You are on page 1of 8

C

View Online
Journal of Dynamic Article Links <

Materials Chemistry
Cite this: DOI: 10.1039/c1jm13277f
www.rsc.org/materials PAPER
New routes to copper sulfide nanostructures and thin films†
Ahmed Lutfi Abdelhady,a Karthik Ramasamy,‡a Mohammad Azad Malik,a Paul O’Brien,*ab Sarah J. Haighb
and James Rafterya
Received 13th July 2011, Accepted 9th September 2011
DOI: 10.1039/c1jm13277f
Published on 05 October 2011 on http://pubs.rsc.org | doi:10.1039/C1JM13277F
Downloaded by University of Alabama at Tuscaloosa on 05 October 2011

The copper(II) complex of 1,1,5,5-tetra-iso-propyl-2-thiobiuret [Cu(SON(CNiPr2)2)2], has been


synthesized and its single crystal X-ray structure determined. The complex has been used as a single
source precursor for the preparation of copper sulfide nanoparticles by solution thermolysis. The
nanoparticles synthesized had various morphologies including spherical, hexagonal disks, and trigonal
crystallites; depending on the reaction temperature, concentration of the precursor and the growth
time. Thermolysis experiments in oleylamine produced Cu7S4 nanoparticles as a mixture of roxbyite
(monoclinic) and anilite (orthorhombic) phases. Pure anilite Cu7S4 nanoparticles were obtained when
a solution of precursor in octadecene was injected into hot oleylamine whereas, djurleite Cu1.94S
nanoparticles were obtained when a solution of the precursor in oleylamine was injected into hot
dodecanethiol. The optical properties of the pure anilite Cu7S4 phase and the djurleite Cu1.94S are
consistent with indirect band gap materials. The complex was also used for the preparation of copper
sulfide thin films by aerosol assisted chemical vapour deposition (AACVD). Powder X-ray diffraction
(p-XRD) patterns of the thin films showed the deposition of a mixture of cubic CuS2 and hexagonal
Cu2S phases at all temperatures. Scanning Electron Microscopy (SEM) of copper sulfide films showed
spherical crystallites from reaction at 280  C, cuboids at 320  C and plate like crystallites between 360
and 400  C. Transmission Electron Microscopy (TEM) of material scratched from thin films showed
nanoplates of copper sulfides with 25–30 nm in width.

Introduction Nanocrystalline Cu9S8, Cu7S4 and CuS were formed in auto-


claves using [Cu(NH3)4]2+ and thiourea.10 In a solvothermal
Copper sulfides thin films and nanoparticles have been investi- method in toluene CuS nanoplatelets were synthesised from
gated for many uses including as: p-type semiconductors in solar copper acetate and CS2.11 In another solvothermal process,
cells,1–3 nanoscale switches4 and cathodic materials for lithium hierarchical doughnut-shaped CuS particles were prepared from
rechargeable batteries.5 Vaughan6 reported that in 1940 only the a copper nitrate and thiourea mixture.12 Korgel et al.13,14 syn-
end member (Cu2S) and CuS were known in the Cu–S system. By thesised nanorods, nanodisks, and nanoplatelets of Cu2S by the
1974 nine more copper sulfide phases had been identified6,7 and in solventless thermolysis of a copper alkylthiolate molecular
2006 a total of fourteen copper sulfide phases were recognized.7 precursor. Chen et al. prepared nanoplates of Cu2S15 and
Some known forms of copper sulfide include: chalcocite (Cu2S), nanoflakes of CuS16 by thermal decomposition of [Cu(acac)2]
djurleite (Cu31S16 or Cu1.94S), digenite (Cu9S5 or Cu1.8S), anilite (acac ¼ acetylacetonate) with elemental sulfur in oleylamine.
(Cu7S4 or Cu1.75S), covellite (CuS) and villamaninite (CuS2).6–8 Recently Cu2S nanocrystals were prepared using cysteine as the
Various methods have been used to synthesise copper sulfide sulfur precursor.17 Thermal decomposition of different precur-
nanoparticles. A hydrothermal method gave digenite nanowires sors such as [Cu(acac)2] or CuCl2 and elemental sulfur or
and djurleite nanotubes from CuCl and thiourea.9 dodecanethiol, produced hexagonal nanodisks of CuS,18,19
Cu7S4,20 and Cu2S.20,21 Nanocrystals of CuS,22,23 Cu7S4,24 and
a
The School of Chemistry, The University of Manchester, Oxford Road, Cu2S24–28 were synthesised by the thermolysis of single source
Manchester, M13 9PL, UK. E-mail: paul.obrien@manchester.ac.uk; precursors including alkyl xanthates,22 mercaptobenzothiazole,23
Fax: +44 161 275 4598; Tel: +44 161 275 4653
b
The School of Materials, The University of Manchester, Oxford Road,
thiobenzoates,24 dithiocarbamates25–27 and dithiolates28 in hot
Manchester, M13 9PL, UK coordinating solvents.
† Electronic supplementary information (ESI) available: TEM images Thin films of copper sulfide have been prepared by various
and UV-Vis spectra of Cu7S4 nanoparticles. CCDC reference number methods including RF-reactive sputtering,29 spray pyrolysis,30
715965. For ESI and crystallographic data in CIF or other electronic
format see DOI: 10.1039/c1jm13277f
successive ionic layer adsorption and reaction,31 chemical bath
‡ Present Address: Center for Materials for Information Technology, deposition32 and chemical vapor deposition.33 In our continuing
University of Alabama, Tuscaloosa, Alabama, USA, 35487. search for potential single source precursors we have studied

This journal is ª The Royal Society of Chemistry 2011 J. Mater. Chem.


View Online

metal complexes of dichalcogenocarbamates,34–39 dichalcoge- Synthesis of nanoparticles


noimidodiphosphinates,40–43 and dichalcogenophosphinates44,45
All the nanoparticles were synthesized by thermal decomposition
as suitable precursors for deposition of metal chalcogenide thin
of the single source precursor. Nine different thermolysis
films by metal organic chemical vapour deposition (MOCVD).
experiments were carried out corresponding to three different
Herein, we report the synthesis of good quality Cu7S4 nano-
temperatures (200, 240 and 280  C) with three different
particles and the AACVD deposition of CuS2 and Cu2S thin
concentrations (5, 10 and 20 mM) at each temperature. In
films from a copper thiobiuret complex46 [Cu(SON(CNiPr2)2)2].
a typical experiment, oleylamine (15 mL) was degassed under
reduced pressure at 100  C for 30 min and then heated to the
Experimental methods desired temperature under nitrogen. The required amount of the
copper sulfide precursor [Cu(SON(CNiPr2)2)2] was dissolved in
All preparations were performed under an inert atmosphere of
oleylamine (5 mL) and injected into the hot oleylamine solution.
dry nitrogen using standard Schlenk techniques. Dodecanethiol
The reaction was maintained for 1 h. The dark solution formed
was purchased from Fluka. All other reagents were purchased
was cooled to approximately 70  C. An excess of methanol was
Published on 05 October 2011 on http://pubs.rsc.org | doi:10.1039/C1JM13277F

from Sigma-Aldrich chemical company and used as received.


then added and the solid was isolated by centrifugation. The
Downloaded by University of Alabama at Tuscaloosa on 05 October 2011

Solvents were distilled prior to use. Mass spectra were recorded


solid was washed several times with methanol then redispersed in
on a Kratos concept 1S instrument. Infrared spectra were
toluene. Any insoluble material was removed by centrifugation.
recorded on a Specac single reflectance ATR instrument (4000–
400 cm1, resolution 4 cm1). Elemental analysis was performed
by the University of Manchester micro-analytical laboratory.
TGA measurements were carried out on a Seiko SSC/S200 model Deposition of films by AACVD
at a heating rate of 10  C min1 under nitrogen.
In a typical experiment, 0.016 M of the precursor in 20 mL
tetrahydrofuran was loaded in a two-necked 100 mL round-
Synthesis of [Cu(SON(CNiPr2)2)2] (1) precursor bottom flask with a gas inlet that allowed the carrier gas (argon)
A solution of di-iso-propylcarbamoyl chloride (1.0 g, 6 mmol) to pass into the solution and aid the transport of the aerosol. The
and sodium thiocyanate (0.49 g, 6 mmol) in acetonitrile (25 mL) round-bottom flask was kept in a water bath above the piezo-
was heated to reflux with continuous stirring for 1 h, during electric modulator of a PIFCO ultrasonic humidifier (Model No.
which time a fine precipitate of sodium chloride formed. To the 1077). This flask was connected to the reactor tube by a piece of
cooled reaction mixture was added di-iso-propylamine (1.49 mL, reinforced tubing. The argon flow rate was controlled by a Platon
12 mmol) followed by stirring for 30 min and addition of copper flow gauge. Seven glass substrates (approx. 1  3 cm) were
(II) nitrate (0.76 g, 3 mmol). The crude product was isolated as placed inside the reactor tube, which is placed in a CARBOLITE
green powder and recrystallized from tetrahydrofuran to give furnace. The aerosol droplets of the precursor thus generated
suitable single crystals for X-ray crystallography. Yield 2.20 g were transferred into the hot zone of the reactor by the carrier
(57%), mpt: 170  C MS (APCI) major fragments: m/z ¼ [M+] gas. Both the solvent and the precursor were evaporated and the
636, [Cu(C14H28N3OS)2], [NSO(CNiPr2)2] 284. IR (vmax/cm1): precursor vapor reached the heated substrate surface where
2964(s), 1500(m), 1434(m), 1359(m), 1284(s), 1133(s),1059(m). thermally induced reactions and film deposition took place.
Elemental analysis: Calc. for C28H56N6O2S2Cu: C, 52.8; H, 8.8;
N, 13.2; S, 10.0; Cu, 9.9%. Found: C, 52.3; H, 9.0; N, 13.3; S, 9.6;
Cu, 9.7%.
Characterization of nanoparticles and thin films
X-ray diffraction studies were performed on a Bruker AXS D8
X-Ray crystallography of precursor
diffractometer using Cu–Ka radiation. The samples were
Single-crystal X-ray diffraction data for the compounds were mounted flat and scanned between 10 to 80 in a step size of 0.05
collected using graphite monochromated Mo-Ka radiation (l ¼ with a count rate of 9 s. Films were carbon coated using
 on a Bruker APEX diffractometer. The structure was
0.71073 A) Edward’s E306A coating system before carrying out SEM and
solved by direct methods and refined by full-matrix least EDX analyses. SEM analysis was performed using a Philips XL
squares47 on F2. All non-H atoms were refined anisotropically. H 30FEG and EDX was carried out using a DX4 instrument. For
atoms were included in calculated positions, assigned isotropic TEM analysis, the scratched thin films were sonicated in toluene
thermal parameters and allowed to ride on their parent carbon and then drop cast onto lacey carbon copper grids. Nanoparticle
atoms. All calculations were carried out using the SHELXTL samples were simply drop cast onto lacey carbon copper grids.
package.48 The details pertaining the data collection of the High resolution transmission electron microscopy (HRTEM)
crystals as follows: C28H56N6O2S2Cu; M ¼ 636.45; Green and selected area electron diffraction (SAED) were performed
 b ¼ 11.127(3) A,
blocks; Triclinic; p-1; a ¼ 7.952(2) A,  c ¼ 11.142 using Tecnai F30 FEG TEM instrument, operating at 300 kV.
 a ¼ 65.992(6) ; b ¼ 72.469 (5) ; g ¼ 79.123(6) ; V ¼ 856.2
(3) A; Electron tomography data was obtained using a JEOL 2100
3; Z ¼ 1; D ¼ 1.234 Mg/m3; T ¼ 100(2) K; reflections
(4) A TEM at 200kV for tilt angles of 60 and tomographic recon-
collected ¼ 5384, unique reflections ¼ 3796 [R(int) ¼ 0.0877]; struction was performed using the IMOD software.49 The
R1 ¼ 0.0471 and wR2 ¼ 0.0682 for [I > 2q(I)]; R1 ¼ 0.1681 and absorption spectra were recorded on a UV-Vis spectrophoto-
wR2 ¼ 0.0896 for all data; largest diff. peak and hole ¼ 0.445 and meter (Thermo Spectronic Helios Beta) in the wavelength range
0.472 e.A3, GOF ¼ 0.491. CCDC reference number 715965. of 400–1000 nm.

J. Mater. Chem. This journal is ª The Royal Society of Chemistry 2011


View Online

Results and discussions and 20 mM in oleylamine) at each temperature. Nanoparticles


obtained from the nine different experiments were analysed by p-
Most of the materials reported for photovoltaic use are either XRD, TEM and UV-Vis spectroscopy. The p-XRD patterns of
toxic or use less-abundant elements such as Cd, Pb, In or Ga. the nanoparticles obtained from all experiments correspond to
Less-toxic, abundant, and thus cheaper materials may be more a mixture of the two Cu7S4 phases, roxbyite (monoclinic) (ICDD
promising even with overall lower efficiencies. Recent estimates card No. 023-0958) and anilite (orthorhombic) (ICDD card No.
of the annual electricity potential as well as material extraction 033-0489) as illustrated in Fig. 2 for a concentration of 10 mM.
costs and environmental friendliness led to the identification of Comparing the intensity of the peaks for roxbyite (086) at 46.83
materials that could be used in photovoltaic applications on and for anilite (224) at 46.23 it is clear that at higher thermolysis
a large scale.3 The most promising materials include iron and temperatures the orthorhombic form dominates. A quantitative
copper sulfide. comparison for concentrations of 10 mM and 20 mM is shown in
supplementary information Fig. S1.† According to Fleet,7 there
X-ray single crystal structure of [Cu(SON(CNiPr2)2)2] have been no reports on the stability of the monoclinic phase.
Published on 05 October 2011 on http://pubs.rsc.org | doi:10.1039/C1JM13277F

The dominance of the orthorhombic form at higher temperatures


The X-ray single crystal structure† of [Cu(SON(CNiPr2)2)2] in
Downloaded by University of Alabama at Tuscaloosa on 05 October 2011

could be due to a lower stability of the monoclinic form. The


Fig. 1 shows the expected square planar copper(II) environment
gradual decrease in the ratio of the monoclinic phase to the
with trans ligation. Within each of the two thiobiuret ligands, the
orthorhombic phase may suggest a metastable monoclinic phase.
four atoms of the urea or thiourea groups show significant
At 280  C with 20 mM solution of the precursor; the nano-
deviation from planarity due to twisting about the central N
particles obtained were principally orthorhombic with only
atom. In both ligands, the pattern of bond distances indicates
minor impurities of the monoclinic phase. TEM images of the
that the formal negative charge is predominately localized on the
nanoparticles grown at different temperatures and concentra-
S atom. The relatively long C–S and short C–O average bond
 are consistent with mostly tions showed remarkable changes in the morphology of crystal-
lengths of 1.763(4) and 1.280(5) A
lites (Fig. 3). At 5 mM solution of the precursor and 200  C,
single and double bond character, respectively. This bond
hexagonal nanodisks and spherical nanoparticles with an
localization is also reflected in the average C–N bond distances to
 in the iso-thiourea group and average diameter of 18.2  2.4 nm were obtained (Fig. 3(a)).
the central N atom: 1.324(5) A
 All the hexagonal particles observed by TEM had large
1.352(5) A in the urea group. In contrast, all the C–NiPr2 bond
 Structure diameter-to-thickness ratios50 and were therefore oriented
lengths are similar, with an average of 1.357(7) A.
parallel to the grid. Hence, in most cases particle thickness had to
refinement data is listed in experimental section. Selected bond
be obtained using tilting experiments. At 240 and 280  C,
angles and bond lengths are given in the caption to Fig. 1.
spherical or nearly spherical nanoparticles were formed with an
average diameter of 9.4  1.3nm and 9.8  1.6 nm, respectively
Thermogravimetric analysis (Fig. 3(b) and 3(c)). On increasing the precursor concentration to
Thermogravimetric analysis shows that [Cu(SON(CNiPr2)2)2] 10 mM at 200  C, hexagonal nanodisks lying flat or stacked face-
decomposes in a single step between 172 and 267  C with an to-face were observed (Fig. 3(d)). The presence of standing and
observed final residue of 14.4% close to the calculated value of flat oriented disks allowed for the easy measurement of diameters
15.0% for copper sulfide (CuS). and thickness by TEM. The average diameter was found to be
22.3  4.6 nm, while the thickness was found to be 4.9  0.35 nm.
When nanodisks are standing upright on the grid they resemble
Characterisation of copper sulfide nanoparticles
rods20 so the shape of the nanodisks was confirmed using electron
All the nanoparticles were synthesized by thermal decomposition tomography (Fig. 4). The disk morphology agrees with that
of the single source precursor. Nine different thermolysis
experiments were carried out at three different temperatures
(200, 240 and 280  C) with three different concentrations (5, 10

 and bond angles


Fig. 1 X-ray structure of (1). Selected bond lengths (A)
( ) C1–N1 1.479(5), C7–S1 1.763(4), C8–O1 1.280(4); N2–C5–S1 124.4 Fig. 2 p-XRD patterns of Cu7S4 at (a) 200, (b) 240 and (c) 280  C using
(3), O1–Cu1–S1 93.1(1). 10 mM. (:) represents roxbyite peaks and (-) represents anilite peaks.

This journal is ª The Royal Society of Chemistry 2011 J. Mater. Chem.


View Online

precursor concentration (20 mM), lead to parallel oriented


hexagonal nanodisks along with some spherical nanoparticles of
an average diameter of 12.8  1.9 nm at 200  C (Fig. 3g).
Trigonal nanodisks were formed at 240  C (Fig. 3(h)), while at
280  C quasi-close-packed spherical nanoparticles with average
diameter of 10.2  2.1 nm were obtained (Fig. 3(i)). A spherical
structure and not a coin structure was confirmed using tilting
experiments.
From these results, it can be suggested that at lower growth
temperature, fewer nuclei are formed which favours the growth
of anisotropic nanocrystals.52,53 By increasing the temperature
more nuclei are formed, due to increasing the reactivity of the
precursor, leaving a lower monomer concentration in solution
Published on 05 October 2011 on http://pubs.rsc.org | doi:10.1039/C1JM13277F

which directs the nanocrystals growth towards the lowest


Downloaded by University of Alabama at Tuscaloosa on 05 October 2011

chemical potential environment and results in the formation of


spherical nanocrystals.52,53
To study the morphology at an early stage of the reaction time,
one or two drops were withdrawn from the reaction mixture after
5 min and dispersed in toluene for TEM. The results for all
Fig. 3 TEM images of Cu7S4 after 1 h. Reactions carried out using samples are shown in Fig. S3. For all thermolysis experiments,
5 mM (a–c), 10 mM (d–f) and 20 mM (g–i) solutions of the precursor at trigonal and hexagonal nanodisks were obtained with no
200  C (a, d and g), 240  C (b, e and h) and 280  C (c, f and i). evidence of any spherical nanoparticles. Fig. 5, shows a study of
the effect of reaction time on morphology. The temperature for
this reaction was fixed at 280  C and the highest concentration
was used (20 mM). At the beginning of the reaction (2 and 5 min)
trigonal and hexagonal nanodisks were obtained as shown in
Fig. 5(a) and 5(b). After 30 min (Fig. 5(c)) the tips of these
nanodisks began to erode and at 1 h (Fig. 5(d)) all the nano-
particles became spherical. From these results, it is clear that at
the higher temperatures and longer enough growth times, only
spherical nanocrystals were obtained.
Hexagonal Cu7S4 nanodisks showed two alignments in TEM
studies: parallel alignment to the grid and upright alignment.
Interestingly, these alignments were found to depend not only
on the diameter-to-thickness ratio of the nanodisk but also on
the concentration of the solution used for preparing the TEM
sample. When a concentrated solution was used (10 mg in

Fig. 4 TEM images from a tomographic data series (60 ) at tilt angles
of (a)+32 , (b) 0 and (c) 36 . Data slices extracted from the complete
tomographic reconstruction are shown for the top surface of the carbon
support film (d), the centre of the film (e), and the bottom surface (f). The
rendered surface of a pair of representative nanoparticles is shown in for
different view directions (g)-(i). The red square in (d) indicates the posi-
tion of these particles in the full reconstruction.

described in previous reports.14 The distance between the face-to-


face stacked nanodisks is about 1.2 nm, which is almost equal to
the length of oleylamine,51 indicating that the nanodisks are
covered with a layer of the amine. In some cases the separation
was double at ca. 2.4 nm, which is equal to the total length of
the two head-to-head oleylamine molecules as shown in Fig. S2.
This stacking may minimize the exposed surface area and hence,
reduce the surface energy of the structure.20,50 At higher
temperatures the particle morphology changes and disks are not
observed. Instead near-spherical or spherical nanoparticles with Fig. 5 TEM images of Cu7S4 produced at (a) 2 min, (b) 5 min, (c) 30 min
diameters of 9.3  2.2 nm at 240  C and 12.8  2.2 nm at and (d) 60 min. Reaction carried at 280  C and using 20 mM solution of
280  C were produced (Fig. 3(e) and 3(f)). Using a higher the precursor.

J. Mater. Chem. This journal is ª The Royal Society of Chemistry 2011


View Online

Fig. 6 TEM of (a) concentrated and (b) diluted Cu7S4 nanodisk


samples.
Published on 05 October 2011 on http://pubs.rsc.org | doi:10.1039/C1JM13277F
Downloaded by University of Alabama at Tuscaloosa on 05 October 2011

1 mL of toluene) the percentage of the upright nanodisks was


very much higher than when the same solution was diluted 10
fold. The upright nanodisks became a minority, as shown in
Fig. 6. Similar results were previously reported for In2S3.54
Fig. 7 (a-d) shows the HRTEM images of samples prepared
using 20 mM solution of the precursor at 280  C for 5, 30 and
60 min. These images clearly show the transformation of the
nanodisks into spherical nanoparticles. The HRTEM images
(Fig. 7 (e and f)) for the sample prepared using 10 mM solution of
the precursor at 200  C for 1h show flat or standing hexagonal
nanodisks. The well resolved lattice fringes indicate the highly
crystalline nature of the nanocrystals. The d-spacings measured
from the lattice fringes of the different crystallites correspond to
either the orthorhombic or monoclinic Cu7S4 phases. For
example spacing of 1.96 A  (Fig. 7(a)) can be indexed to (2 2 4)
plane of orthorhombic phase or to (0 16 0) plane of monoclinic
phase. The SAED pattern (Fig. 7(g)) contains information from
a large number of nanoparticles and the polycrystalline diffrac-
tion rings at a d-spacing of 1.956 A (strong ring) and 1.691 A can
be indexed to either orthorhombic or monoclinic Cu7S4 phases Fig. 7 (a–d) are TEM images for sample prepared using 20 mM solution
of the precursor at 280  C for 5, 30 and 60 min. (e and f) are for sample
within measurement errors.
prepared using 10 mM solution of the precursor at 200  C for 1h, showing
a flat and standing hexagonal nanodisks, respectively. (g) SAED.

Effect of solvent/capping agent


In all previous experiments the precursor was dissolved in
oleylamine and then injected into hot oleylamine for thermolysis.
Oleylamine was used as both a capping agent and stabilising
agent. In order to understand the role of oleylamine in both size
and shape control, two other experiments were carried out. In the
first experiment the precursor was dissolved in oleylamine and
then injected into hot dodecanethiol and in the other experiment,
the precursor was dispersed in octadecene and then injected into
hot oleylamine. Reactions were carried out at 200  C for 1 h.
Using dodecanethiol as a capping agent, spherical nanoparticles
(Fig. S4(b)) of djurleite (Cu1.94S ICDD card No. 23959) (Fig. 8
(a)) with an average diameter of 11 nm were formed. In previous
reports oleylamine has been suggested to direct the anisotropic
growth of copper sulfide.55 Dodecanethiol was described as
a weaker agent24 with poorer binding to the copper sulfide.14
Octadecene gave nonuniform shapes (Fig. S4(c)) of anilite (Cu7S4
ICDD card No. 33489) (Fig. 8(b)). It has been suggested the
solvent used for precursor injection could change the kinetics of
decomposition of the precursor, leading to different phase of the
nanoparticles.24 Fig. 8 p-XRD of (a) Cu1.94S and (b) Cu7S4.

This journal is ª The Royal Society of Chemistry 2011 J. Mater. Chem.


View Online

Optical properties

All samples showed strong absorption in the UV-blue region and


an increase in absorption in the near-IR region (Fig. S5), which is
consistent with previous reports.19,56–58 Although Cu2-xS systems
have been extensively investigated, there remain uncertainties
regarding its electronic structure.59 The near-IR absorption peak,
which corresponds to intraband transition from valence state to
unoccupied middle-gap state,57 was found to depend on both the
stoichiometry57–60 and the structure of the Cu2-xS nanoparticles.56
Drude-like absorption were subtracted from the data in order to
determine the band gap.60–63 Later it was reported that such
correction would make the determination of the nature of the
band gap, whether direct or indirect, impossible.59 Some authors
Published on 05 October 2011 on http://pubs.rsc.org | doi:10.1039/C1JM13277F

conjectured the possibility of a mixture of the two types of


Downloaded by University of Alabama at Tuscaloosa on 05 October 2011

transition.60,64 Our results for the two pure samples obtained


using dodecanethiol or octadecene (djurleite (Cu1.94S ICDD card
No. 23–959) and anilite (Cu7S4 ICDD card No. 33–489)) are
shown in Fig. 9(a-c). These were used for band gap determina-
tion. The best linear fit was obtained on plotting (ahv)0.5 as
a function of photon energy rather than on plotting (ahv)2 (where
a is the absorption). This observation probably indicates an
indirect band gap. The band gap was found to be 1.73 eV for the
Cu1.94S and slightly lower (1.70 eV) for the Cu7S4. This value is
consistent with the value reported by Kuzuya et al.56 for Cu1.72S.
Except that they reported it as a direct transition. Compared to
bulk djurleite (Cu1.96S) with a band gap of 1.3 eV,65 a blue shift of
0.43 eV is observed in our Cu1.94S sample, indicating nanometer-
sized particles. The near-IR absorption, shows a clear stoichio-
Fig. 10 p-XRD pattern of CuS deposited on glass at (a) 280  C (b)
metric dependence (Fig. 9(a)). Moreover, an increase in the 320  C (c) 360  C (d) 400  C from (2) (* hexagonal Cu2S, # cubic CuS2).
orthorhombic composition was associated with an increase in
near-IR absorption (Fig. S5(c)).
characteristic p-XRD peaks arising from impurities were detec-
Copper sulfide thin films from [Cu(SON(CNiPr2)2)2] ted. The X-ray diffractions peaks of films deposited at 280  C
show almost same intensities of cubic and hexagonal phase
Film deposition was carried out at substrate temperatures whereas, at 320  C the cubic phase was predominant. The major
ranging from 280 to 400  C with argon flow rate of 160 sccm. No peaks of p-XRD can be assigned to the copper sulfide (111),
deposition was obtained below 280  C. Dark yellow adherent (200), (210) and (211) reflections of cubic CuS2 (ICDD: 033–
films were deposited at 280  C and 320  C whereas uniform dark 0492) and (100), (101), (110) and (210) reflections of hexagonal
brown films were deposited at 360 and 400  C. p-XRD pattern of Cu2S (ICDD: 046–1195) as shown in Fig. 10.
as deposited films at 280–400  C (Fig. 10) show the formation of
a mixture of hexagonal Cu2S and cubic CuS2 phase. No

Fig. 9 (a) UV-Vis spectra of Cu1.94S (solid) and Cu7S4 (dashed), (b)
direct (solid) and indirect (dashed) band gap for Cu1.94S and (c) direct Fig. 11 SEM images of copper sulfide films deposited on glass at (a)
(solid) and indirect (dashed) band gap for Cu7S4. 280  C, (b) 320  C (c) 360  C and (d) 400  C.

J. Mater. Chem. This journal is ª The Royal Society of Chemistry 2011


View Online
Published on 05 October 2011 on http://pubs.rsc.org | doi:10.1039/C1JM13277F
Downloaded by University of Alabama at Tuscaloosa on 05 October 2011

Scheme 1 A schematic illustration of all growth experiments.

Fig. 12 (a) SEM and b) HRTEM images of cubic copper sulfide (c)
hexagonal copper sulfide films deposited on glass at 400  C and (d) FFT on plotting (ahv)0.5 as a function of photon energy. The
of (c). morphology of the AACVD thin films depended on deposition
temperatures. TEM analysis of material scratched from thin
films showed that all the films were composed of nanoplates of
The SEM images of the films (Fig. 11) show that the copper sulfides with average width of 25–30 nm. A schematic
morphology of the copper sulfide is dependent on the growth illustration of all growth experiments from [Cu(SON(CNiPr2)2)2]
temperature. Films deposited at 280  C consist of spherical (CuTB) is shown in Scheme 1.
crystallites with average size of 500 nm and those deposited at
320  C show block like crystallites. Films deposited at even
higher temperature (360  C and 400  C) give plate like crystallites
Acknowledgements
with average size of 5 mm. EDX analysis shows the films have A. L. A. gratefully acknowledges financial support from the
a Cu:S ratio of 62 : 38 (280  C), 64 : 36 (320 and 360  C) and Egyptian Cultural Affairs and Missions Sector. K. R. is grateful
65 : 35 (400  C) indicating that they are predominantly hexa- to ORS and The University of Manchester for financial support.
gonal Cu2S. S. J. H would like to thank Professor J. H. Neethling for allowing
The TEM images of the thin films grown at 400  C, show the tomographic data to be collected at the Centre for HRTEM,
nanoplates with width ranging from ca. 30 to 35 nm. HRTEM Nelson Mandela Metropolitan University, South Africa and Mr
images of the nanoplates shows lattice fringes with a d-spacing of A.Yarwood for technical assistance. Financial support from
3.4 A (Fig. 12(c)) corresponding to the (100) reflections of JEOL UK and EPSRC grant number EP/G035954/1 is gratefully
hexagonal Cu2S and 2.9 A  (Fig. 12(b)) corresponding to (211) acknowledged. The authors also thank EPSRC, UK for the
reflections of cubic CuS2. The FFT pattern extract in Fig. 12(d) grants to POB that have made this research possible. POB wrote
shows the single crystalline nature of the nanoplates. this manuscript while a Visiting Fellow at IAS University of
Durham. He would like to thank the University for the
Conclusions Fellowship and Collingwood College and its Fellows for being
gracious hosts.
Solution thermolysis of 1,1,5,5-tetra-iso-propyl-4-thiobiuret
complex of copper(II) produced a mixture of orthorhombic and
Notes and references
monoclinic phases of Cu7S4 nanoparticles. Good quality thin
films of CuS2 and Cu2S were deposited by the AACVD method. 1 H. Lee, S. W. Yoon, E. J. Kim and J. Park, Nano Lett., 2007, 7, 778.
2 Y. Wu, C. Wadia, W. Ma, B. Sadtler and A. P. Alivisatos, Nano Lett.,
The morphology of the crystalline nanoparticles obtained by 2008, 8, 2551.
thermolysis varied as a function of the reaction temperature, 3 C. Wadia, A. P. Alivisatos and D. M. Kammen, Environ. Sci.
precursor concentration, reaction time and capping agent used. Technol., 2009, 43, 2072.
Shorter reaction times produced hexagonal and trigonal nano- 4 T. Sakamoto, H. Sunamura and H. Kawaura, Appl. Phys. Lett., 2003,
82, 3032.
disks whereas, longer reaction times at high temperatures yielded 5 J. S. Chung and H. J. Sohn, J. Power Sources, 2002, 108, 226.
spherical nanoparticles. Different phases and different 6 D. J. Vaughan and J. R. Craig, Mineral Chemistry of Metal Sulfides,
morphologies were also obtained by changing the solvent or the ed. W. B. Harland, S. O. Agrell, A. H. Cook and N. F. Hughes,
Cambridge University Press, Cambridge, 1978, pp. 290–292.
capping agent. The distinction between direct or indirect band
7 M. E. Fleet, Rev. Mineral. Geochem., 2006, 61, 365.
gap for these material is complicated, however, our results 8 D. F. A. Koch and R. McIntyre, J. Electroanal. Chem., 1976, 71, 285.
supports an indirect band gap as the best linear fit was obtained 9 Q. Lu, F. Gao and D. Zhao, Nano Lett., 2002, 2, 725.

This journal is ª The Royal Society of Chemistry 2011 J. Mater. Chem.


View Online

10 X. Jiang, Y. Xie, J. Lu, W. He, L. Zhu and Y. Qian, J. Mater. Chem., 38 F. Srouji, M. Afzaal, J. Waters and P. O’Brien, Chem. Vap.
2000, 10, 2193. Deposition, 2005, 11, 91.
11 W. Du, X. Qian, X. Ma, Q. Gong, H. Cao and J. Yin, Chem.–Eur. J., 39 P. O’Brien and J. Waters, Chem. Vap. Deposition, 2006, 12, 620.
2007, 13, 3241. 40 M. Afzaal, D. J. Crouch, P. O’Brien, J. Raftery, P. J. Skabara,
12 J. Liu and D. Xue, Mater. Res. Bull., 2010, 45, 309. A. J. P. White and D. J. Williams, J. Mater. Chem., 2004, 14, 233.
13 T. H. Larsen, M. Sigman, A. Ghezelbash, R. C. Doty and 41 M. Afzaal, D. Crouch, M. A. Malik, M. Motevalli, P. O’Brien,
B. A. Korgel, J. Am. Chem. Soc., 2003, 125, 5638. J. H. Park and J. D. Woollins, Eur. J. Inorg. Chem., 2004, 171.
14 M. Sigman, A. Ghezelbash, T. Hanrath, A. E. Saunders, F. Lee and 42 S. S. Garje, J. S. Ritch, D. J. Eisler, M. Afzaal, P. O’Brien and
B. A. Korgel, J. Am. Chem. Soc., 2003, 125, 16050. T. Chivers, J. Am. Chem. Soc., 2006, 128, 3120.
15 H. T. Zhang, G. Wu and X. H. Chen, Langmuir, 2005, 21, 4281. 43 M. C. Copsey, A. Panneerselvam, M. Afzaal, T. Chivers and
16 H. T. Zhang, G. Wu and X. H. Chen, Mater. Chem. Phys., 2006, 98, P. O’Brien, Dalton Trans., 2007, 1528.
298. 44 C. Q. Nguyen, A. Adeogun, M. Afzaal, M. A. Malik and P. O’Brien,
17 I. J. Plante, T. W. Zeid, P. Yang and T. Mokari, J. Mater. Chem., Chem. Commun., 2006, 2179.
2010, 20, 6612. 45 C. Q. Nguyen, A. Adeogun, M. Afzaal, M. A. Malik and P. O’Brien,
18 A. Ghezelbash and B. A. Korgel, Langmuir, 2005, 21, 9451. Chem. Commun., 2006, 2182.
19 H. Zhang, Y. Zhang, J. Yu and D. Yang, J. Phys. Chem. C, 2008, 112, 46 K. Ramasamy, M. A. Malik, P. O’Brien and J. Raftery, Dalton
Published on 05 October 2011 on http://pubs.rsc.org | doi:10.1039/C1JM13277F

13390. Trans., 2010, 39, 1460.


20 Y. Wang, Y. Hu, Q. Zhang, J. Ge, Z. Lu, Y. Hou and Y. Yin, Inorg. 47 G. M. Sheldrick, SHELXS-97 and SHELXL-97, University of
Downloaded by University of Alabama at Tuscaloosa on 05 October 2011

Chem., 2010, 49, 6601. G€ ottingen, Germany, 1997.


21 A. Tang, S. Qu, K. Li, Y. Hou, F. Teng, J. Cao, Y. Wang and 48 Bruker, SHELXTL Version 6.12, Bruker AXS Inc., Madison,
Z. Wang, Nanotechnology, 2010, 21, 285602. Wisconsin, USA, 2001.
22 N. Pradhan, B. Katz and S. Efrima, J. Phys. Chem. B, 2003, 107, 49 J. R. Kremer, D. N. Mastronarde and J. R. McIntosh, J. Struct. Biol.,
13843. 1996, 116, 71.
23 B. Geng, X. Liu, J. Ma and Q. Du, Mater. Sci. Eng., B, 2007, 145, 17. 50 Y. Chen, L. Chen and L. Wu, Chem.–Eur. J., 2008, 14, 11069.
24 W. P. Lim, C. T. Wong, S. L. Ang, H. Y. Low and W. S. Chin, Chem. 51 S. Yamamuro, D. F. Farrell and S. A. Majetich, Phys. Rev. B:
Mater., 2006, 18, 6170. Condens. Matter, 2002, 65, 224431.
25 Z. Liu, D. Xu, J. Liang, J. Shen, S. Zhang and Y. Qian, J. Phys. Chem. 52 C. Burda, X. Chen, R. Narayanan and M. A. El-Sayed, Chem. Rev.,
B, 2005, 109, 10699. 2005, 105, 1025.
26 N. Revaprasadu, M. Malik and P. O’Brien, S. Afr. J. Chem., 2004, 57, 53 X. Peng, Adv. Mater., 2003, 15, 459.
40. 54 K. H. Park, K. Jang and S. U. Son, Angew. Chem., Int. Ed., 2006, 45,
27 L. Yu, Y. Lv, G. Chen, X. Zhang, Y. Zeng, H. Huang and Y. Feng, 4608.
Inorg. Chim. Acta, 2011, 376, 659, DOI: 10.1016/j.ica.2011.06.046. 55 R. Si, Y. W. Zhang, H. P. Zhou, L. D. Sun and C. H. Yan, Chem.
28 X. S. Du, Z. Z. Yu, A. Dasari, J. Ma, Y. Z. Meng and Y. W. Mai, Mater., 2007, 19, 18.
Chem. Mater., 2006, 18, 5156. 56 T. Kuzuya, K. Itoh and K. Sumiyama, J. Colloid Interface Sci., 2008,
29 Y. B. He, A. Polity, I. Osterreicher, D. Pfisterer, R. Gregor, 319, 565.
B. K. Meyer and M. Hardt, Phys. B, 2001, 308–310, 1069. 57 M. C. Brelle, C. L. Torres-Martinez, J. C. McNulty, R. K. Mehra and
30 S. Y. Wang, W. Wang and Z. H. Lu, Mater. Sci. Eng., B, 2003, 103, J. Z. Zhang, Pure Appl. Chem., 2000, 72, 101.
184. 58 Y. Zhao, H. Pan, Y. Lou, X. Qiu, J. Zhu and C. Burda, J. Am. Chem.
31 S. D. Sartale and C. D. Lokhande, Mater. Chem. Phys., 2000, 65, 63. Soc., 2009, 131, 4253.
32 H. Hu and P. K. Nair, Surf. Coat. Technol., 1996, 81, 183. 59 P. Lukashev, W. R. L. Lambrecht, T. Kotani and van
33 K. Kemmler, M. Lazell, P. O’Brien, D. J. Otway, J. H. Park and M. Schilfgaarde, Phys. Rev. B: Condens. Matter Mater. Phys., 2007,
J. R. Walsh, J. Mater. Sci.: Mater. Electron., 2002, 13, 531. 76, 195202.
34 P. O’Brien, J. R. Walsh, I. M. Watson, L. Hart and S. R. P. Silva, J. 60 L. D. Partain, P. S. Mcleod, J. A. Duisman, T. M. Peterson,
Cryst. Growth, 1996, 167, 133. D. E. Sawyer and C. S. Dean, J. Appl. Phys., 1983, 54, 6708.
35 P. O’Brien, J. R. Walsh, I. M. Watson, M. Motevalli and 61 B. J. Mulder, Phys. Status Solidi A, 1973, 15, 409.
L. Henrikson, J. Chem. Soc., Dalton Trans., 1996, 2491. 62 B. J. Mulder, Phys. Status Solidi A, 1972, 13, 79.
36 T. Trindade and P. O’Brien, J. Mater. Chem., 1996, 6, 343. 63 B. J. Mulder, Phys. Status Solidi A, 1972, 13, 569.
37 M. Afzaal, D. Crouch, M. A. Malik, M. Motevalli, P. O’Brien and 64 E. Eser and J. A. Cambridge, Sol. Cells, 1982, 5, 343.
J. H. Park, J. Mater. Chem., 2003, 13, 639. 65 B. J. Mulder, Phys. Status Solidi A, 1973, 18, 633.

J. Mater. Chem. This journal is ª The Royal Society of Chemistry 2011

You might also like