You are on page 1of 6

Materials Chemistry and Physics xxx (2013) 1e6

Contents lists available at ScienceDirect

Materials Chemistry and Physics


journal homepage: www.elsevier.com/locate/matchemphys

Bio-assisted synthesis and characterization of nanostructured bismuth


(III) sulphide using Clostridium acetobutylicum
Sathish Kumar Kamaraj, Ganesh Venkatachalam, Palaniappan Arumugam,
Sheela Berchmans*
Electrodics and Electrocatalysis (EEC) Division, CSIR e Central Electrochemical Research Institute (CSIR-CECRI), Karaikudi 630006, Tamilnadu, India

h i g h l i g h t s

g r a p h i c a l a b s t r a c t

 Environmentally benign (greener)


route towards synthesis of Bi2S3
nanostructures.
 Bio-assisted synthesis of Bi2S3 at
room temperature using Clostridium
acetobutylicum.
 Extracellular proteins in H2S producing microorganism as stabilizer for
Bi2S3 NPs.
 Hexagonal platelets of Bi2S3 possessing an orthorhombic crystalline
structure.

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 6 May 2012
Received in revised form
19 November 2013
Accepted 22 November 2013

Nanostructured bismuth (III) sulphide is synthesized at room temperature using a hydrogen sulphide
producing microorganism namely Clostridium acetobutylicum. On contrary to chemical routes involving
both the high and room temperature methods, the present experimental procedure involves a bioassisted approach. This method is free from the usage of toxic and hazardous chemicals making it an
environment friendly route. The synthesized bismuth sulphide is characterized using transmission
electron microscope (TEM), powder X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS) and
cyclic voltammetry (CV). From our experiments we nd that bismuth sulphide produced using this bioassisted approach exhibits a hexagonal shaped plate-like structures and is stabilized by the extracellular
proteins present in the culture medium.
2013 Elsevier B.V. All rights reserved.

Keywords:
Electronic materials
Nanostructures
Electrochemical techniques
Semiconductors

1. Introduction
Nanomaterials play a crucial role in technology development;
while the use of nanomaterials become indispensable, it is required
to change the current synthetic methodologies in order to considerably cut down the production cost of these materials [1,2].
Though industries appreciate the cost reduction alternate approaches, it is very important to focus on protecting the environment too. With European Union (EU) emphasizing on controlling
environment pollution and ways to bring down global warming
* Corresponding author. Tel.: 91 4565 241485; fax: 91 4565 227779.
E-mail address: sheelaberchmans@yahoo.com (S. Berchmans).

etc., recent research is mainly focused towards developing greener


routes to obtain nanomaterials [3]. Cost reduction, environment
protection, and global warming guide researchers to adapt new
synthetic methodologies that primarily avoid the usage of toxic
chemicals, organometallic precursors, hazardous reductants, solvents etc., to design, synthesis and fabricate nanomaterials.
One of the well known green chemistry routes in research
community is bio-assisted synthesis and assembly of nanoparticles
(NPs). Though this methodology is well developed for metal NPs, it
is still under an exploration stage for metal oxides, semiconductors
and magnetic NPs. Semiconducting nanomaterials are an important
class of materials, which nd applications in many elds including
electronics, sensors, renewable energy, energy storage etc [4e6]. Of

0254-0584/$ e see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.matchemphys.2013.11.042

Please cite this article in press as: S.K. Kamaraj, et al., Bio-assisted synthesis and characterization of nanostructured bismuth (III) sulphide using
Clostridium acetobutylicum, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.11.042

S.K. Kamaraj et al. / Materials Chemistry and Physics xxx (2013) 1e6

the available semiconductors, bismuth sulphide (Bi2S3) is a unique


material having a direct band gap energy of 1.3e1.7 eV [7]. As far as
the toxicity is concerned Bi2S3 nanomaterials is less toxic when
compared to other semiconductors of technological importance
such as CdS, CdSe, InP, InAs, InGa, PbS, PbSe etc. It has been shown
to possess potential applications in various elds such as photodiode arrays, photovoltaics, thermoelectric technology etc. It is also
known that thin lms of Bi2S3 are a good light harvesting substrate
exhibit absorption in the visible and near-IR region of the solar
spectrum, which allows their use in photodiode arrays and
photovoltaic converters [8e11].
In general, chemical routes to Bi2S3 nanomaterials synthesis
often employ high temperature, toxic reducing agents, hazardous
chemicals and solvents. Recently Rabin et al. reported the use of
polymer coated Bi2S3 nanoparticles as an injectable X-ray
computed tomography imaging agent, which foresees the new
directions for their applications [12]. It has also been reported that
Bi2S3 has a lamellar structure and can exhibit excellent hydrogen
storage properties [13]. As far as the synthesis of Bi2S3 nanorods
and nanowires are concerned, both solvothermal and hydrothermal processes using different sulfur precursors such as sodium
thiosulfate, sodium sulde, thiourea and thioacetamide in various
solvents have been reported [14e21]. Reihaneh Malakooti et al.
used a hot injection method to prepare shape controlled Bi2S3
nanocrystals [22]. The above group and A.P. Alivisatos group have
designed various synthetic procedures to prepare nanostructured
Bi2S3 starting from bismuth salt and sulfur precursors [23e27]. The
use of alkylthiolate precursors is also well established in preparing
nanowires and nanorods of Bi2S3 by solventless synthetic methods
[28]. Chelating reagents such as triethanolamine (TEA) and
ethylenediaminetetra-acetate (EDTA) have also been employed in
the synthetic routes [29,30]. As an alternative, biomolecule assisted synthesis methods have been a new and promising strategy
towards the preparation of various nanomaterials where biomolecules have mostly been exploited as a structure directing
agent. Protein cages were adapted as templates to synthesize
nanocrystals [31,32]. Viruses, peptides and lemon grass were also
used as templates to prepare nanomaterials of primarily transition
metals [33e37]. Nanoparticles and nanowires of semiconductor
metal sulphides have been successfully prepared and assembled
using peptides, viruses, bacteria and fungus in presence of either
Na2S or H2S [38e43]. Flower like patterns of Bi2S3 with well
aligned nanorods were synthesized using a facile solution phase
biomolecule assisted approach in the presence of L-cysteine, which
acts not only as a sulfur source but also as a structure directing
agent [13]. Similarly Glutathione (GSH) was used as a structure
directing medium and as a source of sulfur for the preparation of
snowake like structures of Bi2S3 nanorods [44]. Nanowires of
Bi2S3 have been produced using the protein lysozyme as the
structure directing agent and thiourea as the sulfur source [45].
Careful analysis of the bio-macromolecule based synthesis of Bi2S3
reveals that, in most of the cases, these molecules act as structure
directing agents and in few cases as a source of sulfur. The above
reported methods usually require a very high temperature and
involve the use of expensive and toxic organic reagents consisting
of a series of complicated procedures. The possibility of room
temperature chemical routes for the synthesis of Bi2S3 was also
reported and yet these methods use toxic solvents and carcinogenic reducing agents [46e49].
Hence, in this work an alternative effort to environmentally
benign room temperature route involving biomolecules is reported.
This would be a signicant step towards achieving a greener
methodology. In this paper we describe a green synthetic route for
the preparation of nanostructured Bi2S3 through a bio-assisted
approach. Family of Clostridium is well known to produce hydrogen

sulde [50]. In the present work, one of the Clostridium species


namely, Clostridium acetobutylicum is employed as a source of
hydrogen sulde which brings out the conversion of bismuth sulfate to bismuth sulphide and the process is carried out at room
temperature. Transmission electron microscope (TEM), powder Xray diffraction (XRD), X-ray photoelectron spectroscopy (XPS) and
cyclic voltammetry (CV) are used for the characterization of
structure and composition of Bi2S3. SDS-PAGE gel electrophoresis is
employed to analyze the culture media and the stabilizing groups
present on the surface of Bi2S3.
2. Material and methods
2.1. Cultivation of C acetobutylicum
C. acetobutylicum (NCIM 2337) was obtained from CSIReNational Chemical Laboratory (CSIR-NCL), Pune, India. It was allowed
to undergo anaerobic fermentation in the presence of a cooked
meat medium for about 48 h. Then the culture broth was centrifuged at 8000 rpm for 10 min and the supernatant was collected
separately. Extra cellular protein constituents of this particular
culture medium were later conrmed by SDS-PAGE gel
electrophoresis.
2.2. Synthesis of Bi2S3 from bismuth sulfate using C. acetobutylicum
About 100 mg of bismuth sulfate was added to 45 ml of the
supernatant collected above. Upon addition of Bi2(SO4)3 (white in
color), the color of the supernatant changed from yellow to dark
brown within a few minutes. This color change indicates the
reduction of bismuth sulfate to bismuth sulphide. It is worth
mentioning here that when the bismuth sulfate was dispersed in
either water, phosphate buffer or cooked meat broth without
inoculum the color remains unchanged. And when it was added to
the cooked meat broth with inoculum, the color change from white
to brown was observed. These observations pave a way to understand that C. acetobutylicum releases hydrogen sulde during the
time of its growth and involved in the reduction process. The
resulting brown solution was then incubated under anaerobic
conditions at room temperature for 5e7 days. Then the solution
was centrifuged at 12,000 rpm and 18,000 rpm respectively. A dark
brown pellet was obtained on centrifugation and was found to be
redispersible in water.
2.3. Characterization of bismuth sulphide
The resultant pellet was initially analyzed using the qualitative
chemical tests in order to conrm the presence of sulphide. The
structural morphology and orientation of bismuth sulphide were
characterized using transmission electron microscope (TEM),
powder X-ray diffraction studies (XRD) and X-ray photoelectron
spectroscopy (XPS). Size of the resultant particles was determined
using particle size analyzer. Electrochemical technique namely
cyclic voltammetry (CV) was used for further characterization. A
three-electrode electrochemical cell was used for the study. Pt
wire was used as a counter electrode; mercury-mercurous sulfate
(MMS) was used as a reference electrode and glassy carbon (GC)
was used as a working electrode respectively. An aqueous solution
of 0.5 M sulfuric acid was used as an electrolyte. CV was performed within a potential ranging from 1.5 V to 1.5 V vs. MMS at
a xed scan rate of 50 mV s1. For comparison we have also
studied the electrochemical properties of the starting material,
bismuth sulfate and bare GC in acidic medium. All the electrochemical experiments were carried out using AUTOLAB
equipment.

Please cite this article in press as: S.K. Kamaraj, et al., Bio-assisted synthesis and characterization of nanostructured bismuth (III) sulphide using
Clostridium acetobutylicum, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.11.042

S.K. Kamaraj et al. / Materials Chemistry and Physics xxx (2013) 1e6

Fig. 1. Transmission electron micrographs of bismuth sulphide obtained at 18,000 (A) and 12,000 (B) rpm values. Inset shows the selected area electron diffraction pattern corresponding to 12,000 rpm (B).

2.4. SDS-PAGE experiments

3. Results and discussion

Further, Sodium Dodecyl sulfate-Polyacrylamide Gel Electrophoresis (SDS-PAGE) experiments were performed to identify and
characterize the culture media and the stabilizing groups present
on the surface of the synthesized material. It is a technique, which
involves the separation of proteins based on their molecular
weight and size. SDS is the anionic surfactant that provides a
negative charge to the proteins and polyacrylamide consists of the
chains of polymerized acrylamide units cross linked by a bifunctional agent namely N,N0 -methylenebisacrylamide. We have
used a procedure following the discontinuous gel electrophoresis
i.e. two types of gels were used; of stacking gel (5%) and 3/4
(12%) of separating gel. The SDS plates were cleaned with water
and assembled in the gel stand. Regarding the separating gel 12%
composition of the gel was prepared. In this, the acrylamide gels
are formed by polymerizing 30% acrylamide with a cross-linker
(bisacrylamide) in the presence of a catalyst (TEMED) and an
initiator (ammonium persulfate) with a suitable gel buffer. After
pouring the separating gel in the plate, water was added to prevent the inhibition of polymerization due to the presence of oxygen. Then, the stacking gel was formed above the separating gel.
Before polymerization of stacking gel the comb was placed to get
wells to load the sample. The samples were prepared by the
following procedure; sample was sonicated for 20 min to rupture
the cells. After rupturing it was centrifuged at 8000 rpm for
10 min to settle the cell debris and the supernatant was separated.
About 100 ml of the supernatant was mixed with 25 ml of sample
buffer (containing bromophenolblue e a pinch; 2-mercaptoethanol e 0.5 ml; SDS e 150 mg; glycerol e 1 ml; stacking gel buffer e
1.25 ml; distilled water - 7.25 ml) and kept in water bath at 80  C
for about 4 min. This temperature will denature the proteins
which gives a perfect separation. Then the sample was cooled
suddenly and it was loaded in the lane. We have applied a voltage
of 8 V cm1 for stacking gel and 15 V cm1 for separating gel
during the experiments. After the completion, the gel was stained
for about 2 h using a stain solution (consisting of coomassie
brilliant blue e 200 mg; ethanol e 50 ml, acetic acid e 7 ml;
distilled water e 43 ml. Then, it was de-stained by a de-staining
solution comprising of ethanol e 30 ml; acetic acid e 7 ml;
distilled water e 63 ml) in order to obtain the clear bands for the
presence of proteins.

3.1. Characterization of bismuth sulphide using TEM


Size and shape of bismuth sulphide synthesized using C. acetobutylicum at room temperature was characterized using TEM.
Fig. 1A and B shows the TEM images recorded for bismuth sulphide
obtained using centrifugation carried out at 18,000 rpm and
12,000 rpm respectively. The sizes of these particles were determined to be in the range of 6e10 nm for particles centrifuged at
18,000 rpm and 440e500 nm for particles centrifuged at
12,000 rpm respectively. It can be seen from the images that the
bio-assisted method results in the formation of hexagonally shaped
platelets. The corresponding selected area electron diffraction
(SAED) pattern was shown as the inset. These hexagonally shaped
nanostructures resemble with the structure of layered manganese
oxide prepared through bacterial origin and the naturally occurring
birnessites [51,52]. Particle size analysis which works on the principle of dynamic light scattering was also performed to determine
the size of the resultant particles. It revealed that the centrifugation
under 18,000 rpm yielded particles of size ranging from 50 to
80 nm and centrifugation at 12,000 rpm yielded particles in the size
range of 400e500 nm. The difference in the sizes between these
two measurements was attributed to the phenomena and the
mechanism involved in the respective experiments.
3.2. X-ray diffraction (XRD) studies
Fig. 2 shows the powder X-ray diffraction (XRD) pattern of as
prepared Bi2S3 nanoparticles using a bio-assisted method. The
observed diffraction peaks suggest that the material is of pure
bismuth sulphide. The corresponding planes were indexed and
have been shown in the diagram. The occurrence of various peaks
and planes indicate that bismuth sulphide prepared using this
method possess an orthorhombic crystalline structure [11,13]. The
appearance of broad diffraction peaks is a typical characteristic of
particles being nanoscale in dimensions. Some of the peaks marked
with asterisk (*) are very weak in intensity. This is attributed to the
fact that the bio-conjugated nanoparticles especially with that of
protein materials can affect the intensity of certain peaks [53]. The
observed peaks are in good match with the JCPDS le (65-2431; 751306) for the orthorhombic structure of bismuth sulphide.

Please cite this article in press as: S.K. Kamaraj, et al., Bio-assisted synthesis and characterization of nanostructured bismuth (III) sulphide using
Clostridium acetobutylicum, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.11.042

S.K. Kamaraj et al. / Materials Chemistry and Physics xxx (2013) 1e6

Fig. 2. Powder X-ray diffraction pattern of Bi2S3. The peaks marked with asterisk (*)
are weak in intensity.

Fig. 4. Cyclic voltammograms of bare GC (a), bismuth sulfate (b) and bismuth sulphide
nanoparticles (c) in 0.5 M H2SO4 at a sweep rate of 50 mV s1.

3.3. X-ray photoelectron spectroscopy (XPS) analysis

bismuth sulphide [21,54]. These results clearly show that the


resultant bismuth sulphide nanoparticles were pure and do not
contain any impurities.

Having characterized the size and composition, the presence of


elements with respective oxidation states was determined using a
surface analysis technique namely, X-ray photoelectron spectroscopy (XPS). Fig. 3 shows the respective XPS graphs of Bi 4d (A), Bi 4f
(B) and S 2s (c) regions of Bi2S3. The formation of peaks at 158.8 eV,
164.2 eV corresponds to Bi 4f7/2 and Bi 4f5/2 and at 442 eV corresponds to Bi 4d5/2 respectively. Besides we have also observed a
peak at 226 eV implying the presence of sulfur (S 2s) corresponds to
sulphide region. The positions of these peaks agree very well with
the reported values of Bi3 and S2 indicating the formation of

3.4. Electrochemical characterization


Cyclic voltammetry (CV) was used to study the electrochemical
properties of the synthesized bismuth sulphide nanoparticles. Fig. 4
shows the CVs of a bare glassy carbon (GC) (a), bismuth sulfate (b),
and bismuth sulphide nanoparticles (c) respectively in 0.5 M sulfuric acid at a potential scan rate of 50 mV s1. It can be noticed
from Fig. 4a that the bare GC electrode does not show any peak

Fig. 3. XPS spectra showing Bi 4d (A), Bi 4f (B) and S 2s (C) regions of bismuth sulphide.

Please cite this article in press as: S.K. Kamaraj, et al., Bio-assisted synthesis and characterization of nanostructured bismuth (III) sulphide using
Clostridium acetobutylicum, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.11.042

S.K. Kamaraj et al. / Materials Chemistry and Physics xxx (2013) 1e6

formation and the observed current is just the background current.


On the contrary, both bismuth sulfate (4b) and bismuth sulphide
(4c) show peak formation suggesting the redox behavior of bismuth. Fig. 4b shows the formation of one anodic peak at 0.42 V
and two cathodic peaks at 0.49 V and 0.66 V respectively suggesting the formation and reduction of BiO and (BiO)2SO4 in acidic
medium. The stable forms of bismuth in sulfuric acid medium are
BiO and (BiO)2SO4 [55,56]. The formation of cathodic peak at a
very negative potential (0.66 V) is due to the reduction of
(BiO)2SO4. Interestingly, the bismuth sulphide nanoparticles show
peak formation at 0.44 V and 0.49 V indicating the enhanced
reversibility for formation and reduction of BiO. The disappearance of reduction peak at a very negative potential suggest that
there is no sulfate present in the system and it has been converted
into sulphide by the presence of microorganism during the process
of synthesis.
3.5. SDS-PAGE analysis
Bi2S3 nanoparticles were synthesized from bismuth sulfate using Clostridium species at room temperature. C. acetobutylicum was

used not only as a source of hydrogen sulphide, which acts as a


reducing agent but also as a stabilizing agent. TEM and particle size
analysis were used to determine the size of Bi2S3 nanoparticles
which exhibits the formation of hexagonal nanostructures. XRD,
XPS and CV studies were performed for the structural characterization that reveal the formation of crystalline Bi2S3 possessing an
orthorhombic structure. Further, the stabilization of Bi2S3 nanostructures by extracellular proteins present in C. acetobutylicum was
conrmed through SDS-PAGE gel electrophoresis experiments. The
results were displayed in Fig. 5. These experiments were performed
to analyze the presence of proteins based on their molecular weight
differences. In this gure lane 1 displays the biomarkers used as a
reference for this study. Lane 2 and lane 3 represent the supernatant of C. acetobutylicum and the synthesized bismuth sulphide
nanoparticles respectively. The formation of bands at a particular
region corresponds to the molecular weight of the proteins present
in the sample. Appearance of similar bands from 150 KDa to 38 KDa
in lane 2 and 3 suggests that the synthesized Bi2S3 nanoparticles
were stabilized by these proteins present in the inoculum.
4. Conclusions
Our results clearly demonstrate the feasibility of room temperature synthesis of bismuth sulphide nanostructures employing a
biological route, which is environmentally benign compared to the
other chemical methods reported. C. acetobutylicum acts as a source
of hydrogen sulphide. The extracellular proteins extracted from the
culture broth stabilize the bismuth sulphide nanostructures, as
conrmed by SDS-PAGE studies. Synthesized Bi2S3 nanoparticles
were characterized using TEM, XRD, XPS and CV. Currently we are
working towards using these nanomaterials for electrochemical
and photoelectrochemical applications.
Acknowledgments
Authors thank Council of Scientic and Industrial Research
(CSIR), New Delhi, India for funding this research work under the
network projects scheme (Project no. NWP0035).
References

Fig. 5. SDS-PAGE gel electrophoresis experiment. Lane 1 corresponds to the reference


biomarker. Lane 2 denotes the extracellular extract and lane 3 represents the bismuth
sulphide nano-structured material stabilized by the extracellular proteins.

[1] C.H. Weng, C.C. Huang, C.S. Yeh, H.Y. Lei, G.B. Lee, J. Micromech. Microeng. 18
(2008) 035019.
[2] T. Sato, Fujilms Corporation Technical Reports, vol. 53, 2008, p. 22.
[3] J.A. Dahl, B.L.S. Maddux, J.E. Hutchison, Chem. Rev. 107 (2007) 2228.
[4] J. Black, E.M. Conwell, L. Seigle, C.W. Spencer, J. Phys. Chem. Solids 2 (1957)
240.
[5] M.A. Malik, M. Afzaal, P. OBrien, Chem. Rev. 110 (2010) 4417.
[6] Z. Liu, S. Peng, Q. Xie, Z. Hu, Y. Yang, S. Zhang, Y. Qian, Adv. Mater. 15 (2003)
936.
[7] M.H. Stella, Science 295 (2002) 767.
[8] R.S. Mane, B.R. Sankarpal, C.D. Lokhande, Mater. Chem. Phys. 60 (1999) 158.
[9] R. Suarez, P.K. Nair, P.V. Kamat, Langmuir 14 (1998) 3236.
[10] P.K. Nair, M.T.S. Nair, V.M. Garcia, O.L. Arenas, A.C.Y. Pena, I.T. Ayala,
O. Gomezdaza, A. Sanchez, J. Campos, H. Hu, R. Suarez, M.E. Rincon, Sol. Energy Mater. Sol. Cells 52 (1998) 313.
[11] S. Vitalie, K.H. Whitmire, I. Rusakova, Chem. Mater. 21 (2009) 5456.
[12] O. Rabin, J. Manuel Perez, J. Grimm, G. Wojtkiewicz, R. Weissleder, Nat. Mater.
5 (2006) 118.
[13] B. Zhang, X. Ye, W. Hou, Y. Zhao, Y. Xie, J. Phys. Chem. B 110 (2006) 8978.
[14] X. Cao, L. Li, Y. Xie, J. Colloid Interface Sci. 273 (2004) 175.
[15] S. Yu, Y. Qian, L. Shu, Y. Xie, L. Yang, C. Wang, Mater. Lett. 35 (1998) 116.
[16] D. Wang, M. Shao, D. Yu, G. Li, Y. Qian, J. Cryst. Growth 243 (2002) 331.
[17] Y. Chen, H. Kou, J. Jiang, Y. Su, Mater. Chem. Phys. 82 (2003) 1.
[18] S.H. Yu, L. Shu, J. Yang, Z.H. Han, Y.T. Qian, Y.H. Zhang, J. Mater. Res. 14 (1999)
4157.
[19] Q. Li, M. Shao, J. Wu, G. Yu, Y. Qian, Inorg. Chem. Commun. 5 (2002) 933.
[20] W. Zhang, Z. Yang, X. Huang, S. Zhang, W. Yu, Y. Qian, Y. Jia, G. Zhou, L. Chen,
Solid State Commun. 119 (2001) 143.
[21] R. Chen, M.H. So, C.M. Che, H. Sun, J. Mater. Chem. 15 (2005) 4540.
[22] R. Malakooti, L. Cademartiri, Y. Akakir, S. Petrov, A. Migliori, G.A. Ozin, Adv.
Mater. 18 (2006) 2189.

Please cite this article in press as: S.K. Kamaraj, et al., Bio-assisted synthesis and characterization of nanostructured bismuth (III) sulphide using
Clostridium acetobutylicum, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.11.042

S.K. Kamaraj et al. / Materials Chemistry and Physics xxx (2013) 1e6

[23] L. Cademartiri, R. Malakooti, P.G. OBrien, A. Migliori, S. Petrov, N.P. Kherani,


G.A. Ozin, Angew. Chem. Int. Ed. 47 (2008) 3814.
[24] L. Cademartiri, F. Scotognella, P.G. OBrien, B.V. Lotsch, J. Thomson, S. Petrov,
N.P. Kherani, G.A. Ozin, Nano Lett. 9 (2009) 1482.
[25] J.W. Thomson, L. Cademartiri, M. MacDonald, S. Petrov, G. Calestani, P. Zhang,
G.A. Ozin, J. Am. Chem. Soc. 132 (2010) 9058.
[26] L. Cademartiri, G. Guerin, K.J.M. Bishop, M.A. Winnik, G.A. Ozin, J. Am. Chem.
Soc. 134 (2012) 9327.
[27] J. Tang, A.P. Alivisatos, Nano Lett. 6 (2006) 2701.
[28] M.B. Sigman, B.A. Korgel, Chem. Mater. 17 (2005) 1655.
[29] R.S. Mane, B.R. Sankarpal, C.D. Lokhande, Mater. Res. Bull. 35 (2000) 587.
[30] S.H. Yu, J. Yang, Y.S. Wu, Z.H. Han, Y. Xie, Y.T. Qian, Mater. Res. Bull. 33 (1998)
1661.
[31] T. Douglas, E. Strable, D. Willits, A. Aitouchen, M. Libera, M. Young, Adv. Mater.
14 (2002) 415.
[32] T. Douglas, D.P.E. Dichson, S. Betteridge, J. Charnoch, C.D. Garner, S. Mann,
Science 269 (1995) 54.
[33] W. Shenton, T. Douglas, M. Yong, G. Stubbs, S. Mann, Adv. Mater. 11 (1999) 253.
[34] S.S. Shankar, A. Rai, B. Ankamwar, A. Singh, A. Ahmad, M. Sastry, Nature Mater.
3 (2004) 482.
[35] M. Knez, A.M. Bittner, F. Boes, C. Wege, H. Jeske, E. Mai, K. Kern, Nano Lett. 3
(2003) 1079.
[36] B.D. Reiss, C.B. Mao, D.J. Solis, K.S. Ryan, T. Thomson, A.M. Belcher, Nano Lett. 4
(2004) 1127.
[37] C.B. Mao, D.J. Solis, B.D. Reiss, S.T. Kottmann, R.Y. Sweeney, A. Hayhurst,
G. Georgiou, B. Iverson, A.M. Belcher, Science 303 (2004) 213.

[38] M. Labrenz, G.K. Druschel, T. Thomsen-Ebert, B. Gilbert, S.A. Welch,


K.M. Kemner, G.A. Logan, R.E. Surmmons, G.D. Stasio, P.L. Bond, B. Lai,
S.D. Kelly, J.F. Baneld, Science 290 (2000) 1744.
[39] C.T. Dameron, D.R. Winge, Inorg. Chem. 29 (1990) 1343.
[40] W. Shenton, D. Pum, U.B. Sleytr, S. Mann, Nature 389 (1997) 585.
[41] C.B. Mao, C.E. Flymn, A. Hayhurst, R. Sweeney, J. Qi, G. Georgiou, B. Iverson,
A.M. Belcher, Proc. Natl. Acad. Sci. U. S. A. 100 (2003) 6946.
[42] R.Y. Sweeney, C.B. Mao, X.X. Gao, J.L. Burt, A.M. Belcher, G. Georgiou,
B. Iverson, Chem. Biol. 11 (2004) 1553.
[43] A. Ahmad, P. Mukherjee, D. Mandal, S. Senapati, M.I. Khan, R. Kumar,
M. Sastry, J. Am. Chem. Soc. 124 (2002) 12108.
[44] Q. Lu, F. Gao, S. Komarneni, J. Am. Chem. Soc. 126 (2004) 54.
[45] F. Gao, Q. Lu, S. Komarneni, Chem. Commun. (2005) 531.
[46] L.H. Dong, Y. Chu, W. Zhang, Mater. Lett. 62 (2008) 4269.
[47] X.D. Zhou, H.Q. Shi, B. Zhang, X. Fu, K. Hao, Mater. Lett. 62 (2008) 3201.
[48] W.B. Zhao, J.H. Zhu, Y. Zhao, H.Y. Chen, Mater. Sci. Eng. B 110 (2004) 307.
[49] S.X. Zhou, J.M. Li, Y.X. Ke, S.M. Lu, Mater. Lett. 57 (2003) 2602.
[50] L.M. Mosley, D.S. Sharp, SOPAC Technical Report, 2005, p. 373.
[51] T.G. Spiro, J.R. Bargar, G. Sposito, B.M. Tebo, Acc. Chem. Res. 43 (2010) 2.
[52] S.H. Kim, S.J. Kim, S.M. Oh, Chem. Mater. 11 (1999) 557.
[53] M.J. Meziani, P. Pathak, B.A. Harruff, R. Hurezeanu, Y.P. Sun, Langmuir 21
(2005) 2008.
[54] S.C. Liufu, L.D. Chen, Q. Yao, C.F. Wang, J. Phys. Chem. B 110 (2006) 24054.
[55] W.S. Li, X.M. Long, J.H. Yan, J.M. Nan, H.Y. Chen, Y.M. Wu, J. Power Sources 158
(2006) 1096.
[56] M. Yang, Z. Hu, J. Electroanal. Chem. 583 (2005) 46.

Please cite this article in press as: S.K. Kamaraj, et al., Bio-assisted synthesis and characterization of nanostructured bismuth (III) sulphide using
Clostridium acetobutylicum, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.11.042

You might also like