Sie sind auf Seite 1von 6

Thin Solid Films 496 (2006) 293 298

www.elsevier.com/locate/tsf

Chemical bath deposition of MoS2 thin film using ammonium


tetrathiomolybdate as a single source for molybdenum and sulphur
Poulomi Roy, Suneel Kumar Srivastava *
Inorganic Materials and Nanocomposite Laboratory, Department of Chemistry, Indian Institute of Technology, Kharagpur-721302, India
Received 18 January 2005; received in revised form 14 July 2005; accepted 26 August 2005
Available online 5 October 2005

Abstract
Molybdenum disulphide, MoS2, thin films have been deposited by chemical bath deposition method on glass and quartz substrate using
ammonium tetrathiomolybdate as a single source precursor for Mo and S and subjected to vacuum heat treatment at different temperatures. X-ray
diffraction of as-deposited film indicated its amorphous character and showed the development of poorly crystalline MoS2 thin film on increasing
annealing temperature. The film has been characterized by energy dispersive X-ray analysis, X-ray photoelectron spectroscopy, scanning electron
micrograph and the optical properties also have been studied.
D 2005 Elsevier B.V. All rights reserved.
PACS: 78.55.Hx; 78.40.Fy; 78.20.-e
Keywords: Molybdenum disulphide; Deposition process; X-ray diffraction; X-ray photoelectron spectroscopy

1. Introduction
In the past few years, studies of materials with layered
structures such as transition metal dichalcogenides [1,2], quasi
ternary systems of the type A2X3 M2X3 MVX (A = Ga, In;
M = trivalent metal; MV = divalent metal; X = S, Se) [3,4], metal
oxychlorides [5], graphite [6], clay minerals [7,8] etc. have
received an ever increasing attention. This is mainly because of
their interesting anisotropic behaviour, great diversity in their
other physical properties and their usefulness for various
applications. Among all these, type transition metal dichalcogenides (TX2; T = transition metal of group IVB, VB and VIB,
X = chalcogen, i.e., S, Se and Te), molybdenum and tungsten
dichalcogenides constitutes structurally and chemically a well
defined family of compounds. MoS2, MoSe2, WS2, WSe2 [9
11] appear to be very promising semiconducting materials for
various applications such as photoactive materials in solar
energy conversion purpose because of the main advantage
associated with the prevention of electrolyte corrosion [12],
cathode in solid state secondary lithium batteries due to its
ability to intercalate with lithium ions [13], hydrodesulphurization catalysts [14] and solid lubricants for tribological applica* Corresponding author. Fax: +91 3222 255303.
E-mail address: sunil111954@yahoo.co.uk (S.K. Srivastava).
0040-6090/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.tsf.2005.08.368

tions in high temperature and vacuum environments where the


use of traditional liquid lubricants becomes ineffective or cannot
be tolerated [15]. MoS2 crystallizes in three different types of
structures, namely 2H, 3R and 1T-MoS2 [16]. The lattice
structure consists of a monolayer hexagonal sheet of Mo atoms,
which lies in between two hexagonal sheets of S atoms, i.e. (S
Mo S) layers and are held together strongly by covalent
bonding. In the hexagonal poly-type (2H-MoS2) two such
layers form a unit cell. The interaction between (S Mo S)
layers in the direction [00l] is loosely bound to each other only
by weak van der Waals forces. Such a layered crystallographic
arrangement allows the MoS2 layers to easily shear between
basal planes and is responsible for its excellent lubricity. In
addition to the crystalline MoS2, there also exist three types of
MoS2 structure based on XRD patterns, namely amorphous,
single layer and poorly crystalline MoS2.
For economic reasons from large applications point of view,
obtaining MoS2, MoSe2, and WS2 in thin films are interesting
while for ecological reasons, only sulphides are considered to
be more interesting. Thin films of MoS2 and WS2 have been
prepared as thin film by variety of other methods, such as
chemical bath deposition by drying or heating of chemical
solutions containing liquid Mo/W and S precursors [17,18],
pulse electrodeposition method [19,20], pulsed laser evaporation [21], sulphurization of metal [22], sputtering technique

294

P. Roy, S.K. Srivastava / Thin Solid Films 496 (2006) 293 298

[23], chemical vapour deposition (CVD) [24] and metal


organic CVD [25], spin coating method [26], sulphurization
of predeposited films where chalcogen is brought as elemental
sulphur, elemental selenium H2S or gas mixture containing
H2S while predeposited films are Mo [27] and textured MoTe2
film and polycrystalline MoTe2x Sx film [28].
In recent years, growth of MoS 2 thin films using
ammonium tetrathiomolybdate, (NH4)2MoS4, as a means of a
single source for molybdenum and sulphur have been reported
by electrochemical deposition techniques [29 33]. This is
mainly due to the fact that ammonium tetrathiomolybdate is
commercially available or can be readily prepared by the
reaction of molybdate and hydrogen sulphide. However, no
attempts have been made to deposit this film by any other
means using the same precursor. Among all the deposition
techniques, chemical bath deposition (CBD) method is one of
the most promising and has proved advantageous for
photovoltaic applications because it is an efficient, cost
effective and its suitability for making large area high quality
thin films. In the present work, we report for the first time the
deposition of MoS2 thin films on glass and quartz substrates by
CBD method using the (NH4)2MoS4 as a single precursor. The
film has been characterized by X-ray diffraction for structure
determination and particle size calculations, X-ray photoelectron spectroscopy (XPS), morphological features by scanning
electron micrograph (SEM) studies.

The use of hydrazine hydrate provides the electrons necessary


for reducing the MoVI to MoIV in the precursor similar to
electrochemical depositions where they are supplied by the
electrical current [29 33]. The whole solution was stirred well
to get a homogeneous solution. The pH of the solution was in
the range 10. The solution was transferred into a beaker and
was covered with a Teflon tape so that the ammonia cannot be
evaporated out otherwise the pH of the bulk solution cannot be
controlled. The glass substrate was inserted inside the beaker.
The bath temperature was maintained at 60 -C for 1 h. A
uniform gray-black thin film of thickness 0.6 Am was deposited
onto the substrate, which was washed by distilled water
ultrasonically until heterogeneities were removed. Finally the
deposited film having good adherent property was dried in
nitrogen atmosphere. The deposited films on glass and quartz
substrates were annealed under vacuum of pressure 1 Pa for 1
h at 400, 500 and 800 -C for, respectively. After annealing the
film showed metallic shining.
2.4. Characterization technique

2. Experimental

X-ray diffraction of samples was recorded by using a


Miniflex diffractometer (30 kV, 10 Maq; Rigaku Corp., Tokyo,
Japan) with Cu Ka source. The diffractograms were recorded
between angles 2h = 10- and 65- with a scan rate of 3- per min
at room temperature.
The average crystallite size (D) corresponding to 002
reflections has been calculated by using Scherrer equation as

2.1. Materials used

Ammonium paramolybdate, (NH4)6Mo7O24I4H2O (Merck)


was used for the preparation of (NH4)2MoS4. Hydrazine
hydrate (80% BDH) was used as a reducing agent during the
deposition of MoS2 thin film preparation. All the solutions
were prepared by dissolving appropriate quantity of materials
into de-ionized water.

where, the constant k is a shape factor and has been taken 0.9,
k is the wave length of X-ray (0.15418 nm), b 2h is FWHM in
radians and h is the Braggs angle.
The microstrain (e) developed in MoS2 film is calculated
from the relation:

2.2. Preparation of ammonium tetrathiomolybdate


(NH4)2MoS4
A solution of 5 gram of ammonium paramolybdate, in 15
ml of water was prepared and treated with 50 ml of ammonia
solution (d0.94). Then H2S was introduced. The solution first
turns yellow, later deep red, and after half an hour a cupious
quantity of the crystals, some of them well formed,
precipitate suddenly. The (NH4)2MoS4 so formed crystals
were washed with cold water, then alcohol, and dried in
vacuum [34].
2.3. Thin film deposition using ammonium tetrathiomolybdate
as single source precursor
MoS2 thin films were deposited on glass substrate by
chemical bath deposition technique. 0.2 M (NH4)2MoS4
solution was mixed with 10 ml of ammonia solution. 10 ml
of hydrazine hydrate (80%) was used here as a reducing agent.

kk
b2h cosh

b2h coth
:
4

The dislocation density (q) is estimated according to


Williamson and Smallman [35] using the relation:
q

15e
aD

where a is the lattice constant and D is the particle size.


SEM photograph of gold-coated films were recorded on a
JEOL JSEM-5800 at an acceleration voltage of 20 kV. Electron
microprobe analysis (OXFORD ISIS-300 model) was used to
determine the chemical composition of the films.
The thickness of the deposited film was measured by using
.
Dektak3 ST thickness profiler with accuracy T 10 A
XPS spectra were recorded on a Rigaku XPS 7000
spectrometer using non-monochromatic Mg K(alpha) (1253.6
eV) at a source power of 200 W. Chamber pressure during the
measurement was about 10 7 Pa. The binding energies were
referred to the adventitious C 1s peak at 284.6 eV. Data
acquisition treatment was realized using a computer and a

P. Roy, S.K. Srivastava / Thin Solid Films 496 (2006) 293 298

standard program. The quantitative studies were based on the


determination of Mo 3d, S 2p, C 1s and O 1s peak areas with
2.5, 0.125 and 0.6, respectively, as sensitivity factor provided
by the manufacturer. The wide scan spectra were taken at pass
energy of 50 eV, providing an instrumental resolution of 1 eV.
We observed that the Mo 3d, S 2p peaks are wide and
asymmetric with FWHM of 1.42 eV for Mo 3d5 / 2, 1.39 eV for
Mo 3d3 / 2 and 1.94 eV for S 2p.
Perkin-Elmer Lambda20 UV/VIS spectrometer ranging
from 300 to 1100 nm using a blank substrate as the reference
position examined the optical properties of deposited thin film.
3. Results and discussion
3.1. X-ray powder diffraction results
The deposited film at 60 -C is dark brownish in colour and
is markedly different in appearance to the gray, shiny MoS2.
The X-ray diffraction spectra of this film and when annealed at
400, 500 and 800 -C are shown in Fig. 1. The absence of the
peaks in Fig. 1(a) confirmed the amorphous in nature of the asdeposited film. On increasing the annealing temperature the
crystallinity increases as shown in Fig. 1(a) (d). The 002
interlayer peak of MoS2 is absent in as-deposited film whereas
it appeared in 400 and 500 -C annealed films as expected when
there are two or more S Mo S sandwich layers, otherwise the
interference from within the single sandwich layer makes an
insignificant contribution. That is the case with the samples in
Fig. 1(b) and (c), though, the position of 002 Bragg peaks and
their intensities are quite different from the crystalline MoS2. It
shows the presence of one high intensity peak at 13.79(d = 0.642 nm) at 400 -C and which shifted to 13.85- (d = 0.639
nm) at 500 -C. Interestingly, either of these values of interlayer
spacing not matched very well to that for 2H-MoS2 with a
value of 2h = 14.4- (d = 0.615 nm) [36]. It is also evident from
the Fig. 1(b), (c) that, the overall diffraction patterns of sample
heated at 400 and 500 -C temperatures are not sufficiently
resolved, i.e., an indication towards the formation of poorly

Fig. 1. X-ray diffractograms of MoS2 thin film (a) as deposited and when
annealed at (b) 400 (c) 500 and (d) 800 -C.

295

Table 1
Particle size (D), microstrain (e) and dislocation density (q) data for MoS2 film
annealed at different temperatures
Annealing
temperature (-C)

Particle size
(D) in nm

Dislocation density
(q) in cm 3

Microstrain
(e)

400
500
800

9.1
10.2
26.8

20  1011
16  1011
2.3  1011

3.79  10 3
3.37  10 3
1.29  10 3

crystalline form of MoS2. The as-deposited film when annealed


at 800 -C for 1 h on a quartz substrate, in Fig. 1(d), shows
peaks at 2h = 14.41-, 28.83- and 60.2- which corresponds to
the (002), (004) and (008) planes. It may be mentioned here
that a broad diffraction peak in the range of 2h = 18- 28-,
which begin to appear when the film is annealed at 500 -C, is
subsequently more developed when the film is annealed at 800
-C and could be attributed to the MoO3 (2h 27-)impurity.
Further, it is observed that the film thickness of MoS2 layer
decreases from 0.6 to 0.4 Am upon annealing at 800 -C.
Because of this, the X-rays scattered by the substrate are
absorbed to a much lesser extent by the film. As a result, the
broad peak could be due to both MoO3 as well as substrate
contribution. The (h0k), (hk0) planes are missing here shows
the films are textured with a basal (van der Waals) planes
parallel to the substrate. This property is of particular interest
where films are to be used for their photovoltaic cells or for
their tribological characteristics [29]. An additional peak
(d = 0.194 nm) of relatively low intensity observed on annealing the film at 800 -C corresponds to MoO3 [37]. The existing
literature also confirmed that in post annealed MoS2 films
deposited by various methods are invariably associated with
MoO3 in very minor quantity as additional phase since they are
very much air sensitive [29,38 41].
Table 1 records the particle size (D), microstrain (e) and
dislocation density (q) of the as-deposited film when annealed
at 400, 500 and 800 -C. It is observed that with increase in the
annealing temperature there is a growth of the crystallites. The
size values increases from 9.1 (400 -C) to 26.8 nm (800 -C).
The microstrain values as calculated from the broadening of the
profiles show that the strain values fall considerably with
annealing. It is well known that the smaller crystallites have
greater surface to volume ratio, thus giving rise increase in the
dislocation network, which occurs at the grain boundary
region. Hence as expected with increase in annealing temperature, there is a favourable condition for the crystallite growth.
The thermal energy provides the dislocation movements giving
rise to decrease in their dislocation density. As a consequence
of this, there is a fall in their microstrain values. Further it is
well known that the crystallite size (D) and microstrain (e)
values vary inversely, the proportionality constant depending
on the property of the material and the nature of the dislocation
[42]. In the present case, the product D I e has an almost
constant value. XRD profiles also show that the 002 peaks
become more and more symmetric with increasing annealing
temperature. The faults in the stacking of the planes in the films
annealed at 400 and 500 -C are expected to contribute towards
asymmetric broadening of the 002 peaks. At 800 -C annealing

296

P. Roy, S.K. Srivastava / Thin Solid Films 496 (2006) 293 298

temperature, the faults are expected to cause peak shifts and


symmetrical peak broadening in the diffraction patterns similar
to as observed in other cases [43].
A model similar to that of Wildervanck and Jellinek [44] has
been proposed which describes thermal behaviour of MoS2
film deposited as such on annealing at different temperatures
and is based on the appearance of lines resulting from the
crystallization of MoS2 in X-ray diffractograms. The crystallite
of MoS2 in as-deposited film may be regarded as randomly
oriented fragments of MoS2 layers [Fig. 2(a)]. Further
annealing of this film at higher temperatures results in their
growth by loss of sulphur and progressive orientation. The
crystallization process described is shown in Fig. 2(b) to (d).
3.2. Morphological characterization
The SEM of the annealed film (500 -C) on a glass substrate
is displayed in Fig. 3. It shows that the surface appears like
granular having different size and shapes of grains without any
cracks or holes generated on the surface. These grains are
distributed inhomogeneously and in some cases appeared as
agglomerated. The estimated average grain size is of few
hundred nanometers up to 500 nanometer.
3.3. Energy dispersive X-ray analysis
Typical EDX spectra of MoS2 films as deposited as well as
annealed (500 -C) are examined. The S / Mo atomic ratio in asdeposited film is 2.6, which becomes 1.89 when annealed at
500 -C. It appears that annealing of the film results in loss of
some material by evaporation and which can be attributed by
the lowering of film thickness, increasing density and
compactness of film and finally the change in composition.
Another possibility which would explain the S : Mo less than 2
is likely due to the presence of oxygen in the films where some
MoO42 exist together with MoS42 resulting in the formation
of a Mo O species (or a mixed Mo O S compound) [29].
However, the possibility of adsorption of water from the

(a)

(b)

(c)

Metal atom

Sulfur atom

The atom lie


in one plane

(d)

Fig. 3. Scanning electron micrograph of MoS2 film annealed at 500 -C.

solution during the deposition of the film cannot be ignored.


Formation of such non-stoichiometric MoS1.89 films other than
MoS2 in our case is similar to those as reported earlier for
MoS1.77, MoS1.85 [39], MoS1.85 [40] and MoS1.90 [41].
3.4. X-ray photoelectron spectroscopy study
Fig. 4(a), (b) represents the high resolution XPS spectrum in
Mo and S region of MoS2 film annealed at 500 -C. It shows the
presence of two highly intense peaks, separated by 3.0 eV,
indicating Mo 3d 3 / 2 at 231.9 eV and Mo 3d 5 / 2 at 228.9 eV also
having a shoulder peak at 226.1 eV because of S 2s in MoS2. A
small peak at 235.2 eV corresponds to the Mo 3d 3 / 2 for MoO3
[39]. A high intense peak is observed at 231.9 eV for both Mo
3d 3 / 2 of MoS2 and Mo 3d 5 / 2 of MoO3, as shown in Fig. 4(a).
It suggests that a small fraction of the MoS2 film is oxidized to
MoO3 due to air exposition of the film, an observation, which
further confirmed our X-ray findings earlier. The oxidized part
is shown by shaded portion [Fig. 4(a)]. Fig. 4(b) shows S 2p
doublet peak (2p 1 / 2 and 2p 3 / 2) at 161.9 eV in which 2p 1 / 2 and
2p 3 / 2 orbital peaks coincide with each other. It is to be noted
that the peak positions of S 2p and Mo 3d 5 / 2 for MoS2 in our
experiment is in good agreement with reference MoS2 powder
[39]. Very recently, a more general study on the variation of
MoS2 film stoichiometry versus (Mo 3d 5 / 2 S 2p 3 / 2) binding
energy has been reported [45], which corroborates the present
result. These findings further reaffirmed our contention that
films in particular the annealed one are associated with oxygen
as MoO3 further supporting our inference from EDX that
S : Mo is less than 2. The chemical shifts of Mo 3d 5 / 2 and S 2p
in MoS2 thin film is + 1.0 and  2.2 eV in comparison to
elemental Mo and S [46]. The electronegativity of S being 2.5
is greater than that of Mo being 1.8. Therefore there occurs
some partial electron transfer from molybdenum to sulfur. As a
result, molybdenum becomes positively charged as shown by
the increase in binding energy and sulfur becomes negatively
charged as shown by the decrease in binding energy.
3.5. Optical characterization

Fig. 2. Dependence of crystallite orientations of MoS2 film on annealing


showing (a) fragments of a layer of MoS2; (b) completely random orientation of
one layer crystallites; (c) stacking of the layers in random orientations about the
c-axis and (d) hexagonal MoS2 with stacking faults.

Fig. 5 shows the optical transmission spectra of films as


deposited as well as annealed at 500 -C. The annealed film
shows almost 80% transmission in the visible range of

P. Roy, S.K. Srivastava / Thin Solid Films 496 (2006) 293 298

(a)

297

Mo 3d5/2
of MoS2

Intensity (arb. unit)

(h)2x108 (cm-1eV)2

40
Mo 3d5/2
of MoO3
Mo 3d3/2
of MoS2

MoO3

Mo 3d3/2
of MoO3

35
30
25
20
15
10
5
0

S 2s

h (eV)
Fig. 6. Plots of (ahr)2 versus hm of the MoS2 film annealed at 500 -C.

220

225

230

235

The absorption coefficient (a) is related to the incident


photon en
n=2
K hm  Eg
5
a
hm

Binding Energy (eV)


(b)

Intensity (arb. unit)

S 2p

where, K is constant, E g is the energy gap and n is constant


equal to 1 for direct band gap. Fig. 6 shows the variation of
(ahm)2 versus hm for annealed films. The graph gave the
direct band gap values, obtained by extrapolating the linear
line on hm-axis is 1.74 eV and very close to the
characteristic direct band gap value (1.78 eV) for single
crystal MoS2 [47].
4. Conclusion
160

165

Binding Energy (eV)


Fig. 4. X-ray photoelectron spectra for MoS2 film annealed at 500 -C: (a) closeup survey at Mo-core and (b) close-up survey at S-core.

spectrum. The values of absorption coefficient are calculated


by using this equation.
It I0 exp  at

where, a is the optical absorption coefficient, t is the thickness


of the film, I t and I 0 are the intensity of initial light and
transmitted light, respectively.

% Transmittance

100
80
60

40
20

MoS2 films have been successfully deposited by chemical


bath deposition (CBD) method on a glass/quartz substrate
without stirring at 60 -C using ammonium tetrathiomolybdate
single precursor and hydrazine hydrate as a reducing agent at a
pH level 10. The deposited film showed good adherence
towards the substrate. The XRD patterns of annealed films
show the formation of poorly crystalline 2H-MoS2 with a
preferred orientation along c-axis. The crystallite size
increases while the microstrain values fall remarkably with
increasing the annealing temperature. Additionally, dislocation
density decreases because of more and more thermal energy
supplied. SEM photograph of annealed film shows that the
grain sizes are <500 nm. EDX data indicate that the films are
non-stoichiometric with small sulfur deficiency after annealing. XPS study analyses that the peak positions of Mo and S
are well matched with those of the characteristic MoS2 powder
having some contamination of MoO3 as impurity. The
composition calculated from XPS is also close to that
estimated from EDX analysis for the annealed film. The
annealed film shows about 80% transmission in the visible
range. The direct band gap value is of 1.74 eV, which agrees
well with the value characteristic MoS2 single crystal.

0
200

400

600

800

1000

1200

Wavelength (nm)
Fig. 5. Optical transmission spectra of the film (a) as deposited and (b) annealed
at 500 -C.

Acknowledgement
Authors are grateful to Professor J.C. Bernede, LPSE,
University of Nantes, Nantes, France for SEM of the samples.

298

P. Roy, S.K. Srivastava / Thin Solid Films 496 (2006) 293 298

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]

[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]

F. Hulliger, Struct. Bond. 4 (1968) 83.


S.K. Srivastava, Mater. Res. Bull. 26 (1991) 631.
H. Haeuseler, S.K. Srivastava, Z. Kristallogr. 215 (2000) 205.
S.K. Srivastava, M. Pramanik, D. Palit, H. Haueseler, Chem. Mater. 16
(2004) 4168.
R. Brec, Solid State Ionics 22 (1986) 3.
N. Bartlett, B.W. McQuilllan, in: M.S. Whittingham, A.J. Jacobson (Eds.),
Intercalated Chemistry, Academic Press, New york, 1982, p. 19.
G.W. Brindley, G. Brown (Eds.), Order Disorder in Clay Minerals. In
Crystal Structure of Clay Minerals and their X-ray Identification,
Mineralogical Society, London, 1980.
M. Pramanik, H. Acharya, S.K. Srivastava, Macromol. Mater. Eng. 289
(2004) 562.
J.A. Wilson, A.D. Yoffe, Adv. Phys. 18 (1969) 169.
S.K. Srivastava, B.N. Avasthi, S. Basu, J. Sci. Ind. Res. 41 (1982) 656.
S.K. Srivastava, B.N. Avasthi, J. Mater. Sci. 20 (1985) 3801.
H. Tributsch, Sol. Energy Mater. 1 (1979) 257.
M.S. Whittingha, in: F. Levy (Ed.), Intercalated Layered Materials, Reidel,
Dordrecht, 1979, p. 307.
S.K. Srivastava, B.N. Avasthi, J. Mater. Sci. 28 (1993) 5032.
G.V. Subba Rao, M.W. Schafer, in: F. Levy (Ed.), Intercalated Layered
Materials, Reidel, Dordrecht, 1979, p. 99.
R. Murray, B.L. Evans, J. Appl. Crystallogr. 12 (1979) 312.
P. Pramanik, S. Bhattacharya, Mater. Res. Bull. 35 (1990) 15.
S. Chandra, S.N. Sahu, J. Phys., D, Appl. Phys. 17 (1984) 2115.
S.M. Delphine, M. Jayachandran, C. Sanjeeviraja, Mater. Chem. Phys. 81
(2003) 78.
S.M. Delphine, M. Jayachandran, C. Sanjeeviraja, Mater. Res. Bull. 40
(2005) 135.
A.N. Zelikman, B.P. Lobashev, Y.V. Makarov, G.I. Sevostyanova, Inorg.
Mater. 12 (1976) 1367.
P.D. Fleischauer, Thin Solid Films 154 (1987) 309.
M. Regula, C. Ballif, J.H. Moser, F. Levy, Thin Solid Films 280 (1996)
67.
A.A. van Zomeren, J.H. Koegler, J. Schoonman, P.J. v.d. Put, Solid State
Ionics 53 56 (1992) 333.
A. Jager-Waldau, M. Lux-Steiner, R. Jager-Waldau, R. Burkhardt, E.
Bucher, Thin Solid Films 189 (1990) 339.

[26] J. Putz, M.A. Aegerter, Thin Solid Films 351 (1999) 119.
[27] C. Amory, J.C. Bernede, N. Hamdadou, Vacuum 72 (2004) 351.
[28] S.K. Srivastava, N. Guettari, J.C. Bernede, Solid State Commun. 132
(2004) 601.
[29] E.A. Ponomarev, M. Neumann-Spallart, G. Hodes, C. Levy-Clement,
Thin Solid Films 280 (1996) 86.
[30] E.A. Ponomarev, A. Albu-Yaron, R. Tenne, C. Levy-Clement, J.
Electrochem. Soc. 144 (1997) L277.
[31] E.A. Ponomarev, R. Tenne, A. Katty, C. Levy-Clement, Sol. Energy
Mater. Sol. Cells 52 (1998) 125.
[32] A. Albu-Yaron, C. Levy-Clement, J.L. Hutchison, Electrochem. SolidState Lett 2 (1999) 627.
[33] A. Albu-Yaron, C. Levy-Clement, A. Katty, S. Bastide, R. Tenne, Thin
Solid Films 361 362 (2000) 223.
[34] G. Kruss, Liebigs Ann. 225 (1884) 29.
[35] G.B. Williamson, R.C. Smallman, Philos. Mag. 1 (1956) 34.
[36] JCPDS-International Centre for Diffraction Data, PCPDFWIN, vol. 2.02,
p. 06-0097.
[37] JCPDS-International Centre for Diffraction Data, PCPDFWIN, vol. 2.02,
p. 05-0508.
[38] W. Jaegermann, D. Schmeisser, Surf. Sci. 165 (1986) 143.
[39] J. Pouzet, H. Hadouda, J.C. Bernede, R. Le Ny, Phys. Chem. Solids 57
(1996) 1363.
[40] S.E. Doyle, N. Mattern, W. Pitschke, G. Weise, D. Kraut, H.D. Bauer,
Thin Solid Films 245 (1994) 255.
[41] N. Barreau, J.C. Bernede, J. Phys. D: Appl. Phys. 35 (2002) 1197.
[42] D. Palit, S.K. Srivastava, M.C. Chakravorti, B.K. Samantaray, Mater.
Chem. Phys. 49 (1997) 22.
[43] Charles S. Barrett, T.B. Massalski, Structure of Metals, Crystallographic
Methods, Principles, and Data, 3rd eds., McGraw-Hill Book Company,
New York, 1968, p. 459.
[44] J.C. Wildervanck, F. Jellinek, Z. Anorg. Allg. Chem. 328 (1964) 309.
[45] M.A. Baker, R. Gilmore, C. Lenardi, W. Gissler, Appl. Surf. Sci. 150
(1999) 255.
[46] H.W. Wang, P. Skeldon, G.E. Thompson, Surf. Coat. Technol. 91 (1997)
200.
[47] E. Bucher, Photo voltaic Properties of Solid State Junctions of
Layered Semiconductors, in: A. Aruchany (Ed.), Photoelectrochemistry
and Photovoltaics of Layered Semiconductors, Kluwer, Dordrecht,
1992, p. 36.

Das könnte Ihnen auch gefallen