Sie sind auf Seite 1von 10

19530

J. Phys. Chem. C 2008, 112, 1953019539

Al2O3 Atomic Layer Deposition with Trimethylaluminum and Ozone Studied by in Situ
Transmission FTIR Spectroscopy and Quadrupole Mass Spectrometry
David N. Goldstein, Jarod A. McCormick, and Steven M. George*,,
Department of Chemistry and Biochemistry, and Department of Chemical and Biological Engineering,
UniVersity of Colorado, Boulder, Colorad, 80309
ReceiVed: May 14, 2008; ReVised Manuscript ReceiVed: September 25, 2008

The atomic layer deposition (ALD) of Al2O3 using sequential exposures of Al(CH3)3 and O3 was studied by
in situ transmission Fourier transform infrared (FTIR) spectroscopy and quadrupole mass spectrometry (QMS).
The FTIR spectroscopy investigations of the surface reactions occurring during Al2O3 ALD were performed
on ZrO2 particles for temperatures from 363 to 650 K. The FTIR spectra after Al(CH3)3 and ozone exposures
showed that the ozone exposure removes surface AlCH3* species. The AlCH3* species were converted to
AlOCH3* (methoxy), Al(OCHO)* (formate), Al(OCOOH)* (carbonate), and AlOH* (hydroxyl) species. The
TMA exposure then removes these species and reestablishes the AlCH3* species. Repeating the TMA and O3
exposures in a sequential reaction sequence progressively deposited the Al2O3 ALD film as monitored by the
increase in absorbance for bulk Al2O3 infrared features. The identification of formate species was confirmed
by separate formaldehyde adsorption experiments. The formate species were temperature dependent and were
nearly absent at temperatures g650 K. QMS analysis of the gas phase species revealed that the TMA reaction
produced CH4. The ozone reaction produced mainly CH4 with small amounts of C2H4 (ethylene), CO, and
CO2. Transmission electron microscopy (TEM) was also used to examine the Al2O3 ALD films deposited on
the ZrO2 particles. These TEM images observed conformal Al2O3 ALD films with thicknesses that were
consistent with an Al2O3 ALD growth rate of 1.1 /cycle. The surface species after the O3 exposures and the
mass spectrometry results lead to a very different mechanism for Al2O3 ALD growth using TMA and O3
compared with Al2O3 ALD using TMA and H2O.
I. Introduction
Atomic layer deposition (ALD) is an ideal technique to
deposit ultrathin films with high conformality and precise
thickness control.1,2 Traditional methods to deposit Al2O3 with
ALD involve sequential surface reactions of Al(CH3)3 (trimethylaluminum (TMA)) and water.3-6 These sequential reactions allow conformal Al2O3 film growth with thickness control
on a variety of substrates including polymers,7 porous membranes,3,8
and nanopowders.9 The details of the Al2O3 ALD reaction have
been extensively studied by a variety of techniques, including
the quartz crystal microbalance measurements,10,11 Fourier
transform infrared (FTIR) spectroscopy,3,12 ellipsometry,4,5 and
X-ray photoelectron spectroscopy (XPS).13 Al2O3 ALD is a
model system and serves as a reference point for other ALD
systems.
The semiconductor industry is interested in growing Al2O3
films with ozone instead of water as the oxygen source. Al2O3
is a high-k dielectric that is used as a dielectric film for both
DRAM and MOS-FETs.14 When ozone is used as the oxidant,
the Al2O3 ALD films can have leakage current densities that
are reduced by two orders of magnitude in comparison with
Al2O3 ALD films deposited with water.15 This improvement and
smaller flat band voltage shifts allow Al2O3 ALD films grown
using ozone to make better gate oxides.15,16 There are also other
advantages when replacing H2O with ozone. Water desorbs
slowly from substrates and requires longer purge times.10 Water
can also leave unreacted hydroxyl groups in the Al2O3 ALD
* Corresponding author.
Department of Chemistry and Biochemistry.
Department of Chemical and Biological Engineering.

films.3 The unreacted hydroxyl groups in the films may affect


the dielectric and material properties of the Al2O3 ALD films.
However, no change in equivalent oxide thickness (EOT) of
Al2O3 ALD films was observed when ozone was used as the
oxidant.15
Previous research has been conducted on Al2O3 ALD with
TMA and O3. A growth rate of 0.8 per cycle at 300-450 C
has been measured by several investigations.13,14,17,18 XPS
measurements revealed Al2O3 ALD films that had lower carbon
impurities with ozone compared with water.13 Al2O3 films grown
with ozone also had a reduced percentage of Al-Al defects
that degrade the electrical properties of the Al2O3 ALD films.
These defects have been characterized using XPS by the
presence of a shoulder on the 72.5 eV Al 2p peak.19 Time-offlight secondary ion mass spectrometer (TOF-SIMS) analysis
has probed the bulk of Al2O3 ALD films and revealed different
impurity levels in Al2O3 films grown with ozone compared with
Al2O3 films grown with water.13 Hydrogen impurities were
reduced in the ozone grown films.13
To understand the differences between Al2O3 ALD with TMA
and either H2O or ozone, this study employed in situ transmission FTIR spectroscopy to monitor the surface species formed
and removed during the TMA and O3 exposures. The FTIR
spectra also revealed the growth of Al2O3 bulk vibrational modes
versus the number of ALD reaction cycles. Additional experiments also monitored the gas phase products during O3 and
TMA exposures using a quadrupole mass spectrometer (QMS).
The resulting Al2O3 ALD films on the ZrO2 particles were then
analyzed with transmission electron microscopy (TEM) to obtain
the Al2O3 ALD growth per ALD cycle. These FTIR, QMS, and

10.1021/jp804296a CCC: $40.75 2008 American Chemical Society


Published on Web 11/13/2008

Al2O3 Atomic Layer Deposition

Figure 1. Schematic of W grid in the ALD reactor.

Figure 2. Schematic of the inlet and outlet connections to the ALD


reactor.

TEM studies help to clarify the surface chemistry and thin film
growth mechanism during Al2O3 ALD with TMA and ozone.
II. Experimental Section
The surface chemistry and thin film growth during Al2O3
ALD was studied using sequential exposures of TMA and O3
at various temperatures. Al2O3 ALD films were grown on ZrO2
particles in an ALD reactor designed for in situ FTIR spectroscopy studies.12,20 Figure 1 presents a schematic of the ALD
reactor. The reactor was a warm-wall reactor where the chamber
walls were heated to 350 K while the sample could be
independently heated to >900 K. Figure 2 shows a schematic
of all the inlet and outlet connections to the ALD reactor. Two
argon mass flow controllers (MFC) regulated the flow of argon

J. Phys. Chem. C, Vol. 112, No. 49, 2008 19531


through the reactor at 220 sccm (110 sccm per MFC). This flow
established a base pressure of 1.30 Torr. An Alcatel 2012A
rotary vane pump removed the argon and reaction byproducts
from the reactor.
Pneumatic leak valves with conductance metering valves
allowed accurate exposure of the reactants. A Labview measurement system controlled the reactant exposures and integrated
the area beneath the pressure transients that occurred during
the reactant exposures. The reactant exposures were performed
with use of micropulses that were less than the exposures
required for the reactions to reach completion. The absolute
reactant exposures were determined with no ZrO2 nanoparticles
in the reactor after a sufficient number of micropulses for the
reaction on the reactor walls to reach completion. Under these
conditions, the reaction products do not interfere with the
measurement of the absolute reactant exposure.
The Al2O3 ALD was coated onto ZrO2 nanopowders supported in a 2 3 cm2 tungsten grid.12,20,21 Each W grid was 50
m thick and was photoetched to 100 grid lines per inch.
Tantalum foil was spot-welded on the sides of the grid to
improve current transfer through the grid. The entire grid was
then attached to a copper clamp that was interfaced via an
electrical feedthrough to a Hewlett-Packard 6268B power
supply. Resistive heating was used to heat the sample. A Love
Controls 16A3 PID controller interfaced to a type K thermocouple mounted on the sample grid provided temperature control
of the sample. The feedback loop maintained the sample
temperature at (2 C.
Preparation of the substrate involved pressing ZrO2 nanoparticles into the W grid.12,20,21 Each grid was first sonicated in
deionized water and methanol and then blown dry with ultrapure
nitrogen. The grid was then placed into a stainless steel die and
covered with an excess of nanopowders. Subsequently, a manual
press forced the particles into the W grid until the particles made
a dense matrix with very few pinholes in the sample. Excess
nanopowders lying on the top of the grid were easily removed
with a razor blade. The finished sample contained about 22 mg
of ZrO2 powder. This quantity of ZrO2 powder is equivalent to
a surface area of 0.44 m2. Finally, a type K thermocouple
was attached to the top of the sample grid with Ceramabond
571 Epoxy. This epoxy electrically isolated the thermocouple
and kept the thermocouple firmly attached to the sample during
the experiment.
An infrared beam from a Nicolet Magna 560 FTIR spectrometer was externally aligned to pass through the W grid
sample. The ZrO2 nanopowder substrates provided a large
surface area and improved the signal-to-noise ratio for infrared
absorption. The entire sample stage could be translated along
the vertical z-axis direction to move the sample out of the FTIR
beam. This displacement allowed the background reference
spectra to be measured frequently over the course of these
experiments. A liquid nitrogen cooled MCT-B (mercury cadmium telluride) detector allowed measurement of the infrared
spectra from 400 to 4000 cm-1. During the reactant exposures,
the gate valves on the CsI windows were closed to prevent
deposition on the windows. All FTIR spectra were obtained at
4 cm-1 resolution using 100 averaged scans and were referenced
to the CsI window background. However, most of the FTIR
spectra in this paper are presented as difference spectra.
Mass spectrometry analysis was performed in a rotary reactor
designed for ALD on nanoparticles. The design and operation
of this reactor has been discussed in previous publications.9,22
To provide in situ quadrupole mass spectrometry analysis, a
200 amu quadrupole mass spectrometer with a pressure reduc-

19532 J. Phys. Chem. C, Vol. 112, No. 49, 2008


tion system (PPR200, SRS Inc., Sunnyvale, CA) was attached
to the reactor. During reactant exposures, the QMS scanned the
mass range from 1-75 m/z with 0.1 m/z resolution. A Faraday
cup was used as the detector with no electron multiplier. With
these settings, about 5 s was required to scan the entire mass
range.
Micropulses of both TMA and O3 were used to determine
the exposures required for the reactions to reach completion
with 1.0 g of ZrO2 nanoparticles in the rotary reactor.9 TMA
was dosed into the rotary reactor to a pressure of 0.3 Torr above
the baseline pressure. The O2/O3 mixture was dosed into the
rotary reactor to a pressure of 0.5 Torr. Each reactant reacted
for 60 s in the chamber and then was purged for 60 s before a
final argon pulse flushed the chamber. The reactor then returned
to base pressure before starting the next set of reactant
micropulses. Approximately 20 micropulses of TMA and 60
micropulses of O3 were required for each reaction to reach
completion with 1.0 g of ZrO2 nanoparticles in the rotary reactor.
The ZrO2 particles were obtained from Nanomaterials Research Corporation (Longmont, CO). These ZrO2 particles were
spherical with an average diameter of 50 nm and a surface
area of 20.2 m2/g. TMA was obtained from Aldrich (Milwaukee, WI) and had a purity of 97%. The water was high
performance liquid chromatography (HPLC) grade from Fisher
Scientific (Pittsburgh, PA). Ozone was produced from UHP
grade oxygen (99.9%) obtained from Airgas Ltd. (Cheyenne,
WY). All chemicals were used as purchased, except for water,
which was subjected to 3 freeze-pump-thaw cycles prior to
use.
Ozone was obtained from O2 by flowing 300 sccm of O2
into a DelOzone LC-14 ozone generator (San Luis Obispo, CA).
This flow produced a 6 psi pressure in the generating cell. At
100% power, the ozone concentration at the outlet was 3.7%
with the balance being O2. When the ozone was not going
through the ALD reactor, the ozone was sent through a magnesia
ozone destruct unit and the remaining O2 was evacuated with a
separate rotary vane pump. In the rotary reactor, the O3 was
generated with an Ozonia OZAT CFS-1A ozone generator
(Duebendorf, Switzerland). The ozone generator ran with O2
at a flow rate of 0.2 m3 h-1 and power of 510 W. These
conditions produced an O3 concentration of 12% by mass.
TEM analysis was performed in the Department of Molecular
and Cellular Biology at the University of Colorado at Boulder.
The TEM results were obtained with a Philips CX11 highresolution transmission electron microscope with 80 kV beam
potential. A Gatan slow scan charge-coupled device camera
captured the TEM images. The TEM studies monitored the
conformality and thickness of the Al2O3 films on the ZrO2
particles.
III. Results and Discussion
A. Fourier Transform Infrared Spectroscopy. Studying
the surface chemistry of ALD processes requires a reliable
starting surface. FTIR spectroscopy can determine the initial
surface species on the ZrO2 nanopowders to ensure that they
will be suitable for Al2O3 ALD. A range of absorbances is
observed on the ZrO2 nanoparticles including the following:
O-H stretching vibrations at 3670-3780 cm-1; C-H stretching
vibrations at 2850-3050 cm-1; and the bulk ZrO2 absorbance
at frequencies <800 cm-1. These vibrations agree with vibrational features for ZrOH* species on a tetragonal ZrO2 surface.23
Hydroxyl groups on ZrO2 serve as the reactive sites to initiate
the reactions for Al2O3 ALD. The C-H stretching vibrational

Goldstein et al.

Figure 3. FTIR difference spectra after (a) last water exposure, (b)
next TMA exposure, (c) first ozone exposure, and (d) next TMA
exposure. All exposures were performed at 450 K. C ) carbonate, F
) formate.

features are from adventitious carbon species on the surface


resulting from air exposure that will be removed in the cleaning
step.20
A consistent cleaning procedure was used to prepare the
starting ZrO2 nanoparticles. First, the ZrO2 nanoparticles were
annealed at 1000 K for 1 min under argon flow. After cooling
to the desired reaction temperature, the FTIR spectra revealed
the removal of all C-H and C-C vibrational features between
1200 and 1600 cm-1. Because annealing removed most of the
hydroxyl groups on the ZrO2 nanopowder, a 1 s H2O dose was
used to rehydroxylate the ZrO2 nanopowder. After this treatment,
the infrared spectrum contained just the O-H stretching
vibrations and the bulk ZrO2 absorbance.20 These features
indicated a clean hydroxyl-terminated ZrO2 starting surface.
Subsequently, the ZrO2 nanoparticles were exposed to 10 AB
cycles of TMA (A) and water (B) at 450 K to deposit a starting
Al2O3 ALD surface. The TMA exposure was 0.85 Torr s and
the H2O exposure was 0.60 Torr s. Each exposure was separated
by a 90 s purge to minimize chemical vapor deposition (CVD)
reactions that could increase the growth rate. These 10 AB
cycles deposited 1.2 nm of Al2O3 ALD on the ZrO2
nanopowders.4,9-11 FTIR spectra recorded during these reactions
ensured that each reaction reached completion before switching
to the next reactant. The AlCH3* surface species were observed
by the C-H stretching vibrations between 2820 and 2980 cm-1
and the symmetric Al-CH3 deformation mode at 1212 cm-1.
The AlOH* surface species were observed by the AlO-H
stretching vibrations between 3650 and 3770 cm-1. The FTIR
spectra confirmed growth of Al2O3 after 10 AB cycles by an
increase in the bulk Al2O3 absorbance feature at 900-1000
cm-1.
A comparison at 450 K between the first AB cycle during
Al2O3 ALD with TMA and O3 and the last (10th) AB cycle
during Al2O3 ALD with TMA and water is shown in Figure 3.
These FTIR spectra are difference spectra referenced to the
previous reactant exposure and are displaced from the origin
for clarity in presentation. Figure 3a is the FTIR difference
spectrum taken after the last H2O exposure of 0.75 Torr s at
450 K during Al2O3 ALD with TMA and water. A positive
absorbance feature is observed at 3640-3730 cm-1 corresponding to O-H stretching vibrations from the added AlOH*

Al2O3 Atomic Layer Deposition


hydroxyl surface species. In addition, negative absorbance
features are observed at 2920-2980 cm-1 corresponding to
C-H stretching vibrations and at 1212 cm-1 corresponding to
the Al-CH3 deformation mode. These negative absorbance
features are both from the removal of AlCH3* surface species.
Figure 3b shows the FTIR difference spectra for the next TMA
exposure of 0.85 Torr s at 450 K. This spectrum indicates that
TMA removed the AlOH* surface species and added AlCH3*
surface species. The FTIR difference spectra in Figure 3a,b are
consistent with previous FTIR studies of Al2O3 ALD with TMA
and H2O.3,12
The surface species change dramatically after the first ozone
exposure. Figure 3c shows a markedly different spectrum with
new positive absorbance features visible between 1200 and 1700
cm-1. The most prominent new features were observed at 1388,
1404, and 1597 cm-1. There were also smaller new features
observed at 1320 and 1475 cm-1 and a shoulder at 1720 cm-1.
Not all of the C-H features were eliminated after long ozone
exposures. There were small absorbance features at 2923 and
3016 cm-1 for C-H stretching vibrations that partially result
from slight frequency shifts. In addition, the absorbance for the
O-H stretching vibrations at 3734-3778 cm-1 was reduced in
intensity compared with the intensity observed in Figure 3a after
H2O exposures. In addition, Figure 3d shows that the new
absorbance features added by the O3 exposure were completely
removed by the next TMA exposure. TMA exposures reform
the absorbances from the C-H stretching vibrations at
2820-2970 cm-1 and the methyl deformation at 1212 cm-1.
The absorbance from the O-H stretching vibrations at
3734-3778 cm-1 was also removed by TMA doses.
The new absorbance features appearing after the ozone
exposure are very characteristic of formate and carbonate groups
on the Al2O3 ALD surface. We first reported these new formate
features at the AVS Topical Conference for Atomic Layer
Deposition in 2006 (ALD2006).24 Formate groups are very
common on metal oxide surfaces and can be formed by exposing
metal oxides to a variety of reagents including carbon monoxide,
methanol, and formaldehyde.25-27 The absorption features at
1388 and 1597 cm-1 correspond to the symmetric and antisymmetric OCO modes of bound formate species. In addition,
the 1404 cm-1 shoulder is attributed to the CH bend of formate
species. In the C-H stretching region, the two peaks at 2923
and 3016 cm-1 are identified as the Fermi resonance of the
antisymmetric OCO mode mixing with the lone C-H stretching
vibration in the formate surface group.26 All of these values
match literature values for formate on aluminum oxide.26,28,29
Formate features during Al2O3 ALD with ozone have also
been observed recently by FTIR studies on planar surfaces that
were complemented by DFT calculations.30 These studies
observed the same methoxy modes at 1388 and 1475 cm-1 that
were monitored in this study on the ZrO2 nanoparticles. In
addition, the ratios of the absorbances for the methoxy and
primary formate vibrational features were similar on the planar
surfaces and the ZrO2 nanoparticles. This agreement also rules
out the possibility that ozone decomposition may have prevented
the ozone from reaching the interior surfaces of the ZrO2
nanoparticle sample.
The surface coordination of the formate features is described
by the frequency difference between the symmetric and asymmetric OCO stretching vibrations. Our observed experimental
difference of 212 cm-1 is greater than that of the free formate
ion. Consequently, the formate species are doubly coordinated
to aluminum sites on the surface.28 These doubly coordinated
surface species formed after ozone exposure may contribute to

J. Phys. Chem. C, Vol. 112, No. 49, 2008 19533

Figure 4. FTIR difference spectra after (a) reference ozone exposure,


(b) HCOH (formaldehyde) exposure on hydroxyl-terminated surface,
and (c) next TMA exposure after formaldehyde exposure. All exposures
were conducted at 450 K. F ) formate.

the oxygen-rich stoichiometry of Al2O3 films deposited using


TMA and ozone.31 The remaining shoulders at 1320 and 1720
cm-1 are the symmetric and asymmetric OCO stretching
vibrations of carbonate groups bound to alumina. Some of the
hydroxyl features observed in the difference spectra could also
result from the C-OH group atop surface carbonate species.
These carbonate species can be formed by further oxidation of
the formate species. Carbonate species can be prepared by
reacting CO2 with alumina surfaces. The vibrational frequencies
observed in this study match closely with literature values for
carbonate.25
Control experiments confirm the identity of the formate
species formed during ozone exposures. For these control
experiments, a fresh Al2O3 ALD film was grown at 450 K and
then exposed to 1.0 Torr s of formalin solution. Formalin is a
solution of 37% formaldehyde, 10% methanol, and 53% water.
Exposing aluminum oxide to formaldehyde will produce surface
formate groups.27,32 For reference, Figure 4a shows the FTIR
difference spectrum after an ozone exposure on the Al2O3 ALD
surface. This spectrum is identical to the spectrum shown in
Figure 3c. Figure 4b displays the FTIR difference spectrum after
a formalin exposure on an Al2O3 ALD surface at 450 K. The
formaldehyde leads to the same major absorbance features at
1388, 1404, and 1597 cm-1 observed in Figure 4a. The
differences between these spectra are the peaks at 1098, 1320,
and 1720 cm-1. The first two of these peaks are attributed to
carbonate groups bound on the surface. The last feature is likely
absorbance from C-O stretching vibrations of surface methoxy
groups.25,26
The formate species on the Al2O3 ALD surface formed by
the formaldehyde exposure was then exposed to TMA at 450
K. The FTIR difference spectrum after the TMA exposure is
shown in Figure 4c. The negative absorbances at 1404 and 1597
cm-1 indicate that TMA removes the formate species from the
Al2O3 surface. In addition, the TMA removes additional AlOH*
species as shown by the negative infrared absorbance features
at 3730 and 3770 cm-1. The positive absorbance features from
2920 to 2980 cm-1 in the C-H stretching region and at 1212
cm-1 for the Al-CH3 deformation mode indicate that TMA
has repopulated the Al2O3 surface with AlCH3* species.
The surface chemistry during Al2O3 ALD with TMA and
ozone may depend on the substrate temperature. Figure 5 shows

19534 J. Phys. Chem. C, Vol. 112, No. 49, 2008

Goldstein et al.

Figure 5. FTIR difference spectra after (a) third TMA exposure, (b)
third ozone exposure, (c) fourth TMA exposure, and (d) fourth ozone
exposure. All exposures were performed at 550 K. F ) formate.

FTIR difference spectra after the third and fourth sequential


TMA and ozone exposures at 550 K. The reference spectrum
for each difference spectrum is the FTIR spectrum after the
previous exposure. The absorbance features observed after the
TMA and ozone exposures are very similar for the third and
fourth cycle and also correspond closely to the absorbance
features monitored at 450 K. This close correspondence suggests
that the Al2O3 ALD reaction mechanism is similar at 450 and
550 K.
The FTIR difference spectra in Figure 5 also consistently
show the disappearance and appearance of a small absorbance
feature at 2280 cm-1 with TMA and O3 exposures, respectively.
This feature is attributed to weakly bound CO coordinating to
Al3+ centers on an alumina surface. CO can be formed whenever
the formate species (OCHO) decompose on the alumina
surface.25 This absorbance feature becomes more prominent as
more formate species undergo decomposition. To check this
hypothesis, formate decomposition was examined during the
formaldehyde control experiment. The formate features were
observed to disappear slowly at 550 K. The intensity of the
formate features was reduced by one-third in 22 h. In contrast,
when the temperature was reduced to 473 K, the formate
remained constant over 8 h of scanning. This behavior may
explain why little CO was observed at 450 K. Not enough
formate is decomposing at 450 K to produce CO on the alumina
surface.
Figure 5 also shows the decrease and increase of absorbance
for the bulk Al2O3 absorbance at 900-1000 cm-1 for TMA
and ozone exposures, respectively.33 The increase of this
absorbance during the O3 exposures is always larger than the
decrease of absorbance during the TMA exposures. As a result,
the absorbance of this feature increases gradually versus the
number of AB cycles. This behavior is consistent with the
growth of the Al2O3 ALD film.
The self-limiting nature of the surface reactions during Al2O3
ALD with TMA and ozone can be monitored using the
integrated infrared absorbance for various surface species. If
an ALD reaction is self-limiting, then the surface coverage will
not increase after a certain reactant exposure. For these
experiments, the integrated absorbance was defined for the C-H
stretching vibrations at 2820-2980 cm-1, the Al-CH3 deformation mode at 1175-1250 cm-1, the O-H stretching vibra-

Figure 6. (a) Normalized integrated absorbances during (a) TMA


exposure and (b) ozone exposure at 550 K showing C-H stretch, O-H
stretch, asymmetric OCO stretch, symmetric OCO stretch, and CH3
deformation.

tions at 3600-3800 cm-1, and the two major formate vibrations:


the antisymmetric OCO band at 1575-1625 cm-1 and the
symmetric OCO band between 1350 and 1425 cm-1.
Figure 6a compares the normalized integrated absorbance of
the surface species versus TMA exposure at 550 K. The
absorbances of the O-H stretching vibration and the symmetric
and asymmetric OCO stretching vibrations for the formate
species decrease versus TMA exposure. Hydroxyls react much
more rapidly than the formate features since the absorbance for
the O-H stretching vibrations is reduced before the absorbance
for the formate features. In close correspondence, the absorbance
of the C-H stretching vibrations and the Al-CH3 deformation
mode for the AlCH3* species concurrently increase versus TMA
exposure. The measurements indicate that TMA exposures of
0.9 Torr s are sufficient for the TMA surface reaction to reach
completion. However, the absorbance for the antisymmetric
OCO stretch is not completely extinguished even after 2.0 Torr s
of TMA exposure. This behavior indicates that some formate
groups do not react with TMA at 550 K.
The normalized integrated absorbance of the surface species
during the ozone exposure is presented in Figure 6b. The
absorbance for the various surface species again shows the characteristic signature for self-limiting surface reactions. The
absorbances of the O-H stretching vibration and the symmetric
and asymmetric OCO stretching vibrations for the formate
species increase versus O3 exposure. In close correspondence,
the absorbances of the C-H stretching vibrations and the
Al-CH3 deformation mode for the AlCH3* species concurrently
decrease versus O3 exposure. Fermi resonances from the formate

Al2O3 Atomic Layer Deposition

Figure 7. FTIR difference spectra in the region from 900 to 1900


cm-1 recorded after the second ozone exposure at (a) 363, (b) 450, (c)
550, and (d) 650 K. M ) methoxy, C ) carbonate, F ) formate.

groups leave residual absorbances in the 2820-2980 cm-1 range


and prevent the C-H stretching features from being completely
removed. These measurements indicate that O3 exposures of
1.0 Torr s are sufficient for the O3 surface reaction to reach
completion.
The formate species are dependent on the surface temperature.
Figure 7 displays FTIR spectra that were recorded after the
second ozone exposure at temperatures between 363 and 650
K. Below 650 K, the peaks associated with formate and
carbonate groups were visible between 1200 and 1700 cm-1.
The formate features were obscured when the temperature was
raised to 650 K. This disappearance may result from the
decomposition of the formate and carbonate species into CO
and CO2. At all temperatures, surface methyl groups were
removed based on the negative absorbance of the methyl
deformation feature at 1212 cm-1. The bulk infrared absorbance
of Al2O3 was visible in all the spectra at 900-1000 cm-1.
The strongest bulk infrared absorbance was observed at the
higher temperatures.
Transient species that could produce formate species were
also resolved during the temperature studies. The low-temperature experiments revealed new absorbances at 1089 and 1456
cm-1 that can be attributed to the C-O stretching vibration and
antisymmetric CH3 deformation of methoxy species. The C-O
feature was obscured at higher temperatures by the broad Al-O
bulk absorption mode. In addition, the symmetric CH3 deformation of surface methoxy groups was obscured by the symmetric
deformation of AlCH3* surface species. Methoxy groups are a
potential intermediate to surface formate groups.26 In agreement,
features attributed to methoxy groups decrease and the formate
features increase above 363 K.
The 2000-4000 cm-1 region after the second ozone exposure
at a variety of temperatures is shown in Figure 8. The hydroxyl
region at 3650-3770 cm-1 revealed a greater proportion of
higher frequency peaks at higher temperatures. The hydroxyl
vibrations after ozone exposures have two prominent absorbances at 3718 and 3778 cm-1. In comparison, the hydroxyls
observed during Al2O3 ALD with H2O have a diffuse band
ranging from 3670 to 3730 cm-1.12 This contrast is consistent
with a difference in the basicity of the hydroxyls on the alumina
surface. The hydroxyl vibrations are directly correlated to the
basicity of surface hydroxyl groups.34,35 A higher (OH)

J. Phys. Chem. C, Vol. 112, No. 49, 2008 19535

Figure 8. FTIR difference spectra in the region from 2000 to 4000


cm-1 recorded after the second ozone exposure at (a) 363, (b) 450, (c)
550, and (d) 650 K. M ) methoxy.

vibrational frequency indicates increased basicity of surface


hydroxyl groups. The FTIR spectra suggest that Al2O3 ALD
grown at higher temperatures with ozone produces a larger
proportion of strongly basic hydroxyls than Al2O3 ALD grown
with water.
The C-H stretching features are consistent with the removal
of AlCH3* species and production of new C-H stretching
vibrations corresponding to methoxy species. The absorbance
features around 2850 cm-1 are associated with the CH3
symmetric stretching vibration of methoxy species. These
methoxy features are lost at higher temperatures. As the
temperature is raised to 550 K, Figure 8c reveals an absorbance
at 2280 cm-1 that has already been identified as CO from the
decomposition of formate. In addition, a peak at 2324 cm-1
appeared when the temperature was increased to 650 K. This
peak is attributed to weakly bound CO2 that coordinates to Lewis
base sites on the alumina surface.25 A correlation of the
vibrations of the observed surface species, their referenced
literature values, and observed experimental values is given in
Table 1.
Another control experiment was performed at temperatures
between 450 and 650 K to distinguish the effects of ozone from
oxygen on a surface covered with AlCH3* species following
the TMA reaction. Oxygen was dosed into the reactor chamber
through the ozone generator with the same exposure as ozone
for each growth temperature. The control reactions were
performed one week after the final ozone experiment to ensure
that no residual ozone was left in the ozone generator. Direct
comparison of the control reactions in Figure 9 shows that
oxygen reacts only slightly up to 550 K because very little of
the absorbance for the A1CH3 deformation mode at 1212 cm-1
was removed from the spectrum. At 650 K, oxygen reacts with
AlCH3* species but does not produce any absorbance for O-H
stretching vibrations from hydroxyl species. Al2O3 ALD growth
may also be occurring at 650 K because the bulk Al2O3
absorbance increases after the oxygen exposure. However,
multiple sequential TMA and O2 exposures were not performed
to confirm Al2O3 ALD growth.
The bulk Al-O absorbance feature can be used to monitor
the growth of the Al2O3 ALD film on the ZrO2 nanopowders.
An Al2O3 ALD film was grown using 40 sequential exposures
of 0.9 Torr s TMA and 1.0 Torr s of ozone at 550 K. A 90 s
purge separated the reactants to minimize possible CVD and to

19536 J. Phys. Chem. C, Vol. 112, No. 49, 2008

Figure 9. FTIR difference spectra after an O2 exposure equivalent to


the O2/O3 exposures at (a) 450, (b) 550, and (c) 650 K. (d) FTIR
difference spectra after an O3 exposure at 650 K.

Goldstein et al.

Figure 11. TEM image of ZrO2 particles coated with Al2O3 after 40
reaction cycles of TMA and ozone.

TABLE 1: Vibrational Assignments Comparing Literature


and Observed Experimental Values
assignment

lit. (cm-1)

exptl (cm-1)

Al-O phonon
(s) C-O (m)
Sym. Al-CH3 def
(s) OCO (c)
(s) OCO (f)
CH (f, m)
asym CH3 def (m)
(as) OCO (f)
(a) OCO (c)
(s) CH3 (m)
(as) CH3 (m)
(s) OCO + CH (f)

110026,29,30,32,38
11973,12
131525
138026,29,30,32,38
139526,29,30,32,38
145026,29,30,32,38
159526,29,30,32,38
171025
284626,29,30,32,38
295026,29,30,32,38
297026,29,30,32,38

900-1000
1089
1212
1320
1388
1404
1475
1597
1720
2850
2923
3016

allow desorbing species to be swept from the nanopowders.


FTIR scans were recorded at regular intervals to monitor the
Al2O3 ALD film growth. The absolute FTIR spectra shown in
Figure 10 were acquired after the O3 exposures and are
referenced to the CsI windows. There is a continuous increase
of the bulk Al-O absorbance mode versus number of AB
cycles.

Figure 10. Absolute FTIR spectra showing the growth of the bulk
Al-O bulk absorbance feature after various numbers of TMA and O3
reaction cycles.

After the 40 AB reaction cycles at 550 K, the sample was


removed from the reactor and TEM was performed on the coated
ZrO2 nanopowders to determine the film conformality and ALD
growth rate. The bright field TEM image recorded at 130 000
is shown in Figure 11. The ZrO2 nanopowders have a conformal
coating with a thickness of 65 . This thickness is consistent
with a growth rate of 1.1 /cycle for Al2O3 ALD with ozone at
550 K. This value was determined by subtracting the estimated
12 base layer of Al2O3 ALD grown using TMA and H2O. A
second deposition experiment at 650 K provided the same
growth rate and also showed conformality of the Al2O3 ALD
film.
B. Quadrupole Mass Spectrometry. Identification of the
gas phase species formed after O3 exposures can help determine
how the new surface species were formed on the Al2O3 surface.
The mass spectrum recorded during the first O3 micropulse at
550 K is shown in Figure 12a. This spectrum is shown using a
log intensity scale because the reaction products are very small
compared with the O2 fragmentation pattern. The large peak at
m/z 32 is attributed to a mixture of O2 and O3. Hydrocarbons
evolved in this first ozone micropulse are identified as C2H4
(ethylene) by the three peaks from m/z 26 to 28 and CH4
(methane) by the peaks from m/z 12 to 16. Ethylene peaks are
observed in the QMS scans until the 10th micropulse of ozone
when their signals reach the noise level of <10-9 Torr.
The gas phase species change as the reaction progresses.
Figure 12b shows the mass spectrum during the 55th O3
micropulse. Oxygen and methane are again the majority species.
However, CO and CO2 are also detected at m/z 28 and 44 and
no ethylene was observed at m/z 26-28. Since the 55th
micropulse occurs 118 min after the first O3 micropulse, the
CH4, CO, and CO2 species are likely reaction products from
the slow decomposition of surface formate and carbonate
species.29 In addition, no H2O is observed by the QMS during
ozone exposure. However, any possible product H2O would be
expected to react with the AlCH3* species on the surface. This
reaction between product H2O and AlCH3* species may yield
the surface hydroxyl groups observed in the FTIR spectra.
The gas phase species from the TMA reaction are very similar
to reaction products observed during Al2O3 ALD with TMA
and water.36 Figure 13a reveals the mass spectrum after the first

Al2O3 Atomic Layer Deposition

Figure 12. (a) Mass spectrum during the first O3 micropulse showing
CH4 and C2H4 at 550 K. (b) Mass spectrum during the 55th O3
micropulse showing CH4, CO, and CO2 at 550 K.

TMA micropulse. No peaks from TMA are observed during


this first micropulse. The only major reaction product is CH4
with its characteristic peaks between m/z 12 and 16. In addition,
a slight amount of CO is detected at m/z 28. The appearance of
CO reaction product suggests that formate and carbonate species
are still decomposing on the surface at 550 K.
Figure 13b shows the mass spectrum after the 55th TMA
micropulse. The TMA reaction has reached completion on the
surface and the mass spectrum is consistent with unreacted TMA
with its primary mass cracking fragments centered at m/z 57
and 42. The peaks between m/z 12 and 16 are mass cracking
fragments for TMA. In addition, slight amounts of CO and CO2
are detected at m/z 28 and 44. The appearance of these reaction
products suggests that formate and carbonate species are still
decomposing on the surface at 550 K. Gas phase CO and CO2
are present in the mass spectrum after subtracting the cracking
patterns from both CH4 and TMA. Analysis of the QMS data
shows that the quantity of CO and CO2 steadily decreases versus
time.
C. Mechanism of Al2O3 ALD with Ozone. The FTIR
spectra and the QMS data allow a mechanism to be proposed
for Al2O3 ALD with TMA and ozone. The mechanism during
the ozone reaction is presented in Figure 14. The initial surface
is covered with AlCH3* species resulting from the TMA reaction
on a hydroxylated initial surface. There are two potential
pathways for ozone to react with AlCH3* species. One reaction
pathway involves oxygen insertion into the AlC-H bond. The
second reaction pathway involves oxygen insertion into the
Al-C bond.

J. Phys. Chem. C, Vol. 112, No. 49, 2008 19537

Figure 13. (a) Mass spectrum during the first TMA micropulse
showing CH4 and CO at 550 K. (b) Mass spectrum during the 55th
TMA micropulse showing TMA at 550 K.

Figure 14. Proposed mechanism for O3 reaction during Al2O3 ALD


using TMA and O3.

Oxygen insertion into the AlC-H bond has been predicted


by quantum mechanical simulations and is believed to produce
ethylene as a reaction product.6,17,37 Oxygen insertion into the
AlC-H bond produces AlCH2OH* species. The presence of
these AlCH2OH* species are not ruled out by the FTIR spectra
because the O-H stretching vibration for AlCH2OH* species
would be difficult to distinguish from that of AlOH* species.
These AlCH2OH* species could also be identified by the
presence of methylene C-H stretching vibrations or CH2
rocking modes. However, the intense formate vibrations and
Fermi resonances obscure these vibrations. The transient nature
of these species may also prevent these species from being
isolated by using time- or temperature-dependent measurements.
Adjacent AlCH2OH* species then eliminate ethylene and
leave behind two AlOH* species as shown in Figure 14. This

19538 J. Phys. Chem. C, Vol. 112, No. 49, 2008


reaction pathway is supported by the immediate production of
hydroxyl groups after small ozone doses in Figure 6b and the
ethylene observed by the QMS results as shown in Figure 12a.
The combination of two AlCH2OH* species to produce two
AlOH* species is important because this decomposition pathway
regenerates hydroxyl groups that are needed for the subsequent
TMA reaction.
Oxygen atom insertion into the Al-C bond produces a
methoxy group that is only stable below 473 K according to
the FTIR spectra. Adjacent methoxy groups then combine to
form formate species, releasing CH4 and H2 into the gas phase.
This reaction has been well-characterized in the surface
chemistry literature.29 This reaction is also the likely source of
gaseous CH4 observed by the QMS measurements in Figure
12. Further oxidation of the formate groups yields carbonate
groups. These carbonate features are in a minority as suggested
by their weak infrared absorbances. Oxygen insertion into the
Al-C bond is likely the primary pathway for ozone activation
based on the large amount of CH4 observed in the QMS
compared with the small C2H4 signals that were only monitored
during the initial O3 micropulses. These minority C2H4 reaction
products may only be observed during the initial O3 micropulses
because of higher AlCH3* surface coverages.
As the reaction progresses, the formate and carbonate groups
decompose into CO and CO2, respectively. The CO and CO2
gas phase products were both observed by the QMS measurements during the later O3 micropulses in Figure 12b and during
the TMA micropulses in Figure 13. The formate and carbonate
decomposition leaves behind one hydroxyl group on alumina
per decomposed moiety. This hydroxyl group may be the
isolated hydroxyl observed by the FTIR spectra at 3778 cm-1.
CH4 could also be produced when these surface hydroxyl groups
react with nearby AlCH3* species.
The TMA reaction is believed to be very similar to the
reported TMA reaction for Al2O3 ALD with TMA and water.3
AlOH* species formed during ethylene elimination or formate
or carbonate decomposition in Figure 14 can easily react with
TMA. This reaction reforms the AlCH3* species that are
observed in the FTIR spectra. The efficiency of this TMA
reaction is illustrated in Figure 6a and requires similar exposures
of TMA as the TMA reaction during Al2O3 ALD with TMA
and water. The hydroxyl species are reduced more rapidly than
the formate species during the TMA reaction.
Figure 6a also shows that TMA is unable to remove all the
formate species from the surface. This observation has been
noted by other infrared studies of the TMA reaction during
Al2O3 ALD with ozone.30 On the basis of our results, the
complete removal of the formate species is dependent on the
decomposition of the formate species. If these formate species
are not removed, then carbon may build up in the Al2O3 ALD
films. These residual formate species that are not removed by
TMA may explain the high carbon concentration observed in
Al2O3 ALD films grown with TMA and ozone at lower
temperatures.13
IV. Conclusions
In situ transmission FTIR spectroscopy and QMS were used
to study Al2O3 ALD with sequential exposures of Al(CH3)3 and
O3. The FTIR spectroscopy studies of the surface reactions
occurring during Al2O3 ALD were performed at temperatures
from 363 to 650 K. The FTIR spectra were recorded after
Al(CH3)3 and ozone exposures. These FTIR spectra showed that
the ozone exposure removes surface AlCH3* species and
produces AlOCH3* (methoxy), Al(OCHO)* (formate), Al(O-

Goldstein et al.
COOH)* (carbonate), and AlOH* (hydroxyl) species. The TMA
exposure then removes the methoxy, formate, carbonate, and
hydroxyl species and reestablishes the AlCH3* species. The
identification of formate species in the FTIR spectra was
confirmed by separate formaldehyde adsorption experiments on
Al2O3 ALD surfaces. The formate species were temperature
dependent and were nearly absent from the FTIR spectra at
g650 K.
Repeating the TMA and O3 exposures in a sequential reaction
sequence progressively deposited an Al2O3 ALD film as
monitored by the absorbance increase for bulk Al2O3 infrared
features. TEM was used to examine the Al2O3 ALD films
deposited on the ZrO2 particles. The Al2O3 ALD films were
very conformal to the underlying ZrO2 particles. These TEM
images were consistent with an Al2O3 ALD growth rate of 1.1
/cycle at 550 K. QMS analysis of the gas phase species
revealed that the TMA reaction produced CH4. The ozone
reaction then produced mainly CH4 with a small amount of C2H4
(ethylene), CO, and CO2.
On the basis of the surface species observed by the FTIR
studies and the gas phase species monitored by the QMS
investigations, a very different mechanism is suggested for Al2O3
ALD growth using TMA and O3 compared with Al2O3 ALD
using TMA and H2O. During Al2O3 ALD using ozone, both
Al-C and C-H bond insertion occurs when O3 reacts with
AlCH3* species to create methoxy and AlCH2OH* species. The
methoxy species decompose to formate and carbonate species.
The formate and carbonate species release CO and CO2 to
produce AlOH* species. Two AlCH2OH* species also eliminate
C2H4 and produce two AlOH* species.
Acknowledgment. This work was supported by the National
Science Foundation (CHE-0715552). Some of the equipment
used in this research was provided by the Air Force Office of
Scientific Research. The authors thank Dr. Thomas Giddings
for assistance with the TEM analysis.
References and Notes
(1) George, S. M.; Ott, A. W.; Klaus, J. W. J. Phys. Chem. 1996, 100,
13121.
(2) Ritala, M.; Leskela, M. In Handbook of Thin Film Materials; Nalwa,
H. S., Ed.; Academic Press: San Diego, CA, 2001; Vol. 1, p 129.
(3) Dillon, A. C.; Ott, A. W.; Way, J. D.; George, S. M. Surf. Sci.
1995, 322, 230.
(4) Ott, A. W.; Klaus, J. W.; Johnson, J. M.; George, S. M. Thin Solid
Films 1997, 292, 135.
(5) Ott, A. W.; McCarley, K. C.; Klaus, J. W.; Way, J. D.; George,
S. M. Appl. Surf. Sci. 1996, 107, 128.
(6) Puurunen, R. L. J. Appl. Phys. 2005, 97, 121301.
(7) Wilson, C. A.; Grubbs, R. K.; George, S. M. Chem. Mater. 2005,
17, 5625.
(8) Elam, J. W.; Routkevitch, D.; Mardilovich, P. P.; George, S. M.
Chem. Mater. 2003, 15, 3507.
(9) McCormick, J. A.; Rice, K. P.; Paul, D. F.; Weimer, A. W.; George,
S. M. Chem. Vap. Deposition 2007, 13, 491.
(10) Groner, M. D.; Fabreguette, F. H.; Elam, J. W.; George, S. M.
Chem. Mater. 2004, 16, 639.
(11) Elam, J. W.; Groner, M. D.; George, S. M. ReV. Sci. Instrum. 2002,
73, 2981.
(12) Ferguson, J. D.; Weimer, A. W.; George, S. M. Thin Solid Films
2000, 371, 95.
(13) Kim, S. K.; Lee, S. W.; Hwang, C. S.; Min, Y. S.; Won, J. Y.;
Jeong, J. J. Electrochem. Soc. 2006, 153, F69.
(14) Kim, S. K.; Hwang, C. S. J. Appl. Phys. 2004, 96, 2323.
(15) Kim, J. B.; Kwon, D. R.; Chakrabarti, K.; Lee, C.; Oh, K. Y.; Lee,
J. H. J. Appl. Phys. 2002, 92, 6739.
(16) Avice, M.; Grossner, U.; Monakhov, E. V.; Grillenberger, J.; Nilsen,
O.; Fjellvag, H.; Svensson, B. G. Mater. Sci. Forum 2005, 483-485, 705.
(17) Elliott, S. D.; Scarel, G.; Wiemer, C.; Fanciulli, M.; Pavia, G. Chem.
Mater. 2006, 18, 3764.
(18) Lee, T.-P.; Jang, C.; Haselden, B.; Dong, M. J. Vac. Sci. Technol.
B 2004, 22, 2295.

Al2O3 Atomic Layer Deposition


(19) Jakschik, S.; Schroeder, U.; Hecht, T.; Krueger, D.; Dollinger, G.;
Bergmaier, A.; Luhmann, C.; Bartha, J. W. Appl. Surf. Sci. 2003, 211, 352.
(20) Ferguson, J. D.; Weimer, A. W.; George, S. M. J. Vac. Sci. Technol.
A 2004, 22, 118.
(21) Ballinger, T. H.; Wong, J. C. S.; Yates, J. T. Langmuir 1992, 8,
1676.
(22) McCormick, J. A.; Cloutier, B. L.; Weimer, A. W.; George, S. M.
J. Vac. Sci. Technol. A 2007, 25, 67.
(23) Benfer, S.; Knozinger, E. J. Mater. Chem. 1999, 9, 1203.
(24) Goldstein, D. N.; George, S. M. 6th International Conference on
Atomic Layer Deposition, (ALD2006), Seoul, Korea, July 25, 2006.
(25) Busca, G.; Lorenzelli, V. Mater. Chem. 1982, 7, 89.
(26) Greenler, R. G. J. Chem. Phys. 1962, 37, 2094.
(27) Carlos-Cuellar, S.; Li, P.; Christensen, A. P.; Krueger, B. J.;
Burrichter, C.; Grassian, V. H. J. Phys. Chem. A 2003, 107, 4250.
(28) Busca, G. Catal. Today 1996, 27, 323.
(29) McInroy, A. R.; Lundie, D. T.; Winfield, J. M.; Dudman, C. C.;
Jones, P.; Lennon, D. Langmuir 2005, 21, 11092.
(30) Kwon, J. H.; Dai, M.; Halls, M. D.; Chabal, Y. J. Chem. Mater.
2008, 20, 3248.

J. Phys. Chem. C, Vol. 112, No. 49, 2008 19539


(31) Kim, S. K.; Hwang, C. S. J. Appl. Phys. 2004, 96, 2323.
(32) Busca, G.; Lamotte, J.; Lavalley, J. C.; Lorenzelli, V. J. Am. Chem.
Soc. 1987, 109, 5197.
(33) Frank, M.; Wolter, K.; Magg, N.; Heemeier, M.; Kuhnemuth, R.;
Baumer, M.; Freund, H. J. Surf. Sci. 2001, 492, 270.
(34) Lundie, D. T.; McInroy, A. R.; Marshall, R.; Winfield, J. M.; Jones,
P.; Dudman, C. C.; Parker, S. F.; Mitchell, C.; Lennon, D. J. Phys. Chem.
B 2005, 109, 11592.
(35) Chesters, M. A.; McCash, E. M. Spectrochim. Acta A 1987, 43,
1625.
(36) Juppo, M.; Rahtu, A.; Ritala, M.; Leskela, M. Langmuir 2000, 16,
4034.
(37) Prechtl, G.; Kersch, A.; Icking-Konert, G. S.; Jacobs, W.; Hect,
T.; Boubekeur, H.; Schroder, U. Technical Digest of the IEEE International
Electron DeVices Meeting (IEDM); IEEE: Piscataway, NJ, 2003; p 9.6.1.
(38) Gopal, P. G.; Schneider, R. L.; Watters, K. L. J. Catal. 1987, 105,
366.

JP804296A

Das könnte Ihnen auch gefallen