Anda di halaman 1dari 113

ZnS nanostructures: From synthesis to applications

Xiaosheng Fang
a,b,
, Tianyou Zhai
b,
, Ujjal K. Gautam
b,
, Liang Li
b,
,
Limin Wu
a,
, Yoshio Bando
b
, Dmitri Golberg
b
a
Department of Materials Science, Fudan University, Handan Road 220, Shanghai 200433, China
b
International Center for Young Scientists (ICYS), International Center for Materials Nanoarchitectonics (MANA), National Institute for
Materials Science (NIMS), Namiki 1-1, Tsukuba, Ibaraki 305-0044, Japan
a r t i c l e i n f o
Article history:
Received 19 March 2010
Received in revised form21 September 2010
Accepted 11 October 2010
a b s t r a c t
Zinc sulde (ZnS) is one of the rst semiconductors discovered. It
has traditionally shown remarkable versatility and promise for
novel fundamental properties and diverse applications. The nano-
scale morphologies of ZnS have been proven to be one of the rich-
est among all inorganic semiconductors. In this article, we provide
a comprehensive review of the state-of-the-art research activities
related to ZnS nanostructures. We begin with a historical back-
ground of ZnS, description of its structure, chemical and electronic
properties, and its unique advantages in specic potential applica-
tions. This is followed by in-detail discussions on the recent pro-
gress in the synthesis, analysis of novel properties and potential
applications, with the focus on the critical experiments determin-
ing the electrical, chemical and physical parameters of the nano-
structures, and the interplay between synthetic conditions and
nanoscale morphologies. Finally, we highlight the recent achieve-
ments regarding the improvement of ZnS novel properties and
nding prospective applications, such as eld emitters, eld effect
transistors (FETs), p-type conductors, catalyzators, UV-light sen-
sors, chemical sensors (including gas sensors), biosensors, and
nanogenerators. Overall this review presents a systematic investi-
gation of the synthesis-property-application triangle for the
diverse ZnS nanostructures.
2010 Published by Elsevier Ltd.
0079-6425/$ - see front matter 2010 Published by Elsevier Ltd.
doi:10.1016/j.pmatsci.2010.10.001

Corresponding authors. Addressess: Department of Materials Science, Fudan University, Handan Road 220, Shanghai
200433, China (X.S. Fang, L.M. Wu), National Institute for Materials Science (NIMS), Namiki 1-1, Tsukuba, Ibaraki 305-0044,
Japan (T.Y. Zhai, U.K. Gautam, L. Li). Tel./fax: +86 21 65643795.
E-mail addresses: xshfang@fudan.edu.cn, xshfang@yahoo.cn (X.S. Fang), zhai.tianyou@gmail.com (T.Y. Zhai), Gautam.Ujjal@
nims.go.jp (U.K. Gautam), Li.Liang@nims.go.jp (L. Li), lmw@fudan.edu.cn (L.M. Wu).
Progress in Materials Science 56 (2011) 175287
Contents lists available at ScienceDirect
Progress in Materials Science
j our nal homepage: www. el sevi er . com/ l ocat e/ pmat sci
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
2. Fundamental properties of ZnS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
3. Synthesis of ZnS nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
3.1. 0D nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
3.1.1. 0D nanocrystals (quantum dots) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
3.1.2. 0D core/shell nanocrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
3.1.3. 0D hollow nanocrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
3.2. 1D nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
3.2.1. Nanowires (NWs), nanorods (NRs), and nanotubes (NTs) . . . . . . . . . . . . . . . . . . . . . . . 188
3.2.2. Nanobelts (NBs), nanoribbons (NRs) and nanosheets (NSs). . . . . . . . . . . . . . . . . . . . . . 191
3.2.3. Aligned nanowires and nanobelts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
3.2.4. Complex nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
3.3. 2D nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
4. Luminescence properties of ZnS nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
4.1. Photoluminescence (PL) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
4.1.1. Visible emission of ZnS nanostructures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
4.1.2. UV emission of ZnS nanostructures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
4.2. Cathodoluminescence (CL) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
4.2.1. Multi-angular branched ZnS nanostructures with needle-shaped tips . . . . . . . . . . . . . 216
4.2.2. Single-crystalline ZnS nanobelts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
4.3. Electroluminescence (EL) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
4.4. Electrochemiluminescence (ECL) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
4.5. Thermoluminescence (TL). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
4.6. Luminescence mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
4.6.1. Origin of the green photoluminescence from ZnS nanobelts [148]. . . . . . . . . . . . . . . . 225
4.6.2. Temperature-dependent PL from elemental sulfur species on ZnS nanobelts [345] . . 226
4.7. Optical property tuning by doping. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
4.7.1. Optical properties of Mn-doped ZnS nanobelts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
4.7.2. Optical properties of Cu-doped ZnS nanorods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
4.7.3. Red light PL emission from Mn/Cd-co-doped-ZnS nanostructures . . . . . . . . . . . . . . . . 230
5. Potential applications. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
5.1. Field-emission applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
5.1.1. Field emission phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
5.1.2. Ultrafine ZnS nanobelts as field emitters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
5.1.3. Enhanced field-emission properties by crystal orientation-ordered ZnS nanobelt
quasi-arrays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
5.1.4. Enhanced field-emission properties by multi-angular branched ZnS nanostructures
with needle-shaped tips. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
5.1.5. Multiply enhanced field-emission properties of branched ZnS nanotube-In nanowire
coreshell heterostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
5.1.6. ZnS nanostructures with other novel morphologies . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
5.2. Field effect transistors (FETs) and carrier characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
5.3. p-type conductivity in ZnS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
5.4. Catalytic activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
5.5. UV-light sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
5.5.1. Unique advantages of ZnS nanostructrues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
5.5.2. Individual ZnS nanostructures-based UV-light sensors . . . . . . . . . . . . . . . . . . . . . . . . . 253
5.5.3. Multiple ZnS nanostructures-based UV-light sensors . . . . . . . . . . . . . . . . . . . . . . . . . . 255
5.5.4. Microscale ZnS nanobelts-based UV-light sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
5.6. Gas sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
5.7. Chemical sensors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
5.8. Biosensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
5.9. Nanogenerators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
6. Conclusions and outlooks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
176 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
1. Introduction
Nanostructured materials are not only in the forefront of the hottest fundamental materials re-
search nowadays, but they are also gradually intruded into our daily life [13]. Theres plenty of room
at the bottom, the principles of physics, as far as I can see, do not speak against the possibility of
maneuvering things atom by atom, put the atoms down where the chemist says, and so you make
the substance. . ., this famous statement of legendary Richard Feynman made in 1959 with immense
foresight has been realized in less than half a century by consistent efforts and signicant contribu-
tions from the scientic community across the globe [4].
Nanostructured materials are a new class of materials, having dimensions in the 1100 nm range,
which provide one of the greatest potentials for improving performance and extended capabilities of
products in a number of industrial sectors [5]. Nanostructures can be divided into zero-dimensional
(0D when they are uniform), one-dimensional (1D when they are elongated), and two-dimensional
(2D when they are planar) based on their shapes. The recent emphasis in the nanomaterials research
is put on 1D nanostructures at the expense of 0D and 2D ones, perhaps due to the intriguing possibility
of using them in a majority of short-term future applications. There is a large number of new oppor-
tunities that could be realized by down-sizing currently existing structures into the nanometer scale
(<100 nm), or by making new types of nanostructures. The most successful examples are seen in the
microelectronics, where smaller has always meant a greater performance ever since the invention of
transistors: e.g. higher density of integration, faster response, lower cost, and less power consumption
[1].
Zinc sulde (ZnS) is one of the rst semiconductors discovered [6] and it has traditionally shown
remarkable fundamental properties versatility and a promise for novel diverse applications, including
light-emitting diodes (LEDs), electroluminescence, at panel displays, infrared windows, sensors, la-
sers, and biodevices, etc. Its atomic structure and chemical properties are comparable to more popular
and widely known ZnO. However, certain properties pertaining to ZnS are unique and advantageous
compared to ZnO. To name a few, ZnS has a larger bandgap of -3.72 eV and -3.77 eV (for cubic zinc
blende (ZB) and hexagonal wurtzite (WZ) ZnS, respectively) than ZnO (-3.4 eV) and therefore it is
more suitable for visible-blind ultraviolet (UV)-light based devices such as sensors/photodetectors.
On the other hand, ZnS is traditionally the most suitable candidate for electroluminescence devices.
However, the nanostructures of ZnS have not been investigated in much detail relative to ZnO
nanostructures.
In this article, we will provide a comprehensive review of the state-of-the-art research activities
related to ZnS nanostructures, including their synthesis, novel properties studies and potential appli-
cations. We begin with a historical background of ZnS, description of its structure, chemical and elec-
tronic properties, and the possible reasons for the investigation of this nanomaterial, followed by a
survey of ZnS nanostructures with various morphologies and corresponding synthesis methods and
experimental parameters. Using various facile techniques nanoparticles, nanorods, nanowires, nano-
belts/nanoribbons, nanosheets, nanotubes, core/shell nanostructures, hierarchical nanostructures,
complex nanostuctrues and heterostructures of ZnS have been synthesized under specic growth con-
ditions so far. Subsequently, we will discuss the critical experiments determining the electrical, chem-
ical and physical properties of the nanostructures, in regard of synthetic conditions. This will be
followed by the main objective of the review which is the prospects of ZnS diverse structures in var-
ious functional devices. The recent progress on the improvement of their properties and nding novel
potential applications, such as the latest achievements in using various ZnS nanostructures as eld
emitters, eld effect transistors (FETs), p-type conductors, catalyzators, UV-light and chemical sensors
(including gas sensors), biosensors, and nanogenerators will be highlighted.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 177
2. Fundamental properties of ZnS
ZnS has two commonly available allotropes: one with a ZB structure and another one with a WZ
structure. The cubic form is the stable low-temperature phase, while the latter is the high-tempera-
ture polymorph which forms at around 1296 K [7]. For the purpose of comparison, Fig. 1 shows three
different views of these structures. The differences can be described either in terms of the relative
handedness of the fourth interatomic bond or by their dihedral conformations. Alternatively, ZB con-
sists of tetrahedrally coordinated zinc and sulfur atoms stacked in the ABCABC pattern, while in WZ,
the same building blocks are stacked in the ABABAB pattern. The lattice parameters of ZB are
a = b = c = 5.41 , Z = 4 (space group F4-3 m) and that of WZ are a = b = 3.82 , c = 6.26 , Z = 2 (space
group = P6
3
mc).
Such minute difference in atomic arrangements leads to large difference in properties in these
materials [7], e.g. electronic structures and bandgaps. The WZ phase has a higher bandgap of
3.77 eV [8] while the ZB structure of 3.72 eV [9]. The band structure of a solid describes ranges of en-
ergy that an electron is forbidden or allowed to have and it determines important electronic and
Fig. 1. Models showing the difference between wurtzite and zinc blende crystal structures. (a and b) Show handedness of the
fourth interatomic bond along the right (R) for wurtzite and along the left (L) for zinc blende. (c and d) The respective eclipsed
and staggered dihedral conformations. (e and f) Show atomic arrangement along the close packing axis. Reproduced from Ref.
[7]. Copyright 1992, The American Physical Society.
178 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
optical properties of the material. The optical spectra are related to band structure, its dispersion and
probabilities of inter-band optical transitions. Experimentally, it has been shown that the optical prop-
erties of the ZnS ZB and WZ phases are distinct. In order to understand such differences, electronic
structures of these phases have been widely investigated [1014]. Fig. 2 depicts typical band disper-
sions for the ZnS ZB and the WZ phases obtained using density functional theory (DFT) calculations. It
can be seen that the conduction-band minima are much more dispersive than the valence band max-
ima for both phases. The authors further state that mobility of electrons in these materials is therefore
higher than that of holes [10]. Also, p-electrons forming the topmost valence band states tightly bind
to sulfur and make the valence band holes less mobile. Hence, the contribution of the holes to the con-
ductivity is expected to be smaller. The valence band comprises of three regions: a lower region con-
sists of the s bands of Zn and S, a higher-lying region contains well localized Zn 3d bands, and a top
broader band originating from the Sp states hybridized with Zn 3d states.
Notably, the phase transition temperature for the two allotropes, the bandgap and electronic struc-
ture considerably change when the size of the ZnS particles is of the order of nanometers. In an elegant
and simple experiment, Qadri and co-workers used 2.7 nm ZnS ZB nanocrystals and heated them at
various temperatures [15]. X-ray diffraction (XRD) measurement of these sample showed that ZB
nanocrystals start converting to the WZ phase at a temperature as low as 400 C (Fig. 3). During
Fig. 2. Band dispersion for WZ and ZB ZnS calculated according to LDA (solid lines) and LDA + U (dotted lines). The Fermi level is
set to zero energy Reproduced from Ref. [10]. Copyright 2007, The American Physical Society.
Fig. 3. XRD patterns of the ZnS nanocrystals annealed at various temperatures taken with Cu Ka radiation. Peaks labeled with
ZB and WZ correspond to the zinc blende and wurtzite structures, respectively. Reproduced from Ref. [15]. Copyright 1999,
The American Physical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 179
annealing, the particle size, cell volume, lattice parameters also changed considerably, as depicted in
Table 1.
Akiyama and co-workers using empirical calculations have demonstrated that when the sizes of
the ZnS nanostructures reduce to just a few nanometers, the high temperature WZ structure becomes
stabilized [16]. They calculated cohesive energies for nanowires with ZB and WZ (hexagonal, H) struc-
tures and demonstrated that the stability of a crystal structure depends on the nanowire diameter.
Fig. 4 shows the plot of energy differences between 6H and ZB structures DE
6HZB
, 4H and ZB struc-
tures DE
4HZB
, and WZ and ZB structures DE
WZZB
of ZnS NWs as functions of nanowire diameters.
As seen in the plot, the energy differences converge into those of the bulk phase as diameter in-
creases, indicating the appearance of bulk features at a large diameter. The W structure was found
favorable for diameters less than 4 nm. On the other hand, the ZB structure, which is the most stable
structure in the bulk, is favorable for diameters above 24 nm. The authors explained this behavior on
the basis of two- and three-coordinated surface atoms on nanowire facets. Different from the fourfold
coordination in the bulk solid, the Zn and S atoms at the side surfaces of WZ-ZnS nanowires are all
threefold coordinated with one unsaturated bond. In the case of the ZB-ZnS nanowires, in addition
to the threefold coordinated atom, there are also certain twofold coordinated atoms located at the
edges of the nanowires facets. This makes the surface energy of ZB-ZnS nanowires larger than that
Table 1
Calculated values of the unit-cell parameters, specic volumes, and particle sizes of the ZnS particles obtained at various annealing
temperatures. Reproduced from Ref. [15]. Copyright 1999, The American Physical Society.
Annealing
temperature
Phase Percent
of phase
Lattice parameters
()
Specic volume
(
3
)
Particle size
()
23 C Zinc blende 100 a = 5.42 0.01 38.8 + 0.3 21
c = 5.28 0.02
300 C Zinc blende 100 a = 5.42 0.01 38.7 0.2 29
c = 5.27 0.02
350 C Zinc blende 100 a = 5.41z:0.01 3S.7 0.3 32
c = 5.29 0.02
400 C Zinc blende 72 a = 5.404 0.012 39.5 0.3 74
400 C Wuitzite 28 a = 3.S2 39.5 74
c = 6.26
500 C Zinc blende 72 c = 5.41 39.6 232
500 C Wmtzite 28 a = 3.82 39.6 243
c = 6.26
Fig. 4. Plot of energy differences between WZ and ZB structures DE
WZZB
(triangles), 6H and ZB structures DE
6HZB
(squares), 4H
and ZB structures DE
4HZB
(circles), and for a ZnS nanowire as a function of nanowire diameter. Reproduced from Ref. [16].
Copyright 2007, The Japan Society of Applied Physics.
180 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
of WZ-ZnS nanowires, which results in the lower stability. Therefore, when the size of the nanostruc-
ture is small so that a large number of constituting atoms reside on the surface, it prefers to be in the
WZ phase with a fewer number of dangling bonds.
Different theoretical models have been developed in order to understand the electronic properties
of ZnS nanostructures [1719,8]. Wang and co-workers calculated the energy bandgaps of WZ-ZnS
nanowires using DFT [17]. They showed that the ZnS nanowires have wider bandgaps than that of bulk
ZnS crystal. The bandgap decreases with increasing diameter. More recently, the geometric, energetic,
Fig. 6. Band structures of a WZ-ZnS (a) nanowire; (b) double-walled nanotube with faceted morphology; (c) (9, 0) single-walled
nanotube; (d) bulk crystal. The energies at the Fermi levels are set to zero. Reproduced from Ref. [18]. Copyright 2008,
American Chemical Society.
Fig. 5. Formation energies of ZnS nanowires (solid circles), single-walled nanotubes (up-triangles), double-walled nanotubes
(squares) and triple-walled nanotubes (down-triangle) as a function of tube diameter. The dashed violet lines represent the
formation energies of ZnS nanosheets containing one to four atomic layers, respectively. Reproduced from Ref. [19]. Copyright
2008, Institute of Physics.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 181
and electronic properties of WZ-ZnS nanowires with hexagonal cross sections have been studied by
using rst-principles calculations [18,19,8]. Theoretical investigations were also carried out to study
the properties of different ZnS nanostructures, such as nanowires, nanotubes and nanosheets. In addi-
tion, the evolution of energy of nanowires and nanotubes as a function of diameter and wall thickness
was calculated.
Fig. 5 shows the evolution of formation energy as a function of nanostructure diameter. The forma-
tion energy of the nanowires decreases drastically with increasing diameter. Interestingly, it was also
found that the energy is proportional to the inverse of diameter. Furthermore, both nanowires and
nanotubes with faceted morphologies were found to be energetically more favorable than nanotubes
with round cross sections. The formation energy of the thick-walled nanotubes is very close to that of
nanosheets of the same thickness.
Fig. 6 shows the plot of the electronic band structures of the nanowires, nanotubes and bulk WZ-
ZnS [18]. The nanostructures have wider bandgaps as compared to a bulk crystal due to quantum size
effect. The highest valence band (HVB) and the lowest conduction band (LCB) of the ZnS nanostruc-
tures have relatively high dispersions along the growth direction. Fig. 7 depicts the partial electron
density of states (PDOS) of a ZnS nanowire. Notably, 3p states of both the bulk and the surface S atoms
contribute to the HVB, whereas the LCB arises mainly from the Zn(4s) states of the bulk.
3. Synthesis of ZnS nanostructures
Nanostructures have attracted steadily growing interest due to their fashinating properties, as well
propertymicrostructure correlation [6,20]. They can be divided into three kinds, namely 0D, 1D, and
2D nanostructures based on their shapes.
3.1. 0D nanostructures
In the past decades, signicant progress has been made in the eld of 0D nanomaterials and nano-
structures. A rich variety of methods have been developed for fabricating 0D nanostructures with
well-controlled dimensions [2125]. In this section, we highlight some progress in the synthesis
Fig. 7. Projection of PDOS of a ZnS nanowire onto the surface atoms (S
/
, Zn
/
) and bulk atoms (S, Zn). Red and blue lines represent
S and Zn, respectively. The energies at the Fermi levels are set to zero. The arrows indicate the positions of the peak of the 3p
states of S. Reproduced from Ref. [18]. Copyright 2008, American Chemical Society.
182 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
and characterization of 0D ZnS nanostructures, such as nanocrystals (quantum dots), core/shell nano-
crystals (NCs), and hollow nanocrystals.
3.1.1. 0D nanocrystals (quantum dots)
Initially, Brus and co-workers successfully synthesized ZnS nanocrystals with high colloidal stabil-
ity in aqueous and methanolic media using Na
2
S and Zn(ClO
4
)
2
as precursors. However the size tun-
ability and monodispersity of these nanocrystals were limited [26]. Bawendi and co-workers
developed a ground-breaking nonohydrolytic method to prepare nearly monodisperse CdSe nanocrys-
tals by injecting a solution containing dimethyl cadmium (Cd(CH
3
)
2
) and trioctylphosphine selenide
(TOPSe) into hot trioctylphosphine oxide (TOPO) [27]. Since then this nonohydrolytic chemical route
has been the most widely adopted method for synthesizing high-quality inorganic nanocrystals
including ZnS nanocrystals [28]. Hyeon and co-workers fabricated cubic ZnS nanocrystals with various
sizes and shapes under a thermal reaction of ZnCl
2
and S in oleylamine in the presence of TOPO. A
Fig. 8. (a) TEM image of ZnS nanocrystals fabricated by the thermal reaction of ZnCl
2
and S in oleylamine. Reproduced from Ref.
[29]. Copyright 2003, American Chemical Society. (b) TEM image of hexagonal ZnS nanocrystals (<5 nm) fabricated in a polyol
medium. Reproduced from Ref. [30]. Copyright 2004, American Chemical Society. (c) TEM image of ZnS nanocrystals
fabricated by thermolysing Zn(exan)
2
with TOP as precursor solvent. Reproduced from Ref. [32]. Copyright 2004, American
Chemical Society. (d) TEM image of Zn
x
Cd
1x
S (x = 0.78) fabricated by thermolysing Zn(exan)
2
and Cd(exan)
2
precursors.
Reproduced from Ref. [33]. Copyright 2005, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 183
low-magnication transmission electron microscopy (TEM) image of ZnS nanocrystals (Fig. 8a) shows
uniform nanoparticles with a size of 11 nm. High-resolution TEM (HRTEM) and XRD studies reveal
highly crystalline nature of cubic ZnS nanocrystals [29]. In contrast, Zhao and co-workers synthesized
hexagonal ZnS nanocrystals at temperatures as low as 150 C using ZnCl
2
and S as precursors in a
polyol medium [30]. As demonstrated in Fig. 8b, these ZnS nanocrystals are quite uniform in both
shape and size. An average size is 4.2 nm with a standard deviation of 0.6 nm. The inset of Fig. 8b
shows an HRTEM image of an individual particle where the lattice pattern illustrates the well-struc-
tured phase.
Peng and co-workers developed another synthetic approach to grow ZnS nanocrystals using green-
er chemicals at elevated temperatures. This was achieved by introduction of new terms: activation of
precursors and identical injection/growth temperatures. By using ZnO or Zn-fatty-acid precursors
rather than highly reactive alkyl zinc precursors, the environmentally more benign synthesis was pos-
sible [31]. Recently, a simple synthetic method using single-source molecular precursor was reported
by Li and co-workers (Fig. 8c) [32]. The injection of zinc ethylxanthate (Zn(exan)
2
)/trioctylphosphine
(TOP) solution into a hot hexadecylamine (HDA) + TOP or octylamine (OA) mixed solution yields well-
dened ZnS nanocrystals. The authors demonstrated that in the HDA + OA system, diameter- and as-
pect-ratio-tunable hexagonal ZnS nanorods were attained in the temperature range of 150250 C,
while in the HAD + TOP system, a shape change from rod to spherical particle and a phase transition
from hexagonal to cubic simultaneously occurred with an increase of TOP content in the solution. This
approach can be successfully applied for the synthesis of Zn
x
Cd
1x
S nanocrystals by thermolyzing a
mixture of Cd(exan)
2
and Zn(exan)
2
precursors (Fig. 8d) [33].
3.1.2. 0D core/shell nanocrystals
Besides the development of synthesis techniques to prepare semiconductor NCs with narrow size
distribution, an intense experimental work has been devoted to the molecular modication toward
the improvement of the luminescence efciency and colloidal stability of the NCs. Post-synthesis inor-
ganic surface modication not only passivated the nanocrystals but also improved quantum efcien-
cies (QE) [23]. Thus this may be applicable to a core/shell system, where the bandgap of the core lies
within the bandgap of the shell material and the photogenerated electrons and holes are mainly con-
ned inside the core material. ZnS usually acts as a shell material in such core/shell system. Weller and
co-workers reported the fabrication of CdSe/ZnS core/shell nanoparticles with high reaction yields and
photoluminescence (PL) QE of 5060% [34]. Firstly, monodispersed CdSe nanocrystals were prepared
in a three-component HDATOPOTOP mixture. The PL QE of CdSe nanocrystals was only in the range
of 1025% and had a tendency to decrease with increasing particle size. Secondly, ZnS shell was grown
on the surface of CdSe nanocrystals. The amount of Zn:S stock solution necessary to obtain the desired
ZnS shell thickness was calculated from the ratio between the core and shell volumes using bulk lat-
tice parameters of CdSe and ZnS. Fig. 9a shows a HRTEM image of CdSe/ZnS nanocrystals (1.6 mono-
layers of ZnS) synthesized from -4.0 nm large CdSe cores. Fig. 9b shows a set of absorption and PL
spectra of CdSe/ZnS nanocrystals with different thicknesses of the shells. A maximum of the PL QE
(66%) was observed from a 1.6 monolayer (MLs) thick ZnS shell and was reproducibly 50% or more
for a wide range of the shell thicknesses and core sizes [34].
Combining the advantages of two or more shell materials, Mews and co-workers fabricated CdSe/
CdS/Zn
0.5
Cd
0.5
S/ZnS multishell nanocrystals with high QE of 7085% [35]. This CdSe-NC with a sand-
wich shell-structure consists of 1 MLs CdS followed by 3.5 MLs Zn
0.5
Cd
0.5
MLs and nally 2 MLs ZnS as
the outermost shell, as shown in Fig. 9d. The Zn
0.5
Cd
0.5
S is used as a buffer layer because the bandgap
as well as the lattice parameters can in principle be adjusted by the composition of alloyed materials.
The authors developed a successive ion layer adhesion and reaction (SILAR) technique, and could grad-
ually change the shell composition from CdS to ZnS in the radial direction. Due to the stepwise adjust-
ment of the lattice parameter in the radial direction, the resulting multi-shell NCs showed a high-
crystallinity and were almost perfectly spherical, as was detected by XRD and TEM (Fig. 9c). Also, be-
cause of the radial increase of the respective valence- and conduction-band offsets, the NCs were well
electronically passivated. This led to a high orescence QE of 7085% for the amine terminated mul-
tishell particles in organic solvents and a QE of 50% for mercapto propionic acid (MPA)-covered par-
ticles in water [35].
184 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
3.1.3. 0D hollow nanocrystals
Hollow nanospheres (HNSs) have attracted much attention due to their unique properties includ-
ing low density, high specic surface area and good permeation; and potential applications in cataly-
sis, nanoelectronics, nano-optics, drug delivery systems and lightweight structural materials [3640].
General methods for preparing ZnS HNSs use polystyrene (PS) microspheres [41,42] and SiO
2
spheres
[43] as sacricial templates, but the sizes of the obtained hollow spheres are larger than 100 nm. Xu
and co-workers demonstrated a facile template-free route for mass production of ZnS HNSs with
nano- and submicro-sizes, employing thiocarbamide, (CH
3
COO)
2
Zn2H
2
O and water as the raw mate-
rials [44]. ZnS HNSs were obtained via hydrothermal reaction at 140 C for 5 h. XRD analysis showed
that these hollow spheres had been of hexagonal phase structure. A TEM image of the ZnS spheres
treated with dilute HCl solution to thin thickness of the shell is shown in Fig. 10a. These hollow
spheres are homogeneous and the size ranges from 200 to 300 nm, most of them falling in the range
of 240260 nm. The mean thickness of the shells is about 20 nm. In contrast, Qi and co-workers devel-
oped another synthesis of hollow ZnS nanospheres with diameters smaller than 100 nm using thioac-
Fig. 9. (a) HRTEM image of CdSe/ZnS core/shell nanoparticles (CdSe covered with 1.6 monolayers of ZnS). (b) RT absorption and
emission spectra of CdSe nanocrystals before and after deposition of ZnS shells of different thicknesses (in monolayers, ML).
Reproduced from Ref. [34]. Copyright 2001, American Chemical Society. (c and d) HRTEM image and structural model of a
CdSe/CdS/Zn
0.5
Cd
0.5
S/ZnS multishell nanocrystal. Reproduced from Ref. [35]. Copyright 2005, American Chemical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 185
etamide as the sulfur source in aqueous solutions of the triblock copolymer P123 at RT [45]. Fig. 10b
presents a typical TEM image of hollow ZnS nanospheres, which shows that the hollow spheres are
rather uniform with diameters ranging from 50 to 70 nm and a shell thickness of -15 nm. HRTEM
Fig. 10. (a) TEM image of hollow ZnS spheres. Reproduced from Ref. [44]. Copyright 2005, Institute of Physics. (b) TEM image
of hollow ZnS nanospheres in the presence of P123. Reproduced from Ref. [45]. Copyright 2003, American Chemical Society.
(c and d) TEM image and HRTEM image of Fe
3
O
4
/ZnS core/shell hollow spheres. (e) Schematic diagram of the formation of
Fe
3
O
4
/ZnS HNSs. Reproduced from Ref. [46]. Copyright 2009, American Chemical Society.
186 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
and selected-area electron diffraction (SAED) results indicate that these ZnS nanospheres are cubic,
and the shells consist of ZnS nanocrystals.
Core/shell and hollow nanospheres have also attracted tremendous interest in recent research.
Very recently, Wu and co-workers have fabricated superparamagnetic uorescent Fe
3
O
4
/ZnS HNSs
with diameters of <100 nm using a corrosion-aided Ostwald ripening process [46]. When the mono-
disperse FeS particles were dispersed in a mixture containing zinc acetylacetonate (ZA), poly(vinyl-
pyrrolidone) (PVP), ammonium nitrate, glycol, and water and then reacted at 150 C for 10 h, the
Fe
3
O
4
/ZnS HNSs were obtained. A general view (Fig. 10c) indicates that the as-synthesized HNSs have
typical hollow structures with average internal and external diameters of 66 and 97 nm, respectively.
HRTEM image (Fig. 10d) reveals that the shell is composed of close-packed light and dark particles that
have regions oriented in the outer and inner layers, respectively. The particles have the interplanar
spacings of 0.312 and 0.486 nm, corresponding to the (1 1 1) planes of cubic ZnS and Fe
3
O
4
, respec-
tively. The SAED pattern indicates that Fe
3
O
4
and ZnS are both polycrystalline. Time-dependent exper-
imental results indicate that FeS particles act as in situ templates and ZA as an etching agent in the
growth process of Fe
3
O
4
/ZnS HNSs (Fig. 10e). These multifunctional HNSs or modied particles thereof
have a potential for exciting applications in biomedicine, such as drug delivery, integrated imaging,
diagnosis, and therapeutics [46]. Other 0D ZnS-related nanocrystals have been fabricated, as listed
in Table 2.
Table 2
Representative 0D ZnS-related nanocrystals.
Nanostructure Composition Synthesis method T (C) Ref.
Nanocrystals ZnS Thermal reaction with ZnCl
2
and S in oleylamine 320 [29]
Solution approach in a polyol medium 150 [30]
Chemical reaction in non-coordinating solvents 340, 300 [31]
Thermolysing Zn(exan)
2
with OA and TOP as
precursor solvents
150250 [32]
Liquidsolidsolution (LSS) process 90 [47]
Chemical reduction route RT [48]
Aging reaction mixtures 300 [49]
Core/shell
nanocrystals
CuInS
2
/ZnS (shell) Two-step procedure 230 [50]
Mn-CdS/ZnS (shell) Three-step synthesis method 280 [51]
CdS/ZnS (shell) Two-step procedure 250
280;180
[52]
CdSe/ZnS (shell) Lyothermal technique 250 [53]
CdSe/ZnS (shell) Organometallic synthesis in HDATOPOTOP
mixture
300;220 [34]
CdSe/ZnS (shell) Two-step synthesis 360, 220 [54]
CdSe/CdS/Zn
0.5
Cd
0.5
S/
ZnS (shell)
SILAR technique 200260 [35]
Hollow
nanocrystals
Fe
3
O
4
/ZnS (shell) Corrosion-aided Ostwald ripening process 150 [46]
ZnS Bacteria-template sonochemical route RT [37]
ZnS Hydrothermal route 140 [44]
ZnS Hydrothermal route 120 [38]
Doped
nanocrystals
Mn-ZnS Organometallic and inorganic synthesis method [55]
Mn-ZnS Solution approach RT [56]
Co-ZnS Solution-phase thermal decomposition method 150 [57]
Cu-, Pb-ZnS Organometallic synthesis method 300 [58]
Alloyed
nanocrystals
Zn
x
Cd
1x
S Solution chemical route 270 [59]
Zn
x
Cd
1x
S Thermolyzing a mixture of Cd(exan)
2
and Zn(exan)
2
precursors
210 [33]
Zn
x
Cd
1x
S Chemical reaction of CdO- and ZnO-oleic acid
complexes with sulfur
300 [60]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 187
3.2. 1D nanostructures
1D nanostructures have stimulated an increasing interest due to their importance in basic scientic
research and potential technological applications [61]. It is generally accepted that 1D nanostructures
are ideal systems for exploring a large number of novel phenomena at the nanoscale and investigating
the size and dimensionality dependence of functional properties. They are also expected to play
important roles as both interconnects and the key units in fabricating electronic, optoelectronic, elec-
trochemical, and electromechanical devices with nanoscale dimensions [5,62,63]. Several kinds of 1D
nanostructures have been reported in the literature. Fig. 11 shows the morphologies of ZnS-related 1D
nanostructures, such as nanowires, nanobelts, nanotubes, nanocombs, nanoawls, which have been
synthesized in our laboratory. In this section, we present some typical recently developed processes
for the synthesis of 1D ZnS-related nanostructures and their structural characterizations.
3.2.1. Nanowires (NWs), nanorods (NRs), and nanotubes (NTs)
Nanowires, nanorods, and nanotubes are the most important and popular 1D nanostructures. Typ-
ically, nanowires are relatively long, exible, and have a circular cross-section. Nanorods are shorter
(and therefore stiffer) and though circular cross-sections are prevalent, can often have hexagonal
cross-sections. When hexagonal cross-sections are present, the surfaces of the nanorods are well-fac-
eted [64]. Both nanowires and nanorods have solid structures, whereas nanotubes have hollow
structures.
3.2.1.1. NWs/NRs growth in vapor. Vapor-phase synthesis is probably the most extensively explored ap-
proach to form ZnS NWs/NRs. There are several processing parameters such as temperature, pressure,
carrier gas, substrate and evaporation time period that can be controlled and need to be properly se-
lected before or during the vapor-phase synthesis. Several techniques can be assigned to vapor-phase
methods, such as thermal evaporation, chemical vapor deposition (CVD), molecular beam epitaxy
Fig. 11. Several typical morphologies of ZnS-related 1D nanostructures fabricated in our laboratory.
188 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
(MBE), laser ablation and metalorganic CVD (MOCVD), which are utilized for the growth of ZnS NWs/
NRs, as shown in Table 3. For example, Wang and co-workers reported the fabrication of single-crys-
talline ZnS nanowires by thermal evaporation of ZnS powders with the presence of Au lms as a cat-
alyst at 900 C [65]. Scanning electron microscopy (SEM) image (Fig. 12a) reveals that these
synthesized products consist of a large number of nanowires with typical lengths of several tens of
micrometers, and diameters of 3060 nm. TEM observation shows that most of the ZnS nanowires
terminate with an Au nanoparticle at one end. HRTEM image and SAED pattern indicate that these
nanowires grow along the [1 1 0] direction [65]. Lieber and co-workers demonstrated the rst use
of single-source molecular precursors for the growth of ZnS nanowires at a low temperature [66].
By using well-dened molecular reactants, the authors were able to prepare single-crystalline ZnS
Table 3
Synthetic routes for ZnS nanowires/nanorods.
Nanostructure Synthesis method T (C) Ref.
Nanowires Thermal evaporation method 900 [65]
MOCVD process [66]
Liquid crystal template by c-irradiation RT [68]
Solvothermal route 180 [69]
One-step wet-chemical approach 140 [70]
VLS process 1200 [71]
Thermal evaporation 400500 [72]
Molecular beam epitaxy (MBE) technique 430 [73]
Solvothermal route 180 [74]
Mild-solution chemistry approach 180 [75]
Thermal evaporation method 1250 [76]
Electrochemical-template method RT [77]
Thermal evaporation process 1300 [78]
Vapor phase deposition method 900 [79]
Thermal evaporation of zinc powder and sulfur powder 580, 90 [80]
Electrochemical-template method 120 [81]
One-step CVD method 1200 [82]
Hydrothermal synthesis route 180 [83]
Electrodeposition-template method 120130 [84]
Hydrogen-assisted thermal evaporation method 1100 [85]
Thermal physical evaporation 1000 [86]
Carbothermal CVD process 500 [87]
Organic assistant VLS method 1000 [88]
VLS method 1100 [89]
Thermal evaporation method 700 [90]
Two-step thermal evaporation process 1050 [91]
Pulsed laser vaporization 950 [92,93]
Hydrogen-assisted thermal evaporation 1100 [94]
Intermittent laser ablation-catalytic growth process 950 [95]
Chemical vapor transport and condensation 900950 [96]
Direct reaction of Zn and S powders 750 [97]
Solid-state reaction RT [98]
Thermal evaporation method 1000 [99]
VLS process [100]
Thermal evaporation 1050 [101]
Nanorods Aqua-solution hydrothermal process 95 [102]
Micro-irradiation-assisted growth RT [103]
Solvothermal approach 135250 [104]
Suldation of ZnO nanorod arrays 600 [105]
Radio frequency magnetron sputtering technique 40 to 0 [106]
Solvothermal decomposition 150200 [107]
Solution method in the presence of block copolymer 105 [108]
Plasma-assisted MOCVD process 650 [109]
Thermolysing Zn(exan)
2
with OA and TOP as precursor solvents 150250 [32]
Thermal evaporation 1100 [110]
Thermal evaporation 970 [111]
Hydrothermal synthesis 180 [112]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 189
nanowires with controlled diameters and a high yield. Their synthetic approach follows the nanoclus-
ter-catalysted VLS growth process, where Zn(S
2
CNEt
2
)
2
molecular precursor serves as a source for Zn
and S reactants (Fig. 12b). In this synthetic method, the single-source precursor undergoes thermal
decomposition in the growth region, forms ZnSAu liquid solutions with the Au nanocluster cata-
lysts, and then undergoes nucleation followed by nanowire growth when the nanodroplets become
saturated with reactant. TEM structural analyses show that single-crystalline ZnS nanowires with con-
trolled diameters are obtained. A TEM image of the nanocluster catalyst at the end of a crystalline ZnS
nanowire (Fig. 12c) shows clearly how the nanocluster catalysts dene the diameter of nanowires.
HRTEM image and SAED pattern indicate that these ZnS nanowires are single-crystals growing along
the [1 0 0] direction [66].
3.2.1.2. NWs/NRs growth in solution. One of the disadvantages of high-temperature approaches to
nanowire synthesis is the high cost of fabrication and scaling-up. Recent progress using solution-
phase techniques, such as solvothermal, templating and chemical routes, has resulted in the creation
of 1D nanostructures in high yields (grams scale) using selective ligand precursors or templates [67].
For example, ZnS nanowires were fabricated via a direct templating route in an inverted hexagonal
liquid crystal under c-irradiation [68]. The c-irradiation plays an important role in the formation of
ZnS nanowires using thiourea or CS
2
as the sulfur source. Several close-packed nanowires aggregate
to form bundles with diameters of 1030 nm, which duplicate the hexagonal structure of close-packed
inverted micelles formed by amphiphiles. Moreover, ZnS nanorods were fabricated by a solvothermal
route through thermolysing Zn(exan)
2
using OA as a precursor solvent and HDA as the main ligand
stabilizer [32]. Using hydrazine hydrate (N
2
H
4
H
2
O) as a solvent, ZnS nanowire bundles with diame-
ters of 1025 nm and lengths about 58 lm were prepared in high yield via a solvothermal route. The
nanowires, growing along the [0 0 1] direction, were aligned not only in the length direction but also
in accord with the crystallography rules to form the bundle [69]. Other routes for the synthesis of ZnS
NWs/NRs can be found in Table 3.
3.2.1.3. Nanotubes (NTs). As an important member of the inorganic nanotubes family, recently the ZnS
nanotubes have become of growing interest. Wang and co-workers rst reported the synthesis of ZnS
Fig. 12. (a) SEM image of ZnS nanowires fabricated by thermal evaporation of ZnS powders under controlled conditions with
the presence of Au lms as a catalyst. Reproduced from Ref. [65]. Copyright 2002, Elsevier. (b) Nanowire growth from a
single-source molecular precursor via a gold nanocluster-catalysted vaporliquidsolid mechanism. (c and d) HRTEM images of
ZnS nanowires. Scale bar is 10 nm. Reproduced from Ref. [66]. Copyright 2003, American Chemical Society.
190 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
nanocables and nanotubes by a chemical reaction using as-synthesized ZnO nanobelts as a template
[113]. These structures were composed of ZnS nanocrystallites of -7 nm in size and had a high per-
centage of pores. Yin and co-workers fabricated faceted ZnS nanotubes by a high-temperature ther-
mal-chemical reaction route, in which commercial ZnS powders were used as precursors and H
2
O
vapor was carried by Ar gas and introduced into a graphite crucible to form reductive agents of CO
and H
2
[114]. Fig. 13a and b shows the SEM images of the synthesized ZnS nanotubes. All the products
display a tubular structure. From the cross-sectional morphologies of the ZnS nanotubes, it is clearly
seen that the ZnS nanotubes are perfectly straight and have highly faceted hexagonal cross-section
morphologies. As indicated in Fig. 13b, the nanotubes with hexagonal cross-sections grow along the
[0 0 0 1] direction and are closed by low-index faces. Typically, these nanotubes have a length of sev-
eral micrometers, a uniform outer diameter of 100120 nm, and a tube-wall thickness of about 10 nm.
Fig. 13c and d depict TEM images of two single ZnS nanotubes with hexagonal-faceted open ends
growing along the [0 0 1] direction, which suggest that the nanotubes are very pure, with no impurity
phase coating or lling. The HRTEM image and corresponding SAED pattern (Fig. 13f) again conrm
that these nanotubes are single-crystals growing along the [0 0 0 1] direction, and are closed by
low-index planes. For the formation process of faceted tubular ZnS, no templates or metal catalysts
were utilized. It was shown that the synthesis temperature and reactor gas pressure play an important
role in the formation of ZnS nanotubes, and detailed growth mechanism is discussed in Ref. [114]. By
using MOCVD-template methods, wurtzite-phase ZnS nanotube arrays were synthesized in the pores
of anodic aluminia membranes (AAM)s by Yao and co-workers [115]. Different from the conventional
thermal evaporation and CVD methods in which a ZnS powder/nanopowder or ZnS-based mixture
have always been adopted as source materials, herein low pressure thermal decomposition of zinc
bis(diethyldithiocarbamate) (Zn(S
2
CNEt
2
)
2
) has been employed. The well-ordered and polycrystalline
ZnS nanotubes with 140250 nm in outer diameter and several tens of micrometers in length were
formed [115]. Other synthetic routes for ZnS nanotubes can be found in Table 4.
3.2.2. Nanobelts (NBs), nanoribbons (NRs) and nanosheets (NSs)
Unlike 1D nanostructures (nanotubes, nanowires and nanorods) discussed above, the nanobelts
(NBs)/nanoribbons (NRBs) and nanosheets (NSs) have a rectangular cross section with well-dened
Fig. 13. (a and b) SEM images of the faceted ZnS nanotubes; (c and d) Two single faceted ZnS nanotubes growing along the
[0 0 1] direction; (e) Growth habit for the synthesized ZnS nanotubes; and (f) HRTEM image and SAED pattern of ZnS nanotubes.
Reproduced from Ref. [114]. Copyright 2005, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 191
geometry and high-crystallinity. After discovery by the Zhong Lin Wangs group in 2001, NBs/NRBs/
NSs have become one of the most interesting nanoobjects [123].
Lee and co-workers introduced a convenient hydrogen-assisted thermal evaporation method to
synthesize ZnS nanoribbons. The synthetic reaction was carried out in a quartz tube furnace using
high-purity Ar premixed with 5% H
2
as carrier gas at 1100 C, and sphalerite ZnS as sources [124].
These ZnS nanoribbons have a typical length in the range of several tens to several hundreds of
micrometers, a width ranging from200 nm to 400 nm, and a thickness from one fteenth to one twen-
tieth of the width (Fig. 14a and b). The width and length of the nanoribbons were sensitive to the
Table 4
Synthetic routes for ZnS nanotubes.
Nanostructure Synthesis method T (C) Ref.
Nanotubes Conversion of ZnO nanobelts 1350, 90 [113]
High-temperature thermal-chemical reaction route 15001700 [114]
MOCVD-template method 400 [115]
Ultrasonic chemical solution method RT [116]
Wet-chemistry method 140 [117]
Hydrogel template route RT [118]
Thioglycolic acid-assisted solution route 130180 [119]
Atomic layer deposition 75 [120]
Thermochemical process 1250 [121]
Chemical conversion of ZnO columns 400 [122]
Solution route by using the hard templates of CNTs 120 [39]
Fig. 14. (a and b) SEM images of ZnS nanoribbons fabricated by a hydrogen-assisted thermal evaporation method; (c and d)
TEM image, EDS spectrum, and HRTEM image of ZnS nanoribbons. Reproduced from Ref. [124]. Copyright 2003, Wiley-VCH.
192 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
reaction time and temperature, while the thickness of the ribbons was not sensitive to the growth con-
ditions. X-ray energy-dispersive spectroscopy (EDS) spectrum and XRD pattern show that the sample
is hexagonal-structured ZnS with the lattice constants of a = 3.822 and c = 6.257 . TEM images
show that each nanoribbon is long and straight, and has a uniform width and thickness along its entire
length. Dark lines on the ribbons that appeared in TEM images are due to strain resulting from ribbon
bending. HRTEM image (Fig. 14d) reveals that the ZnS nanoribbons are structurally uniform and single
crystalline. The corresponding SAED pattern indicates that the ribbon grows along [1 2 0], and it is en-
closed by (0 0 1) and (2 1 0) planes.
The hydrogen-assisted thermal evaporation process can be expressed as follows:
ZnS(fcc) H
2

1000

C
Zn(g) H
2
S (1)
Zn(g) H
2
S
900

C
ZnS(hcp) H
2
(2)
Table 5
Synthetic routes for ZnS nanobelts, nanoribbons, and nanosheets.
Nanostructure Synthesis method T (C) Ref.
Nanobelts CVD process 1100 [125]
Vapor-phase transport process 900 [126]
Hydrogen-assisted thermal evaporation 1120 [127]
Under moist gas conditions
Thermal evaporation 1100 [128]
Thermal evaporation 1100 [129]
Conversion reaction of CdS NBs 800 [130]
Rapid CVD process with Au as catalyst and graphite as reductant 1050 [131]
Thermal evaporation 1150 [132]
VLS process 1050 [133]
Thermal evaporation 1150 [134]
Solidvapor phase thermal-sublimation technique 1050 [135]
Thermal evaporation 900 [136]
Diethylenetriamine-assisted solvothermal approach 180 [137]
Controlled thermal process 1150 [138]
Low-temperature thermochemistry route 640 [139]
VLS process 1000 [140]
Thermal evaporation 970 [141]
VPT process 900950 [142]
Thermal evaporation 1200 [143]
H
2
-assisted thermal evaporation approach 9001100 [144]
Thermal evaporation 1050 [145]
Thermal evaporation 1600 [146]
Thermal evaporation 1100 [147]
Solvothermal reaction and subsequent heat treatment 160, 250 [148,149]
Thermal evaporation 1020 [150]
Nanoribbons Au mediated thermal evaporation route 10001100 [151]
CVD method 450 [152]
Thermal evaporation using thiol-capped Au NPs as catalysts 1000 [153]
Thermal evaporation process 1050 [154]
Two-stage temperature-controllable thermal evaporation and condensation
process
1180 [155]
Thermal evaporation 1050 [101]
Microwave-assisted solvothermal synthesis 400 [156]
Hydrogen-assisted thermal evaporation 1100 [124]
Nanosheeets Solvothermal method 180 [157]
Solvothermal method 180 [158]
Solution method in the presence of block copolymer RT [108]
Thermal evaporation process 1050 [159]
Thermal decomposition and reduction 1500 [160]
Solvothermal method 120180 [161]
Solvothermal reaction 120180 [162]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 193
The reaction between the sphalerite ZnS powder and H
2
at high temperature forms Zn vapor and
H
2
S gas, which are transported to the lower temperature zone where they react with each other and
then crystallize as wurtzite ZnS crystals. This is due to the temperature gradient along the tube axis,
which will provide the external driving force for the new phase crystal growth [124]. Other synthetic
routs for ZnS nanobelts, nanoribbons and nanosheets may be found in Table 5.
3.2.3. Aligned nanowires and nanobelts
Growth of aligned nanostructures is important for applications in lasers, LEDs and FETs [163,164].
Aligned growth of nanostructures can be achieved with the use of substrates and catalysts or seeds.
The orientation-aligned ZnS nanowire bundles have been rstly demonstrated on a Si (1 1 1) substrate
with CdSe as a buffer layer after a two-step vapor deposition process [91]. The rst step deposits the
CdSe base and the second step forms the ZnS nanowire bundles. During the growth, bundles of ZnS
nanowires grow on the top of a CdSe layer, which serves as a buffer between the Si(1 1 1) substrate
and the nanowires. All of the ZnS nanowires are orientationally aligned, and they grow uniformly
along the bundle [64,91].
Aligned ZnS nanorods/nanowires have also been successfully grown on several substrates through
heteroepitaxial and homoepitaxial growth [71,76,109,132,148]. The crystal structure of a substrate
was crucial for the orientation of nanowires. Epitaxial relationship between the substrate surface
and ZnS nanowires determines whether there will be an aligned growth or not and how good the
alignment can be. For example, well-aligned ZnS nanorods had been grown on the c-plane Al
2
O
3
sub-
strates by plasma-assisted MOCVD without using any metal catalysts [109]. The lattice mismatch be-
tween the Al
2
O
3
substrate and hexagonal ZnS is only -3%. ZnS nanowire arrays on Zn
3
P
2
crystals were
fabricated via a thermal evaporation method [76]. The fabrication of quasi-aligned ZnS nanobelt arrays
was achieved through a non-catalytic and template-free thermal evaporation process [132]. Well-
aligned ZnS nanobelt arrays on Zn foils were prepared by a simple solvothermal reaction and subse-
quent heat treatment [148]. These well-aligned ZnS nanowire/nanobelt arrays are expected to be
promising candidate for various applications.
3.2.4. Complex nanostructures
Complex nanostructures with modulated compositions, structures and interfaces have recently be-
come of particular interest with respect to potential applications in nanoscale building blocks of future
Fig. 15. Illustration of assembling individual 1D ZnS nanostructures into complex nanostructures.
194 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
optoelectronic devices and systems [165168]. Fig. 15 illustrates assembling of individual 1D ZnS
nanostructures into complex nanostructures, such as longitudinal nanoheterostructures (LONHs),
coaxial (core/shell) nanoheterostructures (LONHs), side-by-side nanostructures, alloyed nanostruc-
tures, doped nanostructures, tetrapodal nanostructures, bicrystalline nanostructures, ZnS/organic hy-
brid nanostructures, and hierarchical nanostructures. In this section, we will highlight some of the
important synthesis of 1D ZnS-based complex nanostructures.
3.2.4.1. Longitudinal nanoheterostructures (LONHs). Silica nanotube-shelled GaZnS nanowire hetero-
structures were prepared through thermal evaporation of ZnS, Ga
2
O
3
, and SiO powder source materi-
als by desired thermo-chemical reactions inside a vertical induction furnace [169]. Fig. 16 shows the
typical TEM images of the heterostructures in a product. The clear contrast variations within the struc-
tures suggest that an individual nanostructure consists of the segments of different materials. More-
over, each heterostructure is sheathed over its entire length within a thin tube. EDS analysis reveals
that the light and dark contrasting segments are composed of ZnS and Ga, respectively, whereas the
Fig. 16. Typical TEM images of the silica-sheathed GaZnS nanowire heterojunctions: (a and b) Junction areas at the
nanostructure center and at the end of a silica tube, respectively. (c and d) Two GaZnS junction areas formed by the Ga and ZnS
nanowire end-to-end contacts within a silica tube. (e) The ends of Ga and ZnS nanowires are not in contact, leaving a narrow
part inside the tube. (f) Several junction areas formed by periodic Ga and ZnS nanowire end-to-end contacts along the tube axis.
Reproduced from Ref. [169]. Copyright 2005, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 195
sheathed tubes are made of SiO
2
. Most of the GaZnS nanowire heterostructures have diameters of
-150250 nm (a few of them have a diameter of -80120 nm); the shielding silica tube walls are
-48 nm thick [169171]. Other ZnS LOHs such as ZnOZnS, ZnScyclohexylamine, or multi-walled
carbon nanotubes (MWCNTs)ZnS LOHs are listed in Table 6.
3.2.4.2. Coaxial (core/shell) nanoheterostructures (CONHs). 3.2.4.2.1. ZnS-core CONHs. Coaxial nanohet-
erostructures (CONHs), e.g. core/shell nanostructures, are fundamentally interesting and have signif-
icant technological potential. CONHs can be fabricated by coating a material with a conformal layer of
the second material. For example, BN/coated ZnS nanoarchitectures (or ZnS/BN core/shell nanostruc-
tures) were synthesized by heating ZnS twinned-crystal whiskers in the presence of BNO vapors in a
N
2
/NH
3
atmosphere in an induction furnace [175]. ZnS twinned-crystal whiskers were preliminary
fabricated through controlled evaporation of ZnS powders at 1200 C under N
2
atmosphere. Fig. 17a
is an SEM image of the as-synthesized products, showing ZnS nanospine arrays grown on the
twinned-crystal whiskers. The nanospines are aligned on both sides with an angle of 59 (Fig. 17b),
forming a shbone-like architecture. A TEM image of the typical structure of ZnS nanospine arrays
is depicted in Fig. 17c, suggesting that the nanospines have sharp tips of only several nanometers thick
and wide bottom of 100 nm at the joint with the whisker. The SAED pattern indicates that each of
Table 6
Representative 1D ZnS longitudinal nanoheterostructures.
Nanostructure Composition Synthesis method T (C) Ref.
Heterostructures GaZnS Thermal evaporation 14001500 [169]
ZnOZnS Thermal evaporation 1100 [172]
ZnSCyclohexylamine Solvothermal method 100180 [173]
MWCNTsZnS Ultrasonic pretreatment and heat treatments 180 [174]
Fig. 17. (a and b) SEM images, (c) TEM image and (d) HRTEM iamge of BN-coated ZnS nanoarchitectures, i.e. ZnS/BN core/shell
nanostructures. Reproduced from Ref. [175]. Copyright 2004, Wiley-VCH.
196 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
nanospines is a single crystal. Fig. 17d is an HRTEM image of the tip of a ZnS nanospine, which clearly
reveals the lattice-resolved image with the interplanar spacing d
001
= 0.626 nm, conrming that [0 0 1]
orientation is the preferred direction for ZnS nanospines. It also clearly displayed that layered BN
sheaths uniformly coat the ZnS nanospines. These BN-coated ZnS, in other words, ZnS/BN core/shell
nanoarchitectures have excellent stability and are promising for a wide range of nanotechnology
usages [175].
3.2.4.2.2. ZnS-shell COHNs. Using ZnO nanowires/nanorods as templates, ZnO/ZnS core/shell nano-
structures have been synthesized by a chemical reaction [113]. Based on the geometrical shape of
the ZnO nanobelt template, Wang and co-workers received ZnO/ZnS core/shell nanostructures by a di-
rect reaction of H
2
S with the surface layer of ZnO within the presence of water, in line with the
reaction:
ZnO H
2
S ZnS H
2
O (3)
Fig. 18a and b shows a comparison of the TEM images taken from ZnO nanobelts before and after
reacting with H
2
S. The interface between the ZnS shell and ZnO core is fairly sharp and there appears
to be no intermediate layer. A clear picture of the core/shell structure is given in Fig. 18c, which dis-
plays a composite core/shell structure with a broken ZnS surface layer. The corresponding SAED pat-
terns (Fig. 18d) and EDS spectra (Fig. 18e and f) indicate that the core is made of hexagonal single-
crystalline ZnO, and the shell consists of cubic nanocrystalline ZnS.
The conversion of ZnO nanobelts into ZnO/ZnS core/shell structures occurs in solution. Due to the
limited solubility of ZnO in water, the reaction is essentially a substitution reaction; thus, these core/
shell nanostructures still preserve the rectangular cross section. The pores in the structures may be
formed due to two factors. First, excess H
2
O produced in the reaction may be present in the structure
and lead to the pores. Second, from the structural viewpoint, cubic ZnO and hexagonal ZnS are incom-
patible in nature. Thus, the substitution reaction is unlikely to produce single-crystalline ZnS. The for-
mation of nanocrystallites is expected especially when the reaction proceeds at room temperature
[175,176].
Using a similar strategy, Xue and co-workers fabricated well-aligned ZnO/ZnS nanocable arrays and
then ZnS nanotube arrays by using ZnO nanorod arrays as templates [119]. First, ZnO nanorod arrays
were grown on a Zn foil substrate through direct oxidation of the foil with ammonium persulfate
oxide in the alkali solution. Such solution-based methods greatly facilitate the approach toward scal-
ing up ZnO nanoarrays (and ZnO/ZnS nanocable arrays and ZnS nanotube arrays) under relatively mild
reaction conditions and at low cost. Second, ZnO/ZnS nanocable arrays with ZnO as the inner core and
ZnS as the outer shell were synthesized via a thioglycolic acid-assisted solution route. Subsequent re-
moval of the ZnO core led to the formation of ZnS nanotube arrays. The evolution from ZnO nanorod to
ZnO/ZnS nanocable or ZnS nanotube arrays is due to the solubility difference between ZnO and ZnS
and to the assistance of thioglycolic acid. When ZnO nanorod arrays are introduced into HSCH
2
COOH
solution, ZnHS
+
complex could be formed between the lone pair electrons of sulfur atom of
HSCH
2
COOH molecule and the vacant d orbital of the Zn
2+
ions on ZnO nanorods, and then ZnS nucle-
ates and grows by dissolution of ZnO nanorods. After reaction over a certain time, ZnO/ZnS nanocables
can be obtained in line with the reactions:
HSCH
2
COOHZn
2
ZnHS

CH
2
COOH

(4)
ZnHS

S
2
2ZnS H
2
S (5)
Furthermore, ZnS-core core/shell nanostructures, such as ZnS/SiC, ZnS/SiO
2
, ZnS/Si, ZnS/C, and ZnS-
shell core/shell nanostructures, such as In/ZnS, Sn/ZnS, Zn
3
P
2
/ZnS, CdS/ZnS, TiO
2
/ZnS, CdSe/ZnS, have
also been synthesized as shown in Table 7.
3.2.4.3. Side-by-side heterostructures. Similar to the concept of creating a uniform sheath around a
nanowire is the idea of coating 1D nanostructures anisotropically, i.e., only along one or two sides
of the nanowire material, which will result in the formation of biaxial or triaxial nanostructures
[67,207]. Hu and co-workers fabricated SiZnS biaxial and ZnSSiZnS triaxial nanowires based on
a catalyst-free thermal evaporation of a mixture of SiO and ZnS powders via a two-stage process under
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 197
entire temperature control [208]. In this process, Si nanowires were rst formed from the dispropor-
tionation of SiO powders, and ZnS nanowires were then grown on the Si nanowire substrates via
thermal evaporation of ZnS powders. This results in the formation of side-to-side SiZnS biaxial and
Fig. 18. TEM image of ZnO nanobelts (a) before- and (b) after reaction with H
2
S, showing the formation of ZnO/ZnS core/shell
nanostructures. (c) A ZnO/ZnS nanocable with a broken ZnS shell, and (d) a corresponding SAED pattern recorded from the
region, showing the presence of a single-crystalline ZnO core and the nanostructured ZnS shell. (e and f) EDS spectra acquired
from the regions indicated in (c), which prove the local chemical composition. Reproduced from Ref. [113]. Copyright 2002,
Wiley-VCH.
198 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
ZnSSiZnS triaxial nanowires. Fig. 19a shows a scanning TEM (STEM) image of the segment of a
straight and long side-to-side biaxial nanowire, which reveals a clear and uniform interface between
the Si subnanowire (light contrast) and ZnS subnanowire (dark contrast). The Si, Zn and S elemental
maps (Fig. 19bd) and line-scanning elemental proles (Fig. 19e) from the composite nanowire di-
rectly feature a side-to-side geometry of the SiZnS biaxial nanowire, which is different from the char-
acteristics of the coreshell nanowire heterostructures [171]. The diameters of Si and ZnS
subnanowires within these biaxial structures are -30 nm and -40 nm, respectively. A HRTEM taken
from the interface domain between Si-side and ZnS-side indicates that the interface is homogeneous
and uniform at the atomic scale. Furthermore, a clear epitaxial relationship, i.e., (1 1 1)Si//(1 1 1)ZnS
and (1 1 1)Si//(1 1 1)ZnS, between the Si- and ZnS-sides is seen, and neither mist dislocations
(the lattice mismatch of 0.40%) nor strains are observed at the interface under HRTEM studies. The
sandwich-like ZnSSiZnS triaxial nanowire heterostructures, which consist of two separated ZnS
Table 7
Representative 1D ZnS-core and ZnS-shell COHs.
Nanostructures Composition Synthesis method T (C) Ref.
ZnS-core
CONHs
ZnS (core)/BN Heating ZnS twinned-crystal whiskers in the present of B
NO vapors
1200,
600
[175]
ZnS (core)/BN One-step CVD method 1200 [82]
ZnS (core)/SiC Two-stage thermal process 1150,
1400
[177]
ZnS (core)/ZnO MOCVD process 450 [178]
ZnS (core)/SiO
2
Thermal evaporation process 900 [179]
ZnS (core)/SiO
2
Thermal evaporation 1100 [180]
ZnS (core)/SiO
2
Volume and surface diffusion VLS process 1000 [181,182]
ZnS (core)/SiO
2
High-temperature VLS process 1350 [183]
ZnS (core)/SiO
2
Thermal evaporation 1200 [184]
ZnS (core)/Si Two-step thermal evaporation method 1060,
1260
[185]
ZnS (core)/Si Thermal evaporation method 1450 [186]
ZnS (core)/C One-step thermal evaporation 780800 [187]
ZnS (core)/C Thermal evaporation method 1300 [188]
ZnS-shell
CONHs
In/ZnS (shell) Carbon-thermal CVD technique 1300 [189]
Sn/ZnS (shell) Sn-nanorod-templated evaporation process 1150 [190]
Zn/ZnS (shell) Thermo-chemical process 1500
1600
[191]
Zn/ZnS (shell) Thermal reduction of ZnS 800,
1300
[192]
Zn-Cd/ZnS
(shell)
Thermal evaporation process 1200 [193]
Zn
3
P
2
/ZnS
(shell)
Thermo-chemical process 1250
1350
[194]
Mn-CdS/ZnS
(shell)
Two-step solvothermal process 160 [195]
TiO
2
/ZnS (shell) Wet chemical method RT [196]
ZnO/ZnS (shell) Hydrothermal process 160 [197]
ZnO/ZnS (shell) Catalyst-free thermal evaporation technique 1100 [198]
ZnO/ZnS (shell) Suldation of ZnO nanorod arrays 400 [105]
ZnO/ZnS (shell) Catalyst-free thermal vapor transport method 1000 [199]
ZnO/ZnS (shell) Thioglycolic acid-assisted solution route 130180 [119]
ZnO/ZnS (shell) Low-temperature reaction in Na
2
S solution 60 [200]
ZnO/ZnS (shell) Two-step chemical reaction 200 [201]
ZnO/ZnS (shell) Conversion of ZnO nanobelts 1350, 90 [113]
CdS/ZnS (shell) MOCVD method 280 [202,203]
Mn-ZnS/ZnS
(shell)
Solvothermal method 200 [204]
CdSe/ZnS (shell) Two-step solution growth 290, 120 [205]
ZnO/ZnS (shell) Low-temperature reaction in Na
2
S solution 60 [206]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 199
subnanowires and one central Si subnanowire, were also prepared. Fig. 19f shows a STEM image of a
segment of ZnSSiZnS triaxial nanowire, where the external surface and SiZnS interfaces are clean
and uniform, and the central Si subnanowire (light contrast) and two separated ZnS subnanowires
(dark parts) have a nearly equal diameter of 2530 nm. The Si elemental mapping (Fig. 19g) shows
that Si is located at the central area along the longitudinal direction of the heterostructure, while
the Zn and S elemental maps (Fig. 19h and i) display that Zn and S species are both distributed at
the left and right sides along the length and are absent in the center, directly revealing a three-layer
sandwich-like nanowire structure with well-dened compositional proles. Elemental proles across
the ZnSSiZnS nanowire heterostructure were also obtained (indicated by a line, Fig. 19f). The prole
of Si, Fig. 19g, has a peak in the center, while the proles of Zn and S both show two peaks at the right
and left sides with a gap in the center, verifying a three-layer sandwiched geometry [171,208].
Recently, two novel semiconducting heterostructures: hetero-crystalline-ZnS/single-crystalline-
ZnO biaxial nanobelts and side-to-side single-crystalline ZnS/ZnO biaxial nanobelts (Fig. 20) have been
synthesized via a simple one-step thermal evaporation method using gold as a catalyst [209]. In the
rst heterostructure a ZnS domain was composed of the hetero-crystalline superlattice (3C-ZnS)N/
(2H-ZnS)M[1 1 1][0 0 0 1] with the atomically smooth interface between WZ and ZB ZnS fragments,
Fig. 19. (a) STEM image and (be) Corresponding elemental mappings and line-scanning elemental proles of SiZnS biaxial
nanowires. (f) STEM image and (gj) Corresponding elemental mappings and line-scanning elemental proles of ZnSSiZnS
triaxial nanowires. Reproduced from Ref. [208]. Copyright 2003, American Chemical Society.
200 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
Fig. 20. (ac) Typical TEM images of two novel ZnS/ZnO biaxial nanobelt heterostructures. (df) HRTEM images recorded from
the ZnO side, the ZnS-side and the interface of hetero-crystalline ZnS/single-crystalline-ZnO biaxial nanobelts. (g and h)
Structural models of WZ-ZnS/ZnO and ZB-ZnS/ZnO interfaces marked with I1 and I2 in (f). Reproduced from Ref. [209].
Copyright 2008, American Chemical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 201
where N and M are the numbers of the atomic layers in the ZB and WZ-ZnS sections. Detailed HRTEM
results suggested that N and M had usually varied from 9 to 11 and from 24 to 28, respectively. The
alternating growth of WZ-ZnS/ZnO and ZB-ZnS/ZnO interfaces might reduce the stresses and system
energy, leading to the formation of hetero-crystalline-ZnS/single-crystalline-ZnO biaxial nanobelts.
For example, the length of 26 layers (Ls) on the WZ-ZnS side is slightly larger than the distance of
19 Ls on the ZnO side, while the other segment 11 Ls on the ZB ZnS side is slightly smaller than the
distance of 10 Ls on the ZnO side. Importantly, the structures remained stable [209,210]. Other
side-by-side heterostructures can be found in Table 8.
3.2.4.4. Doped 1D nanostructures. To utilize semiconductor nanostructures as building blocks of func-
tional nanodevices, it is important to synthesize themby having in the end diverse physical properties.
This could be realized via appropriate doping. Kim and co-workers synthesized Mn/Fe-doped and co-
doped ZnS 1D nanostructures via an Au-assisted chemical vapor transport method [216]. The key in
this synthetic process of co-doping is the use of metal chloride. Typical SEM images of Mn-doped,
Fe-doped and Mn/Fe-co-doped nanowires are shown in Fig. 21a, d and g. These nanowires were pro-
duced at high density. They uniformly covered the entire substrate. The nanowires are tens of microm-
eters long and 70100 nm in diameter. The HRTEM images (Fig. 21b, e and h) of these nanowires show
clear lattice fringes, which conrm their single-crystalline nature,The corresponding SAED patterns
demonstrate that these nanowires all grow along the [0 0 1] direction. The EDS spectra indicate that
Zn and S are major elements, with about 3 at.% Mn for Mn-doped ZnS nanowires (Fig. 21c), 2 at.%
Fe for Fe-doped ZnS nanowires (Fig. 21f), and 1 at.% Mn and 2 at.% Fe for Mn/Fe-co-doped ZnS nano-
wires (Fig. 21i), respectively. The authors suggested that the formation of metal-doped ZnS nanowires
could be explained by the following reaction pathway [216]:
ZnO(power) MS(powder) MCl
2
Zn
1x
M
x
S(nanowires) Cl
2
(gas)
M
3
O
4
(residues in the boat)(M = Mn or Fe) (6)
Except for metal-doped-ZnS nanostructures, non-metal-doped ZnS nanoribbons are also important.
N-doped ZnS nanoribbons were grown via a chemical vapor deposition using ZnS powder as the
source material and high-purity Ar premixed with 5% H
2
as the carrier gas. NH
3
was added to the reac-
tion atmosphere. These N-doped ZnS nanoribbons have a uniform ribbon-like geometry and lengths of
3040 lm, and their morphology remains similar to that of the undoped nanoribbons [217].
Other 1D ZnS-based doped nanostructures, such as N-doped, Cu-doped, Ga-doped, Eu-doped, and
Al-doped-ZnS nanostructures have been fabricated as summarized in Table 9.
3.2.4.5. Alloyed 1D nanostructures. Recent advances in ternary semiconductor nanostructures have
shown that their bandgaps and thus their physical properties can be tuned by changing constituent
stoichiometries [227]. Lee and co-workers fabricated Zn
x
Cd
1x
S nanoribbons with variable composi-
tions (0 6 x 6 1) by combining laser ablation of CdS with thermal evaporation of ZnS at 950 C
[228]. The Zn
x
Cd
1x
S nanoribbons had thicknesses of 5080 nm, widths of 0.55.0 lm, and lengths
up to several hundreds of micrometers, as shown in Fig. 22a and b. They displayed smooth surfaces,
and their morphology varied a little with a substrate temperature in the range of 550800 C. In con-
trast, the composition was highly dependent on the substrate temperature. HRTEM images and SAED
Table 8
Representative 1D ZnS-based side-by-side heterostructures.
Nanostructure Composition Synthesis method T (C) Ref.
Side-by-side heterostructures SiZnS Two-stage thermal evaporation process 1600, 1500 [208]
ZnSSiZnS
ZnOZnS Thermal evaporation 1100 [209]
ZnSSi Thermal co-evaporation 1100 [211]
ZnOZnS VLS technique 1050 [212,213]
ZnOZnS MOCVD method 900, 500 [214]
ZnOZnS Thermal evaporation 1050 [215]
202 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
patterns conrmed the single-crystal quality and could be indexed to a hexagonal structure. These
Zn
x
Cd
1x
S nanoribbons possessed sharp, tunable lasing emission within 340390 nm and 485
515 nm. Chen and co-workers synthesized Zn
x
Cd
1x
Se (x = 0, 0.2, 0.3, 0.5, 0.7, and 1) nanowires via
Fig. 21. SEM image, HRTEM image, and corresponding EDS spectra of (ac) Mn-doped, (df) Fe-doped, and (gi) Mn/Fe-co-
doped ZnS nanowires grown on Au-coated Si substrates. Reproduced from Ref. [216]. Copyright 2009, American Chemical
Society.
Table 9
Representative 1D ZnS-based doped nanostructures.
Nanostructure Composition Synthesis method T (C) Ref.
Doped 1D nanostructures Mn/FeZnS Au-assisted chemical vapor transport method 950 [216]
NZnS NBs CVD method 950 [217]
MnZnS NRs Solvothermal approach 200 [218]
MnZnS NRs Thermal annealing process 700 [219]
CuZnS NRs Solvothermal process 200 [220]
CuZnS NBs
MnZnS NBs
Vapor-phase transport method 950 [221]
GaZnS NWs Thermal evaporation 1350 [222]
Mn, CdZnS NSs CVD method 700 [223]
EuZnS NWs Vapor deposition method 1100 [224]
Cu, MnZnS NWs Controlled thermal process 1150 [138]
Cu, AlZnS NRs Self-assembly method RT [225]
MnZnS NBS Thermal evaporation 900 [226]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 203
a laser-assisted CVD method using sintered mixture of CdSe and ZnSe at 1000 C [229]. The composi-
tions of the alloyed nanowires could be adjusted by varying the precursor ratios of the laser ablated
target and the CVD deposition temperatures. These Zn
x
Cd
1x
Se nanowires had diameters of 60
150 nm and lengths of several tens of micrometers (Fig. 22e). The structures exhibited strong visible
PL from 712 to 463 nm as a nonlinear function of the compositions and could serve as important inte-
grated full-color display elements in nanotechnological applications. Fei and co-workers achieved
ZnS
x
Se
1x
(x = 0.80, 0.71, 0.60, 0.49, 0.35, and 0.21) nanowires via thermal evaporation of the mixtures
of ZnS and ZnSe powders at 1100 C.Various compositions could be easily obtained by changing the
mole ratio of ZnS and ZnSe in the source material [230]. The diameters of the ZnS
x
Se
1x
nanowires
with different compositions x were mostly distributed in the range of 100200 nm, and the lengths
were up to several tens of micrometers (Fig. 22f). The HRTEM and SAED studies indicated that these
nanowires were all single crystalline with a hexagonal structure, but they had different growth direc-
tions for different compositions, such as [2 1 0] for x = 0.73 and [0 0 1] for x = 0.33, respectively. PL
measurements demonstrated the tunable bandgap emission of the alloyed ZnS
x
Se
1x
nanowires
shifted continuously from 340 nm (pure ZnS) to 463 nm (pure ZnSe).
The process of fabricating alloyed nanostructures using ME powders (M = Zn, Cd; E = S, Se) with
high melting points (such as 1830 C for ZnS and 1750 C for CdS) involves vapor generation, transport
and deposition of target materials, and inevitably requires high temperatures or high-vacuum laser-
ablation operations to generate sufcient amounts of vapor for the deposition. Yao and co-workers
have developed a low-temperature route for the preparation of single crystalline ternary Zn
x
Cd
1x
S
nanocombs and zigzag nanowires via a one-step MOCVD approach by heating a mixture of
Zn(S
2
CNEt
2
)
2
and Cd(S
2
CNEt
2
)
2
powders to 420 C [227,233]. Fig. 22c shows typical growth of a
Fig. 22. (a and b) SEM images and EDS spectra of Zn
0.86
Cd
0.14
S and Zn
0.64
Cd
0.36
S nanobelts. Reproduced from Ref. [228].
Copyright 2005, Wiley-VCH. (c) SEM and TEM images of Zn
0.54
Cd
0.46
S nanocombs. Reproduced from Ref. [227]. Copyright
2006, Elsevier. (d) SEM image and EDS spectrum of Zn
0.78
Cd
0.22
S zigzag nanowires. Reproduced from Ref. [233]. Copyright
2006, Institute of Physics. (e) SEM image of Zn
x
Cd
1x
Se nanowires. Reproduced from Ref. [229]. Copyright 2006, American
Chemical Society. (f) SEM image of ZnS
x
Se
1x
nanowires. Reproduced from Ref. [230]. Copyright 2007, Wiley-VCH.
204 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
comb-like structure, with one side at and the other side with a shorter nanotooth. The diameters of
the comb ribbons ranged from 500 to 800 nm, and their lengths were in the range of 250 lm. These
teeth had a length of 300800 nm and a width of 30100 nm. Fig. 22d shows a SEM image of typical
zigzag nanowires. These nanowires have diameters of 200300 nm, widths of 200300 nm and
lengths of up to tens of micrometers. TEM results indicate that these nanowires are single crystalline,
with the growth axis of [0 0 1], by changing the growth direction from [1 1 3a
2
/2c
2
] to [1 1 3a
2
/2c
2
]
(a, c are the respective lattice constants of the a-axis and c-axis). The compositions x of the Zn
x
Cd
1x
S
nanocombs and nanowires are 0.54 and 0.78, respectively [233].
Other ZnS-based alloyed nanostructures, such as Zn
1x
Mn
x
S NWs and NBs, Zn
x
Cd
1x
S
y
Se
1y
NBs
Co
x
Zn
1x
S NWs, and ZnS
x
Se
1x
tetrapods have also been achieved as listed in Table 10.
3.2.4.6. Tetrapodal nanostructures. Since Alivisatoss group rst reported the synthesis of tetrapod-
shaped CdSe nanocrystals using a complicated thermal decomposition of organometallic precursors
[243], several techniques have been developed to prepare tetrapod nanostructures. For example,
ZnS tetrapods were synthesized through the direct reaction of Zn vapor with S vapor using a stainless
network as the collecting substrate [244]. The ZnS tetrapods were almost uniform in size and shape
(Fig. 23a), and they had four legs with equal lengths of 150 nm protruding from the center. HRTEM
observations show that each branch within this architecture has a triangular cross section connected
to form a tetrapod structure; the branches have the wurtzite phase and the core has the zinc blende
phase. These experimental results showed a direct evidence of the zinc blende octahedron nucleus
model, which could give a clue for explaining the formation of ZnS tetrapod nanostructures [244].
Using cubic CdSe nanocrystals as the seeds, Yao and co-workers fabricated size-tunable ZnS tetrapods
via a one-step seed-epitaxial MOCVD approach [245247]. The ZnS tetrapods had a CdSe nanocrystal
core at the center with four wurtzite ZnS arms growing out from the core along four [0 0 0 1] direc-
tions. The diameters of the ZnS tetrapods can be easily tuned by changing the distances between
the substrates and precursors. Furthermore, nanocable-aligned ZnS tetrapods were prepared through
thermal evaporation of ZnS and carbon mixed powders by Zhu and co-workers [248]. The ZnS tetra-
pods were linked by nanocables (Fig. 23d), and the cables passed through the center of the tetrapods
(Fig. 23e). The tetrapods were aligned together with ZnSC nanocables along the [1 0 0] direction. The
Table 10
Representative 1D ZnS-based alloyed nanostructures.
Nanostructures Composition Synthesis method T (C) Ref.
Alloyed 1D
nanostructures
Zn
x
Cd
1x
S NRs Combining laser ablation of CdS with thermal
evaporation of ZnS
950 [228]
Zn
x
Cd
1x
Se NWs CVD method assisted with laser ablation 1000 [229]
Co
x
Zn
1x
S NWs One-step thermal evaporation method 800 [231]
Zn
x
Cd
1x
S
Nanocombs
MOCVD process 420 [227,232]
Zn
x
Cd
1x
S zigzag
NWs
MOCVD process 400 [233]
Zn
x
Cd
1x
S NWs Ethylenediamine-assisted solvothermal approach 175 [234]
Ternary SiZnS
NWs
One-step thermal evaporation method 1150 [235]
Zn
1x
Mn
x
S NWs
and NBs
Vapor phase deposition 900 [236]
Zn
1x
Mn
x
S NWs Thermal evaporation 1020 [237]
Zn
1x
Mn
x
S NWs
and NBs
Ion implanted with Mn 1200 [238]
Zn
x
Mn
1x
S NRs Solvothermal process 200 [239]
ZnS
x
Se
1x
NWs Thermal evaporation 1100 [230]
ZnS
x
Se
1x
Tetrapods
Thermal evaporation 800850 [240]
Zn
x
Cd
1x
Se NWs MOCVD process 550 [241]
Zn
x
Cd
1x
S
y
Se
1y
NBs
Cothermal evaporation route 1050 [242]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 205
nanocable-aligned ZnS tetrapods may have potential applications in nanoelectronics and photonics.
Other synthetic routes for ZnS tetrapodal nanostructures can be found in Table 11.
3.2.4.7. Bicrystalline nanostructures. The periodic twinned structures formed in ZB ZnS with a (1 1 1)
facet as the twin plane [251], whereas the bicrystals were observed in WZ-ZnS. Bicrystalline ZnS nano-
structures composed of (0 1 2)/(1 0 3), (0 1 0)/(1 0 3) [252], (1 0 1)/(0 1 3) [253], (0 0 1)/(0 1 2) or
(0 0 1)/(0 1 3) [254] have been reported. For example, ZnS periodically twinned nanowires (PTNWs)
and asymmetrical polytypic nanobelts (APNBs) were synthesized via Au-assisted VLS processes
[251]. Fig. 24a shows a TEM image of a ZnS PTNW, revealing sequentially bright/dark contrast stripes
throughout the entire length of the wires. The SAED pattern (inset of Fig. 24b) shows a ZB structure
from the [1 1 0] zone axis, indicating the existence of a twin defect along the [1 1 1] growth directions.
A HRTEM image (Fig. 24b) further reveals periodically alternating twins along the [1 1 1] axis of the
wire: atomically sharp twin boundaries appear every 79 ZnS layers. The zigzag angles are 141
(70.5 + 70.5), in accordance with the relative rotational angle (1 1 1) twin crystals in face-centered
cubic (fcc) structures. Fig. 24c depicts a TEM image of a ZnS APNB, showing that straight strips parallel
to the belt axis are embedded inside the belt. The typical widths of the strips are 3050 nm. An SAED
pattern (Fig. 24d) indicates that the nanobelt contains both WZ and ZB phases. HRTEM image
(Fig. 24e) clearly reveals that the strip (as interlayer) is the ZB phase, whereas the left- and right-hand
Fig. 23. (ac) TEM and HRTEM images of ZnS tetrapods. Reproduced from Ref. [244]. Copyright 2006, Wiley-VCH. (df) TEM
images of nanocable-aligned ZnS tetrapods. Reproduced from Ref. [248]. Copyright 2003, American Chemical Society.
Table 11
Representative 1D ZnS-based tetrapodal nanostructures.
Nanostructures Composition Synthesis method T (C) Ref.
Tetrapodal
nanostructures
ZnS Direct reaction of Zn vapor with S vapor 1000 [244]
ZnS Seed-epitaxial MOCVD deposition 420 [245]
Nanocable-aligned ZnS tetrapods Thermal evaporation process 1100 [248]
ZnS Thermal evaporation method 800 [249]
ZnS Thermal evaporation process 1300 [250]
206 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
parts are WZ structures. The formation of ZnS PTNWs and APNBs is based on modulating mass diffu-
sion processes in the catalyst droplet and on the nanowire side surfaces, respectively, and more de-
tailed information can be found in Ref. [251]. Other 1D ZnS-based bicrystalline nanostructures have
also been synthesized as listed in Table 12.
Fig. 24. (a and b) TEM image, HRTEM image and SAED pattern of ZnS periodically twinned nanowires (PTNWs). (ce) TEM
image, SAED pattern and HRTEM image of ZnS asymmetrical polytypic nanobelts (APNBs). Reproduced from Ref. [251].
Copyright 2006, American Chemical Society.
Table 12
Representative 1D ZnS-based bicrystalline nanostructures.
Nanostructures Synthesis method T (C) Ref.
Bicrystalline ZnS NRs Thermal evaporation process 1050 [154]
ZnS Periodically twinned NWs and polytypic NBs VLS growth 1200 [251]
Bicrystalline ZnS MBs Chemical vapor deposition 1300 [252]
Dart-shaped tricrystal-ZnS NRs SiO-assisted thermal evaporation process 1100 [253]
ZnS bicrystal NRs Thermal evaporation 1150 [254]
Bicrystalline ZnS (core/shell) nanocables Physical deposition method 1000 [255]
Periodically twinned ZnS NWs Thermal evaporation 1100 [256]
ZnS heterocrystal and bicrystal structures Plasma enhanced chemical vapor deposition 700 [257]
Bicrystalline ZnS NRs Thermal evaporation 1050 [154]
ZnS tricrystals Thermal evaporation 1250 [258]
ZnS nanosaws Hydrogen-assisted thermal evaporation 1000 [259]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 207
3.2.4.8. ZnS hierarchical nanostructures.
3.2.4.8.1. Homoepitaxial growth. In particular, the hierarchical assembly of nanoscale building blocks
with a tunable dimensional and structure complexity is an essential step towards the realization of
multi-functionality of nanomaterials. Several synthetic approaches have been reported to assemble
3D branched and hyper-branched structures using a variety of materials and techniques including
self-assembled dendritic growth of nanowires [260,261], and growth of multi-branched nanowire
structures via sequential seeding of catalyst [262]. For example, Lee and co-workers reported the
homoepitaxial growth of ZnS nanowires and nanoribbons on the surfaces of micrometer-wide sin-
gle-crystal ZnS nanoribbon substrates. These homo-branched ZnS nanostructures were formed under
a two step thermal evaporation process using ZnS powders as precursors. First, the ZnS nanoribbon
backbones with lengths and widths of tens to hundreds of micrometers, and thickness of -40
200 nm were grown. Second, these ZnS nanoribbon substrates were sputter-coated with a 10 nm thick
Au lm for catalytic growth of ZnS nanowires and nanoribbons. For the products synthesized at
Fig. 25. (ad) SEM images of ZnS nanowires and nanoribbons grown on ZnS nanoribbons. (e) Schematic diagram showing the
alignments of nanowires grown on different surfaces of the substrates. (f) The scheme of the basic cell of hexagonal structure
showing the crystallographic relations of the lattice planes and directions involved in epitaxial growth. (g) Molecular structure
model showing the orientation of the cross-array nanowires with respect to the substrate. Reproduced from Ref. [263].
Copyright 2006, Wiley-VCH.
208 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
nanoribbon substrate temperatures of 800850 C for 1 h, well-aligned ZnS nanowire arrays with high
density were grown on both top and side surface of the substrate, as shown in Fig. 25a and b. The
nanowires grown on the side surface were aligned normal to the side surface, while that grown on
the top surface grew dominantly in two directions with an angle of approximately 60 to each other
and to the substrate surface (Fig. 25c). In sharp contrast, for the products fabricated at about 900 C,
ZnS nanoribbons perpendicular to the substrate surface were obtained (Fig. 25d). The TEM results
show that the nanowires grown on the top surface and side surface grow along the [2 1 0] and
[0 0 1] directions, respectively, implying that the growth direction of ZnS nanowires can be controlled
by using ZnS planes with different orientations as a substrate for homoepitaxy. The orientation rela-
tions between the epitaxial ZnS nanowires and the substrate are shown in Fig. 25eg. On the top sur-
face of the substrate, ZnS nanowires grow along the equivalent [2 1 0] and [1 1 0] directions,
respectively, forming the cross arrays of nanowires at 60 with respect to each other and to the
(2 1 0) surface of the surface. On the side of the substrate, the ZnS nanowires grow along the
[0 0 1] direction, or perpendicular to the side surface [263].
3.2.4.8.2. Heteroepitaxial growth. Agarwal and co-workers reported the fabrication of branched nano-
wire heterostructures, where the backbones and branches had been assembled with ZnS and CdS,
Fig. 26. (a) Schematic illustrating the synthesis of nanowire branched heterostructures. (b and c) SEM and HRTEM images of
ZnS nanowire backbones. (d) TEM image of Au catalysts deposited on ZnS nanowires. (e) SEM image of branched structures
showing multiple branches spread out from each backbone nanowire. (f) SEM image of an isolated branched structure showing
Au catalysts at the tips of branches, marked by arrows. Reproduced from Ref. [264]. Copyright 2007, American Chemical
Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 209
respectively [264]. Growth of branched and backbones with control over the compositions was en-
abled via sequential seeding of gold nanocluster catalyst. The growth method of ZnS (CdS) nanowires
was based on MOCVD process using single molecular precursors, where zinc (cadmium) diethyldithio-
carbamate (Zn(S
2
CNEt
2
)
2
), Cd(S
2
CNEt
2
)
2
)) was used to grow ZnS (CdS) nanowires. The strategy to syn-
thesize branched nanowire heterostructures is illustrated in Fig. 26a. First, ZnS nanowire backbones
with control over their composition and diameters were fabricated by using Zn(S
2
CNEt
2
)
2
) as precur-
sors and monodispersed Au nanoclusters as catalysts via VLS mechanism. The diameters of nanowires
ranged approximately from 80 to 120 nm, and the nanowire lengths were about 10 lm (Fig. 26b). The
TEM results indicate that the majority of the ZnS nanowires growalong [1 0 0]. Next, the ZnS nanowire
backbones were seeded with Au colloidal solution followed by sample air-drying for secondary cata-
lyst deposition. Fig. 26d shows a TEM image of ZnS nanowires after the Au catalyst deposition, reveal-
ing that Au nanoparticles are attached to ZnS nanowires. Finally, the branches were grown by placing
the ZnS nanowires decorated with secondary catalyst in the growth furnace with a ow of CdS precur-
sor material. The nanowires branches were formed on the substrate as observed in Fig. 26e. As seen in
Fig. 26f, the Au nanoparticles were still found at the tips of branches (marked by arrows), and the
diameter distribution of the branches were estimated to be similar to the size distribution of the
Au catalyst. Further microscopy analysis indicated that the growth of heterostructures branches takes
place epitaxially, thus maintaining over all single-crystalline nature of the entire structure. This kind
of new nanowire heterostructures are distinct from axial and radial nanowire heterostructures, pos-
sesses large surface area and is expected also be useful as nanoelectronic and photonic building blocks
to control the generation and transportation of carriers in three dimensions [264]. Subsequently, many
efforts have been focused on the integration of these building blocks into complex functional architec-
tures. Several hierarchical nanostructures, such as ZnO nanorod arrays on ZnS nanobelts [265],
ZnS
x
Se
1x
nanowire arrays on ZnS nanoribbons [266], ZnS nanowire arrays on CdS nanoribbons
[267], ZnS nanowire on ZnO nanobelts [268], have been constructed subsequently, as shown in Table
13.
Table 13
Representative 1D ZnS-based hierarchical nanostructures.
Nanostructure Composition Synthesis method T (C) Ref.
/Hierarchical
nanostructures
ZnS NWsZnS NRs Homoepitaxial growth on ZnS NRs 1200, 1050 [263]
CdS NWsZnS NWs MOCVD process and VLS growth 950, 740 [264]
ZnO NRsZnS NBs Two-step vapor method 1150, 650 [265]
ZnS
x
Se
1x
NWsZnS NRs Metal-catalyzed VLS growth method 1150, 1020 [266]
ZnS NWsCdS NRs Metal-catalyzed VLS growth method 880, 1050 [267]
ZnS NWsZnO NBs Thermal evaporation method 1100 [268]
ZnS nanocantilever structuresZnS
NRs
Catalyst-assisted post-annealing
treatment
1050, 600 [269]
ZnS NWsZnS NWs VLS growth process 950 [270]
ZnS nanoneedlesZnS NWs Thermal evaporation via VLS
mechanism
950 [271]
ZnS tetrapod tree-like
heterostructures
Thermal evaporation method 1200 [272]
ZnS nanosaws Catalyst-free solidvapor deposition
technique
1000 [273]
ZnS hexagonal pyramids Thermal evaporation 1150 [274]
ZnS nanoawls Two-step pressure-controlled vapor
phase deposition
800 [275]
Multi-angular branched ZnS
nanostructures
Thermal evaporation approach 1050 [276]
ZnS nanostructures with
hierarchical architectures
Low-temperature solution route 4 [277]
ZnS hierarchical structures Hydrothermal route 180 [278]
ZnS nanohelices Vapor deposition process 1000 [279]
210 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
3.2.4.9. ZnS/organic 1D nanostructures. Incorporation of inorganic and organic materials into a single
structure may generate inorganicorganic (I/O) hybrid composites that enhance or combine the useful
properties of both components [280]. A new I/O nanocompositie (ZnS/NaSCH
2
COONa) with a ribbon-
like structure was synthesized by using Zn(NO
3
)
2
7H
2
O and thioglycolic acids (TGA) as reactants under
hydrothermal conditions at 160 C [281]. These I/O nanoribbons have typical length of up to tens of
micrometers and widths in the range of 200500 nm (Fig. 27b). TEM and HRTEM observations demon-
strate that ZnS nanoparticles with an average diameter of about 4.5 nmare distributed uniformly in the
composite ribbons. The authors suggested that TGA served not only to promote a sulfur source reaction
with zinc source to form ZnS nanocrystals but also to create nanocrystals to form I/O composites when
subsided fromthe solution by the absolute alcohol. ZnS/CHA (CHA = cyclohexylamine) nanowires were
obtained via a solvothermal method using Zn(CH
3
COO)
2
2H
2
O and (NH
2
)
2
CS as sources and CHA as a
solvent [173], which yielded uniform and ultralong nanowires with lengths of 520 lm (Fig. 27a).
By changing reaction conditions, their diameters can be tailored from 100 nm to 1 lm. Periodic
2 nm-thick subnanometer-layer structures were clearly identied by HRTEM imaging, the latter were
formed by interlaced arrangement of wurtzite ZnS and amorphous CHA. Furthermore, ZnS/amine
nanocomposities (Fig. 27c) and ZnS(NH
2
CH
2
CH
2
NH
2
)
0.5
lamellas (Fig. 27d) were also synthesized
[75,161]. Table 14 summaries some representative works on 1D ZnS/organic nanostructures.
3.3. 2D nanostructures
ZnS thin lms have been fabricated by several techniques, such as vacuum deposition [283], suc-
cessive ionic layer adsorption and reaction process (SILAR) [284286], metalorganic chemical vapor
Fig. 27. (a) SEM image of ZnS/cyclohexylamine inorganicorganic hybrid semiconductor nanowires. Reproduced from Ref.
[173]. Copyright 2006, American Chemical Society. (b) SEM image of ZnS/NaSCH
2
COONa nanoribbons. Reproduced from Ref.
[281]. Copyright 2007, Institute of Physics. (c) SEM image of ZnS-nanowire/amine nanocomposities. Reproduced from Ref.
[75]. Copyright 2007, Wiley-VCH. (d) SEM image of ZnS(NH
2
CH
2
CH
2
NH
2
)
0.5
lamellas. Reproduced from Ref. [161]. Copyright
2002, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 211
deposition (MOCVD) [287,288], molecular beam epitaxy (MBE) [289,290], spray pyrolysis [291,292],
chemical bath deposition (CBD) [293,294], pulse-laser deposition [295,296], atomic layer epitaxy
(ALE) [297,298], sputtering [299], solgel [300,301], and chemical vapor deposition [302,303]. For
example, Herrero and co-workers synthesized ZnS lms by CBD with a decent growth rate using
NH
2
NH
2
as a complementary complexing agent for the classically used NH
3
reaction baths. The
ZnS thin lms had high transmittance, greater than 75%, for energies lower than the bandgap (E
g
). Zink
and co-workers fabricated hexagonal phase pure ZnS thin lms by laser driven CVD from a single-
source precursor Zn(S
2
COCHMe
2
)
2
under mild conditions (e.g. low temperature). The ZnS lm con-
sisted of small granular particles (-1 lm in diameter), and the grain sizes were about 10 times smaller
than those in the thermally-produced lms. It is of interest that the deposits were obtained only with-
in a selected area (irradiated by the laser beam) on the substrate [303].
4. Luminescence properties of ZnS nanostructures
Luminescence is the generation of light. As pointed out in a recent review [304], light can be emit-
ted via a number of luminescent processes, which include photoluminescence (PL), cathodolumines-
cence (CL), electroluminescence (EL), electrochemiluminescence (ECL), and thermoluminescence (TL).
Table 15 lists typical types of luminescence and their origins. In this section, we present some typical
luminescence properties of ZnS nanostructures.
4.1. Photoluminescence (PL)
ZnS exhibits direct bandgaps of 3.72 eV for the cubic ZB phase and 3.77 eV for the hexagonal WZ
phase with a large exciton energy of 39 meV [305]. The strong exciton binding energy, which is much
larger that that of GaN (25 meV) and the thermal energy at room temperature (26 meV) can ensure an
efcient exciton emission at RT under low excitation energy. Although much research has been de-
voted to the PL properties of 1D ZnS nanostructures, very few studies have reported possible UV band-
gap emission at RT. This is mainly due to high sensitivity of the 1D ZnS nanostructure optical
properties to the synthetic conditions, its crystal size and shape, and intrinsic defects such as vacan-
cies and interstitials.
4.1.1. Visible emission of ZnS nanostructures
As the literature documents, for a large number of 1D ZnS nanostructures of various shapes synthe-
sized using different methods, in general, there are two and more visible characteristic optical peaks.
Table 15
Different types of luminescence [304].
Luminescence Caused by
Photoluminescence (PL) Photo-excitation of compounds
Cathodoluminescence (CL) Electron-excitation of compounds
Electroluminescence (EL) Radiative recombination of electrons and holes in a materials after an electrical current
passes through them or a strong electric eld is applied
Electrochemiluminescence
(ECL)
Electrogenerated chemical excitation
Thermoluminescence (TL) Detrapping process caused by heating or thermostimulation
Table 14
Representative 1D ZnS/organic nanostructures.
Nanostructure Synthesis method T (C) Ref.
ZnS NW/Amine nanocomposities Mild-solution chemistry approach 180 [75]
ZnS (NH
2
C
2
H
4
NH
2
)
0.5
lamellas Solvothermal method 120180 [161]
ZnS/cyclohexylamine hybrid NWs Solvothermal method 100180 [173]
ZnS/NaSCH
2
COONa nanoribbons Hydrothermal method 160 [281]
ZnS/octylamine hybrid nanosheets Hydrothermal route 110180 [282]
212 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
Single-crystalline ZnS nanoawls were fabricated in a large scale by a simple two-step pressure-con-
trolled vapor phase deposition at a relatively low temperature of 800 C using a mixture of commercial
ZnS and SnO
2
, and graphite powders. A molar ratio was 5:1:3 and single-crystalline Si wafers covered
with silver were taken as deposition substrates [275]. As shown in Fig. 28a, awl-holders were rectan-
gle-shaped and faceted, and the diameters of the awl tips were in the range of 100200 nm. A RT PL
spectrum of the synthesized ZnS nanoawls is shown in Fig. 28b. A stable and strong green emission
band centered at 513 nm and a weaker blue emission band centered at 447 nm are seen. Based on
their experimental observations and previous studies, the authors suggested that the blue emission
at 447 nm could be ascribed to the transitions involving vacancy states, while the observable green
emission had been caused by some self-activated centers, vacancy states, or interstitial states associ-
ated with the peculiar nanostructures [275]. ZnS nanobelts were synthesized by a simple thermal
evaporation method in a N
2
atmosphere containing a small amount of CO and H
2
gases [146]. The for-
mation of these nanobelts is a spontaneous growth process through the evaporation of a precursor fol-
lowed by the nucleation and growth. The nanobelts have a width in the range of 40120 nm, a
thickness of 20 nm, and a length of several micrometers (Fig. 28c). They are single-crystals with a hex-
agonal wurtzite structure growing along the [0 0 1] direction. The PL spectrum of ZnS nanobelts pre-
sents two emission bands around 450 nm and 600 nm, as shown in Fig. 28d. The authors suggested
that the emission band around 450 nm had been associated with defect-related emission of the ZnS
host, whereas the strong emission around 600 nm was assigned to the emission from the traces of
Mn
2+
impurity [146].
In one of the pioneering works in the eld of controlled growth of nanostructured materials, ZnS
nanostructures with different morphologies, sizes, and microstructures were formed by the evapora-
tion of ZnS nanopowders at 1100 C for 30 min, with Ar served as both a protecting medium and a
Fig. 28. (a) SEM image and (b) PL spectrum of ZnS nanoawls. Reproduced from Ref. [275]. Copyright 2007, American Chemical
Society. (c) TEM image and (d) PL spectrum of ZnS nanobelts. Reproduced from Ref. [146]. Copyright 2003, American Institute
of Physics.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 213
carrying gas [110]. Fig. 29ad shows the typical SEM images of four kinds of ZnS nanostructures,
including nanorods, nanowires, nanobelts, and nanosheets. All experimental results suggested that
the temperature distribution inside the tube furnace and catalyst played the dominant roles in the
structure formation. Within a certain temperature range, products with a specic morphology were
obtained. Therefore, it may be possible to obtain ZnS nanostructures with a specic morphology by
controlling the reaction temperature and a catalyst type. There exist two growth mechanisms for
ZnS nanostructures, VLS mechanism for ZnS nanorods and nanowires, and VS mechanism for ZnS
nanobelts and nanosheets. The details for the controlled growth and the effects of substrate temper-
ature and catalyst on the growth of ZnS nanostructures can be found in Ref. [110]. Fig. 29e depicts a
typical PL emission of the above-mentioned four kinds of ZnS nanostructures. It can be seen that the
ZnS nanorods and nanowires have almost the same PL band positions. Stable and strong green emis-
sion bands centered at -530 nm and a weaker blue emission band centered at -440 nm are detected.
The ZnS nanobelts and nanosheets have almost the same PL band positions, e.g. a stable and strong
green emission band centered at -545 nm and a weaker blue emission band centered at -440 nm
[110].
4.1.2. UV emission of ZnS nanostructures
ZnS nanowires with rectangular cross-section were fabricated by pulsed laser vaporization (PLV) of
ZnS/10% Au targets in a ow of Ar/5% H
2
[92]. These nanowires were at with abrupt ends and tens of
micrometers in length. Most abundant nanowires had a width of 55 nm, while a few wires were as
wide as 200 nm (Fig. 30a). Furthermore, most of the metallic particles at the growth tip of the nano-
wires were observed to be polyhedrons (solid arrows in Fig. 30a), instead of the usual spheres in the
nanowires grown by PLV. HRTEM results indicated that the ZnS nanowires had been single-crystals
with a wurtzite structure, grown along the [0 0 1] or [1 0 0] direction. As shown in Fig. 30b, a strong
Fig. 29. (ad) SEM images of ZnS nanorods, nanowires, nanobelts and nanosheets formed in four different temperature zones,
respectively. (e) PL spectra of ZnS nanostructures. Reproduced from Ref. [110]. Copyright 2005, Wiley-VCH.
214 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
luminescence doublet (3.68 and 3.75 eV) and weaker defect luminescence structure at 2.44, 2.66, 2.86,
3.06 eV were observed for ZnS nanowires. The authors suggested that the 2.44 eV (-510 nm) emission
band might be due to Au impurities, the other three weak emission bands were probably due to the
stoichiometric vacancies or interstitial impurities, and 3.28 and 3.75 eV emission bands were from the
direct transitions between the conduction band and the spinorbit/crystal eld splitting of the valance
bands [92]. ZnS nanowires with periodically alternating twins along the growth direction of wires
were fabricated by thermal evaporation of ZnS nanopowders at 1100 C [256]. The ZnS nanowires
had a length of several to tens of micrometers, and a diameter of 5060 nm. These nanowires were
randomly oriented, and most of them were rather straight (Fig. 30c).TEM and HRTEM images demon-
strated that alternating light/dark contrast appeared in a periodic manner along the axial direction of
the wires. These resulted from alternating twins along the hexagonal ZnS 1 1 0) (cubic 1 1 1)) crys-
talline direction. Fig. 30d shows the PL spectra of the nanopowders and the periodically twinned ZnS
nanowires at RT. A strong UV peak (329 nm) and a weak blue emission are observed in PL emission.
The authors suggested that the weak emission band was probably due to stoichiometric vacancies,
whereas the strong emission band was attributed to the band edge photoluminescence of ZnS
[256]. Table 16 contains a brief summary on PL properties of ZnS nanostructures with various
morphologies.
4.2. Cathodoluminescence (CL)
Cathodoluminescence (CL) is an optoelectrical phenomenon wherein a beam of electrons is gener-
ated by an electron gun and is interacting with a luminescent material, such as phosphor, causing the
Fig. 30. (a) TEM and HRTEM images of rectangular cross-sectional ZnS nanowires. (b) PL spectrum of ZnS nanowires taken at RT
excited by a 266 nm UV laser. Two strong emission bands near band edge and four weak defect luminescence bands were
observed. Blue curves are Lorentzian line shape analyses after a least-squares multi-Lorentzian (red curves). Reproduced from
Ref. [92]. Copyright 2004, American Chemical Society. (c) SEM and TEM images of periodically twinned ZnS nanowires. (d) RT
PL spectra of the ZnS nanopowders and periodically twinned ZnS nanowires. Reproduced from Ref. [256]. Copyright 2006,
American Institute of Physics.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 215
material to emit light. It is a useful technique for characterization of nanostructure optical properties
because of its high spatial resolution and structural information obtained by using secondary electron
(SE) imaging [306,307].
4.2.1. Multi-angular branched ZnS nanostructures with needle-shaped tips
A facile and effective route toward the synthesis of ZnS multi-angular branched nanostructures
with needle-shaped tips has been successfully demonstrated [276]. The ZnS nanostructures were com-
posed of multi-angular branched nanostructures with needle-shaped tips. A diameter of each needle-
shaped branch decreased along its length from the center towards the tip. Each arm had straight
appearance with the needle length reached several micrometers (Fig. 31a). HRTEM image (Fig. 31b)
suggests that the surface of the entire branch is clean and atomically sharp. It is free of dislocations
or stacking faults. Each arm grew along the same direction, which was [0 0 0 1] [276]. High spatial res-
olution CL from individual ZnS multi-angular branched nanostructures was investigated with the
nanometer resolution. The size-dependent optical spectra exhibited sharp UV bandgap emission
and broad visible emission at RT. Fig. 31c and d show a typical individual ZnS multi-angular branched
structure and its CL spectrum obtained at RT. This consists of a sharp UV emission band centered at
-334 nm and a broad visible emission band centered at -404 nm. The inset in Fig. 31d is the enlarged
portion of the sharp UV emission band. The intensity ratio between the UV emission and visible emis-
sion is -1/10. In order to further study spatial variations of the optical properties, the CL spectra were
collected in different spots along the individual tapered arms of varying diameter, as depicted in
Fig. 31e and f. Fig. 31e displays a CL image of another ZnS nanostructure, showing uniform lumines-
cence that entirely arises from the nanostructure alone. The RT CL spectra (Fig. 31f) denoted as C1, C2,
C3, C6, C9, C12, C15, C16, and C17 were acquired at various spots between the two red lines in Fig. 31e.
The CL UV emissions are very strong and sharp in some spots. The centers of the UV emissions shifted
gradually from 337 nm to 333 nm (-0.043 eV blue-shift) with diameter descending, while the broad
visible emissions changed between 398 nm and 425 nm. The CL UV emission from the samples was
Table 16
PL properties of ZnS nanostructures.
a
NSs NBE
emission/nm
DL emission/nm Ref. NSs NBE emission/
nm
DL emission/nm Ref.
NTs 439; 538 [114] NTs 510 [115]
MTs 550; 720 [121] NWs 329 410 [156]
NWs 338.2 [95] NWs 376 464; 505; 571 [83]
NWs 350 397 [72] NWs 450; 520 [65]
NWs 333 [94] NWs 490 [87]
NWs 399; 417; 441; 458; 513 [88] NWs 432440; 563569 [74]
NWs 554; 802 [76] NWs 413; 438 [69]
NWs 421; 530 [97] NWs 339 [96]
NWs 3.75 eV;
3.68 eV
2.44 eV; 2.66 eV; 2.86 eV;
3.06 eV
[82] NWs 394; 458 [100]
NWs 515; 580 [86] NWs 366 467; 515 [79]
NWs 366 467; 515 [79] NWs 450 [138]
NRs 420 [102] NRs 440; 530 [110]
NRs 335 430 [109] NRs 349 385; 427; 489 [103]
NRs 440 [112] NBs 450; 530 [139]
NBs 450; 600 [146] NBs 535 [147]
NBs 513 [136] NBs 439; 517 [129]
NBs 420; 520 [131] NBs 522 [111]
NRs 394; 456 [151] NRs 532 [151]
NRs 414.5; 534.5 [124] NRs 394; 456 [101]
NRs 509 [154] NSs 377; 518.2 [158]
NSs 440; 545 [159] NAs 447; 513 [275]
NSs 424 [157] NTPs 495.2 [249]
a
NSs = nanostructures; NTs = nanotubes; MTs = microtubes; NWs = nanowires; NRs = nanorods; NBs: nanobelts;
NSs = nanosheets; NAs = nanoawls; NTPs = nanotetrapods.
216 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
very stable against a prolonged exposure to air and from one measurement to another. The CL spectra
remained unchanged being consecutively recorded within a period of nearly one year. It is assumed
that the CL UV emission centered at -333337 nm represents the band-edge luminescence, suggest-
ing the band-gap energy of -3.7 eV for the present wurtzite-type ZnS nanostructures [276].
4.2.2. Single-crystalline ZnS nanobelts
Single-crystalline ZnS nanobelts possessing sharp UV-light emission were synthesized via selection
of source materials and by controlling their evaporation and agglomeration rates [125]. The ZnS nano-
belts grew from several separated nucleation sites that seemed to be connected. They had the typical
Fig. 31. (a) TEM image and (b) HRTEM image of multi-angular branched ZnS nanostructures with needle-shaped tips. (c) SEM
image and (d) corresponding CL spectrum of a ZnS multi-angular branched nanostructure. (e) CL image of another ZnS
nanostructure. (f) CL spectra recorded from selective nine spots along the individual arm in (e). Reproduced from Ref. [276].
Copyright 2008, American Chemical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 217
widths in the range of 200 nm to 1 lm and the length could be up to a millimeter. The highly crystal-
line nature of the nanobelts was further veried by an enlarged lattice-resolved HRTEM image and ED
pattern, as shown in Fig. 34b and c. The marked interplanar d-spacings of 0.63 and 0.33 nm correspond
to the (0 0 0 1) and (0 1 1 0) lattice planes of a wurtzite ZnS. Fig. 34d illustrates a model of the nano-
belts and their enclosing facets. Most of the belts take the [0 1 1 0] orientation as the growth direc-
tion and the (2 1 1 0) and (0 0 0 1) planes as the top/bottom and side surfaces, respectively.
By dispersing individual ZnS nanobelts on standard C-coated TEM copper grids, CL properties from
individual ZnS nanobelts were investigated with a nanometer scale resolution. Fig. 32e and f display a
typical SEM image and the corresponding CL image of a ZnS nanobelt. The CL spectrum shows fairly
uniform luminescence that arises from the nanobelt alone. CL spectrum acquired from an individual
ZnS nanobelt (seen in the SEM image in Fig. 32e) is shown in Fig. 32g. It is composed of a narrow and
strong UV peak centered on -337 nm and broad low intense luminescence in the visible region
Fig. 32. (a) TEM image, (b) HRTEM image, (c) SAED pattern of ZnS nanobelts; (d) Structure model of individual nanobelts and
their enclosing facets; (e) SEM image of a ZnS nanobelt and (f) its corresponding CL image; (g) CL spectrum recorded from the
ZnS nanobelt seen in the SEM image in (e); (h) CL spectra acquired at various spots on the nanobelt along the line marked in (e).
Reproduced from Ref. [125]. Copyright 2009, Wiley-VCH.
218 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
(-550 nm). Fig. 32h plots a number of CL spectra, acquired at various spots along the line shown in
Fig. 32e. The shapes of CL spectra are fairly uniform, in which all the peaks are located at the same
positions, while the intensities of the UV peaks are slightly decreased [125]. Table 17 summarizes
the latest achievements in regard of CL properties of ZnS nanostructures.
4.3. Electroluminescence (EL)
Amongst various useful properties of ZnS, EL deserves a special mention, because ZnS is considered
to be one of the best semiconducting functional materials for EL devices. EL is a phenomenon in which
a material emits light in response to an electric current passed through it. This is one of the few in-
stances in which a direct conversion of electric energy into visible light takes place without the gen-
eration of heat (incandescence), chemical reaction (chemiluminescence), or a mechanical action
(mechanoluminescence). The phenomenon was discovered in 1936 by G. Destriau under the observa-
tions that when a ZnS (later found to be slightly doped with Cu) powder was suspended in an insulator
and an intense alternating electric eld was applied with capacitor like electrodes, visible light was
emitted. This has led to a surge in research activities of EL properties of ZnS, which were mostly under-
taken on single-crystals and powder samples [308312]. The light output of thin-lm electrolumines-
cent displays has been very reliable, with little loss after tens of thousands of hours of operation.
The fabrication of an EL device requires several components and the simplest device would require
establishing two electrical contacts to a doped ZnS phosphor layer. ZnS acts as a host for dopant mol-
ecules such as Mg, Cu, Tb and Ag. which are functional centers responsible for emission of light. There-
fore, the electron transport in the device is a complex process associated with charge carrier tunneling
across the electrodesemiconductor interface with subsequent impact excitation of the dopant atoms
and avalanche multiplication of carriers due to impact ionization of intrinsic defects and impurities.
The key parameter determining the luminescence efciency is impact excitation cross section and dis-
tribution of the dopant over the phosphor layer, effective thickness and probability of radiative tran-
sition of dopants. While detailed discussion of these factors is out of the scope of the reviewand can be
found elsewhere [308,313315], it is clear that successful fabrication of a device using ZnS nanostruc-
tures would require systematic efforts and stepwise optimization procedure.
EL in ZnS is largely inuenced by external factors such as temperature and magnetic eld. The EL
brightness increases with increasing temperature due to temperature dependence of the generation of
primary electrons indispensable for luminous center excitation [311]. While the effect of temperature
is well understood and expected, what is more interesting is that exposure to magnetic eld of a ZnS
crystal is found to enhance the EL brightness remarkably [312]. Fig. 33 shows the dependence of EL
intensity of a ZnS crystal doped with 10 ppm Cu on the voltage applied before and after exposure
to the magnetic eld. The authors proposed that the magnetic eld facilitates relaxation of a metasta-
ble state of the structural defects thereby causing this enhancement.
The deep investigation on EL properties of bulk ZnS crystals obviated that colloidal ZnS based semi-
conductor nanocrystals are promising luminophores for a new generation of EL devices. In recent
years, research on semiconductor nanocrystal based light-emitting diodes (LEDs) has made
Table 17
CL properties of ZnS nanostructures.
Nanostructure NBE emission /nm DL emission/nm Ref.
Multi-angular branched ZnS nanostructures 344 404 [276]
ZnS NWs 380 450; 610 [85]
ZnS tetrapod tree-like heterostructures 336; 323 [272]
ZnS hexagonal pyramids 337 700 [274]
ZnS nanowires 340 530 [70]
ZnS nanobelts 337 550 [125]
ZnS nanobelts 2.3 eV; 2.73 eV [143]
ZnS nanosheets 520; 680 [160]
ZnS nanowires 360 510; 580 [257]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 219
remarkable advances. The external quantum efciency has greatly improved by two orders of magni-
tude and highly saturated color emission is being obtained [316]. Thus, alternating current thin-lm
electroluminescent (AC-TFEL) devices are already popular for high resolution, at panel displays.
These displays are robust, having long lifetimes, and offer high output efciency with relatively low
power consumption [317]. AC-TFEL devices consist of a phosphor layer, such as doped zinc sulde
(ZnS:Mn), sandwiched between two insulators. When a sufciently high voltage is applied across
the electrodes, electrons (trapped at interfaces between the layers) are injected into the conduction
band of the phosphor, where they are accelerated by the eld and can excite the luminescent dopant
centers in the phosphor layer via impact excitation and ionization mechanisms [318,319].
Fig. 34 shows a typical structure of an EL device composed of a glass substrate ZnO:Al transparent
columnar electrodes, a Si
3
N
4
insulating layer, a ZnS:Mn NC emission layer, and Al columnar electrodes
[320]. The ZnS:Mn NCs were obtained by a co-precipitation method. The molar ratio of [Zn
2+
]:[Mn
2+
]
was set at 1:0.1. The ZnS:Mn NC powder was dispersed into a mixture of resin and organic solvent,
yielding a printing ink. A layer of this ink was printed on the Si
3
N
4
layer, and the printed layer was
dried at a temperature of 130 C.
Fig. 33. Dependence of the EL intensity of a Cu-doped ZnS crystal on the active voltage U, (m = 5800 Hz): (1) before exposure to a
magnetic eld; (2) 20 min exposure to a magnetic eld (B52 T); and (3) 24 h exposure to the same magnetic eld. Reproduced
from Ref. [312]. Copyright 1999, Springer.
Fig. 34. Schematic showing the construction of an EL device with printed emission layer containing ZnS:Mn NCs. Reproduced
from Ref. [320]. Copyright 2009, Institute of Physics.
220 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
The EL devices were operated at 5 kHz sinusoidal ac voltage. Fig. 35a shows two typical EL spectra
of the ZnS:Mn NCs prepared using different precursor concentrations. Both spectra are identical. The
emission centered at 595 nm can be ascribed to the
4
T
1
?
6
A
1
transition of Mn
2+
ions. It was found
that the EL peak wavelength had remained identical regardless of the operating voltage. Fig. 35b
shows a typical luminescencevoltage (LV) characteristic of the EL device. Above the threshold volt-
age of EL emission, V
th
(dened as the voltage at the luminescence of 0.05 cd m
2
) of 105 V, the lumi-
nescence rises steeply. The characteristic luminescence, L
45
, is dened as the luminescence at a voltage
of V
th
+ 45V. Importantly, this was found to decrease with increase in particle size.
In order to obtain higher luminescence, it is crucial to passivate the ZnS nanocrystal (NC) surface
effectively. This is because the dominant EL excitation mechanism for the doped ZnS NCs is the
hot-electron-induced direct impact excitation [321]. Therefore, the electronphonon interaction of
ZnS NCs is important in gaining the electron energy from a high electric eld. Chan and co-workers
argued that a much weaker phonon coupling strength is expected in the well-passivated NCs than
in bulk ZnS:Mn [322]. The weak phonon coupling strength weakens the electronphonon interaction,
resulting in increased energy of hot electrons or increased luminescence.
During the last two decades, intense research has been carried out in order to fabricate improved
and efcient EL devices by using ZnS based nanocrystals. This has led to remarkable improvement.
And in 2009, the rst commercial transparent AC-TFEL displays have been demonstrated by Sharp,
Inc [323332].
Even though fabrication of AC-TFEL devices have been widely investigated, there still exists many
challenges in order to make viable EL based devices. For example, the development of multicolor dis-
plays with balanced red, green, and blue (RGB) emission has remained difcult as the materials for red,
green, and blue phosphors require different processing conditions. Additionally, the luminous efcien-
cies of different phosphors may vary by an order of magnitude, complicating electronic driving of bal-
anced color displays [333]. Many quantum dot nanocrystals are poor in conduction properties, due to
the presence of an organic capping layer that is necessary for quality size control and size distribution
[334]. Moreover capping agents may decompose with time. Thus another challenge is to stabilize ef-
cient nanocrystals in solution without the use of capping agents.
In a recent breakthrough experiment, Bulovic and co-workers presented a novel technique for
room temperature, solution-based fabrication of AC-TFEL devices using ZnSe/ZnS:Mn/ZnS nanocrys-
tals that exhibited a quantum yield of 65% [323]. Both the nanocrystal and the ceramic lms embrac-
ing them have minimal absorption across the visible light spectrum, enabling them to demonstrate
this transparent AC-TFEL device.
Fig. 35. (a) Normalized EL spectra of the ZnS:Mn NC lms obtained using two different Zn
2+
precursor concentration. (b) LV
characteristic of the EL device. The threshold voltage (V
th
) as well as characteristic luminescence (L45) are indicated here.
Reproduced from Ref. [320]. Copyright 2009, Institute of Physics.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 221
These workers used visible transparent core shell ZnSe/ZnS:Mn/ZnS nanocrystals obtained by a
solution technique (Fig. 36a). The luminescence from these nanocrystals (centered at 590 nm) is char-
acteristic of the Mn
4
T
1
?
6
A
1
phosphor transition. The schematic diagram of the AC-TFEL device, fab-
ricated using radio frequency magnetron sputtering and spin casting (or coating?) method, is seen in
Fig. 36b. First, at, 80 nm of insulating Al
2
O
3
or HfO
2
was sputtered onto an indium tin oxide (ITO)
coated glass. The active layer of the device was fabricated with alternating layers of sputtered ZnS
(12 nm thick) and spin-cast ZnSe/ZnS:Mn/ZnS nanocrystals (30 nm thick). It was found that the active
region of the device had been the interface between the sputtered ZnS lm and nanocrystal, rather
than the stand alone nanocrystal. Besides, the number of layers and their thickness played crucial roles
in the device performance. The detailed characteristics of the device are described in Fig. 37. Impor-
tantly, the PL spectra of the nanocrystals and the EL spectra of the device overlaped (Fig. 37d) sugges-
tive that the Mg dopant is responsible for the EL response.
The another example is a single-layer structure EL device consisting of an indium tin oxide (ITO)
anode, a ZnS nanocrystalspolymer composite (ZnSP) emitter layer and an Al cathode, as shown in
the inset of Fig. 38a. The ZnSP lm was spin-coated onto an ITO-coated glass substrate, and the thick-
ness of the ZnSP layer was 100 nm. Fig. 38a presents the PL and EL spectra of the ZnSP thin lm. The PL
peaks at 415 nm, while the EL reaches a maximum at 435 nm which is red shifted compared to the PL.
The current versus voltage (IV) characteristics are shown in Fig. 38b. The turn-on voltage is <4 V.
Ramping the applied bias, blue light becomes visible just below 5 V forward bias (no light is observed
under reverse bias). Furthermore, there is a little difference between the EL spectra from the fresh
ZnSP and ZnSP aged in the dark at RT for one month, indicating that ZnS NCs prepared in polymer
matrices possess good stability [335].
Although several EL devices have been demonstrated, there are still several challenges that need to
be overcome before these can be put into practical applications. For example, the reported efciencies
are still more than an order of magnitude lower than those of the purely organic LEDs. However, there
are several potential advantages associated with nanocrystal-based devices, such as a spectrally pure
emission color, and stability and robustness of the nanocrystals. Further developments of nanocrystal-
based LEDs will have to be in understanding and controlling the chemical and physical phenomena
across the interfaces, and optimizing carrier mobility. EL properties of bulk ZnS has been studied in
detail. This vast knowledge is now required to be used in developing the new generation of ZnS nano-
structure based EL devices, because nanostructures promise more efciency due to improved crystal-
linity and adjustable surface compositions. In addition to these, it will also require wider efforts to
develop multicolor displays where not only the red, green, and blue emissions exist but also these
are in balance with each other.
Fig. 36. (a) Photographs of ZnSe/ZnS:Mn/ZnS nanocrystals in chloroform in room light (left) and under UV illumination (right)
indicating that these orange-emitting nanocrystals exhibit no absorption in the visible wavelength regime. (b) A schematic of
the doped nanocrystal-based AC-TFEL device structure fabricated by Bulovic and co-workers. Reproduced from Ref. [323].
Copyright 2009, American Chemical Society.
222 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
Fig. 37. (a) Plot showing the EL intensity obtained by using different number of layers of ZnS in the device. At 110 Vpp and
30 kHz, devices with four layers of ZnS give the largest EL response. (b) A photograph of a glass substrate containing 10
unconnected AC-TFEL devices showing near transparency. The inset shows the uniformity of illuminated pixel in the dark,
operated at 170 Vpp and 30 kHz. (c) Plots of the EL response (normalized to the EL response at 170 Vpp and 30 kHz) as a
function of drive voltage and frequency. (d) The EL spectra for devices with Al
2
O
3
and H
f
O
2
insulating layers are presented by
the solid and dashed orange curves, respectively. A photoluminescence spectrum of an incomplete device (ITO/Al
2
O
3
/ZnS/30 nm
nanocrystals) (solid black curve) is compared with that of NCs stabilized in chloroform (solid gray curve). The photolumi-
nescence spectrum of the device where ZnS was sputtered onto the above mentioned device is shown with the dashed black
line. Reproduced from Ref. [323]. Copyright 2009, American Chemical Society.
Fig. 38. (a) PL and EL spectra of a ZnSP thin lm. Curves 1 corresponds to PL spectrum of ZnSP thin lms, 2 and 3 correspond to
EL spectra of ZnSP single-layer structure EL device aged for 1 h and one month, respectively. (b) Currentvoltage characteristics
of the ITO-ZnSP-Al device. Reproduced from Ref. [335]. Copyright 1997, Royal Society of Chemistry.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 223
4.4. Electrochemiluminescence (ECL)
Electrochemiluminescence (ECL) involves the generation of species at electrode surfaces that then
undergo electron-transfer reactions to form excited states that emit light [336339]. Zhang and co-
workers observed a bandgap ECL of ZnS nanoparticles in alkaline aqueous solution at a platinum elec-
trode under the potential between 2.0 V (versus Ag/AgCl, saturated KCl) and +0.86 V [336]. The ECL
peak of ZnS NPs in 0.1 M NaOH aqueous solution appeared at +0.86 V, and the ECL peak wavelength of
ZnS NPs was 460 nm (Fig. 39a) It was found that the surface passivation effect and the core/shell struc-
ture of ZnS/Zn(OH)
2
played a signicant role in the ECL process and that the similarity of the ECL and
PL spectra of semiconductor NPs was dependent on the extent of the surface passivation (schematic
representations of the PL and ECL process of nanoparticles in the core and on the surface are shown
in Fig. 39b). The ECL intensity of ZnS NPs in NaOH aqueous solution was greatly enhanced by an addi-
tion of K
2
S
2
O
8
(Fig. 39d) and the ECL intensity linearly increased along with the concentration of ZnS
NPs in the range from 1.2 10
4
to 1.0 10
3
M. More detailed discussion may be found in Ref. [336].
4.5. Thermoluminescence (TL)
Whenever a semiconductor is irradiated, electrons and holes are created. If electron hole pairs
recombine immediately, they emit a photon, which is known as uorescence; if the electrons and
holes created do not recombine rapidly, but are trapped in some metastable states and separated, they
Fig. 39. (a) Absorption spectrum (A), PL spectrum (B) and the ECL spectrum (C) in the ZnS NPs in NaOH. (b) Schematic
representations of PL and ECL process of NP in the core and on the surface. (c and d) ECL potential curve of ZnS NPs in 0.10 M
NaOH with absence and presence of 0.1 M K
2
S
2
O
8
, respectively. Reproduced from Ref. [336]. Copyright 2007, American
Chemical Society.
224 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
need energy to be released from the traps and recombine to give luminescence. If the detrapping pro-
cess is caused by heating or thermostimulation, the luminescence is called thermoluminescence (TL).
Thermoluminescence is a good way to detect the recombination emission caused by thermal detrap-
ping of carriers [340,341].
Chen and co-workers rst reported the TL of a ZnS nanoparticle [340]. As shown in Ref. [340], the
ZnS nanoparticles were prepared using Zn(NO
3
)
2
and Na
2
S as sources, and these nanoparticles were
deposited on a quartz substrate for measurements. The average sizes of ZnS nanoparticles prepared
at RT, 50, 100, and 200 C, estimating from the Debey-Scherrer formula and using XRD results, were
1.81, 2.50, 2.74 and 3.01 nm, respectively. Fig. 40a shows the luminescence spectra of ZnS nanoparti-
cles, demonstrating that luminescence intensity of trapped emission increases and the emission is
blue-shifted as the size decreases. Fig. 40b depicts the glow curves of the nanoparticles. An obvious
glow peak around 360 K is observed. All samples show the glow peaks at almost the same position
and the TL intensity is consistent with that of the surface uorescence [340].
The thermoluminescence intensity of nanoparticles is given by a simple formula, I = dm=dt =
mnA, where m and n are the density of holes and electrons for recombination, respectively; and A is
the carrier recombination probability. Now, with increase in the content of the surface states, holes
and electrons of the particle become more accessible for the TL recombination, i.e., the m and n in-
creases proportionally with the surface states. As the surface states increases with decrease in the size
of the particle, nanoparticles with smaller size leads to enhancing TL efciency. Furthermore, the wave
functions of electrons and holes are effectively overlapped in nanoparticles, which also result in the
increase of their recombination probability, A. Due to these two factors, the TL of small particles is ex-
pected to be much stronger than that of bulk. Fig. 40d schematically depicts the size dependence of the
surface states. The separation between the electronhole states (similar to the donoracceptor pairs)
increases with the decreasing size because the trap depth does not change much upon decreasing size,
while the bandgap increases [340].
4.6. Luminescence mechanism
4.6.1. Origin of the green photoluminescence from ZnS nanobelts [148]
ZnS nanobelts were obtained by a thermal evaporation method with the assistance of H
2
S in an Ar
atmosphere by Ye and co-workers [147]. These nanobelts had a width ranging from tens of nanome-
ters to hundreds of micrometers (Fig. 41a). The EDS spectra showed that S and Zn were in an atomic
ratio of 52.6047.40% (Fig. 41b), and the X-ray photoelectron spectroscopy (XPS) results indicated that
elemental S species were abundant on the surface of the ZnS nanobelts. Fig. 41c shows the PL spec-
trum of the nanobelts with an excitation wavelength of 325 nm, where an intense green PL band cen-
tered at about 535 nm is clearly observed with no PL peak at around 420 nm. The photoluminescence
excitation (PLE) (curve 1, dotted) exhibits two peaks at 340 nm (corresponding to the bandgap of ZnS)
and 400 nm (corresponding to defect energy levels), respectively, indicating that there is at least an-
other defect states other than S vacancies in the bandgap of ZnS nanobelts that can contribute to the
green PL band. As known, UV light may initiate the photo-oxidation process on the surface of ZnS
Fig. 40. (a) Luminescence spectra and (b) glow curves of ZnS nanoparticles prepared at RT, 50, 100, and 200 C, respectively. (c)
A schematic model for the size dependence of surface states in semiconductor nanoparticles. Reproduced from Ref. [340].
Copyright 1997, American Institute of Physics.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 225
nanobelts and transform elemental S species to sulfur dioxides [342], henceforth, changing the surface
structure of the nanobelts. Meanwhile, the intensity of the PL peak at 535 nm as well as the PLE peaks
decrease greatly after the UV light 30-min illumination, indicating that elemental S species on the sur-
face of the ZnS nanobelts do contribute to the green PL band. Furthermore, the short-nanosecond scale
makes this PL band very efcient against the self-activated blue band, and also precludes the operation
of Cu and Al co-doped emission (Fig. 41d). These results further prove that the origin of 535-nm emis-
sion is related to elemental sulfur species on the surface of the ZnS nanobelts. More detailed discus-
sion can be found in Ref. [147].
4.6.2. Temperature-dependent PL from elemental sulfur species on ZnS nanobelts [345]
Ye and co-workers not only proved that the origin of the green PL band at 535 nm was related to
elemental S species on the surface of the undoped ZnS nanobelts, but also investigated temperature-
dependent PL from elemental S species on the belts. This allowed the authors to clarify on the carrier
recombination mechanism in them [343]. PL and PLE spectra in the temperature range of 10280 K are
displayed in Fig. 42a and b, in which there are two peaks in the blue (B band) and green (G) wave-
length ranges, respectively. These two bands are due to the recombination of electrons from conduc-
tion band and sulfur vacancies to holes on surface species, respectively. Fig. 42c displays an Arrhenius
plot of the integrated PL intensity I(T) over the temperature of 10280 K for the G band and B band,
respectively. The experimental data were tted with a formula [344];
I(T) =
I
0
1 Aexp(E
a
=k
B
T)
(7)
where I
0
is the integrated intensity at zero absolute temperature and A is a constant. Activation ener-
gies of E
a
= 65 1 and 157 5 meV were obtained from tting for the G and B bands, respectively. The
Fig. 41. (a) SEM image and (b) EDS spectrum of ZnS nanobelts. (c) PL and PLE spectra of the ZnS nanobelts. Curves (1) and (2)
correspond to PL recorded immediately and after 30 min UV-light illumination of the ZnS nanobelts (solid curve: PL spectrum,
dotted curve: PLE spectrum). (d) The time decay of the light emission at 535 nm from ZnS nanobelts. The exponential decay can
be separated into a longer part of 7.46 ns and a shorter part of 860 ps. Reproduced from Ref. [147]. Copyright 2004, American
Institute of Physics.
226 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
apparent difference in activation energy indicates that the non radiative channels for these two bands
are also different. The activation energy of the B band is close to the strong localized state in the unsat-
urated orbital of surface S species, as reported by Gibbons and Spear [345]. Thus the non radiative pro-
cess for the B band may correspond mainly to the complete delocalization of the electrons that are
captured by non radiative channels later. The activation energy of the G band is much smaller than
the delocalization energy. Thus the processmay be ascribed to a potential barrier for the electrons
escaping from the sulfur vacancy donor sites to other energy levels and giving off the energy by
non radiative process. Direct excitation of the electrons from the sulfur acceptor levels is efcient
at high temperatures because at a low temperature holes are attracted at this level and the electron
population is low; with an increase in temperature, the holes transfer to other defect levels or ther-
malize to the valence band [6,343].
Fig. 42d shows the temperature dependence of the PL linewidth and peak energy. These demon-
strate that as the temperature increases from 10 to 100 K, the value of the full width at half maximum
(FWHM) decreases from 35.4 to 31.3 meV, and as the temperature further increases, the FWHM value
increases. This temperature dependence can be well explained by the transfer and the thermalization
Fig. 42. (a) PL and (b) PLE spectra of S-rich ZnS nanobelts from 10 to 280 K. (c) Integrated intensity as a function of temperature
for the G band (left column) and the B band (right column), respectively. Squares are experimental data, and curves are the
tting results in line with the formula (1). (d) Peak position as a function of temperature for the G band (left column) and N
band (right column). (e) Diagram of the energy levels responsible for the G band and the B band in the S-rich ZnS nanobelts.
Reproduced from Ref. [343]. Copyright 2006, American Institute of Physics.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 227
of localized states. With an increase in temperature and enhanced thermalization, some localized exci-
tons become mobile and the emission becomes dominated by the recombination of delocalized states
so that the carrier distribution and thus the linewidth narrows. With a further increase in tempera-
ture, thermal broadening of the carrier distribution occurs and the linewidth increases accordingly.
The FWHM of the excitons due to the excitonphoton interaction is given by an equation:
C(T) = C(0) AT
C
ph
exp(E
op
=k
B
T) 1
(8)
where the rst term represents the broadening invoked from temperature-independent mechanism,
the second term is the contribution from acoustic phonon scattering, the last term is caused by the
excitonLO phonon interaction, C
ph
represents the strength of the excitonLO phonon coupling,
and E
op
is the average LO phonon energy. The tting of the FWHM with formula (2), as shown in
Fig. 42d, indicates that the t is good only when setting the coefcient of acoustic phonons to a neg-
ative value: A = 0.23 0.03 meV/K for the G band and 0.8 0.1 meV/K for the B band. The negative
coefcient of the acoustic photons has not been well understood and may imply that the acoustic pho-
tons contribute to the narrowing of the FWHM at a low temperature through the delocalization pro-
cess. Finally, the authors drew the diagram of the energy levels and recombination processes
responsible for the emission bands for the S-rich ZnS nanobelts. More detailed discussion can be found
in Ref. [343].
4.7. Optical property tuning by doping
4.7.1. Optical properties of Mn-doped ZnS nanobelts
Mn-doped ZnS nanobelts were fabricated through a thermal evaporation method starting with a
mixture of acetylacetonates (Zn(acac)
2
and Mn(acac)
2
) and H
2
S at 900 C [226]. The typical widths
of the nanobelts were in a range of 40200 nm and the lengths reached several tens of micrometers.
The TEM results reveal that the ZnS nanobelts are structurally uniform and possess single-crystal hex-
agonal wurtzite structure, growing along the [0 0 1] direction without defects or dislocations. The re-
sults of EDS carried out on several samples and on different regions of the nanobelts showed the molar
fraction of the elemental Mn close to the expected composition. The RT PL spectra taken from bulk
quantities of undoped and Mn-doped ZnS nanobelts under the same experimental conditions showed
that the emission band around 450 nm is attributed to defect-related emission of the ZnS host [55],
whereas the orange band around 580 nm is associated with the Mn
2+ 4
T
1
6A
1
transition due to Mn
doping [346]. The luminescence centers of Mn
2+
may trap electrons and holes, thus the uorescence
efciencies are increased and the glow peaks shift to longer wavelengths. The result also shows that
the defect-related emission of the ZnS nanobelts decreases as the Mn-doped ration increases. This can-
not be often observed in the highly doped cases, because the ZnS phase acts mainly as a host in the
doped case. It absorbs energy and transfers the excited electrons to the dopant. This causes the emis-
sion with various colors [226].
4.7.2. Optical properties of Cu-doped ZnS nanorods
Cu-doped ZnS nanorods were prepared by a solvothermal technique using Zn-nitrate and Cu-ace-
tate as sources and ethylenediamine and water as the solvent. With a gradual increase in the Cu con-
centration, a phase transformation of the doped ZnS nanorods from wurtzite to cubic phase was
observed. All the samples showed luminescence within UV and near IR range with peaks at 370,
492498, 565 and 730 nm (Fig. 43a). The authors suggested that the UV region peak was due to the
near-band-edge transition, whereas, the green peak was related to the emission from elementary S
species on the nanorod surfaces. The orange emission at 565 nm may be linked to the recombination
of the electrons at deep defect levels at the Cu (t
2
) states present near the valence band of ZnS. The
near IR emission possibly originated from transitions due to deep-level defects. The schematic energy
level diagram of the different transitions occurring in Cu-doped ZnS nanorods is shown in Fig. 43b
[220].
228 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
Fig. 43. (a) RT PL spectra of the undoped and Cu-doped ZnS nanorods. (b) The schematic energy level diagram of the nanorods.
Reproduced from Ref. [220]. Copyright 2008, Elsevier.
Fig. 44. (a) PL spectrum of the Mn and Cd co-doped-ZnS nanostructures. Black dotted curves are Gaussian line shape analyzed
plots after a multi-Gaussian t (black solid curve). (b) Schematic energy level diagram for the Mn and Cd co-doped ZnS. The
transition from Mn
2+
4T
1
state to the Cd-related De state results in a red light emission. Reproduced from Ref. [223]. Copyright
2006, Institute of Physics.
Table 18
Optical properties of 1D doped-ZnS nanostructures.
Nanostructures PL peaks Ref.
Mn/Cu-co-doped, Mn/Cu/Fe-
co-doped 1D ZnS NSs
580 nm: Mn
2+ 4
T
1

6
A
1
transition; 520 nm: Cu
2+
CB(ZnS)-t
2
(Cu
2+
(d
9
))
transition; 622 nm: Cu impurities; 450 nm: trapped luminescence from the
surface states; 375 nm: hole traps of unsaturated sp
3
orbitals of the surface S
atoms
[347]
Eu-ZnS NWs 520 nm: Eu
2+
4f
6
5d4f
7
intra-ion transition [224]
Mn-ZnS NBs 580 nm: Mn
2+ 4
T
1

6
A
1
transition; 450 nm: defect-related of ZnS host [226]
Mn/Cd-co-ZnS 1D NSs 580 nm: Mn
2+ 4
T
1

6
A
1
dd transition; 512 nm: E (splitting energy at CB)-De
(Cd vacancies) transition; 636 nm: Mn
2+ 4
T
1
Cd De transition; 440 nm: S
vacancies; 375 nm: hole traps of unsaturated sp
3
orbitals of the surface S atoms
[223]
Mn-ZnS Nanoclusters 580 nm: Mn
2+ 4
T
1

6
A
1
transition [346]
Mn-ZnS NCs 585 nm: Mn
2+ 4
T
1

6
A
1
transition [348]
Cu-ZnS NRs 565 nm: CB DL (deep localized donor level near CB)-t
2
(Cu
2+
(d
9
)) transition;
500 nm: surface S elements; 730 nm: deep trap states in the ZnS lattice
[220]
Ni-ZnS NCs 520 nm: d-d optical transition of Ni
2+
[349]
Co-ZnS NCs 459 nm: Vs(S vacansy)-Dopant level (Co) [57]
Mn, Fe-ZnS NWs 580 nm: Mn
2+ 4
T
1

6
A
1
transition; 520 nm: Fe doping [216]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 229
4.7.3. Red light PL emission from Mn/Cd-co-doped-ZnS nanostructures
Mn/Cd-co-doped ZnS 1D nanostructures were synthesized on the surfaces of Au-coated Si sub-
strates with Zn and S species in the vapor phase. A red light PL emission band was found from the
Mn and Cd co-doped-ZnS nanostructures and this emission could be strengthened by adjusting the
Mn and Cd contents. Fig. 44a shows the PL spectrum of the Mn and Cd co-doped-ZnS nanostructures.
The experimental curves are well tted with multi-peak Gaussian tting, which gives ve emission
bands centered at 3.30 eV, 2.70 eV, 2.40 eV, 2.14 eV and 1.94 eV. The authors demonstrated that the
emission bands around 2.7 eV (455 nm), 2.4 eV (510 nm) and 2.14 eV (580 nm) were associated with
the trapped luminescence arising from the surface, Au impurities and Mn doping, respectively. The
transition from Mn
2+
4T
1
state to the Cd-related De state results in a red light emission. The schematic
energy level diagram is shown in Fig. 44, and more detailed information can be found in Ref. [223].
Table 18 summarizes the detailed optical properties of 1D doped-ZnS nanostructures.
5. Potential applications
5.1. Field-emission applications
5.1.1. Field emission phenomena
Field-emission (FE) (also known as electron eld-emission) is an emission of electron induced by
external electromagnetic elds. FE can take place from solid and liquid surfaces, or individual atoms
into vacuum or open air, or result from promotion of electrons from the valence to conduction band in
semiconductors. FE was explained by quantum tunneling of electrons in the late 1920s, in which elec-
trons pass from an emitting material (which is negatively biased) to the anode through a barrier (vac-
uum) in the presence of a high electric eld. This was one of the triumphs of the nascent quantum
mechanics. The theory of FE from bulk metals was proposed by Fowler and Nordheim, so these days
it is known as FowlerNordheim (FN) tunneling. This phenomenon is highly dependent both on the
properties of the material and the shape of the particular cathode, and the materials with higher as-
pect ratios and sharp edges produce higher eld-emission currents [1].
There are many important questions regarding a FE process. Namely, how would the ideal eld
emitter look like? What kind of factors are the determining ones for eld emitters? Should it be very
long, very thin, sharp tip and single crystalline, made of conductive material with high mechanical
strength, be robust, and cheap and easy to process? The current density produced by a given electric
eld is described by the FN equation [1,350]:
J = (Ab
2
E
2
=/) exp(B/
3=2
=bE); (9)
I = S J; E = V=d; (10)
or ln(J=E
2
) = ln(Ab
2
=/) B/
3=2
=bE (11)
where A and B are constants, (A = 1.54 10
6
A eV V
2
, B = 6.83 10
3
eV
3/2
V lm
1
), S is the emitting
area, V is the applied potential, I is the emission current, b is the eld-enhancement factor, d is the dis-
tance between a sample and the anode and / is the work function of the emitting materials. From FN
model, it is easy to note that the dependence of the emitted current on the local electric eld and the
work function is exponential-like. As a consequence, it is possible that a small variation of the shape or
surrounding of the emitter (geometric eld enhancement) and/or the chemical state of the surface
have a strong impact on the emitted current. The eld-enhancement factor, b, is related to the emitter
geometry (such as aspect ratios), its crystal structure, vacuum gaps, and the spatial distribution of
emitting centers. For some regular structures, b can be expressed as b = h/r, where h is the height
and r is the radius of curvature of an emitting center. Thus, materials with elongated geometry and
sharp tips or edges can greatly increase an emission current.
In accord with our previous review, Fig. 45a shows a schematic illustration of the above-mentioned
FE phenomena [1]. Generally, a rodlike metal probe is used as an anode and the emitted objects serve
as a cathode. A cross section of 1 mm
2
metal aluminum or copper probe was adopted in our
230 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
laboratory. This process involves the extraction of electrons from a solid by tunneling through the sur-
face potential barrier. The emitters can have different tip geometries such as, (i) round tip, (ii) blunt tip
and (iii) conical tip, as marked in Fig. 45b. A different emission current is generated depending upon
the tip geometry. Most of the FE phenomena are studied at room temperature and are barely temper-
ature dependent, so normally it is not a temperature activated process.
The determining factors for the emission current could be the work function of the emitting mate-
rials, the local electric eld, the distance between a sample and the anode, and the tip geometry?
According to Eqs. (1)-(3), the work function of an emitter surface plays a primary role in the emission
current. The work function / is the energy required to remove an electron from the highest lled level
in the Fermi distribution of a solid so that it is stationary at a point in a eld-free zone just outside the
solid, at absolute zero. Table 19 lists the work functions of some important inorganic semiconductors
and metals. It is clear that at a specic eld, a lower work function material can produce a higher elec-
tron emission current. However, not all low work function materials are ideal for constructing eld-
emission cathodes. This is because their other properties may not be suitable. For example, the work
function of cesium (Ce) / is about 1.8 eV, is one of the lowest, however, stable emission and lifetime
obtained from Ce or cesiated (cesium coated) cathodes is a serious concern [1].
Since the studies on carbon nanotubes (CNTs) [351356], much attention has been paid to explore
the prospects of inorganic semiconductor nanostructures as eld emitters due to their low work func-
tions, high aspect ratios, high mechanical stability, and high conductivity. The inorganic semiconduc-
tor nanostructures are ideal systems for exploring a large number of novel phenomena at the
Fig. 45. (a) A schematic illustration of eld emission phenomenon, showing the emission that occurs from the tip of an emitter.
(b) The emitters can have different tip geometry such as, (i) round tip, (ii) blunt tip and (iii) conical tip, and so on. Reproduced
from Ref. [1]. Copyright 2008, Royal Society of Chemistry.
Table 19
The work function / of some important inorganic semiconductors, C nanotubes and metals.
Semiconductors and C nanotubes ZnO C nanotubes AlN WO
3
SiC Si GaN GaAs CdS ZnS
Work function (eV) 5.3 5 3.7 5.7 4.0 3.6 4.1 4.77 4.2 7.0
Metals Au Ag Al W Sn Cu Fe Mo Cd Zn
Work function (eV) 5.1 4.26 4.28 4.5 4.42 4.65 4.5 4.37 4.07 4.3
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 231
nanoscale and investigating the size and dimensionality dependence of their properties for potential
applications [1].
As listed in Table 1, it is easy to note that the work function for ZnS (-7.0 eV) is higher than for
some other popular eld-emission materials, such as Si (3.6 eV), AlN (3.7 eV), SiC (4.0 eV), C nanotubes
(5 eV), ZnO (5.3 eV), CdS (4.2 eV), W (4.5), WO
3
(5.7) and Mo (4.37) [305,357359]. According to FN
theory, ZnS is not the best candidate for FE. Thus the questions are arising: how good are the eld-
emission properties of ZnS nanostructures? Is it possible to make excellent eld emitters out of var-
ious ZnS nanostructures by some facile routes?
In this section, the state-of-the-art research that focuses on FE properties of ZnS nanostructures,
including, ulrtane ZnS nanobelts as eld emitters, enhanced eld-emission properties by crystal ori-
entation-ordered ZnS nanobelt quasi-arrays, enhanced eld-emission properties by multi-angular
branched ZnS nanostructures with needle-shaped tips, multiply enhanced eld-emission of branched
ZnS nanotube-In nanowire coreshell heterostructures and ZnS nanostructures with other novel mor-
phologies, will be presented in detail. Some facile routes to enhance FE will be elucidated. In the end,
we conclude this section with some summaries and comparisons on FE properties between ZnS and
other nanomaterials briey.
5.1.2. Ultrane ZnS nanobelts as eld emitters
For a given material, it has been demonstrated that the eld-emission performance, including
emission current, eld-enhancement factor, etc., can be enhanced through increasing its aspect ratio
(length to thickness ratio), assembling it into arrays, or decorating its surface with a lower work func-
tion materialunder keeping a low turn-on eld [360363]. For example, Ye and co-workers proposed
two alternative approaches to enhance FE performance of standard ZnO nanorods with a diameter of
-150 nm: The rst one is to decrease the tip radius of ZnO nanorods by controlling the growth pro-
cesses, and the second one is to lower the work function of ZnO by modifying it with metal particles
(such as Pt and Ag nanoparticles) with a lower work function. The physical essence of the excellent
performances is the geometrically enhanced local eld near the tip of the nanorods and the lowered
work function of ZnO nanorods by donation of electrons from the nanoparticles [363]. Ren and co-
workers demonstrated that the eld-emission performance of ZnO nanowires can be signicantly en-
hanced through decreasing the density of the nanowires, and increasing their aspect ratio for ZnO
nanobelts [364,365].
In Section 3.2.2 most of the ZnS nanobelts/nanoribbons/nanosheets have a width ranging from tens
of nanometer to several micrometers. This is true for nanostructures synthesized by various methods,
including CVD process, vapor-phase transport process, simple thermal evaporation, thermal decompo-
sition and reduction, conversion reaction of CdS nanobelts, solvothermal reaction and subsequent heat
treatment, two-stage temperature-controllable thermal evaporation and condensation process, micro-
wave-assisted solvothermal synthesis, simple solvothermal method, diethylenetriamine-assisted
solvothermal approach, and solution method in the presence of block copolymer and so on. However,
the large-scale ZnS nanobelt synthesis, especially for those of small size (width < 20 nm) is still a chal-
lenge [128]. Moreover, ultrane nanostructures have been proved to be of particular interest for the
general study of size-dependent electrical and optical properties. For example, Wang and co-workers
reported the synthesis of single crystalline, ultrathin ZnO nanobelts (the average width of the nano-
belts was 5.5 nm with a narrow size distribution within 1.5 nm) using a solidvapor deposition meth-
od. The deposition was performed on Si substrates coated with a thin layer of tin. A 120 meV blue-shift
was observed in the photoluminescence spectra of the ZnO nanobelts. The authors suggested that such
ultrane nanobelts should have interesting electrical and optical properties. Thus the synthesis and
novel property exploration of ultrane ZnS nanostructures are indispensable and important, especially
an ultrahigh aspect ratio is crucial for the excellent eld-emission performance [366].
By checking a lot of pre-existing experiments, we developed a novel method to synthesize ultrane
ZnS nanobelts in a vertical induction furnace. The key point is a control of the evaporation and agglom-
eration rates [128]. A mixture of commercial ZnS powders and C nanopowders, and S powders was
adopted as precursors instead of the previously used pure ZnS powders or nanopowders
[110,111,140]. The evaporation and agglomeration rates were signicantly decreased in our case com-
pared with previous experiments. In fact, the evaporation rate would be very fast if pure ZnS powders
232 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
or nanopowders are used as precursors within a high temperature zone. This would result in numer-
ous nucleation centers and fast agglomeration rates in a relatively low temperature zone. In the later
case, relatively thick and wide ZnS nanobelts/nanosheets have usually been formed.
During the formation of ultrane ZnS nanobelts, rst, the ZnS powders and C nanopowders under-
go a reaction: 2ZnS + C ?2Zn + CS
2
, to gradually produce a Zn vapor. The Zn vapor reacts with S vapor
to form nanobelts at a relatively moderate temperature (-1000 C). Superuous C gas is also trans-
ferred to the relatively low temperature zone and absorbs on the as-grown ZnS nanobelts, blocking
the belt width. Thus the thickness develops less effectively than the belt length, since the growth along
the [0 0 1] direction for ZnS nanobelts is fastest and easiest according to the growth kinetics. This leads
to ultrane ZnS nanobelt formation. Fig. 46 shows SEM images of the as-grown nanobelts. The range of
a
b
10 20 30 40 50 60 70 80
I
n
t
e
n
s
i
t
y

(
a
.
u
.
)
2 (defgree)
(
1
0
0
)
(f)
(
1
0
2
)
(
1
0
1
)
0
0
2
(
1
1
0
)
(
1
0
3
)
(
2
0
0
)
(
1
1
2
)
(
2
0
1
)
(
2
0
2
)
(
2
0
3
)
(
2
1
0
)
(
2
1
1
)
c
Fig. 46. (a) and (b) Low-magnication and high-magnication SEM images, showing a uniform thickness and width of the
structures over the entire length. (c) XRD pattern recorded from the ultrane ZnS nanostructures. Reproduced from Ref. [128].
Copyright 2007, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 233
the belt widths was 530 nm, peaked at 1020 nm, as veried by a low- and high-magnication SEM
images in Fig. 46a and b. The belt thickness varied with the width and was likely of only several nano-
meters. An XRD spectrum taken from the nanobelts is shown in Fig. 46c, where all the diffraction
peaks can be indexed to wurtzite-type ZnS without any impurities [128].
A lattice-resolved HRTEM image taken from the individual nanobelts of -10 nm width is shown
implies that the nanobelts are largely defect-free. The smallest width of ZnS nanobelts during our
TEM observations was found to be -7.5 nm.These particular nanobelts were extremely thin. Fig. 47
shows a UVvis absorption spectrum taken from the present nanobelts, showing a pronounced quan-
tum connement effect (a blue-shift of -0.45 eV compared to bulk ZnS materials) [128].
Fig. 48 shows the FE current density, J, as a function of the applied eld, E, for a JE plot and a ln(J/
E
2
)(1/E) plot measured at the anodecathode distances of 200 lm. The relatively smooth and consis-
tent curves indicate the stability of emission from the belt emitters. The turn-on eld at a current den-
sity of 10 lA/cm
2
was extrapolated to be -3.47 V/lm. The emission current density reached
-11.5 mA/cm
2
at a macroscopic eld of 5.5 V/lm. The enhancement factor b was calculated to be
-2.01 10
3
from the slope of the tted straight line in Fig. 48. These data shows that the present
ultrane ZnS nanobelts are decent eld emitters.
5.1.3. Enhanced eld-emission properties by crystal orientation-ordered ZnS nanobelt quasi-arrays
It is well known that aligned arrays with a proper nanowire density can enhance the material eld-
emission (FE) properties. For example, Singh and co-workers reported enhanced cold eld-emission
from (1 0 0) oriented bW nanoemitters. A highly conned emission current of about 23 lA was ob-
tained at a low extraction voltage of -260 V at -280 nm anodecathode distance with <3% uctua-
tions over -2 h [367]. Chu and co-workers reported enhanced electron FE from oriented columnar
(0 0 2) oriented AlN thin lms with a columnar microstructure fabricated by vapor phase deposition.
They also demonstrated further by atomic force microscopy (AFM) that nanoscale protrusions on the
surface expand the emission area and increase the efciency greatly [368]. Li and co-workers found
that enhanced FE from a nested array of multi-walled carbon nanotubes grown on a Si nanoporous
pillar array might be achieved. The authors proposed that the excellent FE performances had been
attributed to the unique structural and morphological features of the arrays [360].
Although alignment of nanostructures has been proven to be one of the facile routes to enhance FE
performances, the aspect ratio of the emitters, arrays distributions (including arrays distance, density,
etc.), and the morphology of individual nanostructures, etc. should play a different role in various FE
200 250 300 350 400 450 500 550 600
A
b
s
o
r
p
t
i
o
n

(
a
.
u
)
Wavelength (nm)
4.1 4.2 4.3 4.4 4.5 Eg

2

(
a
.
u
.
)
h (eV)
Plot of h
2
-- h
Linear Fit and Prolonged Line
Fig. 47. An UVvis absorption spectrum of ultrane ZnS nanobelts, showing a notable quantum connement effect compared
with the bulk ZnS material (blue-shifted -0.45 eV). Reproduced from Ref. [128]. Copyright 2007, Wiley-VCH.
234 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
performances, including turn-on eld, eld-enhancement factor, current density at a xed eld, FE
stability, etc. For example, Zhai and co-workers synthesized various ordered CdS nanostructure arrays
with different morphologies through a two step MOCVD process. The morphology of CdS nanostruc-
ture arrays was varied from truncated nanocones, nanorods, or cleft nanorods to nanowires arrays via
increasing the distances between precursors and substrates. FE measurements showed that both the
alignments and aspect ratios had a signicant inuence on their FE properties. The quasi-aligned
nanowire arrays emitters with a higher aspect ratio and better alignment exhibited better FE perfor-
mance compared to other CdS nanostructures [164]. Thus, the issue of balancing FE performances and
array distributions seems to be crucial for the emitter optimization.
Crystal orientation-ordered ZnS nanobelt quasi-arrays were successfully fabricated using non-cat-
alytic and template-free thermal evaporation process [132]. Different from the preparation of the
above-discussed ulrtane ZnS nanobelts, the synthesis of quasi-arrays ZnS nanobelts were conducted
in a conventional horizontal tube furnace with a 36 mm inner-diameter quartz tube. The seeding ZnS
sheets, which were prepared by pressing ZnS nanoparticles (mesh -100 nm) under 40 MPa, were
placed on a long alumina plate to act as the deposition substrates. For comparison, we also synthe-
sized randomly-oriented ZnS nanobelts; for this a long alumina plate was used as the deposition sub-
strate, while the other experimental parameters were kept unchanged. Commercial ZnS powders
(-10 lm) were used as a precursor. The growth conditions for both ZnS nanobelts were: the furnace
0 1 2 3 4 5 6
0
2
4
6
8
10
12
a
Electric Field V/m
C
u
r
r
e
n
t

D
e
n
s
i
t
y

m
A
/
c
m
2
0.15 0.20 0.25 0.30 0.35 0.40 0.45
-45
-40
-35
-30
-25
-20
Experiment Date
Linear Fit
l
n
J
/
E
2
A
/
c
m
2
/
V
/

m
2
1/E m/V
b
Fig. 48. FE properties of ultrane ZnS nanobelts recorded at the anodecathode distances of 200 lm. (a) JE curve, showing a
turn-on eld of -3.47 V/lm at a current density of 10 lA/cm
2
. (b) FowlerNordheim plot corresponding to (a). Reproduced
from Ref. [128]. Copyright 2007, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 235
was heated to 1100 C in 10 min and kept at this temperature for 60 min (the furnace was purged with
high-purity Ar for 3 h prior to heating), high-purity Ar served as a protecting mediumand carrying gas,
during the process, the Ar ow rate was kept at 100 sccm [132].
The nanobelts are not only aligned on the micro/macroscopic scales but also display the similar
crystallographic orientation of their growth axes. The observations also revealed that a typical length
of the belts is of several tens-to-hundreds micrometers. Some of them may even reach a millimeter
length. Most of the nanobelts had perfectly parallel ensembles. A typical belt width was 50120 nm
[132]. We found that a deposition substrate certainly plays the key role in the ZnS nanobelt quasi-ar-
rays formation process. The seeding ZnS sheets made of pressed ZnS nanoparticles offered the numer-
ous nucleation centers for the ordered arrays growth, in fact many ZnS nanobelts were found to grow
from the seeding ZnS sheets vertically. The epitaxial growth along the [0 0 1] direction seems to be the
most energetically favorable process since both the ZnS substrate and the as-grown ZnS nanobelts
have the same wurtzite structure. Oppositely, only randomly-oriented ZnS nanobelts were synthe-
sized when a long alumina plate was used as the deposition substrate instead of the seeding ZnS
sheets but the other experimental parameters were kept unchanged [132]. The method reported here
shows the possibility of growing a large area of orientation-ordered ZnS nanobelts. Itthus becomes
possible to grow patterned and uniformly sized ZnS or other aligned nanostructures to be ultimately
utilized in nanoscale devices.
Field-emission measurements show that the arrays are decent eld emitters possessing a current
density of more than 20 times higher than that of randomly-oriented ZnS nanobelt ensembles at a
macroscopic eld of 5.5 V/lm. The comparative plots of a FE current density J versus an applied eld
E for the lms are depicted in Fig. 49a. The inset in Fig. 49a is a magnied JE curve for a randomly-
oriented ZnS nanobelt lm. It can be found that the turn-on elds for the lms made of ordered and
random ZnS nanobelts were extrapolated as -3.55 and 4.19 V/lm, respectively, at a current density of
10 lA/cm
2
. The emission current density for quasi-aligned and well-ordered ZnS nanobelt arrays
reached -14.6 mA/cm
2
at a macroscopic eld of 5.5 V/lm. This value is not only -21.5 times higher
than for randomly distributed ZnS nanobelts (0.68 mA/cm
2
at the same macroscopic eld), but is also
comparable with the massive arrays of monodispersed CNTs nanotubes that are self-oriented on pat-
terned porous Si and plain Si substrates [356].
The eld-enhancement factors b were calculated for the ordered and disordered ZnS nanobelt
ensembles to be -1.85 10
3
and -1.60 10
3
from the slops of ln (J/E
2
)(1/E) plots in Fig. 49b. It
can be concluded that aligned structures with a high density not only useful for enhancing FE current
density, but also it can enhance the material FE properties as compared to random arrays. Fig. 49c
shows variations of emission current density of the ordered ZnS nanobelt ensembles within 4 h at
an applied electric eld of 5 V/lm. It is easy to note that the quasi-arrays display good stability of
the eld-emission. The emission current uctuations were as low as 2.58%. Such stable FE perfor-
mance should be related to a uniform height of aligned ZnS nanobelts, that guarantees a uniform eld
distribution across the arrays.
5.1.4. Enhanced eld-emission properties by multi-angular branched ZnS nanostructures with needle-
shaped tips
With respect to FE applications, it is also well known that the needle-shaped tips can signicantly
enhance the material FE properties. The synthesis of micro/microstructures with needle-shaped/sharp
tips is always the center of FE research and FE application. In 2006, Chen and co-workers presented a
comprehensive review on nanotips [369]. The nansocale tip morphologies have been discussed in this
review. Different growth techniques used for the synthesis of the nanotips have also been outlined and
compared including, etching, pyrolysis, physical, and chemical vapor depositions. Growth models for
hollow and solid nanotips were discussed in detail in the frame of the cone-helix model for the hol-
low tips and the platelet stacking model for the solid tips. Several applications of the nanotips inscan-
ning probe microscopy, FE, broad band antireection, molecular detection, and nanoindentation
demonstrated the enormous promise of this class of materials. For FE properties, the authors sug-
gested that some inherent problems existing in the nanotube or nanowire systems may be automat-
ically overcome by selecting the proper nanotip structure. For example, the eld screening in the case
of densely packed nanotubes is much reduced for nanotips because the tips are spaced apart. This
236 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
spacing also depends on their base diameters [369]. Here, we would like to only list some represen-
tative examples on excellent FE properties based on the needle-shaped nanotips. The complete discus-
sion is beyond the scope of this review. Chen and co-workers reported that Si nanotips with a diameter
and height measuring 1 nm and 1 mm, respectively, and density in the range of 10
9
3 10
11
cm
2
,
could be fabricated monolithically from Si wafers by electron cyclotron resonance plasma etching
technique at a temperature of 200 C. The FE current density was 3.0 mA/cm
2
at an applied eld of
-1.0 V/mm. By the aid of HRTEM and Auger electron spectroscopy analyses, the authors found that
0.15 0.20 0.25 0.30 0.35 0.40
-45
-40
-35
-30
-25
-20
Randomly oriented ZnS nanobelts
Quasi-aligned and oriented ZnS nanobelts
l
n
J
/
E
2
A
/
c
m
2
/
V
/

m
2
1/E m/V
(b)
0 1 2 3 4 5 6
0
2
4
6
8
10
12
14
16
Randomly oriented ZnS nanobelts
Quasi-aligned and oriented ZnS nanobelts
Electric Field V/m
C
u
r
r
e
n
t

D
e
n
s
i
t
y

m
A
/
c
m
2
(a)
0 1 2 3 4 5 6
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0 60 120 180 240
2.5
3.0
3.5
4.0
4.5
5.0
C
u
r
r
e
n
t

D
e
n
s
i
t
y

(
m
A
/
c
m
2
)
Time (min)
(c)
Fig. 49. A comparison of FE properties between quasi-aligned ZnS nanobelt arrays and randomly-oriented ZnS nanobelts. (a) J
E curves and (b) FowlerNordheim plots corresponding to (a). The inset to (a) is a magnied JE curve of random ZnS nanobelts.
A gap between the anode and the cathode was set at 200 lm. (c) Stability measurement for emission current density of the
quasi-aligned ZnS nanobelt arrays ensembles within 4 h at an applied electric eld of 5 V/lm. Reproduced from Ref. [132].
Copyright 2007, Royal Society of Chemistry.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 237
the nanotips had been composed of monolithic Si and nanometer-size SiC caps at the top. The turn-on
eld of 0.35 V/lm turn-on eld to draw a 10 lA/cm
2
current density was reached [370]. Wu and co-
workers fabricated a Si nanotip array by using a Si-based porous anodic alumina membrane (PAAM) as
a mask. This array was obtained by removing the SiO
2
nanoislands formed during anodization of the
Al/Si interface. The authors claimed that the array exhibits excellent eld emission characteristics
with a low turn-on eld of 8.5 V/lm (dened as the electric eld to extract a current density of
10 lA/cm
2
) and a large eld-enhancement factor of -1100 due to the unique surface structure and
morphology [371]. High-density ordered triangular Si nanopillars with sharp tips and varied slopes
were fabricated by reactive ion etching of ordered Ni nanodot arrays, patterned on Siusing nanosphere
lithography. The mixing ratio of CF
4
and O
2
in the etching gas was used to alter the slope and aspect
ratio of the nanopillars. FE measurements indicated that low turn-on eld of 1.6 V/lm (at a current
density of 0.1 lA/cm
2
) and a high eld-enhancement factors (the value of b for low-eld and high-
eld regimes were 4184 and 14,244, respectively) were regular characteristics of the nanopillar arrays.
The authors suggested that sharp tips and a ared base, which are advantageous for a high eld-emis-
sion current and mechanical/thermal stability, respectively, can be tailoreddepending on the applica-
tion requirements [372]. Wang and co-workers presented a VS process for growing well-aligned,
columnar AlN nanorods with multiple-nanotip surfaces. The overall nanostructure had an aligned
columnar morphology, and each nanorod has a hairy surface. This kind of AlN nanorods with multi-
ple-nanotip surfaces showed excellent FE properties with a low turn-on eld of -3.8 V/lm (at a cur-
rent density of 10 lA/cm
2
) [373].
By gradual reducing heating temperature during thermal evaporation, the synthesis of multi-angu-
lar branched ZnS nanostructures with needle-shaped tips was reported by our group [276]. The same
setup was used for the synthesis of quasi-arrays ZnS nanobelts explained in Section 5.1.2. A long alu-
mina plate or a quartz piece (30 cm in length) was placed to act as the deposition substrate, which was
put downstream of high-purity commercial ZnS powders (-10 lm, 99.99%) precursor in a quartz tube.
One of the main experimental parameters to control was the heating regime: the furnace was rst
heated to 1050 C in 9 min and then the reaction temperature was decreased to 700 C in 30 min. A
constant cooling rate of -12 C min
1
was used to reduce the temperature. Then the system was al-
lowed to cool down naturally to RT. Statistics based on numerous TEM observations (more than 100
nanostructures) indicates that ZnS nanostructures of different morphologies all had sharp needle-like
tips as a common characteristic (an average width of the sharp tips was -8 nm). The materials were
mainly composed of bi-angular structures (-20%), tri-angular structures (-50%), tetra-angular struc-
tures (-20%) and some other multi-angular nanostructures (-10%) [276].
Some works demonstrated that quite a number of the electrons are emitted from the tips, which
make the shape of tips and the local work function the most important physical parameters for FE
[374]. Fig. 50 shows a FE current density, J, as a function of an applied eld, E, for a JE plot and a
ln(J/E
2
)(1/E) plot measured at an anodecathode separation of 200 lm. A relatively low turn-on eld
is extrapolated as -3.77 V/lm at a current density of 10 lA/cm
2
from Fig. 50a while a high eld-
enhancement factor (-2182) is calculated from a slope of the tted straight line in Fig. 50b. The emis-
sion current density reached -3.035 mA/cm
2
at a macroscopic eld of 5.5 V/lm. Albeit the possession
of highest b, due to a relatively low density of the present ZnS multi-angular branched nanostructures
compared with the above-discussed works on ulrtane ZnS nanobelts and crystal orientation-ordered
ZnS nanobelt quasi-arrays, the turn-on eld is not the lowest one and the emission current density is
not the highest one among the structures studied. The possession of highest b should be attributed to
the specic sharp tips and high aspect ratios of the present needle-shaped structures [276].
5.1.5. Multiply enhanced eld-emission properties of branched ZnS nanotube-In nanowire coreshell
heterostructures
As described in Section 3, the synthesis of core/shell nano/heterostructures consisting of two or
more important functional materials is essential for developing potential nanoelectronic and optoelec-
tronic devices. Particularly, coreshell heterostructures consisting of two components with distinct
functionality often exhibit enhanced performances, such as emission efciency and high electron
mobility, which are the key factors for many device performances [1,189,375,376]. One can see a
straightforward analogy with the planar semiconductor industry when a complex composition and
238 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
structure modulations could greatly increase the versatility and power of the building blocks in nano-
scale electronic and photonic schemes [377]. Recent advances in the design and control of hetero-
structures and superlattices in 1D nanoscale semiconductors have opened the door to new device
concepts. For example, Ge/Si core/shell nanowire heterostructures have exhibited substantially higher
performance as eld effect transistors (FETs) and low-temperature quantum devices compared with
homogeneous materials [378]. Huo and co-workers reported that quasi-aligned nanoarrays consisting
of TiO
2
nanowire cores/carbon nanocone shells nanostructure arrays produced directly on titanium
foils via a simple one-step thermal reaction under acetone vapor at 850 C. A high eld-enhancement
factor b around 5304 was achieved, and the authors believed that the high b value of the quasi-aligned
TiO
2
/C nanocone arrays stem from the sharp tips, high aspect ratio, low screening effect, and good
electrical contact between the TiO
2
/C nanocones and conducting Ti substrate [379].
Here, we present the synthesis of a novel coreshell metalsemiconductor heterostructure where
In forms the core nanowire and wurtzite ZnS forms the shell nanotube and their multiply enhanced FE
properties, which have been developed by Gautam and co-workers in our Lab [189]. The ZnSIn het-
erostructures were synthesized using a high-temperature vertical induction furnace, which was also
used for the synthesis of Ga-doped ZnS nanowires [222]. The mixture of 0.10 g of In
2
S
3
, 0.25 g of
ZnS, and 0.05 g of activated carbon powder was used as a source material, which was put into a
Fig. 50. FE properties displayed by ZnS multi-angular branched nanostructures with needle-shaped tips. (a) JE plot and (b) the
corresponding FowlerNordheim ln(J/E
2
)(1/E) plot. Reproduced from Ref. [276]. Copyright 2008 American Chemical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 239
graphite crucible and placed at the center of the furnace. Additionally, 0.4 g of S was placed in another
graphite crucible at the furnace bottom where the temperature was estimated to be -400500 C. The
control of S vapor supply was found to be crucial during the synthesis. For example, in the absence of
an excessive S ow a product contained unreacted Zn and In microcrystals and the heterostructures
did not possess branches. The synthesis was made at high temperature (-1300 C) using a carbother-
mal CVD technique where In acts as a catalyst. Carbon was used as one of the source materials in order
to reduce the compounds to elements and to produce CS
2
. Approximately 0.1 g of the product was col-
lected from the deposition tube placed at the top of the crucible [189].
The heterostructures had an average length of few micrometers. The diameter of a core nanowire
was -50100 nm, whereas a ZnS shell was -150 nm thick. The branches had a length of -200 nm and
a width of -20 nm. Fig. 51a and b shows some typical SEM images of the heterostructures at different
magnications. The image in Fig. 51c shows that the sharp and quasi-aligned ZnS tips grow at the het-
erostructure tip, revealing that 1020 nm thick and -100200 nm long 1D branches decorate the en-
tire nanostructure surface [189].
TEM reveals that the core In nanowires are single-crystals, whereas, within a hierarchical shell, the
stem and the branches are separated with a crystalline interface. Fig. 52a shows a typical TEM image
of a single heterostructure, which reveals the clear and uniform contrast variation at the nanostruc-
ture core and shell. The elemental maps displayed in Fig. 52bd shed a light on the distribution of
the constituting elements, i.e. Zn, S, and In elemental maps and their corresponding K-edge signals
demonstrating well-dened compositional proles within the heterostructure. As seen in Fig. 52e, a
thin layer of amorphous carbon creates a coating on the nanostructures [189].
According to FN theory described in the section related to FE phenomena, the heterostructure with
the sharp and quasi-aligned ZnS tips should be attractive as a potential eld emitter. FE measurements
conrmed this idea revealing a remarkably large eld enhancement. This was explained on the basis
of a sequential stepwise enhancement mechanism involving the consecutive stem and branch
Fig. 51. (a) and (b) Typical SEM images of branched ZnS nanotube-In nanowire coreshell heterostructures. (c) The sharp and
quasi-aligned ZnS tips grew at the heterostructure tip. Reproduced from Ref. [189]. Copyright 2008 American Chemical
Society.
240 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
contributions. A relatively low turn-on eld of 5.56 and 5.43 V/lm at a current density of 10 lA/cm
2
was obtained when the anode was placed at a distance of 210 lm and 150 lm, respectively. From
Fig. 53a, it is easy to note that the FE current density can reach -1.1 mA/cm
2
at an applied macro-
scopic eld of 7.45 V/lm. As calculated from Fig. 53b, the eld-enhancement factor b was found to
be 1.6 10
3
at a working anode-sample distance of 210 lm. A variation of emission current density
of the present heterostructure within -3 h at an applied electric eld of 6.5 V/lm is shown in
Fig. 53b (inset). There are no any current degradations or notable uctuations during this period [189].
Compared to the previously reported free-standing ZnS nanorods or nanowires, the present hetero-
structures show a remarkably enhanced eld-enhancement factor b [79,148]. Gautam and co-workers
proposed a new model of eld enhancement in such a hierarchical nanostructure. As shown in
Fig. 53c, the applied eld in structure (i) is enhanced by a stem which acts as a substrate for the sec-
ondary branches rst; and then by a branch (ii), (iii) also highlights FE contribution from the stems of
the nanowires which are lying at of the substrate. This results in multiply enhanced eld-emission
properties. As the authors suggested, the present nanostructures are expected to become valuable
not only in fundamental science but also in device applications [189].
Fig. 52. (a) Typical TEM image of a heterostructure. (be) The Zn, S, In, and C elemental maps and their corresponding K-edge
signals demonstrating well-dened compositional proles within the heterostructure. Reproduced from Ref. [189]. Copyright
2008 American Chemical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 241
5.1.6. ZnS nanostructures with other novel morphologies
The above four examples are the typical FE results of ZnS nanostructures using the same measure-
ment systems. They demonstrated that ZnS nanostructures with different morphologies may be envis-
aged to be important candidates for potential eld emitters. These developed methods are not only
facile to enhance the FE properties of ZnS nanostructure, but also they are general to develop other
inorganic semiconducting nanostructures into potential eld emitters [1].
To date, there have been only few reports that deal with eld-emission properties of ZnS nano-
structures. Table 20 lists the details of almost all eld-emission measurements carried out on them
Fig. 53. (a) and (b) FE properties of branched ZnS nanotube-In nanowire coreshell heterostructures at two different anode-
sample distances of 150 and 240 lm. The inset in (b) is the stability curve over -3 h measurement at a working eld of 6.5 V/
lm. (c) Schematics of eld enhancement in a hierarchical nanostructure. Reproduced from Ref. [189]. Copyright 2008
American Chemical Society.
242 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
so far. It is easy to note that ZnS nanostructures are not only formed in various morphologies, includ-
ing nanobelts or nanoribbons, nanorods, nanowires, multi-angular branched ZnS nanostructures with
needle-shaped tips, branched ZnS nanotube-In nanowire coreshell heterostructures, nanowire/nano-
belt arrays, tetrapod tree-like heterostructures, hexagonal pyramids of zinc blende structured single-
crystals, but indeed these unique ZnS nanostructures are excellent eld emitters.
As discussed in Section 5.1, it is established that a material with a lower work function can produce
a higher electron emission current at a specic eld. Although the work function for ZnS is higher than
for some other popular eld-emission materials (as listed in Table 20), it is possible to turn ZnS nano-
structures into excellent eld emitters via designing various facile routes.
Besides shown examples, some other facile routes have also been developed by several groups. The
differences in FE properties, including turn-on eld, eld-enhancement factors, emission current den-
sity, stability, etc., are not only relative to the synthesis conditions and different morphologies, but also
are tightly connected with the measurement systems, such as a vacuum system, a distance between
sample and tip. Here, we briey list some recent works on FE properties of ZnS nanostructures. Chatto-
padhyay and co-workers reported the FE properties ZnS nanorods synthesized in the thin lm form on
Si substrates using radio frequency magnetron sputtering of a polycrystalline prefabricated ZnS target
Table 20
Key performance parameters of ZnS nanostructure eld emitters reported in the literature.
+
Herein we dene the turn-on eld as
the eld producing emission current density of 10 lA/cm
2
. If the other values are used, it is separately mentioned.
ZnS emitters Synthesis method Turn-on
eld (V/
lm)
Field-
enhancement
factor (b)
Emission
current
density
(mA/cm
2
)
Stability: testing
time and uctuation
Ref.
Nanobelts or
nanoribbons
Thermal-
evaporation
3.55 1850 -14.6 at
5.5 V/lm
6 h, < 3% [132]
Thermal-
evaporation
4.19 1600 -0.68 at
5.5 V/lm
[132]
CVD process 3.47 2010 -11.5 at
5.5 V/lm
[128]
Solvothermal
reaction
3.8 1811 -0.035 at
5.7 V/lm
[148]
Solvothermal
reaction and
subsequent heat-
treatment process
3.8 1839 -0.035 at
4.5 V/lm
[149]
Nanorods Radio frequency
magnetron
sputtering
technique
2.96.3
at
2.452 lA/
cm
2
420105 -1.0 at 9
13 V/lm
[106]
Nanowires Vapor phase
deposition
11.7 at
0.1 lA/
cm
2
522 -0.55 at
15 V/lm
[79]
Multi-angular
branched ZnS
nanostructures
with needle-
shaped tips
Thermal
evaporation
approach
3.77 2182 -3.035 at
5.5 V/lm
[276]
Branched ZnS
nanotube-In
nanowire core
shell
heterostructures
Carbon-thermal
CVD technique
5.43 1600 -1.1 at
7.45 V/lm
No any current
degradations or
notable uctuations
during 3 h
[189]
ZnS nanowire arrays Thermal
evaporation
5.41 at
0.1 lA/
cm
2
1529 -0.01 at
6.5 V/lm
[72]
ZnS tetrapod tree-like
heterostructures
Thermal
evaporation
2.66 2600 -30 at
4.33 V/lm
-12 h, -1% [272]
ZnS Hexagonal
Pyramids
Thermal
evaporation
2.81 -3000 -20 at
5.0 V/lm
-8 h, -0.9% [274]
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 243
at a relatively higher pressure (10
1
mbar) and at a lower substrate temperature (233273 K) without
using catalyst. FE study showed that the turn-on elds were in between 2.9 and 6.3 V/lmfor the diam-
eters in the range of 1040 nmat a current density of 2.452 lA/cm
2
with an electrode distance of 80 lm
[106]. Chang and co-workers synthesized ZnS nanowires througha vapor phase deposition method. The
nanowires had sharp tips and were distributed uniformly on Si substrates. The diameter of the bases
was in the range of 320530 nmand that of the tips was around 2030 nm. FE measurements suggested
a relatively low threshold eld for this kind of ZnS nanowires with sharp tips, rough surfaces and high
crystal quality [79]. Cai and co-workers fabricated oriented wurtzite ZnS nanobelt arrays by a solvo-
thermal reaction and subsequent heat treatment. The nanobelts grew along the [0 0 0 1] direction per-
pendicularly on a Zn substrate. They were -30 nm in thickness, several hundreds of nanometers in
width, and up to 4 lm in length. A turn-on eld of 3.8 V/lm and a eld-enhancement factor of
-1800 were detected [148,149]. Chen and co-workers reported the synthesis of ZnS tetrapod tree-like
heterostructures and ZnS hexagonal pyramids of zinc blende structured single crystal by a thermal
evaporation method. FE measurements showed the emiiter prospects for a potential use [272,274].
5.2. Field effect transistors (FETs) and carrier characteristics
Though eld effect transistors (FETs) based on thin lms are well known for decades, lower sensi-
tivity has limited their applications. Due to the ultrahigh surface-to-volume ratio, NWs are ideal
Fig. 54. (a) Low-magnication and (b) high-magnication TEM image of a ZnS/silica nanocable. (c) I
D
V
D
measurements for a
back-gate nanocable transistor at various gate voltages. The inset shows a SEM image of a typical nanocable transistor.
Reproduced from Ref. [182]. Copyright 2007 American Chemical Society.
244 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
choices for sensors. As ZnS nanostructures with excellent crystallinity can be obtained in high yield
[6,189,222], their uses in FETs based device fabrication have also been demonstrated [182,217,380].
Wang and co-workers used ZnS/SiO
2
core/shell nanocables synthesized using a vaporliquidsolid
growth method to fabricate a FETs nanodevice (Fig. 54a and b). The ZnS core was single crystalline
while the shell consisted of amorphous silica [182]. To assess the electrical characteristics of the
ZnS/SiO
2
nanocables, FETs devices were fabricated using standard back gate geometry. A typical
SEM image of the FETs device is shown in the inset of Fig. 54c. Fig. 54c depicts the drain current
(I
D
) versus drain voltage (V
D
) characteristics of ZnS/SiO
2
nanocable transistors obtained as a function
of different gate voltages (V
g
). The increase in drain current at positive gate voltage is indicative of an
n-type semiconductor. Subsequently in a different conguration, the device was immersed into the
electrolyte solution with an additional electrode, which allowed the application of a gate voltage
through the electrolyte solution.
This device immersed in liquid was further used for biological and chemical sensing, by monitoring
the electrical conductance during protein or chemical additions. Fig. 55a shows the schematic of the
sensor. Protein, such as bovine serum albumin (BSA) has a strong afnity to silica surface, which
makes it possible to utilize the charged BSA as a gate. BSA with an isoelectric point of 4.8 has a partial
negative charge. Since the nanocable is n-type, the conductance of the device decreases when BSA is
added to it. This can be seen in Fig. 55b and c. The conductance of the device, on the other hand, was
found to increase when a positively changed protein was added. The same device concept was later
extended for sensing different chemicals as well as pH of a solution.
Fig. 55. (a) Schematic diagram of a FETs device acting as a sensor; the binding of BSA with a net negative charge is expected to
yield a decrease in the conductance. (b) Plot of the conductance versus time, where region 1 corresponds to the addition of
buffer solution, and region 2 corresponds to the addition of PB solution with dissolved BSA. (c) Plot of the conductance versus
time of a nanocable device during exposure to various concentrations of BSA solution, where region 1 corresponds to the
addition of 3 lL of phosphate buffer (PB) solution with 0.0005 g/L BSA, region 2 corresponds to the addition of 3 lL of PB
solution with 0.001 g/L BSA, and region 3 corresponds to the addition of 3 lL of PB solution with 0.005 g/L BSA. Reproduced
from Ref. [182]. Copyright 2007 American Chemical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 245
5.3. p-type conductivity in ZnS
The ability to dope semiconductors with suitable impurities to obtain n-type or p-type conductiv-
ities is key to the success of electronics. In this regard, a relative ease of IIIV semiconductor dopings
(far better than in IIVI semiconductors) is important. The latter shows a relative ease for either p-type
or n-type doping, but not for both types [381]. Amongst ZnS, ZnSe, and ZnTe, reliable p-type doping
was not observed in ZnS, only N impurities were placed in ZnSe, but for ZnTe most group-V dopants
could be utilized. ZnS can be easily n- doped with Al and Cl [382]. On the other hand, naturally occur-
ring oxide is usually a n-type semiconductor and thats why the full potential of ZnO nanostructures in
electronic devices has not been realized [307,383]. Thus amongst chalcogenides, ZnS carries the max-
imum potentiality for electronic devices due to possible p-type and n-type dopings.
Recently, there has been a great progress in controlling conductivity and carriers in bulk ZnS [384
390]. For example, Iida and co-workers prepared p-type ZnS thin lms by doping with N [384]. The
usually employed dopant elements for p-type conductivity in ZnS are Ag, Tm, N, or combination of
InAg, InAgN, etc. The nature and the possibility of tuning the conductivity of a material depend
not only on its bandgap, but also on the positioning of the band-edges with respect to the vacuum le-
vel. For example, the lower is the valence band maxima (VBM), more difcult is to obtain p-type
behavior. With the lower conduction-band minima (CBM), formation of compensatory defects become
easier. A comparison of calculated energy bands of ZnO and ZnS was calculated by Gai et al. [391].
These authors showed that VBM and CBM can be carried out by tuning the synthesis conditions,
such as creating Zn rich or S rich synthesis atmosphere or using a suitable dopant. Defects such as
Zn vacancies lead to p-type behavior while others like Zn interstitials, or S vacancies kill the p-type
behavior. It is shown that when ZnS is prepared in S rich conditions, the Zn vacancies are stabilized
as compared to ZnO. Thus ther should be a better transfer to the p-type behavior. Further improve-
ment was predicted by using various dopants such as N, Ag, Cu, Li, and Na, wherein N was suggested
to be the most suitable. Despite its obvious advantages over other chalcogenides for electronic devices,
the efforts in controlling conductivity, carrier type and carrier concentrations in ZnS nanostructures
have been limited to date.
Lee and co-workers have demonstrated fabrication of p-type ZnS nanobelts by doping with nitro-
gen [217]. Nitrogen-doped ZnS nanobelts were prepared via a chemical vapor deposition method
using a ZnS powder and high-purity Ar premixed with 5% H
2
as a carrier gas. NH
3
was added to the
reaction atmosphere. The as-synthesized N-doped ZnS nanobelt was assembied into FETs, and the
measurement results revealed that the nanobelts possess p-type conductivity. The effective eld-ef-
fect hole mobility and the hole concentration were estimated as 0.005 cm
2
V
2
s
1
at VDS = 10 V
and 8.6 10
16
cm
3
, respectively. The conductivity of the as synthesized nanobelts was found to en-
hance by 100 times upon annealing at high temperature while the hole mobility increased over 40
times. The measured hole mobility is comparable with the conventional nanomaterials such as p-type
ZnO nanowires (1017 cm
2
V
2
s
1
) and p-type GaN nanowires (12 cm
2
V
2
s
1
) [392,393]. These
observations imply bright future for ZnS based FETs devices.
5.4. Catalytic activities
ZnS is a direct wide-gap semiconductor with remarkable chemical stability against oxidation and
hydrolysis. Importantly, these properties are retained when the particle size steps down to just a few
nanometers. Therefore nanoparticles of ZnS are interesting entities for catalytic activities, where they
would require to be exposed to aggressive environments. The electronic properties and bandgap of
ZnS could be easily modied by doping with several elements. Besides, ZnS is available in abundance
and nontoxic. Therefore ZnS can play an important role as catalyst in environmental protection
through the removal of organic and toxic water pollutants. ZnS nanomaterials have been used for
the photocatalytic degradation of organic pollutants such as dyes, p-nitrophenol, and halogenated
benzene derivatives in wastewater treatment [394402]. These studies revealed that ZnS nanostruc-
tures are good photocatalysts due to rapid generation of electronhole pairs by photo-excitation and
highly negative reduction potentials of excited electrons.
246 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
Hu and co-workers reported a simple procedure for mass production of ZnS nanoporous nanopar-
ticles (NPNPs) comprising of hexagonal wurtzite ZnS nanocrystals, several nanometers in diameter,
and their photocatalytic properties for degradation of eosin B molecules [394]. For the purpose of com-
parison, they have used NPNPs, ZnS nanocrystals obtained by a standard procedure and a commercial
titania based photocatalyst. Fig. 56 shows the absorption spectrum of an aqueous solution of eosin B in
the presence of ZnS NPNPs under exposure to UV light for various durations. The strong absorption at
517 nm corresponding to eosin B gradually diminishes with increasing time indicating its photodeg-
radation. The same can be observed from the color of the solution too, as seen in the inset.
Also to compare catalytic activities, the solution of eosin B was exposed under following condi-
tions: (a) with Degussa P25 titania, in the dark; (b) without catalyst with UV light; (c) ZnS NCs (stan-
dard method) in the dark; (d) with ZnS NPNPs, in the dark; (e) with Degussa P25 titania and UV light;
Fig. 57. Photodegradation of eosin B under different conditions: (a) with Degussa P25 titania, in the dark; (b) without any
catalyst, with UV light; (c) with ZnS NCs, in the dark; (d) with ZnS NPNPs, in the dark; (e) with Degussa P25 titania and UV light;
(f) with ZnS NCs and UV light; (g) with ZnS NPNPs and UV light. Equal amount of catalyst was used in all cases. Reproduced from
Ref. [394]. Copyright 2005, Wiley-VCH.
Fig. 56. Absorption spectrum of a solution of eosin B in the presence of ZnS NPNPs under exposure to UV light. Reproduced from
Ref. [394]. Copyright 2005, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 247
(f) with ZnS NCs (standard method) and UV light; and (g) with ZnS NPNPs and UV light (Fig. 57). Using
conditions (a)(d), the photocatalytic affect on the solution degradation without catalysts but under
exposure to UV light was the same as that with catalyst but without exposure to UV light. The concen-
tration of the eosin B solution did not change considerably. However, when exposed to UV light, as
seen from the curves e (Degussa P25), f (ZnS NCs), and ed(ZnS NPNPs) in Fig. 57, the ZnS NPNP pho-
tocatalyst showed much greater activity than Degussa P25 or ZnS NCs. Exposure of the solution of eo-
sin B to UV light for about 40 min resulted in complete decolorization. The photocatalytic superiority
of ZnS NPNPs over ZnS NCs is due to the inuence of unwanted aggregation of ZnS NCs during the
reaction, which leads to a rapid decrease in the active surface area, while ZnS NPNPs maintain an
excellent porous nanostructure and do not aggregate [394].
The reduction of atmospheric CO
2
is a growing challenge and ZnS nanocrystals have been shown to
be efcient catalyst for photocatalysis of such reduction. Henglein and co-workers reported the ef-
cient photoreduction of CO
2
to formic acid [403]. They used SO
2
stabilized nanocrystals of ZnS as a
photocatalyst and 2-propanol as a sacricial electron donor. The quantum yield for the formation of
formic acid was found to be 0.80. Similar results were reported by Inoue and co-workers. The latter
revealed that when the nanocrystals of ZnS were prepared in the presence of excess Zn
2+
, these exhib-
ited high photocatalytic activity for the CO
2
reduction [404]. Yanagida and co-workers have system-
atically investigated the photo-catalytic activities [395399]. They achieved photoreduction of CO
2
to formic acid at pH 7 using NaH
2
P0
2
as an electron donor with Na
2
S. Photoreduction of CO
2
to formic
acid is more efcient in the presence of Na
2
S. The catalytic activity commences due to the post photo-
excitation high reductive potential of electrons in ZnS (1.75 V relative to a standard hydrogen elec-
trode). The holes have an oxidative potential of about +1.85 V.
ZnS particles were prepared by mixing ZnSO
4
solution and Na
2
S [397]. Once cleaned thoroughly,
the reagents were mixed with required quantities of NaH
2
PO
2
and Na
2
S solution in water. This was
saturated with CO
2
under cooling in an ice bath, when the pH became 7 the sytem wasclosed with
a rubber stopper, and then irradiated with a 500-W high pressure mercury arc lamp. Fig. 58 shows
the rate of the formation of products when the system was exposed to 313-nm radiation. As seen, for-
mate (HCO
2
) forms efciently, accompanied by a small quantity of CO and H
2
. The apparent quantum
yield for formate formation was 0.24 at 313 nm. In this scheme, hydrogen phosphite ion (HPO
2
3
) was
the sole oxidation product. Importantly, although H
2
PO

2
is a good reducing agent with a large nega-
tive redox potential (1.2 V versus SCE at pH 7), the authors found that the reduction never occurred
either in the dark or under the irradiation in the absence of ZnS suspensions.
The possibility of using ZnS particles for attacking two most important global problems, i.e. water
splitting reaction to generate hydrogen and removal of nitrate based pollutants from water, have also
been demonstrated [405407]. What is more important is that these authorshave been able to adjust
Fig. 58. Rates of product formation during photoreduction of CO
2
by ZnS nanocrystals at pH 7, in the presence of SH and
H
2
PO

2
(closed circles, open triangle and open square symbols represent HCO

2
, H
2
and CO, respectively). Reproduced from Ref.
[397]. Copyright 1992 American Chemical Society.
248 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
the photocatalytic activity of ZnS using the visible light. For this purpose, ZnS was either doped with
Cu or Ni. ZnO:957CuO:043S efciently produced H
2
at a rate of 0.45 mol/h under visible light irradi-
ation from an aqueous solution of Na
2
SO
3
, as shown in Fig. 59 [407]. It was found that ZnO:957-
CuO:043S possessed high activity even without a Pt cocatalyst. The quantum yield at 420 nm was
determined to be 3.7%.
Similar photocatalytic activity was achieved by using a Ni-doped ZnS nanocrystalline photocatalyst
with an energy gap of 2.4 eV. In addition to hydrogen production, it was later showed that the same
catalyst could be utilized for the reduction of water. It might dissolve nitrate and nitrite ions to ammo-
nia and dinitrogen under visible light irradiation in the presence of methanol as a reducing reagent
[405]. The reduction of nitrate ions competes with that of water to form molecular hydrogen. The
reactions can be described as follows:
NO
3
2H

2e

NO
2

H
2
O (12)
NO
2

7H

6e

NH
3
2H
2
O (13)
2H

2e

H
2
(14)
The authors investigated the dependence of photocatalytic reduction of nitrate ions upon the con-
centration of KNO
3
, as shown in Fig. 60. At a high concentration of nitrate, the reduction of nitrate ions
predominated over that of water. As the concentration was low the reduction of water became
predominant.
Fig. 60. Plot showing amount of products obtained from photocatalytic reduction of NO
3
over Zn
0:999
Ni
0:001
S in the presence of
CH
3
OH at various concentration of KNO
3
under visible light illumination (wavelength = 420 nm). (a): H
2
, (b): NO
2
, (c): NH
3
, (d):
N
2
. Reproduced from Ref. [405]. Copyright 2002 The Chemical Society of Japan.
Fig. 59. Photocatalytic evolution of H
2
from an aqueous Na
2
SO
3
solution (0.5 mol/l) under visible light irradiation (wavelength
of 420 nm) over (a) ZnO:957CuO:043S and (b) Pt (1 wt.%)/CdS. Reproduced from Ref. [407]. Copyright 1999 Springer.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 249
Amongst others, electrocatalysts is a eld that had not involved ZnS-based materials until recently.
Bredol and co-workers showed that ZnS nanoparticles had catalytic activity for the decomposition of
ethanol, a potentially abundant fuel for mobile electricity generation, since it could be fabricated by
fermentation from a broad range of organic materials [408]. Traditionally in electrochemistry, metal
nanoparticles, such as platinum have been used for catalytic functions. Despite their high catalytic
activity, metals suffer from several problems, such as poisoning of their surface by CO or H
2
. Therefore
these can only work with highly puried materials and alternative energy sources like biogenic alco-
hol cannot be employed in fuel cells. The authors proposed that electrocatalytic chalcogenide semi-
conductors would circumvent such problems due to the lack of specic afnity to CO and the
abundant presence of chalcogenide functions in the catalyst itself.
Electrocatalytic conversion of ethanol in the presence of O
2
to formH
2
O and CO
2
involves the trans-
fer of 12 electrons per molecule of ethanol and proceeds via two intermediates, acetaldehyde and ace-
tic acid. The free enthalpies of combustion associated with the individual species are described below:
C
2
H
5
OH0:5O
2
CH
3
CHO H
2
O 190:5 kJ mol
1
(15)
CH
3
CHO 0:5O
2
CH
3
COOH 268:3 kJ mol
1
(16)
CH
3
COOH2O
2
2CO
2
2H
2
O 849:8 kJ mol
1
(17)
C
2
H
5
OH3O
2
2CO
2
3H
2
O 1308:6 kJ mol
1
(18)
Comparison of these values with that of ZnS shows that ZnS should always be stable in the pres-
ence of ethanol against oxidation or reduction and therefore it could be a stable catalyst against eth-
anol oxidation. It is also important to cap ZnS nanostructures with small ligands in order to facilitate
electron transfer across the surface.
Fig. 61. Schematic of an ethanol fuel cell using ZnS nanoparticles as a catalyst and a FeCl
3
/FeCl
2
counter electrode. Reproduced
from Ref. [408]. Copyright 2010 American Chemical Society.
250 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
Fig. 61 shows a schematic of a fuel cell set up with ethanol fuel and FeCL
3
/FeCl
2
as a counter elec-
trode. Electrodes with carbon supporting ZnS nanoparticles were prepared on a copper plate and KNO
3
was used as a standard electrolyte.
In this setup, oxidation of ethanol took place only in the presence of ZnS particles. Electrodes with-
out ZnS particles generated no current. Repeated ethanol addition increased the owing current step-
wise, as seen from Fig. 62. A pure carbon electrode also led to a sharp current peak (adsorption of
ethanol), but not to continuous decomposition, while with an ITO blank electrode, these effects were
much smaller [408].
This work establishes the usability of ZnS nanoparticles as electrocatalysts for direct conversion of
ethanol in fuel cells. Such catalysts are the need of the hour, since ethanol can be made from a broad
variety of organic materials by fermentation.
5.5. UV-light sensors
5.5.1. Unique advantages of ZnS nanostructrues
UV light is one of the electromagnetic radiations with a wavelength shorter for visible radiation,
but still longer than X-rays. Because the spectrum consists of electromagnetic waves with frequencies
higher than those that humans identify as the violet color, it is named UV light. The irradiation is in the
range 10400 nm, having energies from 3 eV to 124 eV, including VUV (some denitions suggested
that UV light does not include light with a wavelength <200 nm). The UV radiations emitted from
Sun are in the range 200400 nm. According to existing classication of UV light with respect to its
wavelengths, the longest wavelength of the UV-A band (-320400 nm) can reach the Earth surface.
The molecules in sunscreens absorb most of the UV-B (-290320 nm) light and decently protect
the human skin, just as the molecules of the Earth atmosphere completely absorb the UV-C (-200
290 nm) radiation and not allow it to reach the ground. 98.7% of the UV radiation that reaches the
Earths surface is UV-A [3,409].
Many researches evidenced that UV-A light may cause a skin cancer. In our daily life, ordinary glass
is partially transparent to UV-A but is opaque to shorter wavelengths, while silica or quartz glass,
depending on quality, can be transparent even to UV wavelengths in vacuum. Ordinary window glass
passes -90% of the light above 350 nm, but blocks over 90% of the light below 300 nm [409]. UV sen-
sors measure the power or intensity of incident UV radiation. Therefore, it is timely important and
demanding to develop novel effective UV sensors which exhibit the highest UV-A sensitivity but are
blind to standard visible radiation.
As described in the Introduction, with a direct and wide bandgap of 3.72 eV and 3.77 eV for ZB and
WZ forms and a diverse range of possible structures and morphologies, added to superior chemical
Fig. 62. Short-circuit current in a fuel cell with blank ITO electrode, blank carbon electrode, and carbon/ZnS electrode after
ethanol addition. Reproduced from Ref. [408]. Copyright 2010 American Chemical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 251
and thermal stabilities, ZnS provides a novel prospective alternative for UV detectors that would par-
ticularly well serve within the UV-A band [125]. Being compared with all existing alternative materi-
als: an indirect bandgap diamond (-5.5 eV), and a direct bandgap semiconductor ZnO (-3.4 eV), ZnS
has a higher potential as the UV detector in this specic wave-band. Under the illumination with a
photon energy above E
g
of ZnS (-3.7 eV), which is close to the band-gap energy of ZnS, electronhole
pairs are generated under the illumination with the photon energy above E
g
, and one type of carriers is
trapped at the surface, enhancing the sensitivity due to increasing the density of the other carrier type.
Normally, the photoresponse will show a notable increase when the illumination light energy is above
the threshold excitation energy. For example, a photoresponse of an individual ZnS nanobelt-based
sensor increases by over three orders of magnitude upon the illumination with a 320 nm light com-
pared to its response to a visible light. The high spectral selectivity combined with the high photosen-
sitivity implies the bright prospects of using the ZnS nanobelts as visible-blind UV photodetectors in
many areas [125]. Very recently, we fabricated single-crystalline ZnSe nanobelts via an ethylenedia-
mine (en)-assisted ternary solution technique on precursor ZnSe3en nanobelts and subsequent ther-
mal treatment in pure Ar atmosphere at 500 C for 24 h. Individual ZnSe nanobelts were assembled
into nanoscale photodetectors by a microfabrication process. A discrimination ratio of about three or-
ders of magnitude between 250 nm deep UV (DUV) light and 600 nm visible light, and about two or-
ders of magnitude between 400 nm visible light and 600 nm visible light was apparent. Moreover, the
photodetectors showed high spectral selectivity and photocurrent immediate decay ratio (>99%), and
fast time response (<0.3 s), justifying the developed photodetector is not only valuable under UV-light
illumination, but is also applicable for short wavelength blue light-sensitive photodetectors (below
460 nm) [410].
ZnO is emerging as a suitable alternative to GaN in optoelectronic applications, such as light-emit-
ting diodes (LED), laser diodes, and UV photodetectors due to its wide bandgap, low cost and ease of
production [411]. Investigations on ZnO thin lms have been actively pursued in the last 50 years be-
cause of their applications as sensors, transducers and catalysts. In the last decades, research activities
on ZnO nanomaterials/nanostructures have been steadily increasing owing to their promising applica-
tions in nanoscale devices [412414]. Despite the widespread research on ZnO nanostructure photo-
detectors, their sensor performances are still needed to be improved, especially in regard of a slow
response time, both on rise and decay. For example, Lupan and co-workers reported that single-crys-
talline ZnO nanorods might be grown on glass substrates by a hydrothermal method. A single ZnO
nanorod-based photodetector was fabricated using the in situ lift-out technique in a focused-ion-
beam (FIB/SEM) instrument. The UV response time was found to be around a few minutes and the
authors explained that it had been controlled by the adsorption and photodesorption of ambient
gas molecules, such as O
2
or H
2
O [415].
Most of the experiments indicated that the role of oxygen vacancies in the ZnO system is predom-
inant for the electronic properties, similar to the bulk systems [63]. A schematic of the photoconduc-
tion mechanism in the presence of a high density of hole-trap states at the ZnO nanowire surface is
shown in Fig. 63 [411]. When illuminated with a photon energy which is larger than the semiconduct-
ing bandgap (hm > Eg), electronhole pairs are photogenerated (hm ?e

+ h
+
) in the semiconductor. It
is proposed that the holes migrate to the surface due to band-banding, leaving behind unpaired elec-
trons, which increase the carrier concentration and conductivity of the material (Fig. 63a). Pure ZnO is
usually a n-type semiconductor at RT. When oxygen molecules are adsorbed on the oxide surface, they
capture the free electrons present in the n-type oxide semiconductor [O
2
(g) e

2
(ad)[ generating
a low-conductivity depletion layer near the surface (Fig. 63b). When this system is subjected to UV
illumination, photogenerated holes migrate to the surface and combine with oxygen [O

2
(ad)
h

O
2
(g)[. In this way, oxygen is photodesorbed from the surface, leaving behind unpaired electrons
in the nanowire that contribute to the photocurrent (Fig. 63c). The unpaired electrons either go to the
anode or recombine with holes generated by readsorption of oxygen molecules at the surface [411].
This interesting oxygen induced hole-trapping mechanism, however, need to be veried further by
additional experimental and theoretical work. Recent progress on nanostructure photodetectors indi-
cated that other n-type metal-oxide nanostructures have a similar photoconduction mechanism and
sensor performances with ZnO. A comprehensive review on the state-of-the-art research activities
in the eld of 1D metal-oxide nanostructure photodetectors can be found in Ref. [63]. It mainly focuses
252 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
on the metal oxide 1D nanostructures such as ZnO, SnO
2
, Cu
2
O, Ga
2
O
3
, Fe
2
O
3
, In
2
O
3
, CdO, CeO
2
, and
their photoresponses, but not ZnS.
On the other hand, the direct bandgap makes ZnS a promising material for fabricating optoelec-
tronic devices. Being compared with other existing materials, such as indirect bandgap diamond
(-5.5 eV), ZnS has a higher potential as a UV detector in this specic wavelength regime.
In this section, we provide a comprehensive review on the state-of-the-art research activities in the
eld of ZnS nanostructures-based UV-light sensors, including individual ZnS nanobelts-based UV-light
sensors and multiple ZnS nanostructures-based UV-light sensors. The sensors were prepared by a
microfabrication process, e.g. optical lithography techniques with the assistance of a pre-designed
mask and electron beam deposition followed by a lift-off process. Subsequently our latest works on
microscale ZnS nanobelts-based UV-light sensors will be introduced in detail, these were fabricated
via a simple CVD and electron beam deposition method without using any lithography techniques.
5.5.2. Individual ZnS nanostructures-based UV-light sensors
As discussed in the section related to fundamental properties of ZnS nanostructures, the structure
and crystallography within a given ZnS nanostructures could vary (except for perfect single-crystals)
and these can further modify the optical/electrical properties. Our experimental results on multi-
angular branched ZnS nanostructures with needles-shaped tips and single-crystalline ZnS nanobelts
possessing sharp UV-light emission demonstrated that it is possible to achieve strong UV emission
and suppression of the defect related visible emission of 1D ZnS nanostructures at RT by elaborating
the experimental processes. Note that no external S source was used and a control of the evaporation
of source materials was the only demanding parameter [3]. Being motivated by these results we have
systematically investigated ZnS nanostructures-based UV-light sensors. Here, we show that single-
crystalline ZnS nanobelts with sharp UV emission (-337 nm) at RT maybe assembled as UV-light sen-
sors. The high spectral selectivity combined with high photosensitivity and fast time response justify
the effective utilization of the present ZnS nanobelts as visible-blind UV photodetectors in different
areas [125].
The ZnS nanobelts were synthesized by chemical vapor deposition (CVD) using a simple conven-
tional horizontal tube furnace with a 36 mm inner-diameter quartz tube at 1100 C [125]. As de-
scribed in the section devoted to the CL properties, the single-crystalline ZnS nanobelts possessing
sharp UV-light emission were fabricated via selection of source materials and by controlling their
evaporation and agglomeration rates. Typical SEM images of as-synthesized single-crystalline ZnS
nanobelts reveal the nanobelts can grow up to a millimeter length. The distribution of the present
Fig. 63. A schematic of the photoconduction mechanism peculiar to a ZnO nanowire photodetector. (a) Schematic of a nanowire
photoconductor. (b) and (c) Trapping and photoconduction mechanism in ZnO nanowires. The top inset in (b) shows the
schematic of the energy band diagrams of a nanowire in dark, indicating band-bending and surface trap states. VB and CB are
the valence and conduction bands, respectively. Reproduced from Ref. [411]. Copyright 2007, American Chemical Society.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 253
ZnS nanobelts is nearly similar to the aligned ultralong ZnO nanobelts. The latter had an average
length of 3.3 mm, and were synthesized on an Au sheet through a one-step process using molten-
salt-assisted template-free thermal evaporation [365]. TEM observations reveal that ZnS nanostruc-
tures have the belt-like geometry and are transparent, and the typical widths of the nanobelts are
in the range of 200 nm to 1 lm. The highly crystalline nature of the nanobelts was further veried
by a lattice-resolved HRTEM image and ED pattern [125].
Individual ZnS nanobelts were assembled as nanoscale photodetectors using a microfabrication
process. Firstly, the belts were dispersed in ethanol and dropped on a thermally oxidized Si substrate
covered with a 300 nm SiO
2
layer. Subsequently the Cr/Au (10 nm/100 nm) electrodes were patterned
on the top of the nanobelts using optical lithography with the assistance of a pre-designed mask and
electron beam deposition followed by a lift-off process. An optical micrograph of an individual ZnS
nanobelt-based UV sensor array composed of individual structures is shown in Fig. 64a. Fig. 64b
and c present a ZnS nanobelt congured as a photodetector and a representative SEM image of a nano-
belt device, respectively The Cr/Au (10 nm/100 nm) interdigitated electrodes with a -4 lm separation
were deposited on the nanobelts dispersed on a SiO
2
/Si substrate. The gap between electrodes could
be tuned while using different masks and the uncovered part of the belts was exposed to the light. The
currentvoltage (IV) characteristics of the ZnS nanobelt-based sensors were measured in air and RT
using an Advantest picoammeter R8340A and a dc voltage source R6144 under illumination with light
of different wavelengths or in dark conditions. Fig. 64d shows the comparative IV characteristics of a
ZnS nanobelt photodetector illuminated with a light of 320 and 600 nm wavelengths. These reveal
that the present ZnS nanobelts can be used as visible-blind UV-light photodetectors [125]. The spec-
tral responsivity (R
k
) measures the inputoutput gain of a detector system, which is dened as
Fig. 64. (a) An optical micrograph of an individual ZnS nanobelt-based UV sensor array composed of individual structures. (b)
and (c) A schematic illustration of an individual ZnS nanobelt congured as a photodetector and a representative SEM image of a
single-crystalline ZnS nanobelt device. (d) Comparative IV characteristics of the ZnS nanobelt photodetector illuminated with a
light of 320 nm and 600 nm wavelengths, respectively. Reproduced from Ref. [125]. Copyright 2009, Wiley-VCH.
254 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
R
k
= Iph/P, where P is the light power irradiated on an individual nanobelt. External quantumefciency
(EQE) or quantum efciency (QE) is a quantity dened for a photosensitive device such as a photo-
graphic lm or a charge-coupled device (CCD). It relates to the percentage of photons hitting the
photoreactive surface that will produce the electronhole pairs [410]. The R
k
of the present individual
ZnS nanobelts biased at 5 V at 320 nm is -0.12 A/W. This corresponds to an EQE of around 50% if a
photoconductive gain is assumed to be 1. The efciency can be much higher at higher electric elds
due to carrier injection. The present performance is comparable with a near-UV photodetector based
on the hybrid polymer/zinc oxide nanorods prepared by low-temperature solution processes. The lat-
ter device exhibited a response of 0.18 A/W at 300 nm by applying a bias of 2 V [416].
A spectral photoresponse for a bias of 10 V at different wavelengths from 300 to 630 nm was re-
corded by using a 500 W xenon lamp, as shown in Fig. 65a. This indicates that a photoresponse in-
creases by over three orders of magnitude upon the illumination with a 320 nm light compared to
its response to a visible light and the highest sensitivity is observed at the wavelength of -300 nm.
Fig. 65b illustrates a current response of the ZnS nanobelt photodetector upon 320 nm light illumina-
tion measured for the light-on and light-off conditions with a -20 s cycle when a 10 V voltage is ap-
plied, showing that the time response (including rise and decay time) is faster than the limit of our
measurement setup (0.3 s). It is easy to note that although a photoconductivity response of the indi-
vidual ZnS nanobelt-based UV sensor is fully reversible and periodic, the photocurrent is unstable and
relatively weak [125].
By slightly altering the experimental conditions, i.e. the reaction time was prolonged to 90 min
with a minor change at reaction temperature, the width of ZnS nanobelts was increased to -1
2 lm. For comparison, individual saw-like ZnS nanostructures were also assembled into UV-light sen-
sors. The saw-like ZnS nanostructures were synthesized using the similar method but through altering
an Ar ow ux between 100 and 50 sccm(every 5 min during the heating process) [417]. For achieving
a better contact between ZnS nanostructures and Cr/Au electrodes, the width of electrodes (Cr/Au con-
tact designs) was altered from -2 lm to -10 lm. Fig. 66 shows the SEM images of three kinds of indi-
vidual ZnS nanostructure-based UV-light sensors and their corresponding performances, including IV
curves taken under 320 nm light illumination, and a time response to a pulsed incident UV light
(-320 nm) during an onoff cycle. A photocurrent in the range of 1.42.0 pA was detected when a
voltage of 10 V was applied and all sensors show a fast time response (<0.3 s) and a reproducible pho-
tocurrent. Thus the enhanced photocurrent stability and intensity of the present sensors compared to
those reported by us previously is clearly demonstrated [125,417].
5.5.3. Multiple ZnS nanostructures-based UV-light sensors
Only altering the density of ZnS nanobelts solutions with controlled concentration, a multiple ZnS
nanostructures-based UV-light sensors can be fabricated by adopting a similar process followed by
Fig. 65. (a) A spectral photoresponse of a ZnS nanobelt photodetector measured at a bias of 10 V at different wavelengths
ranging from 300 to 630 nm. (b) The reproducible on/off switching of a ZnS nanobelt photodetector upon 320 nm light
illumination measured for the light-on and light-off conditions with a -20 s cycle at a bias of 10 V. Reproduced from Ref. [125].
Copyright 2009, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 255
subsequent optical lithography with the assistance of a pre-designed mask and electron beam depo-
sition. Fig. 67 depicts two multiple ZnS nanostructures-based UV-light sensors with different structure
density and their corresponding performances. Besides similar time response (<0.3 s) and good pho-
tocurrent reproducibility, about 35 and 20 times increase compared to the individual ZnS nanobelt
sensors is apparent in Figs. 64 and 66 [125,417].
An alternative route to develop multiple ZnS nanostructures-based UV-light sensors is the direct
preparation of predesigned patterned substrates. The patterned substrates were produced using the
same fabrication process and the analogous electrodes to fabricate a sensor made of multiple ZnS
nanobelts, as shown in Fig. 68a. Different from the growth of the above-mentioned ZnS nanobelts, a
Cr/Au (3 nm/3 nm) layer was deposited on the SiO
2
/Si substrate. Au was used as a catalyst for the
growth of ZnS nanobelts while a Cr layer was used to protect the SiO
2
layer. During the guided ZnS
growth on the pre-patterned substrates the synthetic conditions within a CVD system were the same
as those during the initial synthesis of ZnS nanobelts for individual ZnS nanobelts-based UV-light sen-
sors. After the CVD system was cooled down to RT, white-colored wool-like products were found
grown selectively on the top and side surfaces of the electrodes, Fig. 68b. It is noted that the interspac-
es between electrodes were densely lled with well-aligned ZnS nanostructures. The lengths of ZnS
nanobelts were enough to merge each two neighboring electrodes in a single circuit. The Cr/Au
(10 nm/100 nm) interdigitated electrodes were repeatedly electron-beam deposited using 50 and
100 lm Au wires as a mask to cover the aligned nanobelts. We found that the nanobelt physical con-
tact to electrodes was signicantly improved with the addition of post-synthesis Cr/Au (10 nm/
100 nm) layers. Fig. 68c displays the spectral photoresponse of a multiple ZnS nanobelt-based UV sen-
sor for a bias of 10 V at different wavelengths from 230 to 630 nm. A similar spectrum was obtained
Fig. 66. SEM images of three kinds of individual ZnS nanostructure-based UV-light sensors and their corresponding
performances, including IV curves taken under 320 nm light illumination, and a time response to a pulsed incident UV light
(-320 nm) during an onoff cycle. Reproduced from Ref. [417]. Copyright 2010, Wiley-VCH.
256 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
with individual ZnS nanobelts based UV-light sensors, justifying that the present single-crystalline ZnS
nanobelts are indeed particularly valuable for newvisible-light-blind UV-light photodetectors, espe-
cially in the UV-A region. Enhanced stability and sensitivity of the present UV-light sensors compared
to the individual nanobelt-based sensors is evidenced by a conductance response under 320 nm UV-
light illumination for the light-on and light-off states (Fig. 68d). It is easy to note that photocurrent
intensity and stability are distinctly improved when a channel distance is altered from several
micrometers to 100 lm, while fast time response is still kept [125].
5.5.4. Microscale ZnS nanobelts-based UV-light sensors
The above-mentioned results demonstrated that individual and multiple ZnS nanostructures can
be used as effective UV-light sensors with high sensitivity and selectivity and fast time response. How-
ever, most of nanostructures-based functional devices and integrated systems have been fabricated by
top-down approach using a combination of lithography, etching, and deposition nowadays [61].
Although the top-down approach has exceedingly been successful in microelectronics, the technique
is complicated, time-consuming and expensive. The most efcient way to improve the ZnS photocur-
rent intensity is the utilization of a large-area single crystal thin lm. This, however, relies strongly on
the growth technique and is costly. In fact, no UV-light sensors using ZnS single crystal thin lms have
been reported to date. An alternative approach is the direct integration of one-dimensional nanostruc-
tures into a thin-lm-like microchip. Very recently, we have developed an effective and low-cost
method to achieve high-performance visible-blind microscale ZnS nanobelts-based UV-light sensors
through enlarging the nanobelts surface areas exposed to light without using any lithography tech-
niques [417].
Microscale ZnS nanobelts-based UV-light sensors of a required size were fabricated by a simple
CVD and electron beam deposition method using Au microwire meshes as masks. 5 nm Au layer-
coated quartz glasses instead of Si wafers were used as deposition substrates. The reaction time
was prolonged to 2 h while keeping all other deposition parameters the same as for the synthesis
of previously reported ZnS nanobelts in Fig. 66. Subsequently, a quartz glass substrate coated with
a layer of ZnS nanobelts was transported to an electric gun deposition system (ULVAC UEP-3000
2C), the Cr/Au (10 nm/100 nm) electrodes were deposited using 100 lm and 50 lm Au wires as a
mask [417].
Fig. 67. Two multiple ZnS nanostructures-based UV-light sensors with different structure density and their corresponding
performances. Reproduced from Ref. [417]. Copyright 2010, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 257
Fig. 68. Another multiple ZnS-nanobelts-based sensors. (a) An optical image of predesigned patterned substrates for the sensor
production. (b) An optical image of the as-grown devices on the the substrate shown in (a). The interspaces between electrodes
were densely lled with well-aligned ZnS nanobelts. (c) and (b) The corresponding sensor performances Reproduced from Ref.
[125]. Copyright 2009, Wiley-VCH.
258 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
The distances between Cr/Au (10 nm/100 nm) electrodes can be altered by changing the diameter
and position of Au wires. In our experiments, we used 100 lm and 50 lm Au wires as a mask, which
means that the distances between electrodes are 100 lm and 50 lm, correspondingly. Considering a
light spot from an optoelectronic detector of a size of 2 2 mm
2
, microscale ZnS nanobelts-based UV-
light sensors with -1 mm and -2 mm electrode lengths were fabricated. Fig. 69 shows an optical
micrograph the sensor performance while using 50 lm electrode distance and -1 mm electrode
length. The optical micrographs and SEM images reveal that a typical belt length is several hundred
micrometers. Some of them may even reach a millimeter in length. The typical width and thickness
of the belts are -5 lm and -100 nm, respectively. Most of the nanobelts are lying on the substrate
providing a good physical contact. As shown in Fig. 69b, when the sensor is illuminated with a
320 nm light and is biased at 30 V, a photocurrent is up to -2 nA. A remarkable improvement was
achieved compared with the above-mentioned individual and multiple ZnS nanobelts based UV-light
sensors. Fig. 69c illustrates very good reproducibility and stability over 10 cycles of 320 nm light-on
and lightoff states during 1000 s at an applied voltage of 20 V. No notable photocurrent degradation
was observed [417]. By using a 350 MHz Tektronix (TDS 5000B) oscilloscope with 50 X impedance
and a 320 nm light pulse chopped at a frequency of 100 Hz, a detailed response time was measured,
as shown in Fig. 70. The rise and fall times were measured to be around 2.57 ms and 1.99 ms,
respectively.
Fig. 69. (a) An optical micrograph of a microscale ZnS nanobelts-based UV-light sensor and the sensor performance using a
50 lm electrode distance and -1 mm electrode length. (b) Typical IV curve when the sensor is illuminated with a 320 nm light
and is biased at 30 V. (c) Time dependent photocurrent response to 320 nm UV-light illumination with a period of -50 s at an
applied voltage of 20 V. The part 1 inserted in (a) represents ZnS nanobelts covered with Cr/Au electrodes, whereas the part 2
shows blank ZnS nanobelts. Reproduced from Ref. [417]. Copyright 2010, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 259
As discussed at the beginning of this section a photoelectrical transport in nanostructures with high
surface-to-volume ratio (SVR) and large surface areas exposed to light is usually signicantly affected
by the surface absorption or desorption processes. For example, several researches showed that trap-
ping at the surface states had drastically affected the transport and photoconduction properties of ZnO
nanowires due to high SVR, and the photoresponse of ZnO is generally governed by surface adsorption
and desorption of oxygen or water molecules [63,411,415,418]. The photoconduction mechanism gen-
erated on a ZnO nanowire photodetector is sketched in Fig. 63. Under this mechanism, the photocur-
rents of a ZnO nanowire photodetector are signicantly different while measured under air and
vacuum conditions. For example, the photocurrent increases by a factor of more than 10 after pump-
ing down for 5 min. It is stated that the increased trapping is due to oxygen desorption as also man-
ifested in IV measurements of single ZnO nanowire devices performed in vacuum (P < 10
4
Torr)
[411]. As shown in Fig. 71a, only a slight uctuation is visible for ZnS sensor under vacuum. The stable
photocurrent implies that the dominated mechanism relates to a bulk effect or that the present ZnS
nanobelts with high SVR and large surface areas exposed to light have a chemically stable surface
[3]. Fig. 71b shows a typical time dependent photocurrent response to different measurement temper-
atures (RT, 60 C and 150 C) and vacuumconditions under 320 nm UV-light illumination for the light-
on and light-off states at an applied voltage of 10 V for a ZnS UV-light sensor. From the curve, it is
noted that all the photocurrents are stable and the sensor has a fast time response. The photocurrent
increases about three times with an increase in measurement temperature from RT to 150 C, which is
probably due to thermal-enhanced photoemission. The photoelectrical behaviors can be fully recov-
ered after changing the measurements conditions to air and RT again. Moreover, all the devices exhibit
low dark and immediate currents after turning off the UV-light. They also have a high stability in air or
vacuum even at 150 C [417]. These results highlight the effective utilization of the present sensors to
work in the UV-A band. Also this low-cost method will be general for the fabrication of other inorganic
semiconductor nanostructures based optoelectronic devices.
The literature documents a large number photodetectors made of nanostructures (with controlled
shapes and sizes) Table 21 lists some representative sensors and their key performances. Although the
measurement conditions are variable (e.g. different bias, etc.), it is easy to note that ZnS nanostruc-
tures guarantee the highest potential for visible-blind UV-light sensors working in the UV-A band. This
is evidenced by a fast time response, low dark current, high ratio of photocurrent and dark current,
high photocurrent immediate decay ratio (not shown here), and high spectral selectivity, etc.
5.6. Gas sensors
ZnS has rarely been used in gas sensors. This is because of its low carrier density and irreversible
reaction with oxygen to convert into ZnO at high temperatures. However, these two problems could
be solved by using ultraviolet UV illumination. The illumination enhances the modulation of conduc-
tance by adsorbed oxygen. The three typical gas-sensing examples are demonstrated below.
Fig. 70. The transient response of a microscale ZnS nanobelts photodetector illuminated with a 320 nm light at a bias of 100 V
along with the reference signal of the chopped light at 100 Hz. Reproduced from Ref. [417]. Copyright 2010, Wiley-VCH.
260 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
The measurement of gaseous oxygen concentration is desirable in environmental monitoring,
chemical process controlling, and personal safety. In semiconductor industry, a slight amount of oxy-
gen in inert gas results in surface degradation of materials by oxidation, therefore, the in situ precise
measurements of oxygen concentration in inert gas is highly required. Wang and co-workers fabri-
cated ZnS nanobelts using a thermal evaporation method and realized oxygen sensing by using an
individual ZnS nanobelt [129]. The response of ZnS nanobelts to different oxygen pressures without
or with UV illumination was measured. The measurement was rst performed under an oxygen pres-
sure of 3 10
3
Pa. The current increased from 0.21 to 1.71 nA within 0.4 s and then in a relative slow
rate to 2.30 nA in 2.65 s. When the light was turned off, the current recovered to the original dark level
within 0.2 s. The fast response and rapid recovery arose from the generation and recombination of the
electronhole pairs in ZnS. The slow increase in response time was attributed to the oxygen desorp-
tion process on the ZnS nanobelt surface. From the conductance of ZnS nanobelt as a function of oxy-
gen pressure in logarithmic scale, it can be seen that the dark level conductance of ZnS nanobelt varied
slightly from 9.8 10
12
to 1.08 10
11
S without UV illumination, and decreased from 1.13 10
10
to 0.13 10
10
S as the oxygen pressure increased from 3 10
3
to 1 10
5
Pa under UV illumination.
From the above results, the conductance of the ZnS nanobelt is roughly proportional to the logarithm
of the oxygen pressure, which indicates that these nanobelts can be promising oxygen sensors [129].
The authors also gave an explanation of room-temperature oxygen sensing by considering a com-
plex process of electronhole generation, recombination, and adsorption on the surface of the
Fig. 71. Time dependent photocurrent responses of a microscale ZnS nanobelts-based UV-light sensor under 320 nm UV-light
illumination for the light-on and light-off states at an applied voltage of 10 V in air and vacuum conditions (a), and at different
temperatures (RT, 60 C and 150 C) and vacuum conditions (b). Reproduced from Ref. [417]. Copyright 2010, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 261
Table 21
Representative parameters of nanostructure-based UV-light sensors and their key performances. A part of the data reproduced
from Ref. [63].
Photodetectors Nanostructures Devices Light of
detection
Bias
(V)
Dark current
or
conductance
Photocurrent
or
conductance
Rise time Decay
time
Ref.
ZnS NB Resistor 320 nm 10 <10
14
A 0.87 pA <0.3 s <0.3 s [125]
NB Resistor 320 nm 10 <10
14
A 12 pA <0.3 s <0.3 s [125]
NB Resistor 320 nm 10 <10
14
A 1.05 nA -2.57 ms -1.99 ms [417]
ZnO NW Resistor 390 nm;
6.3
40 mW/
cm
2
5 110 nA 100 lA [411]
NW Resistor 350 nm;
50 nW
48 mW/
cm
2
5 40 lA for
50 nW/cm
2
0.7 s 1.4 s [418]
NW Resistor 340 nm 1 20 nA 170 s 300 s [419]
NW Resistor 254 nm
(0.1 W/
cm
2
)
0.25 100 nA 580 nA
(254 nm)
Tens of
Seconds
[420]
366 nm
(0.1 W/
cm
2
)
700 nA
(366 nm)
NR FET 254 nm 0.2 2.4 lA 30 min [421]
SnO
2
NW Resistor 370 nm 37.5 kX 12.5 kX [422]
NW FET 254 nm 0.05 0.66 nS 760 nS <0.1 s [423]
NB Resistor 254 nm; 5 0.4 nA 80 lA (air) [424]
10 mW 900 lA
(vacuum)
532 nm; 1.4 nA 57 nA <1 s <1 s
0.135.1
mW
b-Ga
2
O
3
NW Resistor 254 nm; 7
w
8 Several pA Several nA 0.22 s 0.09 s [425]
(15 pA) (10 nA)
In
2
O
3
NW FET 254 nm 0.3 290 nA
(254 nm)
10 s [426]
365 nm 33 nA
(365 nm)
CeO
2
NW lm Resistor 254 nm;
7 W
5 22.8 nA (air) 0.25 nA (air) 300 s
(air)
[427]
0.35 pA
(H
2
O)
0.44 nA
(H
2
O)
2 s (H
2
O)
ZnSnO
3
NW Resistor UV light 0.3 nA 162 nA (UV) 20 s [428]
Green laser 6.6 nA
(green)
ZnGa
2
O
4
NW Resistor UV light 30 8.5 PA 1 nA [429]
RuO
2
/TiO
2
Core/Shell NW Resistor 256 nm 18.5 lA 19.4 lA 307 s 437 s [430]
GaN NW FET 365 nm 1 1.3 nS 23.4 nS [431]
254 nm 258.3 nS
NW Resistor 325 nm 3 -0.7 mA -0.9 mA [432]
BR pn
junctions
Resistor 365 nm 1 9 nA 18 nA [433]
254 nm 17 nA
a-Si
3
N
4
NW Resistor 254 nm 20 0.25 nA 1.5 nA -1 s -1 s [434]
ZnSe NW Resistor UV light 0.1 29 nA 2.7 lA <1 s <1 s [435]
NB Resistor 400 nm 30 <10
14
A 1.7 pA <0.3 s <0.3 s [410]
262 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
nanobelts. Oxygen molecules physically adsorbed on the surface of the nanobelts trap electrons in the
belt and convert to chemisorbed oxygen ions, resulting in a depletion layer which decreases the con-
ductance of the nanobelts. When a ZnS nanobelt is exposed to UV illumination, the carrier density in-
creases because of the electronhole pairs generation. The chemisorbed oxygen ions combined with
the photoexcited holes.This resulted in desorption of oxygen species on the surface of the ZnS nano-
belt. Both of the processes enhance the conductance of the belt. In addition, without UV illumination,
only thermally excited holes are available to liberate the adsorption of oxygen from the surface of ZnS
nanobelt; the low carrier concentration in ZnS limited the modulation of conductance by oxygen. This
can explain the slight change of the conductance without illumination [129].
Based on the same mechanism, Liu and co-workers synthesized ZnS microspheres by a hydrother-
mal process with the tungstosilicate acid (TSA) ion as a soft template, and studied the oxygen-sensing
properties [436]. The detailed morphology characterization of microspheres indicates that it consists
of a large quantity of micrometer-scale spheres with an average diameter of -1.8 lm. The high-mag-
nication SEM image reveals that these particles are spherical and tend to form ordered structures,
and the surface of spheres is composed of plenty of nanoparticles with a size of about 2060 nm.
To investigate the oxygen-sensing properties of ZnS microspheres, electrically contacted ZnS micro-
spheres were fabricated by dropping sample solutions onto 15 nm thick comblike gold electrodes
on Si substrates with a 400 nm oxide layer on top. The separation between two electrodes was about
25 mm. Fig. 72 shows the response of a ZnS microsphere device to different oxygen pressures with or
without UV illumination at room temperature. The measurement was rst performed under an oxy-
gen concentration of 600 ppm (Fig. 72a). The current increased from 0.50 to 0.57 nA in 5 s. When the
oxygen concentration was decreased to 50 ppm, the current increased to 0.73 nA (Fig. 72b). When the
light was turned off, the current recovered to the original dark level within several seconds. Fig. 72
also shows three-cycle switching between low and high-conductance states with no obvious decay
of conductance. It is evident that the microspheres can be reversibly switched to possess low and high
conductance [436].
Cheng and co-workers reported the oriented ZnS nanobelts growth on a Si substrate using H
2
-as-
sisted thermal evaporation under moist gas conditions. These nanobelts were used for detecting H
2
gas [127]. Fig. 73 shows the response of the current owing through the nanobelts at the working tem-
perature of 300 C as the H
2
concentration was cycled from 0 to 500 ppm. Fig. 73a is the response of
the oriented ZnS nanobelts to H
2
and N
2
pulses. It can be seen that the current increases without obvi-
ous hysteresis and the lowest detectable range for the current change is at the concentration of
50 ppm of H
2
. The H
2
gas sensitivity was observed to increase rapidly as the concentration of H
2
in-
creased within the range of 50200 ppm. Fig. 73b is the response time of conductance toward a
detectable H
2
concentration of 100 ppm, which indicates that the H
2
gas sensitivity is very high. It
is well known that the fundamental sensing mechanism of semiconductor based gas sensors relies
on a change in electrical conductivity due to the interaction between the surface complexes such as
O

, O
2
, reactive chemical species (S
2
), and gas molecules which are detected. Here the oriented
Fig. 72. Three cycles of switching between low- and high-conductance states in the presence of oxygen concentrations of
600 ppm (a) and 50 ppm (b) and an applied bias voltage on the microspheres of 10 V. Reproduced from Ref. [436]. Copyright
2009, Wiley-VCH.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 263
ZnS nanostructures have a high surface area and thus provide sufcient ionizable sub-surface oxygen
adsorbed states. When the ZnS nanobelts are exposed to air, the oxygen molecules are adsorbed on the
surface as negatively charged ions by capturing free electrons from the semiconductor ZnS, thereby
creating a depletion layer with a low conductivity near the nanobelts surface. Since the thickness of
nanobelts is very small, the nanobelts create a signicant amount of surface acceptor states, leading
to a high resistance in the normal state without H
2
. The reason for the enhancement in conductance
of the nanostructures in H
2
is due to the reaction on the material surface or removal of chemisorbed
oxygen from the nanostructure surface by H
2
. Oxygen in air can be adsorbed on the surface of a semi-
conductor sensor as different species (O
2
, O
2
, and O

), which may capture electrons. Combustible


gases (e.g. H
2
) react with chemical species like S
2
(s) and O

(s) to form H
2
S and H
2
O, respectively,
resulting in a consequent increase in the conductance.
Carbon tetrachloride, one of important volatile organic compounds, has received much attention in
recent years, since it is widely used as organic solvents, a cleanout and dry-clean reagent. The increas-
ing usage of carbon tetrachloride in industry leads to polluting environment via gaseous and aqueous
wastes. Among the highly volatile organic compounds, carbon tetrachloride is considered as a toxic
and carcinogenic substance, as it could damage liver as well as kidney, even at low concentration lev-
els. In addition, it is a precursor of notorious ozone-depleting compounds chlorouorocarbon 11 (CFC-
11) and 12 (CFC-12), which can ruin the atmosphere and seriously endanger our health. Consequently,
much work is devoted to detect and monitor this toxic compound. Various methods have been estab-
lished such as gas chromatography [437], spectrophotometry [438], high performance liquid chroma-
tography and chemiluminescence [439,440].
In recent years, there has been a growing interest in cataluminescence (CTL), mainly owing to its
unique advantages of high sensitivity, rapidity, simplicity and low cost of instrumentation and main-
tenance. Xu and co-workers fabricated the nanosized ZnS by a simple ultrasonic assisted method.
Strong CTL emission was observed when carbon tetrachloride vapor passed through the surface of
nanosized ZnS in the air [441]. Based on this phenomenon, a gas sensor was developed. Under the
optimized conditions, the linear range of the CTL intensity versus the concentration of carbon
Fig. 73. (a) The response of ZnS nanostructures to H
2
at a working temperature of 300 C to H
2
and N
2
pulses. (b) Conductance
response time towards 100 ppm H
2
. Reproduced from Ref. [127]. Copyright 2008, Institute of Physics.
264 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
tetrachloride was 0.4114 lg mL
1
, with a correlation coefcient of 0.9986 and a limit of detection of
0.2 lg mL
1
. The relative standard deviation for 5.9 lg mL
1
carbon tetrachloride was 2.9% (n = 5).
There was no or weak response to common foreign substances including methanol, ethanol, benzene,
acetone, formaldehyde, acetaldehyde, dichloromethane, xylene, ammonia and trichloromethane.
Thus, for the commonly used organic solvents, the ZnS sensor shows high selectivity for carbon tetra-
chloride. There was no signicant change of the catalytic activity of the sensor for 40 h over 4 days,
with a relative standard deviation of less than 5% by collecting the CTL intensity every hour. The pro-
posed method was simple and sensitive, with a potential of detecting carbon tetrachloride in environ-
ment and industry grounds.
According to widely accepted theory about CL reaction, excited intermediates would be formed
during the catalytic oxidation of carbon tetrachloride on the surface of nanosized ZnS. Furthermore,
the luminescence emission could be obtained due to the excited intermediates falling to ground state.
ZnS is a well known N-type semiconductor and uorescence material. During the reactions, the free
electrons and free holes could be released and they recombined through a certain energy level in a
ZnS crystal with light emission. It was difcult to conrm the exact luminescence processes, the inter-
mediates or recombination radiation species.
From the above examples, one can see that it is still a challenge to employ ZnS nanostructures as
gas sensors. In the future, one should construct core/shell or other complex ZnS nanostructures to en-
hance the carrier density and degrade the effect of adsorbed surface oxygen, thus the improved sen-
sitivity and fast response time can be expected.
5.7. Chemical sensors
Because of the small size of nanowires and nanoparticles, only a few chemical molecules are suf-
cient to change the electrical or optical properties of the nanostructured sensing elements. This al-
lows the detection of a very low concentration of chemical vapors. As an important optoelectronic
semiconductor, ZnS also has this potential application, as discussed below.
Wang and co-workers have synthesized compound semiconductor/insulator (ZnS/silica) core/shell
nanocables using vaporliquidsolid growth, and used them to fabricate single nanowire-based FETs
[182]. After chemical modication, amine- and oxide-functionalized nanocables exhibit linear pH-
dependent conductance. The FET of ZnS/silica core/shell nanostructures are attractive to chemical sen-
sors due to (i) ZnS excellent and reproducible electronic characteristics; (ii) the self-assembled silica
shell natural insulatating and oxidation-protecting behavior on the wire surfaces (iii) deep knowledge
about the chemical modication of silica surfaces.
Chemical sensing in a solution was carried out by monitoring electrical conductance through a sur-
face modied ZnS/silica nanocable device during buffer solution additions with different pH values.
The conductance of the ZnS/silica nanocable-based FET was modulated by an applied gate, and the
surface of the silica shell was modied with 3-aminopropyltriethoxysilane (APTES) to provide a sur-
face that could undergo protonation and deprotonation. The changes in the surface charge could
chemically gate the nanocable-based transistors. The concept of a pH nanosensor is illustrated in
Fig. 74a. Measurements of the conductance as a function of time and pH demonstrate that the conduc-
tance increases in steps in accord with the pH values, as shown in Fig. 74b. A representative curve of
the conductance as a function of pH suggests that a modied ZnS/silica nanocable can be functional-
ized as a nanoscale pH sensor since its pH dependence is linear (Fig. 74c). The proposed mechanism for
pH sensing is described as follows. Covalently linking APTES to the oxide surface of a nanocable results
in a surface terminating in both NH
2
and SiOH groups (Fig. 74a), which have different dissociation
constants, pKa. At a high pH value, SiOH is deprotonated to SiO on the surface of the silica shell and
acts as a negative gate, which depletes electron carriers in the n-type ZnS core and decreases the con-
ductance. At low pH, the NH
2
group is protonated to NH

3
and acts as a positive gate, which corre-
spondingly causes an increase in conductance. The observed linear response can be attributed to an
approximately linear change in the total surface charge density due to the combined acid and base
behavior of both surface groups.
Semiconductor nanocrystals (NCs) have a very good photostability, continuous absorption spectra,
efcient, narrow, and tunable emission, which have been widely exploited for applications in
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 265
biological imaging and in single particle tracking studies. The recent development of water-soluble
CdSe NCs for sensing target analytes using uorescence resonant energy transfer (FRET) via a signal
transduction mechanism renders them as microenvironment sensing probes. However, the previous
studies, have not demonstrated reversible and ratiometric chemical sensing using uorescent NCs.
Snee and co-workers fabricated a reversible NC-based uorescent sensor by conjugating a dye with
an equilibriumresponse to an analyte to the surface of a CdSe/ZnS NC [442]. They observed a ratiomet-
ric response to pH by properly controlling energy transfer between the NC and dye. This engenders a
general method for the development of ratiometric NC sensors, thus providing a means for detecting
analytes with high precision, irrespective of changes in excitation intensity, wavelength, or collection
efciency. The method overcomes several difculties encountered with dye-only-based ratiometric
nanoscopic sensing systems.
Besides the above-mentioned FRET mechanism, the photoinduced electron transfer (PET) mecha-
nism has also been used to control the excited state properties of quantum dots (QDs), although
the number of sensors designed by this approach is signicantly fewer than for FRET. Callan and
co-workers have shown that a ferrocenyl urea receptor, attached to the surface of a pre-formed
CdSe/ZnS QD via a dihydrolipoic acid linker, quenches QD uorescence [443]. Interestingly, the addi-
tion of uoride to the resulting QD-conjugate switched uorescence back on, by modulating the rate
of PET between the ferrocene units and the QD. The observed selectivity was good for uoride over
other tested monovalent anions, such as the chloride, bromide, acetate, hydrogen sulfate, perchlorate
and dihydrogen phosphate.The authors believe that the quenching of QD uorescence upon surface
attachment is an electron transfer process occurring between the ferrocene and the excited QD. The
redox potential of this type of receptor is considerably changed upon anion recognition, which could
be sufcient to modulate the electron transfer process responsible for quenching and switching on
Fig. 74. (a) Schematic illustrating a nanocable-based sensor for pH detection. The APTES-modied ZnS/silica nanocable changes
in the surface charge state with pH due to NH
2
and SiOH groups. (b) Real-time detection of the conductance for an APTE-
modied nanocable. (c) Plot of the conductance versus pH; the points are experimental data, and the solid line is a linear t
through the experimental data. Reproduced from Ref. [182]. Copyright 2007 American Chemical Society.
266 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
uorescence. In addition, it has been shown in organic dye based PET sensors that cations placed in the
path of PET can change its rate. Therefore, it is also possible that the placement of a uoride ion in the
path of PET results in a similar effect.
Copper (Cu) is one of the common metals and essential elements for many living organisms, and
also becomes toxic at high levels. It causes liver damage in infants. So the determination of Cu in
the environment and biological samples is of tremendous interest and importantance for analytical
chemists. Narayanaswamy and co-workers reported the synthesis of l-cysteine-capped ZnS QDs and
used them for Cu detection [444]. Cysteine is a water-soluble amino acid and used as a capping agent
for ZnS QDs. The surface modication of ZnS QDs with cysteine prevents the aggregation of nanopar-
ticles and makes them available for the interaction with the target materials. Furthermore, it increases
the emission quantum yields of QDs and also stabilizes the nanoparticles. These functionalized nano-
particles were used as a uorescence sensor for Cu (II) ion, involved in the uorescence quenching. The
optimum uorescence intensity was found to be at pH 5.0 with a nanoparticle concentration of
2.5 mg L
1
. The uorescence of l-cysteine-capped ZnS QDs has more sensitivity to Cu
2+
ion than to
Ag
+
and Fe
2+
ions in a buffer solution and cations such as K
+
, Mg
2+
, Ca
2+
, Na
+
, Mn
2+
, and Zn
2+
. have
no effect. The limit of detection of l-cysteine-capped ZnS QDs for Cu (II) ion is higher than that of
the l-cysteine-capped CdSe/CdS QDs. But the synthesis procedure of the former is a one-step method
using non-toxic and low cost materials compared with the latter. In addition, l-cysteine-capped ZnS
QDs are more stable than other surface modied QDs. Therefore the development of this sensor is sim-
ple for the detection of the Cu (II) ion. The quenching mechanism was studied and the results showed
the existence of both static and dynamic quenching processes.
Nitroaromatic compounds are highly explosive and environmentally deleterious substances that
have been of serious concern, and therefore the development of robust and sensitive platforms for
their real-time analytical detection has attracted considerable research efforts in recent years
[445,446]. Mn
2+
-doped ZnS nanocrystals with an amine-capping layer have been synthesized and used
for the uorescence detection of ultratrace 2,4,6-trinitrotoluene (TNT) by quenching the strong orange
Mn
2+
photoluminescence [447]. The organic amine-capped nanocrystals can bind TNT species from
solution and atmosphere by the acidbase pairing interaction between electron-rich amino ligands
and electron-decient aromatic rings. The resultant TNT anions bound onto the amino monolayer
can efciently quench the Mn
2+
photoluminescence through the electron transfer from the conductive
band of ZnS to the lowest unoccupied molecular orbital (LUMO) of TNT anions. The amino ligands pro-
vide an amplied response to the binding events of nitroaromatic compounds by the 2- to 5-fold in-
crease in quenching constants. Moreover, a large difference in quenching efciency was observed for
different types of nitroaromatic analytes, dependent on the afnity of nitro analytes to the amino
monolayer and their electron-accepting abilities. The amine-capped nanocrystals can sensitively de-
tect down to 1 nM TNT in solution or several parts-per-billion of TNT vapor in atmosphere. The
ion-doped nanocrystal sensors show a remarkable air/solution stability, high quantum yield, and
strong analyte afnity and, therefore, are well-suited for detecting the TNT traces and distinguishing
different nitro compounds.
5.8. Biosensors
A biosensor is a device for the detection of an analyte that combines a biological component with a
physicochemical detector component. There are many potential applications of biosensors of various
types. The main requirements for a biosensor approach are the identication of a target molecule,
availability of a suitable biological recognition element, and the potential for disposable portable
detection systems preferable for sensitive laboratory-based techniques.
Wang and co-workers used the ZnS/silica core/shell nanocables for label-free, real-time, and sen-
sitive detection of biological species after biotinylation of the nanocables surface since the silica shell
makes the receptor linkage straightforward [182]. The silica shell of a nanocable was functionalized
with biotin. The well-characterized ligand-receptor binding of SA-biotin (dissociation constant on
the order of 10
15
M) may be utilized for specic recognition (Fig. 75a). The biontinylation reagent
was used in order to react with the APTES-modied ZnS/silica nanocable. The conductance-time mea-
surement shows that the conductance of the biotin-modied nanocable decreases rapidly and
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 267
considerably to a constant value upon addition of the SA solution, and this conductance value remains
even after the addition of BSA solution (Fig. 75b). Due to the small dissociation constant and small dis-
sociation rate for biotin-SA, the absence of a conductance change with the addition of BSA was ob-
tained. This work demonstrates that 1D ZnS nanocable-based biosensors are capable of specically
detecting proteins in real-time.
Semiconductor colloidal nanoparticles are under intensive study for potential applications in mate-
rial synthesis, in multiplexed bioassays and in ultrasensitive optical detection and imaging. Colloidal
semiconductor nanoparticle QDs are luminescent inorganic uorophores (comprised of following pro-
cesses: 1 absorption and excitation, 2 energy transfer and 3 emission) that have the potential to
overcome some of the functional limitations encountered by organic dyes in uorescence labeling
applications. Problems with organic uorescent markers include narrow excitation bands and broad
emission spectra. This can make detection of multiple light emitting probes difcult, because of spec-
tral overlap, low resistance to chemical degradation and photodegradation. QD labels exhibit size-tun-
able narrow-band luminescent emission and high resistance to photodegradation. To be a suitable
biosensor, the nanoparticles should have high luminescent efciency and proper surface groups for
coupling with biomolecules. Recently, scientists have concentrated on developing sensors for some
target analytes, wherein QDs are used as electron donors for uorescence resonance energy transfer
(FRET) between the QDs (donor) and an acceptor molecule. Besides FRET, many advantageous proper-
ties of QDs have been exploited for the development of sensors based on the change in the emission
wavelength, voltage or uorescence intensity [448450].
Recently Rhee and co-workers reported that CdSe/ZnS coreshell quantum dots could synthesized
and conjugated with enzymes, glucose oxidase (GOD) and horseradish peroxidase (HRP) [451]. The
complex of enzyme-conjugated QDs was used as QD-FRET-based probes to sense glucose. The QDs
were used as an electron donor, whereas GOD and HRP were used as acceptors for the oxidation/
reduction reactions involved in oxidizing glucose to gluconic acid. Electron transfer between the redox
enzymes and the electrochemical reduction of H
2
O
2
(or O
2
) occurred rapidly, resulting in an increase
of the turn-over rate of the electron exchange between the substrates (e.g. glucose, H
2
O
2
and O
2
) and
the enzymes (GOD, HRP), as well as between the QDs and the enzymes. The transfer of non-radiative
energy from the QDs to the enzymes resulted in the uorescence quenching of the QDs, corresponding
to the increase in the concentration of glucose. The quenching process was based on the transfer of
electrons from the QDs to enzymes (GOD, HRP), which catalyze the oxidation/reduction reactions of
glucose and all of the components involved in the process. The linear detection ranges of glucose con-
centrations were 05.0 g/l (R = 0.992) for the volume ratios of 10/5/5, 0.25.0, g/l (R = 0.985) for the
volume ratios of 10/5/3 and 1.05.0 g/l (R = 0.982) for the volume ratios of 10/5/0. Temperature
(2937 C), pH (610) and some ions (NH4
+
, NO
3
, Na
+
, Cl

) had no interference effect on the glucose


measurement. Tuning inconjugating carboxyl groups attached on the surface of QDs and amine groups
of enzymes as well as the high potential of QDs for electron transfer to neighbor molecules led to high
Fig. 75. Real-time electrical detection of specic biological recognition. (a) Schematic diagram of a biotin-modied ZnS/silica
nanocable (left) and subsequent binding of SA to the surface of the nanocable (right). (b) Conductance-time measurement for a
biotin-modied nanocable, where region 1 corresponds to 3 lL of PB solution, region 2 corresponds to the addition of 3 lL of PB
solution with 1 10
2
g/L SA, and region 3 corresponds to the addition of 3 lL of PB solution with 1 10
2
g/L BSA. Reproduced
from Ref. [182]. Copyright 2007 American Chemical Society.
268 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
efciency of the applied technique, FRET. From these data, the application of CdSe/ZnS QDs in sensing
glucose could replace the uorescent dyes (ruthenium complex, etc.) which are normally used in the
conventional methods. By controlling the PL quenching behaviors, Uematsu and co-workers devel-
oped a highly photoluminescent ZnSAgInS
2
solid solution (ZAIS) NCs for sensing glucose [452]. They
fabricated a glucose-sensing system by combining the redox-dependent quenching of MPA-ZAIS with
a redox enzyme, as illustrated in Fig. 76a.
A pyrroloquinoline quinone glucose dehydrogenase (PQQ-GDH) was employed as the enzyme and
MP
+
was used as the oxidizing agent. When glucose oxidation accompanied by reduction of MP
+
takes
place, it is anticipated that quenched PL emission would be recovered. This was experimentally con-
rmed, as shown in Fig. 76b, where the time course of PL intensity at 655 nm is shown. Addition of
MP
+
(5 lM) to the MPA-ZAIS aqueous solution containing PQQ-GDH (5 unit/mL) and PBS (20 mM)
greatly quenched the PL from ZAIS NCs, and the subsequent addition of an excess amount of glucose
solution (20 lM) made the solution emit once again. Fig. 76c demonstrates PL recovery behavior when
glucose of different concentrations, ranging from 5 to 20 lM, was added to the solution. The emission
intensity increases with glucose concentration, indicating that this system enables quantitative detec-
tion of glucose [452].
Recently, ZnS:Mn nanoparticles were synthesized from quaternary W/O micro-emulsion system
with different Mn% for biosensors [453]. The addition of biotin and the subsequent specic binding
events alter the dielectric environment of the nanoparticle, resulted in a spectral shift of the particle
plasmon resonance. Cyclohexane was used as oil, Triton X-100 as a surfactant, n-hexanol as a co-sur-
factant and mercaptoethanol and a thioglycolic acid for the best linking of the biological part to the
nanoparticle (as linking agents). Surfactant and co-surfactant produced a stable emulsion with
Fig. 76. (a) A schematic illustration of a uorescent glucose sensor. (b and c) Time-course PL spectra (E
x
= 460 nm, E
m
= 655 nm)
of a glucose-sensing solution containing MPA-ZAIS, MP
+
(5 mM), PQQ-GDH (5 unit/mL) and PBS (20 mM, pH = 7.0). Reproduced
from Ref. [452]. Copyright 2009, Royal Society of Chemistry.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 269
connection to the surface of nanoparticles and prevention from contacting together. The results
showed that with reducing particle size, the emission shifted to the lower wavelengths. In addition,
with conjugation between avidin and biotin by mercaptoethanol in biologic media, the spectral emis-
sion decreased. The uorescence spectra for different Mn concentrations showed that during changing
of the concentration of Mn
2+
ions, there was a maximum emission in optimum doping (- 5.5% of Mn).
The uorescence spectra of the doped crystals consist of orange-red emissions. It is believed that spa-
tial interference of the biological part on the nanocrystal caused the decrease emission intensity.
It is known that urea in blood or in urine is an important substance in the diagnosis of renal and
liver diseases. The detection of urea is performed frequently in the medical care. The normal level
of urea in serum is in the range of 2.56.7 mM. In patients suffering from the renal insufciency,
the urea concentration in serum may be as high as 3080 mM. With such an urea level, the hemod-
ialysis is required. For urea detection, biosensors based on the potentiometry and amperometry were
previously investigated. However, the complexity of sensor fabrication is a common drawback often
mentioned in the literature. Huang and co-workers described a simple scheme of preparation of a
urea-sensing system, which was composed of water-soluble mercaptosuccinic acid (MSA)-QDs and
urease. It allows the effective and quantitative detection of urea [454]. By mixing urease and QDs,
the determination of urea can be performed in a quantitative manner. The detection is based on the
enhancement of QD PL intensity, which is correlated to the enzymatic degradation of urea. Fig. 77
shows the variation of PL intensity of MSA-QDs as a function of urea concentrations (0120 mM) un-
der urease catalysis in a 20 mM phosphate buffer with pH 8.0. The emission spectra were recorded
using the excitation wavelength (k
ex
) of 365 nm. The results indicated that the PL intensity of MSA-
QDs was enhanced with increasing urea concentration.
By controlling the buffer concentration and pH, PL enhancement due to the degradation of urea was
linear in the urea concentration ranging from 0.01 to 100 mM. Fig. 78 plots a ratio of PL enhancement
(I I
0
)/I
0
, versus urea concentration. An approximately linear relationship can be observed with the
urea concentration lower than 100 mM [454].
This property makes the urease/QDs promising urea-biosensing system. The newly developed sys-
tem has a superior design and possesses many advantages, including its simple preparation, low cost,
no enzyme immobilization requirement, high exibility, and good sensitivity.
5.9. Nanogenerators
The concept of the nanogenerators was rst introduced by Wang and co-workers through examin-
ing the piezoelectric properties of ZnO nanowires with an atomic force microscope (AFM) [455]. They
Fig. 77. The variation of PL spectra of urease/QDs with urea concentration. Reactions took place in 20 mM pH 8.0 buffers. Note
that, the spectra obtained from the reactions with 100 and 120 mM of urea are nearly superimposed. Reproduced from Ref.
[454]. Copyright 2007, Elsevier.
270 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
converted nanoscale mechanical energy into electrical energy by means of piezoelectric ZnO nanowire
arrays. The presence of polar surfaces and the lack of center symmetry (which results in a piezoelectric
effect) make ZnO one of the hottest candidates for nanogenerators. The coupling of piezoelectric and
Fig. 79. (a) ZnOZnS nanowire arrays grown vertically on a buffer layer, and (b) its three dimensional voltage output image
received after an AFM scan across a heterostructure ZnOZnS nanowire arrays with an area of 30 30 lm
2
. Reproduced from
Ref. [172]. Copyright 2009 American Chemical Society.
Fig. 78. The variation of pH value and PL intensity with urea concentration. Urea was analyzed by the urease/QDs sensing
system in 20 mM pH 8.0 buffers, in which the pH value (h) and the PL perturbation (j) were determined. I
0
and I represent the
PL intensity of MSA-CdSe/ZnS observed at 580 nm with the reaction at zero time and at the end-point, respectively. Reproduced
from Ref. [454]. Copyright 2007, Elsevier.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 271
semiconducting properties in ZnO creates a strain eld and charge separation across the nanowire as a
result of its bending. The rectifying characteristic of the Schottky barrier formed between the metal tip
and the nanowire leads to electrical current generation. The efciency of the nanowire-based piezo-
electric power generator was estimated to be 1730% by Wang and co-workers [455457].
As discussed in the beginning of this review, ZnS has similar fundamental physical properties with
ZnO, including crystal structures, the presence of polar surfaces, lattice constants, and so on. Thus, the
interesting nanoscale morphologies and novel properties of nanoscale ZnS could rival those of ZnO.
Very recently, Wang and co-workers investigated ZnOZnS heterojunctions and ZnS nanowire arrays
for converting mechanical energy into electricity [172]. ZnOZnS heterojunction nanowire arrays were
synthesized by a thermal evaporation process with the presence of residual oxygen. ZnS nanowire ar-
rays were obtained by etching ZnOZnS nanowire arrays with 1 M KOH solution for 120 min. Fig. 79
shows the ZnOZnS nanowire arrays and a three dimensional voltage output image after an AFM scan
across a heterostructure ZnOZnS nanowire arrays with an area of 30 30 lm
2
. Almost all the hetero-
structure nanowire arrays are uniform and grow vertically on the substrate. Piezoelectric responses of
the heterostructure ZnOZnS nanowire arrays were examined using AFM in contact mode with a con-
ducting Pt-coated Si tip. As shown in Fig. 79b, Wang and co-workers explain that each peak represents
an electric voltage/current that is generated by deecting the corresponding nanowire, and output
voltages are around 6 mV. Moreover, the voltage outputs of heterostructure ZnOZnS nanowire arrays
mainly result from the interaction between W-ZnO and the Pt tip as ZnO is at the top part of the het-
erostructure ZnOZnS nanowire arrays [172].
Fig. 80a displays a side-view SEM image of pure ZnS nanowire arrays obtained by removing ZnO
parts of ZnO/ZnS nanowire arrays with a dilute 1 M KOH solution for 120 min [172]. Similar morphol-
ogy and distribution were kept as for the original heterostructure ZnOZnS nanowire arrays. Wang
and co-workers measured the piezoelectric properties of pure ZnS nanowire arrays using AFM under
the same conditions as for heterostructure ZnOZnS nanowire arrays. The results are shown in
Fig. 80. (a) A side-view SEM image of pure ZnS nanowire arrays obtained by removing ZnO parts of ZnO/ZnS nanowire arrays
with a dilute 1 M KOH solution for 120 min and (b) its corresponding three dimensional voltage output image recorded during
an AFM scan across pure ZnS nanowire arrays with an area of 20 20 lm
2
under the same conditions as for the heterostructure
ZnOZnS nanowire arrays. Reproduced from Ref. [172]. Copyright 2009 American Chemical Society.
272 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
Fig. 80b, A three dimensional voltage output image received during an AFM scan across pure ZnS
nanowire arrays with an area of 20 20 lm
2
suggested a generated voltage pulse of 2 mV. Compared
with ZnO nanowire arrays the output signal is signicantly lower because ZnS has smaller piezoelec-
tric constants and the nanowire arrays are shorter [172,458]. Even compared with the heterostructure
ZnOZnS nanowire arrays with a similar height, morphology and structure, a three times decrease was
found.
6. Conclusions and outlooks
In this paper, we provide a comprehensive review on the synthesis of ZnS nanostructures, starting
from 0D nanostructures (0D nanocrystals, 0D core/shell nanocrystals and 0D hollow nanocrystals),
continuing with 1D nanostructures (nanowires, nanorods, nanotubes, nanobelts, nanoribbons, aligned
nanowires and nanobelts, complicated nanostructures, longitudinal heterostructured nanostructures,
coaxial (core/shell) heterostructured nanostructures, side-by-side heterostructures, doped 1D nano-
structures, alloyed 1D nanostructures, tetrapodal nanostructures, bicrystalline nanostructures, hierar-
chical nanostructures, and ZnS/organic 1D nanostructures), and nishing with 2D nanostructures such
as nanosheets. In addition, we also present the novel properties and potential applications of these
materials. Especially the luminescence properties (photoluminescence, cathodoluminescence, electro-
luminescence, electrochemi-luminescence, thermoluminescence, and luminescence mechanism),
chemical and physical properties of the nanostructures produced under different synthetic conditions
and with diverse nanoscale morphologies were introduced by paying special attention to the most re-
cent works as the examples. The tuning on optical properties is highlighted by the introduction of Mn-
doped ZnS nanobelts, Cu-doped ZnS nanorods, and Mn/Cd-co-doped-ZnS nanostructures. The remark-
able level of a synthetic control for ZnS nanostructures and their rich morphologies at the nanoscale
has paved the way to the unique applications in the elds of electronics, optoelectronics, sensors, life
sciences, defense, energy, environmental science, and engineering. The recent achievements in ZnS
nanostructures-based eld emitters, eld effect transistors and analysis of their carrier characteristics,
p-type conductivity, catalytic activities, UV-light sensors, chemical sensors (including gas sensors),
biosensors, and nanogenerators are also highlighted.
The above discussions show that ZnS nanostructures have become one of the popular research pur-
suits owing to the richness of their physical and chemical properties and wide range of potential appli-
cations. The morphology of ZnS at the nanoscale has been proven to be one of the richest among all
inorganic semiconductors. Although there have been many achievements in this eld, several key
challenges has also emerged that need to be overcome for the purpose of practical applications. For
instance, the nanostructures with a clear surface reported in the literatures are usually obtained by
high temperature CVD growth technique and their sizes measure several tens to hundreds of nanome-
ter in width and thickness. There are only few high temperature procedures for the synthesis of ZnS
nanostructures that are small enough to exhibit quantum connement related effects. On the other
hand, low-temperature solution synthesis has remarkable success in obtaining quantum sized ZnS
nanostructures. However in these processes, additional chemical agents are usually employed to re-
strict the size of the nanostructures. Therefore the obtained ZnS nanostructures are covered by a layer
of these constraining agents. This is not desirable, since the surface properties of a nanostructure are
crucial for many applications such as catalysis. This can also inuence optoelectronic applications, by
creating additional surface trap states. Therefore, one of the main challenges is to obtain clean ZnS
nanostructures with much smaller size dimensions, preferably by using high temperature CVD tech-
niques. Apart from the pure nanostructures, reports on the ZnS based heterostructures are rather lim-
ited and should also be pursued earnestly. The coreshell type heterostructures can improve
optoelectronic properties remarkably. The heterostructures can also be used as a part of multi-com-
ponent devices, since assembly of device components is challenging at the nanoscale. Finally, another
important challenge is to stabilize nanocrystals in solution without the use of capping agents. The
quantum dot nanocrystals used in applications such as EL devices are poor in conduction properties,
due to the presence of an organic capping layer that is necessary for quality size control and size
distribution.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 273
As described in the section of synthesis of ZnS nanostructures, as can be seen, a minor difference on
the synthesis condition, the morphologies of as-synthesized ZnS nanostructures will be different. The
growth kinetics and thermodynamics involved in the synthesis of inorganic semiconductor nanostruc-
tures are extremely complex, and presume different mechanisms under different growth conditions. A
deep understanding of the determining factors in the growth of nanostructures may settle down the
uncertainty, and further guild the building of roadmap to the growth of ZnS nanostructures. Although
some researchers have realized the importance of growth mechanism, and tried to explain the mor-
phology derivation and evolution with some experimental parameters, the explanations are rather
supercial and there have been many controversies, such as the inuence of the experimental param-
eters on the morphology variation [459]. Computer simulation methods have proved particularly
effective in modelling and predicting possible structures, morphologies, novel properties and potential
applications [460465], however the big gap between the experimental works and computational
studies on ZnS nanostructures are remained. There should be clearly some new effects to be eluci-
dated, different from traditional theory and innite systems, especially with respect to quantum
behaviors. Therefore, a close collaboration between computational and experimental work is needed.
We have shown in this review that ZnS nanostructures have become one of the popular research
pursuits owing to the richness of their physical and chemical properties and wide range of potential
applications. For technological applications, nanometer precision synthesis, positioning and intercon-
nection of these building blocks for the direct integration, or a reliable post-assembly process of nano-
structures, to implement them in well established manufacturing processes have to be achieved in
order to guarantee a knowledge transfer to an industrial level [466,467]. For example, a great need
is to develop device arrays utilizing the nanowire arrays grown either vertically or laterally on re-
quested (e.g. hard/exiable) substrates with an ideal or compatible contact.
Amongst various properties, the investigation of optical properties of ZnS nanostructures is very
important as it can absorb the fraction of the solar radiation which is carcinogenic in nature. This is
being pursued by many research groups so far and the understanding of the optoelectronic transitions
inside the ZnS nanostructures has improved. However, there are still many unresolved issues, such as
the origin of visible emission, inuence of aerial oxidation and the characterization of surface trap
states. which need to be investigated both experimentally and theoretically. ZnS nanostructures are
considered as one of the most suitable candidate for EL devices. Despite, the development of multi-
color displays with RGB emission has remained difcult as the materials for red, green, and blue phos-
phors require different processing conditions, and therefore their luminous efciencies vary by up to
an order of magnitude. On the other hand, the ZnS based UV sensor have shown remarkable promise.
Recently, simple and inexpensive strategies have been developed for large scale production of these
sensors. The adsorption of gases on the nanostructure surface inuence the current generation in pho-
toinduced processes by interacting with the nascent holes and electrons and this has been explained
by proposing different models. However, more rigorous experimental and theoretical work is also re-
quired to validate these models.
Optimization of nanodevice performances is the key and ultimate goal under their practical appli-
cations. For example, functionalizing the surface of wide bandgap semiconductor with oxides, poly-
mers, nitrides and a catalyst metal is proved to be very useful in enhancing the detection
sensitivity for gases and ionic solutions [468]. The photon detection sensitivity has been enhanced
-58 times with a faster rest time when the contact for CdS nanowire nanosensors was changed from
Ohmic to Schottky operated at the reverse bias mode of 8 V [469]. Although several facile and effec-
tive routes to optimize the device performances based on ZnS nanostructures have been exploited,
more and deep work on tuning the ZnS conductivity, bandgap, surface and optical properties, and
so forth through a control in vacancies formation will be interesting and important.
With a direct wide-band-gap semiconductor, excellent transport properties (reduction of the car-
riers scattering and recombination), an intrinsically n-type semiconductor, good thermal stability and
high electronic mobility, ZnS nanostructures are expected to play a key role on developing novel pho-
tovoltaic solar cells and creating Green renewable energy. The ZnS nanostructures are also ideal ob-
jects for fabricating high-performance nanosensors for biomedical applications, such as force sensors,
blood ow sensors and cancer detection sensors due to their excellent uorescence properties and
nontoxicity.
274 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
The last decade has seen remarkable progress in research activities leading to enormous knowledge
pool of ZnS based nanostructures and guring out of the key challenges. We hope that the systematic
investigation of a synthesis-property-application triangle for the ZnS nanostructures will inspire
more research efforts to address the current challenges and promote an intense interest to the general
study on inorganic semiconducting nanostructures.
Acknowledgements
This work was supported by National Science Foundation of China (Grant Nos. 51002032,
21001028 and 21074023), the innovative team of Ministry of Education of China (IRT0911), Grants-
in-Aid for Scientic Research (B), Japan Society for the Promotion of Science (JSPS) (No. 22760517),
and the World Premier International Research Center (WPI) Initiative on Materials Nanoarchitectonics
(MANA), MEXT, Japan. The authors are indebted to corresponding publishers/authors for the kind per-
missions to reproduce their materials, especially gures, used in this review.
References
[1] Fang XS, Bando Y, Gautam UK, Ye CH, Golberg D. Inorganic semiconductor nanostructures and their eld-emission
applications. J Mater Chem 2008;18:50922.
[2] Zhai TY, Fang XS, Li L, Bando Y, Golberg D. One-dimensional CdS nanostructures: synthesis, properties, and applications.
Nanoscale 2010;2:16887.
[3] Fang XS, Bando Y, Gautam UK, Zhai TY, Zeng HB, Xu XJ, et al. ZnO and ZnS nanostructures: ultraviolet-light emitters, lasers
and sensors. Crit Rev Solid State Mat Sci 2009;34:190223.
[4] Kuchibhatla SVNT, Karakoti AS, Bera D, Seal S. One dimensional nanostructured materials. Prog Mater Sci
2007;52:699913.
[5] Xia YN, Yang PD, Sun YG, Wu YY, Mayers B, Gates B, et al. One-dimensional nanostructures: synthesis, characterization,
and applications. Adv Mater 2003;15:35389.
[6] Davidson WL. X-ray diffraction evidence for ZnS formation in zinc activated rubber vulcanizates. Phys Rev
1948;74:1167.
[7] Yeh CY, Lu ZW, Froyen S, Zunger A. Zinc blende wurtzite polytypism in semiconductors. Phys Rev B 1992;46:1008697.
[8] Chen H, Shi D, Qi J, Jia J, Wang B. The stability and electronic properties of wurtzite and zinc-blende ZnS nanowires. Phys
Lett A 2009;373:3715.
[9] Tran TK, Park W, Tong W, Kyi MM, Wagner BK, Summers CJ. Photoluminescence properties of ZnS epilayers. J Appl Phys
1997;81:28039.
[10] Karazhanov SZ, Ravindran P, Kjekshus A, Fjellvag H, Svensson BG. Electronic structure and optical properties of ZnX (X = O,
S, Se, Te): a density functional study. Phys Rev B 2007;75:155104.
[11] Yeh CY, Wei SH, Zunger A. Relationship between the band-gaps of the zinc blende and the wurtzite modications of
semiconductors. Phys Rev B 1994;50:27158.
[12] Adachi S, editor. Optical constants of crystalline and amorphous semiconductors: numerical data and graphical
information. Boston: Kluwer Academic; 1999.
[13] Karazhanov SZ, Ravindran P, Grossner U, Kjekshus A, Fjellvag H, Svensson BG, et al. J Appl Phys 2006;100:043709.
[14] Karazhanov SZ, Ravindran P, Grossner Kjekshus A, Fjellvag H, Svensson BG. Electronic structure and band parameters for
ZnX (X = O, S, Se, Te). J Cryst Growth 2006;287:1628.
[15] Qadri SB, Skelton EF, Hsu D, Dinsmore AD, Yang J, Gray HF, et al. Size-induced transition-temperature reduction in
nanoparticles of ZnS. Phys Rev B 1999;60:91913.
[16] Akiyama T, Sano K, Nakamura K, Ito T. An empirical interatomic potential approach to structural stability of ZnS and ZnSe
nanowires. Jpn J Appl Phys 2007;46:17837.
[17] Li JB, Wang LW. Band-structure-corrected local density approximation study of semiconductor quantum dots and wires.
Phys Rev B 2005;72:125325.
[18] Li LJ, Zhao MW, Zhang XJ, Zhu HZ, Li F, Li JL, et al. Theoretical insight into faceted ZnS nanowires and nanotubes from
interatomic potential and rst-principles calculations. J Phys Chem C 2008;112:350914.
[19] Zhang XJ, Zhao MW, Yan SS, He T, Li WF, Lin XH, et al. First-principles study of ZnS nanostructures: nanotubes, nanowires
and nanosheets. Nanotechnology 2008;19:305708.
[20] Yang PD, Wu YY, Fan R. Inorganic semiconductor nanowires. Int J Nanosci 2002;1:139.
[21] Alivisatos AP. Semiconductor clusters, nanocrystals, and quantum dots. Science 1996;271:9337.
[22] Jr MB, Moronne M, Gin P, Weiss S, Alivisatos AP. Semiconductor nanocrystals as uorescent biological labels. Science
1998;281:20136.
[23] Peng XG, Manna L, Yang WD, Wickham J, Scher E, Kadavanich A, et al. Shape control of CdSe nanocrystals. Nature
2000;404:5961.
[24] Sun YG, Xia YN. Shape-controlled synthesis of gold and silver nanoparticles. Science 2002;298:21769.
[25] Peng XG. An essay on synthetic chemistry of colloidal nanocrystals. Nano Res 2009;2:42547.
[26] Rossetti R, Hull R, Gibson JM, Brus LE. Excited electronic states and optical spectra of ZnS and CdS crystallites in the ~15 to
50 size range: Evolution from molecular to bulk semiconducting properties. J Chem Phys 1985;82:5529.
[27] Murray CB, Norris DJ, Bawendi MG. Synthesis and characterization of nearly monodisperse CdE (E = S, Se, Te)
semiconductor nanocrystallites. J Am Chem Soc 1993;115:870615.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 275
[28] Jun YW, Choi JS, Cheon J. Shape control of semiconductor and metal oxide nanocrystals through nonhydrolytic colloidal
routes. Angew Chem Int Ed 2006;45:341439.
[29] Joo J, Na HB, Yu T, Yu JH, Kim YW, Wu FX, et al. Generalized and facile synthesis of semiconducting metal sulde
nanocrystals. J Am Chem Soc 2003;125:111005.
[30] Zhao YW, Zhang Y, Zhu H, Hadjipanayis GC, Xiao JQ. Low-temperature synthesis of hexagonal (wurtzite) ZnS nanocrystals.
J Am Chem Soc 2004;126:68745.
[31] Li LS, Pradhan N, Wang YJ, Peng XG. High quality ZnSe and ZnS nanocrystals formed by activating zinc carboxylate
precursors. Nano Lett 2004;4:22614.
[32] Li YC, Li XH, Ye CH, Li YF. Ligand-controlling synthesis and ordered assembly of ZnS nanorods and nanodots. J Phys Chem
B 2004;108:1600211.
[33] Li YC, Ye MF, Yang CH, Li XH, Li YF. Composition- and shape-controlled synthesis and optical properties of Zn
x
Cd
1-x
S
alloyed nanocrystal. Adv Funct Mater 2005;15:43341.
[34] Talapin DV, Rogach AL, Kornowski A, Haase M, Weller H. Highly luminescent monodisperse CdSe and CdSe/ZnS
nanocrystals synthesized in a hexadecylaminetrioctylphosphine oxidetrioctylphospine mixture. Nano Lett
2001;1:20711.
[35] Xie RG, Kolb U, Li JX, Basch T, Mews A. Synthesis and characterization of highly luminescent CdSe-core CdS/Zn0.5Cd0.5S/
ZnS multishell nanocrystals. J Am Chem Soc 2005;127:74808.
[36] Pal S, Goswami B, Sarkar P. Size-dependent properties of hollow ZnS nanoclusters. J Phys Chem C 2008;112:630712.
[37] Zhou H, Fan TX, Zhang D, Guo QX, Ogawa H. Novel bacteria-templated sonochemical route for the in situ one-step
synthesis of ZnS hollow nanostructures. Chem Mater 2007;19:21446.
[38] Shao HF, Qian XF, Zhu ZK. The synthesis of ZnS hollow nanospheres with nanoporous shell. J Solid State Chem
2005;178:35228.
[39] Zhang H, Zhang SY, Pan S, Li GP, Hou JG. A simple solution route to ZnS nanotubes and hollow nanospheres and their
optical properties. Nanotechnology 2004;15:9458.
[40] Wolosiuk A, Armagan O, Braun PV. Double direct tempalting of periodically nanostructured ZnS hollow microspheres. J
Am Chem Soc 2005;127:163567.
[41] Yin JL, Qian XF, Yin J, Shi MW, Zhou GT. Preparation of ZnS/PS micropheres and ZnS hollow shells. Mater Lett
2003;57:385963.
[42] Breen ML, Dinsmore AD, Pink RH, Qadri SB, Ratna BR. Sonochemically produced ZnS-coated polystyrene coreshell
particles for use in photonic crystals. Langmuir 2001;17:9037.
[43] Arul Dhas N, Zaban A, Gedanken A. Surface synthesis of zinc sulde nanoparticles on silica microspheres: sonochemical
preparation, characterization, and optical properties. Chem Mater 1999;11:80613.
[44] Liu HJ, Ni YH, Han M, Liu Q, Xu Z, Hong JM, et al. A facile template-free route for the synthesis of hollow hexagonal ZnS
nano-and submicro-spheres. Nanotechnology 2005;16:290812.
[45] Ma YR, Qi LM, Ma JM, Cheng HM. Facile synthesis of hollow ZnS nanospheres in block copolymer solutions. Langmuir
2003;19:40402.
[46] Wang ZX, Wu LM, Chen M, Zhou SX. Facile synthesis of superparamagnetic uorescent Fe
3
O
4
/ZnS hollow nanospheres. J
Am Chem Soc 2009;131:112767.
[47] Wang X, Zhuang J, Peng Q, Li YD. A general strategy for nanocrystal synthesis. Nature 2005;437:1214.
[48] Wang WZ, Germanenko I, EI-Shall MS. Room-temperature synthesis and characterization of nanocrystalline CdS, ZnS and
Cd
x
Zn
1-x
S. Chem Mater 2002;14:302833.
[49] Yu JH, Joo J, Park HM, Baik S, Kim YW, Kim SC, et al. Synthesis of quantum-sized cubic ZnS nanorods by the oriented
attachment mechanism. J Am Chem Soc 2005;127:566270.
[50] Li L, Daou TJ, Texier I, Chi TTK, Liem NQ, Reiss P. Highly luminescent CuInS
2
/ZnS core/shell nanocrystals: cadmium-free
quantum dots for in vivo imaging. Chem Mater 2009;21:24229.
[51] Yang YA, Chen O, Angerhofer A, Cao YC. Radial-position-controlled doping of CdS/ZnS core/shell nanocrystals: surface
effects and position-dependent properties. Chem Eur J 2009;15:318697.
[52] Steckel JS, Zimmer JP, Coe-Sullivan S, Stott NE, Bulovic V, Bawendi M. Blue luminescence from (CdS)ZnS coreshell
nanocrystals. Angew Chem Int Ed 2004;43:21548.
[53] Cumberland SL, Hanif KM, Javier A, Khirov GA, Strouse GF, Woessner SM, et al. Inorganic clusters as single-source
precursors for preparation of CdSe, ZnSe and CdSe/ZnS nanomaterials. Chem Mater 2002;14:157684.
[54] Dabbousi BO, Rodriguez-Viejo J, Mikulec FV, Heine JR, Mattoussi H, Ober R, et al. (CdSe)ZnS coreshell quantum dots:
synthesis and characterization of a size series of highly luminescent nanocrystallites. J Phys Chem B 1997;101:946375.
[55] Bol AA, Meijerink A. Long-lived Mn
2+
emission in nanocrystalline ZnS:Mn
2+
. Phys Rev B 1998;58:159976000.
[56] Suyver JF, Wuister SF, Kelly JJ, Meijerink A. Synthesis and photoluminescence of nanocrystalline ZnS:Mn
2+
. Nano Lett
2001;1:42933.
[57] Bi C, Pan LQ, Xu M, Yin JH, Qin LQ, Liu JM, et al. Synthesis and characterization of Co-doped wurtzite ZnS nanocrystals.
Mater Chem Phys 2009;116:3637.
[58] Ehlert O, Osvet A, Batentschuk M, Winnacker A, Nann T. Synthesis and spectroscopic investigation of Cu- and Pb-doped
colloidal ZnS nanocrystals. J Phys Chem B 2006;110:231758.
[59] Zhong XH, Liu SH, Zhang ZH, Li L, Wei Z, Knoll W. Synthesis of high-quality CdS, ZnS and Zn
x
Cd
1x
S nanocrystals using
metal salts and elemental sulfur. J Mater Chem 2004;14:27904.
[60] Zhong XH, Feng YY, Knoll W, Han MY. Alloyed Zn
x
Cd
1x
S nanocrystals with highly narrow luminescence spectral width. J
Am Chem Soc 2003;125:1355963.
[61] Lieber CM, Wang ZL. Functional nanowires. MRS Bull 2007;32:99108.
[62] Comini E, Baratto C, Faglia G, Ferroni M, Vomiero A, Sberglieri G. Quasi-one dimensional metal oxide semiconductors:
preparation, characterization and application as chemical sensors. Prog Mater Sci 2009;54:167.
[63] Zhai TY, Fang XS, Liao MY, Xu XJ, Zeng HB, Bando Y, et al. A comprehensive review of one-dimensional metaloxide
nanostructure photodetectors. Sensors 2009;9:650429.
[64] Moore D, Wang ZL. Growth of anisotropic one-dimensional ZnS nanostructures. J Mater Chem 2006;16:3898905.
276 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
[65] Wang YW, Zhang LD, Liang CH, Wang GZ, Peng XS. Catalytic growth and photoluminescence properties of semiconductor
single-crystal ZnS nanowires. Chem Phys Lett 2002;357:3148.
[66] Barrelet CJ, Wu Y, Bell DC, Lieber CM. Synthesis of CdS and ZnS nanowires using single-source molecular precursors. J Am
Chem Soc 2003;125:114989.
[67] Law M, Goldberger J, Yang PD. Semiconductor nanowires and nanotubes. Annu Rev Mater Res 2004;34:83122.
[68] Jiang XC, Xie Y, Lu J, Zhu LY, He W, Qian YT. Simultaneous in situ formation of ZnS nanowires in a liquid crystal template
by c-irradiation. Chem Mater 2001;13:12138.
[69] Chai LL, Du J, Xiong SL, Li HB, Zhu YC, Qian YT. Synthesis of wurtzite ZnS nanowire bundles using a solvothermal
technique. J Phys Chem C 2007;111:1265862.
[70] Shi L, Xu YM, Li Q. Shape-selective synthesis and optical properties of highly ordered one-dimensional ZnS
nanostructures. Cryst Growth Des 2009;9:22149.
[71] Wang M, Fei GT, Zhu XG, Wu B, Kong MG, Zhang LD. Density-controlled homoepitaxial growth of ZnS nanowire arrays. J
Phys Chem C 2009;113:43359.
[72] Biswas S, Ghoshal T, Kar S, Chakrabarti S, Chaudhuri S. ZnS nanowire arrays: synthesis, optical and eld emission
properties. Cryst Growth Des 2008;8:21716.
[73] Chan SK, Lok SK, Wang G, Cai Y, Wang N, Wong KS, et al. MBE-growth cubic ZnS nanowires. J Electron Mater
2008;37:14337.
[74] Liu JY, Guo Z, Jia Y, Meng FL, Luo T, Liu JH. Triethylenetetramine (TETA)-assisted synthesis, dynamic growth mechanism,
and photoluminescence properties of radial single-crystalline ZnS nanowire bundles. J Cryst Growth 2009;311:14239.
[75] Yao WT, Yu SH, Wu QS. From mesostructured wurtzite ZnS-nanowire/amine nanocomposities to ZnS nanowires
exhibiting quantum size effects: a mild-solution chemistry approach. Adv Funct Mater 2007;17:62331.
[76] Shen GZ, Bando Y, Golberg D, Zhou CW. Heteroepitaxial growth of orientation-ordered ZnS nanowire arrays. J Phys Chem
C 2008;112:12299303.
[77] Chang M, Cao XL, Xu XJ, Zhang LD. Fabrication and photoluminescence properties of highly ordered ZnS nanowire arrays
embedded in anodic alumina membrane. Phys Lett A 2008;372:2736.
[78] Liu BD, Bando Y, Wang MS, Zhi CY, Fang XS, Tang CC, et al. Electron-beam irradiation induced conductivity in ZnS
nanowires as revealed by in situ transmission electron microscope. J Appl Phys 2009;106:034302.
[79] Chang YQ, Wang MW, Chen XH, Ni SL, Qiang WJ. Field emission and photoluminescence characteristics of ZnS nanowires
via vapor phase growth. Solid state Commun 2007;142:2958.
[80] Zhang ZX, Yuan HJ, Liu DF, Liu LF, Shen J, Xiang YJ, et al. Nanotechnology 2007;18:145607.
[81] Sun HY, Li XH, Li W, Li F, Liu BT, Zhang XY. Low-temperature synthesis of wurtzite ZnS single-crystal nanowire arrays.
Nanotechnology 2007;18:115604.
[82] Chen ZG, Zou J, Lu GQ, Liu G, Li F, Cheng HM. ZnS nanowires and their coaxial lateral nanowire heterostructures with BN.
Appl Phys Lett 2007;90:103117.
[83] Yue GH, Yan PX, Yan D, Fan XY, Wang MX, Qu DM, et al. Hydrothermal synthesis of single-crystal ZnS nanowires. Appl
Phys A 2006;84:40912.
[84] Xu XJ, Fei GT, Yu WH, Wang XW, Chen L, Zhang LD. Preparation and formation mechanism of ZnS semiconductor
nanowires made by the electrochemical deposition method. Nanotechnology 2006;17:4269.
[85] Ma DDD, Lee ST, Mueller P, Alvarado SF. Scanning tunneling microscope excited cathodoluminescence from ZnS
nanowires. Nano Lett 2006;6:9269.
[86] Zhang XH, Zhang Y, Song YP, Wang Z, Yu DP. Optical properties of ZnS nanowires synthesized via simple physical
evaporation. Phys E 2005;28:16.
[87] Zhang H, Zhang SY, Zuo M, Li GP, Hou JG. Synthesis of ZnS nanowires and assemblies by carbothermal chemical vapor
deposition and their photoluminescence. Eur J Inorg Chem 2005:4750.
[88] Lu HY, Chu SY, Chang CC. Synthesis and optical properties of well-aligned ZnS nanowires on Si substrate. J Cryst Growth
2005;280:1738.
[89] Ding JX, Zapien JA, Chen WW, Lifshitz Y, Lee ST, Meng XM. Lasing in ZnS nanowires grown on anodic aluminum oxide
templates. Appl Phys Lett 2004;85:23613.
[90] Lin M, Sudhiranjan T, Boothroyd C, Loh KP. Inuence of Au catalyst on the growth of ZnS nanowires. Chem Phys Lett
2004;400:1758.
[91] Moore DF, Ding Y, Wang ZL. Crystal orientation-ordered ZnS nanowire bundles. J Am Chem Soc 2004;126:143723.
[92] Xiong QH, Chen G, Acord JD, Liu X, Zengel JJ, Gutierrez HR, et al. Optical properties of rectangular cross-sectional ZnS
nanowires. Nano Lett 2004;4:16638.
[93] Xiong QH, Wang JG, Reese O, Voon LCLY, Eklund PC. Raman scattering from surface phonons in rectangular cross-sectional
w-ZnS nanowires. Nano Lett 2004;4:19916.
[94] Meng XM, Liu J, Jiang Y, Chen WW, Lee CS, Bello I, et al. Structure- and size-controlled ultrane ZnS nanowires. Chem Phys
Lett 2003;382:4348.
[95] Jiang Y, Meng XM, Liu J, Hong ZR, Lee CS, Lee ST. ZnS nanowires with wurtzite polytype modulated structure. Adv Mater
2003;15:11958.
[96] Moon H, Nam C, Kim C, Kim B. Synthesis and photoluminescence of zinc sulde nanowires by simple thermal chemical
vapor deposition. Mater Res Bull 2006;41:20137.
[97] Yang Y, Zhang WJ. Preparation and photoluminescence of zinc sulde nanowires. Mater Lett 2004;58:38368.
[98] Zhou TY, Yuan X, Hong JM, Xin XQ. Room-temperature solid-state reaction to nanowires of zinc sulde. Mater Lett
2006;60:16872.
[99] Lin M, Zhang J, Boothroyd C, Foo YL, Yeadon M, Loh KP. Hollowing mechanism of zinc sulde nanowires in vacuum
induced by an atomic oxygen beam. J Phys Chem B 2004;108:96317.
[100] Kar S, Biswas S, Chaudhuri S. Catalytic growth and photoluminescence properties of ZnS nanowires. Nanotechnology
2005;16:73740.
[101] Kar S, Chaudhuri S. Controlled synthesis and photoluminescence properties of ZnS nanowires and nanoribbons. J Phys
Chem B 2005;109:3298302.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 277
[102] Yu W, Fang PF, Wang SJ. ZnS nanorod arrays synthesized by an aqua-solution hydrothermal process upon pulse-plating
Zn nanocrystallines. Appl Surf Sci 2009;255:570913.
[103] Limaye MV, Gokhale S, Acharya SA, Kulkarni SK. Template-free ZnS nanorod synthesis by microwave irradiation.
Nanotechnology 2008;19:415602.
[104] Biswas S, Kar S. Fabrication of ZnS nanoparticles and nanorods with cubic and hexagonal crystal structures: a simple
solvothermal approach. Nanotechnology 2008;19:045710.
[105] Panda SK, Dev A, Chaudhuri S. Fabrication and luminescent properties of c-Axis oriented ZnOZnS coreshell and ZnS
nanorod arrays by suldation of aligned ZnO nanorod arrays. J Phys Chem C 2007;111:503943.
[106] Ghosh PK, Maiti UN, Jana S, Chattopadhyay KK. Field emission from ZnS nanorods synthesized by radio frequency
magnetron sputtering technique. Appl Surf Sci 2006;253:154450.
[107] Zhang YC, Wang GY, Hu XY, Chen WW. Solvothermal synthesis of uniform hexagonal-phase ZnS nanorods using a single-
source molecular precursor. Mater Res Bull 2006;41:181724.
[108] Zhao ZG, Geng FX, Cong HT, Bai JB, Cheng HM. A simple solution route to controlled synthesis of ZnS submicrospheres,
nanosheets and nanorods. Nanotechnology 2006;17:47315.
[109] Feng QJ, Shen DZ, Zhang JY, Liang HW, Zhao DX, Lu YM, et al. Highly aligned ZnS nanorods grown by plasma-assisted
metalorganic chemical vapor deposition. J Cryst Growth 2005;285:5615.
[110] Fang XS, Ye CH, Zhang LD, Wang YH, Wu YC. Temperature-controlled catalytic growth of ZnS nanostructures by the
evaporation of ZnS nanopowders. Adv Funct Mater 2005;15:638.
[111] Liang CH, Shimizu Y, Sasaki T, Umehara H, Koshizaki N. Au-mediated growth of wurtzite ZnS nanobelts, nanosheets, and
nanorods via thermal evaporation. J Phys Chem B 2004;108:972833.
[112] Chen XJ, Xu HF, Xu NS, Zhao FH, Lin WJ, Lin G, et al. Kinetically controlled synthesis of wurtzite ZnS nanorods through
mild thermolysis of a covalent organicinorganic network. Inorg Chem 2003;42:31006.
[113] Wang XD, Gao PX, Li J, Summers CJ, Wang ZL. Rectangular porous ZnOZnS nanocables and ZnS nanotubes. Adv Mater
2002;14:17325.
[114] Yin LW, Bando Y, Zhan JH, Li MS, Golberg D. Self-assembled highly faceted wurtzite-type ZnS single-crystalline nanotubes
with hexagonal cross-sections. Adv Mater 2005;17:19727.
[115] Zhai TY, Gu ZJ, Ma Y, Yang WS, Zhao LY, Yao JN. Synthesis of ordered ZnS nanotubes by MOCVD-template method. Mater
Chem Phys 2006;100:2814.
[116] Shao HF, Qian XF, Huang BC. Fabrication of single-crystal ZnO nanorods and ZnS nanotubes through a simple ultrasonic
chemical solution method. Mater Lett 2007;61:363943.
[117] Shi L, Xu YM, Li Q, Wu ZY, Chen FR, Kai JJ. Single crystalline ZnS nanotubes and their structural degradation under electron
beam irradiation. Appl Phys Lett 2007;90:211910.
[118] Kalyanikutty KP, Nikhila M, Maitra U, Rao CNR. Hydrogel-assisted synthesis of nanotubes and nanorods of CdS, ZnS and
CuS, showing some evidence for oriented attachment. Chem Phys Lett 2006;432:1904.
[119] Yan CL, Xue DF. Conversion of ZnO Nanorod arrays into ZnO/ZnS nanocable and ZnS nanotube arrays via an in situ
chemistry strategy. J Phys Chem B 2006;110:258505.
[120] Farhangfar S, Yang RB, Pelletier M, Nielsch K. Atomic layer deposition of ZnS nanotubes. Nanotechnology
2009;20:325602.
[121] Shen GZ, Bando Y, Golberg D. Self-assembled three-dimensional structures of single-crystalline ZnS submicrotubes
formed by coalescence of ZnS nanowires. Appl Phys Lett 2006;88:123107.
[122] Dloczik L, Engelhardt R, Ernst K, Fiechter S, Sieber I, Knenkamp R. Hexagonal nanotubes of ZnS by chemical conversion of
monocrystalline ZnO columns. Appl Phys Lett 2001;78:36879.
[123] Pan ZW, Dai ZR, Wang ZL. Nanobelts of semiconducting oxides. Science 2001;291:19479.
[124] Jiang Y, Meng XM, Liu J, Xie ZY, Lee CS, Lee ST. Hydrogen-assisted thermal evaporation synthesis of ZnS nanoribbons on a
large scale. Adv Mater 2003;15:3237.
[125] Fang XS, Bando Y, Liao MY, Gautam UK, Zhi CY, Dierre B, et al. Single-crystalline ZnS nanobelts as ultraviolet-light sensors.
Adv Mater 2009;21:20349.
[126] Li JY, Zhang Q, An L, Qin LC, Liu J. Large-scale growth of millimeter-long single-crystalline ZnS nanobelts. J Soild State
Chem 2008;181:311620.
[127] Chen ZG, Zou J, Liu G, Lu HF, Li F, Lu GQ, et al. Silicon-induced oriented ZnS nanobelts for hydrogen sensitivity.
Nanotechnology 2008;19:055710.
[128] Fang XS, Bando Y, Shen GZ, Ye CH, Gautam UK, Costa PMFJ, et al. Ultrane ZnS nanobelts as eld emitters. Adv Mater
2007;19:25936.
[129] Liu YG, Feng P, Xue XY, Shi SL, Fu XQ, Wang C, et al. Room-temperature oxygen sensitivity of ZnS nanobelts. Appl Phys
Lett 2007;90:042109.
[130] Lee JY, Kim DS, Park J. Chemical conversion reaction between CdS nanobelts and ZnS nanobelts by vapor transport. Chem
Mater 2007;19:46639.
[131] Hu JT, Wang GZ, Guo CX, Li DP, Zhang LL, Zhao JJ. Au-catalyst growth and photoluminescence of zinc-blende and wurtzite
ZnS nanobelts via chemical vapor deposition. J Lumin 2007;122123:1725.
[132] Fang XS, Bando Y, Ye CH, Golberg D. Crystal orientation-ordered ZnS nanobelt quasi-arrays and their enhanced eld-
emission. Chem Commun 2007:304850.
[133] Borchers C, Stichtenoth D, Mller S, Schwen D, Ronning C. Catalyst-nanostructure interaction and growth of ZnS
nanobelts. Nanotechnology 2006;17:106771.
[134] Yang FQ, Jiang CB, Du WW, Zhang ZQ, Li SX, Mao SX. Nanomechanical characterization of ZnS nanobelts. Nanotechnology
2005;16:10737.
[135] Wang ZW, Daemen LL, Zhao YS, Zha CS, Downs RT, Wang XD, et al. Morphology-tuned wurtzite-type ZnS nanobelts. Nat
Mater 2005;4:9227.
[136] Gong JF, Yang SG, Duan JH, Zhang R, Du YW. Rapid synthesis and visible photoluminescence of ZnS nanobelts. Chem
Commun 2005:3513.
278 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
[137] Yao WT, Yu SH, Pan L, Li J, Wu QS, Zhang L, et al. Flexible wurtzite-type ZnS nanobelts with quantum-size effects: a
diethylenetriamine-assisted solvothermal approach. Small 2005;1:3205.
[138] Zhu YC, Bando Y, Xue DF, Golberg D. Oriented assemblies of ZnS one-dimensional nanostructures. Adv Mater
2004;16:8314.
[139] Geng BY, Zhang YG, Wang G, Xie T, Meng GW, Zhang LD. Growth of single-crystal ZnS nanobelts through a low-
temperature thermochemistry route and their optical properties. Appl Phys A 2004;79:17613.
[140] Ding Y, Wang XD, Wang ZL. Phase controlled synthesis of ZnS nanobelts: zinc blende vs wurtzite. Chem Phys Lett
2004;398:326.
[141] Li YJ, You LP, Duan R, Shi PB, Qian GG. Oxidation of a ZnS nanobelt into a ZnO nanotwin belt or double single-crystalline
ZnO nanobelts. Solid State Commun 2004;129:2338.
[142] Li YJ, You LP, Duan R, Shi PB, Du HJ, Qiao YP, et al. Straight ZnS nanobelts with wurtzite structure synthesized by the
vapour phase transport process and their crystallization and photoluminescence properties. Nanotechnology
2004;15:5815.
[143] Li Q, Wang CR. Fabrication of wurtzite ZnS nanobelts via simple thermal evaporation. Appl Phys Lett 2003;83:35961.
[144] Hu PA, Liu YQ, Fu L, Cao LC, Zhu DB. Self-assembled growth of ZnS nanobelt networks. J Phys Chem B 2004;108:9368.
[145] Ma C, Moore D, Li J, Wang ZL. Nanobelts, nanocombs, and nanowindmills of wurtzite ZnS. Adv Mater 2003;15:22831.
[146] Zhu YC, Bando Y, Xue DF. Spontaneous growth and luminescence of zinc sulde nanobelts. Appl Phys Lett
2003;82:176971.
[147] Ye CH, Fang XS, Li GH, Zhang LD. Origin of the green photoluminescence from zinc sulde nanobelts. Appl Phys Lett
2004;85:30357.
[148] Lu F, Cai WP, Zhang YG, Li Y, Sun FQ, Heo SH, et al. Appl Phys Lett 2006;89:231928.
[149] Lu F, Cai WP, Zhang YG, Li Y, Sun FQ, Heo SH, et al. Fabrication and eld-emission performance of zinc sulde nanobelt
arrays. J Phys Chem C 2007;111:1338592.
[150] Yuan HJ, Xie SS, Liu DF, Yan XQ, Zhou ZP, Ci LJ, et al. Formation of ZnS nanostructures by a simple way of thermal
evaporation. J Cryst Growth 2003;258:22531.
[151] Kar S, Biswas S, Chaudhuri S. Nanometre to micrometre wide ZnS nanoribbons. Nanotechnology 2005;16:30748.
[152] Zhang ZX, Wang JX, Yuan HJ, Gao Y, Liu DF, Song L, et al. Low-temperature growth and photoluminescence properties of
ZnS nanoribbons. J Phys Chem B 2005;109:183525.
[153] Zhou XT, Kim PSG, Sham TK, Lee ST. Fabrication, morphology, structure, and photoluminescence of ZnS and CdS
nanoribbons. J Appl Phys 2005;98:024312.
[154] Kar S, Chaudhuri S. Synthesis and optical properties of single and bicrystalline ZnS nanoribbons. Chem Phys Lett
2005;414:406.
[155] Hu JQ, Bando Y, Zhan JH, Golberg D. Growth of wurtzite ZnS micrometer-sized diskettes and nanoribbon arrays with
improved luminescence. Adv Funct Mater 2005;15:75762.
[156] Liu XY, Tian BZ, Yu CZ, Tu B, Zhao DY. Microwave-assisted solvothermal synthesis of radial ZnS nanoribbons. Chem Lett
2004;33:5223.
[157] Wang XY, Zhu YC, Fan H, Zhang MF, Xi BJ, Wang HZ, et al. Growth of ZnS microfans and nanosheets: controllable
morphology and phase. J Cryst Growth 2008;310:252531.
[158] Yue GH, Yan PX, Yan D, Liu JZ, Qu DM, Yang Q, et al. Synthesis of two-dimensional micron-sized single-crystalline ZnS thin
nanosheets and their photoluminescence properties. J Cryst Growth 2006;293:42832.
[159] Fang XS, Ye CH, Peng XS, Wang YH, Wu YC, Zhang LD. Large-scale synthesis of ZnS nanosheets by the evaporation of ZnS
nanopowders. J Cryst Growth 2004;263:2638.
[160] Hu JQ, Bando Y, Zhan JH, Li YB, Sekiguchi T. Two-dimensional micrometer-sized single-crystalline ZnO thin nanosheets.
Appl Phys Lett 2003;83:44146.
[161] Yu SH, Yoshimura M. Shape and phase control of ZnS nanocrystals: template fabrication of wurtzite ZnS single-crystal
nanosheets and ZnO ake-like dendrites from a lamellar molecular precursor ZnS (NH
2
CH
2
CH
2
NH
2
)
0.5
. Adv Mater
2002;14:296300.
[162] Yu SH, Yang J, Qian YT, Yoshimura M. Optical properties of ZnS nanosheets, ZnO dendrites, and their lamellar precursor
ZnS (NH
2
CH
2
CH
2
NH
2
)0.5. Chem Phys Lett 2002;361:3626.
[163] Wang ZL. ZnO nanowire and nanobelt platform for nanotechnology. Mater Sci Eng R 2009;64:3371.
[164] Zhai TY, Fang XS, Bando Y, Liao Q, Xu XJ, Zeng HB, et al. Morphology-dependent stimulated emission and eld emission of
ordered CdS nanostructure arrays. ACS Nano 2009;3:94959.
[165] Li L, Koshizaki N, Li GH. Nanotube arrays in porous anodic alumina membranes. J Mater Sci Technol 2008;24:
55062.
[166] Li L, Li GH, Fang XS. Bi-based nanowire and nanojunction arrays: fabrication and physical properties. J Mater Sci Technol
2007;23:16681.
[167] Verheijen MA, Lmmink G, Smet TD. Borgstrm, Bakkers EPAM. Growth kinetics of heterostructured GaPGaAs nanowires.
J Am Chem Soc 2006;128:13539.
[168] Wu Y, Xiang J, Yang C, Lu W, Lieber CM. Single-crystal metallic nanowires and metal/semiconductor nanowire
heterostructures. Nature 2004;430:615.
[169] Hu JQ, Bando Y, Zhan JH, Golberg D. Fabrication of silica-shielded GaZnS metalsemiconductor nanowire
heterojunctions. Adv Mater 2005;17:19649.
[170] Fang XS, Gautam UK, Bando Y, Golberg D. One-dimensional ZnS-based hetero-, core/shell and hierarchical
nanostructures. J Mater Sci Technol 2008;24:5208.
[171] Hu JQ, Bando Y, Golberg D. Novel semiconducting nanowire heterostructures: synthesis, properties and applications. J
Mater Chem 2009;19:33043.
[172] Yu MY, Song JH, Lu MP, Lee CY, Chen LJ, Wang ZL. ZnOZnS heterojunctions and ZnS nanowire arrays for electricity
generation. ACS Nano 2009;3:35762.
[173] Fan LB, Song HW, Zhao HF, Pan GH, Yu HQ, Bai X, et al. Solvothermal synthesis and photoluminescent properties of ZnS/
Cyclohexylamine inorganicorganic hybrid semiconductor nanowires. J Phys Chem B 2006;110:1294853.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 279
[174] Du JM, Fu L, Liu ZM, Han BX, Li ZH, Liu YQ, et al. Facile route to synthesize multiwalled carbon nanotube/zinc sulde
heterostructures: optical and electrical properties. J Phys Chem B 2005;109:127726.
[175] Zhu YC, Bando Y, Yin LW. Design and fabrication of BN-sheathed ZnS nanoarchitectures. Adv Mater 2004;16:3314.
[176] Wang ZL. Functional oxide nanobelts: materials, properties and potential applications in nanosystems and biotechnology.
Annu Rev Phys Chem 2004;55:15996.
[177] Hu JQ, Bando Y, Zhan JH, Golberg D. Fabrication of ZnS/SiC nanocables, SiC-shelled ZnS nanoribbons (and sheets), and SiC
nanotubes (and tubes). Appl Phys Lett 2004;85:29324.
[178] Lin YF, Hsu YJ, Lu SY, Chiang WS. One-step formation of coreshell suldeoxide nanorod arrays from a single precursor.
Nanotechnology 2006;17:477382.
[179] Zhai TY, Gu ZJ, Dong Y, Zhong HZ, Ma Y, Fu HB, et al. Synthesis and cathodoluminescence of morphology-tunable SiO
2
nanotubes and ZnS/SiO
2
coreshell structures using CdSe nanocrystals as the seeds. J Phys Chem C 2007;111:
1160411.
[180] Su Y, Li L, Liang XM, Zhou QT, Gao M, Chen YQ, et al. Structure and optical properties of ZnS/SiO
2
nanocables. Mater Lett
2008;62:33102.
[181] Moore D, Morber JR, Snyder RL, Wang ZL. Growth of ultralong ZnS/SiO
2
nanowires by volume and surface diffusion VLS
process. J Phys Chem C 2008;112:2895903.
[182] He JH, Zhang YY, Liu J, Moore D, Bao G, Wang ZL. ZnS/silica nanocable eld effect transistors as biological and chemical
nanosensors. J Phys Chem C 2007;111:121526.
[183] Shen GZ, Bando Y, Tang CC, Golberg D. Self-organized hierarchical ZnS/SiO
2
nanowire heterostructures. J Phys Chem B
2006;110:7199202.
[184] Li Y, Ye CH, Fang XS, Yang L, Xiao YH, Zhang LD. Fabrication and photoluminescence of SiO2-sheated semiconducting
nanowires: the case of ZnS/SiO
2
. Nanotechnology 2005;16:5015.
[185] Chen ZH, Tang H, Fan X, Jie JS, Lee CS, Lee ST. Epitaxial ZnS/Si coreshell nanowires and single-crystal silicon tube eld-
effect transistors. J Cryst Growth 2008;310:16570.
[186] Hu JQ, Bando Y, Liu ZW, Zhan JH, Golberg D, Sekiguchi T. Synthesis of crystalline silicon tubular nanostructures with ZnS
nanowires as removable templates. Angew Chem Int Ed 2004;43:636.
[187] Shen XP, Jiang ZY, Gao GL, Xu Z, Xie ZX, Zheng LS. One-step construction of ZnS/C and CdS/C one-dimensional coreshell
nanostructures. J Mater Chem 2007;17:132630.
[188] Shen GZ, Bando Y, Golberg D. Carbon-coated single-crystalline zinc sulde nanowires. J Phys Chem B
2006;110:2077780.
[189] Gautam UK, Fang XS, Bando Y, Zhan JH, Golberg D. Synthesis, structure, and multiply enhanced eld-emission properties
of branched ZnS nanotube-In nanowire coreshell heterostructures. ACS Nano 2008;2:101521.
[190] Hu JQ, Bando Y, Zhan JH, Golberg D. Sn-lled single-crystalline wurtzite-type ZnS nanotubes. Angew Chem Int Ed
2004;43:46069.
[191] Zhu YC, Bando Y, Uemura Y. ZnSZn nanocables and ZnS nanotubes. Chem Commun 2003:8367.
[192] Li Q, Wang CR. Fabrication of Zn/ZnS nanocable heterostructures by thermal reduction/suldation. Appl Phys Lett
2003;82:1398400.
[193] Shen GZ, Bando Y, Gao YH, Golberg D. Synthesis and interface structures of zinc sulde sheathed zinccadmium nanowire
heterojunctions. J Phys Chem B 2006;110:141237.
[194] Shen GZ, Ye CH, Golberg D, Hu JQ, Bando Y. Structure and cathodoluminescence of hierarchical Zn3P2/ZnS nanotube/
nanowire heterostructures. Appl Phys Lett 2007;90:073115.
[195] Kar S, Santra S, Heinrich H. Fabrication of high aspect ratio coreshell CdS-Mn/ZnS nanowires by a two step solvothermal
process. J Phys Chem C 2008;112:403641.
[196] Li H, Zhu BL, Feng YF, Wang SR, Zhang SM, Huang WP. Preparation of TiO
2
/ZnS core/sheath heterostructure nanotubes via
a wet chemical method and their photocatalytic activity. React Kinet Catal Lett 2007;92:23946.
[197] Zhang H, Xie RG, Sekiguchi T, Ma XY, Yang DR. Cathodoluminescence and its mapping of ower-like ZnO, ZnO/ZnS core
shell and tube-like ZnS nanostructures. Mater Res Bull 2007;42:128692.
[198] Sulieman KM, Huang XT, Liu JP, Tang M. One-step growth of ZnO/ZnS coreshell nanowires by thermal evaporation.
Smart Mater Struct 2007;16:8992.
[199] Sun CW, Jeong JS, Lee JY. Microstructural analysis of ZnO/ZnS nanocables through Moir fringe induced by overlapped
area of ZnO and ZnS. J Cryst Growth 2006;294:1627.
[200] Li JH, Zhao DX, Meng XQ, Zhang ZZ, Zhang JY, Shen DZ, et al. Enhanced ultraviolet emission from ZnS-coated ZnO
nanowires fabricated by self-assembling method. J Phys Chem B 2006;110:146857.
[201] Liao HC, Kuo PC, Lin CC, Chen SY. Synthesis and optical properties of ZnOZnS coreshell nanotube arrays. J Vac Sci
Technol B 2006;24:2198201.
[202] Hsu YJ, Lu SY, Lin YF. One-step preparation of coaxial CdSZnS and Cd
1x
Zn
x
SZnS nanowires. Adv Funct Mater
2005;15:13507.
[203] Hsu YJ, Lu SY. One-step preparation of coaxial CdSZnS nanowires. Chem Commun 2004:21023.
[204] Jiang DX, Cao LX, Su G, Liu W, Qu H, Sun YG, et al. Synthesis and luminescence properties of ZnS:Mn/ZnS core/shell
nanorod structures. J Mater Sci 2009;44:27925.
[205] Mokari T, Banin U. Synthesis and properties of CdSe/ZnS core/shell nanorods. Chem Mater 2003;15:395560.
[206] Li JH, Zhao DX, Meng XQ, Zhang ZZ, Zhang JY, Shen DZ, et al. Enhanced ultraviolet emission from ZnS-coated ZnO
nanowire fabricated by self-assembling method. J Phys Chem B 2006;110:146857.
[207] He RR, Law M, Fan R, Kim F, Yang PD. Functional bimorph composite nanotapes. Nano Lett 2002;2:110912.
[208] Hu JQ, Bando Y, Liu ZW, Sekiguchi T, Golberg D, Zhan JH. Epitaxial heterostructures: side-to-side Si-ZnS, Si-ZnSe biaxial
nanowires, and sandwichlike ZnS-Si-ZnS triaxial nanowires. J Am Chem Soc 2003;125:1130613.
[209] Yan J, Fang XS, Zhang LD, Bando Y, Gautam UK, Dierre B, et al. Structure and cathodoluminescence of individual ZnS/ZnO
biaxial nanobelt heterostructures. Nano Lett 2008;8:27949.
[210] Fang XS, Bando Y, Gautam UK, Zhai TY, Gradeck, Golberg D. Heterostructures and superlattices in one-dimensional
nanoscale semiconductors. J Mater Chem 2009;19:56839.
280 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
[211] Zhan JH, Bando Y, Hu JQ, Sekiguchi T, Golberg D. Single-catalyst conned growth of ZnS/Si composite nanowires. Adv
Mater 2005;17:22530.
[212] Murphy MW, Zhou XT, Ko JYP, Zhou JG, Heigl F, Sham TK. Optical emission of biaxial ZnOZnS nanoribbon
heterostructures. J Chem Phys 2009;130:084707.
[213] Murphy MW, Grace Kim PS, Zhou XT, Zhou JG, Coulliard M, Botton GA, et al. Biaxial ZnOZnS nanoribbons
heterostructures. J Phys Chem C 2009;113:47557.
[214] Wang ZQ, Liu XD, Gong JF, Huang HB, Gu SL, Yang SG. Epitaxial growth of ZnO nanowires on ZnS Nanobelts by metal
organic chemical vapor deposition. Cryst Growth Des 2008;8:3911.
[215] Fan X, Zhang ML, Shaq I, Zhang WJ, Lee CS, Lee ST. ZnS/ZnO heterojunctions nanoribbons. Adv Mater 2009;21:23936.
[216] Kang T, Sung J, Shim W, Moon H, Cho J, Jo Y, et al. Synthesis and magnetic properties of single-crystalline Mn/Fe-doped
and Co-doped ZnS nanowires and nanobelts. J Phys Chem C 2009;113:53527.
[217] Yuan GD, Zhang WJ, Zhang WF, Fan X, Bello I, Lee CS, et al. Appl Phys Lett 2008;93:213102.
[218] Kar S, Biswas S. White light emission from surface-oxidized manganese-doped ZnS nanorods. J Phys Chem C
2008;112:111449.
[219] Li YQ, Zapien JA, Shan YY, Liu YK, Lee ST. Manganese doping and optical properties of ZnS nanoribbons by postannealing.
Appl Phys Lett 2006;88:013115.
[220] Datta A, Panda SK, Chaudhuri S. Phase transformation and optical properties of Cu-doped ZnS nanorods. J Solid State
Chem 2008;181:23327.
[221] Yang WQ, Dai L, You LP, Qin GG. Color tuning of photoluminescence from ZnS nanobelts synthesized with Cu and Mn
doping and without intentionally doping. Phys Lett A 2008;372:48314.
[222] Gautam UK, Bando Y, Zhan JH, Costa PMFJ, Fang XS, Golberg D. Ga-doped ZnS nanowires as precursors for ZnO/ZnGa
2
O
4
nanotubes. Adv Mater 2008;20:8104.
[223] Liu JZ, Yan PX, Yue GH, Chang JB, Qu DM, Zhuo RF. Red light photoluminescence emission from Mn and Cd co-doped ZnS
one-dimensional nanostructures. J Phys D: Appl Phys 2006;39:23526.
[224] Cheng BC, Wang ZG. Synthesis and optical properties of Europium-doped ZnS: long-lasing phosphorescence from aligned
nanowires. Adv Funct Mater 2005;15:188390.
[225] Manzoor K, Aditya V, Vadera SR, Kumar N, Kutty TRN. Enhanced electroluminescence properties of doped ZnS nanorods
formed by the self-assembly of colloidal nanocrystals. Solid State Commun 2005;135:1620.
[226] Geng BY, Zhang LD, Wang GZ, Xie T, Zhang YG, Meng GW. Synthesis and photoluminescence properties of ZnMnS
nanobelts. Appl Phys Lett 2004;84:21579.
[227] Zhai TY, Zhang XZ, Yang WS, Ma Y, Wang JF, Gu ZJ, et al. Growth of single crystalline Zn
x
Cd
1x
S nanocombs by metallo-
organic chemical vapor deposition. Chem Phys Lett 2006;427:3714.
[228] Liu YK, Zapien JA, Shan YY, Geng CY, Lee CS, Lee ST. Wavelength-controlled lasing in Zn
x
Cd
1x
S single-crystal nanoribbons.
Adv Mater 2005;17:13727.
[229] Venugopal R, Lin PI, Chen YT. Photoluminescence and Raman scattering from catalytically grown Zn
x
Cd
1x
Se alloy
nanowires. J Phys Chem B 2006;110:116916.
[230] Wang M, Fei GT, Zhang YG, Kong MG, Zhang LD. Tunable and predetermined bandgap emissions in alloyed ZnS
x
Se
1x
nanowires. Adv Mater 2007;19:44914.
[231] Lu MY, Chen LJ, Mai WJ, Wang ZL. Tunable electric and magnetic properties of Co
x
Zn
1x
S nanowires. Appl Phys Lett
2008;93:242503.
[232] Zhang XZ, Zhai TY, Ma Y, Yao JN, Yu DP. Polarity determination for the Cd
x
Zn
1x
S nanocombs by EELS. J Electron Microsc
2008;57:711.
[233] Zhai TY, Gu ZJ, Yang WS, Zhang XZ, Huang J, Zhao YS, et al. Fabrication, structural characterization and
photoluminescence of single-crystal Zn
x
Cd
1x
S zigzag nanowires. Nanotechnology 2006;17:46449.
[234] Biswas S, Kar S, Santra S, Jompol Y, Arif M, Khondaker SI. Solvothermal synthesis of high-aspect ratio alloy semiconductor
nanowires: Cd
1x
Zn
x
S, a case study. J Phys Chem C 2009;113:361724.
[235] Chen YQ, Zhou QT, Zhang XH, Su Y, Jia C, Li Q, et al. Ga-catalyzed growth and optical properties of ternary Si-ZnS
nanowires. Cryst Growth Des 2009;9:72831.
[236] Chang YQ, Yu DP, Zhang HZ, Wang Z, Long Y, Qiang WJ. Fabrication and characterization of single-crystalline
nanostructured of Zn
1x
Mn
x
S. Nanotechnology 2006;17:19992003.
[237] Yuan HJ, Yan XQ, Zhang ZX, Liu DF, Zhou ZP, Cao L, et al. Synthesis, optical, and magnetic properties of Zn
1x
Mn
x
S
nanowires growth by thermal evaporation. J Cryst Growth 2004;271:4038.
[238] Chen L, Niebling T, Heimbrodt W, Stichenoth D, Ronning C, Klars PJ. Dimensional dependence of the dynamics of the Mn
3d
5
luminescence in (Zn, Mn)S nanowires and nanobelts. Phys Rev B 2007;76:115325.
[239] Biswas S, Kar S, Chaudhuri S. Optical and magnetic properties of manganese-incorporated zinc sulde nanorods
synthesized by a solvothermal process. J Phys Chem B 2005;109:1752630.
[240] Xu HY, Liang Y, Liu Z, Zhang XT, Hark S. Synthesis and optical properties of tetrapod-like ZnSSe alloy nanostructures. Adv
Mater 2008;20:32947.
[241] Shan CX, Liu Z, Ng CM, Hark SK. Zn
x
Cd
1x
Se alloy nanowires covering the entire compositional range grown by
metalorganic chemical vapor deposition. Appl Phys Lett 2005;87:033108.
[242] Pan AL, Liu RB, Sun MH, Ning CZ. Quanternary alloy semiconductor nanobelts with bandgap spanning the entire visible
spectrum. J Am Chem Soc 2009;131:95023.
[243] Manna L, Scher EC, Alivisatos AP. Synthesis of soluble and processable rod-, arrow-, teardrop-, and tetrapod-shaped CdSe
nanocrystals. J Am Chem Soc 2000;122:127006.
[244] Gong JF, Yang SG, Huang HB, Duan JH, Liu HW, Zhao XN, et al. Experimental evidence of an octahedron nucleus in ZnS
tetrapods. Small 2006;2:7325.
[245] Zhai TY, Dong Y, Wang YB, Cao ZW, Ma Y, Fu HB, et al. Size-tunable synthesis of tetrapod-like ZnS nanopods by seed-
epitaxial metalorganic chemical vapor deposition. J Solid State Chem 2008;181:9506.
[246] Zhai TY, Zhong HZ, Gu ZJ, Peng AD, Fu HB, Ma Y, et al. Manipulation of the morphology of ZnSe sub-micron structures
using CdSe nanocrystals as the seeds. J Phys Chem C 2007;111:29806.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 281
[247] Zhai TY, Gu ZJ, Zhong HZ, Dong Y, Ma Y, Fu HB, et al. Design and fabrication of rocketlike tetrapodal CdS nanorods by seed-
epitaxial metalorganic chemical vapor deposition. Cryst Growth Des 2007;7:48891.
[248] Zhu YC, Bando Y, Xue DF, Golberg D. Nanocable-aligned ZnS tetrapod nanocrystals. J Am Chem Soc 2003;125:161967.
[249] Deng ZQ, Qi JJ, Zhang Y, Liao QL, Huang YH. Growth mechanism and optical properties of ZnS nanotetrapods.
Nanotechnology 2007;18:475603.
[250] Shen GZ, Bando Y, Hu JQ, Golberg D. High-symmetry ZnS hepta- and tetrapods composed of assembled ZnS nanowire
arrays. Appl Phys Lett 2007;90:123101.
[251] Hao YF, Meng GW, Wang ZL, Ye CH, Zhang LD. Periodically twinned nanowires and polytypic nanobelts of ZnS: the role of
mass diffusion in vaporliquidsolid growth. Nano Lett 2006;6:16505.
[252] Liu BD, Bando Y, Liao MY, Tang CC, Mitome M, Golberg D. Bicrystalline ZnS microbelts. Cryst Growth Des 2009;9:27903.
[253] Fan X, Meng XM, Zhang XH, Shi WS, Zhang WJ, Zapien JA, et al. Dart-shaped tricrystal ZnS nanoribbons. Angew Chem Int
Ed 2006;45:256871.
[254] Meng XM, Jiang Y, Liu J, Lee CS, Bello I, Lee ST. Synthesis and characterization of ZnS bicrystal nanoribbons. Appl Phys Lett
2003;83:22446.
[255] Zhang J, Jiang FH. Temperature-dependent photoluminescence from bicrystalline ZnS (core/shell) nanocables. Phys Lett A
2008;372:547983.
[256] Geng BY, Liu XW, Du QB, Wei XW, Zhang LD. Structure and optical properties of periodically twinned ZnS nanowires. Appl
Phys Lett 2006;88:163104.
[257] Jie JS, Zhang WJ, Jiang Y, Meng XM, Zapien JA, Shao MW, et al. Heterocrystal and bicrystal structures of ZnS nanowires
synthesized by plasma enhanced chemical vapour deposition. Nanotechnology 2006;17:29137.
[258] Zhu YC, Ruan QC, Xu FF. Crystallographically-oriented nanoassembly by ZnS tricrystals and subsequent three-
dimensional epitaxy. Nano Res 2009;2:68894.
[259] Yan J, Wang ZM, Zhang LD. Effect of stacking fault on the formation of the saw-teeth of ZnS nanosaws. Cryst Growth Des
2008;8:17236.
[260] Yan HQ, He RR, Johnson J, Law M, Saykally RJ, Yang PD. Dendritic nanowire ultraviolet laser arrays. J Am Chem Soc
2003;125:47289.
[261] May SJ, Zheng JG, Wessels BW, Lauhon LJ. Dendritic nanowire growth mediated by a self-assembled catalyst. Adv Mater
2005;17:598602.
[262] Dick KA, Deppert K, Larsson MW, Mrtensson T, Seifert W, Wallenberg LR, et al. Synthesis of branched nanotrees by
controlled seeding of multiple branching events. Nat Mater 2004;3:3804.
[263] Jiang Y, Zhang WJ, Jie JS, Meng XM, Zapien JA, Lee ST. Homoepitaxial growth and lasing properties of ZnS nanowire and
nanoribbons arrays. Adv Mater 2006;18:152732.
[264] Jung YW, Ko DK, Agarwal R. Synthesis and structural characterization of single-crystalline branched nanowire
heterostructures. Nano Lett 2007;7:2648.
[265] Zhao JW, Qin LR, Zhang LD. Fabrication of ZnS/ZnO hierarchical nanostructures by two-step vapor phase method. Mater
Res Bull 2009;44:10038.
[266] Shaq I, Sharif A, Sing LC. ZnS
x
Se
1x
nanowire arrays with tunable optical properties growth on ZnS nanoribbons
substrates. Phys E 2009;41:73945.
[267] Li YQ, Tang JX, Wang H, Zapien JA, Shan YY, Lee ST. Heteroepitaxial growth and optical properties of ZnS nanowire arrays
on CdS nanoribbons. Appl Phys Lett 2007;90:093127.
[268] Shen GZ, Chen D, Lee CJ. Hierarchical saw-like ZnO nanobelt/ZnS nanowire heterostructures induced by polar surfaces. J
Phys Chem B 2006;110:1568993.
[269] Li YQ, Zou K, Shan YY, Zapien JA, Lee ST. Catalyst-assisted formation of nanocantilever arrays on ZnS nanoribbons by post-
annealing treatment. J Phys Chem B 2006;110:675962.
[270] Fang XS, Bando Y, Ye CH, Shen GZ, Golberg D. Shape- and size-controlled growth of ZnS nanostructures. J Phys Chem C
2007;111:846974.
[271] Zhang J, Yang YD, Jiang FH, Li JP, Xu BL, Wang XC, et al. Fabrication, structural characterization and photoluminescence of
Q-1D semiconductor ZnS hierarchical nanostructures. Nanotechnology 2006;17:2695700.
[272] Chen ZG, Zou J, Liu G, Yao XD, Li F, Yuan XL, et al. Growth, cathodoluminescence and eld emission of ZnS tetrapod tree-
like heterostructures. Adv Funct Mater 2008;18:30639.
[273] Moore D, Ronning C, Ma C, Wang ZL. Wurtzite ZnS nanosaws produced by polar surfaces. Chem Phys Lett 2004;385:811.
[274] Chen ZG, Zou J, Wang DW, Yin LC, Liu G, Liu QF, et al. Field emission and cathodoluminescence of ZnS hexagonal pyramids
of zinc blende structured single crystals. Adv Funct Mater 2009;19:48490.
[275] Zhai TY, Gu ZJ, Fu HB, Ma Y, Yao JN. Synthesis of single-crystal ZnS nanoawls via two-step pressure-controlled vapor-
phase deposition and their optical properties. Cryst Growth Des 2007;7:138892.
[276] Fang XS, Gautam UK, Bando Y, Dierre B, Sekiguchi T, Golberg D. Multiangular branched ZnS nanostructures with needle-
shaped tips: potential luminescent and eld-emitter nanomaterial. J Phys Chem C 2008;112:473542.
[277] Zhang HJ, Qi LM. Low-temperature, template-free synthesis of wurtzite ZnS nanostructures with hierarchical
architecture. Nanotechnology 2006;17:39848.
[278] Zhao QR, Xie Y, Zhang ZG, Bai X. Size-selective synthesis of zinc sulde hierarchical structures and their photocatalytic
activity. Cryst Growth Des 2007;7:1538.
[279] Moore D, Ding Y, Wang ZL. Hierarchical structured nanohelices of ZnS. Angew Chem Int Ed 2006;45:51504.
[280] Mann S, Ozin GA. Synthesis of inorganic materials with complex form. Nature 1996;382:3138.
[281] Liu YK, Xi GC, Chen SF, Zhang XF, Zhu YC, Qian YT. New ZnS/organic composite nanoribbons: characterization, thermal
stability and photoluminescence. Nanotechnology 2007;18:285605.
[282] Li JP, Xu Y, Wu D, Sun YH. Hydrothermal synthesis of novel sandwich-like structured ZnS/octylamine hybrid nanosheets.
Solid State Commun 2004;130:61922.
[283] Goswami A, Goswami AP. Dielectric and optical properties of ZnS lms. Thin Solid Films 1973;16:17585.
[284] Nicolau YF, Dupuy M, Brunel M. ZnS, CdS, and Zn
1x
Cd
x
S thin lms deposited by the successive inoic layer adsorption and
reaction process. J Electrochem Soc 1990;137:291524.
282 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
[285] Nicolau YF, Menard JC. Solution growth of ZnS, CdS and Zn
1x
Cd
x
S thin lms by the successive ionic-layer adsorption and
reaction process; growth mechanism. J Cryst Growth 1988;92:12842.
[286] Valkonen MP, Kanniainen T, Lindroos S, Leskela M, Rauhala E. Growth of ZnS, CdS and multilayer ZnS/CdS thin lms by
SILAR technique. Appl Surf Sci 1997;115:38692.
[287] Frigo DM, Khan OFZ, Brien PO. Growth of epitaxial and highly oriented thin lms of cadmium and cadmium zinc sulde
by low-pressure metalorganic chemical vapour deposition using diethyldithiocarbamates. J Cryst Growth
1989;96:98992.
[288] Yoshikawa A, Yamaga S, Tanaka K, Kasai H. Growth of low-resistivity high-quality ZnSe, ZnS lms by low-pressure
metalorganic chemical vapor deposition. J Cryst Growth 1985;72:136.
[289] McClean IP, Thomas CB. Photoluminescence study of MBE-growth lms on ZnS. Semicond Sci Technol 1992;7:13949.
[290] Pssler R, Griebl E, Riepl H, Lautner G, Bauer S, Preis H, et al. Temperature dependence of exciton peak energies in ZnS,
ZnSe, and ZnTe epitaxial lms. J Appl Phys 1999;86:440311.
[291] Eildrissi B, Addou M, Regragui M, Bougrine A, Kachouane A, Bernde JC. Structure, composition and optical properties of
ZnS thin lms prepared by spray pyrolysis. Mater Chem Phys 2001;68:1759.
[292] A HH, Mahmoud SA, Ashour A. Structural study of ZnS thin lms prepared by spray pyrolysis. Thin Solid Films
1995;263:24851.
[293] Dona JM, Herrero J. Process and lm characterization of chemical-bath-deposited ZnS thin lms. J Electrochem Soc
1994;141:20510.
[294] Arenas OL, Nair MTS, Nair PK. Chemical bath deposition of ZnS thin lms and modication by air annealing. Semicond Sci
Technol 1997;12:132330.
[295] Murali KP, Vasantha S, Rajamma K. Properties of pulse plated ZnS lms. Mater Lett 2008;62:18236.
[296] Yano S, Schroeder R, Sakai H, Ullrich B. High-electric-eld photocurrent in thin-lm ZnS formed by pulsed-laser
deposition. Appl Phys Lett 2003;82:20268.
[297] Lindroos S, Kanniainen T, Leskel M. Growth of ZnS thin lms by liquid-phase atomic layer epitaxy (LPALE). Appl Surf Sci
1994;75:704.
[298] Ihanus J, Ritala M, Leskel M, Prohaska T, Resch R, Freidbacher G, et al. AFM studies on ZnS thin lms grown by atomic
layer epitaxy. Appl Surf Sci 1997;120:4350.
[299] Shao LX, Chang KH, Hwang HL. Zinc sulde thin lms deposited by RF reactive sputtering for photovoltaic applications.
Appl Surf Sci 2003;212213:30510.
[300] Tang W, Cameron DC. Electroluminescent zinc sulphide devices produced by solgel processing. Thin Solid Films
1996;280:2216.
[301] Kovtyukhova NI, Buzaneva EV, Waraksa CC, Martin BR, Mallouk TE. Surface solgel synthesis of ultrathin semiconductor
lms. Chem Mater 2000;12:3839.
[302] Lee EYM, Tran NH, Russell JJ, Lamb RN. Structure evolution in chemical vapor-deposited ZnS lms. J Phys Chem B
2003;107:520811.
[303] Cheon JW, Talaga DS, Zink JI. Photochemical deposition of ZnS from the gas phase and simultaneous luminescence
detection of photofragments from a single-source precursor, Zn(S
2
COCHMe
2
)
2
. J Am Chem Soc 1997;119:
1638.
[304] Miao WJ. Electrogenerated chemiluminescence and its biorelated applications. Chem Rev 2008;108:250653.
[305] Fang XS, Bando Y, Golberg D. Recent progress in one-dimensional ZnS nanostructures. J Mater Sci Technol
2008;24:5129.
[306] Zhai TY, Fang XS, Bando Y, Dierre B, Liu BD, Zeng HB, et al. Characterization, cathodoluminescence, and eld-emission
properties of morphology-tunable CdS micro/nanostructures. Adv Funct Mater 2009;19:242330.
[307] Gautam UK, Panchakarla LS, Dierre B, Fang XS, Bando Y, Sekiguchi T, et al. Solvothermal synthesis, cathodoluminescence,
and eld-emisision properties of pure and N-doped ZnO nanobullets. Adv Funct Mater 2009;19:13140.
[308] Bringuier E. Tentative anatomy of ZnS-type electroluminescence. J Appl Phys 1994;75:291312.
[309] Frankle DR. Electroluminescence of ZnS single-crystal with cathode barrier. Phys Rev 1958;111:15409.
[310] Piper WW, Williams FE. Electroluminescence of single-crystals of ZnS:Cu. Phys Rev 1952;87:1512.
[311] Pipinys P, Proskura A, Rimeika A. Temperature dependence of electroluminescence in ZnS ceramics. Phys Status Sol (A)
1982;72:5114.
[312] Golovin YI, Morgunov RB, Baskakov AA, Shmurak SZ. Effect of a magnetic eld on the electroluminescence intensity of
single-crystal ZnS. Phys Solid State 1999;41:17835.
[313] Ono YA. In: Trigg GL, editor. Encyclopedia of applied physics, vol. 5. Weinham: VCH; 1993. p. 295325.
[314] Gurin TN, Sabitov OY. Electroluminescence parameters of thin-lm ZnS:Mn electroluminescent devices. Tech Phys
2006;51:101224.
[315] Gurin NT, Shlyapin AV, Sabitov OY. Electroluminescence kinetics thin-lm ZnS-based devices at ultralow frequencies.
Tech Phys 2002;47:21523.
[316] Rogach AL, Gaponik N, Lupton JM, Bertoni C, Gallardo DE, Dunn S, et al. Light-emitting diodes with semiconductor
nanocrystals. Angew Chem Int Ed 2008;47:653849.
[317] Ono YA. Electroluminescent displays. River Edge (NJ): World Scientic; 1995.
[318] Keir JP, Wager JF. Electrical characterization of thin-lm electroluminescent devices. Annu Rev Mater Sci
1997;27:22348.
[319] Dur M, Goodnick SM, Pennathur SS, Wager JF, Reigrotzki M, Redmer R. High-eld transport and electroluminescence in
ZnS phosphor layers. J Appl Phys 1998;83:317685.
[320] Toyama T, Hama T, Adachi D, Nakashizu Y, Okamoto H. An electroluminescence device for printable electronics using
coprecipitated ZnS:Mn nanocrystal ink. Nanotechnology 2009;20:055203.
[321] Adachi D, Takei K, Toyama T, Okamoto H. Excitation mechanism of luminescence centers in nanostructured ZnS:Tb, F
thin-lm electroluminescent devices. Jpn J Appl Phys 2007;47:836.
[322] Chan W, Sammynaiken R, Huang Y, Malm JO, Wallenberg R, Bovin JO, et al. Crystal eld, phonon coupling and emission
shift of Mn
2+
in ZnS:Mn nanoparticles. J Appl Phys 2001;89:11209.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 283
[323] Wood V, Halpert JE, Panzer MJ, Bawendi MG, Bulovic V. Alternating current driven electroluminescence from ZnSe/
ZnS:Mn/ZnS nanocrystals. Nano Lett 2009;9:236771.
[324] Yang H, Holloway PH. Electroluminescence from hybrid conjugated polymer-CdS:Mn/ZnS coreshell nanocrystals
devices. J Phys Chem B 2003;107:970510.
[325] Yang Y, Huang J, Yang B, Liu S, Shen J. Electroluminescence from ZnS/CdS nanocrystal/polymer composite. Synth Met
1997;91:3479.
[326] Yang Y, Huang J, Liu S, Shen J. Preparation, characterization and electroluminescence of ZnS nanocrystals in a polymer
matrix. J Mater Chem 1997;7:1313.
[327] Adachi D, Morimoto T, Hama T, Toyama T, Okamoto O. Orange electroluminescence from chemically synthesized zinc
sulde nanocrystals doped with manganese. J Non-Cryst Solids 2008;354:27403.
[328] Rizzo A, Mazzeo M, Biasiucci M, Cingolani R, Gigli G. White electroluminescence from a microcontact-printing-deposited
CdSe/ZnS colloidal quantum-dot monolayer. Small 2008;4:21437.
[329] Nien YT, Chen IG. Raman scattering and electroluminescence of ZnS:Cu,Cl phosphor powder. Appl Phys Lett
2006;89:261906.
[330] Horii Y, Kitagawa M, Taneoka H, Kusano H, Murakami T, Hino Y, et al. Electroluminescence properties of PVCz
electroluminescent devices doped with nano-crystalline particles. Mater Sci Eng B 2001;85:925.
[331] Chander H. Development of nanophosphors a review. Mater Sci Eng R 2005;49:11355.
[332] Yang H, Holloway PH, Ratna BB. Photoluminescent and electroluminescent properties of Mn-doped ZnS nanocrystals. J
Appl Phys 2003;93:58692.
[333] Li D, Clark BL, Keszler DA, Keir P, Wager JF. Color control in sulde phosphors: turning up the light for electroluminescent
displays. Chem Mater 2000;12:26870.
[334] Leatherdale CA, Kagan CR, Morgan NY, Empedocles SA, Kastner MA, Bawendi MG. Photoconductivity in CdSe quantum dot
solids. Phys Rev B 2000;62:266980.
[335] Yang Y, Huang JM, Liu SY, Shen JC. Preparation, characterization, and electroluminescence of ZnS nanocrystals in a
polymer matrix. J Mater Chem 1997;7:1313.
[336] Shen LH, Cui XX, Qi HL, Zhang CX. Electrogenerated chemiluminescence of ZnS nanoparticles in alkaline aqueous solution.
J Phys Chem C 2007;111:81725.
[337] Richter MM. Electrochemiluminescence (ECL). Chem Rev 2004;104:300336.
[338] Wang XF, Xu JJ, Chen HY. A new electromiluminescence emission of Mn
2+
-doped ZnS nanocrystals. J Phys Chem C
2008;112:175815.
[339] Qi HL, Peng YG, Gao Q, Zhang CX. Applications of nanomaterials in electrogenerated chemiluminescence biosensors.
Sensors 2009;9:67495.
[340] Chen W, Wang ZG, Lin ZJ, Lin LY. Thermoluminescence of ZnS nanoparticles. Appl Phys Lett 1997;70:14657.
[341] Chen W, Wang ZG, Lin ZJ, Lin LY. Absorption and luminescence of the surface states in ZnS nanoparticles. J Appl Phys
1997;82:31115.
[342] Spanhel L, Haase M, Weller H, Henglein A. Photochemistry of colloidal semiconductor. 20. Surface modication and
stability of strong luminescing CdS particles. J Am Chem Soc 1987;109:564955.
[343] Ye CH, Fang XS, Wang M, Zhang LD. Temperature-dependent photoluminescence from elemental sulfur species on ZnS
nanobelts. J Appl Phys 2006;99:063504.
[344] Ogino T, Aoki M. Mechanism of yellow luminescence in GaN. Jap J Appl Phys 1980;19:2395405.
[345] Gibbons DJ, Spear WE. Electron hopping transport and trapping phenomena in orthorhombic sulphur crystals. J Phys
Chem Solids 1966;27:191725.
[346] Sooklal K, Cullum BS, Angel SM, Murphy CJ. Photophysical properties of ZnS nanoclusters with spatically localized Mn
2+
. J
Phys Chem 1995;100:45515.
[347] Zhuo RF, Feng HT, Yan D, Chen JT, Feng JJ, Liu JZ, et al. Rapid growth and photoluminescence properties of doped ZnS one-
dimensional nanostructures. J Cryst Growth 2008;310:32406.
[348] Bhargava RN, Gallagher D. Optical properties of manganese-doped nanocrystal of ZnS. Phys Rev Lett 1994;72:
4169.
[349] Yang P, L M, X D, Yuan D, Chang J, Zhou G, et al. Strong green luminescence of Ni
2+
-doped ZnS nanocrystal. Appl Phys A
2002;74:2579.
[350] Fowler RH, Nordheim L. Electron emission in intense electric elds. Proc Roy Soc Lond A 1928;119:17381.
[351] Rinzler AG, Hafner JH, Nikolaev P, Lou L, Kim SG, Tomanek D, et al. Unraveling nanotubes: eld emission from an atomic
wire. Science 1995;269:15503.
[352] de Heer WA, Chatelain A, Ugarte D. A carbon nanotube eld-emission electron source. Science 1995;270:117980.
[353] Chinnock C. Nanotubes could power eld-emission displays. Laser Focus World 1996;2:224.
[354] Bonard JM, Salvetat JP, Stockli T, de Heer WA, Forro L, Chatelain A. Field emission from single-wall carbon nanotube lms.
Appl Phys Lett 1998;73:91820.
[355] Bonard JM, Kind H, Stockli T, Nilsson LA. Field emission from carbon nanotubes: the rst ve years. Solid State Electron
2001;45:893914.
[356] Jo SH, Tu Y, Huang ZP, Carnahan DL, Wang DZ, Ren ZF. Effect of length and spacing of vertically aligned carbon nanotubes
on eld emission properties. Appl Phys Lett 2003;82:35202.
[357] Zhu HW, Wei BQ. Assembly and applications of carbon nanotube thin lms. J Mater Sci Technol 2008;24:44756.
[358] Fan SS, Chapline MG, Franklin NR, Tombler TW, Cassell AM, Dai HJ. Self-oriented regular arrays of carbon nanotubes and
their eld emission properties. Science 1999;283:5124.
[359] Ruyven LJV, Williams FE. Valence-band bending to the fermi level and radiative recombination in ZnS with liquid
electrodes. Phys Rev Lett 1966;16:88990.
[360] Li XJ, Jiang WF. Enhanced eld emission from a nest array of multi-walled carbon nanotubes grown on a silicon
nanoporous pillar array. Nanotechnology 2007;18:065203.
[361] Baek Y, Song Y, Yong K. Novel heteronanostructure system: hierarchical W nanothorn arrays on WO
3
nanowhiskers. Adv
Mater 2006;18:310510.
284 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
[362] Wang RC, Li CH. Improved morphologies and enhanced eld emissions of CuO nanoneedle arrays by heating ZnO coated
copper foils. Cryst Growth Des 2009;9:222934.
[363] Ye CH, Bando Y, Fang XS, Shen GZ, Golberg D. Enhanced eld emission performance of ZnO nanorods by two alternative
approaches. J Phys Chem C 2007;111:126736.
[364] Banerjee D, Jo SH, Ren ZF. Enhanced eld emission of ZnO nanowires. Adv Mater 2004;16:202832.
[365] Wang WZ, Zeng BQ, Yang J, Poudel B, Huang JY, Naughton MJ, et al. Aligned ultralong ZnO nanobelts and their enhanced
eld emission. Adv Mater 2006;18:32758.
[366] Wang XD, Ding Y, Summers CJ, Wang ZL. Large-scale synthesis of six-nanometer-wide ZnO nanobelts. J Phys Chem B
2004;108:87737.
[367] Singh JP, Tang F, Karabacak T, Lu TM, Wang GC. Enhanced cold eld emission from 1 0 0) oriented bW nanoemitters. J
Vac Sci Technol B 2004;22:104851.
[368] Huang AP, Chu PK, Wu XL. Enhanced electron eld emission from oriented columnar AlN and mechanism. Appl Phys Lett
2006;88:251103.
[369] Chattopadhyay S, Chen LJ, Chen KH. Nanotips: growth, model, and applications. Crit Rev Solid State Mater Sci
2006;31:1553.
[370] Lo HC, Das D, Hwang JS, Chen KH, Hsu CH, Chen CF, et al. SiC-capped nanotip arrays for eld emission with ultralow turn-
on eld. Appl Phys Lett 2003;83:14202.
[371] Huang GS, Wu XL, Cheng YC, Li XF, Luo SH, Feng T, et al. Fabrication and eld emission property of a Si nanotip array.
Nanotechnology 2006;17:55736.
[372] Hsieh HY, Huang SH, Liao KF, Su SK, Lai CH, Chen LJ. High-density ordered triangular Si nanopillars with sharp tips and
varied slopes: one-step fabrication and excellent eld emission properties. Nanotechnology 2007;18:505305.
[373] He JR, Yang RS, Chueh YL, Chou LJ, Chen LJ, Wang ZL. Aligned AlN nanorods with multi-tipped surfaces growth, eld-
emission, and cathodoluminescence properties. Adv Mater 2006;18:6504.
[374] Bai XD, Wang EG, Gao PX, Wang ZL. Measuring the work function at a nanobelt tip and at a nanoparticle surface. Nano
Lett 2003;3:114750.
[375] Nirmal M, Brus L. Luminescence photophysics in semiconductor nanocrystals. Acc Chem Res 1999;32:40714.
[376] Tao F, Liang Y, Yin G, Xu D, Jiang Z, Li H, et al. Concentric sub-micrometer-sized cables composed of Ni nanowires and sub-
micrometer-sized fullerene tubes. Adv Funct Mater 2007;17:112430.
[377] Lieber CM. Nanowire superlattices. Nano Lett 2002;2:812.
[378] Xiang J, Lu W, Hu Y, Wu Y, Yan H, Liebre CM. Ge/Si nanowire heterostructures as high-performance eld-effect
transistors. Nature 2006;441:48993.
[379] Huo KF, Zhang XM, Hu LS, Sun XJ, Fu JJ, Chu PK. One-step growth and eld emission properties of quasialigned TiO
2
nanowire/carbon nanocone coreshell nanostructure arrays on Ti substrates. Appl Phys Lett 2008;93:013105.
[380] Shen GZ, Chen PC, Bando Y, Golberg D, Zhou C. Pearl-like ZnS-decorated InP nanowire heterostructures and their electric
behaviors. Chem Mater 2008;20:677983.
[381] Chadi DJ. Theoretical results of dupability in large band-gap IIVI semiconductors. Jpn J Appl Phys 1999;38:26178.
[382] Chadi DJ. Predictor of p-type doping in IIVI semiconductors. Phys Rev B 1999;59:151813.
[383] Look DC, Clain B, Alivov YL, Park SJ. The future of ZnO light emitters. Phys Status Solid (A) 2004;201:220312.
[384] Iida S, Yatabe T, Kinto H. P-type conduction in ZnS grown by vapor phase epitaxy. Jpn J Appl Phys 1989;28:L5357.
[385] Georgobiani AN, Kotljarevsky MB, Kidalov VV, Rogozin IV, Aminov UA. p-Type IIVI compounds doped by rare-earth
elements. J Cryst Growth 2000;214(/215):5169.
[386] Mitsuishi I, Shibatani J, Kao MH, Yamamoto M, Yoshino J, Kukimoto H. Metalorganic vapor phase epitaxial growth of Li
doped ZnS. Jpn J Appl Phys 1990;29:L7335.
[387] Butkhuzi TV, Tchelidze TG, Chikoidze EG, Kekelidze NP. Silver doped p-type ZnS crystals. Phys Status Sol (B)
2002;229:36570.
[388] Yamamotoa T, Kishimotob S, Iida S. Control of valence states for ZnS by triple-codoping method. Physica B 2001;308
310:9169.
[389] Kohikia S, Suzuka T, Kohikia S, Suzuka T, Yamamoto T, Kishimoto S, et al. Coupling of codoped In and N impurities in
ZnS:Ag: experiment and theory. J Appl Phys 2002;91:7603.
[390] Kishimoto S, Kato A, Naito A, Sakamoto Y, Iida S. Attempts of homo pn junction formation in ZnS by impurity co-doping
with vapor phase epitaxy. Phys Status Sol (B) 2002;229:3913.
[391] Gai Y, Li J, Yao B, Xia JB. The bipolar doping of ZnS via native defects and external dopants. J Appl Phys 2009;105:113704.
[392] Yuan GD, Zhang WJ, Jie JS, Fan X, Zapien JA, Leung YH, et al. p-Type ZnO nanowire arrays. Nano Lett 2008;8:25917.
[393] Zhong Z, Qian F, Wang D, Lieber CM. Synthesis of p-type gallium nitride nanowires for electronic and photonic
nanodevices. Nano Lett 2003;3:3436.
[394] Hu JS, Ren LL, Guo YG, Liang HP, Cao AM, Wan LJ, et al. Mass production and high photocatalytic activity of ZnS
nanoporous nanoparticles. Angew Chem Int Ed 2005;44:126973.
[395] Kanemoto M, Hosokawa H, Wada Y, Murakoshi K, Yanagida S, Sakata T, et al. Semiconductor photocatalysis. Part 20. Role
of surface in the photoreduction of carbon dioxide catalysed by colloidal ZnS nanocrystallites in organic solvent. J Chem
Soc Faraday Trans 1996;92:240111.
[396] Fujiwara H, Hosokawa H, Murakoshi K, Wada Y, Yanagida S. Surface characteristics of ZnS nanocrystallites relating to
their photocatalysis for CO
2
reduction. Langmuir 1998;14:51549.
[397] Kanemoto M, Shiragami T, Pac C, Yanagida S. Semiconductor photocatalysis: effective photoreduction of carbon dioxide
catalyzed by ZnS quantum crystallites with low density of surface defects. J Phys Chem 1992;96:35216.
[398] Yanagida S, Kawakami H, Midori Y, Kizumoto H, Pac C, Wada Y. Semiconductor photocatalysis. ZnS-nanocrystallite-
catalyzed photooxidation of organic compounds. Bull Chem Soc Jpn 1995;68:181123.
[399] Kanemoto M, Shiragami T, Pac C, Yanagida S. Bacteria-like xation of carbon dioxide under UV-light irradiation with
defect-free ZnS quantum crystallites. Chem Lett 1990;19:9312.
[400] Tambwekar SV, Subrahmanyam M. Enhanced photocatalytic H
2
production over CdSZnS supported on super basic
oxides. Int J Hydrogen Energy 1998;23:7414.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 285
[401] Koca A, Sahin M. Photocatalytic hydrogen production by direct sun light from sulde/sulte solution. Int J Hydrogen
Energy 2002;27:3637.
[402] Horner G, Johne P, Kunneth R, Twardzik G, Roth H, Clark T, et al. Semiconductor type A photocatalysis: role of substrate
adsorption and the nature of photoreactive surface sites in zinc sulde catalyzed CC coupling reactions. Chem Eur J
1999;5:20817.
[403] Henglein A, Gutierrez M, Fischer H. Photochemistry of colloidal metal suldes. 6. Kinetics of interfacial reactions at ZnS
particles. Ber Bunsen Phys Chem 1984;88:1705.
[404] Inoue H, Torimoto T, Sakata T, Mori H, Yoneyama H. Effects of size quantization of zinc sulde microcrystallites on
photocatalytic reduction of carbon dioxide. Chem Lett 1990;19:14836.
[405] Hamanoi O, Kudo A. Reduction of nitrate and nitrite ions over NiZnS photocatalyst under visible light irradiation in the
presence of a sacricial reagent. Chem Lett 2002;31:8389.
[406] Kudo A, Sekizawa M. Photocatalytic H
2
evolution under visible light irradiation on Ni-doped ZnS photocatalyst. Chem
Commun 2000:13712.
[407] Kudo A, Sekizawa M. Photocatalytic H
2
evolution under visible light irradiation on Zn
1x
Cu
x
S solid solution. Catal Lett
1999;58:2413.
[408] Bredol M, Kaczmarek M. Potential of nano-ZnS as electrocatalyst. J Phys Chem A 2010;114:39505.
[409] From Wikipedia, The free encyclopedia.
[410] Fang XS, Xiong SL, Zhai TY, Bando Y, Liao MY, Gautam UK, et al. High-performance blue/ultraviolet light-sensitive ZnSe
nanobelt photodetectors. Adv Mater 2009;21:501621.
[411] Sociv C, Zhang A, Xiang B, Dayeh SA, Aplin DPR, Park J, et al. ZnO nanowire UV photodetectors with high internal gain.
Nano Lett 2007;7:10039.
[412] Fang XS, Zhang LD. Controlled growth of one-dimensional oxide nanomaterials. J Mater Sci Technol 2006;22:118.
[413] Fang XS, Ye CH, Zhang LD, Li Y, Xiao ZD. Formation and optical properties of thin and wide tin-doped ZnO nanobelts.
Chem Lett 2005;34:4367.
[414] Dutta S, Chattopadhyay S, Sarkar A, Chakrabarti M, Sanyal D, Jana D. Role of defects in tailoring structural, electrical and
optical properties of ZnO. Prog Mater Sci 2009;54:89136.
[415] Lupan O, Chow L, Chai GY, Chernyak L, Lopatiuk-Tirpak O, Heinrich H. Focused-ion-beam fabrication of ZnO nanorod-
based UV photodetector using the in-situ lift-out technique. Phys Status Sol A 2008;205:2673878.
[416] Lin YY, Chen CW, Yen WC, Su WF, Ku CH, Wu JJ. The inuence of interface modier on the performance of nanostructured
ZnO/polymer hybrid solar cells. Appl Phys Lett 2008;92:233301.
[417] Fang XS, Bando Y, Liao MY, Zhai TY, Gautam UK, Li L, et al. An efcient way to assemble ZnS nanobelts as ultraviolet-light
sensors with enhanced photocurrent and stability. Adv Funct Mater 2010;20:5008.
[418] Li YB, Valle FD, Simonnet M, Yamada I, Delaunay JJ. High-performance UV detector made of ultra-long ZnO bridging
nanowires. Nanotechnology 2009;20:045501.
[419] Prades JD, Jimenez-Diaz R, Hernandez-Ramirez F, Fernandez-Romero L, Andreu T, Cirera A, et al. Towards systematic
understanding of photodetectors based on individual metal oxide nanowires. J Phys Chem C 2008;112:1463944.
[420] Heo YW, Kang BS, Tien LC, Norton DP, Ren F, Laroche JR, et al. UV photoresponse of single ZnO nanowires. Appl Phys A
2005;80:4979.
[421] Park JY, Yun YS, Hong YS, Oh H, Kim JJ, Kim SS. Synthesis, electrical and photoresponse properties of vertically well-
aligned and epitaxial ZnO nanorods on GaN-buffered sapphire substrates. Appl Phys Lett 2005;87:123108.
[422] Mathur S, Barth S, Shen H, Pyun JC, Werner U. Size-dependent photoconductance in SnO
2
nanowires. Small
2005;7:7137.
[423] Liu ZQ, Zhang DH, Han S, Li C, Tang T, Jin W, et al. Laser ablation synthesis and electron transport studies of Tin Oxide
nanowires. Adv Mater 2003;15:17547.
[424] Chen YJ, Zhu CL, Cao MS, Wang TH. Photoresponse of SnO
2
nanobelts grown in situ on interdigital electrodes.
Nanotechnology 2007;18:285502.
[425] Feng P, Zhang JY, Li QH, Wang TH. Individual b-Ga
2
O
3
nanowires as solar-blind photodetectors. Appl Phys Lett
2006;88:153107.
[426] Zhang D, Li C, Han S, Liu X, Tang T, Jin W, et al. Ultraviolet photodetection properties of indium oxide nanowires. Appl
Phys A 2003;77:1636.
[427] Fu XQ, Wang C, Feng P, Wang TH. Anomalous photoconductivity of CeO
2
nanowire in air. Appl Phys Lett 2007;91:073104.
[428] Xue XY, Guo TL, Lin ZX, Wang TH. Individual coreshell structured ZnSnO
3
nanowires as photoconductors. Mater Lett
2008;62:13568.
[429] Feng P, Zhang JY, Wan Q, Wang TH. Photocurrent characteristics of individual ZnGa
2
O
4
nanowires. J Appl Phys
2007;102:074309.
[430] Chueh YL, Hsieh CH, Chang MT, Chou LJ, Lao CS, Song JH, et al. RuO
2
nanowires and RuO
2
/TiO
2
core/shell nanowires: from
synthesis to mechanical, optical, electrical, and photoconductive properties. Adv Mater 2007;19:1439.
[431] Han S, Jin W, Zhang DH, Tang T, Li C, Liu XL, et al. Photoconduction studies on GaN nanowire transistors under UV and
polarized UV illumination. Chem Phys Lett 2004;389:17680.
[432] Kang M, Lee JS, Sim SK, Kim H, Min B, Cho K, et al. Photocurrent and photoluminescence characteristics of networked GaN
nanowires. Jpn J Appl Phys 2004;43:686872.
[433] Son MS, Im SI, Park YS, Park CM, Kang TW, Yoo KH. Ultraviolet photodetector based on single GaN nanorod pn junctions.
Mater Sci Eng C 2006;26:8868.
[434] Zhang JY, Chen YX, Guo TL, Lin ZX, Wang TH. Sub-band-gap photoconductivity of individual-Si
3
N
4
nanowires.
Nanotechnology 2007;18:325603.
[435] Hsiao CH, Chang SJ, Wang SB, Chang SP, Li TC, Lin WJ, et al. ZnSe nanowire photodetector prepared on oxidized silicon
substrate by molecular-beam epitaxy. J Electrochem Soc 2009;156:J736.
[436] Yang L, Han J, Luo T, Li M, Huang J, Meng F, et al. Morphogenesis and crystallization of ZnS microspheres by a soft
template-assisted hydrothermal route: synthesis, growth mechanism, and oxygen sensitivity. Chem Asian J
2009;4:17480.
286 X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287
[437] Papaefstathiou I, Luque de Castro MD. Hyphenated pervaporation solid-phase preconcentration gas chromatography
for the determination of volatile organic compounds in solid samples. J Chromatogr A 1997;779:3529.
[438] Bulter JH, Battle M, Bender ML, Montzka SA, Clarke AD, Saltzman ES, et al. A record of atmospheric halocarbons during the
twentieth century from polar rn air. Nature 1999;399:74955.
[439] Collins CH, Morgano MA. Quantitative determination of several simple perhalogenated compounds by high-performance
liquid chromatography. J Chromatogr A 1999;846:3959.
[440] Liu GH, Zhu YF, Zhang XR, Xu BQ. Chemiluminescence determination of chlorinated volatile organic compounds by
conversion on nanometer TiO
2
. Anal Chem 2002;74:627984.
[441] Luo L, Chen H, Zhang L, Xu K, Lv Y. A cataluminescence gas sensor for carbon tetrachloride based on nanosized ZnS. Anal
Chim Acta 2009;635:1837.
[442] Snee PT, Somers RC, Nair G, Zimmer JP, Bawendi MG, Nocera DG. A ratiometric CdSe/ZnS nanocrystal pH sensor. J Am
Chem Soc 2006;128:133201.
[443] Mulrooney RC, Singh N, Kaur N, Callan JF. An offon sensor for uoride using luminescent CdSe/ZnS quantum dots.
Chem Commun 2009:6868.
[444] Koneswaran M, Narayanaswamy R. L-Cysteine-capped ZnS quantum dots based uorescence sensor for Cu
2+
ion. Sensor
Actuat B 2009;139:1049.
[445] Guan G, Zhang Z, Wang Z, Liu B, Gao D, Xie C. Single-hole hollow polymer microspheres toward specic high-capacity
uptake of target species. Adv Mater 2007;19:23704.
[446] Toal SJ, Trogler WC. Polymer sensors for nitroaromatic explosives detection. J Mater Chem 2006;16:287183.
[447] Tu R, Liu B, Wang Z, Gao D, Wang F, Fang Q, et al. Amine-capped ZnSMn
2+
nanocrystals for uorescence detection of
trace TNT explosive. Anal Chem 2008;80:345865.
[448] Susha AS, Javier AM, Parak WJ, Rogach AL. Colloid Surf A 2006;281:403.
[449] Dai Z, Kawde AN, Xiang Y, Belle JT, Gerlach J, Bhavanandan VP, et al. Nanoparticle-based sensing of glycanlectin
interactions. J Am Chem Soc 2006;128:100189.
[450] Aoyagi S, Kudo M. Development of uorescence change-based, reagent-less optic immunosensor. Biosens Bioelectron
2005;20:16804.
[451] Duong HD, Il Rhee J. Use of CdSe/ZnS coreshell quantum dots as energy transfer donors in sensing glucose. Talanta
2007;73:899905.
[452] Uematsu T, Taniguchi S, Torimoto T, Kuwabata S. Emission quench of water-soluble ZnSAgInS
2
solid solution
nanocrystals and its application to chemosensors. Chem Commun 2009:74857.
[453] Mohagheghpour E, Rabiee M, Moztarzadeh F, Tahriri M, Jafarbeglou M, Bizari D, et al. Controllable synthesis,
characterization and optical properties of ZnS:Mn nanoparticles as a novel biosensor. Mater Sci Eng C 2009;29:18428.
[454] Huang CP, Li YK, Chen TM. A highly sensitive system for urea detection by using CdSe/ZnS coreshell quantum dots.
Biosens Bioelectron 2007;22:18358.
[455] Wang ZL, Song JH. Piezoelectric nanogenerators based on zinc oxide nanowire arrays. Science 2006;312:2426.
[456] Wang ZL. Towards self-powered nanosystems: from nanogenerators to nanopiezotronics. Adv Funct Mater
2008;18:355367.
[457] Wang ZL. Energy harvesting using piezoelectric nanowires a correspondence on energy harvesting using nanowires?
by Alexe et al.. Adv Mater 2009;21:13115.
[458] Xin J, Zheng YQ, Shi E. Piezoelectricity of zinc-blende and wurtzite structure binary compounds. Appl Phys Lett
2007;91:112902.
[459] Ye CH, Fang XS, Hao YF, Teng XM, Zhang LD. Zinc oxide nanostructures: morphology derivation and evolution. J Phys
Chem B 2005;109:1975865.
[460] Wang SQ. First-principles study of the anisotropic thermal expansion of wurtzite ZnS. Appl Phys Lett 2006;88:061902.
[461] Fang XS, Zhang LD. One-dimensional (1D) ZnS nanomaterials and nanostructures. J Mater Sci Technol 2006;22:72136.
[462] Hamad S, Woodley SM, Catlow RA. Experimental and computational studies of ZnS nanostructures. Mol Sumulat
2009;35:101532.
[463] Zhu T, Li J. Ultra-strength materials. Prog Mater Sci 2010;55:71057.
[464] Li YF, Zhou Z, Jin P, Chen YS, Zhang SB, Chen ZF. Achieving ferromagnetism in single-crystalline ZnS wurtzite nanowires
via chromium doping. J Phys Chem C 2010;114:12099103.
[465] Feigl C, Russo SP, Barnard AS. Safe, stable and effective nanotechnology: phase mapping of ZnS nanoparticles. J Mater
Chem 2010;20:497180.
[466] Barth S, Hernandez-Ramirez F, Holmes JD, Romano-Rodriguez A. Synthesis and applications of one-dimensional
semiconductors. Prog Mater Sci 2010;55:563627.
[467] Ikuhara Y. Nanowire design by dislocation technology. Prog Mater Sci 2009;54:77091.
[468] Pearton SJ, Ren F, Wang YL, Chu BH, Chen KH, Chang CY, et al. Recent advances in wide bandgap semiconductor biological
and gas sensors. Prog Mater Sci 2010;55:159.
[469] Wei TY, Huang CT, Hansen BJ, Lin YF, Chen LJ, Lu SY, et al. Large enhancement in photon detection sensitity via Schottky-
gated CdS nanowire nanosensors. Appl Phys Lett 2010;96:013508.
X.S. Fang et al. / Progress in Materials Science 56 (2011) 175287 287

Anda mungkin juga menyukai