You are on page 1of 10

IOP PUBLISHING

NANOTECHNOLOGY

Nanotechnology 22 (2011) 285714 (10pp)

doi:10.1088/0957-4484/22/28/285714

Synthesis of WS2 nanostructures from the


reaction of WO3 with CS2 and mechanical
characterization of WS2 nanotube
composites
M Tehrani1 , C C Luhrs2,5 , M S Al-Haik1 , J Trevino3 and H Zea4
1

Department of Engineering Science and Mechanics, Virginia Tech, Blacksburg, VA 24061,


USA
2
Department of Mechanical and Aerospace Engineering, Naval Postgraduate School,
Monterey, CA 93943, USA
3
Department of Mechanical Engineering, University of New Mexico, Albuquerque,
NM 87131, USA
4
Departamento de Ingeniera Qumica y Ambiental, Universidad Nacional de Colombia,
Bogota, 11001, Colombia
E-mail: ccluhrs@nps.edu

Received 11 January 2011, in final form 19 May 2011


Published 9 June 2011
Online at stacks.iop.org/Nano/22/285714
Abstract
Tungsten disulfide (WS2 ) nanometer sheets, spheres, fibers and tubes were generated by a
synthetic pathway that avoids the use of H2 S as the source of sulfur and employs instead CS2
vapor, carried by an Ar or N2 /H2 stream in a heated tubular furnace, for the reaction with WO3
precursor powders. The experiments were conducted at temperatures between 700 and 1000 C,
while the reaction times expanded between 30 min and 24 h. Characterization methods used to
analyze the products of the synthesis include TEM, SEM, XRD and EDX. We found a strong
correlation between precursor and product microstructure, although the temperature and
reaction times play a critical role in the products microstructural features as well. WS2
inorganic fullerene (IF) nanospheres are generated in a wide window of conditions, while
nanotubes and nanofibers are only produced at high temperatures or long reaction times. A
proposed growth mechanism based on the CS2 synthetic approach is presented.
Nanoindentation and nano-impulse techniques were used to characterize the mechanical
properties of polymer matrixWS2 nanotube composites, finding them superior to equivalent
SWCNT composites. The improvements in toughness of nanocomposites based on WS2 can be
attributed to geometrical and morphological effects that assisted several toughening
mechanisms such as crack pinning and the formation of an immobilized polymeric interphase
around the nanotubes.
(Some figures in this article are in colour only in the electronic version)

posites [15]. WS2 tubes, cages and similar inorganic fullerene


type structures (IF-WS2 ) can be synthesized by diverse
approaches: laser ablation [6, 7], self-assembly [8], template
synthesis [9, 10], hydrothermal reactions [11], metalorganic
chemical vapor deposition (MOCVD) [12, 13], fluidized bed
reactors [1416], spray pyrolysis [17, 18], microwave induced
plasmas [19, 20] and gassolid reactions [3, 6, 21]. The

1. Introduction
Nanomaterials consisting of various tungsten disulfide (WS2 )
structures have proven very useful in several applications including lubricants, catalysts, coatings and shock resistant com5 Author to whom any correspondence should be addressed.

0957-4484/11/285714+10$33.00

2011 IOP Publishing Ltd Printed in the UK & the USA

Nanotechnology 22 (2011) 285714

M Tehrani et al

different synthetic methods produce slightly different WS2


structures, depending on the precursor used and the rate
of WS2 formation. (NH4 )2 WS4 thermal decomposition and
WO3 reaction with sulfur-containing compounds are among
the most common routes employed. All the reactions that
utilize WO3 as reactant have similar mechanisms; first WO3
is reduced with hydrogen, then shearing of atomic planes
at the surface promotes the reaction of tungsten with sulfurcontaining compounds and finally the sulfur atoms replace the
oxygen ones to form WS2 [22]. Gassolid reactions constitute
a very simple approach to generate WS2 ; in most cases WO3
is reacted with a sulfur-containing compound (usually H2 S)
during extended periods of time at high temperature.
In this investigation, we employ a similar approach to
produce WS2 nanostructures except for the use of H2 S. We
studied the reaction of WO3 with CS2 in diverse temperature
atmospherereaction time conditions to determine which
settings will favor the generation of tubular, spherical or sheet
like nanostructures.
To our knowledge, the use of CS2 to generate WS2
has been limited to methods that involve: (i) indirect routes
that require additional sulfur sources, (ii) intermediate steps
entailing the formation of H2 S or (iii) the use of CS2 for
the generation of products with restricted shapes and sizes,
such as plate like micrometer products. Examples of these
include: use of hydrothermal conditions along with ammonia
and water in order to produce H2 S that will act as the sulfur
source [23, 24], use of CS2 along with CCl4 as liquid media in
which S and WO3 will be added to form a mixture that will then
be heated and sonicated before being dried and heated under
H2 atmosphere [3] and W metal powder and WO3 reaction
with flowing CS2 to generate plate like micrometer size WS2
products [25].
Regarding the mechanical properties, WS2 has been
shown to stand shockwaves as high as 25 GPa and elevated
temperatures without any significant structural degradation
or phase change [2]. The single WS2 nanotube has an
elastic modulus of 150170 GPa and strain to fracture of
15% [4]. These outstanding mechanical properties make
WS2 nanostructures very promising for structural applications.
Although the WS2 single nanotube mechanical properties
are not as high as those for single wall carbon nanotubes
(SWCNTs), under similar conditions WS2 nanotubes have
shown to be more stable than SWCNTs when exposed to shock
waves [26]. While there are several studies discussing the
mechanical properties of WS2 structures [27, 28], there is no
cited literature on investigating the mechanical aspects of WS2
nanocomposites at the nanoscale.
We performed nanoindentation and nano-impulse tests on
composites made out of WS2 nanotubes with an epoxy matrix.
The reduced modulus, nanohardness and low velocity impact
resistance (dynamic hardness) of the WS2 -nanotube-based
composites are measured and compared to the ones based on
SWCNTs. Therefore, the current investigation provides, for
the first time, a comparative study of the impact resistance
and quasi-static nanoindentation properties for two polymeric
composite systems based on SWCNTs and WS2 nanotubes,
respectively.

2. Experimental methods
2.1. The generation of WS2 structures
The window of temperatures and time conditions used in
this study was based on preliminary results that revealed the
minimal temperature and time for which WS2 can be obtained
as a pure phase from the reaction of WO3 and CS2 (according
to XRD analysis of products generated in diverse conditions
where no WO3 reflections were detected).
Experiments were conducted between 700 and 1000 C
with treatments that extended from 30 min to 24 h. The gas
solid reaction was performed under inert (argon) and reducing
atmospheres (N2 /H2 : 93%/7%). Flow rates of 20 slpm were
used. Commercial WO3 particles (both micron and nanosizes
<100 nm (TEM), Sigma-Aldrich) were used as tungsten
precursor and CS2 (ACS reagent >99.9%, Sigma-Aldrich) as
sulfur source. The synthesis was carried out as follows. WO3
precursor nanoparticles were placed in a sintered alumina boat,
then positioned into a quartz tube whose ends were equipped
with high temperature o-rings and stainless steel lids for gas
introduction and exhaust. A gas purge step was performed
in order to assure that no oxygen was contained inside the
reactor. The quartz tube was then introduced into a clamp
tubular furnace and the temperature was raised to the desired
set point. A stream of the carrier gas was then directed to the
surface of liquid CS2 which was contained in a trap located
between the gas source and the furnace. The mixture of carrier
gas (Ar or N2 /H2 ) with CS2 was introduced into the quartz
tube for the reaction to occur. The reactor for the synthesis
is shown in figure 1. Once the reaction time was over, the
valve connecting the CS2 trap was closed and the powders were
allowed to cool only under carrier gas.
2.2. The preparation of WS2 nanotube composites
To examine the mechanical properties of WS2 nanotube
polymeric composites, WS2 nanotube epoxy composites were
prepared. SWCNT polymeric composites were prepared in
identical conditions for comparison. Short SWCNTs from
CheapTubes.com with a purity of >95%, outside diameter of
12 nm and average length of 0.52 m were used. The epoxy
comprised PR2032 resin and PH3660 hardener (PTM&W
Industries, Inc.). This epoxy system is a medium viscosity
system, mostly used in structural applications. The system
cures at room temperature for 24 h. The properties of the epoxy
system are published at PTM&W Industries, Santa Fe Springs,
CA 90670-4092, USA.
Due to a high aspect ratio and intrinsic van der Waals
attraction, the WS2 and carbon nanotubes form agglomerated
bundles and ropes. Ultrasonication has been shown to be
an effective way of moderately dispersing carbon nanotubes
in some low viscosity liquids [29]. Hence, the nanotube
samples (WS2 and SWCNTs) were independently added
to an ethanol solution 1:10 by weight, and sonicated for
2 h using an ultrasonic cleaner at 40 kHz and 700 W
power. The suspensions were then separately mixed with
the hardener and sonicated for one more hour, after which
vacuum was applied until the alcohol content was completely
2

Nanotechnology 22 (2011) 285714

M Tehrani et al

Figure 1. Schematic representation and image of WS2 synthesis setup for the direct reaction between WO3 and CS2 .

evaporated. The nanotubes/hardener suspensions were added


to the resin. Further mixing and dispersion of the nanotubes
were performed by simultaneously using a mechanical stirrer
and a sonicator for few minutes. Both the WS2 and SWCNT
composites contained 3 wt% of the corresponding nanotubes.
The short gelling time of the epoxy is a limiting factor for
longer dispersal of the nanotubes; therefore, the samples had
to be degassed right after the resin was added to prevent any
air bubbles from getting trapped in the nanocomposite. The
samples were left to cure for 24 h at room temperature.
The procedure followed for the nanotube dispersion was
based on our previous work [30, 31], where the methodology
proved effective in dispersing CNTs in the same epoxy
system. We limited the loading percentage to 3% with
the goal of having a point for comparison between the two
composites while observing the percolation limit for the
SWCNTs. Many researchers have posed a limit for dispersing
SWCNTs in highly viscous material using sonication or
calendaring [32, 33]; it has been found that beyond a certain
volume fraction (usually around 3%), no further improvements
in the composites mechanical properties are achieved
using the dispersion methods previously mentioned. At
higher volume fractions, dispersion and alignment deteriorate
significantly.

TEM characterization was performed on a JEOL 2010


high resolution transmission electron microscope (HRTEM)
equipped with an EDS analysis detector. SEM analysis was
conducted on a Hitachi S-5200 Nano SEM working at 10 kV.
2.4. WS2 nanotube composite mechanical properties testing
Specimens of each SWCNT and WS2 -nanotube-based composite were tested using a Nano Test 600 mechanical
nanocharacterization system (Micro Materials Ltd, UK).
Nanoindentation tests were performed using a 5 m radius
spherical diamond tip. A total of 10 load-controlled to 3 mN
maximum load indentations were carried out on different
spots on each nanocomposite specimen. The initial load and
loading/unloading rates were set to 0.03 mN and 0.1 mN s1
respectively. A holding period of 60 s at the peak load was used
and the data were corrected for thermal drift. These parameters
were chosen based on the results of the pre-nanoindentation
tests as described in [34]. OliverPharr analysis was used to
extract the reduced modulus and hardness [35]. Contact area
measurements from indenting a fused silica reference sample
at different depths were utilized in the analysis.
A pendulum impulse impact test enables repetitive
impacts to be produced at specified time intervals at the same
location with precisely controlled force, using a solenoid and
timed relay. The nano-impulse module allows for repetitive
impacts of the energized diamond probe into the material
surface. The indenter tip was accelerated in less than a tenth
of a second from a distance of 12.5 m to the surface of
the sample to produce a load of 3 mN. The repetitive impact
was performed 45 times in each spot with the diamond probe
being two seconds in contact and two seconds off the surface.

2.3. Sample characterization


The samples were analyzed using a Scintag Pad V diffractometer/goniometer with Scintillation detector, Datascan software
(Materials Data, Inc.) for diffractometer automation and data
collection, and Jade Software (Versions 9, also from MDI) for
data analysis.
3

Nanotechnology 22 (2011) 285714

M Tehrani et al

Figure 2. Electron microscopy images of diverse WS2 nanostructures generated from the reaction of CS2 , used as sulfur source, with WO3 .
(a) SEM image of nanotubes, (b) TEM image of nanofibers, (c) SEM micrograph of thin sheets, (d) TEM of cage like nanoparticles and
(e) relative amount of each microstructure for samples produced under reducing atmospheres at diverse temperatures and times of treatment.

some nanofibers whose maximum lengths were <70 nm.


Experiments performed employing short reaction times, less
than 3 h, formed only extremely thin sheets of WS2 , that
resembled the structure of carbon materials such as graphene.
The generation of short fibers from micron WO3 and CS2
carried out under inert environments required the use of longer
reaction times, of the order of 24 h, and produced mixtures of
structures: a high yield of thin sheets and plates and very low
yields of fibers. Nanometer hollow spherical or cage like WS2
particles, with the occasional appearance of short fibers, were
observed only when micrometer WO3 was reacted with CS2 in
reducing atmospheres. No nanotubes were obtained by either
the use of inert atmospheres or micron size oxide precursors.

Measurement of the volume change between the first and last


strikes is used as a qualitative metric for the energy absorption
of the different samples. A total of five repetitive nano-impulse
tests were carried out on each sample. Impact resistance is then
measured as the ability of the material to undergo less plastic
deformation under repetitive impact [36, 37].

3. Results and discussion


3.1. Microstructural features of WS2 products
The use of micrometer WO3 powder precursor and inert
atmospheres for the reaction of WO3 with CS2 , rendered
WS2 micrometer plate like structures, thin sheets and
4

Nanotechnology 22 (2011) 285714

M Tehrani et al

Figure 4. XRD analysis of nanotube samples found after 3 h of


reaction at 900 C shows reflections that can be identified as
tungstenite, WS2 .

Similar features to those of the thin sheets presented


herein were observed in quasi-two-dimensional graphene like
WS2 structures, whose synthesis was recently accomplished by
exfoliation and intercalation routes [38].
The use of reducing atmospheres greatly increased the
amount of inorganic fullerene type and nanotubular structures
generated. The microstructure of the synthesized particles was
profoundly affected by the temperature and the duration of the
experiment as well. Thin sheets were more apparent when
the WO3 nano-powder synthesis ran for short periods of time.
Nanotubes were more abundant when the WO3 was exposed
for times longer than 6 h and/or temperatures equal to 800 C
or higher.
Spherical particles were observed to form during all of
the time/temperature combinations and had the tendency to
form micrometer size clusters. WS2 nanofibers co-existed
among these other structures; although never seen as the
predominant microstructure when reducing conditions were
used. Nanotubes were the most abundant microstructural type
for samples annealed for 24 h.
The most abundant structures generated from the reaction
of CS2 , used as sulfur precursor, and WO3 under reducing
These
conditions were IF-WS2 spherical nanoparticles.
cage like materials possessed different morphologies such as
perfect spheres, semispherical particles or structures in which
crystalline directions were evident and resulted in polyhedral
shapes. IF type WS2 nanoparticles were prone to agglomerate
and, while nanometer features were still observable, the
clusters acquired hollow micron size spherical shapes. An
SEM image of WS2 nanospheres as part of a cluster is
presented in figure 3(a). The hollow center of the WS2
spherical nanostructures was consistently observed by TEM.
See figure 3(b).
XRD analysis of the samples revealed that the complete
transformation to WS2 occurs after 3 h of reaction in reducing
environment at the CS2 vaporgas flow conditions used.
Figure 4 shows the x-ray diffraction pattern of a sample in
which nanotubular structures dominated the microstructure.
One could observe that the main peaks corresponded to
tungstenite and only a small intensity peak located close to 24
(*) remained for WO3 .

Figure 3. IF-WS2 spherical nanoparticles with average diameters


close to 50 nm and hollow structures tend to agglomerate to form
hollow micron size spheres. (a) SEM image of WS2 nanospheres
forming clusters and (b) WS2 hollow spherical nanostructures
observed by TEM.

WO3 micron-sized powder did not react at low temperatures


for short periods of time.
When starting with WO3 nanoparticles as precursor,
the nanoscale features in the starting materials aided the
generation of WS2 products with features in the nanometer
range. Electron microscopy observations of samples generated
by the WO3 nanoparticle reaction with CS2 corroborated the
existence of diverse WS2 nanostructures: tubes, spherical IF
particles, fibers and thin sheets. Figure 2 presents electron
microscopy images of some of those features: (a) nanotubes,
with diameters in the nanoscale but lengths in the micron
scale, with hollow centers and multiwall structures, (b) short
fibers of no more than 70 nm in length and with a different
appearance than nanotubes, no hollowness observed, were
occasionally identified as well, (c) thin sheets of WS2 were
spotted under certain experimental conditions, the thickness of
the sheets was in most cases smaller than 3 nm, and (d) cage
like nanoparticles, with diameters of the order of 5070 nm and
the characteristic interplanar spacing for IF WS2 nano-onions.
Figure 2(e) presents the relative amounts of each of these
microstructural features for diverse experimental conditions
using reducing atmospheres. These relative amounts were
calculated from SEM micrographs; images of multiple regions
of each sample were obtained and the percentages of area
occupied by nanoparticles, nanotubes, fibers or thin sheets with
respect to the total determined.
5

Nanotechnology 22 (2011) 285714

M Tehrani et al

Table 1. Nanoindentation test results.


SWCNT-based
WS2 -based composite composite
Hardness (GPa)
0.190 06 0.01363
Reduced modulus (GPa) 4.438 0.234

0.173 88 0.00240
3.487 0.054

first step is usually correlated with the oxide reduction, which


promotes shearing of surface atomic planes, followed by
a sulfidization step that forms closed shell sulfide layers.
Subsequent propagation of hydrogen within the particulate
structure will reduce further oxide layers, followed by a
sulfidization process of the same, with an inward reaction
front.
As mentioned earlier, CS2 will react with WO3 to form
WS2 in inert atmospheres although nanotubes will not be
generated under such conditions. In order to form nanotubes
using CS2 , reducing environments are required, in the case
of our experiments supplied by a mixture of 7% H2 in N2 .
In a similar fashion to that reported for H2 S-based reactions,
the synthesis of WS2 nanotubes and nanospheres from CS2
implies a delicate balance between alternate reduction and
sulfidization steps. TEM observations of nanotubes evidenced
this process: the nanotubes presented multiwall structures
that were commonly divided in sections in which the internal
walls collapsed forming caps at regular spacings, while the
external walls growth continued. The nanotubes internal
walls seem to form the closed structures prior to the growth
of the external layers, as TEM observations of the tubes tips
revealed, supporting the idea of oxide material condensing in
such points, followed by its reduction and sulfidization, see
figure 6. In contrast, synthesis performed in inert atmospheres
will favor the exfoliation of layers and formation of thin WS2
sheets.
Electron microscopy images of the nanotubes show that
they are not aligned with each other and they seem to contain
segments that change their growth direction multiple times,
suggesting a spiral or step winding growth.
3.3. Mechanical properties
Fractured surfaces of the nanocomposites were coated with
a thin layer of gold and observed under SEM. SWCNT and
WS2 nanotube dispersions are presented in figures 7(a) and (b).
Figure 7(c) shows two representative nanoindentation curves
for the two composites based on SWCNT and WS2 nanotubes.
Table 1 presents the results of OliverPharr analysis [35]
of the nanoindentation tests. The results were averaged over 10
tests for each sample. The neat sample has a reduced modulus
of 3.47 0.25 GPa and hardness of 0.1918 0.016 GPa [41].
These results revealed that the hardness and Youngs
modulus of the WS2 -based composite are 10% and 27%
better than the corresponding values for the SWCNTbased composite. Sample nano-impulse tests of the two
nanocomposites are shown in figures 8(a) and (b) while crate
volume changes are presented in table 2.
Based on this analysis the impact resistance of the WS2 based composite is roughly 28% higher than the corresponding
value for the SWCNT-based composite.

Figure 5. (a) WS2 nanotubes are commonly observed by SEM in


samples that also present micron size agglomerates of IF-WS2 ,
(b) the nanotubes lengths are of the order of several micrometers,
with average diameters of 60 nm, (c) the tubes growth proceeds in
multiple directions (not aligned).

WS2 nanotubes were frequently observed by SEM analysis in samples that also presented micron size agglomerates of
IF-WS2 , i.e. reaction times of 12 h and temperatures of 900 C,
figure 5. The nanotubes lengths were of the order of several
micrometers, with average diameters of 60 nm.
3.2. Proposed growth mechanisms
Tenne and co-workers have identified the basic steps in the
WS2 generation from WO3 and H2 S [16, 22, 39, 40]. The
6

Nanotechnology 22 (2011) 285714

M Tehrani et al

Table 2. Nano-impulse test results.


Crate volume change (nm3 )
SWCNTs-based composite
WS2 -based composite

(1.69 0.12) 1011


(1.21 0.15) 1011

micrographs shown in figures 7(a) and (b) a reasonable state of


dispersion for both nanocomposite systems exists, under which
the WS2 performs better than the SWCNTs in both quasi-static
and high strain rate conditions.
It is well known that the mechanical properties of
SWCNTs are much higher compared to those of WS2 . For
example the Youngs modulus for WS2 nanotubes was found
to be in the range of 140175 GPA [7] compared to that
of SWCNTs, 1.2 TPA [8]. The ultimate strength of
WS2 nanotubes with an average diameter of 10 nm was
found to be in the range of 3.716.3 GPa [9] compared
to 250 GPa for SWCNTs [8].
Despite the superior
strength and elastic properties of the SWCNTs [43, 44],
these properties are not directly transmitted to their derived
polymeric composites. Dispersion and deagglomeration are
typically dominating factors in limiting the enhancement of
mechanical properties [4].
Furthermore, several mechanisms that stem from the
size and morphology of the nanofillers control the overall
fracture toughness of the nanocomposites. For example, it is
well recognized that materials that can exhibit crack pinning
typically will possess higher fracture resistance. A crack
pinning mechanism has been invoked for nanoparticles [45].
This is analogous to the restricted movement of dislocations
through metals by incorporating much harder and stronger
particles (dispersion hardening). The crack pinning occurs
when the toughening particles in the polymeric composite are
larger than the crack-opening displacement. It is likely that
the SWCNTs (12 nm in diameter) are much smaller than
the crack-opening displacement and thus they are unlikely to
cause crack pinning. WS2 nanotubes, on the contrary, are
much larger (70 nm diameter) than SWCNTs and thus they are
more likely to pin the cracks movement and thus enhance the
toughness.
Another factor that plays a role in enhancing the toughness
of the nanocomposite is the formation of an interphase or
immobilized layer of polymer around the agglomerates. This
layer can exhibit thicknesses anywhere in the range of tens of
nanometers to 1.4 m [46]. SWCNTs tend to agglomerate
more than WS2 nanotubes upon introducing them into the
epoxy matrix; therefore their interparticle distance in the
derived nanocomposites is larger even for equal loadings.
However, for WS2 , as a consequence of a better dispersion, the
immobilized interphase may be present throughout the epoxy
matrix causing an improved crack resistance.

Figure 6. TEM observations of the nanotube samples show: (a) areas


of the sample in which nanotubes grow from clusters of WS2
semispherical particles that do not present evidence of crystalline
components, (b) the tubes present multiwall structures that
commonly are divided into sections in which the internal walls
collapse forming caps while the external walls are always
continuous. (c), (d) The nanotubes internal walls seem to form the
closed structures prior to the growth of the external layers, as
evidenced in the tubes tips.

4. Conclusions

The strengthening mechanisms in the epoxy systems


strongly rely upon the degree of dispersion of the nanotubes
as well as the extent of cohesive interactions between the
tubes and polymer chains [29, 42]. According to the SEM

The work outlined in this paper shows that the synthesis


of WS2 nanostructures can be successfully performed using
CS2 as sulfur providing agent, in both inert and reducing
7

Nanotechnology 22 (2011) 285714

M Tehrani et al

Figure 7. Scanning electron micrograph of (a) SWCNT/epoxy and (b) WS2 /epoxy nanocomposite fracture surfaces. (c) Loading/unloading
hysteresis of the two nanocomposite systems, WS2 and SWCNTs, respectively.

atmospheres. The former environment used with micron size


WO3 precursors yields mainly plate like and extremely thin
WS2 sheets. Long reaction times, i.e. 24 h, for the same inert
atmosphere produce spherical IF nanoparticles and very few
fibers. No nanotubes are generated from micron precursors
in inert environments. The precursor microstructure played
a large role in determining the outcome regarding the newly
formed WS2 features, such as the particles shapes and sizes.
Use of WO3 precursor nano-powders along with the
reducing environments was indispensable in order to generate
significant amounts of IF WS2 nanospheres and nanotubes.
The microstructure of the synthesized particles was greatly
affected by the temperature and the duration of the experiment.
Nanotubes were detected when the WO3 CS2 reaction
occurred for times longer than six hours and/or temperatures
equal to 800 C or higher. Nanosheets were more apparent
when the WO3 nano-powder synthesis ran for short periods
of time. Nanospheres were formed during most of the
time/temperature combinations and have the tendency to
form micrometer size clusters. Nanofibers were observed
along diverse structures and were never the predominant
microstructure. The growth mechanisms are similar to the
ones reported for H2 S-based reactions, with reduction and
sulfidization process dominating the growth of the nanotubular
structures.
Nanoindentation tests revealed that the WS2 -based
composite hardness and reduced modulus are 10% and

Figure 8. Sample nano-impulse tests of (a) WS2 and (b) SWCNT


nanocomposites, respectively.

Nanotechnology 22 (2011) 285714

M Tehrani et al

27% better than the corresponding values for the SWCNTbased composite, respectively.
The nano-impulse tests
indicate that the impact resistance of the WS2 -nanotubebased composite is roughly 28% higher compared to the
SWCNT-based composite. Despite the superior properties
of SWCNTs, several processing (dispersion), geometrical
and morphological effects can restrict their enhancement
of the fracture toughness.
WS2 -based nanocomposites,
on the contrary, due to the ease of dispersion and the
dimension induced mechanisms such as crack pinning and
immobilized epoxy interphase, are tougher than the SWCNTbased composites.

[16] Margolin A, Rosentsveig R, Albu-Yaron A,


Popovitz-Biro R and Tenne R 2004 Study of the growth
mechanism of WS2 nanotubes produced by a fluidized bed
reactor J. Mater. Chem. 14 61724
[17] Bastide S et al 2004 Synthesis of inorganic fullerenes and
nanoboxes of MOS2 and WS2 by spray pyrolysis Abst.
Papers Am. Chem. Soc. 228 200-COLL U482
[18] Bastide S, Duphil D, Borra J and Levy-Clement C 2006 WS2
closed nanoboxes synthesized by spray pyrolysis Adv. Mater.
18 1069
[19] Vollath D and Szabo D V 1998 Synthesis of nanocrystalline
MoS2 and WS2 in a microwave plasma Mater. Lett.
35 23644
[20] Brooks D J, Douthwaite R E, Brydson R, Calvert C,
Measures M G and Watson A 2006 Synthesis of inorganic
fullerene (MS2, M = Zr, Hf and W) phases using H2S and
N-2/H-2 microwave-induced plasmas Nanotechnology
17 124550
[21] Chen J, Li S, Gao F and Tao Z 2003 Synthesis and
characterization of WS2 nanotubes Chem. Mater. 15 10129
[22] Tenne R, Homyonfer M and Feldman Y 1998 Nanoparticles of
layered compounds with hollow cage structures (inorganic
fullerene-like structures) Chem. Mater. 10 322538
[23] Chen X, Wang X, Wang Z, Yu W and Qian Y 2004 Direct
sulfidization synthesis of high-quality binary sulfides (WS2 ,
MoS2 , and V5 S8 ) from the respective oxides Mater. Chem.
Phys. 87 32731
[24] Afanasiev P 2008 Synthetic approaches to the molybdenum
sulfide materials C. R. Chim. 11 15982
[25] Shin W, Yoon D H and Kim S J 1998 Synthesis of tungsten
disulfide solid lubricant by solidgas reactions J. Mater. Sci.
Lett. 17 17314
[26] Zhu Y Q et al 2003 Shock-wave resistance of WS2 nanotubes
J. Am. Chem. Soc. 125 132933
[27] Roy M, Koch T and Pauschitz A 2010 The influence of
sputtering procedure on nanoindentation and nanoscratch
behaviour of WSC film Appl. Surf. Sci. 256 68508
[28] Voevodin A A, ONeil J P and Sabinski J S 1999
Nanocomposite tribological coatings for aerospace
applications Surf. Coat. Technol. 116 3645
[29] Song Y S and Youn J R 2005 Influence of dispersion states of
carbon nanotubes on physical properties of epoxy
nanocomposites Carbon 43 137885
[30] Choi E S et al 2003 Enhancement of thermal and electrical
properties of carbon nanotube polymer composites by
magnetic field processing J. Appl. Phys. 94 60349
[31] Garmestani H et al 2003 Polymermediated alignment of
carbon nanotubes under high magnetic fields Adv. Mater.
15 191821
[32] Jin L, Bower C and Zhou O 1998 Alignment of carbon
nanotubes in a polymer matrix by mechanical stretching
Appl. Phys. Lett. 73 11979
[33] Gojny F H, Wichmann M H G, Kopke U, Fiedler B and
Schulte K 2004 Carbon nanotube-reinforced
epoxy-composites: enhanced stiffness and fracture
toughness at low nanotube content Compos. Sci. Technol.
64 236371
[34] Tehrani M, Safdari M and Al-Haik M S 2011
Nanocharacterization of creep behavior of multiwall carbon
nanotubes/epoxy nanocomposites Int. J. Plast. 27 887901
[35] Oliver W C and Pharr G M 1992 An improved technique for
determining hardness and elastic modulus using load and
displacement sensing indentation experiments J. Mater. Res.
7 156483
[36] Beake B D and Smith J F 2004 Nano-impact testing-an
effective tool for assessing the resistance of advanced
wear-resistant coatings to fatigue failure and delamination
Surf. Coat. Technol. 188 5948

Acknowledgments
The authors greatly acknowledge the support of the National
Science Foundation (NSF) Awards # EEC-0741525 and NSFCMMI-0846589.

References
[1] Espino J, Alvarez L, Ornelas C, Rico J L, Fuentes S,
Berhault G and Alonso G 2003 Comparative study of WS2
and Co(Ni)/ WS2 HDS catalysts prepared by ex situ/in situ
activation of ammonium thiotungstate Catal. Lett. 90 7180
[2] Zhu Y Q et al 2005 Shock-absorbing and failure mechanisms of
WS2 and MoS2 nanoparticles with fullerene-like structures
under shock wave pressure J. Am. Chem. Soc. 127 1626372
[3] Yang H et al 2006 Synthesis of inorganic fullerene-like WS2
nanoparticles and their lubricating performance
Nanotechnology 17 15129
[4] Kaplan-Ashiri I and Tenne R 2007 Mechanical properties of
WS2 nanotubes J. Cluster. Sci. 18 54963
[5] Rosentsveig R et al 2009 Fullerene-like MoS2 nanoparticles
and their tribological behavior Tribol. Lett. 36 17582
[6] Nath M, Govindaraj A and Rao C N R 2001 Simple synthesis
of MoS2 and WS2 nanotubes Adv. Mater. 13 2836
[7] Schuffenhauer C et al 2005 Synthesis of fullerene-like tantalum
disulfide nanoparticles by a gas-phase reaction and laser
ablation Small 1 11009
[8] Remskar M et al 2001 Self-assembly ofsubnanometer-diameter
single-wall MoS2 nanotubes Science 292 47981
[9] Whitby R L D et al 2002 Complex WS2 nanostructures Chem.
Phys. Lett. 359 6876
[10] Deng H, Chen C, Peng Q and Li Y 2006 Formation of
transition-metal sulfide microspheres or microtubes Mater.
Chem. Phys. 100 2249
[11] Shang Y, Xia J, Xu Z and Chen W 2005 Hydrothermal
synthesis and characterization of quasi-1-D tungsten
disulfide nanocrystal J. Dispers. Sci. Technol. 26 6359
[12] Zink N, Therese H A and Pansiot J 2008 In situ heating TEM
study of onion-like WS2 and MoS2 nanostructures obtained
via MOCVD Chem. Mater. 20 6571
[13] Zink N, Pansiot J and Kieffer J 2007 Selective synthesis of
hollow and filled fullerene-like (IF) WS2 nanoparticles via
metalorganic chemical vapor deposition Chem. Mater.
19 6391400
[14] Rosentsveig R, Margolin A, Feldman Y, Popovitz-Biro R and
Tenne R 2002 Bundles and foils of WS2 nanotubes Appl.
Phys. A 74 3679
[15] Rosentsveig R, Margolin A, Feldman Y, Popovitz-Biro R and
Tenne R 2002 WS2 nanotube bundles and foils Chem. Mater.
14 4713

Nanotechnology 22 (2011) 285714

M Tehrani et al

[42] Zhang S G et al 2003 Use of functionalized WS2 nanotubes to


produce new polystyrene/polymethylmethacrylate
nanocomposites Polymer 44 210915
[43] Yakobson B I and Avouris P 2001 Mechanical properties of
carbon nanotubes Carbon Nanotubes (Topics in Applied
Physics vol 80) pp 287327
[44] Kaplan-Ashiri I et al 2004 Mechanical behavior of individual
WS2 nanotubes J. Mater. Res. 19 4549
[45] Johnsen B B, Kinloch A J, Mohammed R D, Taylor A C and
Sprenger S 2007 Toughening mechanisms of
nanoparticle-modified epoxy polymers Polymer 48 53041
[46] Ragosta G, Abbate M, Musto P, Scarinzi G and Mascia L 2005
Epoxy-silica particulate nanocomposites: chemical
interactions, reinforcement and fracture toughness Polymer
46 1050616

[37] Beake B D, Goodes S R, Smith J F and Gao F 2004 Nanoscale


repetitive impacttesting of polymer films J. Mater. Res.
19 23747
[38] Matte H S et al 2010 MoS2 and WS2 analogues of graphene
Angew. Chem. Int. Edn 49 405962
[39] Rothschild A, Sloan J and Tenne R 2000 Growth of WS2
nanotubes phases J. Am. Chem. Soc. 122 516979
[40] Tenne R and Redlich M 2010 Recent progress in the research of
inorganic fullerene-like nanoparticles and inorganic
nanotubes Chem. Soc. Rev. 39 142334
[41] Al-Haik M S, Trinkle S, Momotyuk O, Roeder B T,
Kemper K and Hussaini M Y 2007 Nanocharacterization of
proton radiation damage on magnetically oriented epoxy Int.
J. Polym. Anal. Charact. 12 41330

10

You might also like