You are on page 1of 67

Progress in Materials Science 54 (2009) 1–67

Contents lists available at ScienceDirect

Progress in Materials Science


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

Quasi-one dimensional metal oxide semiconductors:


Preparation, characterization and application as chemical
sensors
E. Comini, C. Baratto, G. Faglia, M. Ferroni, A. Vomiero, G. Sberveglieri*
SENSOR Lab, CNR-INFM, Dipartimento di Chimica e Fisica per l’Ingegneria e per i Materiali, Brescia University,
via Valotti 9, 25133 Brescia, Italy

a r t i c l e i n f o a b s t r a c t

Article history: The continuous evolution of nanotechnology in these years led to


Accepted 16 June 2008 the production of quasi-one dimensional (Q1D) structures in a
variety of morphologies such as nanowires, core–shell nanowires,
nanotubes, nanobelts, hierarchical structures, nanorods, nanorings.
In particular, metal oxides (MOX) are attracting an increasing
interest for both fundamental and applied science. MOX Q1D are
crystalline structures with well-defined chemical composition,
surface terminations, free from dislocation and other extended
defects. In addition, nanowires may exhibit physical properties
which are significantly different from their coarse-grained poly-
crystalline counterpart because of their nanosized dimensions.
Surface effects dominate due to the increase of their specific sur-
face, which leads to the enhancement of the surface related prop-
erties, such as catalytic activity or surface adsorption: key
properties for superior chemical sensors production.
High degree of crystallinity and atomic sharp terminations make
nanowires very promising for the development of a new genera-
tion of gas sensors reducing instabilities, typical in polycrystalline
systems, associated with grain coalescence and drift in electrical
properties. These sensitive nanocrystals may be used as resistors,
and in FET based or optical based gas sensors.
This article presents an up-to-date review of Q1D metal oxide
materials research for gas sensors application, due to the great
research effort in the field it could not cover all the interesting works

* Corresponding author. Tel.: +39 030 3715771; fax: +39 030 2091271.
E-mail address: sbervegl@sensor.ing.unibs.it (G. Sberveglieri).
URL: http://sensor.ing.unibs.it (G. Sberveglieri).

0079-6425/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pmatsci.2008.06.003
2 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

reported, the ones that, according to the authors, are going to contrib-
ute to this field’s further development were selected and described.
Ó 2008 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Deposition techniques and growth mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Vapor phase growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.1. Vapor–liquid–solid mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.2. Vapor–solid mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2. Solution phase growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1. Template-assisted synthesis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2.2. Template-free methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3. Vertical and horizontal alignment techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.1. Electric field alignment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2. Nanomanipulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4. Doping of quasi 1D metal oxide nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5. Preparation of quasi 1D metal oxide heterostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
6. Applications of metal oxide nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
6.1. Metal oxide gas sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
6.1.1. Surface adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
6.1.2. Detection through surface reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6.1.3. DC resistance transduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6.1.4. Conductometric gas sensors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6.1.5. Single nanowire transistor (SNT) based gas sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.1.6. PL based gas sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.2. Other application fields. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
6.2.1. Lasers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
6.2.2. Solar cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.2.3. Field emitters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.2.4. Li-ion batteries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.2.5. Single nanowire transistors for biosensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
7. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

1. Introduction

The increasing concerns with pollution on health and safety stress the need of monitoring all as-
pects of the environment in real time, and in turn led to a tremendous effort in terms of research
and funding for the development of sensors devoted to several applications [1–9].
As far as chemical sensing is concerned, it has been known, from more than five decades, that the
electrical conductivity of metal oxides semiconductors varies with the composition of the surrounding
gas atmosphere. The sensing properties of semiconductor metal oxides in form of thin or thick films
other than SnO2, like TiO2, WO3, ZnO, Fe2O3 and In2O3, have been studied as well as the benefits from
the addition of noble metals – Pd, Pt, Au, Ag – in improving selectivity and stability.
In 1991 Yamazoe showed that reduction of crystallite size went along with a significant increase in
sensor performance [10]. In a nanosized grain metal oxide almost all the carriers are trapped in surface
states and only a few thermal activated carriers are available for conduction. In this configuration the
transition from activated to strongly not activated carrier density, produced by target gases species,
has a huge effect on sensor conductance. Thus, the technological challenge moved to the fabrication
of materials with small crystallize size which maintained their stability over long-term operation at
high temperature. A huge variety of devices have been developed mainly by an empirical approach
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 3

and a lot of basic theoretical research and spectroscopy studies have been carried out to improve the
well known ‘‘3S” of a gas sensor, namely sensitivity, selectivity and stability.
Nanotechnology is nowadays producing sensing materials such as quasi-1D metal oxides (MOX),
carbon nanotubes, and nano porous materials. In particular, metal oxides are an attractive and heter-
ogeneous class of active materials covering the entire range from metals to semiconductors and insu-
lators and almost all aspects of material science and physics in areas including superconductivity and
magnetism.
After the first publications demonstrating the ability of metal oxide nanowires in detecting a vari-
ety of chemical species [179,188], the interest in this research area was growing exponentially in the
past years as testified by literature.
Significant progress has been made both in terms of our fundamental understanding of the inter-
play between bulk and surface properties and processes in MOX nanowires sensors together with their
development as real world sensing platforms.
Q1D metal oxide nanostructures have several advantages with respect to traditional thin- and thick
film sensors such as very large surface-to-volume ratio, dimensions comparable to the extension of
surface charge region, superior stability owing to the high crystallinity [11], relatively simple prepa-
ration methods that allow large-scale production [14], possible functionalization of their surface with
a target-specific receptor species [190], modulation of their operating temperature to select the proper
gas semiconductor reactions, catalyst deposition over the surface for promotion or inhibition of spe-
cific reactions and finally the possibility of field-effect transistors (FET) configuration that allows the
use of gate potential controlling the sensitivity and selectivity [188].
Preparation and performances of these emerging nanosized structures have been reviewed by a
number of authors [12–15], but this research field is growing so fast that there is still the need of a
review focused on sensing applications.
This review article is focused on the description of metal oxide single crystalline Q1D nanostruc-
tures used for gas-sensing application, specifically on the promising approaches that are going to con-
tribute to the further development of this field. The overview will start from presenting the fabrication
techniques and the growth mechanisms, focusing on their development and improvements, and
pointing out the steps critical for application in real environments. Then the application as chemical
sensors will be addressed. Furthermore an outlook on other possible new applications of metal oxide
single crystalline nanowires will be presented.

2. Deposition techniques and growth mechanisms

Nanocrystalline materials can be classified into different categories depending on the number of
dimensions that are nanostructured (with dimensions lower than 100 nm); we will follow one of
the possible classification: i.e. zero dimensional for clusters, mono dimensional for nanowires and
two dimensional for films.
There are two different approaches to the production of 1D structures: top-down and bottom up
technologies.
The first one is based on standard micro fabrication methods with deposition, etching and ion beam
milling on planar substrates in order to reduce the lateral dimensions of the films to the nanometer
size. Electron beam, focused ion beam, X-ray lithography, nano-imprinting and scanning probe
microscopy techniques can be used for the selective removal processes. The advantages are the use
of the well developed technology of semiconductor industry and the ability to work on planar sur-
faces, while disadvantages are their extremely elevated costs and preparation times.
In the top-down approach highly ordered nanowires can be obtained [16–19], but at the moment
this technology does not fulfil the industrial requirements for the production of low cost and large
numbers of devices. Furthermore the 1D nanostructures produced with these techniques are in gen-
eral not single-crystalline.
The second approach, bottom-up, consists of the assembly of molecular building blocks or chemical
synthesis by vapor phase transport, electrochemical deposition, solution-based techniques or tem-
plate growth. Its advantages are the high purity of the nanocrystalline materials produced, their small
4 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Fig. 1. Schematic drawing of some of the possible morphologies: (a) nanowire, (b) core–shell nanowire, (c) nanotube, (d)
nanobelt, (e) hierarchical structure, (f) nanorod and (g) nanoring.

diameters, the low cost of the experimental set ups together with the possibility to easily vary the
intentional doping and the possible formation of junctions. The main disadvantage regards their inte-
gration on planar substrates for the exploitation of their useful properties, for example transfer and
contacting on transducers can be troublesome.
The bottom-up approach allows low cost fabrication although it could be very difficult to get them
well arranged and patterned [20].
Furthermore more control and insight into the growth process must be achieved for their fruitful
integration in functional devices.
The most promising approach to produce functional nanowires will be the combination of the two
preparation technologies.
This review article will be focused on the bottom-up techniques for the preparation of 1D single-
crystal nanostructures.
Numerous one-dimensional oxide nanostructures with useful properties, compositions, and mor-
phologies have recently been fabricated using bottom-up synthetic routes. Some of these structures
could not have been created easily or economically using top-down technologies.
A nomenclature for these peculiar structures has not been well established. In the literature a lot of
different names have been used, like whiskers, fibers, fibrils, nanotubules, nanocable, etc. The defini-
tion of these 1D nanostructures is not well established. A few classes of these new nanostructures with
potential as sensing devices are summarized schematically in Fig. 1. The geometrical shapes can be
tubes, cages, cylindrical wires, rods, nails, cables, belts, sheets and even more complex morphologies.
When developing 1D nanocrystals the most important requirements are dimensions and morphol-
ogy control, uniformity and crystalline properties. In order to obtain one-dimensional structures a
preferential growth direction with a faster growth rate must exists. Achieving 1D growth in systems
with a isotropic atomic bonding requires a break in the symmetry during the growth and not just stop-
ping the growth process at an early stage (0 and 2D).
In the past years the number of synthesis techniques has grown exponentially. We can divide these
growth mechanisms in different categories, first of all catalyst-free and catalyst assisted procedures
and then we can distinguish between vapor and solution phase growth. As far as metal oxides are con-
cerned the most used procedure is the vapor phase one. But solution phase growth techniques provide
a more flexible synthesis process with even lower production costs.
There are different growth mechanism depending on the presence of a catalyst, i.e. vapor–liquid–
solid (VLS), solution–liquid–solid (SLS) or vapor–solid (VS) process.

2.1. Vapor phase growth

The vapor phase approach was used in the early 60’ for the preparation of micrometer-size whis-
kers. These whiskers were prepared either by simple physical sublimation of the source material or
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 5

through reduction of a volatile metal halide. In the last years this method was used to prepare differ-
ent materials in form of nanowires. The growth was performed in tubular furnace studied to obtain
the proper temperature gradient. The source material once evaporated is transported by a gas carrier
towards the growth site where it nucleates. The nucleation can start from particles or catalyst, follow-
ing the VS, VLS mechanisms.

2.1.1. Vapor–liquid–solid mechanism


The controlled catalytic growth of whiskers, and more recently nanowires, was discovered by Wag-
ner and Ellis in 1964 [21], they found that Si whiskers could be grown by heating a Si substrate cov-
ered with Au particles in a mixture of SiCl4 and H2 and their diameters was determined by the size of
Au particles. Wagner and Ellis named the VLS mechanism for the three phases involved: the vapor-
phase precursor, the liquid catalyst droplet, and the solid crystalline product (Fig. 5).
VLS in the last decades was one of the most important methods for preparing 1D structures, it is
promising as a scalable, economical and controllable growth of different materials (oxide, semicon-
ductors,. . .). Understanding the growth dynamics is important to have a greater control in the nano-
wires shape, diameter and for a selective growth.
In general the presence of a metal particle, of size comparable to the nanowire, at its apex leads to
the conclusion that the growth mechanism followed the vapor–liquid–solid (VLS) process, but this
does not determine the phase of the catalyst during growth.
In most of the catalytic growths, nanowires have uniform diameters. The section can be rounded or
polygonal with atomically sharp lateral terminations. The growth process takes some dead time, a
starting period before the real growth begins, this was experimented also for vapor phase processes
[22].
The catalytic particles can be formed by vapor phase and/or surface diffusion transport or be depos-
ited from the evaporation of a colloidal solution or by deposition of a thin film onto the substrate. If
the metal does not wet the substrate, it will form clusters as the result of Volmer–Weber growth [23]
or when the substrate is kept at the high temperatures required for the growth process, the onset of
Ostwald ripening [24] will lead to a distribution of cluster sizes. In some cases the catalyst clusters
that initiate the NWs growth can also be formed at the initial deposition step; for example when car-
bothermal reduction is used to generate a volatile metal that is transported from a carrier gas and then
condense on the substrate.
Sometimes the catalyst may undergo other processes before becoming active for the growth of
nanowires after its formation or deposition. A mixture of the growth compound and the metal might
be more active for the NW formation than the pure metal catalyst, and may be required to form an
alloy, a true eutectic or some solid/liquid solution. In this case, saturation of the catalytic particle with
the growth material or the formation of the proper composition may explain the dead time period be-
fore growth. The incorporation of a significant amount of growth material into the catalytic particle is
expected to change the volume and, in turn, the diameter of the catalyst from its initial value with a
change in the NW section. Consequently Ostwald ripening and incorporation of growth material con-
tribute in changing the size of the catalytic particles.
A constant section NWs growth may correspond to a condensation on the catalyst surface and dif-
fusion and segregation at the interface between catalyst and nanowire. When the condensation and
incorporation is occurring only on the catalyst and not onto the NW sides, a constant catalyst section
results in a constant nanowire section.
The dimensions of the catalyst clusters can determine the NW section either by direct matching of
the size or by mechanism involving the catalyst curvature in which strain and lattice matching are
important. The NW section will decrease and eventually the growth process will end if the catalyst
is consumed or evaporates during the growth, or when the material is no longer supplied, or if the
temperature is reduced below a critical value necessary for the growth process. Temperature is a
key factor in determining processes such as dissociative adsorption, surface diffusion, bulk diffusion
through the catalyst, solubility and thermodynamic stability of certain phases.
The catalyst cluster can offer a higher sticking coefficient, but the difference in sticking coefficients
alone cannot account for the NWs growth process. Further considerations must be performed to ex-
plain the preferential incorporation at the interface between nanowire and catalyst. For example
6 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

the catalytic particle can lower the energy barrier for the incorporation of new material at the growth
interface compared to the one needed for nucleation of an island on a sidewall or on the substrate.
Adsorption occurs from the fluid (gaseous, liquid or supercritical) phase, it can be molecular or dis-
sociative and may occur on nanowire, catalyst, or substrate. The catalyst can activate the growth with
a sticking coefficient higher on its surface and vanishing elsewhere. After the adsorption there is the
adatoms diffusion onto or into the catalyst, across the substrate, or on the NWs lateral sides. In order
to have the unidirectional growth, the last two processes must be rapid and avoid secondary
nucleation.
The nanowires can grow from the top or the bottom of the catalyst cluster and as reported in Fig. 2,
a catalyst cluster can give rise to single or multiple nanowires growth. The catalyst can be found at the
bottom or top of the nanowire. In single NW growth there is a one-to-one correspondence between
catalyst and nanowires. In single wire growth control over the nanowire diameter should be obtained
controlling the catalyst radius. While in multiple nanowires growth the section must be related to
other factors such as the curvature of the growth interface and lattice matching between the catalytic
particle and the nanowire.
Regardless of the phase of the catalyst, the major requirement is the mobility of the growth mate-
rial that can allow reaching the growth interface with a low probability of nucleation in sites other
than the nanowire–catalyst interface.
The growth activation energy can be related to activated adsorption or with surface or bulk diffu-
sion. The essential role of the catalyst appears to be lowering the activation energy of nucleation at the
interface. There is a substantial barrier associated with the formation of the critical nucleation cluster
at a random position on the substrate or nanowire according to classical nucleation theory. If the cat-
alyst can lower the nucleation barrier at the particle/nanowire interface, then growth may only occur

Fig. 2. The processes that occur during catalytic growth. (a) In root growth, the particle stays at the bottom of the nanowire. (b)
In float growth, the particle remains at the top of the nanowire. (c) In multiple prong growth, more than one nanowire grows
from one particle and the nanowires must necessarily have a smaller radius than the particle. (d) In single-prong growth, one
nanowire corresponds to one particle. One of the surest signs of this mode is that the particle and nanowire have very similar
radii. Reprinted from Ref. [22]. License number 1905961408030.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 7

on the catalyst. The most important role of the catalyst particle is to ensure that the material is pref-
erentially incorporated at the growth interface.
Understanding the dynamics of VLS nanowires growth is essential in order to relate the properties
of the wire to their processing conditions. A theory for VLS growth has been presented in reference
[25] incorporating the surface energy of the solid–liquid, liquid–vapor, and solid–vapor interfaces.
The catalyst concentration profile in the droplet, the degree of supersaturation, and the modification
to the shape of the solid–liquid interface were predicted as functions of the material properties and
process parameters. The calculated growth rate found has the same dependence on diameter as the
flux of growth material at the liquid–vapor interface; thus, for radius independent flux, growth rate
results also radius independent.
It is often found that, the growth rate should decrease with decreasing diameter [26,27].
The effects of size on the growth kinetics of nanowires by the vapor–liquid–solid mechanism were
addressed from the theoretical point of view in [27]. The dependences of the growth rate and the acti-
vation energy of crystallization on size were given quantitatively. The obtained theoretical results
showed that the smaller the nanowire radius, the slower the growth rate, and the activation energy
of crystallization increases with decreasing radius of the nanowire. These theoretical predictions are
in agreement with the experimental cases. However, this conclusion depends on the growth condi-
tions [28] since the extent of supersaturation within the catalyst depends on the temperature and
gas-phase composition. Transitions from smaller diameters having lower growth rates to smaller
diameter having higher growth rates can occur as temperature and gas-phase composition are
changed.
Although it is commonly believed that in the VLS process, the size of the catalyst particles deter-
mines the NWs width, this is not true for all deposition conditions. Experimental studies on ZnO
NWs growth on Al0.5Ga0.5N substrate confirm that this rule only applies when the catalyst particles
are reasonably small (<40 nm) [29].
A linear relationship between the density of the nanowires and the thickness of the catalyst layer
was found, therefore catalyst thickness control could be a very simple and effective way to achieve
density control of aligned nanowires over a large surface area. To reveal why the density varies, but
the width remains constant, the wetting behavior of a gold layer on the substrate was investigated
when heated to the growth temperature.
The results classified the growth processes into three categories: separated dots initiated growth,
continuous layer initiated growth, and scattered particle initiated growth. Because of the wetting sit-
uation between the melted catalyst droplet and the substrate, more energy favorable sites were cre-
ated for nanowire growth with thinner catalyst layers. Moreover, when the catalyst layer was
sufficiently thick, a continuous ZnO network would be deposited simultaneously at the bottom of
the nanowires (Fig. 3).
Another important process controlling the cluster dimensions, that is in general forgotten, is the
thermodynamic limit for the minimum radius of the metal liquid clusters at high temperature
rmin ¼ 2r LV V L =RT ln s
where rLV is the liquid vapor surface free energy, VL is the molar volume of liquid, and s is the vapor
phase supersaturation.
Furthermore as well as the equilibrium vapor pressure of a solid surface also the solubility depends
on the surface curvature. As the size are reduced the solubility increases, as a result higher supersat-
uration in the vapor phase has to be created. Higher supersaturation may promote lateral growth on
the NWs side or homogeneous nucleation in the gas phase.
A procedure for controlling the radial and axial dimensions of SnO2 NWs has been presented in
[30,31] by combining VLS approach with molecule-based chemical vapor deposition. The synthesis
was based on the decomposition of discrete molecular species, which allows growing nanowires at
low temperatures with a precise control over their diameter and length. The precursor chemistry
was chosen to facilitate the stripping of organic ligands and to achieve complete decomposition that
is critical for maintaining the gas phase supersaturation necessary for 1D growth. Axial and radial
dimensions of the NWs were varied by adjusting the precursor feedstock, deposition temperature,
and catalyst size.
8 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ó American Chemical Society 2006

Fig. 3. (a) Variation of density (left-hand vertical axis) and width (right-hand vertical axis) of the aligned ZnO nanowires with
the thickness of gold catalyst layer. Inset: Top-view SEM image of the aligned ZnO nanowires used for density calculation, the
scale bar represents 200 nm. (b, c) TEM images of ZnO nanowires catalyzed by 1 and 8 nm gold layers, respectively. Inset:
Selected area electron diffraction pattern recorded from a nanowire indicated by the circle in image c. catalyzed by the
corresponding gold layers. Reprinted with permission from [29].

Despite the success of all these growth procedures, there have been just few comparative studies on
catalysts and substrates influence. Such studies are valuable because both the catalyst and substrate play
important roles in NWs structure and properties. In reference [32], a case study of ZnO nanowire growth
was performed; four different catalysts and substrates of different materials, structure, and crystal ori-
entation were investigated. It was found that the growth depends on the choice of surface catalysts, e.g.
for the Fe catalysts, the growth of ZnO nanowires may occur via a vapor–solid process, while, for the case
of Au, Ag, and Ni catalysts, the vapor–liquid–solid process usually dominates the wire growth. Further-
more differences in growth were also closely related to the differences in materials properties of these
wires, including the degree of nanowire alignment on substrates and the atomic composition ratio of Zn/
O, as well as the relative intensity of the oxygen vacancy-related emission in PL spectra.
The use of different catalysts provides the versatility of growth for one-dimensional ZnO nano-
structures with different ranges of parameters such as diameters, areal densities, and aspect ratios.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 9

Ó American Physical Society 2006


Fig. 4. (a) Representative bright field TEM images of a Si wire acquired at four successive times during deposition. White arrows
highlight a reference point on the wire sidewall. (b) Length L (open squares) and diameter d (solid circles) of the same wire as a
function of t. The straight line is a least-squares fit to the first 1200 s. Reprinted Fig. 2 with permission from [33]. http://
link.aps.org/abstract/PRL/v96/e096105.

This works suggested that, compared to noble-metal catalysts, growth using transition metal catalysts
occurs at a relatively faster rate and therefore typically yields thicker wires with higher aspect ratio.
However, the NWs have more oxygen vacancies affecting other properties, such as electrical transport
and surface chemistry.
Few in situ studies were performed especially regarding the Si NWs growth. Si nanowires growth
by the vapor–liquid–solid mechanism was monitored using real time in situ ultra high vacuum trans-
mission electron microscopy [33].
A growth rate independent of wire diameter was found. Showing that the irreversible, kinetically
limited, dissociative adsorption of disilane directly on the catalyst surface was the unique rate-limit-
ing step (Fig. 4).
The growth rates were independent of wire diameter, and increased linearly with pressure. From
the growth rate measurements, the reactive sticking probabilities of Si2H6 at the droplet surface
and at the wire sidewall was determined. A novel dependence of growth rate on wire taper, which
was attributed to the deposition of excess Si from the shrinking droplets, was observed.
Many open questions still remain regarding the different experimental evidences on VLS growth,
top or bottom catalyst cluster, single or multiple nanowires growth, the relation between the catalyst
and nanowire size, the possibility of an auto-catalyzed growth. But, regardless of the catalyst phase
(either liquid or solid) the growth dynamics does not change and catalytic growth still appears to
be the most powerful method for producing 1D nanostructures.

2.1.2. Vapor–solid mechanism


The VS growth takes place when the nanowire crystallization originates from the direct condensa-
tion from the vapor phase without the use of a catalyser. At the beginnings the growth was attributed
to the presence of lattice defects, but when defects-free nanowires were observed this explanation
10 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

cannot be any longer accepted. Another peculiar effect registered was a nanowire growth rate higher
than the calculated condensation rate from the vapor phase. A possible interpretation is that all the
faces of the nanowire adsorb the molecules that afterwards diffuse on the principal growth surface
of the wire.
VS process occurs in many catalyst-free growth processes. Quite a few experimental and theoret-
ical works have proposed that the minimization of surface free energy primarily governs the VS pro-
cess. Under high temperature condition, source materials are vaporized and then directly condensed
on the substrate placed in the low temperature region. Once the condensation process happens, the
initially condensed molecules form seed crystals serving as the nucleation sites. As a result, they facil-
itate directional growth to minimize the surface energy.
This self-catalytic growth associated with many thermodynamic parameters is a rather compli-
cated process that needs quantitative modelling [34]. It was reported for indium oxide, In2O3 wires
were synthesized through thermal evaporation of pure In2O3 powders and the effect of substrate seed-
ing was studied for controlling density distribution and lateral dimensions of the wires. The wires ex-
hibit uniform section, atomically sharp lateral facets, and pyramidal termination, typical of a VS
growth mechanism assisted by oxidized nanocrystalline seeds.
Other growth conditions have been reported, for example ZnO NWs were synthesized by a VP pro-
cess using a thin film (10 nm) of tin as catalyst. Carbothermal reduction was used to reduce the source
temperature needed for the vapor phase production. The tip of the NWs resulted without the catalyst
and was attributed to VS process [35].
Many report also the NWs production by simple oxidation of the metal composing the metal oxide
[36], for example [37] report the growth of CuO NWs from copper foils oxidized in wet air at temper-
atures between 300 and 800 °C. Within the temperature range of 400–700 °C, the nanowires formed
have two different morphologies, curved and straight, with diameters between 50 and 400 nm and
lengths between 1 and 15 lm. The growth behavior was explained in terms of kinetics involving
short-circuit diffusion, the strength of the nanowires, and the thickness ratio of the oxide scale and
the metal. The formation kinetic of CuO nanowires was governed by the short-circuit diffusion of
atoms or ions during the reaction. The deformation of thin oxide scale under thermal stresses may also
contribute to the formation of curved nanowires.
The vapor–solid VS growth was attributed also for a two-step high-temperature, catalyst-free,
physical evaporation of tungsten oxide NWs [38]. The procedure consisted in heating tungsten pow-
der, during the heating an oxide layer might be formed on the metal surface; the oxide then can evap-
orate and redeposit on the substrate surface, forming one dimensional nanostructures. Alternatively,
the tungsten metal was vaporized first; a subsequent oxidation during the deposition on the substrate
may also form the nanostructure.

2.2. Solution phase growth

Growth of nanowires, nanorods and nanoneedles in solution phase has been successfully achieved.
These growth methods usually require ambient temperature so that complexity and cost of fabrication
are considerably reduced. To develop strategies that can guide and confine the growth direction to
form Q1D nanostructures, researchers have used a number of approaches that may be grouped into
template-assisted and template-free methods.
The solution-based catalyzed-growth mechanism is similar to the previously described VLS
mechanism, in this case a nanometer-scale metallic droplet catalyze the precursors decomposition
and crystalline nanowire growth. The variants of VLS growth in solutions SLS and supercritical
fluid–liquid–solid (SFLS) growths provide nanowire solubility, control over surface ligation, and
smaller diameters. But the VLS growth in general produce nanowires of the best crystalline
quality.
There exist strong indications that the catalyst droplets in the SLS, as well as in VLS, mechanisms
play a catalytic role in precursor decomposition, in addition to catalyze the NWs growth (Fig. 6). The
early VLS literature claimed such a role on the basis of various experimental observations [21,39,40],
including that VLS crystal growth typically occurs at temperatures several hundreds of degrees lower
than epitaxial film growth from the same precursors.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 11

Fig. 5. Schematic representation of vapor–liquid–solid growth of nanowires. The catalyst is in the liquid phase and precursors
can adsorb and condense to form the nanowire.

Fig. 6. Schematic representation of the solution–liquid–solid growth of nanowires. Precursors in the liquid phase react to form
the nanowire.

Thus, the droplets perform a dual role as ideally rough surfaces for precursor adsorption and
decomposition and as a crystallization solvent supporting semiconductor crystal-lattice formation
and, hence, wire growth.
As well as for VLS, melting points, solvating abilities, and reactivities are the important criteria for
selecting the potential SLS catalyst materials. Moreover, at least one of the components of the product
semiconductor phase must have finite, but limited solubility in the catalyst material, so that high
supersaturations can be achieved. Finally, the catalyst should not react with or form a solid solution
with the target semiconductor phase (unless the catalyst material is the same as one of the constituent
elements of the semiconductor).

2.2.1. Template-assisted synthesis


Anodization growth technique is a well-established process to growth of oxide coatings and pro-
tecting layers on metal surfaces since the past century. Only in the last decades it was used for the
preparation of porous films that later were used also for the production of nanowires and nanotubes
[41–43].
The most used porous material is alumina and in general the starting metallic material is alumin-
ium [44].
There are several references reporting on this template-synthesis strategy for nanofabrication such
as Hulteen and Martin [45] considered as one of the pioneer groups in this subject, particularly in
functional nanowire arrays fabrication.
12 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

The same experimental procedure is recently being used for the preparation of nanowires and
nanotubes of other materials such as TiO2 [46].
The advantages of anodization and electroplating processes for 1D nanostructures production are
low costs, repeatability and potential compatibility with silicon technologies which make these nano-
structure synthesis procedures interesting, one of the main disadvantages is the poor crystallinity of
the produced NWs. Control in nanowire dimensions and the morphology of the ordered arrays can be
achieved. Because the diameter of these nano-channels and the inter-channel distance are easily con-
trolled by the anodization voltage, it provides a convenient way to manipulate the aspect ratio and the
area density of Q1D nanostructures.
The use of two anodization processes to obtain hexagonal close packed highly ordered nanoporous
alumina membranes was reported by Masuda and Fukuda’s [47]. After this report different procedures
to obtain porous templates have been presented for the production of 1D nanostructures ordered
arrays.
With the use of a periodic structured template, such as anodic aluminium oxide (AAO), molecular
sieves, and polymer membranes, 1D nanostructures can be prepared thanks to the confinement effect
or the porous template.
The arrays grow with ordered hexagonal cells with central pores parallel to each other and with a
symmetry axis perpendicularly oriented to the substrate surfaces. The most used procedure to obtain
functionalized AAO is electrochemical growth technique [48–51]. In 2003 great advances were
achieved by using a combination of nano-imprint and lithographic techniques with a subsequent
anodization process of aluminium metal and other metallic or semiconductor substrates [52,53].
Q1D nanostructures can be deposited into the pores using electrodeposition or sol–gel deposition
methods.
In the latter case, as the first step, colloidal (sol) suspension of the desired particles is prepared
from the solution of precursor molecules, afterwards an AAO template is immersed into the sol sus-
pension, so that the sol can aggregate on the AAO template surface. Deposition time must be carefully
chosen in order to allow the sol particles to fill the template pores and form 1D nanostructures. The
final step consists in a thermal treatment to remove the gel.
Finally also the 1D nanostructures can be used as templates for the physical confinement of the
growth, for example the synthesis of LaCoO3 nanowires using carbon nanotubes (CNT) as template
has been reported [54]. The precursor solution for the nanowires was obtained by dissolving La(NO3)3
and Co(NO3)2 in water. CNT were dispersed in the precursor solution by sonication and stirring. Then
the solution was centrifugated, dried and finally calcinated to remove the CNT template. This synthesis
is a combination of solution and templates methods, but without the use of electrochemical deposition.
Despite of its simplicity template based growth is characterized by the production of polycrystal-
line nanowires that can limit their potential for both fundamental studies and applications.

2.2.2. Template-free methods


A big research effort has been reserved to the template-free methods for the deposition of 1D nano-
structures in liquid environment; the most important procedures are surfactant assisted, sonochem-
ical, hydrothermal, organometallic methods and electrospinning.
The use of a surfactant can promote the anisotropic crystal growth required for the production of
1D crystals. The anisotropic growth is in general performed in a three phases system, oil, surfactant
and aqueous phase. These surfactants confine the crystal growth as in microreactors. Key points are
the selection of precursor and surfactants, and parameters such as temperature, pH value, and reac-
tants concentration.
Different surfactants were proposed depending on the material that is addressed, for example oleic
acid, hexylphosphonic acid (HPA), tetradecylphosphonic acid (TDPA), trioctylphosphine oxide (TOPO),
and trioctylphosphine (TOP). Surfactant-assisted methods is a trial-and-error procedure and requires
much effort to select the appropriate capping agents and reaction environment [55–57].
In the sonochemical method, instead ultrasonic wave are exploited to modify the crystal growth
changing the reaction conditions, by acoustic stirring.
During sonication bubbles are formed in the aqueous solution, they grow and then collapse. In such
environments it’s easy to reach extreme reaction conditions (temperature greater than 5000 K, pres-
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 13

sure larger than 500 atm, and cooling rate higher than 1010 K/s) that can allow the formation of 1D
nanostructures [58,59].
As an example, magnetite (Fe3O4) nanorods were synthesized by ultrasonically irradiating aqueous
iron acetate in the presence of beta-cyclodextrin, which serves as a size-stabilizer [60].
A well-known procedure for material synthesis is the hydrothermal process, it has been carried out
to produce crystalline structures since the 1970s. This process begins with aqueous mixture of soluble
metal salt (metal and/or metal-organic) of the precursor materials. The mixed solution is placed in an
autoclave under elevated temperature and relatively high pressure conditions necessary to force the
1D nanostructures formation.
Typically, the temperature ranges between 100 °C and 300 °C and the pressure exceeds 1 atm.
Many work have been reported to synthesize ZnO nanorods by using wet-chemical hydrothermal ap-
proaches [61–64] but other oxides were prepared such as CuO [65], ceria [66] and titania [67].
The preparation of crystalline manganese oxide nanorods is reported in [68], consisting in an easy
one-pot hydrothermal treatment of commercial granular/bulky MnO2 in ammonia solution. Post-cal-
cination treatment of the hydrothermal products was not necessary and no organic solvent was
needed in the process.
The synthesis of Cu2O nanowires was achieved by the reduction of cupric acetate with o-anisidine,
pyrrole, or 2,5-dimethoxyaniline under hydrothermal conditions [69]. The diameter and morphology
of Cu2O nanowires can be easily tuned by the choice of reductant type and synthetic temperature.
It has just been published in [70] the growth of ZnO nanowires aligned to Si and C coated substrates
with the appropriate choice of process conditions without the use of a ZnO thin film.
Another interesting approach for the formation of nanorods is the organometallic method presented
in [71,72]. It has been used for the synthesis, at room temperature, of homogeneous ZnO nanostruc-
tures of isotropic or rod shape. Organometallic complexes have been used as precursors to overcome
the problem of interaction of ionic species with the particles growth process [72]. The synthesis of me-
tal nanoparticles takes advantage of the reactivity of metal-organic precursors to CO, H2, UV irradia-
tion or heat treatment. The method uses both the exothermic reaction of the organometallic precursor
with water to produce crystalline zinc oxide and a kinetic control of the decomposition by long-alkyl-
chain amine ligands in the presence or absence of additional solvents to control size and morphology.
Quantitative yields of ZnO nanostructures is reported. The mechanism of particle growth involves
mass transport of zinc atoms, the amine ligand playing a fundamental role in the process and remain-
ing coordinated to the particles throughout the synthesis. The metal oxide products were dissolved in
common organic solvents forming clean and clear luminescent solutions that could easily be depos-
ited on various surfaces as monolayers or thick layers.
Finally another template-free synthesis process is electrospinning. Electrospinning uses an electrical
charge to form a mat of fine fibers. It may be considered as an electrospray process. A high voltage
induces the formation of a liquid jet. In electrospinning, a solid fiber is generated as the electrified
jet is continuously stretched due to the electrostatic repulsions between the surface charges and
the evaporation of solvent. Whipping due to a bending instability in the electrified jet and concomi-
tant evaporation of solvent (and, in some cases reaction of the materials in the jet with the environ-
ment) allow this jet to be stretched to nanometer-scale diameters. The elongation by bending
instability results in the fabrication of uniform fibers with nanometer-scale diameters.
The first patent that described the operation of electrospinning appeared in 1934, when Formalas
disclosed an apparatus for producing polymer filaments by taking advantage of the electrostatic repul-
sions between surface charges [73].
A number of oxides that include Al2O3, CuO, NiO, TiO2, SiO2, V2O5, ZnO, Co3O4, Nb2O5, MoO3, and
MgTiO3 have been fabricated as fibrous structures [74–83].

3. Vertical and horizontal alignment techniques

The assembly of nanowires into ordered arrays is critical to the realization of integrated electronic
architectures. Methodologies for development of large-scale hierarchical organization of nanowire ar-
rays have been recently developed [84].
14 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

One of the possible strategies to avoid complex and time expensive manipulation of single wires for
suitable alignment is the epitaxial growth of single-crystalline nanowires onto lattice-matched sub-
strates, where the most of the nanowires grow along a precise direction with respect to the substrate
plane. Such techniques can allow direct integration of single nanowires and nanowire arrays into
effective devices. The epitaxial growth on lattice-matched substrates plays a key role in the ordered
assembling of nanowires. Frequently used substrates are crystalline silicon [85,86] and sapphire
[87] as well as yttrium stabilized zirconia (YSZ) [88] and GaN [89,90]. Typically, catalytic seeds of no-
ble metals (Pt and Au) are dispersed either randomly or in ordered arrays onto the substrate, leading
to nanowires growth according to the VLS mechanism. The regular alignment of nanowires being ob-
tained via VLS growth assisted by noble-metal catalyst [91] or through the deposition of a buffer layer
preliminary to the growth of the nanowires [92].
It has been recently reported the possibility of industrial scale patterning using nanolithographic
techniques [93]. This indication suggests the effective possibility of deserving patterned substrates
for large-scale applications, and foresees industrial use of nanowires.
The experimental techniques to measure the relative orientation between the grown nanostruc-
tures and the substrate basically rely on the diffraction of X-rays [94] or electrons [95], which allow
respectively high sampling capability over large sample areas or investigating the crystalline arrange-
ment at a small scale [96].
The direct integration of In2O3 nanowires grown using a bottom-up approach has been applied by
Meyyappan and co-workers [87] to obtain a vertical field-effect transistor (Fig. 7). Optical sapphire (a-
Al2O3) has been used for the growth of the nanowires, due to the good lattice matching with the [0 0 1]
axis of cubic In2O3 (lattice constant a = 10.1 Å, JCPDS 89-4595).
With respect to the traditional VLS mechanism, in which single crystalline nanowire grows from
the catalytically active seed, dynamic simultaneous nucleation and epitaxial growth events can occur
due to the presence of single-crystal substrate, driven by competitive growth mechanisms: namely
nucleation, 2D buffer layer formation over the substrate, nanowire elongation and nanothread
formation.
Epitaxial growth and characterization of perfectly aligned and three-dimensionally branched ITO
nanowire arrays with a controlled crystallographic growth direction has been exploited [97]. Yt-
trium-stabilized zirconia (YSZ) has been used as the substrate because YSZ, with a fluorite crystal

Ó American Chemical Society 2004

Fig. 7. FE-SEM micrographs of vertical In2O3 nanowire arrays on single crystalline optical a-sapphire substrates. (a) A 45°
perspective view of an array of nanowires on a-sapphire. The inset shows a schematic (not to scale) of a typical nanowire with
an aspect ratio a/b  4. (b) A zoom-in 45° perspective view of the nanowires showing orthogonal directionality, nanosized Au
catalytic heads, and nanothread cladding along the square core of the nanowires. (c) A 5° pperspective view of the array,
revealing uniform growth over the surface of the substrate and the rectangular footprint of the pyramidal base. The inset shows
a zoom-in top view of a square columnar nanowire with its pyramidal base. Scale bars: 2 lm, 0.25 lm, and 1 lm for A, B, and C
respectively; 500 nm for the inset of C. Reprinted with permission from Ref. [87].
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 15

structure (cubic, lattice constant a = 0.514 nm), is ideal for heteroepitaxial growth of ITO films with a
bixbyite crystal structure (cubic, lattice constant a = 1.01 nm). The branched ITO arrays were obtained
using sequential Au catalyst seeding and VLS steps. In this case, no ITO buffer layer was observed,
which indicates that the vertically aligned ITO nanowire array was epitaxially grown directly on the
YSZ substrate.
The formation of the buffer layer critically depends on the growth conditions. In Ref. [87] the con-
densation occurs at 1000 °C (argon flux 20–50 sccm), and the formation of a thin film epitaxially
grown over the substrate is key step for nanowires growth. In Ref. [97], instead, condensation occurs
at 800–900 °C (nitrogen flux 150 mL min1).
Detailed analysis of the role of the buffer layer has been carried out in the process of nucleation and
growth of square-shaped SnO2 nanotubules arrays over quartz substrate [98]. Each tubule is single-
crystal of rutile structure with four {1 1 0} peripheral surfaces structure and h0 0 1i growth direction.
The cross sectional dimensions of the tubes increased exponentially with the temperature at which
they were grown, with an activation energy for tube growth of about 0.44 eV. The tubules were found
to grow according to a self-catalyzed, direct VS process, where most new material is incorporated into
the bottom parts of each tubule through surface diffusion. The simple control of the temperature
determines the lateral dimensions of the condensation products. Initially, grains of the condensed bulk
layer are randomly oriented. As condensation continues, grains with energetically favorable crystallo-
graphic planes will preferentially grow larger, while grains with energetically unfavorable surfaces
gradually shrink.
Many studies pointed out the attention on the heteroepitaxial growth of ZnO nanostructures of
various shapes over different substrates.
Aligned growth of ZnO nanorods has been successfully achieved via a VLS process, with the use of
gold [99,100] and tin [101] as catalysts.
A combined technique based on epitaxial growth and substrate template has been investigated.
The self-assembly-based mask technique and the surface epitaxial approach were combined to grow
large-area hexagonal arrays of aligned ZnO nanorods [102]. The synthesis used a gold catalyst tem-
plate produced by a self-assembled monolayer (SAM) of submicron polystyrene spheres (Fig. 8) and
guided VLS growth on a single-crystal sapphire ð2 1 1
 0Þ substrate (Fig. 9). The nanorod grows along
[0 0 0 1] direction and its side surfaces are defined by f2 1 1
 0g. The collective optical properties of
the aligned ZnO nanorods were investigated using PL. It was shown that luminescence was emitted
mainly along the axis of the ZnO nanorods, indicating that collective properties can be tailored under
proper orientation of nanowire arrays.
Aligned ZnO nanowire/nanorod arrays following a predesigned pattern and feature with controlled
site, shape, distribution, and orientation were also obtained by Wang and co-workers [103]. The tech-
Ó American Chemical Society 2004

Fig. 8. Gold pattern produced by an evaporation technique using the shadow provided by the monolayer of self-assembled
polystyrene spheres. Reprinted in part with permission from Ref. [102].
16 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ó American Chemical Society 2004


Fig. 9. (a) Low-magnification top-view SEM image of aligned ZnO nanorods grown onto a honeycomb catalyst pattern. (b) Side
view of the aligned ZnO nanorods at an angle of 30°. (c, d) Top and a 30° view of aligned ZnO nanorods, where the hexagonal
pattern is apparent. (d) Aligned ZnO nanorods at the edge of the growth pattern. Reprinted with permission from Ref.
[102].

nique relies on the integration of atomic force microscopy (AFM) nanomachining with catalytically
activated VLS growth. Single-crystal sapphire (1 1 2  0) substrate was chosen to grow ZnO nanorods.
Si(1 0 0) substrate was also proven effective for aligned growth of ZnO quasi 1D nanostructures
[104]. Well-aligned ZnO nanorod and nanobolt arrays were synthesized on p-type Si(1 0 0) substrates
by a simple physical vapor deposition method (Fig. 10). The effect of the atmosphere on the shape of
the condensation products was investigated: the switch from Ar to air during the growth causes the
change from nanorods into nanobolts completely. The element ratio of Zn to O in the vapor phase dur-
ing the growth discriminates the formation of nanorods and nanobolts. Photoluminescence (PL) spec-
tra indicate a strong emission peak around 3.26 eV attributed to exciton-related emission and another
peak at 2.48 eV related to defects. In case of the nanobolts, the peak at 2.48 is weaker as compared to
the nanorods.
Aside to vertical alignment, horizontal alignment of nanowires via heteroepitaxial growth is being
explored. First attempts are reported for lateral orientated growth of In2O3 nanowire (NW) and nano-
rod (NR) arrays on (0 0 1) and (1 1 1) surfaces of Si substrates [105].The lateral self-aligned In2O3 nano-
wire and nanorod arrays on Si can offer some unique advantages for fabricating parallel nanodevices
that can be integrated directly with silicon technology.
Evidence of epitaxial growth have been recorded mainly in systems where metal catalysts have
been used for promoting the formation of nanowires. Indeed, the VLS mechanism and the formation
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 17

Ó American Chemical Society 2007


Fig. 10. (a) SEM image of the aligned ZnO nanorods, (b) magnified SEM image of the aligned ZnO nanorods, (c) SEM image of the
aligned ZnO nanobolts, (d) top view of the hexagonal structure of the nanotips, (e) single ZnO nanobolt, and (f) ZnO nanobolts
directly grown on a p-type Si(1 0 0) substrate. Reprinted with permission from Ref. [104].

of a nucleation center are expected to magnify the interaction between the crystalline substrate and
the nucleating nanowire [106]. Nagashima et al. report about MgO nanowires prepared through Au
catalyst [107], while ZnO on Si via Pt catalyst was described by [108]. Patterned and epitaxial struc-
tures were investigated by [103,109].
As far as catalyst-free preparation is concerned, Baxter et al. reports that the orientation of ZnO
nanowires prepared on sapphire is governed by epitaxial relations [110]. Fig. 11 highlights the parallel
orientation and dense arrangement of ZnO nanowires, which were produced in absence of metal cat-
alysts. It turned out that the epitaxy is favored when the substrate is oriented normal to the a-plane, as
the [0 0 0 1] direction of ZnO is matched by the ½1 1 2  0 direction of sapphire and vice versa, according
also to the findings of Chen et al. [111].
The epitaxial growth of ZnO nanowires on c-plane of sapphire was investigated by [112], as well as
the orientation of ZnO on m-plane oriented sapphire [113]. Fig. 12 shows the nanowires inclined in
opposite directions at about 30°. This arrangement was preserved over a wide pressure range for
the deposition, resulting in different aspect ratio for the nanowires.
The buffer layer is often homogenous to the composition of the nanowires, a notable example being
reported by Wan et al. [88,114], where vertically aligned tin-doped indium oxide (ITO) nanowires over
a ITO buffer layer were obtained by thermal evaporation on lattice-matched [100] YSZ substrate.
In some cases an heterogeneous buffer layer has been used to implement a functional properties in
addition to the promotion of epitaxial growth: In fact, TiN performed as a good electrode and diffusion
barrier material. Lin et al. [92] developed a ZnO (nanowires)/TiN (buffer layer)/Si (substrate), where

 1 0 k½1 1 1 k½0 1 1
the following lattice distance matching has been determined: ½1 2 ZnO TiN Si and
½0 0 0 1ZnO k½1 1 1TiN k½1 1 1Si . TiN features a lattice mismatch of about 8% with respect to ZnO and fa-
vored a Volmer–Weber growth mechanism.
18 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Fig. 11. SEM view of ZnO nanowires grown on a-plane sapphire. The nanowires grow perpendicular to the substrate and with
nearly perfect rotational alignment with respect to the substrate lattice. Scale bar corresponds to 100 nm. Reprinted from Ref.
[110], Copyright (2005), with permission from Elsevier.

Fig. 12. Tilted view (30°) SEM images of ZnO nanowire samples grown by high-pressure pulsed laser deposition with an oxygen
flow of (a) 0 sccm, (b) 20 sccm and (c) 35 sccm. Reprinted from Ref. [113], with permission from IOP Publishing.

Template alignment of quasi 1D nanostructures can be achieved by suitable application of tem-


plate-assisted growth, as described in Section 2.2.1.

3.1. Electric field alignment

In principle dielectrophoresis (DEP), which is the electrokinetic motion of dielectrically polarized


materials in non-uniform electric fields, could be a powerful tool for self alignment of nanowires into
a well defined space region, in which an external electric field is induced, without mechanical nanom-
anipulation techniques.
DEP has been shown to be capable of aligning metallic nanostructures, like carbon nanotubes and
gold nanowires directly between electrodes [115,116]. Effective manipulation of gold nanowires
[117], synthesized by template electrodeposition in porous aluminium oxide membranes, was ob-
tained. The dielectrophoretic force was modelled on the basis of the interaction between the alternat-
ing applied field and the induced dipole moment of gold nanowires, and allowed the electrical
characterization of the nanowires (see Fig. 13).
Very recently dielectrophoresis was applied also for manipulating semiconducting and oxide nano-
wires to obtain prototype devices.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 19

Fig. 13. Schematic diagram of nanowire assembly, and RC circuit representation. Reprinted from Ref. [117], with permission
from IOP Publishing.

Positive ac DEP has been used to align CdSe semiconductor nanowires near patterned microelec-
trodes Fig. 14 [118]. The induced dipole of the wires is proportional to their conductivity, due to their
large geometric aspect ratio.
AC dielectrophoretic manipulation was applied for fabrication of nanosensors based on tin oxide
nanobelts [119]. Positive and negative DEP was used for the assembly of a nanodevice, which con-
sisted of SnO2 nanobelts attached to castellated gold electrodes defined on a glass substrate, and cov-
ered by a microchannel (Fig. 15).
Controlled assembly of ZnO nanowires was carried out using DEP [120]. A structure similar to a
field-effect transistor with two isolated top electrodes comprising the source and drain and a lower
substrate electrode as the gate was used for the dielectrophoresis-based assembly of zinc oxide nano-
wires. The geometry of the electrodes as well as the magnitude and frequency of the applied electric

Ó American Institute of Physics 2007

Fig. 14. Dielectrophoretically aligned CdSe NWs using an ac electric field (10 V) with electrodes separated by a 20 lm gap. [(a)
and (b)] (1 MHz) Resulting bright field image after alignment and epifluorescence image taken at t = 50 s during the alignment,
respectively. [(c) and (d)] (10 kHz) Resulting alignment in 190 s, under illumination (100 W/cm2 at 488 nm) and in dark,
respectively. Reprinted with permission from Ref. [118].
20 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Fig. 15. SnO2 nanobelts bridging DEP electrode gaps during positive dielectrophoresis. Smaller nanobelts are sticking out of the
electrodes due to the action of the electric field. Reprinted from Ref. [119], with permission from Copyright 2005 Elsevier.

Ó American Institute of Physics 2007

Fig. 16. SEM images of assembled ZnO nanowires between electrodes with different gap distances. An ac voltage of 10Vpp at a
frequency of 20 MHz was applied to the right electrode relative to the left grounded electrode. (a) The left four fingers are all
part of the left electrode and the right four fingers are all part of the right electrode. The four gap distances are 2, 4, 6, and 10 lm
respectively and are enlarged in (b)–(e). The scale bars are 10 mm in (a) and 2 lm in (b)–(e). Reprinted with permission from
Ref. [120].

field was proven significantly affecting the density of the nanowires assembled between the elec-
trodes (Fig. 16).
Dielectrophoretic alignment was employed to design simple structures such as Schottky diodes
formed across Au electrodes using ZnO nanobelts and nanowires [121]. The formation of the Schottky
diodes is suggested due to the asymmetric contacts formed in the dielectrophoresis aligning process.
The detailed IV characteristics of the Schottky diodes have been investigated at low temperatures.
Dielectrophoretic integration of nanodevices with CMOS VLSI circuitry [122] was also achieved, en-
abling the creation of integrated circuits that include readout, signal processing and communication
circuitry. The nanostructures have been manipulated using dielectrophoretic forces, allowing their
individual assembly and characterization (see Fig. 17).
ZnO nanowire-based UV photosensor have been fabricated and characterized using DEP [123]. ZnO
nanowires, which were synthesized by nanoparticle-assisted pulsed laser deposition, were suspended
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 21

Ó IEEE 2006
Fig. 17. SEM photograph of assembly sites initially electrically isolated from the readout circuitry. Assembled nanowires are
clamped and connected to the readout circuit using a post-assembly e-beam lithography and metallization step. Reprinted from
Ref. [122].

in ethanol, and were trapped in the microelectrode gap where the electric field became higher. The
trapped ZnO nanowires were aligned along the electric field line and bridged the electrode gap.

3.2. Nanomanipulation

The manipulation of nanostructures has been pursued in the last years to probe individual nano-
wire-like structures and to fabricate prototypes of electronic micro-sized devices, when direct integra-
tion of the growth of nanowires into a functional substrate has not been achieved.
Various approaches for manipulation of metal oxide nanowires have been proposed, including
mechanical [124], electrostatic [125], and dielectrophoretic [121] methods. Alignment of single nano-
wires or assembling of several identical nanostructures is obtained under vacuum condition or in li-
quid or standard atmospheric environment, most of the manipulation being performed in association
with highly-resolved imaging techniques such as scanning microscopy (SPM and SEM) and even trans-
mission electron microscopy (TEM) [126,127]. The methodology for manipulation of metal oxide
nanostructures has largely benefited from the previous efforts expended for the characterization of
carbon nanotubes and nanostructures [128,129].
Ó American Chemical Society 2003

Fig. 18. Bending of a helical silica nanospring through AFM manipulation. (a) and (b) before and after manipulation. Reprinted
with permission from Ref. [132].
22 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Manipulation has been proved as a powerful technique tool for studying the intrinsic transport
properties of nanowires, and to provide an experimental proof for innovative functional concepts
[130].
The electrical characterization of metal oxide nanowires has been performed via two-contacts or
four-contacts methods, where the conductive paths are provided by either patterned structures
[112], piezo-actuated probes, or metallic deposits obtained by focused ion beam [131]. These ap-
proaches allow the measurement of physical quantities such as free-electron concentration, electrical
mobility, and conductivity in single tin oxide nanowires, accounting for the contribution of the metal-
lic contact, as reported in Ref. [131].
Mechanical manipulation of oxidic structures has been performed in air through AFM [132] for sil-
icon oxide helical structures (see Fig. 18) and has been also implemented in electron microscopes.
Recently, nanoindentation has been introduced to determine the hardness and elastic modulus of
zinc or tantalum oxide nanowires, and to determine the Young modulus of several nanostructures
[133–135].

4. Doping of quasi 1D metal oxide nanostructures

Doping of nanowires is pursued to the controlled modification of the characteristics of the nano-
wires, in terms of morphological features as well as electrical or optical properties. Differently from
heterogeneous systems formed by metallic catalytic particles and metal-oxide nanowires, the intro-
duction of dopants assumes preservation of the crystalline structure for the nanowire and avoids for-
mation of precipitates, segregation phenomena, or nucleation of second phases.
The presence of dopants may introduce a distorsion of the lattice and guide the growth of the nano-
wires to a specific crystallographic direction [136]. As an example, Fan et al. reported about the change
in morphology and crystalline habit in ZnO nanostructures in consequence of introduction of indium.
Fig. 19 summarizes the proposed model that suggest a change in the nucleation behavior of ZnO at the
solid–liquid interface upon supersaturation and oxidation of Zn.
The quantitative description of the correlation between the atomic structure and the properties of
the nanostructure makes the investigation of dopant dispersion a challenging task. Indeed, the small
dimension of the nanowires requires both spatial resolution and chemical sensitivity [137]; therefore
a reliable determination of dopant dispersion may be achieved through a complementary approach,
where the spatial distribution of dopants is associated to the measurement of the electrical activation
of ionized atoms. For these reasons, electron microscopy, scanning probe microscopy, X-ray and syn-
chrotron radiation diffraction techniques are associated to photoluminescence spectroscopy and elec-
trical transport measurements. Elemental analysis was also carried out by Rahm et al. using particle
induced X-ray emission, Rutherford backscattering spectroscopy, and Q-band electron spin resonance

Fig. 19. Schematic model of the growth processes for ZnO nanowire and nanobelt (dimensions not to scale). The main
difference is that the [0 0 0 1]-axial nanowire (a) grows upwards on the top (0 0 0 1) plane of a ZnO columnar/pyramidal nuclei,
whereas the h1 1 2 0i nanobelt (b) grows mainly from the side faces of a quasi-hexagonal indium doped ZnO pad. Reprinted from
Ref. [136] with permission from IOP Publishing.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 23

[138], while analysis of as light an element as Mg was determined by X-ray photoelectron spectros-
copy [139].
In metal oxide semiconductors, modification of the electrical properties can be achieved either by
introduction of foreign elements in the crystal structure or by variation of the oxygen stoichiometry
[140], resulting in large variation of the concentration of carriers, their mobility, and the electrical
resistivity.
Xiang et al. report about the synthesis of doped ZnO nanowires produced via chemical vapor depo-
sition (CVD) and characterized through photoluminescence (PL) and electrical measurements [141].
Their systematic approach highlighted the capability to achieve p-type conduction in phosphor-doped
nanowires (see Fig. 20).
ZnO is usually a n-type semiconducting material, and the p-type behavior of ZnO nanowires is a
remarkable achievement. The inherent difficulty in producing stable p-type behavior and the effect
of P on the photoemission of ZnO has been discussed by Shan et al. [142].
The papers of Xiang et al., and Shan et al. [141,142] also introduce the issue of the doping method-
ology. The concentration of oxygen vacancies can be controlled through variation of the oxygen con-
centration in the Ar/O2 gas carrier during the synthesis of nanowires, and similar results can be
obtained via post-synthesis treatment in reducing atmosphere [140].
Dopant addition is basically carried out by modifying the composition of the precursor in the evap-
oration–condensation process, despite the limited capability to manage the amount of dopant even-
tually introduced in the nanowires. The significant difference between the elemental ratio in the

Ó American Chemical Society 2007

Fig. 20. (left) SEM image of P-doped ZnO nanowires, showing that the nanowires have uniform diameter (about 55 nm). The
marker corresponds to 200 nm. (center). X-ray diffraction pattern of ZnO nanowires with no peaks associated to second phases
or clusters (right) Excitonic peaks of PL spectra at 10 K of n-type (green line), as-grown (red line), and annealed (blue line) P-
doped ZnO nanowires. Reprinted with permission from Ref. [141].
24 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ó American Chemical Society 2007


Fig. 21. Ids–Vds curves for back-gated Ta-doped SnO2 nanowire FET devices fabricated on silicon substrates and operated with
Vgs ranging between 6 and 4 V in 2 V steps from top to bottom. Reprinted with permission from Ref. [149].

precursor material and the composition obtained for the nanowires marks a critical issue of the evap-
oration–condensation approach, that is discussed by Nguyen et al. [143]. Liquid solution syntheses
and hydrothermal methods are promising for good reproducibility, controlled stoichiometry, and pre-
cise doping of the nanowires (as discussed in Section 2.2).
Post-synthesis treatments in controlled conditions of atmosphere and temperature allow an im-
proved dopant addition, as the temperature activated diffusion of dopants into the nanostructure
may be tailored.
The effort of ZnO doping with V-group elements such as As, P, and N has been documented
[144,145,139,146], and also doping with Tm, Yb, and Eu using ion implantation and post annealing
was reported in the literature [147].
Similarly to ZnO nanowires, SnO2 and In2O3 nanowires were doped with Sb, Ta and other elements,
which could favorably substitute the cation in the nanowire crystalline [148–150]. In addition, the
preparation of nanowires of the ternary system of In-stabilized tin oxide has been investigated [143].
Wan et al., together with Dattoli et al. highlight the effect of donor dopants on the electrical prop-
erties of tin oxide nanowires [148,149]: an increased mobility in excess of 100 cm2/V s was statisti-
cally determined from field-effect transistor (FET) devices as shown in Fig. 21, and a metallic
behavior associated to as low a electrical resistivity as 4  104 X cm have been recorded though elec-
trical transport measurements.

5. Preparation of quasi 1D metal oxide heterostructures

The creation of heterostructures is being strongly investigated in the last years in order to exploit
the functional properties arising from the junction of different materials and/or the effect of hierarchi-
cal organization of 1D nanostructures. In Fig. 22 the typical shapes of heterostructures fabricated up to
now are reported (dendritic growth, superlattice in a single nanowire, polycrystals coalescence on a
single backbone, core–shell geometry), each shape enabling exploitation of different functional prop-
erties of these innovative structures. Typically VLS and VS growth mechanisms can be combined for
enabling or inhibiting predefined growth directions and modifications of the crystalline assembly.
As an instance, combined VLS–VLS growth under sequential seeding of the catalyst leads to formation
of dendritic shapes; while VLS–VS process can lead to formation of core–shell structures, or decoration
of nanowires with small crystals.
III–V heterostructures have been created since 2002 by Yang and co-workers [151]. The laser abla-
tion process was applied as a programmable pulsed vapor source, allowing a block-by-block growth of
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 25

Fig. 22. Typical shapes for heterostructures fabricated using various methodologies of synthesis.

Ó American Chemical Society 2002

Fig. 23. (a) SEM image of the heterostructured nanowire array on Si (1 1 1) substrate. The scale bar is 1 lm. The inset shows the
tip of one nanowire. The scale bar is 100 nm. (b) STEM image of two nanowires in bright field mode. The scale bar is 500 nm. (c)
EDS spectrum of the Ge rich region on Si/SiGe superlattice nanowires. (d) Line profile of EDS signal from Si and Ge components
along the nanowire growth axis. The experiments were carried out on a Philip CM200 TEM operated at 200 keV. Reprinted from
Ref. [151].

the nanowire with a well-defined compositional profile along the wire axis via CVD process. Si/SiGe
superlattices in a single-crystalline nanowire have been obtained (Fig. 23).
Dendritic nanowires growth mediated by a self-assembled catalyst was also exploited for produc-
tion of InAs structures [152].
Typically, heterostructures of metal oxides are more difficult to be synthesized with respect to III–V
semiconductors, and in fact the block-by-block mechanism was never reported for oxides.
26 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ó American Institute of Physics 2006


Fig. 24. I–V characteristics of single ZnO vertical nanowires with a junction between intrinsically doped and Al-doped regions
(n–n+ junction) as probed by STM tip. Characteristics from a number of single nanowire junctions are presented in the figure
showing the degree of reproducibility. Broken line is a fit to the general empirical equation. Inset shows I–V’s from intrinsically
doped (pristine) single nanowires. Reprinted with permission from Ref. [153].

The first attempt for creation of heterostructures was based on spatially controlled doping of single
nanowires during NWs growth. Selective doping of single oxide ZnO nanowire was carried out for cre-
ation of nano-junctions by introducing the dopant in vertically grown single-crystalline ZnO nano-
wires. A section of the nanowires was doped with aluminium as donor during crystal nucleation,
resulting in n-n+ junction (Fig. 24) [153].
Heterostructures of one single material (ZnO) of different shapes have been obtained: vertically
aligned 2D and 1D ZnO nanostructures have been grown on electrically conducting, highly oriented
pyrolytic graphite (HOPG) and on insulating (1 1 2 0) sapphire substrates [154]. It was demonstrated
that the simple parameter of time duration of the growth process discriminates the formation of epi-
taxially oriented 1D structures at the junction of the nanowalls. As in most of the growth processes of
heterostructures, key role is played by the catalyst. In this case nanowire growth at the junction of
nanowalls is the result of the interplay between the dynamic wetting of gold and its thermally acti-
vated surface diffusion (see Fig. 25).
One further process for obtaining heterostructures starting from a single material is the sequential
oxidation of metal nanowires. Arrays of metal–metal oxide core–shell nanowires and single-crystal-
line metal oxide nanotubes have been obtained [155]. The process is based on the kinetic control of
the conversion of single-crystalline Bi nanowires to Bi–Bi2O3 core–shell nanowires via a multistep,
slow oxidation method, and then on the control of their further conversion to a single-crystalline
Bi2O3 nanotube array via fast oxidation.
Only very recently complex structures of different oxides have been obtained by properly combin-
ing VLS and VS condensation: core–shell, [156,157] longitudinal, [157] branched heterostructures,
[88,158] decorated nanowires [159]. Epitaxial relationship is typically found between the crystal-lat-
tice of the nanowire acting as the backbone in dendritic growth (or as the core in the core–shell geom-
etry) and the second material, which is applied in the second step of condensation.
Various mechanisms have been set-up for production of core–shell heterojunctions.
ZnO–Al2O3 and ZnO–TiO2 core–shell nanowires have been synthesized using a two-step process
(Fig. 26) [160]. First, ZnO nanowires were grown in aqueous solution using a seeded growth process.
Then atomic layer deposition (ALD) was applied to cover each nanowire with a thin layer of amor-
phous Al2O3 or TiO2. The TiO2 film was amorphous for thickness below 5 nm, while anatase polycrys-
tals were detected for thickness above 5 nm.
Similar technique was applied for creation of indium tin oxide (ITO) nanowires coated with TiO2
[161]. Catalyst-free transport-and-condensation method was applied for nanowires growth, using a
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 27

Fig. 25. Zinc oxide nanowalls and nanowires. (A) SEM image of quasi-3D ZnO nanostructures grown on a sapphire using 4 to
5 nm Au thin film as the catalyst. The inset shows a SEM perspective view. (B) 2D ZnO nanowalls on a sapphire with a height
5 lm. (C) An array of free-standing 1D ZnO nanowires on a HOPG substrate using 15 Å Au ultrathin film as the catalyst. (D)
Schematic illustration showing the growth mechanism of ZnO nanowalls and nanowires. Reprinted from Ref. [154], with
permission form AAAS.

Ó American Chemical Society 2006

Fig. 26. ZnO–Al2O3 core–shell nanowires. (a) Low-magnification image of a wire that has been cleaved in two. Scale bar, 50 nm.
(b) Electron diffraction pattern of the same wire. Only ZnO spots are present. (c) EDS elemental profile along the dashed line in
part a. (d) High-resolution image of the ZnO–Al2O3 interface. Scale bar, 5 nm. Reprinted from Ref. [160].
28 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

pulsed laser ablation for creation of the volatile species [162]. Nanocrystalline TiO2 thin film was then
deposited by RF sputtering to obtain the core–shell structure.
Homojunction has been obtained between b-Ga2O3 nanowires and nanocrystals via vapor phase
epitaxy (VPE): VLS and VS growth mechanism are simultaneously present, the first driving wire elon-
gation, the second being responsible for nanocrystal nucleation [163]. A similar mechanism, namely
the oriented attachment, was investigated for obtaining the formation of heterostructures in a
SnO2–TiO2 system [164].
A non-equilibrium synthesis technique has been developed to produce novel transition metal oxide
single-crystal nanowires, including YBa2Cu3O6.66, La0.67Ca0.33MnO3, PbZr0.58Ti0.42O3, and Fe3O4 [156].
Vertically aligned single-crystalline MgO nanowires, grown via VLS method, were applied as tem-
plates for epitaxial deposition of the desired transition metal oxides using pulsed laser deposition,
and led to core–shell nanowires (Fig. 27).
Radial and longitudinal nanosized In2O3–SnO2 heterostructures were produced by sequential
transport and condensation steps (Fig. 28). Radially shaped heterostructures were achieved via VS–
VS condensation, while VLS–VLS steps led to formation of longitudinal structures, based on the cata-
lytic activity of the gold cluster during the different condensation steps. Lattice matching between the
core In2O3 nanowire and the external SnO2 was demonstrated in the radial heterostructures, similar to
Ref. [156]. However, the lateral dimensions of the crystalline domains of the second phase are limited
by mechanisms of energy reduction, which lead to formation of polycrystals.
Sequential seeding of active catalyst was successfully applied for obtaining the growth of branched
nanowire structures, in which semiconducting In2O3 nanowire arrays with variable densities were
grown epitaxially on metallic ITO nanowire backbones (Fig. 29) [88].
Branched 1D heterostructures have been synthesized by growing vanadium oxide nanostructures
on SnO2 backbones. A sequential two-step strategy allowed formation of hierarchical structures, in

Ó American Chemical Society 2007

Fig. 27. MgO nanowire templates were also used for PZT and Fe3O4 core–shell nanowire synthesis. (a) TEM image of a PZT
nanowire. The catalyst can be seen at the end of the nanowire. Inset: HRTEM image of the MgO/PZT core–shell nanowire at the
sample tilt angle of 45 degrees. (b) SAED pattern of a MgO/PZT nanowire. Red arrows inside are PZT diffraction dots and blue
arrows outside are MgO diffraction dots. (c) TEM image of an Fe3O4 nanowire. (d) SAED pattern of a MgO/ Fe3O4 core–shell
nanowire. Red and blue arrows indicate Fe3O4 and MgO diffraction patterns, respectively. The weak and unmarked spots come
from Fe3O4. Reprinted from Ref. [156].
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 29

Ó American Chemical Society 2007


Fig. 28. Structural characterization of longitudinal heterostructures. (main picture) TEM panoramic view of the In2O3 nanowire,
with the SnO2 nanowire extending lengthwise. The black arrow marks the termination of the SnO2 nanowire and the catalytic
Au nanoparticle, which assisted the VLS growth of heterostructure. (1) SAED pattern showing the cubic single-crystal
arrangement for the indium oxide nanowire. (2) CBED pattern and high-resolution image from the tin oxide nanowire,
demonstrating its single-crystalline tetragonal (cassiterite) arrangement. (3,4) highly magnified TEM view and corresponding
SAED pattern of the heterojunction, where superimposition of both indium and tin oxides has been recorded. Reprinted from
Ref. [157].

Ó American Chemical Society 2006

Fig. 29. (a, b) Schematics of the growth processes of branched In2O3 nanowires, showing the deposition of Au catalysts on ITO
backbones (a) and the subsequent growth of In2O3 nanowire branches in a second VLS growth process (b). (c, d) SEM images of
branched In2O3 nanowires grown on ITO nanowire backbones. The thickness of the Au catalyst is 10 nm in (c) and 2 nm in (d).
Scale bars: 500 nm. Reprinted from Ref. [88].

which a heteroepitaxial growth of vanadium oxide on tin oxide was revealed. Shaping of nanocrystal-
line vanadium oxide was possible in form of buds or nanorods by varying the mass flow of the vana-
dium precursors.
Various ZnO nanostructures, such as nanobelts, nanorods, and nanowires, have been grown on pre-
synthesized SnO2 nanobelts via a simple thermal evaporation of Zn powders, without using any cat-
alysts, producing various heterostructures [165]. The evaporation temperature is the critical
30 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

experimental parameter for the formation of different morphologies of these nanostructures. Room-
temperature photoluminescence spectra of the heterostructures show that the relative intensity of
ultraviolet emission to the green band can be tuned by controlling the morphologies and sizes of the
secondary-grown 1D ZnO nanostructures, suggesting that the nano-heterostructures of these nano-
structures grown on SnO2 nanobelts may have potential applications in nano-optoelectronic devices.

6. Applications of metal oxide nanostructures

6.1. Metal oxide gas sensors

Metal oxides semiconductors are normally high gap metal oxides in which the semiconducting
behavior arises from deviation of stoichiometry [166]. They should always be regarded as compen-
sated semiconductor: cation vacancies are acceptors, yielding holes and negative charged vacancies,
shallow states made up of oxygen vacancies acts as n-type donors, since the bonding electrons on
the adjacent cation are easily removed and donated to the conduction band [167].
The termination of the periodic structure of a semiconductor at its free surface may form surface-
localized electronic states within the semiconductor bandgap and/or a double layer of charge, known
as a surface dipole.
The appearance of surface-localized acceptor states in n-type semiconductors induces charge
transfer between bulk and surface in order to establish thermal equilibrium between the two. The
charge transfer results in a non-neutral region (with a non-zero electric field) in the semiconductor
bulk, usually referred to as the surface space charge region (SCR) [168]. This region depleted of major-
qffiffiffiffiffiffiffiffi
ity carriers extends approximately a few Debye Lengths LD ¼ qekT 2 N into the bulk; typical values of LD
d

for tin oxide ranges from 130 to 10 nm when temperature changes from 400 to 700 K.
In addition to surface states, another important phenomenon associated with a semiconductor sur-
face is the surface dipole [169]. An adsorbate layer may result in a surface dipole, the magnitude of
which depends on the ionicity of the adsorbate–substrate bond. The surface dipole manifests itself
as a step in the electric potential at the surface because the potential changes abruptly over several
monolayers. This is in contrast to the macroscopic dipole created by the surface states and surface SCR.
As for conduction, metal oxide gas sensors are generally operated in air in the temperature range
between 500 and 800 K where conduction is electronic and oxygen vacancies are doubly ionized and
fixed. In quasi 1D gas sensors the current flows parallel to the surface. When the nanowire is fully
depleted, carriers thermally activated from surface states are responsible for conduction. Indeed
when considering nanowires bundles, the conduction mechanism is dominated by the inherent
intercrystalline boundaries at nanowires connections – like in polycrystalline samples – rather than
by the intracrystalline characteristics; the intergranular contact provides most of the sample resis-
tance [170].
The metal semiconductor junction that forms at the interface between the layer and the contacts
can play a role in gas detection, enhanced by the fact that the metal used for the contact acts also as a
catalyst. The contact resistance is more important for single nanowires since it is in series to the semi-
conductor resistance that for bundles where it is connected to a large number of resistances.

6.1.1. Surface adsorption


The first step of association of gas species with a solid surface is physisorption, afterwards the
physisorbed species can be chemisorbed if they exchange electrons with the semiconductor surface,
becoming chemisorbed [171]. When the adsorbate acts as a surface state capturing an electron or
an hole chemisorption is often named ionosorption.
Physisorption is a slightly exothermic process characterized by high coverage at low temperature
and a low coverage at high temperature. If the partial pressure is very low, Henry’s Law applies and the
amount physisorbed is simply proportional to the partial pressure. In contrast of physisorption, that is
a slightly exothermic, inactivated process, chemisorption and desorption are activated processes. The
activation energies can be supplied either thermally or by a non-equilibrium process such as
illumination.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 31

The adsorption isobar, that is the volume adsorbed as a function of temperature at a constant pres-
sure, is characterized at low temperature by physisorption and at high temperature by equilibrium
chemisorption – that decreases exponentially with temperature. Being the desorption greater than
chemisorption energy, in the intermediate region irreversible chemisorption takes place. Identifica-
tion of optimal operating temperature for a MOX sensor is a crucial step, since they must be obviously
operated in a temperature range where reversible chemisorption takes place or alternative ways to
induce desorption must be provided.
As far as the relation between chemisorption and electrical properties is concerned in the simple
charge transfer model CTM [171,172], the physisorbed and chemisorbed atoms and molecules are rep-
resented by localized states in the semiconductor energy gap, whose occupation statistic is given by
the same Fermi-Dirac distribution, physisorption corresponding to unoccupied and chemisorption to
occupied states.

6.1.2. Detection through surface reactions


The process of gas detection is intimately related to the reactions between the species to be de-
tected and ionosorbed surface oxygen. In the temperature range between 400 and 800 K oxygen ion-
osorbs over SnO2 and other oxides in a molecular (O 
2 ) and atomic form (O ) [173]; when a reducing
gas like CO comes into contact with the surface, it oxidizes to CO2 by reacting with ionosorbed oxygen,
releasing electrons from surface states to the conduction band. The overall effect at equilibrium is
shrinking of the density of ionosorbed oxygen, i.e. occupied surface acceptor states. Indeed surface
reactions are still debated: CO sensing, for example, could take place through reaction with hydroxyl
groups, producing atomic hydrogen that recombines with oxygen lattice and releases a free electron
[174] and even by direct adsorption as CO+ [175].
Direct adsorption is also proposed for the gaseous species –like strongly electronegative NO 2
whose effect is to decrease sensor conductance e þ NO2;ads ¡ NO 2;ads . The occupation of surface states,
which are much deeper in the bandgap than oxygen’s, increases the surface potential and reduces the
overall sensor conductance.
An important ubiquitous species that ionosorbs over MOX surfaces is water [173]. The chemisorp-
tion of water onto oxide from air can be very strong, forming an ‘‘hydroxylated surface”, where the
OH ion is bounded to the cation and the H+ ion to the oxide anion. The overall effect of water vapor
is to increase the surface conductance, although the surface reactions are still debated and many
mechanisms have been proposed [176,172,177].

6.1.3. DC resistance transduction


The easiest measurable physical quantity that can be transduced is the sensor conductance in DC
conditions. In laboratory tests it is normally measured by a voltamperometric technique at constant
bias while in commercial gas sensors the layer is usually inserted inside a voltage divider.
The sensor response towards a target gas concentration for metal oxide semiconductor sensors is
usually defined as the (relative) change of conductance, with the obvious relation
GS GS  G0
SG ¼ ¼ þ 1 ¼ SG þ 1
G0 G0
In case of an oxidizing target species, resistance increases following gas introduction and the sensor
response is defined as the (relative) change of resistance
RS RS  R0
SR ¼ ¼ þ 1 ¼ SR þ 1
R0 R0
Starting from the sensor response it is possible to derive the sensor response curve, that is the repre-
sentation of the steady state output as a function of the input concentration [178]. The sensor response
curve is frequently called, erroneously, sensitivity curve.

6.1.4. Conductometric gas sensors


Starting from the first paper that pointed out the feasibility to use bundles of nanowires as gas sen-
sors [179] there is plenty of literature regarding gas-sensing with conductometric nanowires. The aim
of this section is to give a critic review of the work presented in literature about conductometric gas
32 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

sensors. The topic will be divided in section regarding different materials used for nanostructure prep-
aration and gas sensing.
As shown in Table 1, the majority of the works found in literature is based on bundles of nanostruc-
tures. In this case, many of these nanostructures are contacted and their electrical parameters esti-
mated. However, only some mean values of these parameters are determined due to the dispersion
existing among the contacted nanostructures and the grain boundaries among them. Where possible,
we will report separately papers which presents measurements on bundles of nanowires from that on

Table 1
Table of conductometric gas sensors based on nanostructrures of metal oxide classified by sensing oxide
Type of metal Gases tested Multiple or single Wire cross section Working References
oxide nanowire diametera (nm) temperature (°C)b
SnO2 Ethanol, CO, NO2 Mesh of nanobelts 200 wide 20–40 thick 200–400 [179]
SnO2 Ozone Mesh of nanobelts 100 wide 20–40 thick 400 [180]
SnO2 Ethanol Mesh of nanowhiskers 50–200 300 [181]
SnO2 CO, O2 Single nanowire 60 200–300 [182]
SnO2 CO, relative humidity Single nanowire 25, 70 200–295 [184–186]
SnO2 NH3 e CO single nanowire 100 60–300 [187]
SnO2 NO2 Single nanobelt 80–120 wide 10–30 thick UV activated [188]
SnO2–Pt/Ni H2, CO single nanowire 60 400 [190]
SnO2:Sb Ethanol Mesh of nanowire 40–100 300 [191]
ZnO Ethanol,H2 Nanopillars 10–30 200–400 [192]
ZnO Ethanol Bridging nanowires 80 300 [193]
ZnO Ethanol, methane, 2- Tetrapod, paste 20–100 150–400 [195]
butanone,
triethylamine,
isopropanol
ZnO Ethanol Mesh of nanowire 25 300 [194]
ZnO Ethanol, H2S, HCHO, Nanowire, paste 40–80 330 [196]
LPG, 90# gasoline,
CO, ammonia
ZnO Ethanol, H2, CH4, Nanorods, paste 70–110 RT-350 °C [197]
NH3, H2S
ZnO H2, ozone Nanorods 30–150 RT-112 °C [198]
ZnO Methanol, NO/NO2 Aligned nanowires 60–80 325 [199]
ZnO–Pd H2 Mesh of nanorods 30–150 RT [200]
ZnO–Pt H2 Mesh of nanorods 50–100 RT [201]
ZnO–Pt H2, ethanol Paste of nanorods, 30–40 150 [202]
nanowires, nanotubes 20–30
60–100
ZnO–He+ H2S Single nanowire 400 RT [203]
implanted
In2O3 Acetone, NO2 Mesh of nanowires 100–500 200–500 [34]
In2O3 Ethanol, methanol, Nanowires, paste 60–160 370 [204]
CH4
WO3 CO, NO2 Bundles of fibers 20–100 150 [205]
WO2.72 NH3, Ethanol, NO2 Mesh of nanorods 4 50–200-UV [207]
activated
V2O5 Ethanol, methanol, Mesh of nanobelts 60–100 wide 150–400 [208]
H2S, H2, NH3, CO, NOx 10–20 thick
V2O5 NH3, 1-propanol, Mesh of nanobelts 10 wide RT [209]
toluene,1- 1.5 thick
butylamine
SnO2 H2, CO Array of single 100 350 [210]
SnO2–Ni nanowires 100
In2O3 1000
TiO2 1000
a
Diameter is used for wires, while for nanostructure with rectangular cross section width and thickness are reported
(specified).
b
Is indicated when the sensor is operated at room temperature with UV light activation by ‘‘UV activated” sentence.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 33

single nanowires. The technical difficulties in performing reliable electrical contacts on one individual
nanostructure in a controlled fabrication process at the nanoscale level has restricted the number of
works on this topic. Indeed the study on this topic allows reaching a better comprehension of the elec-
trical transport mechanisms which take place in these nanostructures.
Chemical gas-sensing performance of nanostructured MOX will be classified not only on the base of
the target species, but also on the base of a few very important parameters: gas carrier, humidity and
operating temperature.
Few parameters should be carefully controlled for a proper gas sensor characterizations: gas car-
rier, humidity and working temperature.
Humid air should be used instead than nitrogen as gas carrier, because ubiquitous oxygen and
water vapor play a crucial role on the semiconducting surface. Sensors are heated in order to activate
surface reactions with gaseous species. Photochemical activation has also been used.
Since many authors did not follow a correct procedure for gas testing, it is quite difficult to compare
results from different groups. A constant flux of gaseous species at ambient pressure should be em-
ployed instead of vacuum ambient with injection of desired amount of gases. The widespread use
of adhesion agent to realize a gas-sensing paste of nanowires leaves some doubts about the influence
of the agent on gas sensing measurements.

6.1.4.1. Tin dioxide


6.1.4.1.1. Multiple nanowires. Comini et al. [179] were the first to present their results on nanobelts
of SnO2 with a rectangular cross section (200 nm large-20/40 nm thick) in a ribbon-like morphology,
prepared by a thermal evaporation of oxide powders under controlled conditions without the pres-
ence of a catalyst.
After deposition, a bunch of nanobelts was placed onto the interdigitated Pt electrodes to measure
their electric conductance. They obtained a good response to gaseous polluting species like CO and
NO2 for environmental applications, as well as to ethanol ð250 ppm;SG ¼ 41:6) for breath analyzers
and food control applications. All measurements were done at 400 °C working temperature for reduc-
ing gases and 200 °C for oxidizing gases in humid air (30% RH) (see Fig. 30).
They have shown the experimental feasibility of fabricating nanosized sensors using the integrity
of individual nanobelts, even if the size of the nanobelts is quite big in order to have a complete deple-
tion of the conduction path. Moreover the conduction should be governed by the barrier between the
belts contacting each other.
The same group reported later the use of mesh of SnO2 nanowires for ozone sensing at 400 °C of
very low concentration of ozone (70 ppb) [180].

Fig. 30. Response of bunch of SnO2 nanobelts at 400 °C working temperature towards 250 ppm of ethanol [179]. Reprint
permission 1905370465707.
34 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ying et al. [181] reported the synthesis of SnO2 nanowhiskers by thermal evaporation of metal Sn
powder. SnO2 nanowhiskers have a rectangular cross section, and the diameter is in the range of 50–
200 nm. They found that at 300 °C the response to 50 ppm of ethanol is SG ¼ 23.
Given the difference in ethanol concentration and in the working temperature, and the fact that
these measurements are done in dry and not in humid air, the results is consistent with that found
before by Comini et al. [179].

6.1.4.1.2. Single nanowires. Kolmakov et al. [182] realized an array of parallel nanowires of SnO2 by
using self organized highly ordered porous alumina templates. Conductance measurements were per-
formed on isolated individual nanowires with vapor deposited Ti/Au micro-contacts. They analyzed
the behavior of nanowire in N2 and N2 + 10%O2. The nanowire is a fairly good conductor in N2 and
is converted into an insulator in presence of sufficient oxygen. This behavior is attributed to ionized
oxygen vacancy that acts as surface acceptor state.
CO, a reducing gas, reacts with pre adsorbed oxygen species on SnO2 to form carbon monoxide thus
donating few electrons back into the bulk resulting in an increased conductivity, which depends
monotonically on the gas phase partial pressure of CO (Fig. 31).
The electron exchange between the surface states and the bulk takes place within a surface layer
whose thickness is of the order of the Debye length. The key feature is that nanowires have radius of
the order of or less than kD, which for bulk SnO2 is about 43 nm at 500 K [183].
This paper is interesting for what concerns a confirmation of the mechanism of reaction of reducing
gases with nanowires, even if it is difficult to compare with other gas sensors due to peculiar baseline
gas (N2 + 10%O2) used instead than synthetic air. Moreover the CO concentration used here is huge (or
the order of magnitude of 103 ppm).
The papers of Hernandez-Ramirez et al. [184,185] deals with the challenge of performing reliable
electrical contacts on one individual nanostructure in a controlled fabrication process at the nano-
scale level. The single nanowire device was then tested as gas sensor. Platinum stripes were depos-
ited by FIB to realize contacts on single nanowire previously deposited and nano/micro manipulated
onto a silicon wafer, already processed to have an adequate distribution of microelectrodes. The
manipulation defines a better placement of the nanostructure in front of the microelectrodes before
to proceed with the Pt deposition. Pt could be deposited either by ion and electron beam. Using elec-
tron beam deposition, a couple of contacts are fabricated near each extreme of the nanostructure
requiring much less than a half micron (see Fig. 32), remaining a significant length of the nanowire,
more than several couples of microns, free from damage caused by Ga+ ions. Once the influence area
due to the deposition is far away of the nanostructure, ion deposition methodology is used to extend
the stripes to the pre-patterned contacts, because the deposition time is much shorter than using the
electron one.
Measurements at room temperature showed that contact resistance contribution is much more
important than the nanowire resistance; the main part of the measured contact resistance is believed
to be originated in the Pt–SnO2 nanowire junction. In order to measure the electrical response of nano-
wire, they tested the fabricated gas sensors towards different combustion gases such as, CO, relative
humidity and NO2, using AC Impedance Spectroscopy technique. Fig. 33 shows the dependence of the
sensor response to CO of individual nanowire on its radius: for radius equal or smaller than 25 nm an
increase in sensor response and in the sensitivity (slope of the curve in Fig. 33) is observed.
Hernandez-Ramirez et al. [186] studied single nanowire SnO2 device contacted with the same pro-
cedure described above, to give insight on the cross-sensitivity effects between water and carbon
monoxide gases: non linear dependence on the concentration of water vapor was found. Moreover,
they observed higher responses to CO in synthetic air atmosphere compared to nitrogen environ-
ments, due to role of chemisorbed oxygen on the nanowire surface when in air.
Meier et al. [187] presented a technical approach to fabricate practical devices by coupling a single-
crystal SnO2 nanowire sensing element with a microhotplate gas sensor platform (Fig. 34).
Similar to MOX thin film sensors, the response signals and response times for NH3 and CO (100, 50
and 25 ppm) improves with increasing temperature, confirming the thermal activation of the redox
reaction occurring at the nanowire surface. The maximum relative response SG was achieved at
approximately 260 °C, being 0.35 for NH3 and approximately 10 times smaller for CO. The response
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 35

Ó Wiley-VCH Verlag GmbH & Co.GaA 2006


Fig. 31. Top: Response of the SnO2 nanowires towards CO pulses; Down: Change in conductance of individual SnO2 nanowires
as a function of CO concentration at three temperatures. Reproduced with permission. Ref. [182].

time constant decreases as the sensor temperature increases a consequence of the temperature
dependent interaction of the analyte with the oxide depletion layer.

6.1.4.1.3. SnO2 gas sensor activated by light. Law et al. [188] reported the use of light to photoacti-
vate molecule absorption and desorption as an alternative way to the use of temperature. An explosive
environment is the typical application where the use of heated sensors is not favorable. The strong
photoconducting response of individual single-crystalline SnO2 nanoribbons makes it possible to
achieve equally favorable adsorption–desorption behavior at room temperature by illuminating the
devices with ultraviolet (UV) light of energy near the SnO2 bandgap. The active desorption process
is thus photoinduced molecular desorption.
The conductance rise when the device is illuminated with UV light is due both to the generation of
free carriers within the device, and to photodesorption of surface species (mostly O2 and H2O de-
rived) [189] with a concomitant thinning of the electron depletion layer near the nanoribbon surface.
The effect of the 365 nm radiation, which corresponds to an energy lightly smaller than the bandgap of
SnO2 (345 nm), is attributed to the presence of surface states that populate the energy gap, obtaining
results better than using 254 nm radiation.
36 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ó American Physical Society 2007


Fig. 32. SnO2 nanowire (length L = 11 lm, r = 27 ± 3 nm) electrically contacted using dual-beam FIB nanolithography techniques
[131]. Reprinted figure with permission from Ref. [131]. http://link.aps.org/abstract/PRB/v76/e085429.

Fig. 33. Dependence of sensor response SG from CO concentration for single nanowires with different radius. Reprint
permission 1897150915098, Ref. [185].

The device consisted of a single SnO2 nanoribbon (80–120 nmwide 10–30 nm thick) deposited by
dispersion in ethanol onto prefabricated gold electrodes in four terminal configuration, and was cycled
between NO2 (5–100 ppm) and vacuum (see Fig. 35). 100 ppm is a huge concentration for NO2 (alarm
level is usually fixed to 0.2 ppm) and the sensor was not able to distinguish between lower concentra-
tions. UV illumination seemed vital for sensing, as NO2 adsorption is irreversible in the dark.
6.1.4.1.4. Catalyzed and doped SnO2 nanowires. Metal oxide nanowires can be used also as a sup-
port for catalyst metals. Addition of a small amount of noble metals such as Au, Pd, Pt and Ag and
Ni speeds up surface reactions and improves selectivity towards target gas species. The effect of cat-
alyst is to provide a more favorable reaction path. They should be dispersed as small crystallites over
the surface of the oxide in order to be active near the grain boundaries where carrier transport takes
place. Poisoning and mainly coalescence by surface migration of the metal particles at high tempera-
ture could deactivate the supported catalyst and should be carefully avoided.
In order to tune the sensitivity and selectivity of SnO2 nanowires based sensors Kolmakov et al.
[190] examined the influence of the surface sensitization with catalyst particles of Ni/NiO and Pd.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 37

Fig. 34. Top: optical image of sensing micromachined platform. Bottom: SEM image of the individual SnO2 nanowire integrated
as a chemiresistor using FIB technology. Permission 1905370798263, Ref. [187].

Ó Wiley-VCH Verlag GmbH & Co.KGaA 2006

Fig. 35. Cycling SnO2 nanoribbon between decreasing concentrations of NO2 (peaks) and 8  105 mbar vacuum under 254 nm
UV light. Current was held constant at 5 nA. Reproduced with permission. Ref. [188].

The nanowires were placed on the surface of Si/SiO2 wafer without use of any solvents. The deposition
of metal catalyst (Pd, Ni) was performed in situ via thermal evaporation of metal from a miniature
filament.
The authors performed a comparative sensing studies of the SnO2 nanowires as a function of few
sequential Pd and Ni depositions, which were performed in situ in the same chamber.
Few major trends were seen in Fig. 36: (i) the conductance through the nanowire in vacuum and in
oxygen progressively dropped with the increase of the amount of Pd on its surface. This observation
supported the hypothesis of the formation of a catalyst induced depleted region in the nanowire; (ii)
the nanowires with the functionalized surfaces either with Ni or Pd demonstrate an improved re-
sponse to O2, H2 and CO. The major improvement took place at small dosage and then gradually in-
creased with the trend to saturate at higher coverage; (iii) the kinetics of the nanowire sensor
response was improving when the surface was catalyzed.
38 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Fig. 36. Gas-sensing measurements of an individual SnO2 nanowire chemiresistor as a function of Pd deposition. Qualitatively
similar results were obtained for Ni deposition. The dashed lines guide the eye through the major observed trends in the
performance of the chemiresistor. The inserts (linear scale) comparatively demonstrate the responsiveness of the pristine (left)
sensor and the same one functionalized with Pd nanoparticles (on the right) to H2 and CO pulses in the oxygen background.
Reprint from [190].

Wan et al. [191] reported the use of Sb doping to tailor the resistivity of SnO2 nanowires deposited
by thermal evaporation process starting from a mixture of Sn and Sb powders in the ratio 10:1.
The resistance of the undoped SnO2 nanowire film was very high, probably due to their high stoi-
chiometry, which gave a semi-insulating electrical property. After Sb doping the resistance is calcu-
lated to be about 2.5  104 X. The resistance decrease is due to the formation of shallow donor
levels by introduction of Sb+5 into SnO2, which is very favorable for gas sensor application. To test sen-
sor application they have used a mesh of nanowires between two Au electrical contacts and tested it at
300 °C towards ethanol (10–1000 ppm) diluted in dry air.
They claim to obtain very low response and recovery times to 10 ppm of ethanol (1 and 5 s, much
shorter than that reported from Comini et al. [179]) and attribute this effect to Sb doping that would
favor or accelerate the absorption of oxygen molecules and the formation of O 2 ions on the surface of
SnO2 nanowires, which is of great significance to reduce the recovery times of the fabricated sensor.

6.1.4.2. Zinc oxide. As far as zinc oxide is concerned, there is plenty of literature on its usage as gas sen-
sor, mainly as mesh of nanowires and not as single nanowire. It can be obtained also in the form of
aligned nanowires, that permit peculiar contacts geometries.
Nanopillars of ZnO were prepared by Bie et al. [192] by mean of a chemical two-step route. Aligned
zinc oxide nanorods with diameter in the range 10–30 nm grew on Al2O3 tube directly to form a nano-
pillar ZnO array sensor.
The electrical resistance was monitored with hydrogen and ethanol in concentration 10–2500 ppm
diluted in dry air as a function of temperature in the temperature range of 200–440 °C.
The response of the sensor to ethanol is almost linear in the range of 250–2500 ppm, but the re-
sponse to hydrogen did not vary so much in the concentration range from 500 to 2500 ppm. The
ZnO nanorod sensor exhibited response SG = 18.3 and 10.4 to 100 ppm ethanol and hydrogen, respec-
tively. The high response was attributed to the large effective surface area of the aligned nanorods.
Hsueh et al. [193] realized a patterned contact made of 50-nm thick Ga-doped ZnO thin film prior
to the growth of ZnO nanowires. The nanowires were grown subsequently by evaporation in oven,
selectively on Ga doped ZnO film; only some ZnO nanowires (80 nm diameter) grew laterally and
bridged two neighboring fingers to provide electrical paths. With these laterally grown ZnO nano-
wires, the two neighboring electrodes were no longer electrically open and the contact resulted ohmic.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 39

The authors tested response to ethanol by changing between vacuum and ethanol (1500 ppm). The
optimum working temperature was 300 °C and relative response was 0.6, so it is quite small if com-
pared to other ZnO sensor and considering high ethanol concentration used. This could be due to the
use of only few nanowires with large (compared to LD) diameter that contribute to the process, since
only the bridging wires are involved. In this case the effect of complete depletion of the wire as well as
that of enhancement of surface exposed to gas is lacking. Nevertheless the paper presents a peculiar
way to grow single ZnO nanowires between neighboring electrodes.
Wan et al. fabricated ZnO [194] nanowires by thermal evaporation of Zn under controlled condition
without a metal catalyst. ZnO nanowires (diameter 25 nm) were ultrasonic dispersed in ethanol for
30 min and dried by an infrared light on a silicon-based membrane embedded with Pt interdigitated
electrodes and a heater, in order to reduce the power consumption.
They have tested the sensing properties of ZnO nanowires upon exposure to 1, 5, 50, 100, and
200 ppm ethanol in dry air at 300 °C working temperatures. Sensor response SG at 200 ppm ethanol
exposure is measured to be 47, and SG at 1 ppm ethanol exposure to be 1.9, thus giving good perfor-
mances for this sensor. This can be explained considering that if the diameter of the ZnO nanowires is
comparable to LD, it is expected that in presence of adsorbed surface state the SCR occupies the entire
wire, thus switching conductivity properties of the material.
Xiangfeng et al. [195] investigated the effect of preparation atmosphere on sensing properties of
ZnO nano-tetrapod prepared by evaporation and then mixed with PVA (Polyvinyl alcohol) solution
to coat on Al2O3 tube. The gas-sensing properties are strongly dependent on the preparation atmo-
sphere due to the effects of this atmosphere on the intrinsic defect concentration. They found that
the sensors based on tetrapods prepared in humidified Ar flow (leg of the tetrapod has a diameter
20–100 nm) exhibited the best performance to ethanol. They have tested response towards ethanol
(diluted in humid Ar), methane, 2 butanone, triethylamine, isopropanol in the temperature range
150–400 °C (reported in Fig. 37). This sensor proved to be a good methane sensor (1000 ppm) and eth-
anol sensor (100 ppm). Relative response to ethanol peaked at 300 °C. Reponse to other gases tested
was lower at this temperature. Since the gas carrier used is Ar rather than air, the behavior of the sen-
sor could be much different in air.
Xu et al. [196] prepared ZnO via hydrothermal method process. The obtained powders, consisting
of ZnO rods of diameter 40–80 nm, were mixed and ground with adhesion agent to form gas-sensing
paste. The paste was coated on an alumina tube on which a pair of Au electrodes was previously
printed. Response to 100 ppm ethanol in ambient air was SG = 13.0.

Fig. 37. Response SG of ZnO tetrapods to different gases diluted in humidified Ar (ethanol 100 ppm, other gases 1000 ppm).
Reprint permission 1901811229751, Ref. [195].
40 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Fig. 38. Response of ZnO nanowire gas sensor to different gases at 330 °C working temperature. Reprint permission
1901830034138, [196].

The authors also detected the response of these sensors to other gases such as H2S (2 ppm), HCHO
(50 ppm), LPG (500 ppm), and 90# gasoline (50 ppm), CO (500 ppm), ammonia (50 ppm), as reported
in Fig. 38. ZnO sensors showed relative high sensitivity to inflammable gases such as LPG, alcohol, 90#
gasoline, so they can be used as generally sensitive sensor to detect inflammable gases.
Wang et al. [197] fabricated ZnO nanorods by a hydrothermal method with our without adding SDS
surfactant. They obtained nanorods (diameter 70–110 nm) with differences in surface roughness.
Afterwards the authors prepared a paste from a mixture of ZnO nanorods with a PVA solution and
coated it onto an Al2O3 tube on which two gold leads had been installed at each end. They have tested
response towards 1000 ppm of H2, CH, NH3, ethanol and 10 ppm H2S. Gas carrier was dry air. Sensors
exhibited a high response to ethanol when operated at 350 °C and a high response to H2S at 25–200 °C.
Lower response was registered to H2, CH4 and NH3, as one would expect, since these three gases are
less reactive than H2S and ethanol. The responses SG to 1 and 10 ppm of ethanol were as high as 2.6
and 6.3 at 350 °C. Sensor reacts reversibly with very low H2S concentration 0.1 ppm at room
temperature.
Kang et al. [198] reported of ZnO nanorods grown by molecular beam epitaxy. They achieved site-
selective growth of ZnO nanorods (diameter 30–150 nm) by nucleating the nanorods on a substrate
coated with Au islands.
The nanorods are sensitive to the presence of hydrogen or ozone in the measurement ambient for
temperatures as low as 112 °C for H2 and room temperature for O3. However the concentration used
is huge both for ozone (3%) and for H2 (19%), and the gas carrier is dry nitrogen rather than air.
Parthangal et al. [199] reported a novel in situ approach toward connecting and electrically con-
tacting vertically aligned nanowire arrays (60–80 nm diameter) using conductive nanoparticles.
Well-aligned, single-crystalline zinc oxide nanowires were grown through a direct thermal evapora-
tion process at 550 °C on gold catalyst layers. The result is a device, schematically reported in Fig.
39, which is an ensemble of single nanowire devices connected in parallel. For sensor applications
there may be signal to noise advantages in such an arrangement compared to single nanowire devices.
The device is interesting, even if the results obtained in terms of response are not very good for nitrous
oxide, since concentrations reported are very high. They tested sensors at 325 °C towards pulsed con-
centrations (10–50 ppm) of methanol and nitrous oxides in dry air. The sensor response clearly tracks
the pulsed input of the analyte. However, the recovery time for the sensor to retain its original resis-
tance was somewhat high, possibly due to slow desorption rates from the nanowire matrix. Moreover
the stability of a porous electrode is not confirmed by long-term studies.

6.1.4.2.1. Catalyzed and defective ZnO. Wang et al. [200] presented ZnO nanorods deposited by
Molecular Beam Epitaxy on a substrate coated with Au islands. The authors deposited Pd by sputtering
on ZnO surface in order to enhance hydrogen detection properties. The response of Pd-catalyzed sen-
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 41

Fig. 39. Schematic diagram of electrical contact to ZnO aligned nanowire and I–V characteristic measured in air. Reproduced
with permission from [199].

sor to H2 at room temperature (10 ppm–500 ppm) was higher compared to that of non catalyzed sen-
sor, being SG = 2.6 towards 10 ppm of H2 in dry air. An activation energy of 12 kJ/mole was calculated
from the slope of the Arrhenius plot of nanorod resistance, when exposed to 500 ppm of H2 in N2.
The same authors [201] made a comparison of the performances in H2 detection between Pt-coated
single ZnO nanorods and thin films of various thicknesses (20–350 nm), grown by pulsed laser depo-
sition on sapphire substrates. The site-selective growth of ZnO nanorods was achieved in the same
way as described in the previous contribution [200]. Pt, a well known catalyzer for H2 detection,
10 Å thick, was deposited by sputtering on films and nanorods to enhance response to H2.
Pt-coated nanorods had a larger response (SG = 1.08 for 500 ppm H2) than thin film roughly a factor
of 3, at room temperature. This is consistent with the larger surface-to volume ratio of the nanorods
compared with thin film prepared by PLD, which are indeed very flat. However the nanorods showed a
slower recovery than the thin films, most likely due to the relatively higher degree of hydrogen
adsorption.
If we compare the results with that described before, where the ZnO nanowires are catalyzed with
Pd, it seems that results for H2detection are much better with Pd than with Pt, giving higher response
to a H2 concentration 50 times lower.
Rout et al. [202] prepared many different morphologies of ZnO nanostructures: nanorods (diameter
30–40 nm), nanowires (diameter 20–30 nm) and nanotubes (diameter 60–100 nm). They have pre-
pared thick film sensor mixing appropriate quantity of diethyleneglycol to the nanostructures of
ZnO and the mixture ground to form a paste. The paste was coated on to an alumina substrate with
a comb-type Pt electrode on one side, the other side having a heater. Among all, nanowires showed
the best sensing properties towards hydrogen (down to 10 ppm of H2) sensing at relatively low tem-
peratures (150 °C), especially when impregnated with Pt.
Liao et al. [203] presented a work on ZnO used as single nanowire sensor; the diameter of the wire
was quite big (d = 400 nm) and the nanowire was scratched from Si substrate and dispersed in ethanol
to place it over a substrate with Au electrodes. Pt contacts were subsequently grown by FIB. He+
implantation at different doses was used to change the amount of defects in ZnO nanowire. At high
42 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Fig. 40. (a) The schematic diagram of configuration and SEM image of a single ZnO nanowire gas sensor. (b) The I–V curves and
(c) Dynamic response curves for different H2S concentrations of a single ZnO nanowire at different implanted statuses.
Permission 1905371034868, [203].

doses 1  1017 cm2 the photoluminescence UV signal disappeared due to poor crystalline quality; the
material is amorphous with very low conductive electron concentration. It was observed that ZnO im-
planted at 1  1016 cm2 had higher H2S sensitivity than unimplanted (see Fig. 40).
The dependence of gas response on implantation was attributed to the modulation of defects on
the electron concentration in the nanowire. Resistance in air was modulated by SCR due to adsorbed
oxygen ions. After exposing it to H2S the trapped electron was released into ZnO, causing an increase
in electron concentration. Implanted nanowires had much less conductance electron thus the differ-
ence is bigger than for unimplanted nanowire.

6.1.4.3. Indium oxide. Vomiero et al. [34] reported preparation of In2O3 nanowires on alumina sub-
strates by thermal evaporation at 1500 °C. The prepared material showed a broad distribution on
the diameter of the wires from 100 nm to 500 nm. The different size between nanowires and thick
wires is expected to play a key role in determining the amplitude of the response to reactive
gases. Nanowires may also behave differently because their electrical conductivity could depart
from the diffusive regime of ionic crystalline semiconductors. Therefore, variation in size for the
nanowires could significantly affect both electrical conductivity and gas sensitivity. The electrical
conductance is higher for the thick wires. The enhanced surface-to-volume ratio in thin wires
determines a reduction of the conductance, as electron flow is highly reduced in the surface de-
pleted region.
Fig. 41 shows that the response to acetone increases with the operating temperature up to a max-
imum located at about 400 °C, where the nanowires feature a response about 7 times higher than the
one of thick wires. This result was attributed to the higher surface-to-volume ratio of nanowires,
which causes more extended surface interaction with gas molecules, and suggested that the control
of lateral dimension of the wires was the key point for achievement of high sensitivities. For NO2 sens-
ing at very low concentration (500 ppb), as well as in the case of acetone, nanowires exhibited higher
response with respect to thicker ones.
Xianfeng et al. [204] prepared gas sensor based on In2O3 nanowires (diameter ranging from 60 to
160 nm) by carbothermal reduction reaction. The paste formed from a mixture of In2O3 nanowires
with PVA solution was coated onto the alumina substrate. The sensor showed response SG = 2 to
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 43

Ó American Chemical Society 2007


Fig. 41. Response (SG for CO and SR for NO2) of In2O3 nanowires and thick wires as a function of the operating temperature
toward (a) 25 ppm of acetone and (b) 500 ppb of NO2. Reprint with permission from [34].

100 ppm ethanol at 370 °C while almost showed no response to less reactive gases like 1000 ppm CH4
and CH3OH gas when operating at 150–350 °C. Gas carrier was dry air.

6.1.4.4. Tungsten oxide. Polleux et al. [205] reported a soft-chemistry route to prepare crystalline tung-
sten oxide nanofibers self-assembled into bundles. The width of a bundle is between 20 and 100 nm,
while diameter of a single fiber is about 1 nm.
They studied gas-sensing properties of the tungsten oxide nanowire bundles for detection of low
levels of NO2. Dry air was used as gas carrier for the gas-sensing tests. A suspension of the tungsten
oxide nano-bundles in ethanol was deposited onto alumina substrates by drop-coating and subse-
quently calcinated at 500 °C in air, resulting in the formation of a highly porous network as sensing
layers.
The tungsten oxide nanowires were nearly insensitive to CO in the concentration range between 1
and 10 ppm at all operating temperatures studied. Instead, the tungsten oxide nanowire sensors ex-
hibit a high signal (SG = 9) at particularly low NO2 concentrations of about 50 ppb (Fig. 42), which is
much better if compared with results obtained with SnO2 [179] and In2O3 described in Fig. 41, where
SG is 8–500 ppm of NO2. This behavior was explained by high surface area and good crystallinity of
WO3 fibers.
The calibration curves and their dependence on NO2 concentration was approximated by power
laws. The dependence of the sensor signal (S) on the NO2 concentration was described as SG = 1 +
a[NO2]b, where a and b are variables that can be fitted to a = (0.06 ± 0.05) and b = (0.89 ± 0.13). The
experimental results correspond to the model developed by Gurlo et al. [206] thus indicating the for-
mation of surface-trapped states NO2,ads with no evidence of dissociative NO2 adsorption. The iono-
sorption of nitrogen dioxide at the surface of the tungsten oxide nanowires leads to the upward
bending of the bands (120 meV for 50 ppb NO2 at 150 °C) and accordingly to an increase of the resis-
tance, thus giving evidence for the n-type semiconducting properties of the tungsten oxide nanowires.
Kim et al. [207] deposited porous tungsten oxide films onto a sensor substrate with a Si bulk-
micromachined hotplate, by drop-coating isopropyl alcohol solution of highly crystalline tungsten
44 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ó Wiley-VCH Verlag GmbH & Co.KGaA 2006


Fig. 42. Change of the resistance R of the tungsten oxide nanowire bundles upon exposure to NO2 (50 ppb) at 150 °C in dry air;
grey curve: NO2 concentration measured simultaneously with an NOx chemiluminescence analyzer. From [205]. Reproduced
with permission.

oxide WO2.72 nanorods with average 75 nm length and 4 nm diameter. Sensor capabilities have been
tested against NH3, Ethanol, NO2 diluted in dry air from room temperature to 200 °C. The sensor
showed p-type behavior for temperatures less than 100 °C, n-type behavior at 200 °C. Obviously at
room temperature recovery is a problem that can be solved only by applying UV light (365 nm) to give
sufficient energy the molecule to desorb. No clear improvement with respect to a polycrystalline
nanosized film is here found, even if the very small dimension of the rods should give enhanced prop-
erties to the sensor.

6.1.4.5. Vanadium oxide. Liu et al. [208] reported the preparation of V2O5 nanobelts by hydrothermal
method, with rectangular cross section 60–100 nm wide and 10–20 nm thick.
Sensor showed good response to ethanol from 1000 to 10 ppm (SG = 1.7–200 ppm ethanol), while
response to other gases like methanol, H2S, H2, NH3 was very low and response to CO and NOx was
absent. Gas carrier was 40% humid air. Gas containing hydroxyl group (like alcohol) tended to be ad-
sorbed on V–O nanobelts by oxygen bonds and then reacted with the oxygen negative ion to change
the conductivity of the sensor: this explained good response to ethanol.
Raible et al. [209] have reported synthesis of V2O5 nanofibers with rectangular section (1.5–
10 nm) by mean of a wet-chemical process. V2O5 nanofibers are reported to have high affinities
to organic amines and they have tested gas-sensing properties towards ammonia, 1-propanol, tolu-
ene, 1-butylamine diluted in ambient air. Response to 10 ppm of 1-butylamine and to 10 ppm of
ammonia was SG ¼ 42 and SG ¼ 1:8, with a response a order of magnitude lower to 1000 ppm of
other gases.

6.1.4.6. Electronic nose approach. Sysoev et al. [210] have followed an electronic nose approach with an
array of single nanowires. They presented a comparative study of gas-sensing properties of individual
SnO2, surface doped (Ni)–SnO2 nanowires, In2O3, and TiO2 nano- and mesowires placed on the same
chip and wired as an array of chemiresistors. Conductance was measured in pulses of H2 and CO
reducing gases in oxygen as background gas. The nanostructures were shown to be n-type semicon-
ductors possessing high sensitivity to the target gases (except In2O3 that has been withdrawn from the
array due to low response to gas). Following the ‘‘electronic nose” concept, correlation analysis of re-
sponse of three-chemiresistor array is shown to be sufficient to discriminate between H2 and CO
signals.
The different morphology of these structures is a positive factor that, along with the compositional
divergence, contributes to a specificity of the response to a particular gas. The effective diameters of
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 45

Ó American Chemical Society 2006


Fig. 43. Current–voltage characteristics taken on individual metal oxide nanostructures in a vacuum at 400 K. The insert
diagram represents the effective cross section of the nanostructure where W is the width of the depletion zone. Reprinted with
permission from [210].

the nanostructures, D, were in the order of 100 nm for tin oxide nanowires, and ca. 1000 nm for the
titanium and indium oxide whiskers.
The measurements of the nanostructures’ conductance were performed: (i) under high vacuum
conditions, (ii) while admitting pure oxygen as oxidizing background gas up to a pressure of
1.3  102 Pa, and (iii) during pulses of H2 and CO having oxygen as a background gas.
The I–V curves were measured for each nanostructure to ensure the Ohmic character of the con-
tacts (Fig. 43). As one can see from Fig. 43, the thicker structures have higher absolute values of the
conductance roughly obeying a I  D2 trend.
The differences in the observed gas responses would obey roughly the degree of adsorbate induced
depletion/accumulation in the nanostructure’s SCR with respect to its diameter.
By acceptance of the coaxial geometry as a first approximation of the conducting channel (see in-
sert in the Fig. 43), the conductance through the nanostructure in an oxidizing environment is gov-
erned by SCR induced by adsorbed surface states.
As in the case of thin film sensors, the selectivity of the individual nanostructure was not sufficient
to discriminate between them. However the gas recognition was achieved using well-established
methods realized for ‘‘electronic noses” using three nanowires array.
In conclusions, the good performances reported in literature on conductometric gas sensors based
on nanowires are usually derived from a high surface to volume ratio due to nanosized wires or belt. A
comment that is often lacking in literature is the comparison of the results obtained with nanowires
with that obtained with thin or thick films. As a general feeling, results are similar, provided they have
similar surface to volume ratio.
Even if a big enhancement of the stability of the device at high working temperature is expected in
single-crystalline nanowires, no tests are reported in literature. The main problem is the long-term
stability of the contact when working at high temperature needed for gas-sensing, instead than coa-
lescence of grains inside polycrystalline films.
There is still a lot of work to do in the field of nanowire-based gas sensor, to reach a stable, inno-
vative device.

6.1.5. Single nanowire transistor (SNT) based gas sensors


One-dimensional metal oxide nanowires have been introduced in electronic devices like field-ef-
fect transistors to serve as the fundamental building blocks of highly integrated nanowires electronics
46 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Fig. 44. Scheme of a single nanowire transistor.

and chemical sensors. Nanowires, after being scratched from the substrates and dispersed in solvents
like ethanol or isopropanol, are deposited over SiO2/Si heavily doped substrates. Afterwards single
nanowires are contacted, usually by e-beam lithography, and Si substrate operates as back contact.
The scheme of a Single Nanowire Transistors is reported in Fig. 44.
The electrical properties of MOX Single Nanowire Transistors have been the subject of many arti-
cles [211–217]; in order not to introduce additional unpredictable effects produced by uncontrolled
gas adsorption, the NWs surface is not exposed to the gaseous environment, but it is often covered
by a protecting layer which passivates the surface states. A typical output characteristic (IDS versus
VDS) and transfer characteristic (IDS–VGS) of an n-type MOX NWs FET is reported in Fig. 45 [213].
The device works the same way as n-channel Thin Film Transistors (TFT)[218–220] operating in
accumulation mode upon application of a positive gate bias (see Fig. 46).
Application of a negative gate bias VGS depletes the channel of electrons and shuts the device off; a
positive VGS greater than VT creates a majority carrier accumulation conducting channel. At low VDS,
the SNT shows typical transistor behavior as IDS increases linearly with VDS (Fig. 46 left). Current sat-
uration is observed at high VDS as the accumulation layer is pinched off at the drain electrode (Fig. 46
right).
Assuming a cylindrical wire of radius r and length L and a gate oxide of thickness h  r, the capacity
C between the channel and the back gate when VGS is greater than VT and the n-channel builds up, is
given by [221,149]
2pee0

lnð2h=rÞ
Carrier concentration density n can be estimated by the relation

Q ¼ qnpr 2 L ¼ CðV GS  V T Þ

The relation between IDS and VDS is given by [218]


" #
ln C V2
IDS ¼ ðV GS  V T ÞV DS  DS
L 2

Mobility ln in the linear regime is estimated by calculating transconductance gm in the linear zone
@IDS lC
gm ¼ ¼ n2 V DS
@V GS L
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 47

Ó American Institute of Physics 2007


Fig. 45. Output characteristics (a) and transfer characteristics (b) for n-channel depletion-mode FET of ZnO NWs. The inset
shows the semilogarithmic plot of the IDS–VG curve at VDS = 0.1 V. (Reprint with permission from Ref. [213].

Fig. 46. Schematic cross section of a Single Nanowire n-type Transistor device when VGS > VT and a majority carrier
accumulation conducting channel forms. Left: linear region; Right: saturation region, channel is pinched off at the drain.

In saturation (VDS > VGS–VT)


ln C
IDS ¼ ðV GS  V T Þ2
2L
pffiffiffiffiffiffi
Mobility in the saturation regime can be calculated from the slope of the plot of IDS versus VGS. Due to
dependence of mobility from the gate potential the two calculated mobilities, in linear and saturation
regime, can be different and are considered only rough estimations.
The above reported treatment assumes that drain and source contacts are ohmic and discards con-
tact resistance; Go et al. [217] have developed a more complete model which take into accounts con-
tact resistance and short gate effects.
As for transistor type, enhancement (VT > 0) and depletion-mode (VT < 0) single nanowires transis-
tors has been built by employing nanowires prepared by different routes [213].
Single Nanowire Transistors prepared and studied as chemical sensors were based on n-type In2O3
[222–225], SnO2[226,14,227] and ZnO [228,229].
As long as the surface of the nanowires channel is exposed to the gaseous environment, acceptor
surface states induced by oxygen adsorption fill with electrons as schematically shown in Fig. 46.
When gate potential is floating, the surface space charge region (SCR) extends through all the nano-
48 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

wire nanometer size section, bands are flat (Debye Length LD for tin oxide ranges from 130 to 10 nm
when temperature changes from 400 to 700 K) and carriers thermally activated from surface states are
responsible for conduction. By applying a gate potential, the position of the electrochemical potential
and the availability of electrons for surface reactions can be modulated, exploiting the well known
electroadsorbitive effect [230].
The first observed effect of the gaseous environment has been the shift of VGS necessary to induce
the conducting channel inside the nanowires. Zhou et al. [222,225] first showed that at room temper-
ature VT of a SNT based on In2O3 increases following the introduction of NO2. As showed in Fig. 47 VT
shifted rightward monotonically at increasing NO2 concentration, implying a gradually suppressed
carrier concentration inside the nanowire with NO2 concentration.
NO2 directly ionosorbs over In2O3 and other metal oxides trapping conduction band electrons. The
occupation of NO2 related surface states, the energy of which is much deeper in the bandgap than oxy-
gen’s, decreases carriers density inside the nanowires and lowers electrochemical potential. Further-
more even physisorbed strongly electronegative NO2 may result in a surface dipole which manifests
itself as a step in the electric potential at the surface. As a consequence a greater threshold gate voltage
VT is required in order to bend down the conduction band at the gate/nanowire interface to build up
the majority carrier accumulation conductive channel.
The same behavior was determined by Fan et al. [228] in ZnO SNTs: VT shifted from 7.9 V in pure
Ar to 5.2 V in 5 ppm NO2; moreover they established that in the linear regime a linear relationship
between conductance and NO2 partial pressure exists. In addition they demonstrated another impor-

Ó American Chemical Society 2004

Fig. 47. (a) Channel conductivity of a In2O3 based SNT as a function of VGS at growing concentration (left to right) of NO2 (b) IDS–
VDS curves at VGS = 0 (reprinted with permission from Ref. [225]).
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 49

Ó American Institute of Physics 2004


Fig. 48. ZnO SNT sensing response at RT to 10 ppm NO2 in Ar and the conductance recovery process caused by a 60 V gate
voltage pulse (reprinted with permission from Ref. [228]).

tant usage of gate voltage, by employing a strong negative VGS = 60 V field to refresh the ZnO SNT
inducing NO2 electrodesorption, as showed in Fig. 48.
As known, NO2 adsorption at room temperature is hardly reversible because the thermal energy is
usually lower than the activation energy for desorption. A high gate voltage induces an electric field
perpendicular to the surface: hole migration to the surface driven by the negative field leads to recom-
bination and discharge of NO 2 . In addition, it is likely that the negative gate induced repulsive field
stretches and weakens the bonding between NO2 surface dipoles and adsorption sites. Conclusively,
these factors lead to the electrodesorption process and the fast conductance recovery at room
temperature.
Fan et al. [228] found that VT increases as a function of O2 concentration in Ar for a ZnO based SNT,
as expected since O2 – like NO2 – ionosorbs as O 
2 and O over metal oxide surfaces. In addition they
found that sensor response is greater when VGS is close to VT: when the VGS  VT, electron concentra-
tion in the channel is quite high and surface adsorbed O2 molecules capture only a small portion of the
available electrons. Therefore, the relative conductance change is very small. However, when VGS is
close to VT, channel electrons are substantially depleted and the conductance change caused by O2
adsorption becomes much more significant.
Zhou et al. observed an intriguing interplay between the surface treatment and the nanowire dop-
ing concentration in presence of NH3 [223,224], a species which is known to inject electrons in con-
duction band when adsorbed. For In2O3 nanowires with surfaces cleaned via UV illumination in a
vacuum, VT has been found to decrease for lightly doped nanowires and to increase for heavily doped
nanowires upon exposure. On the contrary in ambient atmosphere VT has been consistently observed
to decrease upon NH3 exposure, regardless of the doping concentration.
The group of Moskovits carried out a detailed study of the effects of the gate voltage in a SnO2
based SNT following O2 introduction in N2 and subsequent CO detection [226,14,227] at 553 K, con-
firming that the rates and extent of oxidation and reduction reactions taking place at the surface of
a SnO2 nanowire, configured as a field-effect transistor, can be modified by changing the electron den-
sity in the wire with a gate voltage.
As shown in Fig. 49, at VGS < VT = 6.2 V in N2 the conductive channel is not present and the device
is almost insulating. When VGS is made progressively more positive, the conductance increases and
then achieves an approximately constant value as expected. The conductance decreases greatly when
O2 is introduced, however the amount of this reduction is lower at decreasing VGS. Almost immedi-
ately upon the introduction of CO, the conductance increases, eventually achieving a new steady state.
Indeed the observed conductance increase on admitting CO is not monotonic with VGS but achieves a
maximum value in the range VGS = 2 to 0 V. Interestingly, even at VGS = 6 V, when oxygen ionosorp-
tion no longer takes place, the introduction of CO causes the conductance to increase.
50 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ó American Chemical Society 2004


Fig. 49. IDS kinetic response (VDS = 0.2 V) at 553 K of a SnO2 based SNT to the addition of 10 sccm of O2 to 100 sccm of flowing N2
gas at time t1 followed by the addition of 5 sccm of CO at time t2 at various values of the gate potentials (reprint permission
from Ref. [226]).

By considering oxygen surface ionosorption, CO oxidation through ionosorbed oxygen, surface lat-
tice oxygen and hydroxyl groups, and CO direct ionosorption as a donor, Moskovits group was able to
estimate surface coverage with ionosorbed oxygen h – that is the fraction of nanowires surface cov-
ered with ionosorbed oxygen – as a function of VGS as showed in Fig. 50.
The most striking evidence from Fig. 50 is that the rate and extent of oxygen ionosorption and the
resulting rate and extent of catalytic CO oxidation reaction on the nanowire’s surface could be con-
trolled and even entirely halted by applying a negative enough gate potential.
In conclusion, it has been demonstrated by many independent studies how the additional gate volt-
age tool, which is introduced by configuring a single nanowire in a field-effect configuration, allows
tuning the availability of electrons for surface adsorption and desorption reactions and the sensing
and recovering performance of the devices.

6.1.6. PL based gas sensors


Nanowires of ZnO and SnO2 show photoluminescence (PL) spectra at room temperature. A recent
work by Faglia et al. [231] evidenced that SnO2 nanowires exhibit interesting gas-sensitive photolu-
minescence properties: it was found that exposure to low concentrations of nitrogen dioxide (few
ppm concentrations) significantly quenched the visible PL emission of nanowires, which was more-
over not affected by variation of relative humidity, and by other interfering gases. The challenge to
realize a reliable contact on nanowires for conductometric gas sensor, increases the interest in an
all-optical gas sensor, where contacts are not needed. Moreover the possibility to use the sensor at
room temperature is interesting when dealing with gas detection in explosive environment. Since
PL is different for different oxides a separate paragraph will be dedicated to each material.

6.1.6.1. SnO2 based PL sensors. SnO2 is a direct wide-gap semiconductor (3.6 eV). Bulk oxygen vacan-
cies are responsible for two donor states localized at 30 meV and 150 meV from the bottom of the con-
duction band [167].
Nanostructured SnO2 typically exhibits strong PL emission in the visible range from 400 to 600 nm
[232–237]. Analyzed nanostructures exhibited high degree of crystallinity, and broad PL emission
peaked in the green1 or yellow range (1.9–2.5 eV) was reported in Fig. 51. Such emission band cannot
be assigned to band-edge or to exciton recombination once one considers the wide bandgap energy of

1
For interpretation of color in Figs. 1, 2, 5, 6, 14, 16, 18, 19, 20, 22, 24-27, 29, 33, 34, 38, 39, 41, 43-47, 53, 60 and 61 the reader is
referred to the web version of this article.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 51

Ó American Chemical Society 2005


Fig. 50. The relative steady state ionosorbed oxygen coverage calculated from the observed conductance changes as a function of
gate potential before (squares) and after (triangles) CO gas is admitted (reprint permission from Ref. [227]).

Ó Wiley-VCH Verlag GmbH & Co.KGaA 2006

Fig. 51. Typical PL spectra of SnO2 nanorods (black) and nanorods of uniform diameter (grey). Reproduced with permission
from [233].

SnO2. Thus, transitions involving defective states within the bandgap have to be invoked. The key role
played by oxygen vacancy states in such PL emission was highlighted in all cases on which the oxygen
defectivity of a sample was intentionally altered by means of annealing under vacuum and/or under
oxygen flow: it is found as a general trend that the PL efficiency is reduced (or eventually almost sup-
pressed) once oxygen vacancies are removed [237,235]. Thus, the role of radiative states for the visible
PL emission is commonly attributed to OV states.
Faglia et al. studied PL quenching of SnO2 nanowires deposited on Si substrate caused by NO2
impurities diluted in dry air [231]. NO2 at the ppm level strongly quenches the spectral intensity of
the PL emission band by adsorbing over the surface and creating competitive non-radiative recombi-
nation paths. They examined also the kinetics of the involved processes, taking the PL spectrum every
5 s as the gas fluxes into test chamber; the area under the broad PL band peak was extracted as a fea-
ture for every sampling. Fig. 52 reports the dynamic response of the peak area toward 1 ppm of NO2 in
dry air, keeping relative humidity constant at 0%, 70% and 30% at room temperature; the sensor was
52 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Fig. 52. Response of optical sensor based on PL quenching of SnO2 nanowires. Area under the PL peak was plotted as a function
of time, after the introduction of 1 ppm of NO2. Humidity was kept constant at 0%, 70% and 30% during three pulses of NO2,
giving no influence on sensor response. The sample was kept at 120 °C. After Ref. [231], permission 1905361485263.

kept at T = 120 °C. Compared to electrical results [179] performed on the sample produced through the
same technique, the dynamic is much faster: response time to NO2 needed to reach 90% of the steady
state step response was equal to about 30 s (RH = 70% 20 °C), while the conductivity one was of the
order minutes. The adsorption process is reversible: 90% of PL baseline is recovered after restoring
air in about 600 s.
Therefore, a very important feature is that NO2 can be detected at low concentration with almost
no interference from water vapor. On the contrary, it was well known that humidity adsorption over
metal oxides produces an hydroxylated surface and modifies electrical properties by acting as an elec-
tron donor. Besides, the authors investigated the effect on PL of NH3 (50 ppm) and CO (1000 ppm)
without observing any appreciable quenching effect. The same gaseous species are active in affecting
electrical transport.
Lettieri et al. [238] investigated the gas adsorption induced PL quenching in SnO2 nanowires,
mainly focusing on the physical mechanisms controlling the phenomenon by time-resolved (TR)
and continuous wave (CW) PL experiments performed in controlled nitrogen dioxide (NO2)
environment.
While the CWPL characterization gives phenomenological information on the overall material,
more details on the light emission properties can be obtained by TRPL technique, which allows the
direct observation of the electronic excited states lifetime and, therefore, to obtain useful information
on the possible creation of non-radiative paths due to adsorbate–nanostructure interaction.
The authors observed overall PL signal reduction of the order of about 5% due to the NO2 (5 ppm)
introduction. In Fig. 53, effective decay time (defined as the time at which the PL signal is at 10% of its
initial value) is plotted versus the emission wavelength for the two sets of measurements (air and
NO2). It was observed that the measured decay times were essentially the same within the experimen-
tal resolution. Assuming that radiative recombination of surface electrons was responsible for green
emission also in SnO2 nanobelts, the results of Fig. 53 suggested that the adsorbed NO2 reduced the
overall number of radiative recombinations but did not affect the decay rates, at least at the nanosec-
ond scale.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 53

Fig. 53. Photoluminescence decay time (defined as 10% decay time) as a function of emission wavelength for air (open squares)
and 5 ppm NO2 (open circles). Sample was kept at T = 120 °C. Permission number 1901830530722, [238].

Since decay times do not depend on the quencher presence, the quenching is said to be ‘‘static” and
time dynamics are not influenced by the quencher presence.

6.1.6.2. ZnO based PL sensors. In ZnO oxygen vacancies are responsible for doping of the high gap metal
oxide seminconductor that has high exciton binding energy (60 meV). The interest in ZnO PL was
mainly driven by the existence of room temperature UV peak and to lasing activity of ZnO [109].
UV peak is attributed to free exciton emission. However some variation of the position of the PL
peak was observed for different nanostructures (see Fig. 54 [239]).
ZnO nanowires also show PL spectrum in the visible range, with broad peak depending strongly on
the preparation conditions in the range from green to yellow [240,241] to blue. While blue emission is
attributed to amine ligand coming from organometallic nanoparticles preparation [242], concerning to
PL response in the visible-green range, still, there is no consensus on the positions of the peaks of ZnO
nanostructures and thin film, and their origin. The green emission has been assigned to the transition
between the photoexcited holes and a single ionized oxygen vacancy [243] attributed to an antisite
oxygen [244] and donor–acceptor complexes [245,246]. Surface states have also been identified as
a possible cause of the visible emission in ZnO nanowires [247].
Ó Wiley-VCH Verlag GmbH & Co.KGaA 2006

Fig. 54. Room-temperature PL spectra of various nanostructures un UV range: (1) tetrapods, (2) needles, (3) nanorods, (4)
shells, (5) highly faceted rods, (6) ribbon/combs. Ref. [239]. Reproduced with permission.
54 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

The ratio between UV and visible emission, which is sometimes used to estimate sample quality for
optoelectronic application [248], is dependent on fabrication conditions and on excitation area and
power [249]. Djurisic et al. reported that different fabrication methods resulted in needles, shells
and rods of ZnO with different defect types and concentrations resulting in different luminescence
spectra. The authors say that it was reasonable to assume that they were not dependent on morphol-
ogy but rather on dominant defect type [250].
A very interesting result, pointed out by Shalish et al. [251], is the progressive increase of the green-
light emission intensity determined as the diameter of the wires decreases, by the increase of the sur-
face, rich in oxygen vacancies, to volume ratio (see Fig. 55). The ratio between the PL contributions at
below-bandgap and at band-edge was adequately fitted by the authors, as a function of the wire ra-
dius, assuming a surface layer wherein the surface-recombination probability is 1.
Moreover, annealing treatments in different atmosphere influence the emission in ZnO nanorods
[252].
Lettieri et al. [238] studied PL spectrum of ZnO nanowires in air and in NO2 environment. The first
observation regards PL spectra of ZnO as a function of temperature collected in pure synthetic air.
They have noticed that PL quenching occurs as the temperature increases, as expected if thermally-
activated non-radiative recombination paths are introduced.
A significant quenching of the green PL was observed after introduction of 10 ppm NO2 diluted in
dry air, similar to the temperature-induced quenching. Such results suggest that the origin of PL green
quenching lied in the introduction of non-radiative recombination paths for surface electrons, which
competed with the radiative surface recombination ones.
Differently from the case of SnO2, here the higher quenching was recorded at room temperature.
The authors have performed time-resolved photoluminescence measurement to gain more detail
on the possible mechanism responsible for quenching. Even at room temperature nitrogen dioxide
molecules acted as a strong quencher for ZnO but, at the microscopic scale, their interaction with
the metal oxide nanostructures seemed to take place on a time scale of less than 1 ns (that is beyond
the resolution of instrumentation used). While the experimental findings did not allow identifying the
microscopic origins of the PL green emission, the results were consistent with the hypothesis that it
was related to radiative recombination of surface-localized electrons, while PL quenching was sup-
posed to be due to charge capture from gas adsorbed onto the surface.
Comini et al. [253] studied the dynamic of PL quenching of ZnO induced by NO2: after a 1-hour sta-
bilization period in dry air, 16 ppm of NO2 in dry air are introduced into the test chamber. Then, after

Fig. 55. Photoluminescence spectra obtained from ZnO wires of three different sizes (shown above in SEM micrographs).
Reprinted figure with permission from [251].
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 55

another hour, dry air flow was restored and maintained for almost 1 h. In order to study the dynamic
behavior of PL quenching the spectra were taken every 5 s and they calculated the area under the
broad PL band and plotted its variation as a function of time. The visible and UV bands gave the same
quenching behavior. Fig. 56 reports the quenching behavior with 16 ppm of NO2. The response time
was less than 30 s, while recovery time was longer. The presence of nitrogen dioxide on the nanowire
surface influenced recombination processes, destroying radiative recombination paths.
Fast (of the order 10 min) desorption of NO2 from surface at room temperature was possible due to
energy given by UV light. Since NO2 should adsorb over the surface creating a new surface state NO 2,
like oxygen do, the mechanism with light should be the same as reported for O2 on ZnO surface when
illuminated with UV light [254,255].
When the sample is irradiated by light near the band gap, hole-electron pairs are formed. The hole
formed at the surfaces find itself in an electric field which pulls the hole towards the surface where it
recombines with a NO2 ion to form a neutral physically absorbed atom which is in equilibrium with
NO2 in the ambient. When NO2 is removed from the test chamber and replaced by air, NO2 tends to
desorb.
The amount of quenching follows NO2 concentrations, as reported in Fig. 57. The test ranging from
1 to 16 ppm diluted in dry air shows that the IAIR =INO2 follows a Stern Volmer model,
IAIR kNO2
¼1þ ½NO2 
INO2 kRD;AIR þ kNRD;AIR
where kRD,AIR is the radiative decay rate constant in air; kNRD,AIR is the non radiative decay rate constant
in air, [NO2] is the partial pressure of NO2, kNO2 is the non-radiative decay rate constant introduced by
NO2; IAIR is the PL emission in air (calculated as the area under the peak); INO2 is the PL emission in NO2.
Comini et al. [253] have also studied the effect of other gases like ethanol, relative humidity and CO
on PL quenching. No response was observed for CO, while a quenching of 4% was observed changing
from dry air to 30% relative humidity and an increase of 7% was observed for 900 ppm of ethanol.
Therefore they stated that the optical sensor seemed quite insensitive to common interfering gases
for NO2 detection, like CO, even if there was still a dependence of optical signal from humidity and
ethanol that must be taken into account. The effect of humidity should be ascribed to room temper-
ature operation of the sensor. This strongly suggests the application of this property to the develop-
ment of an optical gas sensor working at RT, with no need to create low reliability electrical
contacts on nanowires.

Fig. 56. Dynamic of PL quenching on ZnO nanowires caused by NO2 introduction. A fast reversible quenching is observed at
room temperature. Reprint after permission from [253].
56 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

1.4
T = 20°C
dry air

Peak Area0/Peak Area


1.2

1.0
0 2 4 6 8 10 12 14 16 18
NO2 (ppm)

Fig. 57. Dependence of IAIR =INO2 from NO2 concentration.

6.2. Other application fields


Metal oxide nanowires are being successfully used to a variety of application, leading to increased
interest for investigation of their functional properties.
Among all, we report some cutting edge technological applications, in which metal oxide nano-
wires can lead to breakthrough with respect to the state-of-the-art devices.

6.2.1. Lasers
The lasing properties of ZnO dendritic nanowires have been successfully exploited [256].
Under optical excitation, an ultraviolet laser array was demonstrated for these comb structures (see
Fig. 58).
The interest in ZnO nanostructures has been recently increased due to the new applications in the
room-temperature nanolasing in the ultraviolet range [109]. The lasing dynamics of single ZnO nano-
wires have been also studied as a function of the ultraviolet excitation intensity in Ref. [257].

Ó American Chemical Society 2003

Fig. 58. Left: Comb structures made of ZnO nanowires. Right: (A) Far-field optical image of spatially resolved light emission
from individual nanowires of the comb structures. The spacing between the wires is ca. 1 lm. (B) Power-dependent emission
spectra recorded on the comb structure. From bottom-up, the excitation energy densities are 252, 505, 580, 707, 883 nJ/cm2,
respectively. Lasing behavior is evidenced by the appearance of sharp cavity modes in the top two traces. Reprinted with
permission from [256].
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 57

6.2.2. Solar cells


Application of MOX nanowires in the field of solar cells is experiencing growing interest in the last
years.
The first results on a hybrid solar cell composed by CdSe nanorods and a conjugated polymer were
obtained by Alivisatos and co-workers [258]. In this case the band gap of CdSe nanorods was modified
by altering the nanorod radius to optimize the overlap between the absorption spectrum of the cell
and the solar emission spectrum.
After the first work, MOX nanowires (mainly ZnO) have been applied as promising candidates for
electron transport in DSC [259–261].
The main advantage of single-crystalline nanowire, with respect to traditional polycrystalline net-
work acting as electron transducer in the traditional Graetzel cells [262] is the increased diffusion
length in the anode by replacing the nanoparticle film with an array of oriented single-crystalline
nanowires [260]. One further advantage is the possibility of precisely tailoring of the electronic prop-
erties of the nanowires and nanorods for tuning the absorption spectrum of the hybrid system dye/
nanowires with the solar emission spectrum.
Innovative geometries and structures are being continuously developed, such as fully-inorganic cell
composed by ZnO nanowires sensitized with CdSe quantum dots acting as light absorbers [263].
Core-shell heterostructures have been applied for DSC [160,161]. Better electron transport in such
structures is a product of their single crystalline assembly and of the presence of a radial electric field
within each nanowire. This field assists carrier collection by repelling the photoinjected electrons from
the surrounding electrolyte. Moreover, application of branched nanowires allows one-dimensional
electronic transport, while electrodes can be configured in complex geometries in order to increase
the energy conversion density of the cell.

6.2.3. Field emitters


Quasi 1D nanostructured zinc oxide was the first and most extensively tested material for field
emission studies [264–269].
The field emission properties of well-aligned ZnO nanowires are reported in Fig. 59 from Ref. [264].
Nanowires [264,265], nanoneedles [266], nanopyramids [267], nanorods [268], and screw-shape
InZnO nanorod arrays [269] have been successfully tested as field emitters, as well as other oxides:
tungsten oxide nanowires [270],SnO2 nanowires doped with RuO2 [271] or Sb [272], TiO2 nanotubes
[273], ZnS nanobelts [274].
Ó American Institute of Physics 2002

Fig. 59. Emission current density from ZnO nanowires grown on silicon substrate at 550 °C. The inset reveals that the field
emission follows FN behavior. Reprinted with permission from Ref. [264].
58 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

Ó American Chemical Society 2007


Fig. 60. Room-temperature I–V characteristics of an individual SnO2–In2O3 and SnO2 nanowire. Reprinted in part from Ref.
[159].

Ó American Chemical Society 2006

Fig. 61. Real time selective detection of DNA molecules. Reprinted from Ref. [277].
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 59

6.2.4. Li-ion batteries


Lithium-alloy anodes for rechargeable ambient temperature lithium batteries have been studied
since the early 1970s. Nanosystems are potentially useful as electrodes for Li-ion batteries, because
the small size of these structures is usually correlated with their high surface area and large sur-
face-to-volume ratio, which favors the physical or chemical interactions of the electrodes and lithium
ions. The benefits of enhanced active sites and short diffusion distances, which arise from 1D nano-
structures, will thus result in a high capacity and fast kinetics.
In2O3-SnO2 heterostructures have been exploited for application in Li-ion batteries [159]. The elec-
tronic conductivity of the individual SnO2-In2O3 nanowires during cycling was 2 orders of magnitude
better than that of the pure SnO2 nanowires (Fig. 60), due to the formation of Sn-doped In2O3 caused
by the incorporation of Sn into the In2O3 lattice during the nucleation and growth of the In2O3 shell
nanostructures. This effect could improve significantly the electrochemical performance of the
battery.

6.2.5. Single nanowire transistors for biosensing


More recent applications of ZnO nanowires have been proposed in the field of electricity generation
from mechanical energy [275]. Starting from the work of Lieber et al., the fabrication of hybrid inor-
ganic–organic systems for ultrasensitive biomolecule sensors has been extended to metal-oxide nano-
structures [276,277]. The fast detection of DNA molecules has been demonstrated, as shown in Fig. 61;
the perspective of highly-sensitive detection of single molecules is also discussed.

7. Conclusions

This article reviewed the research on Q1D nanostructures of metal oxide for sensing application,
although it could not cover all the interesting works reported due to the great research effort in the
field: the authors were selecting the ones that, according to their opinion, are going to contribute to
this field’s further development.
The fundamental properties of nanosized materials have been studied over the last years with par-
ticular focus on the possibility to exploit the preparation of new devices such as gas sensors, biosen-
sors, FET, LED, lasers, solar cells, field emitters, Li-ion batteries.
A great effort has been done to understand and control the growth process for the production of
high quality quasi-one dimensional nanostructures with bottom-up techniques.
The starting point was the VLS mechanism that is still widely employed for growing most of the
nanometric wires produced nowadays. In addition, various new techniques have been reported for
the preparation of high quality nanowires and to control the alignment of these structures towards
device integration.
Solution-based techniques can be a promising alternative approach for mass production of metal
oxide nanosized materials with good control of shape, composition and reproducibility, but its main
limitation is the lack in full understanding of growth mechanisms. Most of the methods are based
on trial-and-error procedure leading to difficulties in prediction and engineering of NWs production.
Last but not least, not all the solution-based techniques can produce single-crystal nanowires, neces-
sary to exploit peculiar properties and advantages of nanowires in functional devices.
During the last years the understanding of quasi-one dimensional structures has grown thanks to
the experimental and theoretic studies reported in literature, leading to a higher control in the produc-
tion (composition, crystallinity, orientation, alignment) and reproducibility, but a great effort is still
necessary for the comprehension of some morphologies and processes. One key issue, still not fulfilled,
is the growth of p-type metal oxide nanowires (ZnO for example) interesting for fabrication of p–n
junctions.
Concerning the gas sensor applications, MOX NWs demonstrate improved sensitivity to various
gases, even if selectivity remains a concern. Beside the additional gate voltage, which is introduced
by configuring a simple nanowire in a field-effect configuration, increases the sensing and recovery
performances of the device.
A possible way to differentiate the NWs response may be surface coating with chemical selective
membrane, surface modification by specific functional groups, or combining multi-component sensing
60 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

modules coupled with signal processing functions, acting as an ‘‘electronic nose” to differentiate min-
ute quantities in a complex environment.
The main gain with MOX NWs is their improved stability thanks to their high crystallinity degree.
An open problem, which deserve particular attention, remains the need of reliable and stable electrical
contacts that can guarantee the preparation of stable devices with single crystalline structures.

Acknowledgements

The authors want to thank first of all the member of the SENSOR lab group in Brescia, and further-
more the European Commission that partially funded the research in the nanostructures field for the
author’s institution with the project NMP4-CT-2003-001528 ‘‘Nanostructured solid-state gas sensors
with superior performance” (NANOS4). This work was partially supported, within the EU FP6, by the
ERANET project ‘‘NanoSci-ERA: NanoScience in the European Research Area”.

References

[1] Gopel W, Hesse J, Zemel JN. Sensors: a comprehensive survey. New York: VCH; 1995. p. 1–7.
[2] Moseley PT, Tofield BC, editors. Solid-state gas sensors. Bristol/Philadelphia: Hilger; 1987.
[3] Sberveglieri G, editor. Gas sensors: principles, operation and developments. Boston: Kluwer; 1992.
[4] Mandelis A, Christofides C. Physics, chemistry and technology of solid state gas sensor devices. New York: Wiley-
Interscience; 1993.
[5] Appell D. Nanotechnology: wired for success. Nature 2002;419:553.
[6] Samuelson L. Self-forming nanoscale devices Mater. Today 2003;6:22.
[7] Duan XF, Huang Y, Cui Y, Wang JF, Lieber CM. Indium phosphide nanowires as building blocks for nanoscale electronic and
optoelectronic devices. Nature 2001;409:66.
[8] Cui Y, Wei Q, Park H, Lieber CM. Nanowire nanosensors for highly sensitive and selective detection of biological and
chemical species. Science 2001;293:1289.
[9] Xia YN, Yang PD, Sun YG, Wu YY, Mayers B, Gates B, et al. One-dimensional nanostructures: synthesis, characterization,
and applications. Adv Mater 2003;15:353.
[10] Yamazoe N. New approaches for improving semiconductor gas sensors. Sensors Actuat B 1991;5:7.
[11] Wang ZL. Characterizing the structure and properties of individual wire-like nanoentities. Adv Mater 2000;12:1295.
[12] Wang N, Cai Y, Zhang RQ. Growth of nanowires. Mater Sci Eng R 2008;60:1–51.
[13] Lu JG, Chang P, Fan Z. Quasi-one-dimensional metal oxide materials—synthesis, properties and applications. Mater Sci Eng
R 2006;52:49–91.
[14] Kolmakov A, Moskovits M. Chemical sensing and catalysis by one-dimensional metal-oxide nanostructures. Annu Rev
Mater Res 2004;34:151–80.
[15] Lieber CM, Wang ZL. Nanowires as building blocks for bottom-up nanotechnology. MRS Bull 2007;32:99–104.
[16] Haghiri-Gosnet AM, Vieu C, Simon G, Mejıas M, Carcenac F, Launois H. Nanofabrication at a 10 nm length scale: limits of
lift-off and electroplating transfer processes. J Phys 1999;IV(9 Pr2):133–41.
[17] Marrian CRK, Tennant DM. Nanofabrication. J Vac Sci Technol A 2003;21:S207–15.
[18] Candeloro P, Comini E, Baratto C, Faglia G, Sberveglieri G, Kumar R, et al. SnO2 lithographic processing for nanopatterned
gas sensors. J Vac Sci Technol B 2005;23:2784–8.
[19] Candeloro P, Carpentiero A, Cabrini S, Di Fabrizio E, Comini E, Baratto C, et al. SnO2 sub-micron wires for gas sensors.
Micro Eng 2005;78-79:178–84.
[20] Hanrath T, Korgel BA. Nucleation and growth of germanium nanowires seeded by organic monolayer-coated gold
nanocrystals. J Am Chem Soc B 2002;124:1424–9.
[21] Wagner RS, Ellis WC. Vapor–liquid–solid mechanism of single crystal growth. Appl Phys Lett 1964;4:89–90.
[22] Kolasinski KW. Catalytic growth of nanowires: vapor–liquid–solid, vapor–solid–solid, solution–liquid–solid and solid–
liquid–solid growth. Curr Opin Solid State Mater Sci 2006;10:182–91.
[23] Kolasinski KW. Surface science: foundations of catalysis and nanoscience. Chichester: John Wiley and Sons; 2002.
[24] Kolasinski KW. Growth and etching of semiconductors. In: Hasselbrink E, Lundqvist I, editors. Handbook of surface
science, vol. 3. Amsterdam: Elsevier, 2008 (in press).
[25] Roper SM, Davis SH, Norris SA, Golovin AA, Voorhees PW, Weiss M. Steady growth of nanowires via the vapor–liquid–
solid method. J App Phys 2007;102:034304.
[26] Givargizov E. Fundamental aspects of VLS growth. J Cryst Growth 1975;31:20–30.
[27] Chen Z, Cao CB. Effect of size in nanowires grown by the vapor–liquid–solid mechanism. Appl Phys Lett 2006;88:143118.
[28] Johansson J, Svensson CPT, Martensson T, Samuelson L, Seifert W. Mass transport model for semiconductor nanowire
growth. J Phys Chem B 2005;109:13567–71.
[29] Wang X, Song J, Summers CJ, Ryou JH, Li P, Dupuis RD, et al. Density-controlled growth of aligned ZnO nanowires sharing a
common contact: a simple, low-cost, and mask-free technique for large-scale applications. J Phys Chem B
2006;110:7720–4.
[30] Mathur S, Barth S, Shen H, Pyun JC, Werner U. Size-dependent photoconductance in SnO2 nanowires. Small 2005;1:713–7.
[31] Mathur S, Barth S. One-dimensional semiconductor nanostructures: growth, characterization and device applications. Z
Phys Chem 2008;222:307–17.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 61

[32] Zhu Z, Chen TL, Gu Y, Warren J, Osgood RM. Zinc oxide nanowires grown by vapor-phase transport using selected metal
catalysts: a comparative study. J Chem Mater 2005;17:4227–34.
[33] Kodambaka S, Tersoff J, Reuter MC, Ross FM. Diameter-independent kinetics in the vapor–liquid–solid growth of Si
nanowires. Phys Rev Lett 2006;96:096105.
[34] Vomiero A, Bianchi S, Comini E, Faglia G, Ferroni M, Sberveglieri G. Controlled growth and sensing properties of In2O3
nanowires. Cryst Growth Design 2007;7:2500–4.
[35] Wang X, Ding Y, Summers CJ, Wang ZL. Large-Scale synthesis of six-nanometer-wide ZnO nanobelts. J Phys Chem B
2004;108:8773–7.
[36] Zhi CY, Bai XD, Wang EG. Synthesis of semiconductor nanowires by annealing. Appl Phys Lett 2004;85:1802–4.
[37] Xu CH, Woo CH, Shi SQ. Formation of CuO nanowires on Cu foil. Chem Phys Lett 2004;399:62–6.
[38] Zhou J, Gong L, Deng SZ, Chen J, She JC, Sheng N, et al. Growth and field-emission property of tungsten oxide nanotip
arrays. Appl Phys Lett 2005;87:223108.
[39] Wagner RS. In: Levitt AP, editor. Whisker technology. New York: Wiley; 1970 (Chapter 3).
[40] Givargizov EI. In: Kaldis E, editor. Current topics in materials science, vol. 1. Amsterdam, The Netherlands: North-Holland;
1978 (Chapter 3).
[41] Despic A, Parkhutik VP. In: Bockris JO’M, White RE, Conway BE, editors. Modern aspect of electrochemistry, vol.
20. NY: Plenum Press; 1989 (Chapter 6).
[42] Foss CA, Hornyak GL, Stockert, Martin CR. Optical-properties of composite membranes containing arrays of nanoscopic
gold cylinders. J Phys Chem 1992;96:7497–9.
[43] Cepak VM, Hulteen JC, Che G, Jirage KBJA, Lakshmi BB, Fisher ER, et al. Chemical strategies for template syntheses of
composite micro- and nanostructures. Chem Mater 1997;9:1065.
[44] Jessensky O, Muller F, Gosele U. Self-organized formation of hexagonal pore arrays in anodic alumina. Appl Phys Lett
1998;72:1173–5.
[45] Hulteen JC, Martin CR. A general template-based method for the preparation of nanomaterials. J Mater Chem
1997;7:1075–87.
[46] Mor GK, Varghese OK, Paulose M, Mukherjee N, Grimes CA. Fabrication of tapered, conical-shaped titania nanotubes. J
Mater Res 2003;8:2588–93.
[47] Masuda H, Fukuda K. Ordered metal nanohole arrays made by a 2-step replication of honeycomb structures of anodic
alumina. Science 1995;268:1466–8.
[48] Martin CR. Membrane-based synthesis of nanomaterials. Chem Mater 1996;8:1739–46.
[49] Muller F, Birner A, Schilling J, Li AP, Nielsch K, Gosele U, et al. High aspect ratio microstructures based on anisotropic
porous materials. Microsyst Technol 2002;8:7–9.
[50] Choi J, Nielsch K, Reiche M, Wehrspohn RB, Gosele U. Fabrication of monodomain alumina pore arrays with an interpore
distance smaller than the lattice constant of the imprint stamp. J Vac Sci Technol B 2003;21:763–6.
[51] Masuda H, Abe A, Nakao M, Yokoo A, Tamamura T, Nishio K. Ordered mosaic nanocomposites in anodic porous alumina.
Adv Mater 2003;15:161.
[52] Asoh H, Ono S, Hirose T, Nakao M, Masuda H. Growth of anodic porous alumina with square cells. Electrochim Acta
2003;48:3171–4.
[53] Choi J, Wehrspohn R, Gosele U. Moire pattern formation on porous alumina arrays using nanoimprint lithography. Adv
Mater 2003;15:1531.
[54] Teng F, Liang S, Gaugeu B, Zong R, Yao W, Zhu Y. Carbon nanotubes-templated assembly of LaCoO3 nanowires at low
temperatures and its excellent catalytic properties for CO oxidation. Catal Commun 2007;8:1748–54.
[55] Xu CK, Xu GD, Liu YK, Wang GH. A simple and novel route for the preparation of ZnO nanorods. Solid State Commun
2002;122:175–9.
[56] Xu CK, Zhao XL, Liu S, Wang GH. Large-scale synthesis of rutile SnO2 nanorods. Solid State Commun 2003;125:301–4.
[57] Xu CK, Xu GD, Wang GH. Preparation and characterization of NiO nanorods by thermal decomposition of NiC2O4
precursor. J Mater Sci 2003;38:779–82.
[58] Gao T, Li QH, Wang TH. Sonochemical synthesis, optical properties, and electrical properties of core/shell-type ZnO
nanorod/CdS nanoparticle composites. Chem Mater 2005;17:887–92.
[59] Miao JJ, Wang H, Li YR, Zhu JM, Zhu JJ. Ultrasonic-induced synthesis of CeO2 nanotubes. J Cryst Growth 2005;281:525–9.
[60] Kumar RV, Koltypin Y, Xu XN, Yeshurun Y, Gedanken A, Felner I. Fabrication of magnetite nanorods by ultrasound
irradiation. J Appl Phys 2001;89:6324–8.
[61] Liu B, Zeng HC. Hydrothermal synthesis of ZnO nanorods in the diameter regime of 50 nm. J Am Chem Soc
2003;125:4430–1.
[62] Wang JM, Gao L. Wet chemical synthesis of ultralong and straight single-crystalline ZnO nanowires and their excellent UV
emission properties. J Mater Chem 2003;13:2551–4.
[63] Guo M, Diao P, Cai SM. Hydrothermal growth of well-aligned ZnO nanorod arrays: dependence of morphology and
alignment ordering upon preparing conditions. J Solid State Chem 2005;178:1864–73.
[64] Sun Y, Ndifor-Angwafor NG, Riley DJ, Ashfold MNR. Synthesis and photoluminescence of ultra-thin ZnO nanowire/
nanotube arrays formed by hydrothermal growth. Chem Phys Lett 2006;431:352–7.
[65] Cao MH, Wang YH, Guo CX, Qi YJ, Hu CW, Wang EB. A simple route towards CuO nanowires and nanorods. J Nanosci
Nanotechnol 2004;4:824–8.
[66] Zhou KB, Wang X, Sun XM, Peng Q, Li YD. Enhanced catalytic activity of ceria nanorods from well-defined reactive crystal
planes. J Catal 2005;229:206–12.
[67] Yuan ZY, Su BL. Titanium oxide nanotubes, nanofibers and nanowires. Colloids Surf A Physicochem Eng Aspects
2004;241:173–83.
[68] Yuan ZY, Ren TZ, Du G, Su BL. A facile preparation of single-crystalline a-Mn2O3 nanorods by ammonia-hydrothermal
treatment of MnO2. Chem Phys Lett 2004;389:83–6.
[69] Tan Y, Xue X, Peng Q, Zhao H, Wang T, Li Y. Controllable fabrication and electrical performance of single crystalline Cu2O
nanowires with high aspect ratio. Nano Lett 2007;7:3723–8.
62 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

[70] Sun Y, Fox NA, Riley DJ, Ashfold MNR. Hydrothermal growth of ZnO nanorods aligned parallel to the substrate surface. J
Phys Chem C. CCC American Chemical Society Published on Web 05/30/2008. doi:10.1021/jp8019107.
[71] Kahn M, Monge M, Snoeck E, Maisonnat A, Chaudret B. Spontaneous formation of ordered 2D and 3D superlattices of ZnO
nanocrystals. Small 2005;1:221–4.
[72] Kahn ML, Monge M, Colliere V, Senocq F, Maisonnat A, Chaudret B. Size- and shape-control of crystalline zinc oxide
nanoparticles: a new organometallic synthetic method. Adv Funct Mater 2005;15:458–68.
[73] Formalas A. US patent 1 975 504; 1934.
[74] Dai H, Gong J, Kim H, Lee D. A novel method for preparing ultra-fine alumina–borate oxide fibres via an electrospinning
technique. Nanotechnology 2002;13:674–7.
[75] Viswanathamurthi P, Bhattarai N, Kim HY, Lee DR, Kim SR, Morris MA. Preparation and morphology of niobium oxide
fibres by electrospinning. Chem Phys Lett 2003;374:79–84.
[76] Shao C, Kim HY, Gong J, Ding B, Lee DR, Park SJ. Fiber mats of poly(vinyl alcohol)/silica composite via electrospinning.
Mater Lett 2003;57:1579–84.
[77] Guan H, Shao C, Chen B, Gong J, Yang X. A novel method for making CuO superfine fibres via an electrospinning technique.
Inorg Chem Commun 2003;6:1409–11.
[78] Guan H, Shao C, Chen B, Gong J, Yang X. Preparation and characterization of NiO nanofibres via an electrospinning
technique. Inorg Chem Commun 2003;6:1302–3.
[79] Yang X, Shao C, Guan H, Li X, Gong J. Preparation and characterization of ZnO nanofibers by using electrlospun PVA/zinc
acetate composite fiber as precursor. Inorg Chem Commun 2004;7:176–8.
[80] Ding B, Kim H, Kim C, Khil M, Park S. Morphology and crystalline phase study of electrospun TiO2–SiO2 nanofibres.
Nanotechnology 2003;14:532–7.
[81] Viswanathamurthi P, Bhattarai N, Kim HY, Lee DR. Vanadium pentoxide nanofibers by electrospinning. Scripta Mater
2003;49:577–81.
[82] Sawicka KM, Prased AK, Gouma PI. Metal oxide nanowires for use in chemical sensing applications. Sensor Lett
2005;3:31–5.
[83] Dharmaraj N, Park HC, Lee BM, Viswanathamurthi P, Kim HY, Lee DR. Preparation and morphology of magnesium titanate
nanofibres via electrospinning. Inorg Chem Commun 2004;7:431–3.
[84] Whang D, Jin S, Lieber CM. Large-scale hierarchical organization of nanowire arrays for integrated nanosystems. Nano Lett
2003;3:1255–9.
[85] Yong-Jin K, Chul-Ho L, Young JH, Gyu-Chul Y, Sung Soo K, Hyeonsik C. Controlled selective growth of ZnO nanorod and
microrod arrays on Si substrates by a wet chemical method. Appl Phys Lett 2006;89:163128.
[86] Zhang Y, Jia H, Yu D. Metal-catalyst-free epitaxial growth of aligned ZnO nanowires on silicon wafers at low temperature.
J Phys D: Appl Phys 2004;37:413–5.
[87] Nguyen P, Ng HT, Yamada T, Smith MK, Li J, Jie Han J, et al. Direct integration of metal oxide nanowire in vertical field-
effect transistor. Nano Lett 2004;4:651–7.
[88] Wan Q, Dattoli EN, Fung WY, Guo W, Chen Y, Pan X, et al. High-performance transparent conducting oxide nanowires.
Nano Lett 2006;12:2909–15.
[89] Fan HJ, Fuhrmann B, Scholz R, Syrowatka F, Dadgar A, Krost A, et al. Well-ordered ZnO nanowire arrays on GaN substrate
fabricated via nanosphere lithography. J Crystal Growth 2006;287:34–8.
[90] Tseng Y, Chia C, Tsay C, Lin L, Cheng H, Kwo C, et al. Growth of epitaxial needlelike ZnO nanowires on GaN films. J
Electrochem Soc 2005;152:G95–8.
[91] Kirkham M, Wang X, Wang RL, Snyder ZL. Solid Au nanoparticles as a catalyst for growing aligned ZnO nanowires: a new
understanding of the vapour–liquid–solid process. Nanotechnology 2007;18:365304.
[92] Lin Y, Tseng Y, Yang S, Wu S, Hsu C, Chang S. Buffer-facilitated epitaxial growth of ZnO nanowire. Cryst Growth Design
2005;5:579–83.
[93] Sellacci F. Towards industrial-scale molecular nanolithography. Adv Funct Mater 2006;16:15–6.
[94] Campos LC, Dalal SH, Baptista DL, Magalhães-Paniago R, Ferlauto AS. Determination of the epitaxial growth of
zinc oxide nanowires on sapphire by grazing incidence synchrotron X-ray diffraction. Appl Phys Lett 2007;90:
181929.
[95] Li C, Fang C, Fu Q, Su F, Li G, Wu X, et al. Effect of substrate temperature on the growth and photoluminescence properties
of vertically aligned ZnO nanostructures. J Crystal Growth 2006;292:19–25.
[96] Eymery JI, Rieutord F, Favre-Nicolin V, Robach O, Niquet Y, Froberg L, et al. Strain and shape of epitaxial InAs/InP nanowire
superlattice measured by grazing incidence X-ray techniques. Nano Lett 2007;7:2596–601.
[97] Wan Q, Wei M, Zhi D, MacManus-Driscoll JL, Blamire MG. Epitaxial growth of vertically aligned and branched single-
crystalline tin-doped indium oxide nanowire arrays. Adv Mater 2006;18:234–8.
[98] Liu Y, Liu M. Growth of aligned square-shaped SnO2 tube arrays. Adv Funct Mater 2005;15:57–62.
[99] Ng HT, Chen B, Li J, Han J, Meyyappan M, Wu J, et al. Optical properties of single-crystalline ZnO nanowires on m-sapphire.
Appl Phys Lett 2003;82:2023–5.
[100] Zhao QX, Willander M, Morjan RR, Hu QH, Campbell EEB. Optical recombination of ZnO nanowires grown on sapphire and
Si substrates. Appl Phys Lett 2003;83:165–7.
[101] Gao PX, Ding Y, Wang ZL. Crystallographic orientation-aligned ZnO nanorods grown by a tin catalyst. Nano Lett
2003;3:1315–20.
[102] Wang X, Summers CJ, Wang ZL. Large-scale hexagonal-patterned growth of aligned ZnO nanorods for nano-
optoelectronics and nanosensor arrays. Nano Lett 2004;4:423–6.
[103] He JH, Hsu JH, Wang CW, Lin HN, Chen LJ, Wang ZL. Pattern and feature designed growth of ZnO nanowire arrays for
vertical devices. J Phys Chem B 2006;110:50–3.
[104] Guo Z, Zhao D, Shen D, Fang F, Zhang J, Li B. Structure and photoluminescence properties of aligned ZnO nanobolt arrays.
Cryst Growth Design 2007;7:2294–6.
[105] Hsin CL, He JH, Lee CY, Wu WW, Yeh PH, Chen LJ, et al. Lateral self-aligned p-type In2O3 nanowire arrays epitaxially grown
on silicon substrates. Nano Lett 2007;7:1799–803.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 63

[106] Johansson J, Wacaser BA, Dick KA, Seifert W. Growth related aspects of epitaxial nanowires. Nanotechnology
2006;17:S355–61.
[107] Nagashima K, Yanagida T, Tanaka H, Kawai T. Epitaxial growth of MgO nanowires by pulsed laser deposition. J Appl Phys
2007;101:124304.
[108] Tak Y, Yong K, Park C. Growth of aligned ZnO nanorods on Pt buffer layer coated silicon substrates using metallorganic
chemical vapor deposition. J Electrochem Soc 2005;152:794–7.
[109] Huang MH, Mao S, Feick H, Yan H, Wu Y, Kind H, et al. Room-temperature ultraviolet nanowire nanolasers. Science
2001;292:1897–9.
[110] Baxter JB, Aydil ES. Epitaxial growth of ZnO nanowires on a- and c-plane sapphire. J Crystal Growth 2005;274:407–11.
[111] Chen C, Xu W, Chern M. Low-temperature epitaxial growth of vertical In2O6 nanowires on a-plane sapphire with
hexagonal cross-section. Adv Mater 2007;19:3012–5.
[112] Park WI, Kim DH, Jung SW, Yi GC. Metalorganic vapor-phase epitaxial growth of vertically well-aligned ZnO nanorods.
Appl Phys Lett 2002;80:4232–4.
[113] Zùñiga-Pèrez J, Rahm A, Czekalla C, Lenzner J, Lorenz M, Grundmann M. Ordered growth of tilted ZnO nanowires:
morphological, structural and optical characterization. Nanotechnology 2007;18:195303.
[114] Wan Q, Feng P, Wang TH. Vertically aligned tin-doped indium oxide nanowire arrays: epitaxial growth and electron field
emission properties. Appl Phys Lett 2006;89:123102.
[115] Smith PA et al. Electric-field assisted assembly and alignment of metallic nanowires. Appl Phys Lett 2000;77:1399–401.
Alberto.
[116] Nagahara LA et al. Directed placement of suspended carbon nanotubes for nanometer-scale assembly. Appl Phys Lett
2002;80:3826–8.
[117] Boote JJ, Evans SD. Dielectrophoretic manipulation and electrical characterization of gold nanowires. Nanotechnology
2005;16:1500–5.
[118] Zhou R, Chang H-C, Protasenko V, Kuno M, Singh AK, Jena D, et al. CdSe nanowires with illumination-enhanced
conductivity: induced dipoles, dielectrophoretic assembly, and field-sensitive emission. J Appl Phys 2007;101:073704.
[119] Kumar S, Rajaraman S, Gerhardt RA, Wang ZL, Hesketh PJ. Tin oxide nanosensor fabrication using AC dielectrophoretic
manipulation of nanobelts. Electrochim Acta 2005;51:943–51.
[120] Wang D, Zhu R, Zhou Zhaoying, Ye X. Controlled assembly of zinc oxide nanowires using dielectrophoresis. Appl Phys Lett
2007;90:103110.
[121] Lao CS, Liu J, Gao P, Zhang L, Davidovic D, Tummala R, et al. ZnO nanobelt/nanowire Schottky diodes formed by
dielectrophoresis alignment across Au electrodes. Nano Lett 2006;6:263–6.
[122] Narayanan A, Dan Y, Deshpande V, Di Lello N, Evoy S, Raman S. Dielectrophoretic integration of nanodevices with CMOS
VLSI circuitry. IEEE Trans Nanotechnol 2006;5:101–9.
[123] Suehiro J, Nakagawa N, Hidaka S-I, Ueda M, Imasaka K, Higashihata M, et al. Dielectrophoretic fabrication and
characterization of a ZnO nanowire-based UV photosensor. Nanotechnology 2006;17:2567–73.
[124] Lin X, He X, Lu J, Gao L, Huan Q, Deng Z, et al. Manipulation and four-probe analysis of nanowires in UHV by application of
four tunneling microscope tips: a new method for the investigation of electrical transport through nanowires. Surf Inter
Anal 2006;38:1096–102.
[125] Harnack O, Pacholski C, Weller H, Yasuda A, Wessels JM. Rectifying behavior of electrically aligned ZnO nanorods. Nano
Lett 2003;3:1097–101.
[126] Wang ZL, Gao RP, Poncharal P, de Heer WA, Dai ZR, Pan ZW. Mechanical and electrostatic properties of carbon nanotubes
and nanowires. Mater Sci Eng C 2001;16:3–10.
[127] Wang ZL. New developments in transmission electron microscopy for nanotechnology. Adv Mater 2003;15:1497–514.
[128] Yu MF, Lourie O, Dyer MJ, Moloni K, Kelly TF, Ruoff RS. Strength and breaking mechanism of multi walled carbon
nanotubes under tensile load. Science 2002;287:637–40.
[129] Walton AS, Allen CS, Critchley K, G’orzny MŁ, McKendry JE, Brydson RMD, et al. Four-probe electrical transport
measurements on individual metallic nanowires. Nanotechnology 2007;18:065204.
[130] Xiao L, Xiaobo H, Junling L, Li G, Qing H, Zhitao D, et al. Manipulation and four-probe analysis of nanowires in UHV by
application of four tunneling microscope tips. Surf Interface Anal 2006;38:1096–102.
[131] Hernandez-Ramirez F, Tarancon A, Casals O, Pellicer E, Rodriguez J, Romano-Rodriguez A, et al. Electrical properties of
individual tin oxide nanowires contacted to platinum electrodes. Phys Rev B 2007;76:085429.
[132] Zhang HF, Wang CM, Buck EC, Wang L. Synthesis, characterization, and manipulation of helical SiO2 nanosprings. Nano
Lett 2003;3:577–80.
[133] Desai AV, Haque MA. Mechanical properties of ZnO nanowires. Sensors Actuat A-Phys 2007;134:169–76.
[134] Feng G, Nix WD, Yoon Y, Lee CJ. A study of the mechanical properties of nanowires using nanoindentation. J Appl Phys
2006;99:074304.
[135] Fang TH, Chang WJ. Nanolithography and nanoindentation of tantalum-oxide nanowires and nanodots using scanning
probe microscopy. Phys B: Condens Mat 2004;352:190–9.
[136] Fan HJ, Fuhrmann B, Scholz R, Himcinschi C, Berger A, Leipner H, et al. Vapour-transport-deposition growth of ZnO
nanostructures: switch between c-axial wires and a-axial belts by indium doping. Nanotechnology 2006;17:S231–9.
[137] Allen JE, Hemesath ER, Perea DE, Lensch-Falk JL, Li ZY, Yin F, et al. High-resolution detection of Au catalyst atoms in Si
nanowires. Nat. Nanotech. 2008;3:168–73.
[138] Rahm A, Kaidashev EM, Schmidt H, Diaconu M, Pöppl A, Böttcher R, et al. Growth and characterization of Mn- and Co-
doped ZnO nanowires. Microchim Acta 2007;156:21–5.
[139] Lee CY, Tseng TY, Li SY, Lin P. Effect of phosphorus dopant on photoluminescence and field-emission characteristics of
Mg0.1Zn0.9O nanowires. J Appl Phys 2006;99:024303.
[140] Lei B, Li C, Zhang D, Tang T, Zhou C. Tuning electronic properties of In2O3 nanowires by doping control. Appl Phys A
2004;79:439–42.
[141] Xiang B, Wang P, Zhang X, Dayeh SA, Aplin DPR, Soci C, et al. Rational synthesis of p-type zinc oxide nanowire arrays using
simple chemical vapor deposition. Nano Lett 2007;7:323–8.
64 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

[142] Shan CX, Liu Z, Hark SK. Temperature dependent photoluminescence study on phosphorus doped ZnO nanowires. Appl
Phys Lett 2008;92:073103.
[143] Nguyen P, Ng HT, Kong J, Cassell AM, Quinn R, Li J, et al. Epitaxial directional growth of indium-doped tin oxide nanowire
arrays. Nano Lett 2003;3:925–8.
[144] Lee W, Jeong MC, Myoung JM. Optical characteristics of arsenic-doped ZnO nanowires. Appl Phys Lett 2004;85:6167.
[145] Lin CC, Chen HP, Chen SY. Synthesis and optoelectronic properties of arrayed p-type ZnO nanorods grown on ZnO film/Si
wafer in aqueous solutions. Chem Phys Lett 2005;404:30–4.
[146] Lee W, Jeong MC, Myoung JM. Catalyst-free growth of ZnO nanowires by metal-organic chemical vapour deposition
(MOCVD) and thermal evaporation. Acta Mater 2004;52:3949–57.
[147] Geburt S, Stichtenoth D, Müller S, Dewald W, Ronning C, Wang J, et al. Rare earth doped zinc oxide nanowires. J Nanosci
Nanotechnol 2008;8:244–51.
[148] Wan Q, Dattoli EN, Lu W. Transparent metallic SnO2 nanowires. Appl Phys Lett 2007;90:222107.
[149] Dattoli EN, Wan Q, Guo W, Chen Y, Pan X, Lu W. Fully transparent thin-film transistor devices based on SnO2 nanowires.
Nano Lett 2007;7:2463–9.
[150] Huang J, Lu A, Zhao B, Wan Q. Branched growth of degenerately Sb-doped SnO2 nanowires. Appl Phys Lett
2007;91:073102.
[151] Wu Y, Fan R, Yang P. Block-by-block growth of single-crystalline Si/SiGe superlattice nanowires. Nano Lett 2002;2:83–6.
[152] May SJ, Zheng J-G, Wessels BW, Lauhon LJ. Dendritic nanowire growth mediated by a self-assembled catalyst. Adv Mater
2005;17:598–602.
[153] Pradhan B, Batabyal SK, Pal AJ. Rectifying junction in a single ZnO vertical nanowire. Appl Phys Lett 2006;89:233109.
[154] Ng HT, Li J, Smith MK, Nguyen P, Cassell A, Han J, et al. Growth of epitaxial nanowires at the junctions of nanowalls.
Science 2003;300:1249.
[155] Li L, Yang Y-W, Li G-H, Zhang L-D. Conversion of a Bi nanowire array to an array of Bi–Bi2O3 core–shell nanowires and
Bi2O3 nanotubes. Small 2006;2:548–53.
[156] Han S, Li C, Liu Z, Lei B, Zhang D, Jin W, et al. Transition metal core–shell nanowires – generic synthesis and transport
studies. Nano Lett 2007;4:1241–6.
[157] Vomiero A, Ferroni M, Comini E, Faglia G, Sberveglieri G. Preparation of radial and longitudinal nanosized
heterostructures of In2O3 and SnO2. Nano Lett 2007;7:3553–8.
[158] Mathur S, Barth S. Molecule-based chemical vapour growth of aligned SnO2 nanowires and branched SnO2/V2O3
heterostructures. Small 2007;3:2070–5.
[159] Kim D-W, Hwang I-S, Kwon SJ, Kang H-Y, Park K-S, Choi Y-J, et al. Highly conductive coaxial SnO2–In2O3 heterostructured
nanowires for Li ion battery electrodes. Nano Lett 2007;7:3041–5.
[160] Law M, Greene LE, Radenovic A, Kuykendall T, Liphardt J, Yang P. ZnO–Al2O3 and ZnO–TiO2 core–shell nanowire dye-
sensitized solar cells. J Phys Chem B 2006;110:22652–63.
[161] Joanny E, Savu R, Go’es MS, Bueno PR, de Freitas JN, Nogueira AF, et al. ZnO–Al2O3 and ZnO–TiO2 core–shell nanowire dye-
sensitized solar cells. Scripta Mater 2007;57:277–80.
[162] Savu R, Joanni E. Low-temperature, self-nucleated growth of indium–tin oxide nanostructures by pulsed laser deposition
on amorphous substrates. Scripta Mater 2006;55:979–81.
[163] Shin TI, Lee HJ, Song WY, Kim S-W, Park MH, Yang CW, et al. A homojunction of single-crystalline b-Ga2O3 nanowires and
nanocrystals. Nanotechnology 2007;18:345305.
[164] Ribeiro C, Longo E, Leite ER. Tailoring of heterostructures in a SnO2/TiO2 system by the oriented attachment mechanism.
Appl Phys Lett 2007;91:103105.
[165] Sun S, Meng G, Zhang G, Zhang L. Controlled growth and optical properties of one-dimensional ZnO nanostructures on
SnO2 nanobelts. Cryst Growth Design 2007;7:1988–91.
[166] Tsuda N, Nasu K, Fujimori A, Siratori K. Electronic conduction in oxides. 2nd ed. Berlin: Springer; 2000.
[167] Samson S, Fonstad CG. Defect structure and electronic donor. Levels in stannic oxide crystal. J Appl Phys
1973;44:4618–21.
[168] Wolkenstein T. Electronic processes on semiconductor surfaces during chemisorption. New York: Plenum press; 1991. p.
31–4.
[169] Kronik L, Shapira Y. Surface photovoltage phenomena: theory, experiment, and applications. Surf Sci Rep 1999;37:1–206.
[170] Orton JW, Powell MJ. The Hall effect in polycristalline and powdered semiconductors. Rep Prog Phys 1980;43:1265–306.
[171] Morrison SR. The chemical physics of surfaces. New York: Plenum Press; 1978.
[172] Madou MJ, Morrison SR. Chemical sensing with solid state devices. San Diego: Academic Press Inc.; 1989.
[173] Barsan N, Weimar U. Conduction model of metal oxide gas sensors. J Electroceram 2001;7:143–67.
[174] Barsan N, Weimar U. Understanding the fundamental principles of metal oxide based gas sensors; the example of CO
sensing with SnO2 sensors in the presence of humidity. J Phys: Condens Mat 2003;15:R813–39.
[175] Hahn SH, Barsan N, Weimar U, Ejakov SG, Visser JH, Soltis RE. CO sensing with SnO2 thick film sensors: role of oxygen and
water vapour. Thin Solid Films 2003;436:17–24.
[176] Heiland G, Kohl D. Physical and chemical aspects of oxidic semiconductor gas sensors. In: Sn Seiyama T, editor. Chemical
sensor technology, vol. 1. Tokyo: Kodansha; 1988. p. 15–38 (Chapter 2).
[177] Henrich VE, Cox PA. The surface science of metal oxides. Cambridge: University Press; 1994.
[178] D’Amico A, Di Natale C. A contribution on some basic definition of sensors properties. IEEE Sensors J 2001;1:183–90.
[179] Comini E, Faglia G, Sberveglieri G, Pan Z, Wang Z. Stable and highly sensitive gas sensors based on semiconducting oxide
nanobelts. Appl Phys Lett 2002;81:1869–71.
[180] Sberveglieri G, Baratto C, Comini E, Faglia G, Ferroni M, Ponzoni A, et al. Synthesis and characterization of semiconducting
nanowires for gas sensing. Sensors Actuat B 2007;121:208–13.
[181] Ying Z, Wan Q, Song ZT, Feng SL. SnO2 nanowhiskers and their ethanol sensing characteristics. Nanotechnology
2004;15:1682–4.
[182] Kolmakov A, Zhang Y, Cheng G, Moskovits M. Detection of CO and O2 using tin oxide nanowire sensors. Adv Mater
2003;15:997–1000.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 65

[183] Barsan N, Weimar U. J Electroceram 2001;7:143.


[184] Hernández-Ramı ´ rez F, Tarancón A, Romano-Rodrı́guez A, Morante JR. Electrical contacts and gas sensing analysis of
individual metal oxide nanowires and 3-D nanocrystal networks. IEEJ Trans Sensors Micromach 2006;126:537–47.
[185] Hernández-Ramı ´ rez F, Tarancón A, Romano-Rodrı́guez A, Casals O, Arbiol J, Morante JR. High response and stability in CO
and humidity measures using a single SnO2 nanowire. Sensors Actuat B 2007;121:3–17.
[186] Hernandez-Ramirez F, Barth S, Tarancon A, Casals O, Pellicer E, Rodriguez J, et al. Water vapor detection with individual
tin oxide nanowires. Nanotechnology 2007;18:424016.
[187] Meier DC, Semancik S, Button B, Strelcov E, Kolmakov A. Coupling nanowire chemiresistors with MEMS microhotplate gas
sensing platforms. Appl Phys Lett 2007;91:063118.
[188] Law M, Kind H, Messer B, Kim F, Yang P. Photochemical sensing of NO2 with SnO2 nanoribbon nanosensors at room
temperature. Angew Chem Int Engl 2002;41:2405–8.
[189] Carrey J, Kahn ML, Sanchez L, Chaudret B, Respaud M. Synthesis and transport properties of ZnO nanorods and
nanoparticles assemblies. Eur Phys J Appl Phys 2007;40:71–5.
[190] Kolmakov A. The effect of morphology and surface doping on sensitization of quasi-1D metal oxide nanowire gas sensors.
Proc SPIE 2006;6370:63700X1–X8.
[191] Wan Q, Wang TH. Single-crystalline Sb-doped SnO2 nanowires: synthesis and gas sensor application. Chem Commun
2005:3841–3.
[192] Bie L-J, Yan X-N, Yin J, Duan Y-Q, Yuan Z-H. Nanopillar ZnO gas sensor for hydrogen and ethanol. Sensors Actuat B
2007;126:604–8.
[193] Hsueh T-J, Hsu C-L, Chang S-J, Chen I-C. Laterally grown ZnO nanowire ethanol gas sensors. Sensors Actuat B
2007;126:473–7.
[194] Wan Q, Li QH, Chen YJ, Wang TH, He XL, Li JP, et al. Fabrication and ethanol sensing characteristics of ZnO nanowire gas
sensors. Appl Phys Lett 2004;84:3654–6.
[195] Xiangfeng C, Dongli J, Djurisic AB, Leung YH. Gas-sensing properties of thick film based on ZnO nano-tetrapods. Chem
Phys Lett 2005;401:426–9.
[196] Xu J, Chen Y, Li Y, Shen J. Gas sensing properties of ZnO nanorods prepared by hydrothermal method. J Mater Sci
2005;40:2919–21.
[197] Wang C, Chu X, Wu M. Detection of H2S down to ppb levels at room temperature using sensors based on ZnO nanorods.
Sensors Actuat B 2006;113:320–3.
[198] Kang BS, Heo YW, Tien LC, Norton DP, Ren F, Gila BP, et al. Hydrogen and ozone gas sensing using multiple ZnO nanorods.
Appl Phys A 2005;80:1029–32.
[199] Parthangal PM, Cavicchi RE, Zachariah MR. A universal approach to electrically connecting nanowire arrays using
nanoparticles-application to a novel gas sensing architecture. Nanotechology 2006;17:3786–90.
[200] Wang HT, Kang BS, Ren F, Tien LC, Sadik PW, Norton DP, et al. Hydrogen selective sensing at room temperature with ZnO
nanorods. Appl Phys Lett 2005;86:243503.
[201] Tien LC, Sadik PW, Norton DP, Voss LF, Pearton SJ, Wang HT, et al. Hydrogen sensing at room temperature with Pt-coated
ZnO thin films and nanorods. Appl Phys Lett 2005;87:222106.
[202] Rout CS, Hari Krishna S, Vivekchand SRC, Govindaraj, Rao CNR. Hydrogen and ethanol sensors based on ZnO nanorods,
nanowires and nanotubes. Chem Phys Lett 2006;418:586–90.
[203] Liao L, Lu HB, Li JC, Liu C, Fu DJ, Liu YL. The sensitivity of gas sensor based on single ZnO nanowire modulated by helium
ion radiation. Appl Phys Lett 2007;91:173110-1–0-3.
[204] Xianfeng C, Caihong W, Dongli J, Chenmou Z. Ethanol sensor based on indium oxide nanowires prepared by carbothermal
reduction reaction. Chem Phys Lett 2004;399:461–4.
[205] Polleux J, Gurlo A, Barsan N, Weimar U, Antonietti M, Niederberger M. Template-free synthesis and assembly of single
crystalline tungsten oxide nanowires and their gas sensing properties. Angew Chem 2006;45:261–5.
[206] Gurlo A, Barsan N, Ivanovskaya M, Weimar U, Gopel W. In2O3 and MoO3–In2O3 thin film semiconductor sensors:
interaction with NO2 and O3. Sensors Actuat B 1998;47:92–9.
[207] Kim YS, Ha S-C, Kim K, Yang H, Choi S-Y, Kim YT. Room temperature semiconductor gas sensor based on
nonstoichiometric tungsten oxide nanorod film. Appl Phys Lett 2005;86:213105.
[208] Liu J, Wang X, Peng Q, Li Y. Vanadium pentoxide nanobelts: higly selective and stable sensor materials. Adv Mater
2005;17:764–7.
[209] Raible I, Burghard M, Schlecht U, Yasuda A, Vossmeyer T. V2O5 nanofibres: novel gas sensors with extremely high
sensitivity and selectivity to amines. Sensors Actuat B 2005;106:730–5.
[210] Sysoev VV, Button BK, Wepsiec K, Dmitriev S, Kolmakov A. Toward the nanoscopic ‘‘Electronic Nose”: hydrogen vs carbon
monoxide discrimination with an array of individual metal oxide nano- and mesowire sensors. Nano Lett 2006;6:1584–8.
[211] Liu F, Bao M, Wang KL, Li C, Lei B, Zhou C. Low T one-dimensional transport of In2O3 nanowires. Appl Phys Lett
2005;86:213101.
[212] Yun YS, Park JY, Oh H, Kim JJ, Kim SS. Electrical transport properties of size-tuned ZnO nanorods. J Mater Res
2006;21:132–6.
[213] Hong WK, Hwang DK, Park IK, Jo G, Song S, Park SJ, et al. Realization of highly reproducible ZnO nanowire field effect
transistors with n-channel depletion and enhancement modes. Appl Phys Lett 2007;90:243103.
[214] Dattoli EN, Wan Q, Guo W, Chen YB, Pan XQ, Lu W. Fully transparent thin-film transistor devices based on SnO2
nanowires. Nano Lett 2007;7:2463–9.
[215] Chang PC, Fan Z, Chien CJ, Stichtenoth D, Ronning C, Lu JG. High-performance ZnO nanowire field effect transistors. Appl
Phys Lett 2006;89:133113.
[216] Park WI, Kim JS, Yi GC, Bae MH, Lee HJ. Fabrication and electrical characteristics of high-performance ZnO nanorod field-
effect transistors. Appl Phys Lett 2004;85:5052–4.
[217] Jo G, Maeng J, Kim TW, Hong WK, Choi BS, Lee T. Channel-length and gate-bias dependence of contact resistance and
mobility for In2O3 nanowire field effect transistors. J Appl Phys 2007;102:084508.
[218] Dimitrakopoulos CD, Malenfant PRL. Organic thin film transistors for large area electronics. Adv Mater 2002;14:99–117.
66 E. Comini et al. / Progress in Materials Science 54 (2009) 1–67

[219] Mitzi DB, Kosbar LL, Murray CE, Copel M, Afzali A. High mobility ultrathin semiconducting films prepared by spin coating.
Nature 2004;428:299–303.
[220] Fortunato EMC, Barquinha PMC, Pimentel ACMBG, Goncülves AMF, Marques AJS, Pereira LMN, et al. Fully transparent ZnO
thin film transistor produced at room temperature. Adv Mater 2005;17:590–4.
[221] Martel R, Schmidt T, Shea HR, Hertel T, Avouris P. Single- and multi-wall carbon nanotube field-effect transistors. Appl
Phys Lett 1998;73:2447–9.
[222] Li C, Zhang DH, Liu XL, Han S, Tang T, Han J, et al. In2O3 nanowires as chemical sensors. Appl Phys Lett 2003;82:1613–5.
[223] Zhang DJ, Li C, Liu XL, Han S, Tang T, Zhou CW. Doping dependent NH3 sensing of indium oxide nanowires. Appl Phys Lett
2003;83:1845–7.
[224] Li C, Zhang DH, Lei B, Han S, Liu XL, Zhou CW. Surface treatment and doping dependence of In2O3 nanowires as ammonia
sensors. J Phys Chem B 2003;107:12451–5.
[225] Zhang DH, Liu ZQ, Li C, Tang T, Liu XL, Han S, et al. Detection of NO2 down to ppb levels using individual and multiple
In2O3 nanowire devices. Nano Lett 2004;4:1919–24.
[226] Zhang Y, Kolmakov A, Chretien S, Metiu H, Moskovits M. Control of catalytic reactions at the surface of a metal oxide
nanowire by manipulating electron density inside it. Nano Lett 2004;4:403–7.
[227] Zhang Y, Kolmakov A, Lilach Y, Moskovits M. Electronic control of chemistry and catalysis at the surface of an individual
tin oxide nanowire. J Phys Chem B 2005;109:1923–9.
[228] Fan ZY, Wang DW, Chang PC, Tseng WY, Lu JG. ZnO nanowire field-effect transistor and oxygen sensing property. Appl
Phys Lett 2004;85:5923–5.
[229] Fan ZY, Lu JG. Gate-refreshable nanowire chemical sensors. Appl Phys Lett 2005;86:123510.
[230] Bogner M, Fuchs A, Scharnagl K, Winter R, Doll T, Eisele I. Electrical field impact on the gas adsorptivity of thin metal oxide
films. Appl Phys Lett 1998;17:2524–6.
[231] Faglia G, Baratto C, Sberveglieri G, Zha M, Zappettini A. Adsorption effects of NO2 at ppm level on visible
photoluminescence response of SnO2 nanobelts. Appl Phys Lett 2005;86:011923.
[232] Hu J, Bando Y, Liu Q, Goldberg D. Laser-ablation growth and optical properties of wide and long single-crystal SnO2
ribbons. Adv Funct Mater 2003;13:493–6.
[233] He JH, Wu TH, Sin CL, Li KM, Chen LJ, Chueh YL, Chou LJ, et al. Beaklike SnO2 nanorods with strong photoluminescent and
field-emission properties. Small 2006;2:116–20.
[234] Luo S, Chu PK, Liu W, Zhang M, Lin C. Origin of low-temperature photoluminescence from SnO2 nanowires fabricated by
thermal evaporation and annealed in different ambients. Appl Phys Lett 2006;88:183112.
[235] Luo SH, Fan JY, Liu WL, Zhang M, Song ZT, Liu CL, et al. Synthesis and low-temperature photoluminescence properties of
SnO2 nanowires and nanobelts. Nanotechnology 2006;17:1695–9.
[236] Calestani D, Zha M, Zappettini A, Lazzarini L, Salviati G, Zanotti L, et al. Structural and optical study of SnO2 nanobelts and
nanowires. Mater Sci Eng C 2005;25:625–30.
[237] Zhou JX, Zhang MS, Hong JM, Yin Z. Solid State Commun 2006;138:242–6.
[238] Lettieri S, Bismuto A, Maddalena P, Baratto C, Comini E, Faglia G, et al. Gas sensitive light emission properties of tin oxide
and zinc oxide nanobelts. J Non-Cryst Solids 2006;352:1457–60.
[239] Djurisic AB, Leung YH. Optical properties of ZnO nanostructures. Small 2006;2:944–61.
[240] Huang MH, Wu Y, Feick H, Tran N, Weber E, Yang P. Catalytic growth of zinc oxide nanowires by vapor transport. Adv
Mater 2001;13:113–6.
[241] Sun Y, Fuge GM, Fox NA, Riley DJ, Ashfold MNR. Synthesis of aligned arrays of ultrathin ZnO nanotubes on a Si wafer
coated with a thin ZnO. Film Adv Mater 2005;17:2477–81.
[242] Kahn ML, Cardinal T, Bousquet B, Monge M, Jubera V, Chaudret B. Optical properties of zinc oxide nanoparticles and
nanorods synthesized using an organometallic method. ChemPhysChem 2006;7:2392–7.
[243] Vanhausden K, Warren W, Seager CH, Tallant DR, Voigt JA, Gnade BE. J Appl Phys 1996;79:7983.
[244] Hu JW, Bando Y. Growth and optical properties of single-crystal tubular ZnO whiskers. Appl Phys Lett 2003;82:1401.
[245] Lin B, Fu Z, Jia Y. Green luminescent center in undoped zinc oxide films deposited on silicon substrates. Appl Phys Lett
2001;79:943.
[246] Studenikin SA, Cocivera M. Time-resolved luminescence and photoconductivity of polycrystalline ZnO films. J Appl Phys
2002;91:5060.
[247] Yao BD, Chan YF, Wang N. Formation of ZnO nanostructures by a simple way of thermal evaporation. Appl Phys Lett
2002;81:757.
[248] Sun Ye, George Ndifor-Angwafor N, Jason Riley D, Ashfold Michael NR. Synthesis and photoluminescence of ultra-thin
ZnO nanowire/nanotube arrays formed by hydrothermal growth. Chem Phys Lett 2006;431:352–7.
[249] Sun Y, Ashfold MNR. Photoluminescence from diameter-selected ZnO nanorod arrays. Nanotechnology 2007;18:245701.
[250] Djurisic AB, Leung YH, Tam KH, Ding L, Ge WK, Chen HY, et al. Green, yellow, and orange defect emission from ZnO
nanostructures: influence of excitation wavelength. Appl Phys Lett 2006;88:103107.
[251] Shalish I, Temkin H, Narayanamurti V. Size dependent surface luminescence in ZnO nanowires. Phys Rev B
2004;69:245401.
[252] Kwok WM, Djurisic AB, Leung YH, Li D, Tam KH, Phillips DL, et al. Influence of annealing on stimulated emission in ZnO
nanorods. Appl Phys Lett 2006;89:183112.
[253] Comini E, Baratto C, Faglia G, Ferroni M, Sberveglieri G. Single crystal ZnO nanowires as optical and conductometric
chemical sensor. J Phys D 2007;40:7255–9.
[254] Melnick DA. Zinc oxide photoconduction, an oxygen adsorption process. J Chem Phys 1957;26:1136–46.
[255] Shapira Y, McQuistan RB, Lichtman D. Relationship between photodesorption and surface conductivity in ZnO. Phys Rev B
1977;15:2163–9.
[256] Yan H, He R, Johnson J, Law M, Saykally RJ, Yang P. Dendritic nanowire ultraviolet laser array. J Am Chem Soc
2003;125:4728–9.
[257] Song JK, Willer U, Szarko JM, Leone SR, Shihong L, Zhao Y. Ultrafast upconversion probing of lasing dynamics in single ZnO
nanowire lasers. J Phys Chem C 2008;112:1679–84.
E. Comini et al. / Progress in Materials Science 54 (2009) 1–67 67

[258] Huynh WU, Dittmer JJ, Alivisatos AP. Hybrid nanorod-polymer solar cells. Science 2002;295:2425.
[259] Baxter JB, Aydil ES. Nanowire-based dye-sensitized solar cells. Appl Phys Lett 2005;86:053114.
[260] Law M, Greene LE, Johnson JC, Saykally R, Yang P. Nanowire dye-sensitized solar cells. Nat Mater 2005;4:455.
[261] Baxter JB, Aydil ES. Dye-sensitized solar cells based on semiconductor morphologies with ZnO nanowires. Sol Energy
Mater Solar Cells 2006;90:607–22.
[262] O’Regan B, Grätzel M. A low-cost, high efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature
1991;353:737.
[263] Leschkies KS, Divakar R, Basu J, Enache-Pommer E, Boercker JE, Carter CB, et al. Photosensitization of ZnO nanowires with
CdSe quantum dots for photovoltaic devices. Nano Lett 2007;7:1793.
[264] Lee CJ, Lee TJ, Lyu SC, Zhang Y, Ruh H, Lee HJ. Field emission from well-aligned zinc oxide nanowires grown at low
temperature. Appl Phys Lett 2002;81:3648.
[265] Zhang. Photoluminescence field-emission characteristics of ZnO nanowires synthesized by two-step method. Vacuum
2008;82:30.
[266] Lee. Field emission characteristics of ZnO nanoneedle array cell under ultraviolet irradiation. Phys Lett A 2007;370:345.
[267] Chen H, Qi J, Zhang X, Liao Q, Huang Y. Controlled growth and field emission properties of zinc oxide nanopyramid arrays.
ASS 2007;253:8901.
[268] Ashanulhaq Q, Kim J-H, Hahn Y-B. Controlled selective growth of ZnO nanorod arrays and their field emission properties.
Nanotechnology 2007;18:485307.
[269] Lin S, Ye Z, He H, Zeng Y, Tang H, Zhao B, et al. Catalyst-free synthesis of vertically aligned screw-shape InZnO nanorods
array. J Cryst Growth 2007;306:339.
[270] Kojima Y, Kasuya K, Ooi T, Nagato K, Takayama K, Nakao M. Effects of oxidation during synthesis on structure and field-
emission property of tungsten oxide nanowires. J Appl Phys 2007;46:6250.
[271] Bhise AB, Late DJ, Ramgir NS, More MA, Mulla IS, Pillai VK, et al. Field emission investigations of RuO2-doped SnO2 wires.
Appl Surf Sci 2007;253:9159.
[272] Bhise AB, Late DJ, Walke PS, More MA, Pillai VK, Mulla IS, et al. Sb-doped SnO2 wire – highly stable field emitter. J Cryst
Growth 2007;307:87.
[273] Miyauchi M, Tokudome H, Toda Y, Kamiya T, Hosono H. Electron field emission from TiO2 nanotube arrays synthesized by
hydrothermal reaction. Appl Phys Lett 2006;89:043114.
[274] Fang X, Bando Y, Shen G, Ye C, Gautam UK, Costa PMFJ, et al. Ultrafine ZnS nanobelts as field emitters. Adv Mater
2007;19:2593.
[275] Wang ZL, Song JH. Piezoelectric nanogenerators based on zinc oxide nanowire arrays. Science 2006;312:242–56.
[276] Zheng G, Patolsky F, Cui Y, Wang WU, Lieber CM. Multiplexed electrical detection of cancer markers with nanowire sensor
arrays. Nat Biotechnol 2005;23:1294–301.
[277] Patolsky F, Zheng G, Lieber CM. Nanowire-based biosensors. Anal Chem 2006;78:4260–9.

You might also like