Sunteți pe pagina 1din 17

Aerosol Science 33 (2002) 1309 – 1325

www.elsevier.com/locate/jaerosci

Approaches to increasing yield in evaporation=condensation


nanoparticle generation
Yogendra Singha , Julie R.N. Javiera , Sheryl H. Ehrmana; ∗ , Martin H. Magnussonb ,
Knut Deppertb
a
Department of Chemical Engineering, University of Maryland, College Park, MD 20742 2111 USA
b
Solid State Physics, Lund University, Box 118, S221 00 Lund, Sweden

Received 1 June 2001; received in revised form 19 April 2002; accepted 22 April 2002

Abstract

With the recent interest in the chemical, electronic and optical properties of nanometer scale metal particles,
there is now interest in manufacturing these materials in larger quantities. Since both small particle size and
high particle number concentrations are sought, there is a need for improved particle generation reactors
that can realize both goals. Here, results are presented for the synthesis of indium metal nanoparticles in an
evaporation=condensation aerosol generator. Size distributions were measured for metal nanoparticles formed
using a standard 2ow con3guration, as well as using several variations on the standard con3guration. The aim
of the modi3cations is to increase the cooling rate and thus, to increase the nucleation rate of the nanoparticles.
An increase in the number concentration of particles and, in some cases, a signi3cant decrease in average
particle size was observed when the modi3ed reactor con3gurations were used. These results can be explained
by the changes in the time–temperature history of the nanoparticles resulting from the modi3cations to the
aerosol generator. A monodisperse model of nanoparticle formation and growth, accounting for nucleation,
condensation and coagulation, was used to describe particle formation in the standard 2ow con3guration,
to guide the modi3cations, and to describe particle formation in one of the modi3ed con3gurations, with
qualitative agreement seen between measured and predicted particle sizes. ? 2002 Elsevier Science Ltd. All
rights reserved.

1. Introduction

Metal nanoparticles are of interest for a variety of uses including catalysis, as masks for nanofab-
rication techniques, and as precursors in the formation of compound semiconductor nanoparticles.


Corresponding author. Tel.: +1-301-405-1917; fax: +1-301-314-9126.
E-mail address: sehrman@eng.umd.edu (S.H. Ehrman).

0021-8502/02/$ - see front matter ? 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 2 1 - 8 5 0 2 ( 0 2 ) 0 0 0 7 2 - 1
1310 Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325

Since many of the material properties of interest are size dependent, most applications require unag-
glomerated particles with a controlled size distribution, and these requirements limit the scalability
of many production processes.
Various processes for synthesizing metal nanoparticles from the gas phase have been developed
at the laboratory scale. In one of the earliest systematic studies of metal nanoparticle formation by
evaporation and condensation in a low pressure (0.5 –4 Torr) inert gas atmosphere, Granqvist and
Buhrman (1976) found that the size distributions of various metal nanoparticles were log-normal in
shape, if the metal particles were spherical (¡ 20 nm in diameter). The size distribution of particle
populations containing larger particles with well de3ned crystal habits were found to deviate from a
log-normal shape. Scheibel and PorstendGorfer (1983) described a device based upon evaporation and
condensation in a horizontal tube furnace at atmospheric pressure for the generation of monodisperse
silver and NaCl aerosols in the size range of 2–300 nm. With this generator, they studied particle
formation by measuring particle size distributions of the condensation aerosols as a function of
furnace temperature, carrier gas 2ow rate, and position of the boat containing the metal.
Some more recent eHorts have been focused on increasing the production rate of metal nano-
particles and on developing continuous processes for synthesis and collection. Mahoney and
Andres (1995) developed a versatile arc evaporation reactor for producing nanoscale unagglom-
erated, equiaxed metal clusters, and metal-based ceramics. In this process, production rates of grams
per hour are possible because of the enhanced evaporation rate of the metal. In another approach
to metal nanoparticle synthesis, termed the aerosol 2ow condenser, Haas, Birringer, Gleiter, and
Pratisinis (1997) used radio frequency (RF) heating to enhance metal evaporation, combined with
jets of carrier gas 2owing over the metal source to produce palladium particles in a continuous
process. Compared to particles formed by traditional inert gas condensation methods (Gleiter, 1992),
the palladium particles formed were considerably smaller, and a reduction in the geometric standard
deviation of the particle size distribution was observed.
The reactor system used here, an evaporation=condensation aerosol generator, was selected because
it is widely used at the laboratory scale, and can be easily modi3ed, allowing a comparison of
the eHect of our modi3cations on the 3nal particle size distribution and number concentration to
results obtained using the original, unmodi3ed 2ow con3guration. The goal of our modi3cations is
to simultaneously increase the number concentration of nanoparticles produced and to decrease the
average particle size. We have chosen to study the formation of indium metal nanoparticles, but the
results obtained here are generally applicable to any material formed by the evaporation=condensation
route.
To obtain insight into the physical phenomena occurring during nanoparticle growth from the gas
phase, and to guide our modi3cations to the aerosol generator, theoretical predictions of particle
formation rates and growth dynamics are helpful. For these purposes, various models have been
developed for the case of instantaneous coalescence, an assumption that, as described later, we
believe is reasonable for our system. Friedlander (1983) reported a solution for the growth of
particles by nucleation and condensation for isothermal conditions, with no assumptions regarding
the particle size distribution, using the 3rst three moments of the particle size distribution. Pratsinis
(1988) developed a moment-based model accounting for the eHects of nucleation, condensation, and
coagulation on particle growth dynamics, assuming a unimodal log-normal particle size distribution.
This moment-based approach was later extended to incorporate thermophoresis and diHusion as well
as two-dimensional variations in concentration, temperature, and velocity (Phanse & Pratsinis, 1989).
Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325 1311

Other more computationally intensive approaches, including those developed by Girshick and Chiu
(1989) and Xiong and Pratsinis (1991), use discrete-sectional techniques to describe the evolution
of the particle size distribution.
In the simple case of particles uniform in size that coalesce instantaneously, Panda and Pratsinis
(1995) have shown that particle formation and growth can be described by balances on the metal
vapor atoms, the particle number concentration and the particle volume concentration. The model we
have used in this study is based upon their approach, developed originally to describe the growth of
aluminum particles by an evaporation=condensation process. We selected this methodology because
it has been shown to give results for metal nanoparticle formation and growth that are comparable
to those obtained using much more computationally intensive models (Panda & Pratsinis, 1995).
The estimates for the 3nal primary particle size and number concentration of indium nanoparticles
obtained using the model were 3rst compared with the experimental results for diHerent furnace
temperatures, obtained using a standard con3guration. The model was then used to estimate the
eHect of increasing the cooling rate on particle number and average size for several hypothetical
temperature pro3les. Guided by our numerical results, we made two simple modi3cations to the
reactor to increase the rate of cooling, and the particle number concentration and size distributions
were compared to those obtained using the original standard con3guration.

2. Experimental methods

In the standard con3guration, shown in Fig. 1, puri3ed nitrogen carrier gas (¿ 99:9999%), passes
through the reactor, an alumina tube, 60 cm in length and 1:8 cm in inner diameter, heated externally
by a 22 cm long horizontal tube furnace. For these experiments, the set point of the furnace was
between 1173 and 1373 K. A boron nitride boat containing indium metal was placed inside the tube
at the point of maximum temperature to obtain the highest possible metal evaporation rate. At these
temperatures, the metal is in the liquid phase, and as it evaporates, it provides a source of metal
vapor. The carrier gas 2ow rate was kept constant at 1.68 standard liters per minute (slpm). The
temperature inside the furnace was measured along the centerline as a function of axial distance
using a type N thermocouple, and the temperature pro3les are reported in Fig. 2. The Reynolds
numbers for this system calculated at the various furnace set point temperatures do not exceed 20,
indicating laminar 2ow.

Fig. 1. Schematic of the standard evaporation=condensation setup.


1312 Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325

Fig. 2. Measured temperature pro3le inside the reactor tube for various maximum furnace set point temperatures and for
the QTAO setup at 1373 K. The zero position, with respect to axial distance, corresponds to the inlet of the alumina tube.
For reference, the location of the furnace is shown with dashed lines.

The particle size distribution was measured downstream of the alumina tube using the following
home built instrumentation: an ultraviolet particle charger, a Vienna-type diHerential mobility analyzer
(DMA), and an electrometer. The ultraviolet charger uses the photoelectric eHect to produce a stream
of positively charged particles and free electrons by irradiating the particle-laden stream with 222 nm
wavelength light. The free electrons can then charge the neutral particles, with the 3nal result being
a stream of positively and negatively charged particles (Burtscher, Scherrer, Siegmann, Schmidt-Ott,
& Federer, 1982). In the DMA, diHerences in mobility of charged particles in an electric 3eld as a
function of particle size are used to select a desired nanoparticle size range (Winkelmayer, Reischl,
Lindner, & Berner, 1991). The electrometer measures the current carried by the stream of particles,
thus giving the number concentration of a particular size range assuming singly charged particles.
The reported size distribution was corrected for the charging eOciency of the ultraviolet charger, an
experimentally determined function of particle diameter (Deppert & Magnusson, 1999). The charging
probability using the ultraviolet charger is greater for smaller particles as compared to the charging
probability (Boltzmann distribution) obtained using the more common Kr-85 charger. The mass of
metal evaporated was determined by weighing the boat before and after each experiment using an
analytical balance.
The modi3cations to the furnace are shown in Figs. 3 and 4. In the 3rst modi3ed setup, shown
in Fig. 3, a quartz tube of inner diameter 4 mm, and 25:4 cm in length, was placed inside the
alumina tube downstream of the boat containing the metal. We refer to this con3guration as the
quartz-tube-as-outlet (QTAO) setup. The 2ow over the boat was equal to 1:51 slpm. To reduce
losses of the vapor from diHusion to the reactor tube walls, a carrier gas stream with a 2ow rate
of 0:17 slpm was introduced from the opposite side, giving a total 2ow rate of 1:68 slpm. The
entire 2ow through the furnace exits through the quartz tube, and the vapor condenses to form
nanoparticles within the quartz tube. The temperature pro3le inside the quartz tube for a furnace set
point temperature of 1373 K was measured using a type N thermocouple, and is shown in Fig. 2. The
Reynolds number for the 2ow of nitrogen through the small quartz tube was higher, approximately
65, but still well below the transition to turbulent 2ow.
Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325 1313

Fig. 3. Schematic, not to scale, of the QTAO modi3ed setup.

Fig. 4. Schematic, not to scale, of the QTAB modi3ed setup.

In the other modi3cation reported here, shown in Fig. 4, a quartz tube 4 mm in inner diameter
and 35:4 cm in length was placed inside the alumina tube at the inlet side. The outlet of the
quartz tube was located downstream of the boat, just before the end of the tube furnace. Based
upon observations of the location of metal deposition in the alumina tube in experiments using the
standard con3guration, it is believed that nucleation occurs a few centimeters downstream of where
the alumina tube exits the furnace, with the exact location depending upon the furnace temperature
pro3le. Hence, it is believed that with the quartz tube in this position, the time–temperature history
in the region of nucleation will be aHected. We refer to this arrangement as the quartz-tube-after-boat
(QTAB) setup. In this case, the carrier gas 2ows through both the quartz tube and the annular space
between the quartz and alumina tubes. The total 2ow rate was kept constant at 1:68 slpm, but the
2ow rates of the two streams were varied so that the ratio of the average velocity in the quartz tube
to the average velocity in the annular space ranged from 1:1 to 20:1. The details of the 2ow rates
are given in Table 1. At the higher velocity ratios, a con3ned co2owing annular jet is formed, with
a jet Reynolds number of 33 for the 20:1 velocity ratio, indicating that the jet is laminar.

3. Theory

In the evaporation–condensation reactor, the carrier gas passes over the boat, carrying away metal
vapor. As the mixture of vapor and carrier gas 2ows downstream and cools, particles form and
grow by the processes of nucleation, condensation and coagulation. Here, the term nucleation is used
to describe the formation of stable, chemically homogeneous metal clusters from a supersaturated
1314 Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325

Table 1
Details of 2ow rates for the QTAB experiments

Velocity ratio
(quartz tube:annular region) 1:1 2:1 20:1
−1
Average velocity (cm min )
Quartz tube 711 1354 7356
Annular region 711 677 368

Volumetric 2ow rate ratio


(quartz tube:annular region) 1:20 1:9.5 1.02:1
Flow rate (slpm)
Quartz tube 0.08 0.16 0.85
Annular region 1.60 1.52 0.83

Jet Reynolds number at exit of quartz tube 3 6 33

vapor. Condensation is the process of cluster growth via collisions with vapor phase metal atoms
(monomers). Coagulation is the process of cluster growth via collisions between metal clusters,
followed by coalescence of the clusters (Friedlander, 1977).
An important consideration is whether the particles are spheroids or dendritic aggregates. We can
look to estimates of characteristic times for interparticle collision and particle coalescence to give in-
sight into the 3nal particle morphology (Flagan & Lunden, 1995; Lehtinen, Windeler, & Friedlander,
1996). The characteristic time for collisions is de3ned as the time required for the cluster number
concentration to decrease by a factor of two, and the characteristic time for coalescence is de3ned as
the time required for two coalescing spheres to form a single sphere. From a modeling point of view,
it is important to determine which of these two mechanisms, collision or coalescence, limits the rate
of particle growth. If particles coalesce faster than they collide, the particles will be spherical, but if
the time required for coalescence is greater than the time between interparticle collisions, then the
particles will form aggregates (Flagan & Lunden, 1995; Lehtinen et al., 1996). Transmission electron
microscope images of indium particles obtained for a furnace maximum temperature of 1373 K, an
example of which is shown in Fig. 5, suggest that the particles are spherical. The faint core-shell
pattern visible in the image is believed to result from shell melting of the nanoparticles under the
electron beam. The following simple calculation serves to support the assertion that the particles are
coalescing completely between collisions.
The characteristic coalescence time, de3ned as the time for two equal size spheres to coalesce via
a viscous 2ow mechanism, is given by (Frenkel, 1945)
dp
coal = p ; (1)

where coal is the coalescence time, p is the particle viscosity, dp is the particle diameter and
is the surface tension of the metal. The collision time, de3ned as the time required for the initial
number concentration of the particles to be reduced by half by collisions between particles, is given
by (Friedlander, 1977)
2
coll = ; (2)

N g
Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325 1315

Fig. 5. Transmission electron microscope image of indium nanoparticles.

where coll is the collision time,


is the collision frequency function, N is the number of particles
per unit mass of carrier gas, and g is the density of the carrier gas.
The collision frequency function for particles of uniform diameter, valid from the free molecular
to the continuum regime, is given by (Seinfeld, 1986)
 √ −1
dp 4 2D

= 8 Ddp √ + ; (3)
dp + 2g cdp

where the particle diHusion coeOcient D, particle velocity c, and the transmission parameter g are
given by the following equations:
 
kB T 5 + 4Kn + 6Kn2 + 18Kn3
D= ; (4)
3 g dp 5 − Kn + (8 + )Kn2
where T is temperature, Kn is the Knudsen number (Kn = 2=dp ) and  is the mean free path of
the gas.

8kB T
c= ; (5)
p p
1
g= [(dp + p )3 − (d2p + p2 )3=2 ] − dp ; (6)
3dp p
where vp is particle volume, and the mean free path for the particles given by
8D
p = : (7)
c
1316 Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325

From experimental results for the average particle size as a function of furnace temperature, char-
acteristic times for collisions and coalescence were estimated, using Eqs. (1) and (2), at conditions
corresponding to an axial location 13 cm downstream of where the reactor tube exits the furnace.
The temperature was taken to be 459 K, the measured temperature at that point for a furnace set
point temperature of 1373 K. The particles were assumed to be monodisperse, with a diameter equal
to the measured volume average diameter, 54 nm, and a number concentration estimated from the
measured metal sublimation rate assuming all of the material condensed to form nanoparticles. The
viscosity, p and surface tension, , of indium at 459 K are 1:6 × 10−3 kg m−1 s−1 and 0:55 N m−2 ,
respectively (Smithells, 1992; CRC, 1996). The characteristic collision time was relatively long, in
the order of a second. However, the characteristic time for coalescence was very short, in the order
of 10−10 s, suggesting that coalescence is eHectively instantaneous under these conditions. These
results suggest that the assumption that the particles are spherical (at least in the furnace) is reason-
able, a result which is not surprising, considering the low melting point of indium, 428 K, and the
process temperatures. Here, the use of a model in which coalescence is assumed instantaneous is jus-
ti3ed, but this has to be veri3ed for other temperature histories or for materials with higher melting
points.
Particle formation and growth can be described, for the simple case of particles that are uniform
in size, and that coalesce instantaneously, by balances on the metal vapor atoms, the particle num-
ber concentration and the particle volume concentration (Panda & Pratsinis, 1995). Following their
approach, the equation for the rate of change of the number of monomers of the condensing metal
per unit mass of gas, nm , can be written as

dnm Ig∗ 2 kB T
=− − dp nms (S − 1) Nf(Kn); (8)
dt g 2 m1
where I is the nucleation rate and f(Kn), a function of Knudsen number (Warren & Seinfeld,
1984). The 3rst term on the right hand side (RHS) of Eq. (8) represents the loss of monomers by
nucleation, while the second RHS term describes the loss of monomers by condensation. The carrier
gas density g , is calculated assuming ideal gas behavior. The saturation ratio is given by
n m g
S= ; (9)
nms
where the monomer concentration per unit volume of gas, nms , at saturation is
ps
nms = (10)
kB T
with ps equal to the equilibrium vapor pressure of the condensing metal.
To describe nucleation, a variety of improvements to the classical theory of Becker and DGoring
(1935) have been proposed (see for example Girshick & Chiu, 1989; McClurg & Flagan, 1998;
Martinez, Ferguson, Heist, & Nuth, 2001 and references therein). There is still considerable discus-
sion as to the most accurate approach. Here, because we cannot measure nucleation rates directly,
and thus cannot evaluate which model gives the most exact results, we simply use the nucleation
rate derived from classical nucleation theory (Friedlander, 1977):
   1=2  
p1 2=3 12=3 16 3 21
I =2 (nm 1 ) exp − ; (11)
(2 m1 kB T )1=2 kB T 3(kB T )3 (ln S)2
Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325 1317

where p1 is the partial pressure of the nucleating species and m1 , and 1 are the mass and volume
of the monomer, respectively. The number of monomers (g∗ ) in the particle of critical size is
 3
∗ 2
g = ; (12)
3(ln S)
where the dimensionless surface tension group  is
s1
= ; (13)
kB T
where s1 is the monomer surface area.
Similarly, the equation describing the change in the number concentration of particles with time
can be written as
dN I 1
= −
N 2 g (14)
dt g 2
with the 3rst RHS term representing the formation of particles by nucleation and the second RHS
term accounting for the loss of particles by coagulation.
The particle volume concentration per unit mass of gas is given by

dV Ig∗ 2 kB T
= 1 + dp nms 1 (S − 1) Nf(Kn); (15)
dt g 2 m1
where the 3rst term on the RHS describes the volume contribution by nucleation and the second
term stands for the gain in particle volume via condensation of monomers. Assuming particles of
spherical shape, the average particle diameter, dp is
 1=3
6V
dp = : (16)
N
A continuous expression for temperature as a function of axial distance is obtained by 3tting a
polynomial function to the experimentally measured temperature pro3les. The residence time, t, can
be related to the axial position, x, and the reactor diameter, d, assuming a uniform plug 2ow velocity
pro3le:
dt d2
= (17)
dx 4Q
with the gas 2ow rate given by
Qrt T (x)
Q= ; (18)
Trt
where the subscript rt refers to reference conditions.

4. Results and discussion

4.1. Experimental and model results—standard con4guration

The measured particle size distributions, adjusted for charger eOciency, are shown for diHerent
furnace maximum set point temperatures in Fig. 6. It can be seen that an increase in the furnace
1318 Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325

Fig. 6. Experimentally determined particle number concentration versus particle diameter for various furnace set point
temperatures, using the standard (STD) setup.

Table 2
Comparison of experimental and numerical number average particle diameter for diHerent furnace temperature

Furnace set point Volume average particle diameter (nm)

temperature (K) Experimental Predicted


1173 15 7
1273 27 13
1373 54 36

set point temperature results in an increase in the mean particle size. These results are in agreement
with results reported previously for other metallic material systems by Scheibel and PorstendGorfer
(1983), Panda and Pratsinis (1995), and Magnusson, Deppert, Malm, Bovin, and Samuelson (1999),
among others.
From the experimental size distribution data, obtained for various furnace set point temperatures
in the manner described previously, the volume average particle diameter was calculated and was
compared with the particle diameter estimates obtained using the model. The experimental and the
predicted particle diameter for diHerent furnace temperatures, tabulated in Table 2, agree qualitatively.
It should be mentioned here that our measurements re2ect the number concentration of particles as
a function of mobility diameter measured at the electrometer, and a signi3cant amount of metal may
have been lost prior to this point via deposition to the reactor walls and downstream in the process
tubing. Some coagulation may also be occurring between the reactor and the DMA. These processes
will reduce the total measured number concentration, and, because small particles are more mobile,
the size distribution may be aHected. With these considerations, it is presumable that the true volume
average particle diameter at the reactor exit is less than our measured values. Uncertainty in the use
of the classical expression for the nucleation rate, use of a monodisperse model, as well as the use
of simple one-dimensional temperature and velocity pro3les (rather than two- or three-dimensional
pro3les) may also contribute to the gap between measured and predicted particle size. Predictions
also were made using a nucleation rate based upon the nucleation theory proposed by Girshick and
Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325 1319

Fig. 7. Average particle number concentration pro3les, N , for various cooling rates as a function of axial distance, estimated
using the monodisperse model.

Chiu (1990), and the agreement between observed and predicted values did not improve. Regardless,
the model is still useful for estimating the sensitivity of particle size and number concentration to
various experimental parameters.
As shown by Girshick and Chiu (1989) and Panda and Pratsinis (1995) among others, a decrease
in particle size will be observed for high cooling rates, low evaporation temperatures, low system
pressures, and low metal vapor concentration. If the concentration of metal vapor is decreased by
lowering the evaporation temperature, or by some other modi3cation to reduce the metal vapor
concentration, the production rate of particles will decrease. Because our objectives are to increase
the rate of production of particles as well as to decrease their size, we focus on the cooling rate as
our adjustable experimental parameter.

4.2. Modeling results for hypothetical time–temperature histories

A hypothetical base case is considered in which indium vapor is generated at a rate of 8:3 ×
10−3 mg s−1 , corresponding to the experimentally measured evaporation rate for a furnace set point
temperature of 1373 K. Three diHerent cooling rates, 2300, 5000, and 10 000 K s−1 were considered,
with 2300 Ks−1 corresponding to the maximum measured cooling rate for the standard 1373 K case.
Fig. 7 shows the evolution of particle number concentration, reported as number of particles per
kilogram of carrier gas, for the various cooling rates. As the cooling rate increases, the particle
number concentration increases, since an increase in the cooling rate will result in a higher rate
of nucleation and a greater total number of particles formed. Correspondingly, the 3nal particle
diameter, as shown in Fig. 8, decreases with increasing cooling rate. The location of the onset of
particle formation is also aHected by the cooling rate as shown in both Figs. 7 and 8, with slower
cooling resulting in particle formation occurring further downstream.
1320 Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325

Fig. 8. Evolution of the average particle diameter, dp , for various cooling rates as a function of axial distance, estimated
using the monodisperse model.

Fig. 9. Experimental particle size distribution for the quartz tube as outlet (QTAO) case, and the standard (STD) case,
for a furnace temperature of 1373 K.

4.3. Experimental and model results—modi4ed con4gurations

The measured particle size distributions, for the QTAO and standard setup at a set point tempera-
ture of 1373 K are shown in Fig. 9. Experimentally determined yields for the modi3ed and standard
con3gurations as well as metal evaporation rates are given in Table 3. The yield was estimated
by 3rst estimating the number concentration of particles before the charger based upon charging
eOciency as a function of size, converting the number size distribution to a mass distribution and
summing the particle mass in each size range. This, multiplied by the gas 2ow rate, was then divided
by the mass evaporated from the boat to determine yield.
Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325 1321

Table 3
Metal evaporation rates and estimated yields, for all runs at 1373 K

Flow con3guration Temperature (K) Rate of metal evaporated Yield (%)


from boat (mg min−1 )

Standard 1373 0.50 3.4


QTAO 1373 0.50 3.5
QTAB 1:1 1373 0.28 9.9
QTAB 2:1 1373 0.28 4.6
QTAB 20:1 1373 0.26 29

Fig. 10. Temperature in the cooling zone as a function of residence time, calculated based upon temperature measurements
as a function of distance, for both the standard (STD) con3guration and the QTAO con3gurations at a furnace set point
temperature of 1373 K.

Comparing the two setups, the total particle number concentration obtained for the QTAO setup
was approximately twice the number concentration obtained for the standard setup. This result is all
the more signi3cant, considering that the evaporation rate is approximately the same. The volume
average particle size was also smaller, 35 nm, versus 54 nm for the standard case, as would be ex-
pected since the overall yield is unchanged. The geometric standard deviation (GSD) was unchanged,
1.5 for each case. It should be mentioned that, though the size distributions were not measured as
a function of residence time as that was outside the limitations of our apparatus, the GSD is sim-
ilar to that obtained for self-preserving log-normal distributions (Lai, Friedlander, Pich, & Hidy,
1972). Fig. 10 shows temperature versus residence time, calculated from experimentally measured
temperature pro3les downstream of the maximum temperature assuming a 2at velocity pro3le, for
the standard 1373 K and the QTAO case. As apparent in Fig. 10, the cooling rate for the QTAO
case is much greater than the cooling rate for the standard 1373 K case. Using the quartz tube as the
reactor outlet increases the average velocity in the cooling region by a factor of 20. Hence, while the
magnitude of the rate of change of temperature with position is slightly less than the standard case,
the rate of change of temperature with time is much greater and this results in a higher nucleation
rate.
1322 Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325

Fig. 11. Experimental particle size distributions for the QTAB and the standard (STD) cases, for a furnace temperature
of 1373 K.

The rate of nucleation as a function of axial distance was calculated as a function of axial
distance for the QTAO case and for the standard 1373 K setup. The maximum nucleation rate for
the modi3ed setup was 2 × 1020 kg−1 s−1 whereas for the standard 1373 K case, it was found to
be 2 × 1017 kg−1 s−1 . Estimated nucleation rates are reported in units of per mass of carrier gas
per time to allow for comparison of rates obtained for diHerent temperature histories. If the same
concentration of vapor is condensing, a greater nucleation rate will result in a smaller average particle
size. The experimental rates of metal evaporation were approximately the same for the two setups,
29:8 mg h−1 for the standard 1373 K case and 29:4 mg h−1 for the QTAO modi3cation, suggesting
the same concentration of vapor is condensing in each case, and the observed diHerences are only
due to the diHerences in cooling rate.
Because the velocity pro3les were similar to the standard case, we used the model to predict
particle size for these experimental conditions. As done previously, a uniform velocity pro3le was
assumed, and a polynomial 3t to the measured temperature pro3le reported for the QTAO case in
Fig. 2 was used. From the model calculations, an average 3nal particle size of 11 nm was predicted,
a much greater shift to smaller size than was observed experimentally (35 nm). Again, uncertainty
in the nucleation rate, neglecting the eHect of wall deposition, and use of only one-dimensional
temperature and velocity pro3les may be responsible for the disagreement between the observed
results and estimates based upon theory.
Experimental results for the second modi3cation, the QTAB setup, are shown in Fig. 11. From
the results in Table 3, this modi3cation leads to a decrease in the metal evaporation rate, under-
standable as convection over the boat decreases. At the lowest velocity ratio, 1:1, the total number
concentration is lower than the standard case, possibly re2ecting the decrease in the amount of
available metal vapor because of the slower evaporation rate. This phenomena is even more obvious
for the 2:1 velocity ratio. The GSD of the size distributions also vary somewhat, larger for the 1:1
velocity ratio (1.7), versus 1.5 for the 2:1 and 20:1 velocity ratios. Most notably, at the highest
velocity ratio, 20:1, there is a signi3cant increase in the total number concentration of nanoparticles,
Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325 1323

approximately 2.5 times the total number concentration obtained for the standard case. Again, this
is a notable result because of the slower evaporation rate, and thus, less available metal vapor.
Correspondingly, the yield is the highest for this case. It may be possible that wall losses imme-
diately after particle formation are signi3cantly reduced because of the jet-like 2ow in the region
where nucleation is believed to occur. Considering particle size, a decrease in the volume average
particle diameter was observed for the 20:1 velocity ratio, 40 nm versus 54 nm for the standard
setup.
The velocity, temperature, and concentration pro3les that result from this modi3cation are much
more complex than for the previously mentioned con3gurations. It is beyond the capability of the
one-dimensional model used here to describe them, but the eHects of this con3guration on the time–
temperature history can be qualitatively described. As measured with a type N thermocouple, the
center jet gas temperature at the jet outlet is, within the experimental uncertainty of the temperature
measurement, the same as the annular gas stream temperature for each of the velocity ratios. This
indicates that the gas that forms the center jet has suOcient residence time in the furnace to reach
the temperature of the surrounding gas prior to exiting the quartz tube, and the jet does not act
to cool the metal vapor=carrier gas mixture. The center jet initially contains no metal vapor, but
as the center jet and the coaxial stream come into contact, the velocity of the gas in the outer
stream adjacent to the jet will increase as momentum is transferred to it from the jet. There may
also be some diHusion of metal vapor into the faster 2owing center jet. The result is that some
of the metal vapor will experience an increased rate of cooling, and this eHect increases as the
velocity ratio increases. A drawback to this modi3cation is that it results in a broader range of
possible time–temperature histories than either the standard con3guration or the QTAO setup, and
thus, as observed here, this would not be the best approach for reducing the spread of the size
distribution.

5. Conclusions

Variations on the evaporation=condensation method were used to alter the time–temperature history
during the synthesis of indium nanoparticles. Particle size distributions were measured for indium
nanoparticles using both the standard and modi3ed setups. The results obtained were compared with
respect to the particle concentration and particle diameter. An increase in the number concentration
of particles and, in some cases, a signi3cant decrease in average particle size was observed when
the modi3ed reactor con3gurations were used. The smallest average size was achieved using the
QTAO reactor con3guration, but the QTAB con3guration, with the 20:1 velocity ratio, resulted in
the greatest increase in number concentration. A one-dimensional monodisperse model was used to
describe the dynamics of particle synthesis and growth and to give insight into the mechanisms
leading to an increase in the number concentration and decrease in average particle size. Our model
results suggest that the experimental observations can be attributed to the increase in the nucleation
rate that arises from the steeper temperature gradient in the modi3ed con3gurations. Our experimental
observations demonstrate that simple changes to evaporation=condensation aerosol generators will
result in a decrease in the average particle size, without sacri3cing particle yield, measured in terms
of total number concentration.
1324 Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325

Acknowledgements

The authors gratefully acknowledge the support of the United States National Science Foundation
(grant CTS-9973845), and the Nanometer Consortium, Lund University, Sweden. J.R.N. Javier also
acknowledges the partial support of the ASPIRE program at the University of Maryland, College
Park. We thank Jan-Olle Malm for supplying the TEM image.

References

Becker, R., & DGoring, W. (1935). Kinetische Behandlung der Keimbildung in uG bersGattigen DGampfen. Ann allen des
Physik, 24, 719–752.
Burtscher, H., Scherrer, L., Siegmann, H. C., Schmidt-Ott, A., & Federer, B. (1982). Probing aerosols by photoelectric
charging. Journal of Applied Physics, 53, 3787–3791.
CRC Handbook of Chemistry and Physics (77th ed.) (1996). Cleveland, Ohio: Chemical Rubber Publishing Co.
Deppert, K., & Magnusson, M. H. (1999). Unpublished results.
Flagan, R. C., & Lunden, M. M. (1995). Particle structure control in nanoparticles synthesis from the vapor phase.
Materials Science and Engineering, A204, 113–124.
Frenkel, J. (1945). Viscous 2ow of crystalline bodies under the action of surface tension. Journal of Physics, 9, 385–391.
Friedlander, S. K. (1977). Smoke, dust and haze, fundamentals of aerosol behavior. New York: Wiley.
Friedlander, S. K. (1983). Dynamics of aerosol formation by chemical reaction. Annals of the New York Academy of
Sciences, 404, 354–364.
Girshick, S. L., & Chiu, C.-P. (1989). Homogeneous nucleation of particles from the vapor phase in thermal plasma
synthesis. Plasma Chemistry and Plasma Processing, 9, 355–369.
Girshick, S. L., & Chiu, C. P. (1990). Kinetic nucleation theory—a new expression for the rate of homogeneous nucleation
from an ideal supersaturated vapor. Journal of Chemical Physics, 93, 1273–1277.
Gleiter, H. (1992). Nanostructured materials. Advances in Materials, 4, 474–481.
Granqvist, C. G., & Buhrman, R. A. (1976). Ultra3ne metal particles. Journal of Applied Physics, 47, 2200–2219.
Haas, V., Birringer, R., Gleiter, H., & Pratisinis, S. E. (1997). Synthesis of nanostructured powders in an aerosol 2ow
condenser. Journal of Aerosol Science, 28, 1443–1453.
Lai, F. S., Friedlander, S. K., Pich, J., & Hidy, G. M. (1972). Self preserving particle size distribution for Brownian
coagulation in free molecule regime. Journal of Colloid and Interface Science, 39, 395–405.
Lehtinen, K. E. J., Windeler, R. S., & Friedlander, S. K. (1996). Prediction of nanoparticle size and the onset of dendrite
formation using the method of characteristic times. Journal of Aerosol Science, 27, 883–896.
Magnusson, M. H., Deppert, K., Malm, J.-O., Bovin, J.-O., & Samuelson, L. (1999). Gold nanoparticles: Production,
reshaping, and thermal charging Journal of Nanoparticle Research, 1, 243–251.
Mahoney, W., & Andres, R. P. (1995). Aerosol synthesis of nanoscale clusters using atmospheric arc evaporation.
Materials Science and Engineering, A204, 160–164.
Martinez, D. M., Ferguson, F. T., Heist, R. H., & Nuth, J. A. III (2001). Application of scaled nucleation theory to
metallic vapor condensation. Journal of Chemical Physics, 115, 310–316.
McClurg, R. B., & Flagan, R. C. (1998). Critical comparison of droplet models in homogeneous nucleation theory. Journal
of Colloid and Interface Science, 201, 194–199.
Panda, S., & Pratsinis, S. E. (1995). Modeling the synthesis of aluminum particles by evaporation–condensation in an
aerosol 2ow reactor. Nanostructured Materials, 5, 755–767.
Phanse, G. M., & Pratsinis, S. E. (1989). Theory for aerosol generation in laminar 2ow condensers. Aerosol Science and
Technology, 11, 100–119.
Pratsinis, S. E. (1988). Simultaneous nucleation, condensation, and coagulation in aerosol reactors. Journal of Colloid
and Interface Science, 124, 416–427.
Scheibel, H. G., & PorstendGorfer, J. (1983). Generation of monodisperse Ag- and NaCl-aerosols with particle diameters
between 2 and 300 nm. Journal of Aerosol Science, 14, 113–126.
Seinfeld, J. H. (1986). Atmospheric chemistry and physics of air pollution. New York: Wiley.
Y. Singh et al. / Aerosol Science 33 (2002) 1309 – 1325 1325

Smithells, C. J. (1992). Smithells metals reference book. Oxford: Butterworth-Heinman.


Warren, D. R., & Seinfeld, J. H. (1984). Nucleation and growth of aerosol from a continuously reinforced vapor. Aerosol
Science and Technology, 3, 135–154.
Winkelmayer, W., Reischl, G. P., Lindner, A. O., & Berner, A. (1991). A new electromobility spectrometer for the
measurement of aerosol size distribution in the range from 1 to 100 nm. Journal of Aerosol Science, 22, 289–296.
Xiong, Y., & Pratsinis, S. E. (1991). Gas phase production of particles in reactive turbulent 2ows. Journal of Aerosol
Science, 22, 637–655.

S-ar putea să vă placă și