Sie sind auf Seite 1von 12

Journal of

Welcome Comments Help


MATERIALS RESEARCH

Electrically controlled flame synthesis of nanophase TiO2 , SiO2 ,


and SnO2 powders
Srinivas Vemury,a) Sotiris E. Pratsinis,b) and Lowinn Kibbeyc)
Department of Chemical Engineering, University of Cincinnati, Cincinnati, Ohio 45221-0171
(Received 20 April 1996; accepted 27 October 1996)

Nanophase particles with precisely controlled characteristics are made by oxidation of


their halide vapors in electrically assisted hydrocarbon flames using needle-shaped or
plate electrodes. The particle size and crystallinity decrease with increasing field strength
across the flame. The field generated by the electrodes across the flame decreases the
particle residence time in the high temperature region of the flame. Furthermore, it
charges the newly formed particles, resulting in electrostatic repulsion and dispersion
that decreases particle growth by coagulation. Electric fields reduced the primary
particle size of TiO2 , the agglomerate size of SnO2 , and both the agglomerate and
primary size of SiO2 .

I. INTRODUCTION particles unipolarly during their synthesis can reduce the


Nanosized particles have distinctly different proper- particle size and narrow the particle size distribution.
ties compared to bulk materials because the number of Recently, Vemury and Pratsinis7 showed that the
atoms on the particle surface is comparable to that inside average primary particle size of TiO2 particles made
the particle.1 As a result, these particles are characterized by flame oxidation of TiCl4 was decreased from 50 to
by lower melting point, better light absorption, and struc- 30 nm, and its crystallinity was drastically altered by
tural properties. Nanosized particles are also used to form applying a unipolar electric field with needle electrodes
catalysts with high specific surface area and large density across a CH4 -air diffusion flame. They attributed the
of active sites. Though a number of processes have been observed changes to particle charging and repulsion
developed for synthesis of nanoparticles,2 their produc- and the ionic wind from the needles across the flame,
tion cost remains high, limiting, thus, development of which decreased the particle residence time at high
their applications. Flame reactors, on the other hand, temperatures. Though this study nicely revealed the
are routinely used in industrial synthesis of submicron potential of electrical discharge on flame synthesis of
powders with relatively narrow size distribution and high powders, it was confined to a single material and a
purity.3 As such, the flame route has high potential diffusion flame which is hard to characterize for its broad
for relatively inexpensive synthesis of finer, nanosize variation of temperature and gas velocities. Furthermore,
particles at high yields and production rates. the ionic wind can have a dramatic effect on the diffusion
Charging particles during their formation can have flame itself since it may affect reactant gas mixing,8
a profound effect on primary particle size, crystallinity, which, in turn, can alter the particle diameter by as much
degree of aggregation, and agglomerate size. Hardesty as a factor of 10.
and Weinberg4 showed that the silica primary particle Here, the effects of externally controlled electric
size can be reduced by a factor of three when an electric fields on the characteristics of flame-generated powders
field is applied across a counterflow CH4 -air diffusion are investigated in a burner-stabilized flat flame and a
flame. They attributed it to the rapid deposition of nonstabilized laminar flame. These fields are created
particles on the electrodes, thus decreasing the particle either by needle electrodes that introduce ions in the
residence time in the high temperature region of the flame or by plate electrodes that merely attract flame-
flame. Likewise, Katz and Hung5 showed that the size of generated ions. Three materials with distinctly different
TiO2 , SiO2 , and GeO2 particles made in a similar reactor sintering rates and applications, TiO2 , SiO2 , and SnO2 ,
was greatly influenced by the presence of electric fields. are produced here. Titania is used as a pigmentary
Xiong et al.6 showed theoretically that charging titania material,9 photocatalyst,10 and as a catalyst support.11
Fumed silica particles are widely used for optical fibers,
catalyst supports, and as fillers and dispersing agents.12
a)Currently at Lucent Technologies Inc., Murray Hill, New Jersey Nanosized tin oxide powders are used as semiconductors
07974. and gas sensors.13 The objective of the present study
b)
Author to whom correspondence should be addressed.
c)
Currently at Protein Technologies International, St. Louis, Missouri is to explore the potential of electric fields for flame
63164. synthesis of nanophase materials with closely controlled

J. Mater. Res., Vol. 12, No. 4, Apr 1997  1997 Materials Research Society 1031

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

characteristics. The effects of material properties, electric precursors. The particles are collected on glass fiber
field intensity, electrode geometry, and position along filters (Gleman Scientific) for subsequent analysis in an
the flame axis are investigated. open-faced filter (Graseby/Anderson) kept 10.5 cm from
the burner face, by the aid of a vacuum pump.
II. EXPERIMENTAL Stainless steel, needle-like electrodes with sharp tips
(ø100 mm) or plate electrodes [3.8 3 2.5 3 0.38 cm
Figure 1 shows a schematic of the experimental (long plates) or 2.5 3 2.5 3 0.38 cm (short plates)] are
setup. A laminar premixed burner-stabilized flame re- used to create the electric field across the flame.7 One
actor is used to make titania, tin oxide, and silica electrode is connected to the DC power supply (Gamma
particles in the presence of externally controlled electric High Voltage Research, Inc.; Spellman) with reversible
fields. The flame is stabilized on the burner mouth polarity and the second one to the ground. For the
using a 2 cm long mullite monolith (Corning) honey- powders made in the burner-stabilized, premixed, flat
comb with 48 openingsycm2 . The burner [alumina tube flame, the distance between the electrode tips is 5 cm,
(Coors), 1.875 cm ID] is packed to 3y4 of its volume while in the nonstabilized flame, it is 4 cm. The plate
with glass beads (6 mm in diameter) supported on a electrodes are arranged in such a way that the bottom
screen to provide good mixing of the reactants before edge of the plate is in the same plane as the burner face.
they enter the honeycomb. The advantage of using a The specific surface area of the powder, A, is meas-
burner-stabilized, premixed, flat flame is that most of ured by nitrogen adsorption at 77 K (Gemini 2360,
the particles experience similar temperatures and gas Micromeritics) using the BET equation. The average
velocities across the flame.14 The above powders are primary particle size, dp , is obtained from dp ­ 6yrp A,
also made in a nonstabilized, laminar flame reactor. This where rp is the particle density. The particle morphology
reactor is also an alumina tube (Coors, 1.25 cm ID), but is obtained by transmission electron microscopy (TEM)
without the monolith. on Philips CM 20 or Philips EM 400 microscopes operat-
Clean, dry argon gas (Wright Brothers, 99.8%) is ing at 100 kV. The TEM primary particle size is obtained
bubbled into a gas-washing bottle containing TiCl4 by visually counting at least forty primary particles.15
(Aldrich, 99.9%), SiCl4 (Aldrich, 99%), or SnCl4 The weight fractions of the anatase and rutile phases
(Aldrich, 99%), premixed with particle-free nitrogen in the titania samples are obtained by x-ray diffraction7
(Wright Brothers, 99.8%), oxygen (Matheson, 99.9%), (D500, Siemens; using Cu Ka radiation). The XRD crys-
and methane (Wright Brothers, 99.8%) and sent through tallite size was not determined since this estimate may be
the burner. A check valve is used before the precursor is inconclusive for powders with large particle size differ-
mixed with the mixture of air and methane to prevent any ences.16 For flame-generated powders it has been shown
back flow of the flame. In the case of the nonstabilized that XRD crystallite sizes are consistently smaller than
flame reactor, air is used as the oxidant. Oxide particles the BET determined particle sizes.15 A 0.038 cm gauge
are formed in the flame by oxidation/hydrolysis of the Pt–Rh thermocouple (Omega Engineering) insulated
with a mullite sheath is used for measuring the flame
temperature and is corrected for radiation losses.17

III. RESULTS AND DISCUSSION


A. Flame synthesis of oxides in the
absence of electric field
First, titania, tin oxide, and silica powders are made
in both flames in the absence of electric fields. At the
burner-stabilized flat flame reactor, the room temperature
flow rates of CH4 , N2 , and O2 streams are 445, 3250,
and 1300 cm3ymin, respectively, while 250 cm3ymin of
argon are bubbled through TiCl4 and SnCl4 . The SiCl4
bubbler is kept in ice (at 0 ±C) to reduce its vapor
pressure, and only 200 cm3ymin of argon are bubbled
through SiCl4 . To keep the total gas flow rate the
same in all experiments, the N2 flow rate is set to
3300 cm3ymin during silica synthesis. Thus, the TiCl4 ,
SnCl4 , and SiCl4 concentrations leaving each bubbler
are 1.75 3 10–4 , 2.9 3 10–4 , and 8 3 10–4 molymin,
FIG. 1. Schematic of the experimental setup. respectively. The specific surface areas of products TiO2 ,

1032 J. Mater. Res., Vol. 12, No. 4, Apr 1997

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

SnO2 , and SiO2 are 50, 20, and 148 m2yg, respec- burner (5205 cm3ymin) results in less gas dilution and
tively. The differences in the specific surface area of longer exposure at high temperatures.8 Consequently,
these oxides synthesized under similar conditions stem particles made in the nonstabilized, laminar flame are
from their different sintering rates.18 Tin oxide has low sintered more effectively having lower specific surface
melting point (1127 ±C) and sinters by evaporation- area than those made in the flat flame. This is supported
condensation,19 while titania sinters by grain boundary also by XRD of titania powders that were mostly rutile
diffusion20 and silica sinters by viscous flow.21 when they were made in the nonstabilized flame while
When a nonstabilized flame is used, the flow rates they were exclusively anatase (rutile-free) in the flat
of the CH4 and air streams are 237 and 1100 cm3ymin, flame. Rutile is the thermodynamically stable form of
respectively, while 200 cm3ymin of argon are bubbled TiO2 , while anatase and brookite are metastable at all
through the precursor. The concentrations of TiCl4 , temperatures and transform to rutile upon heating.22
SnCl4 , and SiCl4 leaving the bubbler are 1.4 3 10–4 ,
2.3 3 10–4 , and 8 3 10–4 molymin, respectively. The B. Flame characteristics in the
specific surface areas of product TiO2 , SnO2 , and presence of electric fields
SiO2 powders are 16, 13, and 130 m2yg, respectively. Figure 2 shows pictures of the burner-stabilized
The lower overall flow rate through the latter burner flame during synthesis of SnO2 particles at various
(1537 cm3ymin) compared to that through the flat flame positive field strengths and electrode configurations.

FIG. 2. (a – f) Burner-stabilized, premixed, laminar flames producing SnO2 particles under the influence of various electric fields created by needle
(b –d) and plate (e, f) electrodes. Gas flow rates: CH4 ­ 445, O2 ­ 1300, N2 ­ 3250, and Ar ­ 250 cm3ymin; SnCl4 ­ 2.9 3 10–4 molymin.

J. Mater. Res., Vol. 12, No. 4, Apr 1997 1033

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

Figure 2(a) shows this flame in the absence of any strength extinguishes the flame. The latter property is
electric field. The needle electrodes are kept at 0.1 cm used for extinction of pool flames.26 It should be noted
above the burner face [Figs. 2(b)–2(d)]. When an that the maximum (“breakdown”) field strength that can
electric field of 1.4 kVycm is applied, the flame is visibly be used is limited by sparking between electrodes, which
affected by the ensuing corona discharge [Fig. 2(b)]. is a function of electrode tip diameter, gas composition,
At low field strengths, the flame is tilted toward the and electrode distance. For the employed electrodes and
negative electrode (cathode). The established current SnO2 aerosol, this limit was 2.75 kVycm, which is of
across the electrodes assures equal movement of positive little significance here since the flame is extinguished by
and negative charges; the positive ions must equal the ionic wind at lower field intensities. However, this is
the negative ions and electrons. As the positive ions an important limit for burner-stabilized flames that are
are larger than the electrons, the motion of positive not extinguished by ionic wind or flames that are not
ions “drags” neutral gas molecules, creating a net gas affected by burner deposits, as it was the case with the
movement toward the cathode.14,23 Further increasing flames forming silica and titania.
the field strength up to 2.2 kVycm drastically reduces Figure 4 shows the temperature along the axis of
the flame height [Figs. 2(c) and 2(d)]. At the highest the burner-stabilized, flat flame in the absence of any
field strength, the flame front is nearly on the burner precursor, but with the carrier gas argon flowing through
face, resulting in substantial particle deposition on the burner along with CH4 , N2 , and O2 . Precursor gases
the monolith. This represents the upper limit of the were not used to avoid particle deposition on the tip of
employed field strength. The reduction of the flame the thermocouple.27 In the absence of an electric field,
height is attributed to the ionic wind (also known as the maximum temperature is about 1110 ±C, slightly
the Chattock wind or corona wind) flowing across the above the luminous zone of the flame (open circles). The
flame.24,25 When a corona discharge is created across a temperature is measured 0.5 cm away from the flame
flame, ions flow from the discharging electrode toward center up to 4 cm above the burner face. The measured
the ground electrode. The flow of ions increases with temperature across the burner face in a horizontal plane
increasing field strength. As these ions move, convection is almost the same (±20 ±C) as that measured at the
is created across the flame, reducing the flame height.7 center. The temperature quickly drops away from the
Figure 2 also shows pictures of the burner-stabilized maximum along the flame axis. Figure 4 also shows
flame producing SnO2 particles under the electric field measured flame temperatures with 21.6 kVycm applied
created by long plate electrodes [Figs. 2(e) and 2(f)]. In across the flame using the needle electrodes (filled cir-
this configuration the flame ions are attracted toward the cles). So as not to disturb the field across the flame
plate electrodes. Under a field intensity of 1 kVycm, the with the thermocouple, temperatures are measured 5 cm
flame height is similar to that in the absence of fields, above the electrodes and higher. In the presence of
with a slight inclination of the flame toward both cathode the corona discharge, the downstream temperature is
and anode. As the field intensity is increased to 2 kVycm, about a hundred degrees lower than without the field.
the flame is pulled toward the electrodes and its height Figure 4 also shows flat flame temperatures measured in
further decreases. Literally, flame 2(a) can be viewed as a the presence of a negative electric field (21.6 kVycm)
flower bud that opens up [Figs. 2(e) and 2(f)] as the field created by long plate electrodes (filled diamonds). The
intensity increases and plate electrodes attract more and temperature downstream of the flame is not influenced by
more flame ions. When plates are used as the electrodes, the electric field created by the plate electrodes, implying
the ionic wind is reduced and the flame is substantially that the ionic wind with this configuration does not have
more stable than when needle electrodes are used. as strong an impact as with the needles.
Figure 3 now shows pictures of the nonstabilized Figure 4 also shows the temperature profile of the
flame producing SnO2 particles at various positive field nonstabilized flame with a maximum temperature of
intensities created across the flame by needle electrodes 1000 ±C at the tip of the flame (open squares). Also
at 1 cm above the burner face. Since the gas flow rates in shown are the measured temperatures of the nonstabi-
this flame are lower, the onset of the ionic wind appears lized flame under an electric field created by the needle
at a slightly lower field strength (1.2 kVycm). Again, as electrodes of 11.75 Vycm (filled squares). The electric
the field strength increases, the flame decreases in height field has a profound effect on the flame temperature. The
and leans toward the cathode [Figs. 3(b) –3(e)]. The ensuing ionic wind across the flame dissipates heat by
ionic wind has a stronger impact on this flame since the convection.7 This reduces the measured temperature by
gas velocity through the burner is lower and nonuniform as much as 200 ±C away from the burner face.
compared to that of the burner-stabilized flat flame. As a
matter of fact, at a field strength of 2.5 kVycm, the flame C. Synthesis of titania powders
is barely holding onto the burner [Fig. 3(f)] as the flame Figure 5(a) shows the specific surface area of TiO2
is vigorously oscillating, and a slight increase in the field particles as a function of the applied positive or negative

1034 J. Mater. Res., Vol. 12, No. 4, Apr 1997

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

FIG. 3. Nonstabilized, laminar flames producing SnO2 particles under the influence of various electric fields created by needle electrodes. Gas
flow rates CH4 ­ 237, air ­ 1100, Ar ­ 200 cm3ymin; SnCl4 : 2.3 3 10–4 molymin.

field strength created by the needle electrodes across of electrons is 1000 times greater,23 resulting in more
the burner-stabilized, flat flame. In the absence of elec- efficient particle charging.30 Furthermore, the ionic wind
tric field, the specific surface area of titania particles is sets in at lower field strengths with negative than with
50 m2yg and increases to about 85 m2yg at 21.8 kVycm positive polarity because of the greater mobility of nega-
and to 75 m2yg at 11.8 kVycm. The specific surface tive ions. As a result, particle growth by coagulation and
area increases or the primary particle size decreases with coalescence is retarded more effectively with negative
increasing applied potential across the flame, regardless than with positive coronas. The small reduction in the
of the polarity of the electric field. When a corona dis- specific surface area at the highest applied electric field
charge is applied, the temperature of the flame decreases is attributed to the proximity to the “breakdown” field
by the ionic wind (Fig. 4) which decreases the particle intensity (1.8 kVycm) that destabilizes this flame and
sintering rate, resulting in smaller primary particles or the onset of particle deposition on the burner face. The
powders with high specific surface area. Also, the parti- monotonic increase of specific surface area between 0.8
cles are charged unipolarly, causing them to repel each and 1.7 kVycm shows that electric fields can be used to
other, resulting in reduced coagulation rates. Further- control the TiO2 primary particle size from 30 to 18 nm
more, electrostatic dispersion of unipolarly charged parti- at these conditions.
cles reduces their concentration,28 slowing down further Needle electrodes introduce new ions in the flame
coagulation and, subsequently, particle growth. by corona discharge. An alternative is to create an
The negative electric field is more efficient than electric field using the ions generated by the flame
the positive one in reducing particle growth since it itself. Thus, needle electrodes are replaced by plate
results in smaller particles. A similar effect was observed electrodes. As shown in Figs. 2 and 4, the ionic wind
in diffusion flames.7 The velocity of negative ions is is reduced substantially when plate electrodes create
20 –50% greater29 than that of positive ions, while that the electric field. Figure 5(b) shows the specific surface

J. Mater. Res., Vol. 12, No. 4, Apr 1997 1035

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

particle size of 55 nm. At lower field strengths, the


position of the needle electrodes along the axis of the
flame does not influence the particle specific surface area.
However, at an applied potential of 12.75 kVycm, TiO2
particles have a specific surface area of 21 m2yg when
the electrodes are at 1 cm from the burner face compared
to 27 m2yg when the electrodes are at 0.1 cm. Creating
the electric field close to the particle formation region
(near the burner face) results in convection and particle
charging at earlier stages of their inception, slowing
down, thus, particle growth by coagulation.
Figure 6(a) shows a micrograph of titania particles
synthesized in the nonstabilized, laminar flame in the
absence of an electric field corresponding to the data
in Fig. 5(c). The primary particles are polyhedral, with
very well-defined edges. The primary particle size varies
from about 20 to 100 nm, with an average primary
particle size of about 75 nm. At a field strength of
12 kVycm, with electrodes places at 0.1 cm, above the
FIG. 4. Axial temperature profiles of the burner-stabilized (circles)
burner face, the primary particle size decreases to about
and nonstabilized (squares) premixed flames in the absence (open 60 nm [Fig. 6(b)] and to about 45 nm at a field strength
symbols) and presence (filled symbols) of electric fields. of 12.25 kVycm [Fig. 6(c)]. The primary particles still
retain their polyhedral shape. Increasing the field strength
area of titania particles as a function of applied field to 12.75 kVycm decreases the primary particle size
strength across the burner-stabilized, flat flame, for pos- to about 40 nm, and the variation of the particle size
itive (circles) and negative (squares) electric fields using decreases [Fig. 6(d)]. The decreasing TEM primary par-
short (open symbols) and long (filled symbols) plates. ticle size with field intensity is consistent with the BET
At a positive field strength of 1.6 kVycm, the specific data [Fig. 5(c)]. The necking between primary particles
surface area of titania particles is about 85 m2 gyg and results in larger average primary particles by BET than
70 m2yg using the long and short plates, respectively. TEM or XRD as is commonly observed with flame-made
The fact that the flame height and temperature do not particles.3,15,26
decrease substantially downstream of the flame (Figs. 2 The titania particles made in the burner-stablized,
and 4) indicates that the observed increase in the particle premixed, flat flame are .99.9 wt. % anatase, and the
specific surface area of titania particles can be attributed phase composition of these powders does not change
largely to electrostatics rather than convection. as the field strength across the flame is increased. On
At a given field strength, the increase in specific the other hand, TiO2 powders made in the nonstabilized,
surface area of titania particles is larger with long laminar flame in the absence of an electric field are
electrodes (115 m2yg) than with short ones (85 m2yg) 82 wt. % rutile and 18 wt. % anatase. Figure 7 shows the
at 21.6 kVycm. By exposing the particles to an electric rutile content of TiO2 powders made in the nonstabilized
field for longer residence times (long plates), particle flame as a function of the applied positive field strength
collisions and growth are further inhibited. However, fur- across the flame. The rutile content of TiO2 decreases
ther downstream, the temperatures decrease drastically down to 65 wt. % with increasing field strength across
(Fig. 4), and no further coalescence and sintering of the the flame, as observed in our earlier diffusion flame
particles takes place. studies,7 though, here, powders with much higher rutile
Figure 5(c) shows the specific surface area of the content are obtained: 82 vs 20 wt. %. Of course, this is
TiO2 particles made in the nonstabilized, laminar flame attributed to the different flame structures: coflow diffu-
as a function of the applied positive field strength for sion flame versus premixed flame. As the field strength
two needle-electrode positions along the flame axis. across the flame is increased, the flame height and the
In the absence of any electric field, TiO2 particles particle residence time in the high temperature region
made at these conditions have a specific surface area of the flame are reduced, inhibiting the anatase to rutile
of 16 m2yg, corresponding to primary particle size of phase transformation.31
90 nm. The specific surface area is about 18 m2yg at
11.5 kVycm with the needles positioned at 0.1 cm from D. Synthesis of silica powders
the burner face and further increases to 27 m2yg at Figure 8(a) shows the specific surface area of SiO2 par-
12.75 kVycm, corresponding to an average primary ticles made in the burner-stabilized, premixed flat flame

1036 J. Mater. Res., Vol. 12, No. 4, Apr 1997

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

(a) (b)

(c)
FIG. 5. Specific surface area of TiO2 particles as a function of the applied field strength across the (a) burner-stabilized, premixed flame (needle
electrodes), (b) burner-stabilized, premixed flame (long and short plate electrodes), and (c) nonstabilized, laminar flame (needle electrodes).

as a function of applied field strength. The specific sur- in the presence of 1.6 kVycm negative and positive elec-
face area of SiO2 particles increases from 148 m2yg in tric fields, respectively, across the flat flame. As with
the absence of an electric field to about 240 and 220 m2yg titania, the negative electric field is more influential at

J. Mater. Res., Vol. 12, No. 4, Apr 1997 1037

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

FIG. 6. TEM micrographs of TiO2 particles made in a nonstabilized, laminar flame as a function of the applied positive field strength, with the
needle electrodes kept at 0.1 cm from the burner face: (a) No electric field, (b) 12 kVycm, (c) 12.25 kVycm, and (d) 12.75 kVycm.

increasing the specific surface area. Figure 8(a) typi-


fies the capacity of electric fields to facilitate precise
control of the average primary particle size from 20
to 12 nm.
Figure 8(b) shows the specific surface area of SiO2
particles made in the burner-stabilized, flat flame as
a function of applied field strength for positive and
negative polarities, using both short (open symbols) and
long (filled symbols) plate electrodes. As with titania,
the specific surface area of silica particles increases
with increasing field strength, with the negative electric
field again being most effective in arresting the particle
growth. Creating the electric field using longer plate elec-
trodes results in a higher specific surface area at the same
applied field strength. As for TiO2 , extending the length
of the electric field downstream of the flame further
influences the particle growth. TEM analysis of silica
particles reveals that the agglomerate size decreases with
increasing field strength across the flame.32
Figure 8(c) shows the specific surface area of SiO2
particles made in the nonstabilized, laminar flame as a
FIG. 7. Rutile content of TiO2 particles made in a nonstabilized, function of the applied positive field strength for two
laminar flame as a function of the applied positive field strength for different locations of the electrodes along the axis of the
two locations of the needle electrodes along the flame axis. flame. Silica particles made in the absence of an electric

1038 J. Mater. Res., Vol. 12, No. 4, Apr 1997

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

(a) (b)

(c)
FIG. 8. Specific surface area of SiO2 particles as a function of the applied positive and negative field strength across the (a) burner-stabilized,
premixed flame (needle electrodes), (b) burner-stabilized, premixed flame (long and short plate electrodes), and (c) nonstabilized, laminar
flame (needle electrodes).

field have a specific surface area of 130 m2yg. With an 0.1 and 1 cm, respectively, away from the burner face.
applied field strength of 2 kVycm, this increases to about Again, the position of the electrodes affects the product
230 and to 200 m2yg when the electrodes are placed at particle size as with TiO2 particles.

J. Mater. Res., Vol. 12, No. 4, Apr 1997 1039

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

(a)

FIG. 9. Specific surface area of SnO2 particles as a function of the


applied field strength across the nonstabilized, laminar flame. The
electric field is created by the needle electrodes.

E. Synthesis of tin oxide powders


In contrast to the results of TiO2 and SiO2 , in the
case of the burner-stabilized, premixed, flat flame, the
specific surface area of SnO2 particles did not increase
and remained at about 20 m2yg at all field strengths.
This is observed when the electric field is created by
both needle and plate electrodes. As mentioned earlier,
tin oxide sinters much faster than titania and silica. The (b)
fact that the specific surface area of SnO2 particles did
FIG. 10. TEM micrographs of SnO2 particles made in a nonstabilized,
not change within experimental error indicates that the laminar flame as a function of the applied field strength, with the
sintering of tin oxide particle is completed at a very needle electrodes kept at 0.1 cm from the burner face: (a) No electric
early stage in the flame. As the grain size increases, the field and (b) 12.25 kVycm.
sintering rates decrease and grain growth is very slow.33
In the presence of electric fields particles are charged.
However, the high sintering rates of tin oxide dominate across the flame, and the specific surface area of SnO2
the electrostatic forces, and hence, no change in particle particles does not change when they are synthesized in
characteristics is observed. a burner-stabilized premixed flame. As the electrodes
On the other hand, in the case of the nonstabilized, are moved further up to 1 cm from the face of the
laminar flame, the specific surface area of SnO2 particles nonstabilized, laminar flame, the effect of the electric
increased from 13 to 24 m2yg at a positive field strength field diminishes. Figure 9 shows that the SnO2 particle
of 12.5 kVycm with the needles kept at 0.1 cm from specific surface area increases from 13 to 17 m2yg only
the burner face (Fig. 9). Particles made in this flame at 12.5 kVycm field strength when the electrodes are
experience a broader spectrum of temperatures than placed at 1 cm above the burner face, while it increases
in the burner-stabilized flame. As a result, the ionic to 24 m2yg when they are just 0.1 cm above the burner.
wind more drastically reduces the particle residence time Figure 10(a) shows the TEM micrograph of SnO2
at high temperatures in the nonstabilized than in the particles made in the nonstabilized, laminar flame in the
burner-stabilized flame (Fig. 4). The high temperature absence of any electric field across the flame. The pri-
encountered by the particles during the early stages of mary particles are polyhedral, with well-defined edges.
particle growth dominates the influence of the ionic wind The primary particle size ranges from 50 to 200 nm,

1040 J. Mater. Res., Vol. 12, No. 4, Apr 1997

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

FIG. 11. TEM micrographs of SnO2 agglomerates made in a nonstabilized, laminar flame as a function of the applied field strength, with the
needle electrodes kept at 0.1 cm from the burner face: (a) No electric field, (b) 12 kVycm, (c) 12.25 kVycm, and (d) 12.75 kVycm.

with an average of about 100 nm. When a positive in smaller agglomerates with increasing field strength
electric field of 12.25 kVycm is applied across the flame across the flame. Even though the primary particle size
by electrodes kept at 0.1 cm above the burner face, of SnO2 agglomerates does not change substantially with
the average primary particle size decreases to about increasing field strength across the flame, the agglom-
65 nm, but with a wide size variation from 10 to 200 nm erate size decreases. The high sintering rates of SnO2
[Fig. 10(b)]. This is in agreement with the BET results makes the length scales at which electrostatics become
(Fig. 9). The broad size distributions are attributed to dominant comparable to the size range of agglomerates.
the various temperature histories of particles made in Similar trends of decreasing agglomerate size are ob-
the nonstabilized, laminar flame. served for SnO2 particles made in the burner-stabilized,
Figure 11(a) shows typical agglomerates of SnO2 premixed flame.32
particles made in the nonstabilized, laminar flame in
the absence of any electric field. The particles are
highly agglomerated. The agglomerates are chain-like IV. CONCLUSIONS
structures, again with a wide range of primary particles. The effect of electric fields on the flame synthesis
When an electric field of strength 12.0 kVycm is applied of nanosized titania, silica, and tin oxide particles was
across the flame using needle electrodes kept at 0.1 cm investigated in two flames: a burner-stabilized premixed,
above the burner face, the extent of agglomeration of flat flame and a nonstabilized, laminar flame. Negative
SnO2 particles decreases [Fig. 11(b)]. The extent of electric fields are most effective in affecting the product
agglomeration further decreases as the field strength powder characteristics. When the electric field is created
across the flame is increased to 12.25 and 12.75 kVycm by needle electrodes, ionic wind is generated across
[Figs. 11(c) and 11(d)]. the flame, decreasing the flame height and resulting in
The ionic wind across the flame decreases the parti- shorter particle residence times at high temperatures.
cle residence time and temperature of the flame, resulting Furthermore, charging of the particles results in elec-
in decreased coagulation rates. Furthermore, electrostatic trostatic repulsion and dispersion. These two effects
repulsion and dispersion contribute to a further de- result in reduced coagulation rates. The position of
crease in the coagulation rates of the particles, resulting electrodes along the flame axis has a strong influence

J. Mater. Res., Vol. 12, No. 4, Apr 1997 1041

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194


S. Vemury et al.: Electrically controlled flame synthesis of nanophase TiO2 , SiO2 , and SnO2 powders

on the particle characteristics. Placing the electrodes 3. G. D. Ulrich, C&EN, 62(8), 22 (1984).
closer to the particle inception region (near the burner 4. D. R. Hardesty and F. J. Weinberg, Fourteenth Symposium (Inter-
national) on Combustion (The Combustion Institute, Pittsburgh,
face) results in reduced particle growth. Replacing the PA, 1973), p. 1365.
needle with plate electrodes substantially reduced the 5. J. K. Katz and C-H. Hung, Twenty-Third Symposium (Interna-
ionic wind making electrostatics the dominant reason tional) on Combustion (The Combustion Institute, Pittsburgh, PA,
for reduced particle growth. Increasing the plate length 1990), p. 1733.
reduced particle size since the particles were exposed to 6. Y. Xiong, S. E. Pratsinis, and S. V. R. Mastrangelo, J. Colloid
Interface Sci. 153, 106 (1992).
electric fields for increased residence times. 7. S. Vemury and S. E. Pratsinis, Appl. Phys. Lett. 66, 3275 (1995).
The specific surface area of TiO2 and SiO2 increased 8. S. E. Pratsnis, W. Zhu, and S. Vemury, Powder Technol. 86, 87
with increasing field strength in both flames. On the other (1996).
hand, the specific surface area of SnO2 powders was not 9. E. J. Mezey, in Vapor Deposition, edited by C. F. Powell, J. H.
affected when they were made in the burner-stabilized, Oxley, and J. M. Blocher, Jr. (John Wiley & Sons, New York,
1966), p. 423.
premixed, flat flame, while it increased with increasing 10. D. F. Ollis, E. Pelizzetti, and N. Serpone, Environ. Sci. Technol.
field strength when they were made in the nonstabilized, 25, 1523 (1991).
laminar flame. The difference in the temperature history 11. M. R. Bankmann, R. Brand, B. H. Engler, and J. Ohmer, Catal.
of the particles in these two flames is the reason for Today 14, 225 (1992).
this differing behavior. On the other hand, the sintering 12. J. R. Bautista and R. M. Atkins, J. Aerosol Sci. 22, 667 (1991).
13. E. U-K. Kim and I. Yasui, J. Mater. Sci. 23, 637 (1988).
rates of TiO2 and SiO2 are slow enough for charging 14. R. Vijayakumar and K. T. Whitby, Aerosol Sci. Technol. 3, 17
to influence the primary particle size. The agglomerate (1984).
size of SnO2 particles is decreased in the presence of the 15. S. Vemury and S. E. Pratsinis, J. Am. Ceram. Soc. 78, 2984 – 2992
electric field, even though the effect of electrostatics on (1995).
the primary particle size is not substantial. The length 16. A. Kobata, K. Kusakabe, and S. Morooka, AlChE J. 37, 347 –359
(1991).
scale at which electrostatics influence SnO2 particles 17. S. Vemury, Flame Synthesis of Nanoparticles: Effect of Charging,
is on the agglomerate size rather than the primary Ph.D. Thesis, University of Cincinnati (1996).
particle size. 18. T. Matsoukas and S. K. Friedlander, J. Colloid Interface Sci. 146,
In conclusion, external electric fields can greatly 495 (1991).
facilitate synthesis of ceramic and possibly other par- 19. P. G. Harrison and M. J. Willett, J. Chem. Soc., Faraday Trans.
1, 85 1921 – 1932 (1989).
ticles with precisely controlled specific surface area, 20. M. Astier and P. Vergnon, J. Solid State Chem. 19, 67 (1976).
crystallinity, and agglomerate structure. Given the cur- 21. W. D. Kingery, H. K. Bowen, and D. R. Uhlmann, Introduction to
rent low cost of flame aerosol technology, this tech- Ceramics (Wiley-Interscience, New York, 1976).
nique has strong potential for large scale manufacture of 22. R. D. Shannon and J. A. Pask, J. Am. Ceram. Soc. 48, 391 (1965).
nanoparticles. 23. K. G. Payne and F. J. Weinberg, Proc. R. Soc. A250, 316 (1959).
24. D. Bradley and M. A. M. Jamel, Comb. Flame 58, 115 (1984).
25. D. Bradley, in Advanced Combustion Processes, edited by F. J.
ACKNOWLEDGMENTS Weinberg (Academic Press, Orlando, FL, 1986), p. 331.
We acknowledge support from the National Sci- 26. E. Sher, G. Pinhasi, A. Pokryvailo, and R. Bar-on, Combustion
Flame 94, 244 (1993).
ence Foundation, Presidential Young Investigator Award 27. G. P. Fotou, S. E. Pratsinis, and P. A. Baron, Chem. Eng. Sci. 49,
(S.E.P.), Grant CTS-8957042, and recently, Grant CTS- 1651 (1994).
9612107. 28. G. Kasper, J. Colloid Interface Sci. 81, 32 (1981).
29. M. Adachi, Y. Kousaka, and K. Okuyama, J. Aerosol Sci. 16,
REFERENCES 109 (1985).
30. A. Wiedensohler, J. Aerosol Sci. 19, 387 (1988).
1. R. P. Andres, R. S. Averback, W. L. Brown, L. E. Brus, W. A. 31. M. K. Akhtar, Y. Xiong, and S. E. Pratsinis, AlChE J. 37, 1561 –
Goddard III, A. Kaldor, S. G. Louie, M. Moscovits, P. S. Peercy, 1570 (1991).
S. J. Riely, R. W. Siegel, F. Spaepen, and Y. Wang, J. Mater. 32. S. Vemury and S. E. Pratsinis, J. Aerosol Sci. 27, 951 (1996).
Res. 4, 704 (1989). 33. W. Koch and S. K. Friedlander, J. Colloid Interface Sci. 140, 419
2. N. Ichinose, Y. Ozaki, and S. Kashu, Superfine Particle Technol- (1990).
ogy (Springer-Verlag, London, 1992).

1042 J. Mater. Res., Vol. 12, No. 4, Apr 1997

http://journals.cambridge.org Downloaded: 14 Aug 2014 IP address: 128.120.194.194

Das könnte Ihnen auch gefallen