectively


(Marchese et al. 1988). The Verlets leapfrog algorithm for integrating Newtons equations of motion
was used for all MD simulations. When the sintering
processes were simulated in the microcanonical
ensemble (i.e., with the NVE property), no periodic
boundary conditions was applied, and the isolated
nanoparticle system was used.

Results and discussion


As mentioned early the current study focused on the
formation of iron nanoparticles in the small model
range (with the average diameter of 3 nm in the
premixed flame). The sintering processes of 2-nm
nanoparticles through 2-body and 3-body collisions
were investigated and expected to produce the
nanoparticle size of about 3 nm.
Calculation of the melting point of 2-nm iron
nanoparticles
The sintering process occurs in a fast manner,
when the temperature is near the melting temperature

123

of nanoparticles. On the other hand, the determination of the melting point of corresponding
nanoparticles helps to analyze the causes of these
coalescence events through either the phase (liquid
and/or solid)-related process or the intracluster
interactions (as discussed later). It has been well
accepted that nanoparticles have lower melting
temperatures than their bulk state, and the melting
point decreases with the declining nanoparticle size.
In order to identify the melting point of 2-nm
nanoparticles, the calculations were conducted
using the NVT MD method, and done for a
temperature range between 400 and 1811 K (the
melting point of the bulk iron). The simulation time
was set to 500 ps for all simulation cases. Figure 3
shows the computed structure of nanoparticles,
when they reached the equilibrium states at different temperatures.
Figure 3 clearly shows that a significant change in
the shape and structure of iron nanoparticles occurs
above 1500 K, and the phenomenon of melting
appears remarkably at 1600 K. This indicates the
melting point of 2-nm iron nanoparticles is about
1550 K. When it is widely accepted that the sintering
process occurs close to the melting point, for the
following simulation on sintering, the temperature
range between 1400 and 1600 K was studied.

J Nanopart Res (2011) 13:803815

809

Fig. 3 Initial and equilibrium structures of iron nanoparticles at different temperatures

Simulation of iron nanoparticle sintering


Theoretically, an iron nanoparticle can be considered
as a larger iron cluster (Wen et al. 2007). According
to the kinetic theory in gas phase, the simultaneous
collision of four molecules (i.e., fourth order reaction) and higher ones are hardly possible. For this
reason, the sintering processes between two and
among three iron nanoparticles were investigated in
this study. In addition, at equilibrium states near the
melting temperature, the iron nanoparticles are nearly
spherical and hence the effect of the directional
structure of nanoparticles was not taken into account.
The sintering of iron nanoparticles was treated as
chemical reactions with the original nanoparticles
being the reactant and the sintered nanoparticle as the
product. When the original nanoparticles combine
together to form a larger nanoparticle through
sintering, the process is exothermic due to the release
of surface energy from the system (the surface area of
the produced nanoparticle is smaller than the combination of surface areas of original nanoparticles).
This energy release will heat up the system and the
temperature of nanoparticles will increase (by assuming an adiabatic condition due to the fast-occurring
sintering process). The following approach was used
to calculate the final temperature of the nanoparticle
using the aforementioned principle.
Given a bulk energy of the system as Ebulk and the
total energy for the same system with a surface as
Esurface, the surface energy ESE is then defined as
ESE

Esurface  Ebulk
A

where A is the surface area. Suppose that during the


sintering process, the energy release is DE (J/mol)
and the heat capacity of iron, CP(T), can be calculated
in the range of computed temperatures with an initial
temperature of T1, the energy balance gives the final

temperature of the nanoparticle T2 under the adiabatic


condition
Z T2
DE
CP TdT
5
T1

where CP(T) is the heat capacity function and its


parameters are taken from the literature (Lide 2010).
CP T 776:74 0:92T  3:84  104 T 2 5:71
 108 T 3 2:42  108 T 2 J=mol  K
6
Combining Eqs. 5 and 6 gives


DE 776:74T2  T1 0:92  0:5  T22  T12


 3:84  104  1=3  T23  T13


5:71  0:25  108  T24  T14


 2:42  108 T21  T11
7
By using the Eq. 7, the final temperature T2 can be
determined if the energy release DE can be obtained
from MD simulations. Please note that Eq. 7 provides a
simplified model for the temperature calculation and is
based on the assumption of fast sintering close to the
melting temperature. This is the case when the nanoparticle formation is studied during the flame synthesis
where a steep temperature profile presents prior to
reaching the maximum flame temperature. Equation 7
allows a quick calculation of the temperature T2 from a
given T1 and known DE (as shown later). MD simulations can be elaborated to calculate the temperature T2
directly. The process, however, is very time-consuming
and needs a significant computing resource.

Sintering between two iron nanoparticles


In order to simulate the coalescence process, two
identical nanoparticles at their equilibrium states

123

810

J Nanopart Res (2011) 13:803815

were placed into the computational domain. The

distance between two nanoparticles was set to 3 A


which is within the effective range of intermolecular
forces. As proposed in the earlier study (Weber and
Friedlander 1997), the interaction between the
primary particles occurs over a shorter distance than
. The distance between neighthe bulk value of 4 A
boring nanoparticles was chosen in this study to
provide the sufficient intermolecular forces (corre in the FS
sponding to the cut-off distance of 3.57 A
potential) which lead to effective sintering. Because
the activation energy of sintering processes is associated with, rather than proportional to, the bonding
strength of neighboring nanoparticles, this chosen
distance would not affect the study of sintering
processes as long as the melting points are predicted
reasonably. The simulation tasks were performed up
to 250 ps, when a good statistics of data was shown.
The snapshots of two typical simulation cases for
1400 and 1500 K are shown in Fig. 4.

Figure 4 shows the strong dependence of the


sintering rate on the initial temperature T1. When
the sintering rate at 1400 K was very slow and the
aggregate with a neck connection was formed, the
sintering process at 1500 K was completed at 250 ps
and caused the formation of a new spherical nanoparticle. The newly produced nanoparticle due to
sintering has a diameter of 2.72 nm in comparison
with the original size of 2 nm. The surface energy
ESE and the total energy Esurface computed for the
original nanoparticle and the new nanoparticle are
shown in Table 3.
As shown in Table 3, the amount of energy release
can be calculated per mole of Fe atoms as
Fe342 Fe342 ! Fe684 DE 1:5963 kJ/mol

And the final temperature T2 can be calculated


using Eq. 7 by adding the energy release and the
initial temperature T1. The values of T2 calculated by
this method are shown in Table 4 for various

Fig. 4 Snapshots of typical MD computed structures during sintering for two initial temperatures: 1400 K (a) and 1500 K (b)
Table 3 The surface energy, total energy, and energy release DE computed for equilibrium structures through 2-body and 3-body
sintering processes
2 nm (primary particle)

2.72 nm (2-body sintering)

3.15 nm (a ? c)

3.15 nm (b ? c)

2)
Surface energy (eV/A

0.1845

0.1946

0.1985

0.1985

Total energy (eV)

-1232.0041

-2475.3248

-3858.6044

-3858.6044

-1.5963

-5.7923

-10.4017

DE (kJ/mol)
The energy release was calculated for per mole Fe atoms

123

J Nanopart Res (2011) 13:803815

811

Table 4 The final temperature T2 (K) and the initial temperature T1 (K) obtained from Eq. 7 and the MD simulation
(values in parentheses) after the system reached its equilibrium
state
Fe342 Fe342 ! Fe684 , DE = -1.5963 kJ/mol
T1

1400

1500

1600

T2

1441.3 (1457.5)

1540.4633 (1524.2)

1639.3403

temperatures. For the comparison purpose, the final


temperatures T2 calculated directly using MD are
shown in parentheses.
The results from Table 4 show that, if the sintering
process starts at 1500 K, the temperature of the
nanoparticle will eventually increase to about
1550 K. The locally generated excess heat resulting
from this temperature change (about 40 to 50 K) over
a very short period (about a couple hundreds ps)
could promote the further growth of newly formed
nanoparticles from the secondary processes such as
surface condensation of gaseous atoms and secondary
coagulation with sintering. Under flame synthesis
conditions, the energy release from sintering will not
significantly affect the flame temperature because the
concentration of iron nanoparticles is usually small
and the rate of sintering is quite fast. In fact, if the
adiabatic condition is not valid during the sintering
process and there is heat transfer between the
nanoparticle and the carrier gas, as studied in
the literature (Lummen and Kraska 2005b, 2006),
the temperature change of the nanoparticle will be

affected by the heat transfer rate which is dependent


on the thermodynamic properties of the carrier gas.
This interesting phenomenon has been explored by
other researchers (Eremin et al. 2008) during a laser
photolysis for producing iron nanoparticles. Table 4
shows the consistent values calculated using Eq. 7
and the MD simulation. This validates the Eq. 7 as a
simplified approach to determine the final nanoparticle temperature following sintering.
Sintering among three iron nanoparticles
In addition to sintering between two nanoparticles,
the collision among three nanoparticles is possible
especially when the concentration of iron nanoparticles is extremely high. The three-body sintering
process, occurring with less probability than that
between two particles, can be written as the chemical
reaction shown below and with a smaller probability
than two-body sintering.
Fe342 Fe342 Fe342 ! Fe1026

The three-body sintering process can also be


described as reaction of the third nanoparticle with
a sintering structure resulting from an ongoing
reaction between two other nanoparticles.
As shown in Fig. 5, there are basically two
pathways for Reaction 9. By means of the first
pathway, three initial iron nanoparticles located on a
line (a) experience sintering and form a new iron
nanoparticle of 3.15 nm (c). Alternatively, three initial

Fig. 5 Two possible


pathways for three-body
sintering processes

123

812

J Nanopart Res (2011) 13:803815

iron nanoparticles located as a triangle (b) experience


sintering and form a new iron nanoparticle of
3.15 nm (c).
The sintering pathways were investigated by
placing the initial nanoparticles at their designated
locations and setting the distance among nanoparti . The simulation tasks were performed up
cles to 3 A
to 250 ps, when a good statistics of data was shown.
The computed surface energy and the energy release
are shown in Table 3. Please note that in this study,
the energy release was calculated using the potential
energy U which depends on the initial and final
structures of nanoparticles. And for the different
pathways, the energy releases to produce the same
equilibrium structure are different due to different
initial structures.
The results from Table 3 clearly show that the
pathway (b) ? (c) possesses the greater chemical
potential than the pathway (a) ? (c), and hence,
results in a larger reaction rate. This agrees with the
previous experimental findings on size-dependent
activation energies (Weber and Friedlander 1997).
Since the linear geometry (a) has the larger mobility
diameter than the triangle geometry (b), the activation energy for sintering of (a) is greater than the
value for (b). In other words, the sintering process
passing through (b) ? (c) has a larger rate.
It is interesting to compare the sintering processes
between two nanoparticles and among three nanoparticles. Because the difference in mobility diameters of
two-nanoparticle agglomerate and the close-packed
three-body geometry (c) is insignificant, the MD
computed equilibrium temperatures were used to
conduct this investigation. Using the Eq. 7 and the
computed energy release, the final temperatures T2
were calculated for different initial temperatures T1
and are shown in Table 5 [for the pathway (b) ? (c)
only].
Table 5 shows that, in comparison with the twobody sintering process with a small temperature rise
(about 50 K) during sintering, the three-body

sintering by (b) ? (c) causes a dramatic increase


(about 250 K) in the temperature of the final equilibrium state. This significant temperature increase,
resulting from the large energy release from sintering
among three nanoparticles, led to melting of the
newly generated structure at the much lower initial
temperature (1300 K compared to 1500 K for twobody sintering). This proposes a supplemental finding
to the previously published relationship between the
particle size and the sintering activation energy
(Weber and Friedlander 1997). For smaller nanoparticles (less than 18 nm), the sintering activation
energy (corresponding to the fractal initial geometries
studied by Weber and Friedlander (1997)) increases
with increasing agglomerate diameter. This was
suggested in the previous study and also shown in
this study on the comparison of sintering between two
nanoparticles and among three linear nanoparticles.
When the agglomerates are close-packed (such as the
geometry b) and the sizes of agglomerates are
comparable, the activation energy will be affected
more critically by the bonding strength of neighboring particles. The agglomerate with a stronger
bonding energy (sum over all bonds between neighboring primary particles) requires less excess energy
to initiate its sintering towards a more compact
structure at the condensed state. This agrees with a
recent modeling study which showed the nanoparticles with larger fractal dimensions (hence closedpacked structures) possess faster sintering rates
(Hawa and Zachariah 2007). A further validation of
this statement by experimental measurements, however, would be quite useful. In order to check the
accuracy of the Eq. 7, a MD simulation was
conducted for the (b) ? (c) pathway and performed
at the initial temperature of 1200 K. The trajectory of
the system temperature is shown in Fig. 6. The curve
eventually reaches the temperature of 1420 K which
agrees well with the calculated data in Table 5. A
similar result was found for the two-body reaction.
Implication to the flame synthesis of iron
nanoparticles

Table 5 The final temperature T2 (K) and the initial temperature T1 (K) obtained from the Eq. 7 for (b) ? (c) process
Fe342 Fe342 Fe342 ! Fe1026 , DE = -10.4017 kJ/mol_Fe
T1

1200

1300

1400

1500

T2

1467.2

1567.4

1662.3

1755.6

123

The above studies on the sintering process between


two and among three iron nanoparticles show the
following characteristics for nanoparticle formation
during the flame synthesis. First, following the
thermal-decomposition of Fe(CO)5 and iron vapor

J Nanopart Res (2011) 13:803815

813

Finally, at the post-flame zone where the temperature


is gradually cooled down, the secondary nucleation of
iron nanoparticle from the vapor phase occurs.
Combined with a relatively less intensive coagulation
process and insignificant sintering among these
new nanoparticles, the formation of the small-mode
nanoparticles is desired. The above analysis explains
the different formation mechanisms of two different
size ranges of iron nanoparticles during the flame
synthesis and is supported by the MD investigation
on sintering in this study.
Fig. 6 The trajectory of temperature calculations during the
MD simulation for T1 = 1200 K

formation at about 700 K, the nucleation of the


smallest iron nanoparticles occurs, based on (Wen
et al. 2007). Secondly, the large population of newly
nucleated iron nanoparticles enters into the hightemperature zone of the flame where the significant
coagulation processes among small nanoparticles
occur. Meanwhile, these aggregated nanoparticles
experience a fast heating-up process over a large
temperature range [from 700 to 1700 K as shown in
Wen et al. 2008a, b]. The sintering will then
dominate in the nanoparticle growth mechanism,
based on this MD study. Due to the high concentration of iron nanoparticles in this high-temperature
zone, both two-body and three-body sintering can
occur. In addition, the large energy release from these
sintering processes over a very short reaction period
will contribute to the increase of nanoparticles
temperature, which in return enhances the sintering
process and promotes the rapid growth of iron
nanoparticles. The similar phenomenon regarding to
the temperature difference between the grown nanoparticles and the carrier gas has been observed both
experimentally and theoretically in previous studies
(Lummen and Kraska 2006; Eremin et al. 2008). The
fast sintering processes in this high-temperature flame
zone lead to the formation of the large-mode
nanoparticles which survive over the entire flame.
Thirdly, due to the high temperature, the iron vapor
coexists with the condensed phase near the maximum
flame temperature, which is supported by the thermodynamic study (Wen et al. 2008a) and could
become more significant when the energy release
from the nanoparticle sintering is taken into account.

Conclusions
The sintering of iron nanoparticles with a designated
size (2 nm) was studied using the classic MD
method. The FinnisSinclair (FS) potential was used
for predicting the melting point and the structural
transformation of nanoparticles for various temperatures. The detailed MD simulations suggested that,
while the two-body sintering can occur at a temperature which is slightly below the melting point, the
three-body sintering, especially when three nanoparticles form a near close-packed shape, can occur at a
much lower temperature. Both sintering processes
will generate significant amounts of heat which affect
the initiation temperature of sintering process. For
smaller nanoparticles (less than 18 nm), the sintering
activation energy (corresponding to the fractal initial
geometries) increases with increasing agglomerate
diameter. When the agglomerates are close-packed
and the sizes of agglomerates are comparable, the
activation energy will be affected, however, more
critically by the bonding energy of agglomerates. The
agglomerate with a larger bonding strength requires
less excess energy (hence, the lower temperature) to
initiate its sintering. A simplified model was proposed for calculating the final equilibrium temperature of nanoparticles after experiencing the sintering
process. The model prediction agrees with the MD
result with significant saving in computational costs.
The MD simulations were used to explain the flame
synthesis of iron nanoparticles. The formation of
large-mode and small-mode nanoparticles was
explained by two distinctive nanoparticle formation
mechanisms. Before the flame reaches its maximum
temperature, the nucleation of nanoparticles followed
with an intensive sintering process leads to the

123

814

formation of large-mode nanoparticles. When the


flame temperature declines in the post-flame zone, the
secondary nucleation occurs with slower sintering
rates and results in the formation of small-model
nanoparticles. The aforementioned findings on the
characteristics of nanoparticle formation will help
designing conditions to produce the right-sized
nanoparticles allowing for the efficient growth of
SWNT.
Acknowledgments This study is supported by the Natural
Sciences and Engineering Research Council of Canada.

References
Anisimov AS, Nasibulin AG, Jiang H, Launois P, Cambedouzou J, Shandakov SD, Kauppinen EI (2010) Mechanistic investigations of single-walled carbon nanotube
synthesis by ferrocene vapor decomposition in carbon
monoxide. Carbon 48:380388
Debry E, Sportisse B, Jourdain B (2003) A stochastic approach
for the numerical simulation of the general dynamics
equation for aerosols. J Comput Phys 184:649669
Ding LF, Davidchack RL, Pan JZ (2009) A molecular
dynamics study of sintering between nanoparticles.
Comput Mater Sci 45:247256
Drakon AV, Emelianov AV, Eremin AV (2008) Nonequilibrium radiation and ionization during formation of iron
clusters in shock waves. J Phys D Appl Phys 41:135201
Duan HM, Ding F, Rosen A, Harutyunyan AR, Curtarolo S,
Bolton K (2007) Size dependent melting mechanisms of
iron nanoclusters. Chem Phys 333:5762
Elliott JA, Shibuta Y, Wales DJ (2009) Global minima of
transition metal clusters described by finnis-sinclair
potentials: a comparison with semi-empirical molecular
orbital theory. Philos Mag 89:33113332
Engin C, Sandoval L, Urbassek HM (2008) Characterization of
fe potentials with respect to the stability of the bcc and fcc
phase. Model Simul Mater Sci Eng 16:035005
Eremin A, Gurentsov E, Schulz C (2008) Influence of the bath
gas on the condensation of supersaturated iron atom
vapour at room temperature. J Phys D Appl Phys
41:055203
Finnis MW, Sinclair JE (1984) A simple empirical n-body
potential for transition-metals. Philos Mag A 50:4555
Ford IJ (1997) Nucleation theorems, the statistical mechanics
of molecular clusters, and a revision of classical nucleation theory. Phys Rev E 56:56155629
Giesen A, Herzler J, Roth P (2003) Kinetics of the fe-atom
condensation based on fe-concentration measurements.
J Phys Chem A 107:52025207
Giesen A, Kowalik A, Roth P (2004a) Iron-atom condensation
interpreted by a kinetic model and a nucleation model
approach. Phase Transit 77:115129
Giesen B, Orthner HR, Kowalik A, Roth P (2004b) On the
interaction of coagulation and coalescence during gas-

123

J Nanopart Res (2011) 13:803815


phase synthesis of fe-nanoparticle agglomerates. Chem
Eng Sci 59:22012211
Guo B, Kennedy IM (2007) Gas-phase flame synthesis and
characterization of iron oxide nanoparticles for use in a
health effects study. Aerosol Sci Technol 41:944951
Hawa T, Zachariah MR (2007) Development of a phenomenological scaling law for fractal aggregate sintering from
molecular dynamics simulation. J Aerosol Sci 38:793806
Hecht C, Kronemayer H, Dreier T, Wiggers H, Schulz C
(2009) Imaging measurements of atomic iron concentration with laser-induced fluorescence in a nanoparticle
synthesis flame reactor. Appl Phys B 94:119125
Height MJ, Howard JB, Tester JW, Sande JBV (2004) Flame
synthesis of single-walled carbon nanotubes. Carbon
42:22952307
Kazakov A, Frenklach M (1998) Dynamic modeling of soot
particle coagulation and aggregation: Implementation
with the method of moments and application to highpressure laminar premixed flames. Combust Flame 114:
484501
Knight PC, Seville JPK, Kamiya H, Horio M (2000) Modelling
of sintering of iron particles in high-temperature gas fluidisation. Chem Eng Sci 55:47834787
Kock BF, Kayan C, Knipping J, Orthner HR, Roth P (2005)
Comparison of LII and TEM sizing during synthesis of
iron particle chains. Proc Combust Inst 30:16891697
Lide DR (2010) CRC handbook of chemistry and physics, 90th
edn. CRC Press/Taylor and Francis, Boca Raton, FL
Lummen N, Kraska T (2005a) Homogeneous nucleation of
iron from supersaturated vapor investigated by molecular
dynamics simulation. J Aerosol Sci 36:14091426
Lummen N, Kraska T (2005b) Molecular dynamics investigations of the coalescence of iron clusters embedded in an
inert-gas heat bath. Phys Rev B 71:205403
Lummen N, Kraska T (2006) Influence of the carrier gas on the
formation of iron nano-particles from the gas phase: a
molecular dynamics simulation study. Comput Mater Sci
35:210215
Marchese M, Jacucci G, Flynn CP (1988) Isotope effect of
vacancy diffusion in bcc alpha-fe. Philos Mag Lett
57:2530
McCarthy DN, Brown SA (2009) Evolution of neck radius and
relaxation of coalescing nanoparticles. Phys Rev B
80:064107
McGraw R (1997) Description of aerosol dynamics by the
quadrature method of moments. Aerosol Sci Technol
27:255265
Merchan-Merchan W, Saveliev AV, Kennedy L, Jimenez WC
(2010) Combustion synthesis of carbon nanotubes and
related nanostructures. Prog Energy Combust Sci
36:696727
Moisala A, Nasibulin AG, Shandakov SD, Jiang H, Kauppinen
EI (2005) On-line detection of single-walled carbon
nanotube formation during aerosol synthesis methods.
Carbon 43:20662074
Moniruzzaman CG, Park HG, Park KY (2007) Analysis of iron
particle growth in aerosol reactor by a discrete-sectional
model. Korean J Chem Eng 24:299304
Nasibulin AG, Pikhitsa PV, Jiang H, Kauppinen EI (2005)
Correlation between catalyst particle and single-walled
carbon nanotube diameters. Carbon 43:22512257

J Nanopart Res (2011) 13:803815


Oxtoby DW (1992) Homogeneous nucleationtheory and
experiment. J Phys Condens Matter 4:76277650
Pikhitsa PV, Choi M (2007) Fast fragmentation of metal oxide
nanoparticles via reduction in oxyhydrogen flame. Appl
Phys Lett 90:163106
Pope CJ, Howard JB (1997) Simultaneous particle and molecule modeling (spamm): an approach for combining sectional aerosol equations and elementary gas-phase
reactions. Aerosol Sci Technol 27:7394
Richter H, Treska M, Howard JB, Wen JZ, Thomasson SB,
Reading AA, Jardim PM, Vander Sande JB (2008) Large
scale combustion synthesis of single-walled carbon
nanotubes and their characterization. J Nanosci Nanotechnol 8:60656074
Shibuta Y, Suzuki T (2008) A molecular dynamics study of the
phase transition in bcc metal nanoparticles. J Chem Phys
129:144102
Shibuta Y, Watanabe Y, Suzuki T (2009) Growth and melting
of nanoparticles in liquid iron: a molecular dynamics
study. Chem Phys Lett 475:264268
Vander Wal RL, Hall LJ (2001) Flame synthesis of fe catalyzed single-walled carbon nanotubes and ni catalyzed
nanofibers: Growth mechanisms and consequences. Chem
Phys Lett 349:178184
Weber AP, Friedlander SK (1997) In situ determination of the
activation energy for restructuring of nanometer aerosol
agglomerates. J Aerosol Sci 28:179192

815
Wen JZ, Goldsmith CF, Ashcraft RW, Green WH (2007)
Detailed kinetic modeling of iron nanoparticle synthesis
from the decomposition of fe(co)(5). J Phys Chem C
111:56775688
Wen JZ, Richter H, Green WH, Howard JB, Treska M, Jardim
PM, Sande JBV (2008a) Experimental study of catalyst
nanoparticle and single walled carbon nanotube formation
in a controlled premixed combustion. J Mater Chem
18:15611569
Wen JZ, Celnik M, Richter H, Treska M, Sande JBV, Kraft M
(2008b) Modelling study of single walled carbon nanotube formation in a premixed flame. J Mater Chem
18:15821591
Yasuoka K, Matsumoto M (1998) Molecular dynamics of
homogeneous nucleation in the vapor phase. I. Lennardjones fluid. J Chem Phys 109:84518462
Yu SY, Yoon Y, Muller-Roosen M, Kennedy IM (1998) A
two-dimensional discrete-sectional model for metal aerosol dynamics in a flame. Aerosol Sci Technol 28:185196
Zhao B, Yang ZW, Johnston MV, Wang H, Wexler AS, Balthasar M, Kraft M (2003) Measurement and numerical
simulation of soot particle size distribution functions in a
laminar premixed ethylene-oxygen-argon flame. Combust
Flame 133:173188
Zhao H, Liu XF, Tse SD (2009) Effects of pressure and precursor loading in the flame synthesis of titania nanoparticles. J Aerosol Sci 40:919937

123

You might also like