Bakke 2010
Bakke 2010
Pure, polycrystalline CdS deposited by atomic layer deposition (ALD) on Si(100) or glass using
dimethyl cadmium and in situ generated H2S is investigated in detail. This ALD system follows
saturation behavior typical of ALD systems, and the growth rate monotonically decreases with
temperature from 100 C-300 C; by 400 C linear growth rate behavior is no longer seen. The crystal
structure as determined by X-ray diffraction and transmission electron microscopy gradually
transitions from zincblende to wurtzite with increasing temperature until the film is primarily
wurtzite by 400 C. Further, the average grain size increases with temperature. Transmission electron
microscopy images and selected area diffraction patterns confirm the presence of zincblende and
wurtzite crystals because of stacking faults and demonstrate that {111} crystal planes are more
oriented parallel to the substrate at lower temperatures. Ultraviolet-visible spectroscopy shows that
the bandgap is 2.3-2.42 eV in the 100 C-400 C range with a slight increase occurring with
temperature. The roughness of the films is found to increase both with temperature and cycle number
as observed with atomic force microscopy and scanning electron microscopy. Density functional
theory calculations were used to understand observations concerning the growth rate and the
bandgap of the films deposited at different temperatures.
ALE using a molecular beam epitaxy (MBE) setup at spectroscopy (XPS), and transmission electron microscopy (TEM)
340 C. This system produced epitaxial c(200) CdS with measurements was a Czochralski grown n-type Si(100) wafer with
stacking faults and dislocations.8 Luo et al. deposited CdS by a resistivity of 1-5 Ω-cm and a native silicon oxide layer which is
ALE at room temperature on ZnSe(100) using a UHV approximately 2 nm thick. Films were also deposited on glass for
system as well.9 The precursors dimethyl cadmium and H2S comparison of crystal orientation and bandgap using X-ray diffrac-
tion (XRD) and ultraviolet-visible spectroscopy (UV-vis), res-
were used to deposit up to 15 bilayers of stoichiometric CdS.
pectively. The substrates were cleaned via 5 min sonication steps
Annealing at 250 C produced zincblende CdS, which
each in acetone and ethanol with deionized water rinses between
matched the substrate crystal structure. The reaction mecha- each solvent bath. Residual organics were then removed with a
nism was also studied in detail, and the results showed that piranha clean (70% sulfuric acid and 30% hydrogen peroxide) for
the DMCd precursor etched the surface to release DMZn. 15 min followed by a deionized water rinse and drying with
Another similar II-VI material, CdSe has been grown via pressurized air. Fresh piranha is hot and corrosive, and extreme
ALE using cadmium and selenium elemental sources.15 caution must be used when handling this solution.
In this study, we detail the deposition of CdS via ALD After ALD, resulting film thicknesses were measured by a
on hydroxyl terminated surfaces of Si(100) or glass using Gaertner L116C single-wavelength ellipsometer using 632.8 nm
DMCd and an in situ H2S source in an ALD flow reactor. light at a 70 angle of incidence and with the polarizer set to 45.
In depth characterization of the growth rates of the films At least two 1 cm 1 cm samples from each run were used, and
measurements were performed on three spots for each sample to
is performed, and DFT calculations are used to support
account for any non-uniformity. Elemental composition of the
observed growth rates. The material properties of the
films was determined by XPS with a Surface Science Instru-
films including the crystal structure, bandgap, roughness, ments S-Probe monochromatized spectrometer using Al KR
and surface morphology are carefully documented. In 1486 eV radiation at a pressure of 6.7 10-11 kPa (5 10-10
addition, we make direct comparisons between CdS ALD Torr). The Argon sputter was performed at a pressure of 1.3
and the related ZnS ALD processes. ZnS is another II-VI 10-8 kPa (1 10-7 Torr) using 5 keV Arþ at 2 mm 2 mm raster
semiconductor which has been studied more extensively at 45 incidence to the sample. Survey scans were performed
when deposited by ALD. Analogous precursors dimethyl with a step size of 1 eV. Surface morphology and roughness were
zinc (DMZn)16 and diethyl zinc (DEZn)17-19 have been characterized with a Veeco Multimode AFM in tapping mode
used with the counter reactant H2S to yield ZnS films, and with MikroMasch NSC16 tips. Crystal structure was deter-
density functional theory calculations have also been mined with a PANalytical X’Pert PRO XRD system in parallel
beam mode using Cu KR radiation. Surfaces were imaged by
carried out to explore the growth mechanisms of ZnS.20
SEM using an FEI XL30 Sirion SEM at a 5 kV operating
Hence, it serves as an ideal basis for comparison.
voltage. Bandgaps were determined by UV-vis with Varian’s
Cary 6000i UV-vis-NIR spectrophotometer. TEM samples
Experimental and Computational Details with a thickness of ∼80 nm were prepared using a focused ion
ALD growth of CdS was performed in a custom built warm beam (FIB, FEI Strata 235DB dual-beam FIB/SEM) lift-out
wall reactor which has been described in a previous publication.17 Omniprobe technique with a Gaþ ion beam at 30 keV. Cross-
The precursors for this reaction were dimethyl cadmium (DMCd) sectional bright-field and high resolution transmission electron
(Strem)14 and in situ generated H2S which is produced by heating microscopy images and selected area diffraction (SAD) patterns
thioacetamide to 150 C under an inert atmosphere; this method were taken by an FEI Tecnai G2 F20 X-TWIN operated at an
has been previously described in detail.17 DMCd is a liquid at accelerating voltage of 200 kV. Fast Fourier Transformation
room temperature with a melting point of -4.5 C and a boiling (FFT) images were obtained by DigitalMicrograph from the
point of 105.5 C at atmospheric pressure. DMCd is a highly high resolution TEM (HRTEM)
reactive chemical and should only be handled in a glove box free Periodic hybrid density functional theory (DFT) calcula-
of water and oxygen. The H2S is mostly pure with small amounts tions were performed to investigate the electronic character-
of acetonitrile. Nitrogen was used as the carrier and purge gas at a istics of the deposited films and the temperature dependence
constant flow rate of 60 sccm. The substrate temperature was of the growth rate. The calculations were carried out by
varied for the deposition of CdS with the temperature affecting CRYSTAL200621 quantum chemistry software using the PBE0
parameters such as growth rate, crystal structure, root-mean- density functional,22 which is one of the better performing hybrid
square (rms) roughness, and bandgap. For all studies except pulse functionals for solid state and for late-transition-metal reac-
and purge length dependencies, the optimized cycle of ALD tions.23,24 Default optimization convergence thresholds and
consisted of 0.4 s DMCd, 10 s purge, 0.4 s H2S, and a 10 s purge. an extra large integration grid were utilized in the calcula-
All precursors were maintained at room temperature (∼22 C). tions. Karlsruhe split-valence basis set with polarization func-
Needle valves controlled the rate of dosing of each precursor. tions (def2-SVP)25 was used for C and H, while a standard
The substrate for ellipsometry, scanning electron microscopy 6-31G* basis was adopted for S. For Cd a triple valence quality
(SEM), atomic force microscopy (AFM), X-ray photoelectron
(21) Dovesi, R. S.; Saunders, V. R.; Roetti, C.; Orlando, R.; Zicovich-
(15) Mikhaevich, D. P.; Ezhovskii, Y. K. Russ. J. Appl. Chem. 2003, 76, Wilson, C. M.; Pascale, F.; Civalleri, B.; Doll, K.; Harrison, N. M.;
1197–1200. Bush, I. J.; D’Arco, Ph.; Llunell, M. CRYSTAL2006 User’s
(16) Hunter, A.; Kitai, A. H. J. Cryst. Growth 1988, 91, 111–118. Manual; University of Torino: Torino, Italy, 2006.
(17) Bakke, J. R.; King, J. S.; Jung, H. J.; Sinclair, R.; Bent, S. F. Thin (22) Adamo, C.; Barone, V. J. Chem. Phys. 1999, 110, 6158–6170.
Solid Films 2010, 518, 5400–5408. (23) Paier, J.; Marsman, M.; Kresse, G. J. Chem. Phys. 2007, 127,
(18) Kim, Y. S.; Yun, S. J. Appl. Surf. Sci. 2004, 229, 105–111. 24103.
(19) Stuyven, G.; De Visschere, P.; Hikavyy, A.; Neyts, K. J. Cryst. (24) Quintal, M. M.; Karton, A.; Iron, M. A.; Boese, A. D.; Martin,
Growth 2002, 234, 690–698. J. M. L. J. Phys. Chem. A 2006, 110, 709–716.
(20) Tanskanen, J. T.; Bakke, J. R.; Bent, S. F.; Pakkanen, T. A. (25) Weigend, F.; Ahlrichs, R. Phys. Chem. Chem. Phys. 2005, 7, 3297–
Langmuir 2010, 26, 11899-11906. 3305.
Article Chem. Mater., Vol. 22, No. 16, 2010 4671
9-7-6-311-d631G basis26,27 with the outermost exponent vari- Figure 2. Film thickness as a function of temperature for CdS ALD.
ationally optimized for bulk CdS (0.1739 -> 0.1797) was used.
A modified basis set was necessary for the periodic calculations, Using the optimized operating parameters at 150 C
and the basis reproduced the characteristics of bulk CdS well. described above, the behavior of the CdS deposition
As an example, the optimized lattice parameter of zincblende process was tested over a temperature range of 100 C-
CdS was within 2% of the experimental value. The active 400 C. Figure 2 demonstrates that the growth is linear
growth surface was modeled by a four atomic layer thick for the 100 C-300 C range. The growth rates obtained
hydrogenated S-terminated (111) slab of zincblende CdS, and by linear regression decrease with temperature, and
the outermost atomic layer of the slab was allowed to relax in DFT calculations provide an explanation for this effect
the optimizations. Frequency calculations were performed to
(vide infra). The growth rate monotonically decreases
investigate the effect of thermodynamics on the film growth
characteristics. The electronic characteristics of the deposited
from approximately 1.96 Å/cycle at 100 C to 0.72 Å/
films were analyzed on the basis of PBE0-calculated density of cycle at 300 C. These growth rates are equivalent to 0.58
states-plots (DOS) and by using 2 2 2 and 3 3 3 CdS monolayers/cycle and 0.21 monolayers/cycle, respec-
supercells with a stoichiometry of Cd8S8 and Cd27S27, respec- tively, when referenced to the c(111) interplanar distance
tively. The expansion of the DOS plots was performed using 18 of 3.36 Å. At 300 C a significant portion of the film has
Legendre polynomials. the h(103) phase, and the growth rate is 0.38 monolayers/
cycle using its interplanar distance of 1.9 Å; consequently,
Results and Discussion the actual growth rate at 300 C is in the range of 0.21-
0.38 monolayers/cycle. No incubation period is evident
1. ALD Growth Characterization. CdS growth rates as
from the growth rate studies indicating that nuclea-
a function of precursor pulse time and nitrogen purge
tion occurs fairly readily on the SiO2/Si(100) substrate;
time were studied, keeping all other process parameters
nevertheless, further experiments specifically focusing on
constant. The resulting growth rate curves are shown in
nucleation are of interest for future studies.
Figure 1 as a function of DMCd pulse time and N2 purge
The boiling point of DMCd is 105.5 C, and an
time, respectively. The data in Figure 1 show that satura-
increased residence time in the reactor at lower tempera-
tion behavior, which is a characteristic of ALD, with
tures may account for the significantly increased growth
respect to DMCd is reached by 0.4 s, and the growth is
rate at 100 C because the remaining reactant could react
self-limiting. Also shown, after 10 s of purging with N2,
with H2S during the following pulse, leading to a small
excess reactants have been removed. The growth rate
CVD component. However, as shown by Groner et al.
dependence on H2S pulse time was not measured since
for deposition of Al2O3 using trimethylaluminum at low
the gas was delivered in excess to account for a decrease in
temperatures, the growth rate can be held constant
the pressure of H2S gas during deposition. Unless other-
by increasing the purge time at lower temperatures.28 At
wise stated, experiments in this paper were performed at
400 C, the growth rate does not display the linear behavior
the optimum conditions for 150 C, which is a pulse
expected from ALD: the growth rate increases with cycle
sequence of 0.4 s DMCd, 10 s N2 purge, 0.4 s H2S, and
number and is not uniform across a 2 in. sample range.
10 s N2 purge with a total nitrogen flow rate of 60 sccm. A
This observation is likely due to instability or decomposition
purge time of around 7 s is sufficient in this reactor for
of the ligands. Thus, true ALD does not occur at 400 C, but
ALD growth; however, a 10 s purge time was used to
pure CdS is still deposited without contamination from
ensure that no CVD component is incorporated during
ligands. The cross-section TEM images in Figure 3a confirm
the detailed study of ALD growth rate and properties.
the growth rate for the 150 C film. The thickest portion of
(26) CRYSTAL basis set library at http://www.crystal.unito.it/Basis_ the 400 C film was analyzed by TEM, and it verifies that the
Sets/Ptable.html. film is very rough.
(27) Dou, Y.; Egdell, R. G.; Law, D. S. L.; Harrison, N. M.; Searle, B.
G. An experimental and theoretical investigation of the electronic
structure of CdO. J. Phys.: Condens. Matter 1998, 10(38), 8447– (28) Groner, M. D.; Fabreguette, F. H.; Elam, J. W.; George, S. M.
8458. Chem. Mater. 2004, 16, 639–645.
4672 Chem. Mater., Vol. 22, No. 16, 2010 Bakke et al.
Figure 3. TEM images of 500 cycles of CdS at (a) 150 C and (b) 400 C. Figure 4. PBE0-calculated dissociative DMCd chemisorption energies
per precursor (ΔG, 1 atm) on hydrogenated S-terminated (111) surface of
zincblende CdS for a precursor fractional surface coverage of 0.25, 0.50,
The growth rate curves obtained for CdS display many and 1.00 as a function of temperature.
similarities to the diethylzinc and H2S ALD process.17
The ZnS system displays a growth rate that monotoni- monolayer with respect to the lower fractional coverages,
cally decreases with temperature from a rate of 0.44 and thus entropy increases at an increasing rate for the lower
monolayers/cycle at 100 C to 0.24 monolayers/cycle coverages with temperature. (2) In addition, the effective size
at 300 C (compared to 0.58 monolayers/cycle and 0.21 of the chemisorbed species increases with temperature be-
monolayers/cycle for CdS). Moreover, density functional cause of the increased atomic motion, which leads to more
theory calculations by Tanskanen et al.20 suggest that a significant ligand repulsion at higher coverages and to the
higher growth rate occurs with DMZn compared to observed preference for the less densely packed surfaces at
DEZn. Consequently, the increased growth rate in terms higher temperatures.
of Å/cycle is not unexpected compared to the DEZn 2. Film Characterization. XPS was used to analyze the
system because of the increased lattice constant and the composition of the ALD films and shows that the films
difference in ligands. Also, neither system shows a “tradi- are stoichiometric to within the error of the XPS method;
tional” ALD window in which the growth rate is flat.17,19 the chemical composition as calculated from XPS data
To understand the origin of the decreasing growth rate as is bounded by a few atomic percent because of the many
a function of increasing temperature, the film growth char- sources of error inherent in quantitative analysis by
acteristics were investigated by periodic PBE0 calculations. XPS;30-32 moreover, the cross-section (sensitivity) for
We focused on studying the reaction energies for the dis- Cd in XPS is much larger than for S.33 Thus, films
sociative chemisorption of DMCd according to the reaction deposited above 150 C are likely stoichiometric as the
mechanism SH*(s) þ Cd(CH3)2(g) f S-Cd-CH3*(s) þ bandgap and crystal lattice parameters match that of pure
CH4(g), which has been suggested to be the major pathway CdS (vide infra). Films deposited at 100 C display
of DMCd dissociation with a hydrogenated sulfur surface behaviors which may be due to sulfur vacancies or
where “*” refers to the surface species.29 The SH*(s) growth hydrogen impurities (vide infra) and are not unexpected
surface was represented by a hydrogenated (111) slab of from literature.28,34 Figure 5 shows an XPS spectrum of a
cubic CdS, and precursor surface coverages of 0.25, 0.50, and film deposited at 200 C: only Cd and S are visible in the
1.00 monolayers were considered. The calculated Gibbs- film. Similar to the ZnS deposition in a previous study,17
corrected reaction energies at temperatures between 50 and no noticeable carbon or nitrogen from the acetonitrile
400 C are illustrated in Figure 4. The energies are given per byproduct of the in situ H2S generation is incorporated
mole of precursor to allow direct comparison of different into the film. No contaminants were seen in any analyzed
fractional coverages. At low temperatures, the DMCd dis- samples in the temperature range of 100 C-400 C, and
sociation energies for the different coverages are within 5 the composition did not change with temperature. Final-
kcal/mol of each other. On the other hand, low-coverage ly, the composition was analyzed at various sputter times,
systems become clearly favored over a monolayer coverage and the composition was constant, confirming that the
as a function of increasing temperature so that by 400 C the sputtering did not have a noticeable effect on the mea-
surface coverages of 0.25 and 0.50 are favored over the full sured composition.
coverage system by approximately 10 kcal/mol per chemi- CdS is well-known to exist in either the zincblende
sorbed DMCd. As a consequence, the slow growth rates (cubic) or the wurtzite (hexagonal) crystal struc-
at high temperatures are attributed to partial surface cover- tures, and the phase strongly depends upon factors
age by the metal precursor because of steric hindrance at
(30) Analytical Instrumentation Handbook; Marcel Denker: Las Vegas,
highly covered surfaces. The calculated energetic trends NM, 1990.
originate primarily from (1) entropy and (2) steric effects: (31) Ohring, M., The Materials Science of Thin Films; Academic Press
(1) Namely, atomic motion, which increases as a function Limited: New York, 1992.
(32) Wagner, C. D. Anal. Chem. 1977, 49, 1282–1290.
of temperature, is more limited for the tightly packed (33) Wagner, C. D.; Davis, L. E.; Zeller, M. V.; Taylor, J. A.; Raymond,
R. H.; Gale, L. H. Surf. Interface Anal. 1981, 3, 211–225.
(34) Weber, M.; Krauser, J.; Weidinger, A.; Bruns, J.; Fischer, C. H.;
(29) Han, M.; Luo, Y.; Moryl, J. E.; Osgood, R. M. Surf. Sci. 1999, 425, Bohne, W.; Rohrich, J.; Scheer, R. J. Electrochem. Soc. 1999, 146,
259–275. 2131–2138.
Article Chem. Mater., Vol. 22, No. 16, 2010 4673
Table 1. Summary of CdS XRD Intensities Relative to the h(002)/c(111) Peak and a Comparison to an Analogous ZnS ALD Processa
400 C 300 C 225 C 150 C 100 C
cubic hexagonal CdS ZnS CdS ZnS CdS ZnS CdS ZnS CdS ZnS
100 0.14
111 002 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00
101 1.20 0.14
200 0.02
102 0.30 0.13
220 110 0.06 0.05 0.01 0.003 0.02 0.03
103 0.41 0.06 0.44 0.07 0.13 0.003 0.06 0.002 0.08
311 112 0.23 0.10 0.14 0.02 0.003
201 0.07
222 004 0.08 0.01 0.03 0.01 0.03 0.01 0.05
202 0.06
%h(002)/c(111) 29% 79% 52% 90% 86% 98% 90% 99% 86% 100%
a
The %h(002)/c(111) row is the percentage of total intensity belonging to that peak.
Figure 7. Grain sizes of the main grains attributed to the principal cubic
and wurtzite peaks. The solid markers are measured after a 500 C anneal
in H2S for 10 h.
Figure 9. High resolution TEM image of 150 C CdS film: (a) overview of a grain containing stacking faults and (b) magnified view of the region in the
white box in (a). Intrinsic and extrinsic stacking faults (SFs) form by a/6 Æ211æ type displacement of plane position on zincblende (111) planes to form local
wurtzite grains. Closely spaced SFs may construct a larger wurtzite structure.
and wurtzite were assigned on the basis of the previously spots in Figure 8d as opposed to Figure 8b presumably
reported lattice parameters36,38 by the relationship of indicates a higher portion of wurtzite in that local area.
Rd = Lλ = fixed constant, which is the camera length An HRTEM image of the 400 C film and its FFT image
equation of a TEM diffraction pattern, where R is the (not shown) elucidate that both zincblende and wurtzite
distance from the center spot to a specific spot corre- CdS are present. The SAD pattern matches the FFT
sponding to a specific plane, d is the interplanar spacing image in that the c(111) plane is coincident with the
between the specific planes, L is the camera length, and h(002) plane.
λ is 2.51 pm, which is the wavelength of the electron beam A high resolution TEM image of 150 C CdS film
with a 200 kV acceleration. The Si spots in the SAD (Figure 9a) confirms that wurtzite grains are present in
patterns of Figure 8a and Figure 8c were used as a zincblende grains because of stacking faults. A magnified
reference because Si has a well-known lattice parameter view of the HRTEM image in Figure 9b shows mixed
(aSi = 5.4308), and direct comparison of the spots bet- local zincblende and wurtzite form from a combination of
ween Si and CdS gives better verification of CdS spots. intrinsic and extrinsic stacking faults, and the stacking
Comparatively, the 400 C film does not show a strong sequence is clearly defined and labeled. c(111) and h(002)
alignment of the h(002)/c(111) planes and is primarily planes are stacked parallel and a/6 Æ211æ type displacement
wurtzite in character. The XRD patterns and Table 1 of zincblende (111) planes yields intrinsic and extrinsic
show that many crystal phases are visible because of the stacking faults which provide transformation from zinc-
random orientation. The SAD patterns such as those in blende (thermodynamically stable phase at 150 C) to
Figure 8c,d show that the crystals are randomly oriented wurtzite (metastable phase at 150 C). This indicates that
compared to the silicon substrate. Thus, the reduction in wurtzite AB stacking forms because of stacking faults from
peak intensities of the main peak as seen in XRD is due to zincblende ABC stacking. Finally, closely spaced stacking
both the increasing wurtzite content and the decreasing faults can construct a larger local wurtzite structure as seen
texture. The clear wurtzite diffraction array with higher in Figure 9b.
index planes is seen in the SAD pattern of the CdS region Optical characterization of the CdS films on glass was
in Figure 8d. c(111) and h(002) have the same diffraction performed with UV-vis to determine the bandgap. CdS
spot position because the ABC stacking direction of the is a direct bandgap semiconductor, and literature values
c(111) planes is the same as the AB stacking direction of of the bandgap range from 2.25 eV-2.45 eV depending
h(002) planes. This coincidence of c(111) and h(002) on deposition method, crystal structure, grain size, and
suggests the presence of stacking faults which local form quantum effects;1,7,12,36,39 however, the accepted band-
hexagonal stacking in the cubic structure. Random ori- gap for pure CdS is 2.42 eV.1 CdS quantum dots display
entation of the spots without coincidence would signify much larger bandgaps; however, they approach bulk
that wurtzite forms freely without orientation relative to values at >10 nm, which means significant quantum
zincblende; thus, the coincidence of those spots is con- confinement effects are not expected for these films based
sistent with wurtzite AB stacking resulting from stacking on the Scherrer equation calculations.40,41 In this work,
faults in the zincblende ABC stacking. Further, the
clearer and stronger intensity of the wurtzite diffraction (39) Zelayaangel, O.; Hernandez, L.; Demelo, O.; Alvaradogil, J. J.;
Lozadamorales, R.; Falcony, C.; Vargas, H.; Ramirezbon, R.
Vacuum 1995, 46, 1083–1085.
(38) Benkabou, F.; Aourag, H.; Certier, M. Mater. Chem. Phys. 2000, (40) Wang, Y.; Herron, N. Phys. Rev. B 1990, 42, 7253.
66, 10–16. (41) Weller, H. Angew. Chem., Int. Ed. Engl. 1993, 32, 41–53.
4676 Chem. Mater., Vol. 22, No. 16, 2010 Bakke et al.
Figure 12. rms roughness versus cycle number of CdS deposited by ALD
as determined by AFM. Films were deposited at 150 C on Si(100).