Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Colloids and Surfaces A: Physicochem. Eng.

Aspects 257–258 (2005) 277–282

Nanocrystalline CuO films prepared by pyrolysis of


Cu-arachidate LB multilayers
M. Parhizkara,1 , Sukhvinder Singha , P.K. Nayaka , Nigvendra Kumarb , K.P. Muthec ,
S.K. Guptac , R.S. Srinivasad , S.S. Talwara , S.S. Majora,∗
aDepartment of Physics, Indian Institute of Technology Bombay, Mumbai 400076, India
b Maharashtra College, Bombay University, Mumbai 400008, India
c TPPED, Bhabha Atomic Research Centre, Mumbai 400085, India
d Department of Metallurgical Engineering and Materials Science, Indian Institute of Technology Bombay, Mumbai 400076, India

Available online 18 November 2004

Abstract

Nanocrystalline CuO films have been formed by the pyrolysis of copper arachidate LB multilayers. FTIR and UV–vis spectroscopic
investigations showed that the pyrolysis was initiated at ∼100 ◦ C and was complete at ∼300 ◦ C. XPS and electron diffraction studies confirm
the films to be single-phase copper (II) oxide. Analysis of the bright-field images showed that the film consists of uniformly distributed particles
of size 2.5 ± 1.5 nm. The optical absorption edge of the CuO film did not show a measurable blue shift attributable to the nanocrystalline nature
of the films. The absorption features are analysed in the background of the uncertainty prevailing in the nature and value of the absorption
edge of CuO.
© 2004 Elsevier B.V. All rights reserved.

Keywords: CuO; Langmuir–Blodgett; Copper arachidate; Nanocrystalline; Structure

1. Introduction interesting size-dependent effects on the structure, optical,


electrical and magnetic behaviour of nanocrystalline CuO.
Cupric Oxide (CuO) is a semiconductor that has been Recently, very thin and homogenous metal oxide films
studied for photoconductive, photothermal and photoelectro- have been obtained by the oxidation of a precursor fatty acid
chemical applications [1–4]. It has also been known to be salt Langmuir–Blodgett (LB) multilayers [15–18]. In most
antiferromagnetic [5]. In recent years, it has attracted wide cases, the precursor LB multilayer is exposed to heat or UV
interest due to its monoclinic unit cell and square-planar co- radiation in air, resulting in oxide formation and removal of
ordination [6], which is similar to that in high Tc cuprate organic components. Using this method, very thin films of
superconductors. Owing to the current interest in nanomate- CuO could be obtained by UV exposure followed by heat
rials and their unique properties, the formation of nanocrys- treatment in air at 365 ◦ C of a precursor copper arachidate
talline CuO has been achieved by rapid precipitation [7], LB multilayer [19]. It was also reported [19] that a direct heat
spin coating [8], solid-state reaction [9], sonochemical re- treatment of the copper arachidate (CuA) multilayer resulted
action [10], sol–gel techniques [11], solvothermal route [12], in a discontinuous film. Elastic recoil detection analysis was
electrochemical route [13] and controlled oxidation of cop- used to calculate the Cu to O ratio.
per nanoparticles within silica gels [14]. These studies reveal In view of the increasing interest in transition metal ox-
ide films, in particular copper oxide, a controlled heat treat-

ment of CuA LB multilayers in oxygen ambient has been
Corresponding author. Tel.: +91 22 2576 7550; fax: +91 22 2576 7552.
E-mail address: syed@phy.iitb.ac.in (S.S. Major).
carried out in the present work. The process of conversion
1 Present address: Department of Physics, Tabriz University, Tabriz 51664, of the arachidate multilayer into CuO was studied by X-ray
Iran. diffraction, infrared spectroscopy and UV–vis spectroscopy.

0927-7757/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2004.10.029
278 M. Parhizkar et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 257–258 (2005) 277–282

The results show that the oxidation process and the removal
of organic components are completed at about 300 ◦ C. The
formation of a single-phase, nanocrystalline CuO film was
confirmed by transmission electron microscopy and X-ray
photoelectron spectroscopy studies.

2. Experimental

LB multilayers of copper arachidate were prepared by the


conventional LB deposition technique using a KSV 3000 in-
strument in a clean room. Arachidic (Aldrich, 99%) with
HPLC grade chloroform as solvent (1 mg/ml) was spread
on an aqueous subphase containing CuCl2 (10−4 M). Deion-
ized and ultra-filtered water (Millipore) having a resistiv-
ity of 18.2 M cm was used to prepare the subphase. The
subphase temperature was kept constant at 20 ◦ C. The sub-
phase pH of 10−4 M CuCl2 solution was found to be 6.1. The
monolayer was compressed with a constant barrier speed of
3 mm/min. The isotherm (not shown) exhibited a condensed
nature, in which the liquid condensed region was absent, in-
dicating complete ionization of arachidic acid. The limiting
mean molecular area (LMMA) was found to be ∼20 Å2 com-
parable to the reported values for arachidate monolayers [20].
The multilayer deposition was carried out at a surface pres-
sure of 20 mN/m. The compressed monolayer was transferred
by vertical dipping method at a speed of 3 mm/min. Typically
about 100–200 monolayers were transferred on quartz and
CaF2 substrates, respectively.
Fig. 1. X-ray reflection patterns of a copper arachidate multilayer on a quartz
The as-deposited CuA multilayers were annealed in a substrate. In (a) as-deposited state and after heat treatments at (b) 100 ◦ C,
quartz tubular furnace in the temperature range 100–600 ◦ C (c) 190 ◦ C, (d) 200 ◦ C.
in a stream of oxygen at atmospheric pressure. A tempera-
ture controller (±5 ◦ C) was used to set the temperature at the strate are shown in Fig. 1. The XR pattern (a) for the as-
required value. deposited multilayer exhibits well-defined Bragg peaks cor-
IR spectra were obtained with a Perkin Elmer Spectrum1 responding to (0 0 l) reflections. The average bilayer period
FTIR spectrometer and UV–vis spectra with a Shimadzu UV- calculated from the peak positions was found to be 55 Å. Con-
160A spectrophotometer. X-ray reflection (XR) studies were sidering the total length of arachidate molecule to be ∼27.5 Å,
performed with Philips PW1820 powder diffractometer us- the bilayer period of 55 Å indicates molecular packing with
ing Cu K␣ radiation in the 2θ range of 4–16◦ . Transmission alkyl chains of copper arachidate nearly perpendicular to
electron microscopy (TEM) studies were carried out using the layer [21]. Heat treatment of the CuA multilayer up to
a PHILIPS Model CM200 Supertwin Transmission Electron ∼100 ◦ C (Fig. 1b) showed an increase in the intensity of the
microscope operated at 200 kV. For TEM studies, CuO films Bragg peaks, without any change in their positions. Further
were lifted off from quartz substrates in a dilute HF solution heat treatments in the range 100–190 ◦ C showed that the in-
and transferred onto copper grids. X-ray photoelectron spec- tensities of the Bragg peaks decreased. Fig. 1c shows the XR
troscopic measurements were carried out with an XPS set- pattern of the CuA multilayer heat-treated at 190 ◦ C. How-
up incorporated in the RIBER MBE EVA 32 R&D system. ever, when the CuA multilayer was heat-treated at 200 ◦ C,
The XPS system comprises of a twin anode (Mg/Al) source the Bragg peaks disappeared (Fig. 1d) suggesting complete
(Model CX 700) and an electron energy analyzer (Model destruction of the layered structure and this is attributed to
MAC 2). The excitation used was Mg K␣ (1253.6 eV) and melting of the arachidate salt. Similar observations have ear-
the binding energies were referenced using Au 4f7/2 peak. lier been reported on LB multilayers of several fatty acid salts
[22].
Fig. 2a and b shows the FTIR spectra in the regions,
3. Results and discussion 2800–3000 and 1300–1700 cm−1 , respectively, for the CuA
multilayer on CaF2 substrate in the as-deposited state and af-
The XR patterns in the 2θ range of 4–16◦ for the as- ter oxidation treatment at different temperatures. The spectra
deposited and heat-treated CuA multilayers on quartz sub- of the as-deposited multilayer show intense bands at 2916 and
M. Parhizkar et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 257–258 (2005) 277–282 279

Fig. 3. UV–vis transmittance (T) and reflectance (R) spectra of (a) as de-
posited copper arachidate multilayer and after heat treatment at (b) 100 ◦ C,
(c) 200 ◦ C, (d) 300 ◦ C, (e) 400 ◦ C, 500 ◦ C.

Fig. 3 shows the transmission and reflection spectra of


the as-deposited and heat-treated CuA multilayers on quartz
substrates, in the wavelength range 300–1100 nm. The as-
deposited CuA multilayer is transparent in the visible region
and exhibits an absorption edge at ∼300 nm. A similar be-
haviour is seen with a small reduction in transmission af-
ter heat treatment at 100 ◦ C. However, the heat treatment at
200 ◦ C results in a noticeable drop in transmission in the re-
Fig. 2. FTIR transmittance spectra of an as-deposited copper arachidate
gion 300–400 nm, with an increase in overall reflection, indi-
multilayer () and after heat treatment at 100 ◦ C (䊉), 200 ◦ C () and 300 ◦ C cating the beginning of oxide formation. The films subjected
(—). to heat treatment at 300 ◦ C show significant loss of transmis-
sion in the wavelength region 300–700 nm. Subsequent heat
2849 cm−1 due to CH2 asymmetric and symmetric vibrations treatments at 400 and 500 ◦ C did not show any significant
of the alkyl chain of the CuA. The strong absorption band at change in the UV–vis spectra. These observations suggest
1587 cm−1 is assigned to the asymmetric stretching vibra- that the oxidation process starts at ∼200 ◦ C and is completed
tions of the carboxylate group. The presence of this band and at ∼300 ◦ C. This conclusion is consistent with the FTIR data
the complete absence of C O stretching band of unionized discussed above.
carboxylic acid at ∼1700 cm−1 confirm that the as-deposited The structure and composition of the CuA multilayer heat-
multilayer consists of CuA and not a mixture of arachidic treated at temperatures of 300 ◦ C and above was investigated
acid and salt. The bands at ∼1470, ∼1413 and ∼1315 cm−1 by transmission electron microscopy (TEM) and X-ray pho-
are assigned to CH2 scissoring, COO− symmetric stretch and toelectron spectroscopy (XPS). The XR patterns of the heat-
CH2 wagging modes of arachidate salt, respectively [23]. The treated multilayers were featureless, which is attributed to the
asymmetric COO− band is normally observed for divalent small thickness of the oxide film formed by the heat treatment.
fatty acid salts, such as cadmium arachidates at ∼1540 cm−1 The bright-field image and the diffraction pattern of the
[23]. However, it is known to shift to higher frequencies de- CuA film obtained by heat treatment at 400 ◦ C are shown in
pending on the size, mass and electronegativity of the metal Fig. 4a and b, respectively. The diffused rings in the diffrac-
[24]. tion pattern were indexed to reflections from the planes of
The spectrum of the CuA multilayer subjected to oxida- CuO indicating the presence of CuO phase in the oxidized
tion treatment at 100 ◦ C shows that in general, the bands be- film. The presence of a small quantity of Cu2 O could not be
come sharper and more intense compared to the as-deposited ruled out on the basis of TEM data, since all the reflections
multilayer, suggesting improvement in structural order due from cubic Cu2 O lie close to a large number of reflections
to the annealing effect of heat treatment [25]. On oxidation from monoclinic CuO. Further, the diffused nature of the
treatments at higher temperatures, the intensities of the bands diffraction rings makes it quite difficult to unambiguously
decrease and eventually disappear for multilayers treated at rule out the presence of reflections from Cu2 O. The bright-
300 ◦ C or above. This indicates the removal of all organic field image of the film shows a uniform array-like distribution
moieties by pyrolysis in the films subjected to oxidation treat- of nanoparticles and the image analysis showed the particle
ment at 300 ◦ C or above. size to be 2.5 ± 1.5 nm. The nanocrystalline nature of the film
280 M. Parhizkar et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 257–258 (2005) 277–282

Cu 2p3/2 peak in CuO is reported at 933.2 eV as compared to


932.4 eV in Cu2 O [26]. The separation between the Cu 2p3/2
peak and its first satellite is about 8 eV and the 2p peaks are
broad, with 2p3/2 having a peak width of ∼3.5 eV. In compar-
ison, the 2p core levels of Cu2 O are reported to be sharp (Cu
2p3/2 peak width ∼1.9 eV) [26]. These features are charac-
teristics of materials such as copper dihalides or CuO having
a d9 configuration in ground state [26]. Thus the existence
of strong satellite peaks for Cu 2p confirms the presence of
CuO phase in the film. Furthermore, the symmetric nature of
the 2p core level peaks and the absence of any shoulder on
the lower energy side confirms the absence of Cu2 O in the
film.
The absorption spectrum of the CuO film obtained by heat
treatment at 400 ◦ C of a CuA multilayer consisting of 200
monolayers was calculated from the transmission and reflec-
tion data shown in Fig. 3. The absorption coefficient (α) was
calculated using the simple relation for a strongly absorb-
ing film, neglecting reflections from the film–substrate and
substrate–air interfaces:

T = (1 − R)e−αt

The thickness (t) of the heat-treated oxide film was taken


to be 40 nm based on earlier data on oxidation of copper
Fig. 4. (a) Bright-field image and (b) the selected area diffraction pattern
of a copper oxide film obtained by heat treatment of a copper arachidate arachidate LB multilayers [19]. The plots of α versus hν in the
multilayer at 400 ◦ C. energy range of 1−5 eV for films heat-treated at temperatures
above 300 ◦ C were found to be similar. Hence, only a typical
is responsible for the diffused nature of the rings seen in the plot has been shown in Fig. 6a for the film heat-treated at
electron diffraction pattern of the oxide film. 400 ◦ C. Fig. 6b shows an expanded version of this plot in the
In order to confirm the absence of Cu2 O phase, XPS stud- energy range 1.55–1.65 eV. The values of α ≤ 1.6 eV were
ies were carried out on the above film. Fig. 5 shows the found to be smaller than 1000 cm−1 , the experimental error
Cu 2p core level XPS spectrum. The peaks corresponding in the measurement of α. A significantly sharp increase in
to the core level 2p3/2 and 2p1/2 transitions of copper at 933.5 α is clearly seen in Fig. 6b for energy values greater than
and 953.2 eV, respectively, along with the satellites on higher ∼
=1.6 eV.
binding energy side are seen. These values compare well with Limited studies have been reported on the electronic struc-
the reported values for Cu 2p levels in CuO. For example, the ture of CuO and its optical properties. Most of the results on
polycrystalline and thin-film samples report an energy gap in
the range 1.2–1.85 eV, with some early reports mentioning
the absorption edge to be too shallow to define an energy
gap [3,4,27–29]. Ghijsen et al. [26] have studied the elec-
tronic structure of CuO by electron spectroscopic techniques
to demonstrate that CuO is a charge transfer gap insulator
with a bandgap of 1.4 eV. A limited number of band structure
calculations have also been carried out for CuO, of which,
Ching et al. [30] have reported that CuO has a direct gap of
1.6 eV, which agrees well with the present results. The absorp-
tion spectrum of CuO films obtained in the present work also
shows a good agreement with the spectroscopic ellipsometry
results of Ito et al [31]. There is a general agreement, both in
the nature of the α versus hν curve as well as the magnitude
of α in the energy range studied. It may be mentioned that
owing to the ambiguities associated with the electronic struc-
ture and optical properties of CuO, we have not attempted to
Fig. 5. Cu 2p core level X-ray photoelectron spectrum of a copper oxide describe the absorption edge through the usual dependence
film obtained by heat treatment of a copper arachidate multilayer at 400 ◦ C. of α ∼ (ν − ν0 )x .
M. Parhizkar et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 257–258 (2005) 277–282 281

4. Conclusions

Single-phase copper (II) oxide thin films have been formed


by the pyrolysis of copper arachidate LB multilayers. FTIR
and UV–vis spectroscopic investigations showed that the ox-
idation and the removal of organic moieties were complete
at ∼300 ◦ C. TEM images showed the film to consist of a
uniformly distributed particles of size 2.5 ± 1.5 nm. The ab-
sorption edge of the CuO films was found to be 1.6 eV and did
not show any noticeable blue shift due to the nanocrystalline
size of the films. This absence of blue shift is attributed the
relatively large effective mass of the holes and electrons in
CuO and the uncertainty in the value of the absorption edge
of CuO prevailing in literature.

Acknowledgments

The financial support for this work from the Department


of Science and Technology, Government of India, is grate-
fully acknowledged. One of the authors Sukhvinder Singh is
thankful to CSIR, New Delhi, for junior research fellowship.

References

[1] A.H. Pfund, Phys. Rev. 3 (1916) 289.


[2] B.O. Seraphin, in: B.O. Seraphin (Ed.), Solar Energy Conversion –
Solid State Physics Aspects, Springer Verlag, Berlin, 1979.
[3] B. Karlsson, C.G. Ribbing, A. Roos, E. Valkonen, T. Karlsson, Phys.
Fig. 6. (a) Absorption coefficient (α) vs. photon energy plot for a cop- Scripta 25 (1982) 826.
per oxide film obtained by heat treatment of a copper arachidate multi- [4] F.P. Koffyberg, F.A. Benko, J. Appl. Phys. 53 (1982) 1173.
layer at 400 ◦ C and (b) an expanded version in the photon energy range of [5] N. Brockhouse, Phys. Rev. 94 (1954) 781.
1.55–1.65 eV. [6] S. Asbrink, L.-J. Norrby, Acta Cryst. B 26 (1970) 8.
[7] V.R. Palkar, P. Ayyub, S. Chattopadhyay, M. Multani, Phys. Rev. B
53 (1996) 2167.
[8] M.A. Brookshier, C.C. Chusuei, D.W. Goodman, Langmuir 15
(1999) 2043.
It is also interesting to note that the absorption edge of [9] J.F. Xu, W. Ji, Z.X. Shen, S.H. Tang, X.R. Ye, D.Z. Jia, X.Q. Xin,
CuO films measured in this work is not too different from the J. Solid State Chem. 147 (1999) 516.
[10] R.V. Kumar, R. Elgamiel, Y. Diamant, A. Gedanken, Langmuir 17
corresponding values of 1.3–1.6 eV reported in recent litera- (2001) 1406.
ture for CuO, as described above. This implies that the CuO [11] B. Orel, F. Svegl, N. Bukovec, M. Kosec, J. Non-Cryst. Solid 159
films obtained in the present work do not exhibit a noticeable (1993) 49.
blue shift in spite of their nanocrystalline nature. This is in [12] Z. Hong, Y. Cao, J. Deng, Mater. Lett. 52 (2002) 34.
contrast to the recent studies on CuO nanoparticles [12,13] [13] K. Borgohain, S. Mahamuni, J. Mater. Res. 17 (2002) 1220.
[14] D. Das, D. Chakravorty, Appl. Phys. Lett. 76 (2000) 1273.
that have reported a blue shift in the absorption features de- [15] W. Meisel, P. Gütlich, P. Tippmann-Krayer, H. Möhwald, Fresenius’
pending on the particle size. It may be noted here that very J. Anal Chem. 341 (1991) 289.
limited data are available in literature on the electron and [16] D.T. Amm, D.J. Johnson, T. Laursen, S.K. Gupta, Appl. Phys. Lett.
hole effective masses in CuO, which report relatively large 61 (1992) 522.
values of effective masses of electrons (0.4–0.95m0 ) [30,32] [17] D. Brandl, Ch. Schoppmann, Ch. Tomaschko, J. Markl, H. Voit, Thin
Solid Films 249 (1994) 113.
and holes (7.9m0 ) [4] for CuO. Thus considering the uncer- [18] D.V. Pranjape, M. Sastry, P. Ganguly, Appl. Phys. Lett. 63 (1993)
tainties in the exact location of absorption edge of CuO as 18.
described earlier, and a large effective mass of electrons and [19] M. Schurr, M. Seidl, A. Brugger, H. Voit, Thin Solid Films 342
holes, it is unlikely that a significant blue shift can be unam- (1999) 266.
biguously observed. This explains the apparent absence of a [20] G.L. Gaines, Insoluble Monolayers at Liquid–Gas Interface, Inter-
science, New York, 1966.
significant blue shift in the absorption edge of the CuO films [21] G. Roberts, Langmuir–Blodgett Films, Plenum Press, New York,
in spite of their nanocrystalline nature. 1990.
282 M. Parhizkar et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 257–258 (2005) 277–282

[22] W. Mahler, T.A. Barberka, U. Pietsch, U. Hohne, H.J. Marle, Thin [27] V.F. Drobny, D.L. Pulfrey, Thin Solid Films 61 (1979) 89;
Solid Films 256 (1995) 198; A.E. Rakhshani, F.K. Barkat, Mater. Lett. 6 (1987) 37.
P. Tippmann-Krayer, R.M. Kenn, H. Mohwald, Thin Solid Films [28] K. Santra, C.K. Sarkar, M.K. Mukherjee, B. Ghosh, Thin Solid Films
210–211 (1992) 577. 213 (1992) 226.
[23] J.F. Rabolt, F.C. Burns, N.E. Burns, N.E. Schlotter, J.D. Swalen, J. [29] H. Wieder, A.W. Czanderna, J. Appl. Phys. 37 (1966) 184;
Chem. Phys. 78 (1983) 946. T. Tanaka, Jpn. J. Appl. Phys. 18 (1979) 1043;
[24] K. Nakamoto, Infrared Spectra of Inorganic and Coordination Com- K.L. Hardee, A.J. Bard, J. Electrochem. Soc. 124 (1977) 215.
pounds, Wiley, New York, 1963. [30] W.Y. Ching, Y. Xu, K.W. Wong, Phys. Rev. B 40 (1989) 7684.
[25] J. Cha, H.S. Shin, Y.K. See, J. Lee, T. Chang, S.B. Kim, Synth. [31] T. Ito, H. Yamaguchi, T. Masuni, S. Adachi, J. Phys. Soc. Jpn. 67
Met. 117 (2001) 169. (1998) 3304.
[26] J. Ghijsen, L.H. Tjeng, J. van Elp, H. Eskes, J. Westerink, G.A. [32] S. Warren, W.R. Flavell, A.G. Thomas, J. Hollingworth, P.L. Wincott,
Sawatzky, M.T. Czyzyk, Phys. Rev. B 38 (1988) 11322. A.F. Prime, S. Downes, C. Chen, J. Phys. 11 (1999) 5021.

You might also like