Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Full Paper

CVD of Al
2
O
3
Thin Films Using Aluminum Tri-isopropoxide**
By Sabine Blittersdorf, Naoufal Bahlawane, Katharina Kohse-Hinghaus,* Burak Atakan, and Jrgen Mller
A stagnation point cold-wall reactor was used for the CVD of corundum alumina (a-Al
2
O
3
) on metallic substrates. Deposi-
tions were carried out under low pressure using the thermally induced pyrolytic oxidation of aluminum tri-isopropoxide
(ATI). The effects of the substrate temperature (3001080 C) and the total pressure (50250 mbar) on the growth rate and
morphology of the deposits were investigated. An excess of oxygen facilitates the formation of dense alumina films. Precursor
depletion was prevented using high gas velocity, low ATI concentration, and a high temperature gradient. X-ray diffraction
(XRD) analysis provided evidence of corundum alumina deposition at substrate temperatures above 1000 C.
Keywords: a-Al
2
O
3
, Alternative precursors, Aluminum tri-isopropoxide
1. Introduction
Alumina is a material of choice wherever hardness, wear
resistance, and thermal and chemical stability are desired.
By the application of alumina coatings, these properties
can, in principle, be used to enhance the performance, life-
time, or field of application of many materials. Alumina
exists in several metastable crystalline phases, however, all
of them transform to the stable corundum (a-Al
2
O
3
) phase
at high temperatures. The continued interest in corundum
alumina is based on its high performance in wear and
corrosion resistance,
[13]
its high temperature insulating
properties,
[4]
and its diffusion barrier properties.
[5,6]
Thin
films of a-Al
2
O
3
are deposited using wet processes such as
solgel,
[2,3]
spray coating,
[7]
or physical vapor deposi-
tion,
[1,8,9]
the thermal CVD process being the most often
used. The classic way of producing alumina thin films by
CVD is based on the hydrolysis of AlCl
3
with a mixture of
H
2
and CO
2
gases at temperatures between 700 C and
1000 C.
[1014]
The main disadvantage of this process is the
corrosive by-product (HCl), which is suspected of facilitat-
ing whisker growth, instead of dense layers, on nickel-base
superalloys.
[6]
Alternative precursors have been investigated, in particu-
lar aluminum alkoxides which undergo a pyrolytic decom-
position forming alumina even at low temperatures, and in
the absence of an oxidizing gas.
[10]
Efforts have been made
to synthesize highly volatile and non-pyrophoric liquid pre-
cursors to deposit dense alumina films free from car-
bon.
[1517]
The potential of ATI to produce a suitable film
structure with high yield was noted,
[1719]
but, these investi-
gations
[1520]
concentrated on the low temperature process
resulting in amorphous alumina thin films. The deposition
of crystalline alumina requires the use of high temperatures,
however homogeneous reaction of the reactive intermedi-
ates may then result in undesired gas-phase nucleation. This
limitation was overcome using inductive heating, which of-
fers a high temperature gradient around the substrate.
[21,22]
Another approach for selective heating of the substrate was
adopted by Tago et al.
[23]
using an infrared (IR) furnace in a
quartz reactor. Pauer et al.
[21]
investigated the high temper-
ature (above 900 C) pyrolysis of various aluminum alkox-
ides (including ATI) with no further oxygen source. This
process leads to the deposition of non-closed a-Al
2
O
3
films
with a dome-like structure. Niska et al.
[22]
attained a closed
a-Al
2
O
3
structure using the high temperature pyrolysis of
ATI in an oxidizing atmosphere, under reduced pressure.
Well-adherent and fine-grained films with high density
were achieved by maintaining low deposition rates. The
limitation of the inductively-heated process is caused by a
thermal convection effect that induces an increase in the
deposition rate at the substrate edges. A similar dome-like
structure was obtained by Tago et al.
[23]
with the pyrolysis
of ATI in a non-oxidizing atmosphere using a vertical reac-
tor. They have performed a systematic investigation into
such deposition control parameters as the precursor con-
centration, the substrate position from the inlet gas, and the
substrate temperature. They have also examined the possi-
194 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/cvde.200306248 Chem. Vap. Deposition 2003, 9, No. 4

[*] Prof. K. Kohse-Hinghaus, S. Blittersdorf, Dr. N. Bahlawane


Physikalische Chemie I
Fakultt fr Chemie, Universitt Bielefeld
Universittsstr. 25, D-33615 Bielefeld (Germany)
E-mail: kkh@pc1.uni-bielefeld.de
Prof. B. Atakan
Thermodynamik, Institut fr Verbrennung und Gasdynamik
Gerhard-Mercator-Universitt Duisburg
Lotharstr. 1, D-47057 Duisburg (Germany)
J. Mller
Lehrstuhl fr Werkstoffchemie
Rheinisch-Westflische Hochschule Aachen
D-52056 Aachen (Germany)
[**] This work was performed within the DFG research structure Schwer-
punktprogramm SPP 1119 Anorganische Materialien durch Gasphasen-
synthese: Interdisziplinre Anstze zu Entwicklung, Verstndnis und
Kontrolle von CVD-Verfahren and funding by the DFG under contract
KO 1363/13-1 and NE 351/30-1 is gratefully acknowledged. N.B. is grate-
ful for a fellowship by the Alexander von Humboldt Foundation. We
thank I. Hattesohl (Physikalische Chemie I) and T. Tak (Anorganische
Chemie I) for their assistance with the surface analysis.
Full Paper
bility of further decreasing gas-phase reactions via reduc-
tion of the substrate surface area compared to the reactor
diameter, however, neither the film crystalline structure
nor the carbon content was provided.
With regard to the decomposition mechanism of ATI,
Pauer et al.
[21]
assumed that the first step involves the hy-
drolysis of ATI. The complete hydrolysis leading to
Al(OH)
3
takes place only when the residence time is long
enough, otherwise intermediate molecules or radicals
formed during the decomposition may absorb onto the sur-
face and integrate into the growing Al
2
O
3
interface by the
elimination of the excess oxygen. The necessary water for
the hydrolysis is made available by the further pyrolytic de-
composition of the precursor. Haanappel et al.
[24]
assumed
that ATI decomposes, to a minor extent, by a free radical
or an a-hydride elimination mechanism and suggested b-
hydride elimination as the most plausible mechanism.
Their investigations show that ATI undergoes the elimina-
tion of isopropanol and propene by b-hydride elimination
at 200 C, followed by a second pyrolysis at 280 C forming
AlO(OH) molecules which absorb onto the growing film
interface. Considering these observations and the film mi-
crostructures,
[21,22]
it is assumed that the presence of oxygen
in the atmosphere induces the oxidation of hydrocarbon
groups and causes their depletion to species which do not
incorporate into the film (CO and CO
2
), accompanied by
an excess formation of water molecules that react with the
fresh ATI. These conditions are thought to favor the forma-
tion of dense film structures with a small amount of de-
fects.
In this paper, we describe the possibility of depositing
a-Al
2
O
3
thin films on nickel-containing alloys (stainless
steel and CMSX4) by the oxidant pyrolysis of ATI under
reduced pressure. Both high temperature gradient and fo-
cussed high flow of the inlet gas were used in a stagnation
point flow reactor in an attempt to effectively suppress the
gas phase reaction and produce dense films. The substrate
temperature and the total pressure were systematically var-
ied for the microstructural optimization of a-Al
2
O
3
films.
Deposited films were characterized using scanning electron
microscopy (SEM) together with energy-dispersive X-ray
(EDX) analysis. Phases were determined using grazing-in-
cidence XRD.
2. Results and Discussion
The effect of the substrate temperature was investigated
under a reactor pressure of 100 mbar using stainless steel
substrates. The growth rate shown in Figure 1 is almost in-
dependent of the substrate temperature. The lowest tem-
perature in these experiments is considerably higher than
the decomposition threshold of ATI (200280 C), and it is,
therefore, plausible that the precursor decomposes in suffi-
cient yield before reaching the heated surface. The ob-
served independence of temperature indicates that, under
these conditions, the transition regime from a surface- to a
transport-limited process, generally indicated by a change
of activation energy, occurred at temperatures below
400 C. This result is in agreement with that of Saraie et
al.,
[18]
where the temperature of the transition regime in-
creased, with the ATI concentration in the inlet gas, from
270 C to 350 C. Under 10 mbar, using a gas mixture of
N
2
/ATI with a total flow rate of 1 slm (1.67 10
5
m
3
s
1
),
these authors observed a constant growth rate with increas-
ing temperature, which was explained by the saturation in
the surface-controlled regime.
[18]
The temperature of the
transition regime was registered at 420 C by Morssinkhof
et al.,
[19]
and Go et al.,
[25]
in spite of distinct differences in
deposition conditions. Morssinkhof et al. deposited the alu-
mina films under atmospheric pressure using an N
2
/ATI
gas mixture with a total flow rate of 10 slm, while Go et al.
deposited films under reduced pressure (20 mbar) using
both N
2
/ATI and O
2
/ATI gas mixtures with a total flow
rate of less than 1 slm. The concentration of ATI in the in-
let gas thus appears to be an important parameter influenc-
ing the temperature of the transition regime. The nature of
the accompanying gas, the total flow rate, and the total
pressure do not seem to have a major influence on the tran-
sition regime temperature, in spite of their confirmed effect
on the activation energy of the ATI decomposition.
The homogeneous reaction of reactive intermediates
leading to particle growth appears to be the main disadvan-
tage in the transport-limited deposition process. This disad-
vantage can be overcome by decreasing the collision prob-
ability during the mass transfer process. With this strategy,
Tago et al.
[23]
reported the possibility of maintaining a sur-
face-controlled-like process at temperatures as high as
9001150 C by lowering the concentration of ATI in the
inlet gas. In the present study, low ATI concentration is
complemented by a high temperature gradient and a high
gas velocity towards the substrate.
The SEM image in Figure 2a demonstrates the typical
microstructure of Al
2
O
3
thin films deposited at low tem-
perature. At the surface, the dense cauliflower morphology
shows a spherical shape with an average diameter of
1.8 lm for a deposition temperature, T
dep
, of 300 C,
Chem. Vap. Deposition 2003, 9, No. 4 http://www.cvd-journal.de 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 195
Fig. 1. Deposition rate (Arrhenius plot) as a function of the substrate tem-
perature; deposition on stainless steel under 100 mbar.
Full Paper
0.9 lm at 600 C, and 0.75 lm at 900 C. Above
T
dep
=900 C, the microstructure is typical for crystalline
films, with average crystalline sizes of 0.5 lm at 1080 C
(Fig. 2b). Thus, the film structure is finer at higher temper-
atures.
Figure 3 shows the corresponding XRD analysis of the
Al
2
O
3
films deposited at various temperatures. Only two
crystalline phases of alumina are deposited, i.e., c-Al
2
O
3
starting at T
dep
=800 C and a-Al
2
O
3
at T
dep
>1000 C. The
formation of (Fe,Cr)
2
O
3
results from the oxidation of the
stainless steel substrate. The XRD patterns are in agree-
ment with the SEM observation of a fine crystal structure
at deposition temperatures near 1000 C and above.
The effect of the total pressure in the reactor during de-
position was investigated at a substrate temperature of
1000 C, using the same inlet gas composition and flow rate.
Decreasing the pressure reduces the residence time, limit-
ing gas-phase nucleation, and favoring thin film growth at
the substrate. In Figure 4, the observed linear decrease of
the growth rate with increased pressure indicates that the
deposition process is limited by the gas-phase reactions,
and is consistent with partial depletion of the precursor in-
duced by homogeneous reactions. Near 250 mbar, deposi-
tion was marginal due to almost complete consumption of
ATI in the gas phase, before reaching the substrate.
The SEM analysis of the deposited films in Figure 5
shows that low pressure (50 mbar) results in a homoge-
neous and fine microstructure (Fig. 5a), while, after deposi-
tion under 250 mbar, the surface exhibits a rough and het-
erogeneous structure (Fig. 5b). The corresponding XRD
analysis in Figure 6 demonstrates that the major phase de-
posited under low pressure is c-Al
2
O
3
. The relatively low
intensity of diffraction peaks related to the substrate (see
the peak at 2 h=75) indicates the high thickness of the
film. No significant oxidation of the substrate was noted
upon deposition under 50 mbar. Under pressures between
100 mbar and 200 mbar, the films are composed of a mix-
ture of c-Al
2
O
3
and a-Al
2
O
3
. Under the highest pressure,
250 mbar, there was no apparent films growth since no
peaks related to a-Al
2
O
3
or c-Al
2
O
3
were observed, how-
ever, high intensity peaks related to (Fe,Cr)
2
O
3
and Fe
2
O
3
were detected, indicating the heavy oxidation of the sub-
strate.
The deposition conditions of a-Al
2
O
3
on super alloys
(CMSX4) were chosen, based on the observations during
deposition on stainless steel, selecting a total pressure of
100 mbar and a substrate temperature of 1050 C to
achieve both a high yield and pure a-Al
2
O
3
deposition.
The low and high magnification SEM images in Figure 7
196 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cvd-journal.de Chem. Vap. Deposition 2003, 9, No. 4
a)
b)
Fig. 2. SEM images of Al
2
O
3
films deposited at a) 300 C, and b) 1080 C;
deposition on stainless steel under 100 mbar.
Fig. 3. XRD pattern of Al
2
O
3
films at various deposition temperatures;
deposition on stainless steel under 100 mbar. The reference spectra are
extracted from the powder diffraction files (PDF) of the International Cen-
ter for Diffraction Data (ICDD), a-Al
2
O
3
(PDF 10-0173), c-Al
2
O
3
(PDF
29-0063), and (Cr,Fe)
2
O
3
(PDF 02-1357).
50 100 150 200 250
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.50
D
e
p
o
s
i
t
i
o
n

r
a
t
e

(
m
g

c
m
-
2

h
-
1
)
Pressure (mbar)
Fig. 4. Deposition rate as a function of pressure; deposition on stainless steel
at 1000 C.
Full Paper
show the homogeneity of the coated surface and the de-
fect-free closed structure of the layer, with an average crys-
tal size of 1.3 lm.
The ATI precursor offers a viable alternative to the de-
posited a-Al
2
O
3
films using the classic (AlCl
3
/CO
2
/H
2
)
CVD process.
[6]
This may be a significant issue whenever
the substrate is not able to withstand aggressive by-prod-
ucts, such as HCl.
3. Conclusions
Alumina thin films with a closed structure were depos-
ited on stainless steel and a nickel-based alloy using a cold-
wall CVD reactor. ATI was used as the precursor in a
stagnation point-flow reactor geometry. ATI precursor de-
pletion was prevented by limiting the collision probability
during the mass transfer process, using a low ATI concen-
tration together with low pressure and high gas velocity.
These conditions resulted in a temperature-independent
deposition process between 400 C and 1080 C, however
pressure was observed to have a significant effect on the
growth rate. Amorphous c-Al
2
O
3
and a-Al
2
O
3
were depos-
ited, depending on the substrate temperature. Above about
1000 C, the use of ATI as the precursor shows a promising
potential for producing a-Al
2
O
3
coatings. However, further
efforts are being directed to the effective suppression of
the surface oxidation of the substrate and the co-deposition
of the c-Al
2
O
3
phase.
4. Experimental
The deposition system consisted of an aluminum cold-wall CVD reactor
in a stagnation point geometry. Deposition of Al
2
O
3
films was performed on
stainless steel (316 L, type 1.4404 {x 2 CrNiMo 17 13 2}) and super alloy
(CMSX4) substrates, the latter with a composition of 64 % Ni, 9.7 % Co,
7.5 % Cr, and 12.7 % Al as the major constituents. After an ultrasonic clean-
ing in soak cleaner and ethanol, the substrates were heated by a resistive
heater (Bach Resistor Ceramics) at a rate of 10 Cmin
1
in an argon atmo-
Chem. Vap. Deposition 2003, 9, No. 4 http://www.cvd-journal.de 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 197
b)
a)
Fig. 5. SEM images of Al
2
O
3
films deposited under a) 50 mbar and b)
250 mbar; deposition on stainless steel at 1000 C.
Fig. 6. XRD patterns of Al
2
O
3
films under various pressures; deposition on
stainless steel at 1000 C, Fe
2
O
3
(PDF 21-0920).
b)
a)
Fig. 7. SEM images of a-Al
2
O
3
film deposited on CMSX4 at 1050 C under
100 mbar.
Full Paper
sphere. The temperature was measured at the surface of the heater and in
the gas phase using Ni/CrNi thermocouples. The pressure in the reactor was
set and stabilized using an electronic valve (RS Components). The inlet gas
was composed of two separated flows. The central flow consisted of 0.1 slm
ATI carried in 1 slm argon, with an outer flow of 1.4 slm of a 1:1 O
2
/Ar mix-
ture. The two gas streams were mixed 1 cm above the substrate surface, at
the limit of both inlet tubes. The flow rates were adjusted using electronic
mass flow controllers (Millipore/Tylan General 2900 series), while the
amount of ATI was controlled by the adjustment of its temperature. Due to
the lack of vapor pressure data, the flow rate of ATI was estimated from the
mass loss of the precursor by weighing the container before and after a pro-
longed use. To prevent precursor condensation, all inlet pipes were main-
tained at temperatures above 100 C. After a deposition time of 1 h, the in-
let gas was stopped and the substrate cooled under vacuum at a rate of
10 Cmin
1
. Total pressure was varied in the range 50250 mbar, substrate
temperature between 300 C and 1080 C.
The temperature gradient was estimated under atmospheric pressure by
measuring the temperature within the central inlet tube at various heights
above the substrate surface, heated at 1050 C. These measurements were
performed under both static (ambient atmosphere), and dynamic conditions
using an argon flow corresponding to the deposition conditions (Fig. 8). In
the static case, the temperature is higher than the decomposition threshold
of ATI over a distance of 5 cm from the heated substrate, while under typi-
cal flow conditions, the gradient is higher and the decomposition tempera-
ture of ATI is attained at a distance of less than 23 cm. This distance should
be even smaller under low pressure conditions where the gas velocity is
higher. From the gas velocity (calculated to be >20 mmin
1
), and the dis-
tance along which the temperature exceeds the decomposition threshold of
ATI, the resulting time for the ATI fragments to reach the substrate is less
than 0.1 s. This reactor geometry thus allows a short residence time of the
ATI fragments, minimizing homogeneous reactions in the gas phase, and the
heterogeneous reaction is assumed to be the major path for this deposition
process as long as the low pressure is maintained.
Analysis of the deposited films was performed using SEM, EDX analysis
(Philips XL-30 ESEM), and XRD (Siemens D500 diffractometer). The de-
position rate was estimated from the corresponding mass gain using the
Mettler (ME 320) microbalance. The contribution of the digital error was
calculated at 0.02 mg cm
2
for 1 h of deposition.
Received: October 2, 2002
Final version: December 23, 2002

[1] J. M. Schneider, W. D. Sproul, A. Matthews, Surf. Coat. Technol. 1998,


98, 1473.
[2] B. Felde, A. Mehner, J. Kohlscheen, R. Glbe, F. Hoffmann, P. Mayr,
Diamond Relat. Mater. 2001, 10, 515.
[3] N. Bahlawane, Thin Solid Films 2001, 396, 126.
[4] H. Nakai, O. Harasaki, J. Shinohara, Mater. Chem. Phys. 1998, 54, 131.
[5] R. Cremer, M. Witthaut, K. Reichert, M. Schierling, D. Neuschtz,
Surf. Coat. Technol. 1998, 108109, 48.
[6] J. Mller, M. Schierling, E. Zimmermann, D. Neuschtz, Surf. Coat.
Technol. 1999, 120121, 16.
[7] A. Otsuka, K. Kitagawa, N. Arai, paper IJPGC2000-15028, in Proc.
2000 Int. Joint Power Generation Conf., The American Society of Me-
chanical Engineers, Fairfield, NJ 2000.
[8] F. Fietzke, K. Goedicke, W. Hempel, Surf. Coat. Technol. 1996, 86-87,
657.
[9] O. Zywitzki, G. Hoetzsch, Surf. Coat. Technol. 1995, 77, 754.
[10] CVD of Nonmetals (Ed: W. S. Rees, Jr.), VCH, Weinheim 1996, p. 282.
[11] B. Lux, C. Colombier, H. Altena, K. Sternberg, Thin Solid Films 1986,
138, 49.
[12] E. Fredriksson, J.-O. Carlsson, Surf. Coat. Technol. 1993, 56, 165.
[13] E. Fredriksson, J.-O. Carlsson, Chem. Vap. Deposition 1993, 1, 333.
[14] V. J. Silvestri, C. M. Osburn, D. W. Ormond, J. Electrochem. Soc.
1983, 125, 902.
[15] D. Barreca, G. A. Battiston, G. Carta, R. Gerbasi, G. Rosseto, E. Ton-
dello, P. Zanella, J. Phys. IV 2001, 11, Pr3-539.
[16] G. A. Battiston, G. Carta, G. Cavinato, R. Gerbasi, M. Porchia,
G. Rossetto, Chem. Vap. Deposition 2001, 7, 69.
[17] J. H. Kim, G. J. Choi, J. K. Lee, S. J. Sim, Y. D. Kim, Y. S. Cho, J. Ma-
ter. Sci. 1998, 33, 1253.
[18] J. Saraie, K. Ono, S. Takeuchi, J. Electrochem. Soc. 1989, 136, 3139.
[19] R. W. J. Morssinkhof, T. Fransen, M. M. D. Heusinkveld, P. J. Gel-
lings, Mater. Sci. Eng. A 1989, 121, 449.
[20] D.-H. Kuo, B.-Y. Cheung, R.-J. Wu, Thin Solid Films 2001, 398399,
35.
[21] G. Pauer, H. Altena, B. Lux, J. Ref. Hard Metals 1986, 5, 165.
[22] R. H. Niska, A. P. Constant, T. Witt, O. J. Gregory, J. Vac. Sci. Technol.
A 2000, 18, 1653.
[23] T. Tago, M. Kawase, Y. Masaki, K. Hashimoto, Kagaku Kogaku Ron-
bunshu 1998, 24, 81 (in Japanese).
[24] V. A. C. Haanappel, H. D. van Corbach, R. Hofman, R. W. J. Mors-
sinkhof, T. Fransen, P. J. Gellings, High Temp. Mater. Proc. 1996, 15,
245.
[25] T. Go, N. Hara, K. Sugimoto, J. Jpn Inst. Met., Sendai 1993, 57, 1041
(in Japanese).
______________________
198 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim http://www.cvd-journal.de Chem. Vap. Deposition 2003, 9, No. 4
Fig. 8. Temperature as a function of distance from the substrate
(T
dep
=1050 C) under atmospheric pressure; a) ambient atmosphere, b) Ar
flow of 2.5 slm.

You might also like