Vous êtes sur la page 1sur 130

CuO nanostructures: Synthesis, characterization,

growth mechanisms, fundamental properties,


and applications
Qiaobao Zhang
a
, Kaili Zhang
a,
, Daguo Xu
a
, Guangcheng Yang
b
, Hui Huang
b
,
Fude Nie
b
, Chenmin Liu
c
, Shihe Yang
d
a
Department of Mechanical and Biomedical Engineering, City University of Hong Kong, 83 Tat Chee Avenue, Kowloon, Hong Kong
b
Institute of Chemical Materials, China Academy of Engineering Physics, Mianyang 621900, China
c
Nano and Advanced Materials Institute, Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong
d
Department of Chemistry, William Mong Institute of Nano Science and Technology, Hong Kong University of Science and
Technology, Clear Water Bay, Kowloon, Hong Kong
a r t i c l e i n f o
Article history:
Received 21 May 2013
Received in revised form27 September 2013
Accepted 29 September 2013
Available online 5 October 2013
a b s t r a c t
Nanoscale metal oxide materials have been attracting much atten-
tion because of their unique size- and dimensionality-dependent
physical and chemical properties as well as promising applications
as key components in micro/nanoscale devices. Cupric oxide (CuO)
nanostructures are of particular interest because of their interest-
ing properties and promising applications in batteries, supercapac-
itors, solar cells, gas sensors, bio sensors, nanouid, catalysis,
photodetectors, energetic materials, eld emissions, superhydro-
phobic surfaces, and removal of arsenic and organic pollutants
from waste water. This article presents a comprehensive review
of recent synthetic methods along with associated synthesis mech-
anisms, characterization, fundamental properties, and promising
applications of CuO nanostructures. The review begins with a
description of the most common synthetic strategies, characteriza-
tion, and associated synthesis mechanisms of CuO nanostructures.
Then, it introduces the fundamental properties of CuO nanostruc-
tures, and the potential of these nanostructures as building blocks
for future micro/nanoscale devices is discussed. Recent develop-
ments in the applications of various CuO nanostructures are also
reviewed. Finally, several perspectives in terms of future research
on CuO nanostructures are highlighted.
2013 Elsevier Ltd. All rights reserved.
0079-6425/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.pmatsci.2013.09.003

Corresponding author.
E-mail addresses: kaizhang@cityu.edu.hk, kaili_zhang@hotmail.com (K. Zhang).
Progress in Materials Science 60 (2014) 208337
Contents lists available at ScienceDirect
Progress in Materials Science
j our nal homepage: www. el sevi er . com/ l ocat e/ pmat sci
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
2. Synthesis of CuO nanostructures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
2.1. Solution-based methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
2.1.1. Hydrothermal synthetic method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
2.1.2. Solution-based chemical precipitation methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
2.2. Solid-state thermal conversion of precursors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
2.3. Electrochemical method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
2.4. Thermal oxidation method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
2.5. Other synthetic methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
3. Growth mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
3.1. Oriented attachment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
3.2. Ostwald ripening process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
3.3. Scroll of Cu(OH)
2
nanosheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
3.4. Stress and grain-boundary diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
3.5. Stress-induced cracking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
4. Fundamental properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
4.1. Crystal structures and phase transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
4.2. Electronic band structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
4.3. Optical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
4.4. Electrical conductivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
4.5. Photoelectrochemical (PEC) properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
4.6. Magnetic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
5. Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
5.1. Application in LIBs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
5.2. Application in supercapacitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
5.3. Application in sensors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
5.3.1. Application in gas sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
5.3.2. Application in enzyme-free glucose electrochemical sensors . . . . . . . . . . . . . . . . . . . . 288
5.4. Application in solar cells. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
5.5. Application in photodetectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
5.6. Application in catalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
5.7. Application in photocatalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
5.8. Application in enhancement of thermal conductivity of nanofluid . . . . . . . . . . . . . . . . . . . . . . . 307
5.9. Application in nEMs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
5.10. Application in field emission displays (FEDs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
5.11. Application in superhydrophobic surfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
5.12. Application in removal of arsenic (As) and organic pollutants from waste water . . . . . . . . . 316
5.13. Toxicity of CuO nanoparticles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
6. Conclusion and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
1. Introduction
Nanostructured transition metal oxides (MOs), a particular class of nanomaterials, are the indisput-
able prerequisite for the development of various novel functional and smart materials. These transi-
tion MO nanocrystals have been attracting much attention not only for fundamental scientic
research, but also for various practical applications because of their unique physical and chemical
properties [128]. These physical and chemical properties are strongly dependent on the sizes, shapes,
compositions, and structures of the nanocrystals. Interesting phenomena such as remarkable increase
in surface-to-volume ratio, signicant change in surface energy, and quantum connement effects
occur when transition MOs are reduced to nanoscale dimension [7,20,21]. These phenomena result
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 209
in a variety of new physical and chemical properties that are not feasible for materials with bulk
dimensionality. Therefore, the manipulation of well-controlled synthesis and fabrication of nanostruc-
tured transition MOs with different sizes, shapes, chemical compositions, and structures is crucial in
the advancement in nanoscience and nanotechnology. Consequently, various nanostructured transi-
tion MOs have been synthesized by diverse chemical, physicochemical, and physical strategies
[17,9,1417,20,21,25,28]. Compared with their micro or bulk counterparts, nanostructured transition
MOs exhibit unique structural characteristics and size connement effects as well as novel properties.
These properties contribute to the potential of transition MOs as candidates for both theoretical
studies and practical applications in micro/nanodevices.
Cupric oxide (CuO) has been a hot topic among the studies on transition MOs because of its inter-
esting properties as a p-type semiconductor with a narrow band gap (1.2 eV in bulk) and as the basis
of several high-temperature superconductors and giant magneto resistance materials [25,2935]. CuO
nanostructures with large surface areas and potential size effects possess superior physical and
chemical properties that remarkably differ from those of their micro or bulk counterparts. These nano-
structures have been extensively investigated because of their promising applications in various elds.
CuO nanostructures are also considered aselectrode materials for the next-generation rechargeable
lithium-ion batteries (LIBs) because of their high theoretical capacity, safety, and environmental
friendliness [36]. They are also promising materials for the fabrication of solar cells because of their
high solar absorbance, low thermal emittance, relatively good electrical properties, and high carrier
concentration [37]. Furthermore, CuO nanostructures are extensively used in various other applica-
tions, including gas sensors [38], bio-sensors [39], nanouid [40], photodetectors [41], energetic mate-
rials (EMs) [42], eld emissions [43], supercapacitors [44], removal of inorganic pollutants [45,46],
photocatalysis [47], and magnetic storage media [48]. Recent studies have demonstrated that nano-
scale CuO can be used to prepare various organicinorganic nanocomposites with high thermal con-
ductivity, high electrical conductivity, high mechanical strength, high-temperature durability, and so
on [32,33,49,50]. Moreover, the nanoscale CuO is an effective catalyst for CO and NO oxidation as well
asin the oxidation of volatile organic chemicals such as methanol [5153]. In addition, some reports
have demonstrated the excellent activities of nanoscale CuO as catalyst in the CN coupling and
CS cross-coupling of thiols with iodobenzene reactions [51,54,55]. The superhydrophobic properties
of CuO nanostructures render these materials as promising candidates in Lotus effect self-cleaning
coatings (anti-biofouling), surface protection, textiles, water movement, microuidics, and oilwater
separation [56]. Thus, nanoscale CuO with different shapes and dimensions, such as zero-dimensional
(0D) nanoparticles, one-dimensional (1D) nanotubes, 1D nanowires/rods, two-dimensional (2D)
nanoplates, 2D nanolayers, and several complex three-dimensional (3D) nanoowers, spherical-like,
and urchin-like nanostructures have been synthesized using numerous methodologies. More interest-
ing applications of CuO nanostructures are being explored.
Cuprous oxide (Cu
2
O), another important copper (Cu)-based oxide, is also one of the rst known
p-type semiconductor materials [57]. However, Cu
2
O and CuO have striking contrasting colors, crystal
structures, and physical properties [58]. Cu
2
O is a reddish p-type semiconductor of both ionic and
covalent nature with cubic structure (space group, O
4
h
pn3m) that exhibits various excitonic levels.
By contrast, CuO has an iron-dark color with a more complex monoclinic tenorite crystallographic
structure (space group, C2/c) and displays promising antiferromagnetic ordering [58,59]. Cu
2
O is
expected to have an essentially full Cu 3d shell with a direct forbidden band gap of 2.17 eV in bulk,
which can only absorb light up to the visible region. CuO has an open 3d shell with a direct band
gap (1.2 eV in bulk) of charge-transfer type, which can absorb light up to the near infraredregion
[59,60]. Recent reports have demonstrated that CuO has higher conductivity than Cu
2
O but with lower
carrier mobility [61].
Although these two Cu-based oxides have contrasting properties, both oxides are of considerable
interest in photovoltaics, gas sensors, CO oxidation catalysts, various heterogeneous catalysts, and
LIBs, because of their low band-gap energy, high optical absorption, high catalytic activity, nontoxic
nature, and low-cost [30,31,62,63]. In recent years, the size- and morphology-controlled synthesis
and application of Cu
2
O and CuO have been intensively investigated [25,2831]. However, CuO is
more stable than Cu
2
O because Cu(II) ions are much more stable in ambience, which makes it more
important in practical applications. Furthermore, the synthesis, properties, and applications of various
210 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
su thao tac
hoat tinh xuc tac cao
khong dong nhat
moi truong
Cu
2
O nanostructures have been extensively reviewed [28,31,6466]. Therefore, the recent advance-
ment in Cu
2
O will not be covered in this article to avoid overlapping reviews.
Additionally, compared with other MO nanostructures, such as TiO
2
[7,9], ZnO [14], WO
3
[21], and
SnO
2
[17], CuO nanostructures have more interesting magnetic and superhydrophobic properties.
Additionally, these nanostructures demonstrate unique applications in heterogeneous catalysis in
the complete conversion of hydrocarbons into carbon dioxide, enhancement of thermal conductivity
of nanouid, nanoenergetic materials (nEMs), and superhydrophobic surfaces. CuO nanostructures as
anode materials for LIBs have not been paid as much attention as SnO
2
[17,67] and TiO
2
[67,68].
However, the simplicity of preparation, scalability, non-toxicity, abundance, and low-cost of CuO
nanostructures is expected to increase the application of these nanomaterials as anode materials
for LIBs. MOs, including SnO
2
, ZnO, TiO
2
along with their various sub-stoichiometric forms [38], are
widely considered for gas sensor applications. Thus, the study of CuO for gas sensors is expected
to increase rapidly because of the easy synthesis of high-quality and single-crystalline CuO
nanostructures.
However, only few reports have described the synthesis strategies adopted for CuO nanostructures
along with the introduction of their related applications [25,29,31]. Furthermore, most of these review
papers only focused on the 1D CuO nanostructures [25,30,31]. No review for the systematic introduc-
tion of the recent progresses of various CuO nanostructures has been published. This article will begin
with a systemic discussion on the synthesis of different CuO nanostructures. For each synthetic meth-
od, critical comments will be provided based on our knowledge and related research experience. Next,
the associated synthesis mechanisms for controlling the size, morphology, and structure of CuO nano-
structures will be addressed. The fundamental properties of CuO nanostructures will also be intro-
duced. The promising applications of 0D CuO nanoparticles, 1D CuO nanotubes, 1D nanowires/rods,
2D CuO nanostructures, and several complex 3D CuO nanostructures along with perspectives in terms
of future research on CuO nanostructures will be highlighted. This review aims to provide a critical
discussion of the synthesis of CuO nanostructures. The potential of CuO nanostructures as functional
components for fabrication of micro/nanodevices are also evaluated and highlighted. In particular, we
focus on the fundamental properties and various nanostructured forms of CuO that have been re-
ported in the literature to date and summarize the various synthetic strategies. Promising selections
Fig. 1. Schematic diagram of a typical hydrothermal synthesis for CuO nanostructures.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 211
he so ty luong
hoat chat be mat
and interesting applications are presented, and nally some perspectives on the future research and
development of CuO nanostructures are provided.
2. Synthesis of CuO nanostructures
The development of synthetic methods has been widely accepted as an area of fundamental impor-
tance to the understanding and application of nanoscale materials. It allows scientists to modulate dif-
ferent parameters such as morphology, particle size, size distributions, and composition. Numerous
methods have been recently developed to synthesize various CuO nanostructures with diverse mor-
phologies, sizes, and dimensions using various chemical and physical strategies. In this review, we
present the most common synthetic strategies and associated mechanisms for tuning the morphology,
size, and structure of the CuO nanostructures along with the studies of the effects of these parameters
on the chemical and physical properties of the synthesized nanostructures.
2.1. Solution-based methods
Solution-based synthetic methods are very common and effective ways to prepare various MO
nanostructures with good control of shape, composition, and reproducibility. They usually have rela-
tively low reaction temperature and are exible and suitable for large-scale production. Moreover, the
synthesis parameters can be rationally tailored throughout the entire process, which is benecial for
more precise control of compositions, sizes, and dimensions of the resulting materials
[12,20,22,30,6972]. Among a variety of solution-based synthetic methods, hydrothermal and chem-
ical precipitation techniques have been widely used to synthesize CuO nanostructures.
2.1.1. Hydrothermal synthetic method
Hydrothermal synthetic method, in which the reactions are conducted in water in a pressurized
sealed container and reaction temperature over the critical point of a solution, has been widely used
to generate different nanomaterials because its reaction system is simple/green, with convenient post-
treatment [1,12,69]. Furthermore, the hydrothermal method exhibits the following advantages: (i)
numerous inorganic salts can be well dissolved in water, allowing a very exible adjustment of the
source of the metal ions depending on the requirements; (ii) water is low-cost, non-toxic, and envi-
ronment friendly; (iii) small coordinating molecules can be easily applied to modulate the growth
of the nal nanocrystals; and (iv) the strong polarity of water may be favorable to the oriented growth
of nanocrystals [69]. A schematic diagram of a typical hydrothermal synthesis for CuO nanostructures
is shown in Fig. 1.
Fig. 2. Temperature contour diagram of a T-type micro mixer [75]. Copyright 2011 Elsevier.
212 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
co the lap lai
hop ly lm cho phu hop
lang dong ket tua
bieu do bien nhiet
phan cuc thuan tien
2.1.1.1. Synthesis of CuO nanoparticles. The hydrothermal synthesis of CuO nanoparticles is generally
based on a two-step process. First, cupric hydroxide [Cu(OH)
2
] particles are formed by the reaction
of a cupric salt precursor with a basic solution, such as sodium hydroxide (NaOH) or ammonium
hydroxide. The Cu(OH)
2
particles are then thermally dehydrated in an autoclave at xed temperatures
to obtain the nal CuO nanoparticles. With the hydrothermal techniques, experimental parameters
such as cupric concentration, pH, growth time, or growth temperature determine the nal dimension,
size, and quality of the CuO nanoparticles.
Neupane et al. [73] synthesized ake-like CuO nanoparticles by simply controlling the precipitation
reaction temperature between the copper nitrate trihydrate [Cu(NO
3
)
2
3H
2
O] and NaOH under a
hydrothermal process. The results showed as-prepared samples with sizes of 37 nm, regular
ake-like morphology, and uniform size distribution. A series of controlled experiments conrmed
that the temperature and the partial pressure inside the autoclave changed the morphology and phase
of CuO nanoparticles during the hydrothermal process. To study the effect of temperature on the mor-
phology and phase control, Neupane et al. xed the reaction time to 2 h by only adjusting the temper-
ature in the reaction system. When the reaction temperature was xed at 100 C, impure phases of
Cu(OH)
2
and Cu
2
O appeared. Increasing the reaction temperature to 300 C resulted in the formation
of pure metallic Cu. The uniformly dispersed pure CuO nanoparticles were obtained at an optimum
temperature of 200 C. These results indicated that the temperature and the partial pressure inside
the autoclave are essential in controlling the morphology and phase of CuO nanoparticles during
the hydrothermal process.
Chakraborty et al. [74] synthesized CuO nanoparticles by the hydrothermal route using two differ-
ent organometallic and inorganic precursors of copper acetylacetonate [Cu(C
5
H
7
)
2
:Cu(AA)
2
] and
Cu(NO
3
)
2
3H
2
O. The resulting ower-like CuO nanoparticles are both in single phase. However, the
synthesized nanoparticles from the two different precursors had different the vibrational properties.
The appearance of the Raman peak at 218 cm
1
was observed only in the CuO nanostructures synthe-
sized using Cu(AA)
2
as the precursor, which was attributed to a certain concentration of the precursor
resulting in particular defect states which induced the structural changes in the synthesized product.
Sue et al. [75] used a T-type micro mixer (Fig. 2) at 673 Kand 30 MPa to synthesize the CuO, nickel oxide
(NiO), and iron oxide (Fe
2
O
3
) nanoparticles by the hydrothermal method. The effects of the various
experimental and sample parameters such as residence time, metal species on conversion, crystal
structure, particles size, and mass loss were investigated. The time variation of conversion, average
particle size, and coefcient of variation of the particles showed diverse behavior depending on the me-
tal species, that is, differences in MO solubilities where Fe
2
O
3
< NiO < CuO. For CuO nanoparticles,
Fig. 3. Transmission electron microscope (TEM) images and SAED patterns of CuO nanorods prepared at (a) room temperature
and (b) 100 C [79]. Copyright 2004 American Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 213
dang tam vay
nucleation did not occur at 0.002 s, but it remarkably proceeded to 0.157 s with increasing conversion.
The average particle size of CuO was found to increase quickly from 23.7 nm to 28.7 nm during the
early reaction stage (60.257 s), and then it gradually increased to 34.3 nm by increasing the residence
time to 2 s.
In addition, Outokesh et al. [76] reported the hydrothermal synthesis of CuO nanoparticles under
near-critical and supercritical conditions. During synthesis, three targets of the sample parameters,
namely, yield of the reaction, size of the nanoparticles, and purity of the products, were optimized
by the authors through a series of controlled experiments. Results show that the optimization of
the three parameters was obtained under the following reaction conditions: T = 500 C, time = 2 h,
[Cu(NO
3
)
2
] = 0.1 mol dm
3
, and at pH 3. The appropriate mechanisms of the formation of CuO
nanoparticles were proposed by the authors as follows. First, nanoparticles are suggested to be
in the liquid phase similar to Cu(OH)
2
. Second, in the presence of HNO
3
under relatively high
temperature, some of the initially formed Cu(OH)
2
or CuO are transformed to Cu
2
(OH)
3
NO
3
. Finally,
Cu(OH)
2
and Cu
2
(OH)
3
NO
3
are decomposed to CuO.
2.1.1.2. Synthesis of CuO 1D nanostructures. Hydrothermal method can also synthesize CuO 1D nano-
structures including CuO nanorods, nanotubes, and nanoneedles. Cao et al. [77] reported the synthesis
of Cu, Cu
2
O, and CuO nanotubes as well as nanorods. In a typical process, a 0.8 mM CuCl
2
and 3 M
NaOH were dissolved to prepare CuOH
2
4
solution. Then, a surfactant cetyltrimethyl ammonium bro-
mide (CTAB) was added to the solution under vigorous stirring at 50 C for 30 min to ensure complete
dissolution of CTAB. After adding 0.8 mM glucose, the solution was transferred into a stainless steel
autoclave and kept at room temperature for 1 h. Finally, after the CuO nanotubes were collected, they
were washed by ethanol and water, centrifuged, and dried. The concentration of CuOH
2
4
was just
adjusted to 15 mM to synthesize the CuO nanorods. Cheng [78] recently synthesized CuO nanorods
in a large scale using the same method and proved that the concentration of surfactant CTAB critically
inuences the morphology of CuO nanorods.
Gao et al. [79] reported that the temperature in the hydrothermal treatment during synthesis
remarkably inuences the crystalline structures and morphology of CuO nanorods. Similarly, Cu(OH)
2
was prepared from NaOH and CuCl
2
and dispersed in NaOH solution. Then, the mixture was trans-
ferred into a Polytetrauoroethylene (PTFE) container in an autoclave and kept at room temperature
at 48 h. The temperature was then increased to 100 C for another 48 h. After washing and drying, the
nal products were collected. The ne CuO nanorods prepared at room temperature have higher as-
pect ratio and smaller diameters compared with the bulk CuO nanorods prepared at 100 C (Fig. 3).
However, the SAED images show that the ne CuO nanorods are polycrystalline, in contrast to the
monocrystalline bulk CuO nanorods.
Shrestha et al. [80] used the following chemical combinations to synthesize CuO nanorods by the
hydrothermal method: (1) Cu(NO
3
)
2
, lactic acid, and NaOH; (2) CuSO
4
, sodium lactate, and NaOH; and
(3) Cu(NO
3
)
2
and NaOH. The chemical reagents were mixed and stirred, and then transferred into an
Fig. 4. Scanning electron microscope (SEM) images of CuO nanorods (a), (b), and (c) synthesized with chemical combinations
(1), (2), and (3), respectively [80]. Copyright 2010 American Chemical Society.
214 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
quay ly tam
autoclave for 24 h at 140 C. The as-synthesized CuO nanorods have similar morphologies (Fig. 4).
However, unlike the other two cases, the CuO nanorods synthesized from combination (1) assembled
in a spherical structure without separation of individual nanorods (Fig. 4a) The CuO nanorods
achieved from combination (2) separated, but still tended to aggregate. Interestingly, some of the
CuO nanorods obtained from combination (3) show rectangular cross-sections (Fig. 4c). Although
the chemical combinations were different, the shapes of the CuO nanorods did not remarkably change.
However, when the concentration of NaOH, aging period, and temperature were altered, the resulting
products varied from plate-like structure to octahedral structures instead of nanorods.
Dar et al. [81] reported a hydrothermal method of synthesizing CuO nanoneedles by the reaction
between Cu(NO
3
)
2
3H
2
O and NaOH under continuous stirring followed by heating from 120 C to
180 C for 20 h to 60 h. The CuO nanoneedles had very sharp tips and large-diameter bottoms. In
Fig. 5. (a) Schematic representation of the synthesis of CuO using alcohols, EG, and nonionic polymeric surfactants [83].
Copyright 2011 Elsevier. (b) The schematic growth of CuO nanostructures with diverse morphologies [47]. Copyright 2012
Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 215
addition, the nanoneedles were grown by surfactant-free approach. Yang et al. [82] recently
introduced microwave-assisted hydrothermal method for the synthesis of CuO nanorods. Similarly,
CuSO
4
5H
2
O, polyethylene glycol 400 (PEG-400), and urea were dissolved and heated to 80 C in an
ultrasonic bath. NaOH was added to the solution and aged for 5 min under the same condition. Then,
the mixture was transferred to an autoclave, sealed, and treated in the microwave digestion system at
120 C for just 20 min to obtain the nal CuO nanorods. The size of the as-synthesized CuO nanorods
was smaller than those from normal hydrothermal method. During the microwave-assisted
hydrothermal stage, CuO nanorod formation just required 20 min.
Fig. 6. Typical SEM images of (a) buttery-like [90] (Copyright 2011 Indian Academy of Sciences), (b) gear wheel bundles [105]
(Copyright 2007 American Association of Nanoscience and Technology), (c) ower-like assemblies [32] (Copyright 2007
American Chemical Society), (d) dendrite-like [101] (Copyright 2007 Elsevier), (e) nanobat-like [108] (Copyright 2009 Elsevier),
(f) layered hexagonal discs [112] (Copyright 2011 Institute of Physics), (g) honeycomb-like [32] (Copyright 2007 American
Chemical Society), (h) self-assembled leaf-like [99] (Copyright 2010 Springer), (i) hierarchical peachstone-like [109] (Copyright
2010 Elsevier), (j) shrimp-like [47] (Copyright 2006 American Chemical Society), (k) sheaf-like [111] (Copyright 2009 Elsevier),
and (l) urchin-like microspheres [104] (Copyright 2009 Elsevier).
216 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Table 1
Summary of CuO nanostructures obtained hydrothermally and the various synthesis conditions [PEG, sodium dodecyl sulfate
(SDS), CTAB, EG, sodium dodecylbenzenesulfonate (SDBS), dodecylsulfate (DS), hexamethylenediamine (HMDA), poly-sodium
4-styrenesulfonate (PSS)].
Morphology Size (nm) Starting materials Additives Temperature
(C)
Duration
(h)
Flower-like, boat-
like, plate-like,
and ellipsoid-like
[85]
Various Cu(NO
3
)
2
3H
2
O PEG 100180 0.52
NH
3
H
2
O
Honeycomb-like
[32]
90100 nm in diameter, tens of
micrometers long
Cu foil Na
2
WO
4
,
Na
2
MoO
4
,
SDS
160 24
Nanoplatelets [9698] 400 nm to 2 mm in width and 1 mm to
several micrometers long
Cu
x
S 150, 200, 230 1.5
NaOH
1 lm in width Cu foil H
2
O
2
150 12
120300 nm in width and 480700 nm
long
Cu(DS)
2
120 12
NaOH
Leaf-like [99] 0.51.5 lm in length, 80110 nm long Cu(OH)
2
Urea 150 612
Momordica-like
[100]
Less than 100 nm long and 3050 nm in
width
CuSO
4
5H
2
O 180 15
NH
3
H
2
O
Dendrite-like [101] About 100 nm in width and several
micrometers in length
CuSO
4
5H
2
O EG 200 20
NaOH
Branch-like and
ake-like [102]
50200 nm width and 300400 nm
length for ake-like nanostructure;
hundreds of nanometers long and
diameters of 20100 nm for branch-like
nanostructures
CuSO
4
5H
2
O Sodium
citrate
160 12
Spherical-like [103] 500 nm in diameter Cu(CH
3
COO)
2
Urea 121 0.67
Hierarchical hollow
microspheres [87]
3.5 lm in diameter Cu(CH
3
COO)
2
H
2
O 120 24
Urchin-like [104] 3 lm in diameter CuCl
2
2H
2
O EG 100 12
Gear wheel and
clew like [105]
10 lm in diameter for gear wheel
nanostructure
Cu(NO
3
)
2
3H
2
O PEG 150180 10, 21
Flower-like
[91,106]
2.53 lm long Cu threads K
2
Cr
2
O
7
,
H
2
SO
4
140 12
6 lm in diameter CuCl
2
2H
2
O CTAB 150 12
Hierarchical
ower-like [107]
510 lm in diameter CuSO
4
5H
2
O H
2
O
2
120 6
NaOH
Nanobat-like [108] 70 nm in diameter, 170 nm long Cu(NO
3
)
2
3H
2
O Urea 100150 615
Hierarchical
buttery-like [90]
6 lm long and 24 lm in width CuCl
2
2H
2
O SDBS 80150 115
Nanobundles [89] 2030 nm in diameter and 350500 nm
long
CuCl
2
2H
2
O SDBS 130 1824
Peachstone-like
[109]
4 lm long and 3 lm in width CuCl
2
2H
2
O [Omim]TA 100 24
Hollow micro/
nanostructures
[110]
1.53 lm in diameter CuSO
4
Tyrosine 130 4
NH
3
H
2
O
Shuttle-like [88] 200300 nm in width and several
micrometers long
Cu(CH
3
COO)
2
H
2
O CTAB 120 12
Sheaf-like [111] 2 lm long Cu(NO
3
)
2
3H
2
O Urea 120 8
Layered hexagonal
discs [112]
34 lm long and 1.52.0 lm in width Cu(NO
3
)
2
3H
2
O HMDA 130 310
NH
3
H
2
O
Hierarchical
dandelion-like
microspheres [113]
36 lm in size Cu(NO
3
)
2
3H
2
O EG 130 16
NH
3
H
2
O
NaOH
Twinned-
hemisphere-like
[114]
22.5 lm in size Cu(CH
3
COO)
2
H
2
O PSS 180 2
NaF
Hierarchical
dendrite-like
[115]
69 lm in size Cu(NO
3
)
2
3H
2
O Urea 160 12
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 217
2.1.1.3. Synthesis of CuO 2D/3D nanostructures. The simplicity of the hydrothermal method facilitated its
performance at low temperatures and in a large scale. Additionally, it can be used for the production of
complex nanostructured CuO with diverse morphologies and sizes. In most cases, the hydrothermal
synthesis of CuO nanostructures starts with the formation of intermediate compound Cu(OH)
2
precip-
itationor other intermediate phases, whichare preparedfromcupric salt/cupric acids/Cufoil precursors
in alkaline media. Then, the solution is kept at an elevated temperature in an autoclave for a certain per-
iod, allowing the decomposition of Cu(OH)
2
or other intermediate phases into the nal product of CuO.
Therefore, the manipulation of well-controlled synthesis of CuO nanostructures with various shapes is
possible by choosing different solutions and by adjusting the concentrations of precursors. To obtain di-
verse morphologies and dimensions of CuO nanostructures, several surfactant and structure-directing
agents are normally added into the precursor solution. Systematic studies of the experimental param-
eters reveal that the Cu source, reaction temperature, reaction time, and surfactant along with the pH
value of the precursor solution inuence the morphology, growth, size, and dimensions of the resulting
CuO nanostructures (Fig. 5) [25,47,83,84].
The various morphologies of the CuO nanostructures achieved using the hydrothermal technique is
shown in Fig. 6. The obtained CuO nanostructures exhibit diverse optical, electrical, and catalytic prop-
erties [8587].
The hydrothermally synthesized CuO nanostructures fabricated with and without additives exhibit
evidently different morphologies and properties. However, the exact mechanism for the formation of
these architectures by the addition of inorganic or organic materials has not been fully understood. In
terms of organic additives, several reports have suggested that certain ions in the additives are ad-
sorbed on the CuO surface, which alters the growth mechanism [78,8890]. Other studies have dem-
onstrated that additives can act as a template for the formation of CuO nanostructures with different
morphologies [83,91]. The addition of basic media such as NaOH, urea, NH
3
H
2
O, and (CH
2
)
6
N
4
leads to
the formation of intermediate compound Cu(OH)
2
that is transformed into CuO under heat treatment
by the oriented attachment growth mechanism as detailed in Section 3. The nal morphology of CuO
was determined by the internal crystallographic structure of Cu(OH)
2
. Cudennec et al. [92] showed
that the transformation of Cu(OH)
2
into CuO is a reconstructive transformation involving a dissolution
reaction followed by the precipitation of CuO according to the following scheme: CuOH
2
2s

2OH

aq
!CuOH
2
4aq
$CuO
s
2OH

aq
H
2
O. Moreover, the application of microwave power
during the hydrothermal process to synthesize CuO nanostructure has emerged as an important topic
in the scientic community because of its low energy consumption, rapid heating process, and fast
kinetics of crystallization [93]. The use of hydrothermal microwave to synthesize CuO nanostructure
leads to small diameter at short annealing time with high yield compared with the conventional
hydrothermal process [94,95].
In summary, various CuO nanostructures, including CuO nanoparticles, 1D CuO nanowires/rods/
tubes, 2D, and several complex 3D CuO nanostructures, have been hydrothermally synthesized by
adding various reactants to the cupric salt/cupric acids/Cu foil precursor solution in the presence of
some capping agents. Moreover, morphologies of the synthesized CuO nanostructures can be con-
trolled by selecting certain types of structure-directing and dispersing modifying agents. A brief sum-
mary of the obtained CuO nanostructures that are hydrothermally synthesized with or without
different additives is shown in Table 1.
The possible growth scheme for the formation of rectangular-shaped nanobat-like CuO nanostruc-
tures through a typical hydrothermal synthetic process is shown in Fig. 7 [108]. The reagents are
Cu(NO
3
)
2
3H
2
O and urea, where urea acts as pH buffer to control the supply of OH

ions. During
the initial stage, water reacts with urea to form ammonia and CO
2
. In addition, ammonia further reacts
with water to produce ammonium and OH

ions. The reaction slowly generates solid building units


into solution (b). The concentration of solid building units in the solution increases continuously
and nucleation starts (c) after a crucial super-saturation level is reached. The crystallographic struc-
tures of the nuclei and seeds greatly affect the nal shapes of the nanocrystals (d). The variation in
the CuO nuclei is caused by the increasing pH of the reaction solution (c). The small CuO crystal
increases and forms a bigger crystal with increasing reaction time. O

and Cu
2+
are stacked alternately
along with the specic directions to form alternating planes, which causes the anisotropic growth of
the CuO crystal. The different growth rates are due to the crystallographic faces, where the growth rate
218 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
da dang
su thao tac
is sequenced as [010]
length
> [100]
breadth
> [001]
height
. The slowest growing (010) planes dominate the
crystal into the typical rectangular-shaped structure (e). The formed nanobat-like structure then
forms a circular fashion via self-assembly to build hollow microspheres. This phenomenon was caused
by the formation of several unbalanced charge centers as a result of the dissimilarities in the surface
charges. These charge centers attract the nanobats in the solution through rotating adjacent structure
(e) to share identical orientation. Finally, sphere-like CuO microsphere assemblies are formed with
increasing reaction time.
2.1.2. Solution-based chemical precipitation methods
Chemical precipitation synthesis is similar to the hydrothermal method with a reaction also occur-
ring in the solution, but the chemical reactions can be conducted in an open container with a relatively
low reaction temperature (normally below 100 C). This process can be simply dened as the chemical
reaction between the precursors to produce monomers that subsequently aggregate into nal result-
ing materials [69]. A schematic drawing of the solution-based chemical precipitation to produce
nanoscale CuO is shown in Fig. 8.
2.1.2.1. Synthesis of CuO nanoparticles. Cupric salt (normally nitrate or sulfate) and alkaline compounds
(normally NaOH) are often used in the synthesis of CuO nanoparticles. A typical process of chemical
precipitation is as follows [116]: Cu(NO
3
)
2
solution (300 mL 0.02 M) is prepared by dissolving
Cu(NO
3
)
2
3H
2
O in deionized water. The solution is placed into a round-bottom ask equipped with
a reuxing device. The Cu(NO
3
)
2
solution is kept at an appropriate temperature (6100 C) with vigor-
ous stirring. Then, 0.5 g solid NaOH (platelet) is rapidly added into the solution, resulting in the pro-
duction of a large amount of blue or black precipitates and the crystallization temperature is
maintained for 10 min. Next, the precipitates are heated at 100 C for another 10 min. After complete
reactions, the resulting products are centrifuged, washed with water and ethanol for several times,
Fig. 7. Schematic illustration of the growth scheme for the formation of rectangular-shaped nanobat-like CuO nanostructures
by the hydrothermal synthesis method [108]. Copyright 2009 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 219
and then dried in air at room temperature. Zhu et al. [116] synthesized CuO nanoparticles by this
route. The average diameter of the as-prepared particles is 10 nm.
The proportion of reactive compounds and reactive temperature markedly affects the size and mor-
phology of the synthesized CuO nanoparticles. Using a similar method and three sets of recipes, Zhou
et al. [52] obtained three types of CuO products with distinct sizes and morphologies: nanoparticles,
nanobelts, and nanoplatelets (Fig. 9). The recipes were (a) Cu(NO
3
)
2
3H
2
O + NaOH, (b) Cu(OAc)
2
H
2
O
+ NaOH, and (c) Cu(NO
3
)
2
3H
2
O + Na
2
CO
3
.
The synthesized CuO nanoparticles using ordinary chemical precipitation methods usually have a
common problem, that is, the achieved nanoparticles are apt to agglomerate. To solve this limitation,
chemical precipitation methods have been extensively investigated to improve the separation of the
nanoparticles through the application of some external energy such as ultrasonic or high pressure. The
sonochemical method is used to apply ultrasonic cavitation during the synthesis procedure. Kumar
et al. [48] synthesized well-separated 26 nm CuO nanoparticles by irradiating the reactive solution
with a high-intensity ultrasonic horn under 1.5 atm of argon at room temperature for 3 h. However,
this method requires expensive apparatus and excessive organic solvent as well as severe reaction
conditions. In addition to the sonochemical synthesis of dispersed CuO nanoparticles, Zhu et al.
[117] developed a simple quick-precipitation procedure to prepare highly dispersed CuO nanoparti-
cles with the size of about 6 nm in aqueous solution. The authors noted that the tendency of CuO
nanoparticles to aggregate during preparation may have been caused by the low nucleation and
growth rates of CuO particles at mild reaction condition. A large amount of well-dispersed CuO nano-
particles were obtained by the rapid addition of the NaOH solid to the mixture of an aqueous solution
of Cu(CH
3
COO)
2
and a glacial acetic acid at 100 C. This result indicates that higher temperatures cause
Fig. 8. Schematic of a typical chemical precipitation synthetic process for CuO nanostructures.
Fig. 9. TEM images of CuO (a) nanoparticles, (b) nanobelts, and (c) nanoplatelets [52]. Copyright 2008 Institute of Physics.
220 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
higher reaction rates, resulting in large amounts of nuclei to form in a short period and the inhibition
of the aggregation of crystals. Consequently, well-dispersed CuO nanoparticles with small sizes were
achieved at relatively high reaction temperature. However, Mahapatra et al. [118] introduced a wet
chemical method to synthesize ultrane dispersed CuO nanoparticles. The authors claimed that the
size of as-prepared CuO nanoparticles can be controlled by simply changing the concentration of
the basic Cu precursor. In addition, hot-solution decomposition and microemulsion process, which
have been widely applied to synthesize well-dispersed quantum dots, may be applied in the prepara-
tion of dispersed CuO nanomaterials with various shapes and sizes. Other synthetic technologies such
as by introducing sonic waves or microwaves in the synthetic process can also be tried to obtain well-
dispersed CuO nanoparticles in the future [119].
2.1.2.2. Synthesis of CuO 1D nanostructures. Wang et al. [120] reported a simple wet-chemical method
for preparing CuO nanorods. In a typical process, a non-ionic surfactant, polyethylene glycol (PEG; Mw
20,000), and CuCl
2
2H
2
O are dissolved in water. After stirring for 15 min, NaOH is added into the solu-
tion to generate Cu(OH)
2
precipitate. The Cu(OH)
2
is placed into a steam trace for 30 min and is trans-
formed into CuO precipitate, followed by washing, ltering, and drying. The as-synthesized CuO
nanorods have monoclinic structure up to 400 nm long with a diameter ranging from 5 nm to
15 nm. Lu et al. [121] synthesized CuO nanowires by dehydration of the precursor of Cu(OH)
2
nano-
wires. First, KOH solution was added dropwise into CuSO
4
solution under rigorous stirring, followed
by the dropwise addition of ammonia solution. Then, the Cu(OH)
2
nanostructures were heated at
120 C for 2 h and at 180 C for another 3 h. The resulting CuO nanowires have structures similar to
those of the precursor Cu(OH)
2
nanowires. Ethiraj and Kang [122] introduced an organic molecule thi-
oglycerol (TG) as a stabilizer in the synthesis of CuO nanowires. TG was rst added in a copper acetate
solution under stirring. Then, the NaOH solution was added dropwise into the mixture, followed by
immediate addition of water under continuous stirring for a few minutes. The precipitate was col-
lected by centrifuging, washing and overnight drying. By comparing the samples with and without
the use of TG, the authors found that a small amount of TG led to well-dispersed CuO nanowires,
otherwise the CuO nanowires would assemble into a ower-like structure. Zhang et al. [44] reported
the preparation of CuO nanobelts by adding ammonia to a Cu(NO
3
)
2
solution directly under constant
stirring for 15 min. After the formation of a blue precipitate, the mixture was then sealed and heated
at 60 C for 4 h. The product was collected by repeated washing and centrifugation. The widths and
lengths of the as-obtained CuO nanobelts were 510 nm and 13 lm, respectively. Moreover, the
thickness of the CuO nanobelts was estimated to be 25 nm. Interestingly, Yangs group [123] synthe-
sized vertically aligned CuO nanorod arrays on a Cu substrate. In a typical synthesis, a Cu foil reacted
with NaOH under the presence of some oxidants or surfactants. Then the foil was removed, washed,
and dried. The CuO nanorod arrays were found to uniformly cover the surface of the Cu foil.
Remarkably, by using PEG 200 as the capping agent, ultralong CuO nanowire bundles with lengths
ranging from tens to hundreds of micrometers as shown in Fig. 10a were selectively synthesized on a
large scale at room temperature by a facile solution-phase method by Li et al. [124]. Transmission
electron microscope (TEM) characterizations (Fig. 10b) demonstrated that the obtained CuO nanowire
bundles are polycrystalline. Moreover, a series of controlled experiments performed by the authors
revealed that the presence of PEG 200 and the concentration of OH

affected the morphology and


phase control of CuO nanostructures during the reaction process. No CuO nanowire bundles can be
obtained, but CuO nanoleaves (NLs), without PEG 200 (Fig. 10c). These NLs were single crystals and
grew along the [111] crystal plane with diameters ranging from 200 nm to 500 nm (Fig. 10d). The
CuO nanowire bundles could only be synthesized when the molar ratio of OH

/Cu
2+
was higher than
four, indicating that PEG200 and the concentration of OH

are essential in the formation of CuO nano-


wire bundles.
Wang et al. [125] synthesized large-scale ultralong CuO nanowires with an average diameter of
8 nm and lengths of up to several tens of micrometers by a facile room temperature solution-phase
chemical route without any capping agent, as depicted in Fig. 11. Cu(OH)
2
nanowires were rst formed
and subsequently served as template to direct the formation of CuO nanowires. Compared with the
commercial CuO powders, the obtained ultralong CuO nanowires exhibit enhanced photocatalytic
activity for RhB degradation and the potential for applications in LIBs and catalysis.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 221
2.1.2.3. Synthesis of CuO 2D/3D Nanostructures. The solution-based chemical precipitation method can
also be used to synthesize more complex 2D/3D CuO nanostructures. A brief summary of the obtained
CuO nanostructures prepared by the solution-based chemical precipitation method with different
additives are listed in Table 2. Some examples of CuO nanostructures with various morphologies
achieved by this technique are shown in Figs. 13 and 14. In the whole synthetic process, two key
points, namely, nucleation and growth, dominate the formation of different CuO nanoarchitectures.
An appropriate precursor, a rational condition for the reaction system together with additional ligands
to adjust the surface energies will signicantly inuence the nucleation and growth of the CuO
Fig. 10. (a) SEM image of the CuO nanowire bundles. (b) TEM image and the corresponding SAED pattern taken in a zone rich in
nanowires. (c) SEM image and TEM image (inset). (d) High-resolution (HRTEM) image of one CuO NL and the corresponding
SAED pattern, indicating growing along the

111 direction [124]. Copyright 2010 Elsevier.


Fig. 11. (a) TEM image of CuO nanowires. (b) A histogram of the CuO nanowires, revealing the diameter distribution [125].
Copyright 2012 Royal Society of Chemistry.
222 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
nanostructures. Thus, to achieve the controlled synthesis of CuO nanoarchitectures by solution-based
chemical precipitation, various experimental parameters including the type of precursors, reaction
temperature, reactant concentration, and surfactant can be manipulated to control the morphology
and structure of the obtained CuO nanocrystals. Moreover, recent reports have determined that the
pH value together with concentration of OH

ions can signicantly affect the nucleation and growth


behavior (for example, the number of nuclei and concentrations of growth units) of the CuO crystals,
resulting in the formation of the diverse morphologies of CuO nanostructures as shown in Fig. 12. A
similar phenomenon is also observed in controlling the hydrothermal synthesis of CuO
nanostructures.
In addition, the various additives used in the preparation process have distinct functions. Several
studies have suggested that ionic surfactants, such as SDBS and CTAB, can be absorbed onto the sur-
faces of CuO nanocrystals [130,131]. This effect can reduce the interfacial tension between the crys-
tallizing phase and the surrounding solutions, which strongly affects the growth rate and
orientation of the crystals, resulting in the formation of CuO nanostructures with diverse morpholo-
gies. Non-ionic surfactants, polyvinylpyrrolidone (PVP) and PEG were thought to serve as templates
for the formation of CuO nanostructures. However, the mechanism responsible for the growth of
CuO nanostructures by solution-based chemical precipitation method remains unclear. Some
researchers have used oriented attachment as the possible statistical growth mechanism as detailed
in Section 3 [128,138].
The schematic growth mechanism for the formation of ower-shaped CuO nanostructures through
a typical solution-based chemical precipitation process is shown in Fig. 15 [33]. When Cu nitrate is
Table 2
Summary of CuO nanostructures obtained by solution-based chemical precipitation methods and their synthesis conditions (PEG,
CTAB, SDBS, DS, HMDA, and PVP).
Morphology Starting Materials Additives Temperature
(C)
Duration
(h)
Nanowires, rectangles, and seed-, belt- and sheet-
like [132]
Cu(NO
3
)
2
3H
2
O 80 0.5
NaOH
Urchin-like [51] Cu(NO
3
)
2
3H
2
O Urea 100 6
Flower-shaped [33] Cu(NO
3
)
2
3H
2
O HMTA 100 3
NaOH
Belt-, bamboo leaf- and shuttle-like [130] CuCl
2
2H
2
O PEG, CTAB 75 0.5
NaOH
Leaet-like nanosheet [133] CuSO
4
5H
2
O 25 2
KOH, NH
3
H
2
O
Flower-like hierarchical assemblies [134] Cu(NO
3
)
2
3H
2
O 25, 100 1272
NaOH
Carambola-like [135] Cu(NO
3
)
2
3H
2
O HMTA 95 3
NaOH
Shuttle- and ower-like [136] Cu(NO
3
)
2
3H
2
O HMTA, CTAB 100 3
NaOH
Nanoribbons and nanorings [137] CuCl
2
2H
2
O SDBS 120 0.5
NaOH
NLs [138] CuCl
2
2H
2
O 35 40
NH
3
H
2
O
Hierarchical nanochains [139] CuCl
2
2H
2
O PVP,
(NH
4
)
2
SO
4
100 12
NaOH
Hierarchical nanosheets [140] Cu(CH
3
COO)
2
H
2
O 30 72
NH
3
H
2
O
Oval nanosheets and nanoellipsoid[128] Cu(CH
3
COO)
2
H
2
O 65 24
NH
3
H
2
O
Flower-like microspheres [141] Cu power, NaOH (NH
4
)
2
S
2
O
8
25 20
Lenticular-like, pseudo-like and elliptical [84] Cu(CH
3
COO),
NaOH
25 24
Spindle-like and plate-like [142] Cu(NO
3
)
2
, NaOH 80 0.5
Flower-like hierarchical nanostructures [143] Cu(NO
3
)
2
3H
2
O 80 4
NH
3
H
2
O
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 223
mixed only with HMTA (a), a turbid solution containing the building units (b) is obtained. After the
turbid solution is reuxed at 100 C for >1 h, Cu(OH)
2
nuclei are formed from the origination of
OH

ions by the hydrolyzation of HMTA, and then nuclear aggregation occurs (c). With increasing
reaction time, the Cu(OH)
2
is converted into small CuO crystals. Moreover, the CuO crystals are ar-
ranged with time and nally form petal-like structures (d). Interestingly, if the Cu(NO
3
)
2
is mixed with
both HMTA and NaOH, the solution immediately changes to blue color because of the instant forma-
tion of Cu(OH)
2
nuclei (e) caused by the fast formation of OH

ions from NaOH. The Cu(OH)


2
nuclei are
transformed into CuO through a simple chemical reaction of Cu(OH)
2
?CuO + H
2
O. With increasing
reaction time, the initially formed CuO nuclei are assembled, forming individual petals (f) and nally
ower-like morphologies (g).
As mentioned in Section 2.1, the solution-based synthesis of CuO nanostructures is commonly per-
formed in solutions containing Cu(II) salts, and the obtained nal products are normally free-standing.
In addition, this simple method can also be used to synthesize CuO nanostructures directly on the sur-
face of Cu substrates. Generally, Cu substrates are washed in an HCl solution for a few minutes and
subsequently rinsed with deionized water and absolute ethanol to remove the surface impurities
along with oxide layers. Then, the treated Cu substrates are immersed into alkaline solutions with
or without additives of oxidative reagents (e.g., (NH
4
)
2
S
2
O
8
or K
2
S
2
O
8
) at certain temperature for a
xed period [129,144147]. The reaction system without additives of oxidative reagents commonly
requires a longer time to yield the nal CuO nanostructures on the Cu surface [146,147].
In 2003, Yang et al. [129] synthesized sheet- and whisker-like CuO nanostructures on a Cu sub-
strate surface by a simple liquidsolid reaction under alkaline and oxidative conditions at room tem-
perature. They demonstrated an evolution of the lm structures as a function of the solution
treatment time and the concentration of NaOH aqueous solution from the bers of Cu(OH)
2
onto
Fig. 12. Schematic illustration of the growth of CuO nanostructures under different pH conditions and the corresponding SEM
images [127]. Copyright 2010 Elsevier.
224 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
the scrolls of Cu(OH)
2
onto the sheets or whiskers of CuO. The formation of Cu(OH)
2
and CuO nano-
structures on Cu surfaces involved inorganic polymerization (polycondensation) reactions. The ob-
tained CuO nanostructures are phase-pure single crystallites. This method for the preparation of Cu
compound lms with ultrane structures is advantageous because of its simplicity, high yield, and
mild reaction conditions. It can also produce single crystal nanomaterials in array form, which offers
an attractive and convenient path to the large-scale engineering of ordered inorganic nanostructures
on metal electrodes.
Large-area ower-like 3D CuO nanostructures were also successfully synthesized on a Cu surface
by a template-free solution route under alkaline and oxidative conditions at 70 C [145]. The forma-
tion of ower-like nanostructures was found to strongly depend on the concentration of oxidant
K
2
S
2
O
8
. Liu et al. [146] reported the fabrication of CuO hierarchical nanostructures on Cu substrates
by the oxidation of Cu in alkaline conditions at 60 C without the addition of oxidative reagents. They
Fig. 13. Typical SEM images of (a) urchin-like nanostructure[51] (Copyright 2009 American Chemical Society), (b) hierarchical
nanochains [139] (Copyright 2011 Elsevier), (c) Nanosheets [128] (Copyright 2006 American Chemical Society), (d) ower-like
[32] (Copyright 2008 American Chemical Society), (e) ower-like assembles [134] (Copyright 2007 Elsevier), (f) nanoellipsoid
[128] (Copyright 2006 American Chemical Society), (g) spindle-like [142] (Copyright 2012 Royal Society of Chemistry), (h)
microplates [84] (Copyright 2012 Royal Society of Chemistry), (i) ower-like microspheres [141], (Copyright 2012 Royal Society
of Chemistry), (j) lenticular-like [84] (Copyright 2012 Royal Society of Chemistry), (k) Pseudo-like [84] (Copyright 2012 Royal
Society of Chemistry), and (l) elliptical-like nanostructure[84] (Copyright 2012 Royal Society of Chemistry).
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 225
demonstrated that CuO ower-like structures composed of hierarchical 2D nanosheets and spherical
architectures constructed by ultrathin nanowalls could be selectively generated by controlling the
alkaline reactant. When NaOH was used, 3D CuO ower-like structures on a Cu foil was obtained
Fig. 14. Typical TEM images of (a) sheet-like [136] (Copyright 2009 Springer), (b) ower-like [136] (Copyright 2009 Springer),
(c) NLs [138] (Copyright 2007 American Chemical Society), (d) spindle-like [142] (Copyright 2012 Royal Society of Chemistry),
(e) lenticular-like [84] (Copyright 2012 Royal Society of Chemistry), and (f) elliptical-like nanostructures [84] (Copyright 2012
Royal Society of Chemistry).
Fig. 15. Schematic growth mechanism for the formation of ower-shaped CuO nanostructures [33]. Copyright 2008 American
Chemical Society.
226 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
(Fig. 16a). Whereas NaOH was replaced by NH
3
H
2
O, CuO with 3D spherical architectures was formed
on the Cu foil as illustrated in Fig. 16b. In addition, by reducing the concentration of the alkaline solu-
tion, well-dened 2D nanosheets and nanowall arrays can be fabricated on a large scale on the Cu foil
accordingly. A new hierarchical CuO microcabbage architecture [147] consisting of densely packed
nanoplates and nanoribbons was also directly fabricated on Cu foils via a similar synthetic process
at room temperature for 6d. In addition, CuO nanostructures such as nanoneedles, nanoowers, and
stacking of ake-like structures on Cu foils can also be obtained by Cu oxidation under alkaline con-
dition by a simple wet chemical route [144].
2.2. Solid-state thermal conversion of precursors
Various CuO nanostructures can be generated through the thermal conversion of the precursors,
and the morphological features of the precursors can be well-preserved in the nal products given
that heat treatment is appropriately performed. Cu(OH)
2
and basic Cu salts have become favorable
precursor candidates to exploit the interesting morphologies of CuO because of their unique and
well-known layered structure [148,149]. The process usually starts with the synthesis of the Cu
precursors via the reaction of cupric salt (normally nitrate or chloride) with alkaline compounds
(normally NaOH). The obtained cupric precursors are then centrifuged and washed with distilled
water and absolute ethanol. Finally, these cupric precursors are calcined in solid state to obtain the
nal CuO nanostructures. Moreover, the corresponding Cu(OH)
2
particles are formed and precipitated
in H
2
O by adding a basic solution (usually NaOH) to the obtained cupric precursors solution. The
resulting nitrate or chloride salts are then washed away, and the corresponding Cu(OH)
2
particles
are thermally dehydrated after ltration and washed to obtain the nal CuO nanostructures.
This method is similar to the solution-based chemical precipitation method, but the thermal dehy-
dration of cupric precursors, such as Cu
2
(OH)
2
CO
3
, Cu
2
Cl(OH)
3
, CuC
2
O
4
, and Cu(OH)
2
is in solid state
with relatively higher treatment temperature. Additionally, the morphological features of the cupric
precursors can be well-preserved in the nal CuO nanostructures. The mechanism of shape-reserved
transformation from Cu(OH)
2
nanowires to CuO nanowires as illustrated in Fig. 17 clearly demon-
strates that the morphology can be properly reserved during the transformation from Cu(OH)
2
to
CuO because of the topotactic transformation in the dehydration process [151].
In addition, a schematic diagram of the metamorphosis of CuO as given by Dey et al. [152] illus-
trates the formation process (Fig. 18). This diagram presents a simple depiction of the transformation
that occurs at the atomic level for the rst time.
Fig. 16. SEM images of (a) 3D CuO ower-like structures on Cu foil and (b) 3D spherical CuO architectures on Cu foil [146].
Copyright Royal Society of Chemistry 2006.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 227
However, synthesis parameters such as alkaline content (pH) of the solutions during hydrolysis
play a key role in the formation of various morphologies and dimensions of the Cu(OH)
2
nanostruc-
tures, leading to a series of fascinatingly shaped nanostructures of nal CuO by subsequent heat
treatment.
Dey et al. [152] reported that different morphologies of CuO nanostructures have been synthesized
by a simple chemical route, in which the Cu(OH)
2
nanostructures are rst synthesized, and the
precipitate is subsequently annealed at 130 C. A variety of shapes such as seed-like, ellipsoidal, rods,
and leaves of CuO can be obtained by simply varying the pH value during synthesis (Fig. 19), indicating
that alkaline content (pH) of the solutions are essential in the formation of various CuO
nanostructures.
Fig. 17. (a) Crystal structure of Cu(OH)
2
. (b) When the temperature rises, long CuO and HO bonds break. The structure of
Cu(OH)
2
changes to a layered Cu(OH)
4
structure. (c) Projection along the a-axis for (b), (d), and (e). The dehydration process in
the b, c plane of Cu(OH)
2
. (f) Crystal structure of CuO [151]. Copyright 2012 Royal Society of Chemistry.
228 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Fig. 18. Schematic representation of CuO formation at the molecular level from a crystallographic perspective. The blue balls
represent Cu atoms, and the red balls represent O atoms. (a) FCC lattice of Cu and (b) monoclinic cell of CuO. (c) Two unit cells
joined by a common face, showing the stacking of the 111 planes [152]. Copyright 2012 Royal Society of Chemistry.
Fig. 19. Diagram of the formation of various shapes of CuO under different synthesis conditions [152]. Copyright 2012 Royal
Society of Chemistry.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 229
With the method mentioned above, CuO with different morphologies such as nanoribbons
[153], hierarchical sphere-like [154], peanut-shaped superstructures [150], sisal-like [148], sh-
bone-like [155], pillow-shaped [156], perpendicularly cross-bedded microstructure [157], NC
line-assembled bundle-like [158], buttery-sheet-like or nanotubes [159], and worm-like [160]
were obtained through thermal conversion of their corresponding Cu salts. In addition, other
CuO nanostructures, including CuO nanoribbon arrays, CuO nanotube arrays, nanotube arrays with
special nanoplate wall structure, and quasi-aligned submicrometer CuO ribbons formed on Cu foils,
achieved by heating the corresponding Cu(OH)
2
precursors have also been reported [161163].
Chen et al. [164] fabricated novel hierarchically mesoporous nanosheet-assembled gear-like pillar
CuO arrays directly grown on Cu foil by the vapor-phase corrosion approach and subsequent heat
treatment [164]. The hierarchically nanosheet-assembled gear-like pillar arrays (HNGPAs) of
Cu(OH)
2
were rst formed on Cu foil. Hierarchical micro/nanostructure CuO was obtained by the
solid-state thermal transformation of the formed Cu(OH)
2
arrays on Cu foil. The obtained CuO
nanostructures were found to inherit the intact hierarchical superstructures with retained radial
symmetry and nanosheets subcomponents from Cu(OH)
2
HNGPAs without collapse and aggrega-
tion (see Fig. 20).
Novel CuO mesoporous nanosheet cluster arrays that are directly grown on a Cu substrate via the
same ammonia vapor-phase corrosion route were also found by Chen et al. [165]. This simple and
effective fabrication strategy shows promising potential for the preparation of other nanoarchitec-
tured materials for both high-energy and high-power applications.
The advantages of solid-state thermal conversion of precursors are simple, easy and safe to use,
controllable, allows the mass-production of CuO nanostructures with unique superstructures, practi-
cal, and promising for various applications ranging from catalytic reaction to sensing.
Fig. 20. (a) Schematic illustration of the fabrication of CuO HMNGPAs via a vapor-phase corrosion route; SEM images of
Cu(OH)
2
HNGPAs and (c) CuO HNGPAs; and (d) schematic of the growth process of Cu(OH)
2
HNGPAs [164]. Copyright 2012
American Chemical Society.
230 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
2.3. Electrochemical method
Electrochemical method is widely used for the preparation of nanoporous MOs because of its sim-
plicity, low-temperature operation process, and viability of commercial production. This method is
also advantageous because the growth orientation, morphology, and size of the resulting products
can be modestly controlled by adjusting the deposition parameters (deposition voltage, current den-
sity, temperature, etc.). The setup of an electrochemical deposition facility is illustrated in Fig. 21.
A typical fabrication process for the electrochemical synthesis of CuO nanostructures with different
shapes at room temperature was described by Yuan et al. [166]. In this fabrication, a two-electrode
system was used with Cu plates as the anode and stainless steel plate as the cathode. The electrolyte
contained 1 M NaNO
3
dissolved in distilled water. Systematic studies of the experimental parameters
such as electrolyte, composition of electrolyte solution, and current density reveal that the composi-
tion of electrolyte solution and the use of current density during synthesis remarkably inuence the
size and morphology of the resulting CuO nanostructures. When H
2
OEtOH mix was used as an elec-
trolytic solvent, CuO nanorods with sharp-end morphology and with 2050 nm in diameter and
200300 nm in length were obtained. The products obtained with pure distilled water as an electro-
lytic solvent are almost uniform and mono-disperse spindle-like nanoparticles (i.e., nanospindles)
with 80100 nm in diameter and 200300 nm in length. However, by simply increasing the current
density from 5 mAcm
2
to 10 mAcm
2
and then to 20 mAcm
2
, the morphologies of CuO nanostruc-
tures tend to vary from nanospindles to nanorods and then to irregular nanoplates. These results sug-
gest that electrochemical shape- and size-controlled synthesis of CuO nanostructure could be easily
realized by controlling the current density or changing the electrolytic solvent.
Toboonsung et al. [167] synthesized CuO nanorods and their bundles on a glass substrate using an
electrochemical dissolution and deposition process. The deposition time, electrode separation, and
voltage were found to play key roles in the formation of CuO nanorods and the ratio of bundles to
Fig. 21. Schematic for typical setup of an electrochemical deposition facility for synthesizing CuO nanostructures.
Fig. 22. TEM images of the leaf-like CuO mesocrystals under (a) low and (b) high magnication [171]. Copyright 2012 Royal
Society of Chemistry.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 231
co tinh kha khi thuong mai
nanorods. In addition, Xu et al. [168] synthesized large-scale CuO honeycombs using two-step electro-
chemical deposition and subsequent heat-treatment. The Cu(OH)
2
precursor was rst fabricated by
two-step electrochemical deposition of a Cu foil in an aqueous solution of KOH. The CuO honeycombs
were then synthesized by heating the Cu(OH)
2
precursor at 160 C for 3 h under nitrogen protection.
The applied potential and deposition time were found to be the most important parameters affecting
the honeycomb-like CuO growth. The CuO honeycombs comprised well-oriented nanowires with uni-
form average diameters of approximately 100 nm and lengths of tens of micrometers. In addition to
the nanorods and honeycombs, ower-shaped CuO nanostructures can also be achieved electrochem-
ically. Lu and Huang [169] recently reported the synthesis of ower-like CuO microspheres from nano-
akes by electrochemical anodic dissolution of pure Cu in an NaOH aqueous solution at room
temperature. Whisker-shaped CuO nanostructures electrochemically fabricated from a metallic Cu
precursor, followed by annealing at 600 C for 30 min in air, have also been recently reported [170].
Xu et al. [171] reported the synthesis of leaf-like CuO mesocrystals using the electrochemical pro-
cess (Fig. 28), in which Cu foils were used as the working and counter electrodes. The electrodes were
submerged into an aqueous solution of NaNO
3
. The distance between the two electrodes was main-
tained at 25 mm with moderate magnetic stirring being applied throughout the process. The CuO
mesocrystals were electrochemically grown at a constant voltage of 3 V for 200 s at 70 C. The ob-
tained precipitates were nally harvested from the solution by centrifugation and dried at 70 C.
TEM images (Fig. 22) reveal that the width of the CuO NLs is 50 nm, and the length is estimated
to be approximately several hundred nanometers. Each CuO NL comprised numerous small particles,
causing the very rough surfaces of the obtained CuO NLs. When usedas LIB anode materials, the ob-
tained leaf-like CuO mesocrystals showed high specic capability and good cycle performance because
of the novel feature of the CuO mesocrystals.
2.4. Thermal oxidation method
Different methods have been employed to synthesize 1D CuO nanostructures [25,30,31]. The most
commonly used method is to directly heat Cu substrates in air, during which the reaction between Cu
Fig. 23. SEM images of the CuO nanowires synthesized by directly heating Cu grids in air at 500 C for (ac) 4 h and (d) 2 h
[172]. Copyright 2002 American Chemical Society.
232 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
and oxygen (O
2
) results in the growth of CuO nanowires [31]. Typically, the process of thermal oxida-
tion just involves heat treatment of pure Cu substrates in either ambient air or O
2
atmosphere. The
morphology of the grown CuO nanowires depends on the oxidation temperature, growth time, and
gas ow rate.
In 2002, Jiang et al. [172] synthesized CuO nanowires by directly heating Cu substrate in air. Typ-
ically, the Cu substrate was cleaned using HCl solution to remove the native oxidation layer and rinsed
with distilled water. After blow-drying with N
2
gas, the Cu substrate was placed in a furnace. Then, the
substrate was heated at 500 C for 4 h and naturally cooled to room temperature. The CuO nanowires
with large aspect ratio grew on the surface of the Cu grid or wire (Fig. 23). These nanowires had diam-
eters ranging from 30 nm to 100 nm and can be up to 15 lm long. The authors also reported that CuO
nanowires only grow within the temperature range from 400 C to 700 C.
Various synthesis parameters, including oxidation temperature, oxidation time, O
2
ow rate, and
different type of Cu substrates, have been extensively studied to achieve better results from thermal
oxidation. Chen et al. [173] annealed a Cu foil in air using various temperatures and growth times.
Their results showed that the density and the length of the CuO nanowires increased as the growth
time was prolonged at 400 C. Instead of directly heating in ambient air, Kumar et al. [174] synthe-
sized CuO nanowires by thermal annealing Cu foil in an oxygen atmosphere. The results showed that
the oxygen ow rate and the annealing temperature both affected the aspect ratio and density of CuO
nanowires, but annealing time primarily affected only the aspect ratio. Compared with the direct oxi-
dation in air, the presence of oxygen ow during thermal oxidation can reduce the necessary growth
time for CuO nanowires to no more than 1 h. Mema et al. [175] improved the growth density of CuO
nanowires by applying bending stresses on the Cu surface. The Cu foil was initially bent at a 10 mm
radius, followed by typical cleaning, annealing in 200210 Torr oxygen pressure, and cooling. The
upper surface of the Cu foil was shortened by dL, whereas the bottom surface was elongated by dL,
Fig. 24. A schematic diagram illustrating the generation of stresses at the upper and bottom surfaces of the Cu foil [175].
Copyright 2011 Elsevier.
Fig. 25. SEM images of the oxide surface for (a) unbent Cu, (b) upper surface of the bent Cu, and (c) bottom surface of the Cu
[175]. Copyright 2011 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 233
khi moi truong
causing the compressive and tensile stresses in the upper and bottom surfaces, respectively (Fig. 24).
The results shown in Fig. 25indicated that the tensile stress in the bottom surface promoted the
growth density of the CuO nanowires. However, the length of the CuO nanowires is not signicantly
dependent on tensile stress.
Yuan and Zhou [176] reported that the growth density and length of the CuO nanowires could be
enhanced by increasing the surface roughness of the Cu substrate. The Cu substrates were rst sand-
blasted with different durations to generate different surface roughness, followed by typical thermal
oxidation process under 200 Torr oxygen pressure. Their results proved that surface roughness signif-
icantly increases the length and density of the grown CuO nanowires. However, the synthesis of CuO
nanowires directly on Cu substrates (foils, plates, or grids) is not suitable for device integration with
the current semiconductor industry. Therefore, novel techniques to directly grow CuO nanowires on
other substrates (especially semiconducting silicon) are desired to achieve CuO nanowire-based func-
tional devices. Zhang et al. [177] reported large-scale and aligned CuO nanowires synthesized onto a
silicon substrate by thermal oxidation of a Cu thin lm deposited onto silicon. Comparative results of
two Cu thin lms deposited by thermal evaporation and electroplating show that a uniform and large
amount of CuO nanowires grew on the electroplated thin lm only, which provided higher roughness
and larger surface grain size. Given that the CuO nanowires were synthesized onto silicon, a basic
material for microelectronics and microsystems, integrating CuO nanowires into silicon-based micro-
systems is more convenient to achieve promising functional devices. The introduction of high-purity
N
2
and O
2
gas during annealing results in more vertically aligned and uniform CuO nanowires [178].
Furthermore, Zhang et al. [179] presented a novel localized thermal oxidation method in ambient
air instead of heating the entire Cu foils/lms to synthesize the CuO nanowires. The CuO nanowires
only grew on the surface of the heated area realized by localized joule heating (Fig. 26). This method
is CMOS-compatible and can potentially integrate CuO nanowires with conventional microelectronics.
Vertically aligned CuO nanowires by thermal oxidation of a Cu thin lm deposited onto 30 nm
Cu/Ti lm-coated silicon substrates were fabricated by Cheng and Chen [180]. The length of the
Fig. 26. (a) SEM image of the suspended microheaters, (b) highly magnied SEM image of the tip in (a), (c) highly magnied
SEM image of CuO nanowires in (b), and (d) optical image of one tip heater being heated [179]. Copyright 2010 Institute of
Physics.
234 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
CuO nanowires was observed to be tuned from several to tens of micrometers by tailoring the
oxidation temperature and time. The obtained CuO nanowires were single-crystalline with different
axial crystallographic orientations, and the average length of CuO nanowires produced at each
temperature followed a parabolic relationship with the oxidation time (Fig. 27).
Tsai et al. [181] synthesized uniformly aligned single-crystal CuO nanowires using a Co WP capping
layer as a nanolter to catalyze CuO nanowires on Cu (100 nm)/TaN/Ta/SiO
2
/Si blanket substrates by
thermal oxidation. Their results suggested that the obtained CuO nanowires can grow at a relatively
higher growth rate than those reported and become longer and denser with increasing calcination
time. Park et al. [182] demonstrated that CuO nanowire growth can be achieved by thermal oxidation
of Cu metal deposited on CuO (20 nm)/SiO
2
/Si substrate. The CuO nanowires were grown by a contin-
uous supply of both Cu from the Cu lms and O
2
from air. The growth of CuO nanowires by heating Cu
lm deposited on glass substrates in the air was realized by Hsueh et al. [183] and Chang and Yang
[184] (Fig. 28). To avoid cracking of the obtained CuO nanowires/lm during oxidation, a 100-nm-
thick CuO lm was rst deposited onto the glass substrate to serve as an adhesion layer (Fig. 28).
Moreover, the average length of the CuO nanowires was determined by the initial Cu lm thickness.
In summary, various methodologies have been reported to synthesize nanostructured CuO. The
hydrothermal synthesis, a wet chemical process with a reaction occurring in solution, is well known
for low temperature, simple equipment, environmentally safe process, and good potential for
high-quantity production. The solution-based chemical precipitation method that utilizes chemical
solutions is another promising synthetic route because of its high efciency, relatively low-cost,
and its advantages in adjusting the size and morphology of the CuO nanostructures. The electrochem-
ical method for the formation of nanostructured CuO is also of particular interest because of its many
merits over other methods including low-temperature, ease of process, and viability of commercial
production. Furthermore, with the right selection of pH level and/or potential, the Cu or CuO phase
can be well controlled. Thermal oxidation is a simple, efcient, and low-cost method for synthesizing
1D CuO nanostructures, and it is suitable for batch fabrication and mass production. The achieved 1D
CuO nanostructures are uniform and vertically aligned with low impurity. In addition, the morphol-
ogy, density, diameter, and length ofthe1D CuO nanostructures can be easily tailored by adjusting
the synthesis parameters.
Fig. 27. Photographs of as-deposited Cu/Ti thin lm and electrodeposited Cu lm and SEM images of nanowires. Photographs of
the (a) as-deposited Cu/Ti thin lm on a silicon substrate and the electrodeposited Cu lm-coated silicon samples (b) before and
(c) after thermal oxidation. (d) A typical cross-sectional SEM image of nanowires grown on an oxidized Cu lm-coated silicon
substrate. (e) A typical high magnication SEM image of an individual nanowire [180]. Copyright 2012 Springer.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 235
dinh huong
gia tri
dien ap
2.5. Other synthetic methods
In addition to the methods described in Section 2.4, other synthetic techniques have also been used
to prepare CuO nanostructures. These techniques include sonochemical synthesis [48,50,185189],
microwave irradiation synthesis [190196], template-assisted method [197205], solgel [206],
microemulsion [207210], electrospinning technique [211213], synthesis combined with some
physical methods (e.g., spray pyrolysis) [214,215], thermal-based chemical vapor deposition [216],
and c-irradiation [217]. For example, nanorods, nanoparticles, and submicrospheres from CuO have
been obtained by sonochemical synthesis. Various CuO nanostructures, which include nanorods, shut-
tle-like, ower-like, plate-like, leaf-like, dandelion-like, and hollow structures, together with CuO
nanoparticles have been synthesized by microwave irradiation synthesis. CuO nanowires and nanof-
ibers have been achieved by template-assisted method and electro-spinning technique.
3. Growth mechanisms
The development of nanotechnology has resulted in the fabrication of CuO nanostructures with
various morphologies and sizes using different synthetic methods. However, the growth mechanisms
responsible for the formation of CuO nanostructures with various morphologies during syntheses are
still not fully understood, and extensive studies have been conducted to determine the growth mech-
anisms of different CuO nanostructures. In this section, we briey review the most important mech-
anisms that are proposed for the growth of CuO nanostructures.
3.1. Oriented attachment
Oriented attachment is dened as a special kind of crystal growth in which small crystallites attach
to each other through their suitable crystal planes or facets along the same crystallographic directions.
In this sense, the nal aggregates can be considered as large single crystals built from the pristine crys-
tallite in an irreversible and highly oriented manner [218,219].
Zhang et al. [220] demonstrated an anisotropic aggregation-based crystal growth of a few hundred
monoclinic CuO nanoparticles into uniform ellipsoidal monocrystalline architectures by taking advan-
tage of the oriented attachment (Fig. 29). Stepwise orientation and aggregation in three dimensions
were observed at room temperature, caused by the formation of primary CuO nanoparticles in a
mother solution via the preferential 1D [001] orientation of a limited number of nanoparticles at
an early stage. Then, the 3D-oriented aggregation of a few hundred nanoparticles results in monocrys-
talline structure as illustrated in Fig. 30. The selective absorption formamide molecules on various
crystallographic planes of monoclinic CuO nanoparticles may play a critical role in the anisotropic
growth of uniform ellipsoidal monocrystalline architectures.
Fig. 28. Procedure used to grow CuO nanowires and fabricate CuO nanowire-based pH sensors [184] Copyright 2012 ESG.
236 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Liu et al. [128] found that 2D-layered oval CuO nanosheets and 3D CuO nanoellipsoids consisted of
nanosized subunits that essentially have the same size as the primary crystals obtained at the early
growth stage. This phenomenon indicates that the nal products were actually built from the original
small precursors by oriented attachment growth mechanism. The individual oval nanosheet is com-
posed of layered nanoribbons and exhibits an almost single-crystal diffraction pattern with its spot-
like appearance along the [001] axis of crystalline (Fig. 31), indicating that the CuO nanosheets are
formed through the oriented attachment of small nanoribbons along the [010] direction. The HRTEM
images in Fig. 31b taken from two attached CuO nanoribbons reveal that the nanoribbons are single-
crystals. The fringe spacing is measured to be approximately 2.7 , which corresponds to the [110]
lattice fringe of the monoclinic CuO, suggesting that the growth direction of the nanoribbons is
[010]. This result conrms the supposition that the CuO nanosheets are formed through oriented
attachment of small nanoribbons along the [010] direction. A typical TEM image of an isolated CuO
Fig. 29. Schematic illustration of the assembly of CuO built from aggregated nanoparticles [220]. Copyright 2005 Wiley.
Fig. 30. (ac) Schematic illustration of stepwise orientation and aggregation of a large number of primary nanoparticles into a
monocrystalline structure. (df) TEM and SAED examination of the temporal evolution of stepwise orientation and aggregation
growth of CuO nanoarchitectures. (d) The formation of primary nanoparticles and the one-dimensional [001] orientation of
primary nanoparticles during the early reaction stage (rst 2 h). (e) Oriented aggregation of a few hundred primary
nanoparticles into a loosely organized aggregate during the intermediate stage (24 h aging). (f) Further 3D rearrangement of
primary nanoparticles leads to the formation of single-crystalline nanoarchitectures after subsequent aging for 1 wk [220].
Copyright 2005 Wiley.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 237
nanoellipsoid in Fig. 31c shows a rough surface, indicating that the structure is assembled from small
nanoparticles. The SAED of the whole ellipsoid (inset of Fig. 31c) reveals that the entire structure of
ellipsoid is single-crystal with the preferential [010] growth direction. The HRTEM image taken from
the head part and the central surface of an individual CuO nanoellipsoid as illustrated in Fig. 31d and e
indicate that both fringe spacings are 2.3 , which corresponds to the (200) plane of monoclinic CuO,
conrming that the long-axis direction is along [010]. This result implies that the subunits oriented-
assemble with each other and nally form a single-crystal structure.
Interestingly, Liu and Zeng [221] reported a hydrothermal method for the formation of hollow CuO
microspheres with dandelion-like structures using the oriented attachment mechanism. Hollow
microspheres were obtained from primary small crystal strips that contain even smaller 1D nanorib-
bons (Fig. 32). A two-tiered organizing scheme for the construction of CuO microspheres with interior
space has been elucidated (Fig. 32f): (1) mesoscale formation of rhombic building units from smaller
Fig. 31. (a) TEM image of an individual CuO nanosheet. The inset gives the corresponding SAED. (b) HRTEM image of two
attached nanoribbons, showing the [010] growth direction. (c) TEM image of an individual CuO ellipsoid. The inset gives the
SAED. (d) HRTEM image of the head part of one ellipsoid. From the enlarged image in the inset, the [010] growth direction can
be determined. (e) HRTEM image of the central surface of CuO ellipsoids [128]. Copyright 2006 American Chemical Society.
238 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
nanoribbons via oriented aggregation and (2) macroscopic organization of these units into the CuO
microspheres, which facilitate the development from primary nanoribbons to nal dandelion-like hol-
low CuO formation.
Moreover, Xu et al. [138] experimentally conducted for the rst time a large-scale transition pro-
cess from 1D Cu(OH)
2
nanowires to 2D CuO NLs through hierarchical-oriented attachment. The poly-
crystalline Cu(OH)
2
nanowires rst evolved into single crystalline Cu(OH)
2
NLs by an oriented
attachment. Then, the formed crystalline Cu(OH)
2
NLs were converted into single crystalline CuO
NLs via a reconstructive transformation. The transformation process consisted of CuO nucleation, fol-
lowed by a two-step oriented attachment of the CuO particles to 1D CuO nanoribbons and then to 2D
CuO NLs. This phenomenon indicates that oriented attachment plays a critical role in the entire NC
growth for the formation of CuO nanostructures with different morphologies. The assembly of the
1D ribbons and 2D leaves were attributed to the thermodynamic force driven by the reduction of sur-
face energies and the sequential nature, resulting from the difference in surface energy among the
crystal planes. Volanti et al. [93] also attributed the formation of sea urchin-like CuO nanostructure
synthesized by a microwave-assisted hydrothermal process to the particle-by-particle aggregation
growth via oriented attachment. They observed a stepwise aggregation and orientation of primary
CuO nanoparticle cores into 3D sea urchin-like CuO nanostructure. The primary CuO nanoparticle
cores were reported to be rst formed in the early stage and then transformed to the early triangular
points. After consecutive reactions, these triangular points experience CuO aggregation from 3D cross-
like structures to ower-like crystals and transform to the nal sea urchin-like CuO nanostructure.
Through analyzing the products of the different stages during the crystal growth process, the authors
noted that a mesoscale self-assembly followed by fusion of the adjacent crystallites is responsible for
the growth mechanism of CuO microcrystals with sea urchin-like morphology via oriented
attachment.
Fig. 32. (a and b) SEMimages of two crashed CuO microspheres. (c and d) TEM images of two rhombic CuO crystal strips formed
from smaller 1D nanoribbons. (e) SAED pattern of the crystal strip shown in (d). (f) Two-tier organization with multiple-length
scales: (1) oriented aggregation of CuO nanoribbons and (2) concentric alignment of the preformed rhombic building blocks
from (1) [221]. Copyright 2004 American Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 239
As illustrated in the scheme shown in Fig. 33, Li et al. [222] developed an ammonia (NH
3
)-evapo-
ration-induced synthetic method to synthesize hierarchical CuO spheres from primary CuO nanopar-
ticles via oriented attachment. Considering an analysis of the reaction process, they suggested that
NH
3
molecules can passivate the surface to form coordinate bonds with the Cu ions on the surface.
The density of the adsorbed NH
3
was found to be dependent on the density of the Cu
2+
ions on the
crystal plane in the order of (001) > (100) > (010). When the initial Cu(NO
3
)
2
concentration is only
2.5 mM with excessive NH
3
, all the primary CuO NC planes are densely passivated, resulting in a weak
driving force for further aggregation into nanowires or nanoplates. Increasing the Cu(NO
3
)
2
concentra-
tion to 10 mM results in less dense passivation on the (010) plane, but the passivation layers on the
(100) and (001) planes remain largely intact. Thus, oriented attachment along the (b) axis occurs to
form nanowires. Further increasing the concentration to 20 mM can lead to signicant depassivation
on the (100) plane. As a result, both the (a) and the (b) axes in Fig. 33 become possible via oriented
attachment growth. This result indicates that the observed diamond CuO nanoplates with dimensional
sizes along the different axes are in the order of (b) > (a) > (c). However, by increasing the concentra-
tion up to 100 mM, from the single-crystalline nature of the thick plate and its smooth surface, the
author noted that Ostwald ripening is the major growth mechanism. This result suggests that the
growth rates along (a)- and (c)- axes are higher than that along the (b)-axis, resulting in the formation
of a plate with the (b)-axis as the shortest dimension. The results indicated that the NH
3
molecules can
passivate the surface and thereby inuence the growth and aggregation behavior of the particles. With
a gradual decrease in ammonia concentration, the primary particles attach with each other and form
the hierarchical CuO nanostructures.
Xu et al. [171] recently demonstrated the use of electrochemical approach for the synthesis of leaf-
like CuO mesocrystals that were found to be built up from primary CuO nanoparticles by oriented
attachment growth mechanism as illustrated in Fig. 34. The TEM image of an individual CuO NL
and the corresponding SAED pattern are shown in Fig. 35a and b. Individual diffraction spots shown
in Fig. 35b indicate that the obtained leaf-like CuO is a single crystal. The HRTEM image in Fig. 35c
taken from the marked area in Fig. 35a shows a clear and continuous lattice-fringe indicating that
the CuO NLs have the same crystallographic orientation (like a single crystal). Hence, the authors con-
cluded that the primary particles are orderly attached with each other. Moreover, clear parallelism of
the lattice-fringes in Fig. 35c further conrmed that each attached nanoparticle shares the same crys-
tallographic orientation. TEM images of another CuO NL are shown in Fig. 36. The insets of Fig. 36c
present the Fast Fourier Transform (FFT) patterns, which correspond to the respective marked areas.
The FFT patterns originating from the marked areas are similar, indicating that the nanocrystals within
Fig. 33. Growth mechanism of CuO at various NH
3
/Cu
2+
ratios. (A) Oriented attachments of CuO primary nanocrystals along the
[010] direction to form nanowires at a high NH
3
/Cu
2+
ratio; (B) Oriented attachments along the [010] and [100] directions to
formnanoplates at a mediumNH
3
/Cu
2+
ratio; (C) Ostwald ripening of the CuO primary nanoparticles to formthick nanoplates at
a low NH
3
/Cu
2+
ratio [222]. Copyright 2009 American Chemical Society.
240 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
the leaf-like CuO share the same crystallographic orientation. These results provide evidence that the
CuO mesocrystals are formed through the oriented attachment of small nanocrystals.
Sun et al. [142] demonstrated the signicant evidence on a green synthesis for the ordered-
aggregation-driven growth from surfactant-free 1D CuO nanosubunits into dimension-controlled
mesostructures (3D spindles and 2D plates) by a facile additive-free complex precursor solution
method through oriented attachment. An oriented nanoparticle-aggregation with tailoring shapes in
different dimensions can be achieved in different concentrations of reactants at high reaction temper-
ature (Fig. 37). The 3D layer-by-layer growth of mesostructural CuO spindles was achieved in
low-concentration reagents, whereas the 2D shoulder-by-shoulder growth of mesostructural CuO
plates was obtained in high-concentration reagents.
In addition to the aforementioned examples, an increasing variety of CuO nanostructures
[88,223225] are being identied using the oriented attachment as the growth mechanism. This
mechanism can facilitate the design of various types of inorganic materials and even more complex
hybrid materials with unique structures.
Although the application of oriented attachment mechanism in controlling the shape of various
CuO nanostructures have been observed and conrmed, a commonly accepted understanding to elu-
cidate both the chemical reaction route and the oriented attachment mechanism is necessary. Xu et al.
Fig. 34. Schematic illustration for the formation of leaf-like CuO mesocrystals [171]. Copyright 2012 Royal Society of Chemistry.
Fig. 35. (a) TEM image, (b) SAED pattern, and (c) HRTEM image of an individual CuO NL [171]. Copyright 2012 Royal Society of
Chemistry.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 241
[84] recently experimentally provided strong evidence that the assembled CuO particles are composed
of CuO primary single crystal nanoparticles, which are bounded by three {100} planes. The morphol-
ogy of these CuO primary crystals was determined by the surface energy of the exterior crystallo-
graphic planes and inuenced by NaOH concentration in the solution. They are oppositely charged
on the opposite surfaces. The electrical polarity, thus the static attraction force, was observed to be
benecial to the selective and aligned self-assembly of these CuO primary crystals. The surface charge
densities, the selective adsorption of OH

ions, and the shape of CuO primary crystals determine the


selectivity of self-assembly on the three {100} planes. Governed by the selectiveness of self-assembly,
the morphology of the assembled CuO particles evolves into various microstructures (Fig. 38).
3.2. Ostwald ripening process
For more than a century, the Ostwald ripening process is a physical phenomenon has been well
known to elucidate the formation of a crystal. This process refers to the growth of larger crystals from
smaller crystals that have higher solubility than the larger ones [219,226228].
An interesting example found by Zou et al. [229] showed that well-aligned arrays of CuO nanoplat-
elets synthesized through a hydrothermal route without template assistance were grown via Ostwald
ripening process. After investigating the hydrothermal reaction under different reaction times (6, 20,
and 40 h), the authors suggested that the nanoplatelets did not grow by vaporliquidsolid (VLS) or
solutionliquidsolid (SLS) mechanisms because of the absence of liquid droplet located at the top
of CuO nanostructure as the catalytic active site to urge the growth of CuO nanoplatelets. The evolu-
tion process of the nanoplatelets was extensively observed, and the authors reasonably presumed that
the CuO nanoplatelets were formed through the Ostwald ripening process. At the initial stage of the
hydrothermal reaction, both small and large nanopatches could be produced in non-equilibrium solu-
tion of autoclave. Then, these small nanopatches gradually dissolved to generate solvated h-ions in the
solution, and these solvated h-ions spontaneously transferred onto the surfaces of some large nano-
patches. Next, the large CuO crystals (nanoakes) served as the seeds for the growth of nanoplatelets
at the expense of the small nanopatches through Ostwald ripening. Finally, large amount of CuO
nanoplatelets with wider, longer, and thicker sizes were formed.
Fig. 36. (a) Low-magnication, (b) high-magnication, and (c) HRTEM images of the CuO mesocrystal. The insets are the FFT
patterns originated from the marked areas [171]. Copyright 2012 Royal Society of Chemistry.
242 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Guan et al. [87] recently demonstrated that hierarchical CuO hollow microspheres prepared by
hydrothermal process without any surfactants or templates can be formed via a combined process
of self-assembly and Ostwald ripening. During the rst stage, tiny crystalline nuclei were formed in
the supersaturated solution, which grew into nanoparticles (Fig. 39). Then, these primary nanoparti-
cles quickly generated and further spontaneously aggregated to form spherical aggregates with min-
imizing interfacial energies. Subsequently, grain growth was initiated preferentially from the most
thermodynamically active clusters on the surface of the spherical aggregates. Next, symmetric micro-
spheres that were assembled by singular nanorods may be formed because of the supersaturated
Cu(Ac)
2
solution surrounding the active clusters. Given that the randomly oriented growth is physi-
cally limited, the growth fronts could radiate towards the exterior of the microsphere with the driving
force of the natural tendency of polar crystal growth in CuO. Finally, compared with those in the exte-
riors, the crystallites located in the cores may have smaller crystallite sizes and higher surface ener-
gies. The energy non-equilibrium between the large particles located in the exteriors and the small
particles in the cores provides the driving force for Ostwald ripening process. Consequently, the Ost-
wald ripening process may dominate the subsequent growth of the microspheres in the second stage.
In addition to the aforementioned examples, the formations of other inorganic compounds, such as
Cu
2
O nanospheres and nanocubes, TiO
2
hollow spheres, and coreshell structures of ZnS and Co
3
O
4
,
have also been identied to adopt the Ostwald ripening as the growth mechanism [227].
Fig. 37. (A) Schematic illustration of the reaction pathway and formation mechanism of the ordered-aggregation-driven growth
from surfactant-free 1D CuO nanocrystals into dimension-controlled mesostructure 3D mesospindle and 2D mesoplate. (B) TEM
images of an individual CuO mesospindle and the as-prepared CuO mesoplates [142]. Copyright 2012 Royal Society of
Chemistry.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 243
The combination of both oriented attachment and Ostwald ripening processes has also been dem-
onstrated in the synthesis of ultralong CuO nanowire bundles and hierarchically porous CuO hollow
spheres. Li et al. [124] demonstrated that ultralong CuO nanowire bundles were rst formed through
oriented attachment of colloidal particles, then through Ostwald ripening side self-assembly process
leading to nanowire bundles, and nally to CuO NLs by further Ostwald ripening process (Fig. 40a).
However, without PEG 200, the nanoparticles aggregated quickly to form single crystal CuO NLs
directly through Ostwald ripening (Fig. 40b).
Qin et al. [230] reported hierarchically porous CuO hollow spheres synthesized by a simple one-pot
template-free method. The CuO hollow spheres were formed via an oriented-attachment growth step
followed by an Ostwald ripening process. The quasi-micropores and mesopores were formed by the
primary nanograin attachment and the nanosheet assembly, respectively. The macropores (hollow
interior) of the CuO spheres were produced by self-hollowing driven by the Ostwald ripening process
(Fig. 41). These two growth processes have also been observed in the synthesis of hollow nanocubes of
Cu
2
O and Cu metal [231].
3.3. Scroll of Cu(OH)
2
nanosheets
CuO 2D nanostructures, such as nanosheets and nanoribbons, can also be formed directly on the Cu
lm by dehydration of the pre-formed Cu(OH)
2
nanotubes on a Cu foil under alkaline oxidative
conditions [129,163,232,233]. The proposed mechanism is as follows:
Cu
2
2OH

!CuOH
2
; CuOH
2
!CuO H
2
O
Fig. 38. Schematic of the formation for CuO primary crystals to self-assemble into 2D CuO microplates [84]. Copyright 2012
Royal Society of Chemistry.
Fig. 39. Schematic illustration of the formation mechanism proposed for CuO hollow spheres [87]. Copyright 2011 Elsevier.
244 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
The formation of Cu(OH)
2
and CuO nanostructures on Cu surfaces involves inorganic polymeriza-
tion (polycondensation) reactions under alkaline and oxidative conditions [129]. First, by dipping
the Cu foil into an alkaline oxidant solution, the Cu foil is rapidly oxidized to Cu
2+
by the oxidant. Then,
the highly alkaline conditions favor the square planar coordination of OH

groups to Cu
2+
(Fig. 42a),
which leads to an extended chain along [100] (Fig. 42b). The chains can be connected through the
coordination of OH

to dz
2
of Cu
2+
, forming a 2D structure (Fig. 42c). Finally, the 2D Cu(OH)
2
layers
are stacked through the relatively weak hydrogen bond interactions to forma 3D crystal.
Under highly basic conditions, the interlayer H-bond linkage at the sheet edges may be weakened.
This condition, together with the asymmetric layer structure, causes stresses in the layers. Conse-
quently, the nanosheets roll to relieve the stresses, forming the nal Cu(OH)
2
tubular structure
[32,163,232]. The corrugated layer structure of Cu(OH)
2
(Fig. 43) is unstable against oxolation because
oxygen atoms are either penta coordinated or tricoordinated. Consequently, under certain reaction
conditions such as high pH or long reaction time, the interplanar hydrogen bonds will be broken,
and the cleavage of the interplanar hydrogen bonds causes the Cu(OH)
2
bers and scrolls to crack into
pieces of CuO sheets/ribbons [129].
3.4. Stress and grain-boundary diffusion
As stated in Section 2.4, the thermal oxidation method is currently widely used to synthesize CuO
nanowires from Cu substrates (foils, lms, wires, etc.) because ots simplicity, low-cost, high quality
Fig. 40. Schematic illustration of (a) the formation of Cu(OH)
2
and CuO nanowire bundles and nally CuO NLs with PEG200. (b)
Formation of CuO NLs without PEG200. The thread-like object in (a) represents PEG200, and the spheres represent Cu(OH)
2
nanoparticles [124]. Copyright 2012 Elsevier.
Fig. 41. Schematic illustration for the formation of the hierarchical structure and pores of the CuO spheres [230]. Copyright
2012 American Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 245
of realized nanowires, and suitability for large-scale fabrication. However, the growth mechanism of
CuO nanowires via thermal oxidation method remains controversial. The VLS and vaporsolid (VS)
models have been excluded by many researchers [173,174,234,235]. Kumar et al. [174] attributed
the CuO nanowire formation to the relaxation of compressive stress at the CuO/Cu
2
O/Cu interface
caused by volume and structural changes. Annealing the Cu foil under an oxygen ow rate of
150 mL/min resulted in the appearance ofCuO nanowires during annealing at 400 C for just 5 min
and at 350 C for 30 min. However, no nanowire was observed when annealing at 350 C for only
10 min or at a lower temperature, e.g., such as at 300 C. Hence, the authors believed that a period
for the accumulation of stress is critical for the growth of CuO nanowires. Kuar et al. [234] reported
that no CuO nanowires were found when directly heating Cu
2
O under a similar condition, which indi-
cated that the formation of Cu
2
O is not a necessary condition for the growth of CuO nanowires during
thermal oxidation of Cu. The authors believed that the accumulated stress was not sufcient to cause
the growth of CuO nanowires. Shao et al. [235] recently provided evidence for the growth of CuO
nanowires caused by the relaxation of compressive stress. The Cu microcontainers were fabricated
on a glass substrate and heated at 400 C for 3 h in air. TheCuO nanowires grew in the inner surface
of the microcontainers only as shown in Fig. 44. The authors believed that as oxygen atoms diffused
into the Cu microcontainer, tensile stress was generated at the outer surface whereas compressive
stress was generated at the inner surface. Consequently, the CuO nanowires grow as a result of relax-
ation of compressive stress. However, the manner by which the compressive stress impels the growth
of CuO nanowires has not been elucidated.
Gonalves et al. [236] presented a systematic study on the growth of CuO nanowires by thermal
oxidation of commercial Cu foils under atmospheric pressure in air. The authors proposed that the
CuO nanowire growth occurs via grain-boundary diffusion of Cu ions through the Cu
2
O layer and
the oxygen ions through the top CuO layer. Zhong et al. [237] also believed that given a highly porous
and defective Cu
2
O layer that formed initially because ofthe compressive stress, the Cu atoms diffused
to the surface by grain-boundary diffusion during the continuous oxidation process. Li et al. [238] re-
ported the use of external electric eld to affect the growth of CuO nanowires. The electric eld for
driving Cu ions outwards with the strength of 3000 V/m resulted in a length ve times longer than
Fig. 42. Schematic showing the coordination assembly growth of Cu(OH)
2
nanoribbons [233]. Copyright 2002 American
Chemical Society.
246 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
that without external electric eld. This result proved that the diffusion of Cu ions plays a key role in
the formation of CuO nanowires. Based on grain-boundary diffusion, Yuan el al. [239] proposed that
the formation of CuO nanowires resulted from the CuO/Cu
2
O interface reaction which produced com-
pressive stress in the CuO layer and thus forced Cu ions to diffuse outwards along CuO grain bound-
aries. A kinetic model was also proposed and is shown in Fig. 45. In this model, the CuO and Cu
2
O
layers also grow during the oxidation process, consequently leading to the burying of CuO nanowires
found by Zhu et al. [240]. This model explained the observed time-dependent length, bi-crystalline
nature observed in Refs. [172,175,177,239,241]. However, further improvement in the model is
needed to explain the formation of mono-crystalline CuO nanowires as observed in Refs.
[174,175,182,235,237,242245].
Liang et al. [243] also stated that the growth of CuO nanowires occurred to reduce the stress gen-
erated during Cu oxidation by grain-boundary diffusion process. The smaller grain size can increase
the effective diffusion coefcient of Cu ions and oxygen in a smaller crystallite size. Consequently, a
higher density of CuO nanowires can be observed. The compressive stress serves as the driving force
for Cu ion diffusion, which is dominated by grain-boundary diffusion at intermediate temperature
(300600 C). However, at a higher temperature, such as 800 C, the oxidation is dominated by lattice
Fig. 43. Theac projection of the (010) plane of orthorhombic Cu(OH)
2
. A corrugated layer formed by edge sharing of distorted
Cu(OH)
6
octahedra [129]. Copyright 2003 American Chemical Society.
Fig. 44. Tilted SEM images of CuO nanowires grown in microcontainers by thermal oxidation in air [235]. Copyright 2011
Springer.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 247
diffusion, and thus no nanowires form. Mema et al. [175] increased the in-plane tensile stress by bend-
ing Cu foils leading to smaller grain sizes at the early stage of the oxidation process, thus increased the
density of CuO nanowires. Evidence was provided in the report of Yuan and Zhou [176] in which the
density of grain-boundaries was increased by sandblasting on the Cu substrate, resulting in a larger
density of CuO nanowires.
Hansen et al. [246] obtained dense ultra-long and vertically aligned CuO nanowires by oxidation of
Cu substrate with surface mechanical attrition treatment (SMAT) as illustrated in Fig. 46. They ob-
served substantial increase in the oxide growth rate and tremendous enhancement inthe density
and length of the nanowires. Detailed TEM analysis of single CuO nanowire demonstrated the pres-
ence of a bi-crystal structure that extends along the entire length of nearly every nanowire. A model
of bi-crystal boundary of the CuO nanowire based on TEM analysis is illustrated in Fig. 46, suggesting
that bi-crystal boundary of the CuO nanowire comprised a Cu-rich boundary of (112)/(001). These
results provide strong evidence that the growth of CuO nanowires occurs through short-circuit grain
boundary diffusion of Cu ions across the Cu
2
O layer, followed by short-circuit diffusion along the CuO
nanowire bi-crystal grain boundary and to the nanowire tip, where subsequent oxidation and growth
occur.
3.5. Stress-induced cracking
Thermal oxidation method, a simple, scalable, and low-cost method, has been extensively used to
synthesize high-quality CuO nanowires from different Cu substrates (foils, lms, wires, etc.).
CuO nanowires had been broadly applied using the thermal oxidation method. However, the
mechanical adhesion between the CuO nanowire/lm and the substrates is very weak. Cracking and
aking of the CuO nanowire/lm or even exfoliation from the substrates is still a common problem
[177,247249]. This important limitation has not been resolved. It severely affects the properties
and practical applications of the CuO nanowires, especially if the CuO nanowires should be used under
liberating, rubbing, and impacting environments where the adhesion between the nanowires and the
substrates are crucial.
The mist stress between Cu and CuO generated during the high-temperature oxidation process
causes the cracking of CuO nanowire/lm. According to the understanding of the stress-based cracking
mechanism, several methods have been proposed to alleviate the cracking problem. Zhang et al. [177]
employed photolithography to form patterned and separated Cu patches on a silicon substrate. During
the thermal oxidation process, the discontinuous surface allowed stress release at the edges. Conse-
quently, the detaching of the CuO nanowires/lm from the substrate is prevented (Fig. 47b).
Wang and Li [248] reported that the cracking and exfoliation of CuO nanowires from a Cu foil sub-
strate can be eliminated by the chemical predeposition of a ZnO layer on a Cu foil before the thermal
oxidation process. During thermal oxidation, the ZnO layer causes the generated Cu
2
O to further react
to CuO. Consequently, a 100-nm-thick CuO layer is formed between the ZnO and the Cu substrate for
Fig. 45. (a) Initial growth of CuO nanowires on the outer surface of CuO grains. (b) Growth of the CuO substrate buries the root
of the nanowires. (c) Continued decomposition of the CuO layer at the CuO/Cu
2
O interface leading to the direct contact between
the nanowire roots and the Cu
2
O layer [240]. (d) Cross-sectional image from the oxidation of Cu showing that CuO whiskers are
buried by the CuO layer, and that their roots have direct contact with the Cu
2
O layer [240]. Copyright 2005 Elsevier.
248 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Fig. 46. (a) Schematic illustration of the SMAT processing of Cu and the subsequent CuO nanowire growth process. (b) TEM
image showing bi-crystal structure along the entire length of the nanowire. (c) A dark eld TEM image showing the bi-crystal
structure. (d) TEM image of the CuO nanowire tip. (e) Model of the CuO-monoclinic nanowire bi-crystal interface. (f) Schematic
illustration of the proposed CuO nanowire growth model by grain boundary diffusion [246]. Copyright 2011 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 249
the ZnO-coated Cu foil (Fig. 48b). The mist stress between the oxide and Cu substrate can be easily
released by the thin CuO layer, and thus the initiation of cracks is prohibited. The authors observed
that eld emission of CuO nanowires is enhanced by eliminating the cracks of CuO nanowire/lm.
Mumm and Sikorski [249] presented an easier method to prevent CuO nanowire/lm cracking
using very thin Cu foils (25 lm). After thermal oxidation, the thickness of the forming oxide layer
was signicantly larger than that of the remaining Cu layer. The difference in thickness caused the re-
lease of the mist stress by deforming the remaining Cu layer sandwiched within the foil and leaving
the oxide layers untouched. A photograph of the supporting data in [249] is shown in Fig. 49, in which
no cracking is found in the 25-lm-thick Cu foil.
Fig. 47. (a) CuO nanowires/lm without patterning. (b) CuO nanowire/lm with patterning [219]. Copyright 2007 Institute of
Physics.
Fig. 48. CuO nanowire/lm synthesized by (a) directly heating a Cu foil and (b) heating a ZnO-precoated Cu foil [248]. Copyright
2009 American Chemical Society.
Fig. 49. Comparison of thermally oxidized Cu foils with different thicknesses from left to right: 25, 100, 250, and 800 lm [250].
Copyright 2011 Institute of Physics.
250 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Overall, a deep understanding of the growth mechanisms of various CuO nanostructures may lead
to large-scale and well-controlled synthetic processes of nanostructures. These processes provide new
insights into the morphology-controllable design of high-performance materials for nanodevices.
4. Fundamental properties
Table 3 lists the key physical properties of bulk CuO. However, the reduction of CuO dimensions to
the nanoscale or even smaller scale results insignicant deviation of some of its physical properties
from its bulk counterpart because of the quantum-size effects. Therefore, a thorough understanding
of the fundamental properties of CuO nanostructures is crucial to their synthesis and applications and
a key to the rational design of CuO nanostructure-based functional devices. In this section, we will fo-
cus on the fundamental physical properties of CuO nanostructures including crystal structures and
phase transition, optical, electrical, photoelectrochemical (PEC), and magnetic properties.
4.1. Crystal structures and phase transition
The original crystal structure of CuO was rst determined by Tunnel in 1933 and was then rened
by single-crystal X-ray methods in 1970 [252]. Contrary to the usual rock-salt structure of other 3d
Table 3
Key physical properties of CuO at room temperature (300 K) [251].
Properties Value
Density 6.31 g/cm
3
Melting point 1200 C
Stable phase at 300 K Monoclinic
Dielectric constant 18.1
Refractive index 1.4
Bang gap (E
g
) 1.211.55 eV direct
Hole effective mass 0.24 mo
Hole mobility 0.110 cm
2
/V s
Fig. 50. Crystal structure of CuO (tenorite). The special atomic positions for Cu are (1/4, 1/4, 0), (3/4, 3/4, 0), (1/4, 3/4, 1/2), and
(3/4, 1/4, 1/2) and for oxygen are (0, y, 1/4), (0, 1/2 + y, 1/4), (0, y, 3/4), and (1/2, 1/2 y, 3/4) with y = 0.416(2). The small light
spheres and large dark spheres represent Cu and oxygen atoms, respectively [255]. Copyright 2003 Electrochemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 251
transition-metal monoxides, the CuO crystal structure is monoclinic with C2/c symmetry and four for-
mula units per unit cell. The Cu
2+
ions are at centers of inversion symmetry in a single fourfold site 4c
(1/4, 1/4, 0), and the oxygen ions occupy site 4e (0, y, 1/4) with y = 0.416(2) (Fig. 50) [253256]. The
structural parameters as summarized by Meyer et al. [256] are listed in Table 4.
Bourne et al. [257] proved that CuO has no phase transition at pressures up to 700 kbar and tem-
peratures up to 3000 K, in contrast to other MOs, in which crystal phase transitions can occur during
annealing and cooling. Azam et al. [258] successfully synthesized and annealed CuO nanoparticles at
various temperatures. X-ray diffraction (XRD) measurement of these samples indicated that no crystal
phase transitions are found during the whole annealing process as illustrated in Fig. 51. However, the
Table 4
Structural parameters of CuO [256].
Space group C 2/c (No.15)
Unit cell a () = 4.6837
b () = 3.4226
c () = 5.1288
b () = 99.54
a, = 90
Cell volume 81.08
Cell content 4 [CuO]
Formula weight 79.57
Distances
CuO 1.96
OO 2.62
CuCu 2.90
Fig. 51. XRD spectra of CuO nanoparticles annealed at different temperatures [258]. Copyright 2012 Dove Medical Press.
Table 5
Variation in the crystallite size and lattice parameters with annealing temperature [258].
Temperature (C) a () b () c () Crystallite size (nm)
400 4.688 3.427 5.132 20
500 4.711 3.429 5.133 21
600 4.715 3.430 5.135 25
700 4.719 3.432 5.136 27
252 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
particle size and lattice parameters of these samples increase accordingly with an increase in anneal-
ing temperature as shown in Table 5. Vidyasagar et al. [259] also observed similar experimental re-
sults. Their ndings further conrm that no crystal phase transitions of CuO occur during annealing
and cooling, but particle size and lattice parametersvary.
Several recent theoretical studies based on the density functional theory have focused on atetrag-
onal phase of CuO (i.e., elongated rocksalt cell along one crystal axis), which has recently been success-
fully deposited on the substrates of SrTiO
3
thin lms [260,261]. Himmetoglu et al. [261] also
investigated cubic CuO as a theoretical model, which has not been reported experimentally. For nano-
structured CuO, the unit-cell volume increases with a decrease in particle size, and the lattice becomes
distorted as the crystal symmetry tends to increase. This phenomenon can cause a structural transi-
tion to a phase with a more symmetric structure. Palkar et al. [262] obtained different sizes of CuO
nanoparticles using two distinct routes, namely, rapid liquid dehydration and precipitation, where
the crystallite size is controlled by changing the solution concentration and the calcination condition.
The observed variation in the normalized unit-cell volume (per formula unit) with the XRD domain
size in the CuOCu
2
O system is presented in Fig. 52. The size-induced transition from the low-symme-
try CuO (monoclinic structure) phase to the high-symmetry Cu
2
O (cubic structure) phase is accompa-
nied by a sharp increase in the unit-cell volume (calculated per formula unit) (Fig. 52). This result
indicates that high-symmetry crystal structures (Cu
2
O) are more likely to be stable at smaller sizes
(below 25 nm).
4.2. Electronic band structure
CuO is a narrow-bandgap p-type transition MO semiconductor material. The electronic bandgap of
CuO corresponds to the difference between the energy levels of the top of the valence band (VB) de-
rived from the Cu 3d orbital and the bottom of the conduction band (CB) derived from the Cu 3d orbi-
tal [263]. The electronic structure of CuO was theoretically investigated using cluster model
calculations [60] and other ab initio calculations based on the local spin-density approximation (LSDA)
[58,257,264]. However, these calculations normally fail to predict the semi-conducting ground state
because of the strong electronelectron interactions of CuO that have not been considered in the cal-
culations. To overcome this problem, Anisimov et al. [265,266] obtained the rst correct electronic
density of states (DOS) of CuO by LSDA+Hubbard (U) calculation (Fig. 53), but they failed to provide
a detailed CuO band structure.
Wu et al. [263] recently performed a rst-principle calculation using the LSDA+U method to thor-
oughly investigate the electronic band structure of CuO. For comparison purposes, the LSDA calcula-
tion results were also demonstrated by the authors. The DOS of CuO from the calculations based on
Fig. 52. Variation in the normalized unit-cell volume (per formula unit) with the XRD domain size in the CuOCu
2
O system
[262]. Copyright 1996 American Physical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 253
both the LSDA and LSDA+U methods are shown in Fig. 54. The DOS calculated from the LSDA method
exhibit a metallic nonzero spin-polarized DOS at the Fermi level. The LSDA+U approach leads to a dif-
ferent outcome of the DOS of CuO, with zero DOS at the Fermi level and an indirect energy gap of
1.0 eV. This result suggests that CuO can be successfully predicted to be a semiconductor, consistent
with the experimental results, by introducing Hubbard (U) into the calculation.
Fig. 53. Total and partial DOS of CuO [266]. Copyright 1997 Institute of Physics.
Fig. 54. Comparison of DOS of CuO from LSDA (dotted lines) and LSDA+U (solid lines) calculations. The positive and negative
DOS refer to spin up and spin down calculations, respectively. Zero-point energy is selected at the Fermi level [263]. Copyright
2006 American Physical Society.
254 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
The band structures of CuO along the high-symmetry directions of the rst Brillouin zone calcu-
lated by the LSDA and LSDA+U methods are shown in Fig. 55. Two conduction bands dropped and
crossed the Fermi level (Fig. 55a), that is, no energy gap appeared for CuO, indicating a metallic or
semimetallic ground state for CuO. By contrast, the band structure by the LSDA+U method
(Fig. 55b) showed the absence of a conduction band crossing the Fermi level. An indirect energy
gap of 1.0 eV appeared, which suggests that CuO is a semiconductor in its ground state. Based on
the analysis of the wave functions, the authors indicated that the top VB and the two bottom CBs
are attributed to the 3d orbitals of Cu atoms. They also noted that Cu vacancies are the most stable
defects in CuO in Cu- and O-rich environments by studying the formation energies of native point de-
fects in CuO, indicating that CuO is intrinsically a p-type semiconductor.
Fig. 55. (a) Band structure of CuO from LSDA calculations. (b) Band structure of CuO from LSDA+U calculations. The Fermi level
(dashed line) is set at 0 eV [263]. Copyright 2006 American Physical Society.
Fig. 56. Structural models of CuO surfaces. For clarity, the thicknesses of slabs are much smaller than those used in calculations
[267]. Copyright 2010 American Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 255
The stability and electronic structures of various CuO surfaces were also systemically studied based
on rst-principle calculations by Hu et al. [267] as depicted in Fig. 56. Table 6 shows that CuO(111) is
the most stable surface with a lowest surface energy of 0.74 J/m
2
under ambient condition. By con-
trast, the surface free energy of CuO(100) is 2.28 J/m
2
, which is much larger than those of other ener-
gies because of the existence of a high electrostatic eld between the cationic and anionic surfaces.
Considered a common phenomenon in oxides, oxygen pressure can signicantly inuence the surface
energies, especially for nonstoichiometric surfaces. The surface energies of O- and Cu-terminated CuO
(110) and CuO(100) slabs were also calculated by the authors (Fig. 56). They found that Cu(110)
O
and
CuO(110)
Cu
are relatively more stable in O- and Cu-rich conditions, respectively. Cu(110)
O
is even
more stable than CuO(111) in a very narrow range near the limit of O-rich condition. The density
of state values of CuO(111) and CuO

111 are also plotted in Fig. 57. These surfaces maintain a semi-
conducting behavior similar to that of bulk CuO, but their E
g
values are reduced to 0.79 eV for
CuO(111) and 0.55 eV for CuO

111, compared with bulk CuO (E


g
= 1.1 eV).
A number of studies have focused on the electronic properties of CuO systems. However, informa-
tion on band structure and bandgap energies of CuO remains insufcient. In addition to the theoretical
studies, Ghijsen [60] demonstrated the photoelectron, auger electron, and bremsstrahlung isochro-
matic spectroscopy techniques to study the electronic structure of CuO. They noted that the Cu 3d
Coulomb interaction is larger than the O 2pCu 3d charge transfer energy (D) in CuO, making CuO
a B-type charge transfer semiconductor [E
g
D(proportional to the electronegativity of the anion); holes
are light (anion VB), and electrons are heavy (d bands)] [268]. The bandgap of CuO is 1.4 eV.
To further study the electronic properties of CuO and provide more information on band structure
and bandgap energies of CuO that theoretical studies have not provided, Tahir and Tougaard [269]
determined the electronic properties of CuO by quantitative analysis of photoelectron spectroscopy
(XPS) and reection electron energy-loss spectroscopy. XPS results shown in Fig. 58a indicate that four
Cu 2p peaks exist for CuO, which can clearly be distinguished from those of Cu and Cu
2
O. The peak
positions of Cu 2p/
2/3
, Cu (LVV), O1s, and O (KLL) for CuO are at 934.0, 917.1,529.9, and 509.1 eV,
respectively. These peak positions are different from those of Cu and Cu
2
O. Considering the results
from Ref. [270] together with UV photoelectron spectra (UPS), the CuOband structures are presented
in Fig. 58b.
Table 6
Surface free energies (C) and work functions (U, eV) of stoichiometric CuO surfaces [267].
(111)

111
(011) (101) (110) (010) (100)
C (J/m
2
) 0.74 0.86 0.93 1.16 1.29 1.37 2.28
U (eV) 5.8 5.8 5.6 5.3 6.3 5.6 7.6
Fig. 57. (A) Surface energies of CuO (110) with different terminations as a function of oxygen chemical potential. The values of
1.6 and 0.0 Ev correspond to the limits of O-rich and O-poor conditions, respectively. (B) DOS of (a) bulk CuO, (b) CuO (111),
and (c) CuO 111. The vertical dashed lines highlight the range of band gap of bulk CuO [267]. Copyright 2010 American
Chemical Society.
256 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Several authors have also reported on the positions of the photochemically determined CB and VB
edges of CuO [271274]. Their results indicated that the VB and CBs are located at 4.80 eV to
5.22 eV and 3.23 eV to 3.80 eV with respect to vacuum, respectively (Fig. 59).
An interesting property of these nanostructured CuO materials is the control of the potential en-
ergy levels of the CBs and VBs along with the bandgap by adjusting the size and shape of CuO. This
property is often experimentally observed as a blueshift of the optical absorption band edge as the
dimensions of the nanostructures are reduced. The observed blueshift can be ascribed to the quantum
connement (QC) effect [21,275,276]. The QC in semiconductors results from the geometric conne-
ment of electrons and holes as independent wave-particles or as bound pairs known as excitons. The
normal size of an exciton in bulk, expressed as an exciton Bohr radius, provides an approximate
dimension for the onset of QC effects [277]. When the crystal size is reduced to much smaller than
the Bohr radius for the materials (from 6.6 nm to 28.7 nm for CuO), a strong QC effect occurs resulting
in a larger bandgap. Rehman et al. [278] achieved different sizes of CuO nanoparticles by the precip-
itationpyrolysis method. The sizes are controlled by various annealing temperatures in the desired
atmosphere for 2 h. The sizes obtained ranged from 11 nm (T = 250 C) to 20 nm (T = 600 C). Bandgap
Fig. 58. (a) Cu 2p and O 1s XPS spectra and Cu LVV and O KLL Auger spectra for Cu, CuO, and Cu
2
O [269]. (b) Energy band
structure of CuO [269,270]. Copyright 2012 Institute of Physics.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 257
(direct) edge variation with crystallite size is presented in Fig. 60, in which an obvious blueshift in the
direct band edge is observed as the particle size is reduced. Given that the particle sizes (1120 nm)
are within the range of the Bohr radius of CuO, a strong QC effect occurs, leading to a larger bandgap.
Given that the size of the crystal is larger than the Bohr radius, a weak QC effect occurs resulting in
more subtle changes in the bandgap energy. The bandgap energy changed by the strong and weak QC
effects can be attributed to the direct changes and indirect perturbation to the electron wave function,
respectively [21]. In addition, Buhro et al. [277] indicated that the inuence of the NC shape on QC is
similar to that of the size. Yang and He [85] recently obtained various shapes of nanostructured CuO
by controlling the reactants, reaction temperature, and reaction duration. The observed UVvisible
diffuse reectance spectra and estimated bandgap energy for each shape of CuO is presented in
Fig. 61. The band gaps of the ower-, boat-, plate-, and ellipsoid-like CuO products were estimated
by the authors to be 1.425, 1.429, 1.447, and 1.371 eV, respectively. The effect of the morphology of
CuO nanostructure on its band gap is clearly demonstrated.
Yang et al. [134] also synthesized different scales and shapes of nanostructured CuO by a simple
solution-based method. Crystallite size and morphology are controlled by molar ratio modulation of
Fig. 59. Estimated position of CuO band edges for the study [274] performed at pH 14 and prior reports in the literature
adjusted to pH 14 [274]. Copyright 2012 American Chemical Society.
Fig. 60. Band gap (direct) variation with crystallite size of samples annealed for 2 h [278]. Copyright 2011 Springer.
258 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
NaOH to Cu(NO
3
)
2
, reaction temperature, and concentration of the starting NaOH solution. As shown
in the absorption spectra of the as-prepared CuO products in Fig. 62, the absorption edge is blueshifted
to shorter wavelength with reduced size and changed morphology. This result indicates that not only
the size but also the morphology of the CuO nanocrystals affects the bandgap.
4.3. Optical properties
The optical properties of CuO in the absorption region are dominated by the absorption threshold,
which is dened by the bandgap of the materials. Compared with bulk CuO, the bandgap of nanostruc-
tured CuO is blueshifted, with reported values ranging widely from 1.2 eV to 2.1 eV [279]. Other
researchers have also reported larger bandgap up to 4.13 eV for 10 nm quantum dots [280] and
3.02 eV for well-aligned arrays of CuO nanoplatelets [86]. Therefore, CuO absorbs strongly throughout
the visible spectrum with a slight transparency for bigger bandgap nanostructured samples, which
absorb in the UV region. Notably, the spectral dependence of the absorption at the edge is described
by the usual dependence [278]:
Fig. 61. (A) UVvisible diffuse reectance spectra of (a) ower-like, (b) boat-like, (c) plate-like, and (d) ellipsoid-like CuO. (B)
Plots of (aE
phot
)
2
vs. E
phot
for (a) ower-like, (b) boat-like, (c) plate-like, and (d) ellipsoid-like CuO nanostructures [85].
Copyright 2011 Elsevier.
Fig. 62. UVvisible absorption spectra of the as-prepared CuO products. (a) CuO nanorods prepared at 100 C (length < 1 lm,
breadth = 50100 nm, straight), (b) CuO nanoribbons prepared at 50 C (length > 1 lm, breadth = 3050 nm, curved), (c) CuO
nanoribbons prepared at room temperature (length < 1 lm, breadth = 1530 nm, curved), (d) CuO spherical assemblies
prepared at 100 C (12 lm long), (e) CuO ower-like assemblies prepared at 100 C (1.52 lm long), and (f) CuO spherical
assemblies prepared at room temperature (0.51 lm long) [134]. Copyright 2007 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 259
ahm Ahm E
g

n
where E
g
is the band gap of the material, m is the frequency of the incident radiation, h is Plancks con-
stant, a is the absorption coefcient in cm
1
, A is a constant related to the material and the matrix ele-
ment of the transition, and coefcient n depends on the nature of the transition (n = 1/2 for the direct
allowed transition or 2 for an indirect transition). The absorption coefcient a was obtained using the
relation ad = ln(1/T), where the transmittance T was calculated from the measured absorbance using
the BeerLambert law, B = log
10
(T), where d represents the path length of the wave in centimeter and
was set equal to the cuvette length of 1 cm [278]. To obtain the absorption onset, (ahm)
2
was plotted
against energy hm. Extrapolation of the linear part until its intersection with the hm axis provides the
values of E
g
(Fig. 61B). Typical optical absorption spectra for nanostructured CuO can be observed in
Fig. 61A with a clear absorption edge obviously at the visible region to UV region covering almost the
entire visible spectrum. Importantly, the optical properties of CuO are strongly dependent on sample
preparation and the measurement technique used along with the test temperatures [102,281,282].
In addition to the absorption properties of CuO, photoluminescence (PL) spectra of CuO nanostruc-
tures are also comprehensively reported [167,282288]. Gaashani et al. [282] achieved different mor-
phologies and sizes of CuO nanostructures by rapid thermal decomposition of Cu nitrate [Cu(NO
3
)
2
]
under ambient conditions, in which the size and morphology of the nanostructures were controlled
by changing the temperature and the duration of the decomposition process. The observed PL spectra
of (a) the CuO samples prepared at a xed time (20 min) for various temperatures and (b) the CuO
samples prepared at a xed temperature (400 C) for different times are illustrated in Fig. 63.
The similar PL spectra of all samples, which have three main broad emission bands centered at
305 (4.07 eV), 505 (2.46 eV), and 606 nm (2.05 eV), are shown in Fig. 63. The PL peak at 305 nm
(4.07 eV) is related to the band-edge emission of CuO nanostructures. The three strong emission peaks
located at 489 (2.54 eV), 505 (2.46 eV), and 525 nm (2.37 eV) are due to the band edge emission from
the new sublevels at 300 K or maybe due to the defects present in the CuO nanostructures. The emis-
sion bands extending from 585 nm to 625 nm correspond to deep level defects of CuO. Other emission
peaks centered at 680 (1.81 eV) and 714 nm (1.74 eV), as well as an IR band at 760 nm, are
obtained from CuO nanoparticles, nanorods, and nanobrils, respectively [167,282285]. The PL peak
located at 685 nm (1.81 eV) is related to the interstitial in CuO, whereas the peak at 714 nm (1.74 eV)
is related to the recombination of electrons and holes at oxygen vacancies. The peak at 760 nm is
ascribed to the specic surface effect, which results in the red-shift of the PL emission. In summary,
the PL properties of CuO nanostructures can be controlled by their shape, dimension, and morphology.
QC effect and specic surface effect are the two most reported mechanisms, which can result in the
blueshift and redshift of the PL peak, respectively [284]. Thus, considering these parameters in the
design of CuO-based photo-electronic devices is important.
Fig. 63. The PL spectra of all CuO samples synthesized by rapid thermal decomposition of Cu nitrate [282]. Copyright 2011
Elsevier.
260 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Huang et al. [288] reported the PL properties of a single CuO nanowire grown on a Cu grid by a sim-
ple thermal oxidation method and PL mapping along the length of the nanowire. They found that the
PL peak at 856 nm was only observed at the root region of CuO nanowire as illustrated in Fig. 64a and
b, which was attributed to the Cu
2
O crystals presented at the bottom position of a single CuO nano-
wire. Moreover, the main broad emission peak at 520 nm gradually becomes narrower toward the
tip of the nanowire. Finally, the peak shifted to 500 nm at approximately 13.8 lm. Broadening of
the emission peak at the bottom of the CuO nanowire was caused bythe presence of impurities, such
as Cu/Cu
2
O, which also had emission in this range. As shown in Fig. 64c, the blueshift in the PL peaks
occurs linearly up to 6 lm of the nanowire and remains constant up to 11 lm. Then, a further blueshift
in the PL was observed up to 14 lm. Considering that the diameter of the nanowire changes from the
root position (44 nm) to the tip position (25 nm), the authors ascribed the initial shift to the grad-
ual transformation from Cu
2
O to CuO structure. The constant region is attributed to the intermediate
state, and the ultimate blueshift is directly attributed to the PL of pure CuO structure with enhanced
nanoeffect because of the decrease in the diameter (19 nm) at the tip the CuO nanowire. This phe-
nomenon may be related to the consequence of crystallization process from an impure mixed phase
state to pure crystalline state along the growth direction in accordance with a steady decease in the
nanowire diameter with enhanced surface defects.
Raman spectroscopy is also widely applied to investigate the optical properties of nanoscale mate-
rials because of its sensitive probe to the local atomic arrangements and vibrations of the materials
[125]. CuO with a monoclinic structure belongs to the space group of the C
6
2h
and shows three acoustic
modes (Au + 2Bu), six IR-active modes (3Au + 3Bu), and three Raman-active modes (Ag + 2Bg)
[29,125]. A typical Raman spectrum of CuO with different particle sizes is shown in Fig. 65. Three Ra-
man peaks for sample (a) at 288, 330, and 621 cm
1
are observed. The peak at 288 cm
1
can be as-
signed to the Ag mode, and the peaks at 330 and 621 cm
1
to the Bg modes [289]. However, a
decrease in grain sizes results in broader Raman peaks of the samples and a slight redshift, which
are mainly due to the size effects. This phenomenon is widely observed for nanoscale CuO
[29,125,289].
Overall, understanding the optical properties of CuO nanostructures is very important in their
applications in optical electronic devices.
Fig. 64. (ab) Typical PL spectra collected at three different positions along the growth direction using a step distance of 1 lm
for the scan. The spectra reveal the main luminescence emission properties of the CuO nanowire. (c) PL peak vs. distance for a
single CuO nanowire. The length-dependence of the PL peak is plotted against the length of the investigated nanowire. The inset
shows an optical image of an individual CuO nanowire [288]. Copyright 2010 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 261
4.4. Electrical conductivity
For p-type MOsemiconductors, the charge carriers based on a considerable concentration of the free
holes existing in their VBs are widely accepted. The concentration of the free holes in these materials is
mainly determined by the metal decit concentration (or excess oxygen) within the crystallite sites of
the materials [61,290,291]. This phenomenon is attributed to the deviation from the stoichiometric
composition of the components, which can be induced by regulating the preparation condition of
the material [291]. Given that oxide crystals usually cannot accommodate a large oxide ion, the
nonstoichiometry of p-type CuO is expected to be cation decient. The defect reaction may be repre-
sented by the following equations [292]:
2CuO 1=2O
2
2Cu Cu V
00
Cu
3O
o
Cu Cu CuP
where O
o
, V
Cu
, and P denote lattice oxygen, Cu vacancy, and hole, respectively. Superscripts h,
00
, and
represent effective neutral, negative, and positive charge states, respectively. According to the defect
equation, the concentration of holes and thus the electrical carrier concentration and conductivity of
CuO can be controlled by the partial oxygen pressure during growth and will be highly sensitive to the
presence of absorbed molecules, which shows great potential for chemical and environmental sensing
[293]. For example, resistivity decreases signicantly and carrier concentration increases rapidly with
an increase in oxygen concentration when the oxygen concentration is more than 1.2% (note that Cu
2
O
is the major phase when the oxygen concentration is below 1.2%) [294]. These results were conrmed
by Meyer et al. [256], who obtained Cu oxides with different stoichiometries, particularly the three
crystalline phases, namely, Cu
2
O, Cu
4
O
3
, and CuO, when they varied the oxygen ow during synthesis.
The resistivity of the obtained CuO lms changed from high to low by two orders of magnitude with
increasing oxygen ow as shown in Fig. 66. Moreover, the lowest carrier densities of CuO are
10
17
cm
3
, which also reach 10
20
cm
3
with increasing oxygen ow.
Structural factors, including grain size, grain boundary, lm thickness, specic phase, and dopants,
also remarkably inuence the conductivity of a material [21,292,295297]. Consequently, both the
synthetic techniques and growth conditions can strongly affect the electrical properties of CuO. For
example, CuO lm prepared from acetate and sintered at 300600 C by a chemical solution deposi-
tion technique has the highest conductivity of 10
2
S/cm to 10
3
S/cm. By contrast, lm conductivities
prepared using Cu 2-ethylhexanoate and naphthenate are 10
6
to 10
4
and 2 10
3
to 10
5
S/cm,
Fig. 65. Raman spectra of (a) sample A with a grain size of 10 nm, (b) sample B with a grain size of 30 nm, and (c) sample C with
a grain size of >100 nm at room temperature excited by 488 nm radiation from an argon-ion laser [289]. Copyright 1999 Wiley.
262 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
respectively [292]. These results showed that the conductivity of doped CuO lm increases when
doped with Li
+
and decreases when doped with Al
3+
. Similar results are shown in other MOs, e.g.,
WO
3
[21].
Fig. 66. (a) Resistivity and carrier mobility vs. oxygen concentration. (b) Measured carrier concentrations as a function of
oxygen concentration [294]. Copyright 2008 Elsevier. (c) Specic resistance from van der Pauw measurements as a function of
the oxygen ow for the series of sputtered Cu-oxide samples. Within the different Cu oxide phases, the specic resistance
decreases with increasing oxygen ow, whereas an increasing resistance indicates a phase change (lines are guides to the eye).
(d) Carrier concentrations of the series of sputtered Cu oxide samples determined from Hall measurements. The carrier
concentration increases with increasing oxygen ow and reaches saturation before dropping sharply at each phase change
(dashed lines are guides to the eye) [256]. Copyright 2012 Wiley.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 263
Changing the synthesis parameters, thickness, and roughness of samples can also vary the resistiv-
ity of the prepared CuO samples. Resistivity reportedly increases along with thickness and deposition
temperature of CuO sample because of the irregular grains (or aggregates) that contributes to more
trapping and scattering of free charge carriers [21,295,296]. Jundale et al. [298] also observed a similar
trend in the electrical conductivity of CuO prepared by solgel method and subsequently sintered at
various temperatures ranging from 300 C to 700 C. The room temperature electrical conductivity of
CuO increased from 10
6
(O cm)
1
to 10
5
(O cm)
1
with an increase in annealing temperature from
300 C to 700 C because of the removal of H
2
O vapor, which may resist conduction among CuO grains
(Fig. 67). Moreover, electron density (n) and mobility (l) also increased with the increase in annealing
temperature and were estimated to be of the order of 4.6 10
19
cm
3
to 7.2 10
19
cm
3
and
3.7 10
5
cm
2
V
1
s
1
to 5.4 10
5
cm
2
V
1
s
1
, respectively. The enhancement was due to the tem-
perature-dependent carrier scattering mechanism, which was driven by the intergrain barrier poten-
tial. As illustrated in Fig. 67, the intergranular potential values (scattering potential) was between
0.35 eV and 0.48 eV for CuO lm annealed at 400700 C.
Moreover, carrier motilities tend to signicantly decrease when going from single crystals to poly-
crystalline and nanocrystalline materials by the scattering at grain boundaries and energy barriers at
these boundaries [299]. Other factors, such as lattice strain and crystal distortions, can also affect the
charge movement, resulting in an increase in the resistivity and a decrease in conductivity [21].
Fig. 67. (a) Arrhenius plot of dc log conductivity vs. 1000/T f CuO thin lm annealed at different temperatures (300700 C). (b)
Variation in the log n and log l as a function of temperature for CuO thin lms annealed at 700 C. (c) Plot of loglT
1/2
vs. 1000/T
[298]. Copyright 2012 Springer.
264 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
The fundamental investigation of the electrical properties of CuO nanostructures is crucial for their
future applications in nanoelectronics. Electrical transport measurements have been performed on
individual CuO nanobers and nanowires [236,300302]. Field-effect transistors (FET) based on single
CuO nanobers were congured following several procedures [300]. Two silver plates placed in a side-
by-side parallel arrangement were used as collectors. The electrodes were set onto the surface of heav-
ily doped p-type Si chips capped with a 500 nm oxide layer, where the underlying semi-conducting
silicon was used as a back gate. A single PVA/Cu acetate ber was deposited across the two electrodes
within 10 sand calcined at 500 C for 4 h to obtain the CuO nanober. The deposited ber was posi-
tioned perpendicular to the edge of the electrode. The schematic of the structure of a fabricated nano-
ber FET combined with the measurement circuit along with the currentvoltage (IV) characteristics
under different voltages is shown in inset of Fig. 68a. This gure also presents well-dened transfer
characteristics, showing the p-type semiconductor behavior with weak gating effect. The conductivity
of individual CuO nanober without applied gate voltage V
g
was measured to be 3 10
4
S/cm. Gon-
alves et al. [236] fabricated FET based on a single CuO nanowire. Intrinsic p-type behavior and weak
eld effect response of single CuO nanowire were observed. The conductivity of individual CuO nano-
wire without applied gate voltage V
g
, carrier concentration, and mobility are measured to be
2.5 10
4
S cm
1
, 6.32 10
4
cm
2
V
1
s
1
, and 10
16
cm
3
, respectively. Li et al. [303] also fabri-
cated similar FET based on individual CuO nanowires. The eld effect mobility and carrier concentra-
tion were estimated to be 2.51 10
3
cm
2
V
1
s
1
and 9.04 10
19
cm
3
, respectively. Liao et al. [302]
reported FET based on single and multifunctional CuO nanowires. The IV characteristics of the FET
Fig. 68. (a) Gate-dependent IV measurement of a single CuO nanober contacted with Ag electrodes. Upper inset: schematic
view of a nanober FET conguration. The source and drain contacts are silver stripes, and heavily doped p-type Si substrate
serves as the back gate. Bottom inset: SEM image of a 100-nm-long CuO nanober with a diameter of 60 nm suspending across
the Ag electrodes [300]. (b) I
ds
V
gt
curves of a single CuO nanowire FET. The inset shows I
ds
V
ds
curves of a single CuO nanowire
FET. (c) I
ds
V
ds
curves of the CuO nanowire-TFT device. (d) I
ds
V
gt
curves of the CuO nanowire-TFT device [302]. Copyright 2009
Institute of Physics.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 265
based on single CuO nanowires under different back-gate voltages are shown in Fig. 68b. The conduc-
tance of the nanowire clearly decreases monotonically as the gate potential increases, demonstrating
that the CuO nanowire is a p-type semiconductor. The eld effect mobility (l) is estimated to be
2 cm
2
V
1
s
1
to 5 cm
2
V
1
s
1
at V
g
= 0 V for a single CuO nanowire. However, compared with the
FET based on a single CuO nanowire, the multifunctional nanowire FET device displays the enhance-
ment-mode p-type transistor behavior, with clear linear regions observed in the I
ds
V
ds
output curves
(Fig. 68c). Notably, from the I
ds
V
gt
shown in Fig. 68d, a large on-current (I
on
) of 0.3 lA, transconduc-
tance of 0.2 lS, mobility of 15 cm
2
V
1
s
1
at V
g
= 0 V, and on/off ratio of 100 were obtained, indicating
that the multifunctional nanowire FET device shows better performance because of more conductance
channels in a device.
4.5. Photoelectrochemical (PEC) properties
PEC cells are widely used to investigate the PEC properties of semiconductor materials because
they can convert solar energy into electricity [304,305]. They consist of an anode and a cathode im-
mersed in a redox electrolyte and connected in an external circuit. The redox electrolyte, which sep-
arates the anode and cathode, regenerates the photoactive component of the electrodes. The anode or
the cathode typically consists of a semiconductor that absorbs sunlight, and the other electrode is typ-
ically a metal. A simple schematic illustration of the PEC cells is given in Fig. 69 [305]. When semicon-
ductors are illuminated, they can absorb the photons with energy greater than their band gap, creating
electronhole pairs, which are split by the electric eld in the space-charge region between the semi-
conductor and the electrolyte. A semiconductor to be used as a photoelectrode in a PEC cell must be
chemically stable and should have an optimum band gap so that it may efciently absorb bulk of solar
radiations [305]. CuO, a p-type semiconductor with a bandgap energy from 1.2 eV to 2.1 eV, is very
attractive as a photoelectrode in PEC cells. Compared with TiO
2
[306], which has been the most stud-
ied photoelectrode material, the main advantage of CuO is that it can absorb throughout the visible
region because its bandgap energy is within 1.22.1 eV [304]. To evaluate the PEC properties of the
semiconductors, the photocurrentvoltage measurements along with the SchottkyMott method
are extensively used in semiconductor particle systems with electrochemical charge collection tech-
niques [307].
Nakaoka et al. [272] performed currentvoltage (IV) and MottSchottky curve studies of the PEC
properties of the CuO. The IV experiments showed that the onset of the cathodic photocurrent be-
cause of the reduction of O
2
occurs at +0.21 V, giving a positive potential shift of 0.2 V from the dark
current increase at the same electrode. This phenomenon indicates that the as-prepared CuO lmelec-
trode shows a p-type semiconductor behavior. In the MottSchottky curves, the atband potentials of
CuO lm electrode were found to be 0.31 (pH 7.0), 0.16 (pH 9.2), and 0.01 V (pH 11.0) with a charge
carrier density of 4.0 10
20
cm
3
.
However, PEC properties often strongly depend on surface morphology and particle size in nano-
structured MOs. This phenomenon is evident in ZnO [308]. Reports have indicated that vertically
Fig. 69. Photoelectrochemical cell for conversion of light into electricity [305]. Copyright 2010 American Chemical Society.
266 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
aligned branched ZnO nanorods with higher surface areas are important in obtaining increased inter-
action with light to maintain good electron transport compared with ZnO nanoparticles. Chauhan et al.
[304] investigated the PEC properties of nanostructured CuO lms and their suitability for splitting of
water. They found that small-sized CuO lms with larger grain agglomerates sintered at lower tem-
peratures (400500 C) yield higher photocurrent and are more efcient for photosplitting of water
because of lesser scattering effect and smaller resistance, which leads to better absorption of incident
photon and enhanced photoeffects. Their use for PEC splitting of water is possible only with an exter-
nal bias. This nding is attributed to their improperly aligned band edges to redox levels correspond-
ing to hydrogen and oxygen evolution, respectively. Therefore, a denite bias voltage (V
bias
) is required
to allow spontaneous transfer of charge carriers across CuOelectrolyte junction.
The lm thickness and uniformity of the particles can also strongly affect the PEC properties of the
PEC cell and photon-to-hydrogen generation efciency. Chiang et al. [273] fabricated a low-cost PEC
cell by spin coating CuO nanoparticles on an ITO substrate, where the CuO nanoparticles were made
by ame spray pyrolysis. The results shown in Fig. 70 provide strong evidence that an optimum lm
thickness exists for obtaining the highest photocurrent. An increase in lm thickness results in an in-
crease in the resistance of the lm and greater tendency for the electrons and holes to recombine be-
cause transporting the charge carrier to the external circuit is more difcult. A considerable decrease
in lm thickness may also lead to a decrease in the photocurrent density because of the lack of CuO to
efciently absorb enough light. To observe the effect of uniformity of particle size on the performance
of CuO PEC cell, Chiang et al. compared the as-prepared CuO samples with uniform size to CuO nano-
particles with poor uniformity purchased from SigmaAldrich. The CuO samples and nanoparticles,
with similar thicknesses of 580 nm, were manufactured using the same method as the CuO electrodes.
The comparative results indicated that the CuO photocurrent is 4.27.8 times higher than that of the
purchased CuO samples. The difference between the two samples is attributed to the larger particles
inside the purchased samples, which may lead to an increased probability for the electron and hole to
recombine. The poor sintering in the small particles also leads to a higher resistance for the electron
and hole to transport. However, photocurrent performances reported for CuO nanoparticles of differ-
ent sizes and lm thicknesses achieved by calcination at various temperatures should be compared
cautiously. Chiang et al. [273] demonstrated that a better photocurrent density is obtained at higher
sintering temperature. However, Chauhan et al. observed an opposite result [304], which indicates
that a better photocurrent density is obtained in lower sintering temperature.
Structural factors, such as surface area, crystallinity, defects inside the materials, porosity, and do-
pants, also markedly inuence the PEC properties of nanoscale CuO. The PEC properties of nanoscale
Fig. 70. Net photocurrent density of different thicknesses of CuO/ITO electrodes sintered at 600 C for 3 h [273]. Copyright 2011
Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 267
CuO are strongly dependent on the synthesis techniques and growth conditions. A summary of pho-
tochemical properties of CuO synthesized by different methods as photocathodes for PEC application
is presented in Table 7 [274]. The CuO doped with 2.0 at% Li prepared by ame spray pyrolysis/spin
coating has the highest charge carrier density of 4.2 10
21
cm
3
along with the highest efciency
of 1.30% among other studies. However, nanoparticles from the intrinsic CuO samples have high
charge carrier density ranging from 4.0 10
20
cm
3
to 1.5 10
21
cm
3
and thus lead to a higher
photocurrent density.
Table 7
CuO in PEC application according to Ref. [274].
System CuO 2.0 at% LiCuO CuO CuO CuO 2.0 at%
LiCuO
CuO
Light source 150 W
solar
simulator
150 W xenon arc
lamp
500 W xenon
lamp
150 W xenon
arc lamp
150 W solar
simulator
150 W solar
simulator
150 W
solar
simulator
I
0
(W/m
2
) 1000 8100 8100
d
1000 1000 1000
Filter AM 1.5 G Below 420 nm
c
AM 1.5 G AM 1.5 G AM 1.5 G
Preparation
method
Spinning
disk
reaction/
spin
coating
Grind
power + LiNO
3
solution
a
Electrodeposition Solgel dip
coating
Flame spray
pyrolysis/spin
coating
Flame spray
pyrolysis/
spin coating
Wet
chemical/
spin
coating
Primary
particle
size (nm)
55 17 12 (crystalline
size)
8130
e
150 nm
crystalline
110 nm
65 11 178 48
Porosity (%) 57% 85% of the
theoretical density
38% to 45% of
theoretical
density of CuO
38% 38% 57%
Surface area
(m
2
)
2.34 10
3
1.43 10
3
4.2 10
3
2.31 10
3
Conductivity
(S/cm)
(6.3% 10% 10
2
)
b
(7.9 10
3
) 1.6 10
6
to
2.7 10
6
7 10
5
to
3 10
4
Electrolyte
pH
14 14 14 14 14 14 14
Flatband
potential
(V vs. Ag/
AgCl)
0.57 0.33 0.05 0.13 0.58 0.23 0.1 0.21
Charge
carrier
density
(cm
3
)
9.0 10
20
4 10
18
4 10
20
(4.2 0.6)10
19
1.5 10
21
4.2 10
21
6.1 10
20
Photocurrent
density
(mA/cm
2
)
1.58
(0.5 V vs.
Ag/AgCl)
0.44
(0.4 V
vs. SCE)
0.08
(0.2 V
vs. AgCl)
20.2
(0.5 V
vs. SCE)
1.20
(0.55 V
vs. Ag/AgCl)
1.69
(0.55 V
vs. Ag/AgCl)
1.2
(0.55 V
vs. Ag/AgCl)
Efciency (%) 1.20 0.045 0.2
d
; 1.59
f
0.91 1.3 0.91
Conduction
band edge
(eV)
3.54 3.80 3.24 3.57 3.23 3.23 3.63
VB edge (eV) 5.22 5.15 4.8 5.34 4.87 4.87 4.98
Reference [274] [271] [272] [304] [273] [309] [310]
a
CuO powder from Fischer Reagent grade moistened with appropriate quantity of LiNO
3
solution and pressed into discs with
an area of 1.54 cm
2
.
b
Calculated based on their date.
c
By a colored glass lter.
d
The power is assumed to be similar to Koffyberg and benko.
e
Film is not covered well by the particles.
f
Based on power 1000 W/m
2
.
268 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
4.6. Magnetic properties
The bulk transition metal monoxides, including MnO, FeO, CoO, NiO, and CuO, are all antiferromag-
nets, where CuO occupies a special place because its unique magnetic properties. Unlike MnO, FeO,
CoO, and NiO, CuO has a low-symmetry monoclinic crystal structure and a magnetic susceptibility
temperature behavior that is unusual for 3d antiferromagnets [252]. Bulk CuO is antiferromagnetic
with Nel temperatures from 213 K to 230 K [311,312]. However, although the preliminary magneti-
zation measurements performed by Avoni et al. in 1992 suggest the presence of weak ferromagnetism,
which may explain the unusual behavior of the single-crystal bulk CuO susceptibility [313].
Anomalous ferromagnetic behavior is observed for different CuO nanostructures. Punnoose et al.
[314] investigated the magnetic properties of CuO nanoparticles of nominal size range from 6.6 nm
to 37 nm synthesized by the solgel method combined with high-temperature annealing. For nano-
particles with a diameter (d) < 10 nm, a more rapid lattice expansion exists and the magnetic suscep-
tibility varies as 1/d (Fig. 71), accompanied by a weak ferromagnetic component and hysteresis loops.
For nanoparticles greater than 10 nm, the magnetic ordering is essentially similar to the antiferromag-
netic ordering of the bulk CuO [314]. For CuO nanoparticles <10 nm, the surface spins are expected to
dominate the measured magnetization because of their lower coordination and uncompensated ex-
change couplings. Consequently, a weak ferromagnetic behavior is shown. Rao et al. also studied
the magnetic properties of CuO nanoparticles synthesized by thermal annealing of Cu(OH)
2
with aver-
age particle diameters ranging from 13 nm to 33 nm. Their magnetic measurements reveal the pres-
ence of weak ferromagnetic interaction and the blocking behavior in these CuO nanoparticles, which is
caused by the existence of uncompensated surface pins [315].
According to Refs. [314,316,317], a diameter of 10 nmis a critical size for CuOnanoparticles to show
ferromagnetic behavior. However, the magnetic properties of nanostructures depend not only on size
but also on the morphology and anisotropy of nanostructures. Xiao et al. investigated the magnetic
properties of CuO nanorods with 3040 nm diameter and 100200 nm long hydrothermally synthe-
sized in the presence of sodium citrate [318]. The CuO nanorods with a much larger diameter than
10 nm exhibit an obvious anomalous ferromagnetic behavior in the 5300 K range, in which the effect
of the morphology of CuO nanostructures has an important function in the magnetic properties. Dar
et al. studied the magnetic properties of CuO nanoneedles synthesized by simple hydrothermal
method. Their measurements using a superconductor quantuminterference device magnetometer also
reveal the ferromagnetic behavior of the CuO nanoneedles, in which the coercivity of CuO nanoneedles
at 3 K was estimated to be 42 Oe [81]. The magnetic behavior of CuO nanowires with lengths ranging
from several micrometers and diameters from 50 nm to 120 nm grown by thermal oxidation of Cuis
investigated in Ref. [287]. Hysteresis loops of the nanowires show ferromagnetic behavior from 5 K
to room temperature.
Fig. 71. Plot of magnetic susceptibility at 7 and 350 K vs. 1/d (d is the particle size in nm) [314]. Copyright 2009 American
Institute of Physics.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 269
Room-temperature ferromagnetism is an interesting property because of its various applications.
Following the theoretical predictions by Dietl [319], numerous studies have searched for magnetic
semiconductors with Curie temperatures above room temperature using various host semiconductor
materials doped with 3d transition metals. Recently, room temperature ferromagnetism has been ob-
served in pure (undoped) MOs, such as ZnO, MoO
2
, and CaO, because of moments induced by defects
[320322]. Room-temperature ferromagnetism is also observed in pure CuO nanostructures. Shang
et al. synthesized straw-like CuO nanostructures (nanostraws) on a Cu foil substrate through a simple
solution method and demonstrated that the room-temperature ferromagnetism is associated only
with the nanostructures (not microscale CuO lm) [323]. Gao et al. studied the room-temperature fer-
romagnetism of ower-like CuO nanostructures prepared by the co-precipitation method with post-
annealing in air conrming that the observed room-temperature ferromagnetism in ower-like CuO
nanostructures might originate from oxygen vacancies [324]. In another study from the same group,
room-temperature ferromagnetism is observed in pure CuO nanoparticles prepared by the precipita-
tion method without any ferromagnetic dopant. The ferromagnetism is believed to be caused by the
oxygen vacancies at the surface/or interface of the nanoparticles. The CuO nanoparticle-based ferro-
magnet can be a suitable option for a class of spintronics [325].
5. Applications
This section we focus on the recent developments in the different CuO nanostructures as building
blocks for applications in a wide range of elds. These elds include LIBs, supercapacitors, sensors, so-
lar cells, photodetectors, catalysis, nanouid, nanoenergetic materials (nEMs), eld emissions, super-
hydrophobic surfaces, and removal of arsenic and organic pollutants from waste water. The toxicity of
CuO nanoparticles is also briey addressed.
5.1. Application in LIBs
Nanostructured materials have attracted considerable interest both in fundamental as well as in
applied research areas because of their outstanding physical and chemical properties. In recent years,
using nanostructured materials as electrodes in LIBs (Fig. 72) has been promisingas both energy
density and power density of the battery [36,326,327].
Fig. 72. Schematic illustration of a typical LIB and the electrochemical processes during chargedischarge [326]. Copyright
2013 Royal Society of Chemistry.
270 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
To date, nanosized transition MOs (where the metal may be Fe, Co, Ni, or Cu) have been considered
as promising anodes for LIBs because of their high theoretic capacity and excellent cycling capacity
retention property [36,127,326329]. Among these MOs, CuO is extensively investigated because of
high theoretic capacity (670 mA h/g), lowcost, environmental benignity, easy production, and easy
storage, among others [87,110,127,329]. The large volume variation during the lithium uptake/release
process is one of the most important issues restricting its application in LIBs, which results in severe
mechanical strains and very rapid capacity decay [87,110]. In this regard, considerable efforts have
been made to use nanoarchitectured electrodes to overcome this drawback and improve the electro-
chemical performance. Various CuO nanostructures, including 0D nanoparticles [330], 1D nanowires
and nanorods [79,331], and other complex 2D/3D hollow nanostructures [110,329], have been devel-
oped using different synthetic strategies to optimize the electrochemical performance of CuO. Well-
dened unique hierarchical or hollow nanostructured CuO has been veried as a good candidate for
LIBs because ots high surface area, low material density, and surface permeability [87,332].
Hierarchical CuO nanostructures with various morphologies, such as leaf, shuttle, ower, dande-
lion, and caddice clew, were obtained by a simple self-assembled synthetic method and used as anode
materials for LIBs [127]. The electrochemical performances of electrodes are tightly related to their
morphology. The leaf- and shuttle-shaped CuO nanostructures exhibit high initial coulombic efcien-
cies (69.1% and 69.5%) but relatively low discharge capacities (740 and 773 mA h/g) because of their
small specic surface. These characteristics limit the large-scale formation of inactive solid electrolyte
interphase (SEI) lm and simultaneously reduce Li
+
accommodation. The dandelion- and caddice
clew-like CuO have superior discharge (927 and 952 mA h/g) and cycling properties to the leaf- and
shuttle-shaped CuO, especially at high rates, although their initial coulombic efciencies (62.1% and
62.5%) are slightly little lower. The high specic surface and porosity of hierarchical CuO are suggested
to result in large contact area for CuO/electrolyte and shorter diffusion length of Li
+
.
Guan et al. [87] hydrothermally synthesized hierarchical CuO hollow microspheres at 120 C for
24 h without using any surfactants or templates. The as-synthesized CuO microspheres were used
as the anode material in LIBs. The cycling performances of the prepared samples are shown in
Fig. 73. The CuO hollow sphere electrodes exhibited superior electrochemical performance to the
CuO solid sphere or CuO/Cu
2
O hollow sphere at a high current of 670 mA h/g. After 30 cycles, CuO hol-
low spheres delivered a reversible capacity of 457 mA h/g with a high coulombic efciency of 99%,
which is much higher than the theoretical capacity of graphite (372 mA h/g). The reversible capacity
of CuO solid spheres rapidly dropped from 373 mA h/g to 193 mA h/g, whereas the CuO/Cu
2
O hollow
sphere only delivered a reversible capacity of 409 mA h/g. This performance is mainly attributed to the
synergetic effect of small diffusion lengths in building blocks of nanorods and proper void space that
buffers the volume expansion.
Fig. 73. Cycling performances of CuO solid sphere, CuO hollow sphere, and CuO/Cu
2
O hollow sphere at a current density of
670 mA h/g [87]. Copyright 2011 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 271
Moreover, 3D hierarchical CuO hollow micro/nanostructures were realized using a bimolecular-as-
sisted green method using CuSO
4
, NH
3
H
2
O, and tyrosine instead of toxic and dangerous reagents
[110]. The 3D hierarchical CuO hollow micro/nanostructure-based anode material for LIBs exhibited
a high initial discharge capacity of 560 mA h/g and improved electrochemical performance toward
Li uptakerelease. The improved electrochemical performance is primarily caused by the novel hollow
micro/nanostructures of CuO. First, the 3D conguration of hierarchical hollow micro/nanostructures
can be considered as a 3D current collector network, which provides negligible diffusion times (short
diffusion length) and enhances electronic conductivity. Second, these novel nanostructures can also
act as an elastic buffer to relieve the strain associated with the volume variations during Li uptakere-
lease, suppress particle pulverization, maintain electronic contact, and guarantee good power perfor-
mance and capacity retention. These ndings prove that novel hollow micro/nanostructure electrode
materials are advantageous for future high-performance LIBs.
Hu et al. [333] were the rst to achieve 3D dendrite-shaped CuO hollow micro/nanostructures
using a Kirkendall effect-based approach. These micro/nanostructures composed of nanocubes outside
and a dense lm inside and demonstrated as anode material for LIBs that exhibited signicantly im-
proved cyclability and high capacity. As shown in Fig. 74a, the dendrite-shaped CuO hollow structures
obtained at 350 C for 0.25 h displayed an initial discharge capacity of 1503.9 mA h/g at a current den-
sity of 335 mA h/g, which was greater than the theoretical value of 670 mA h/g. After extended cycling
up to 50 cycles, the dendrite-shaped CuO hollow structures retained a capacity of 300 mA h/g, that is,
57.9% retention of the discharge capacity. Remarkably, the average coulombic efciency of 50 cycles
was up to 97.0%. The rate capability was also studied as shown in Fig. 74b. These observations clearly
verify the advantages of novel CuO hollow structure, which consisted of nanocubes with structural
robustness. They lead to fast electrolyte diffusion and provide higher electrode/electrolyte contact
area and a structural buffer for large volume expansion.
Wang et al. [334] used the porous microspheres of CuO with hydrothermally prepared dandelion-
like hollow structure as anode materials to characterize their LIB performance. The LIBs had an initial
discharge capacity of 1274 mA h/g at a current density of 100 mA h/g and showed stable good capacity
retention with a reversible capacity of over 600 mA h/g even after 50 cycles at a current rate of C/5
(completing the charge or discharge process in 5 h; 1 C = 670 mAg/h), which is much higher than
the commonly used graphite materials. These results indicate that the CuO hollow microspheres
are promising anode materials for LIBs. Other complex hollow structures [332] and hollow nano-
spheres assembled by nanoparticles [335] are also reported as anode materials for LIBs. Initial dis-
charge capacities of 1167 and 1134 mA h/g are achieved, respectively. Furthermore, both show
superior cycle performance to 10 nm CuO nanoparticles. These results further prove the advantages
Fig. 74. (a) Variation in discharge/charge capacity vs. cycle number for sample prepared at 350 C for 0.25 h (S350-1) at
different rates of 0.5 and 1 C. (b) Capacity delivered upon cycling at various rates by sample S350-1 cycled between 0.02 and 3 V
[333]. Copyright 2010 Elsevier.
272 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
of hollow structures, which can provide enhanced accommodation of the strain energy associated
with lithium insertion/removal and offer a suitable electrode/electrolyte contact area, resulting in
good retention of charge/discharge capacity.
Although improved LIB performance can be achieved for CuO as anodes through synthesis of un-
ique nanostructures, a high irreversible capacity loss is generally observed during the initial
chargedischarge cycles (especially during the rst cycle). This capacity loss is the main bottleneck
that restricts the application of CuO nanostructures as anode materials [336,337]. Dbart et al.
[338] reported that CuO is reduced to Cu nanograins during the charge process through an interme-
diate Cu
2
O, and the reduced metallic Cu re-oxidized back to Cu
2
O rather than to CuO during discharge.
This phenomenon results in capacity fading because of the lower theoretical capacity of Cu
2
O
(375 mA h/g) compared with that of CuO (670 mA h/g). Several other studies have also observed
and conrmed this charge/discharge mechanism [222,337,339]. Wang et al. [336] recently investi-
gated the lithiationdelithiation processes of CuO nanowire-based LIBs using an individual CuO nano-
wire anode to construct an LIB prototype within a TEM (Fig. 75). Their results demonstrated that both
the irreversible volume expansion and the incomplete conversion of Cu to CuO hampered the revers-
ible electrochemical reaction and led to detrimental capacity fading during the rst cycle (Fig. 76). By
contrast, good reversibility was observed during the second chargingdischarging cycle (Fig. 77). The
reversible volume and phase changes and SEI layers of nearly the same thicknesses after the rst cycle
indicate a complete electrochemical reaction, resulting in high reversibility.
Therefore, combining CuO nanostructures with other active nanomaterials with high capacity (e.g.,
carbon-based materials) is necessary to form novel nanocomposite electrodes, which possess the
advantages and mitigate the limitations of both components. In CuO/carbon nanocomposite elec-
trodes, the carbon nanostructures possess excellent electronic conductivity, which is benecial for
charge transport of CuO. Moreover, they also serve as physical support to prevent the breakdown of
CuO nanostructures during Li
+
insertion/extraction [340].
Zheng et al. [329] investigated the LIB performance of CuO nanomicrospheres and carbon nanotube
(CNT) composite. Their results showed that the CuOCNT composite spheres exhibit remarkably en-
hanced cycling and rate performance compared with CuO spheres when used as anode materials in
LIBs. The enhanced LIB performance is attributed to the 3D CNT network, which acts as a 3D current
collector network and an elastic buffer to relieve the strain during Li uptakerelease. Wang et al. [340]
compared the electrochemical performance of self-assembled CuO and CuO/graphene urchin-like
structures used as anode materials in LIBs and prepared by a simple solution method. The CuO/graph-
ene nanocomposite exhibit remarkably enhanced cycling and rate performances. The enhancement of
the electrochemical performance of CuO/graphene nanocomposite is attributed to the 3D electrically
conductive networks of graphene as well as the unique nanomicrostructure of the CuO/graphene
nanocomposite in which the CuO nanomicroowers are enwrapped by a thin layer of graphene as
an elastic buffer.
Ko et al. [341] studied the electrochemical performance of mesoporous CuO particles and mesopor-
ous CuO particles/CNT nanocomposites synthesized by a simple solution-based method. TEM analyses
indicated that the synthetic processes involve the rst formation of Cu(OH)
2
on the surface of CNTs.
Subsequently, these formed Cu(OH)
2
/CNTs were transformed into the mesoporous CuO particles
Fig. 75. (a and b) Schematic illustration and corresponding TEM image of an electrochemical device constructed inside the TEM.
This device consists of a CuO nanowire anode, a droplet of the ionic liquid electrolyte (ILE), and a Li metal electrode [336].
Copyright 2012 Royal Society of Chemistry.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 273
wrapping the CNTs. The fringe spacing of 0.23 nm in the HRTEM image in Fig. 77 is attributed to the
spacing of the (111) plane of the CuO cubic structure, whereas the 0.33 nm fringe spacing is the inter-
layer distance of the multi-wall CNTs (MWCNT). These results clearly indicate the successful forma-
tion of threaded bead structure of the CuO/CNT nanocomposite. Raman spectra of the relative
intensity (D/G) of the D and G bands of MWCNTs revealed that the surfaces of the MWCNTs were
chemically bonded with the CuO surfaces. Under the same electrochemical testing conditions,
CuO/CNT nanocomposite electrodes showed higher reversible capacity and remarkably enhanced rate
capabilities compared with mesoporous CuO particles/C blend (Fig. 78). This excellent performance of
the CuO/CNT nanocomposite electrode is due to the strong binding of the CuO to the CNTs, and their
micropores have morphological stability during the repeated battery cycles and high electrical
conductivity.
In addition tothe aforementioned examples of CuO nanostructures and their related nanocompos-
ites, various other CuO nanostructures, including 0D nanoparticles, 1D nanotube, 1D nanowires, 1D
nanorods, 2D nanosheets, 2D nanoplatelets, and 3D nanoowers and their nanocomposites, are exten-
sively investigated as anode materials for LIBs. The electrochemical properties of other CuO nanostruc-
tures and their nanocomposites used as active materials of LIB anodes are summarized in Table 8.
5.2. Application in supercapacitors
Pseudocapacitors (Fig. 79) is a type of supercapacitors, in which fast and reversible faradic pro-
cesses occur and involve the passage of charge across a double layer, resulting in faradaic current pass-
ing through the supercapacitor cell [1,363,364,365].
Pseudocapacitor has attracted increasing interest because of its higher power density, longer life
cycle, and higher energy efciency than that of secondary batteries. Two major types of materials,
including conducting polymers and MOs, have been widely used as electrode materials for application
in pseudocapacitors [363,365]. Given their favorable capacitive characteristics and environmental
friendliness, MOs (RuO
2
, SnO
2
, MnO
2
, NiO, and CuO) were considered as promising electrode materials
for pseudocapacitors [363,365]. Among these MOs, CuO can be an attractive candidate because of its
lowcost, chemical stability, and easy preparation in diverse shapes with nanosized dimensions [44].
Zhang et al. [366] studied the electrochemical properties of CuO with different morphologies syn-
thesized by chemical precipitation method as electrode materials used in supercapacitors. Specic
capacitance was reported to be higher in ower-like CuO than those of CuO sheets and globular
Fig. 76. Schematic illustration of Li-insertion and Li-extraction processes and corresponding phase changes at different reaction
stages: (a) rst and (b) second cycles [336]. Copyright 2012 Royal Society of Chemistry.
274 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
CuO. They added that ower-like CuO also exhibited excellent cycle performance at a high current
density of 10 mA/cm
2
. The excellent performance is related to the special ower-like structure of
CuO, which can lead to high surface area resulting in remarkable enhancement of electrochemical
properties. Zhang and Zhang [367] also achieved similar morphology-dependent electrochemical
properties of CuO being used in supercapacitor. Wang et al. [368] studied the supercapacitance of
Fig. 77. (a) Schematic illustration showing the preparation process for the CuO/CNT nanocomposite with the threaded bead
structure. (b) TEM image of the Cu(OH)
2
coating on the CNTs. (c) TEM image of the CuO/CNT nanocomposite produced by
transformation of the Cu(OH)
2
coating layer into the CuO particles. (d) HRTEM image of Fig. 77(c). (e) Raman spectra of the as-
received CNTs and the CuO/CNT nanocomposite [341]. Copyright 2012 Wiley.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 275
CuO nanosheets grown directly on Ni foam synthesized by template-free growth method. The as-pre-
pared CuO nanosheet electrode exhibited very high specic capacitance of 569 F/g at a current density
of 5 mA/cm
2
. The high capacitance was due to the unique structure of the CuO, which has high utili-
zation and superior properties for electrolyte diffusion. Huang et al. [369] reported a remarkable
enhancement in supercapacitance based on CuO nanosheets supported on Ni foam by doping CuO
nanosheets with Ag. Their results indicated that the specic capacitance of CuO nanosheets electrode
after Ag doping was 689 F/g at 1 A/g and 299 F/g at 10 A/g, respectively. These ndings were higher
than those of unmodied CuO nanosheet arrays (418 F/g at 1 A/g and 127 F/g at 10 A/g). The enhance-
ment was attributed to the addition of Ag particles, which facilitated the fast electron transport be-
tween the current collectors and the active materials and thereby enhanced electrical conductivity.
To further improve the electrochemical performances of CuO-based pseudocapacitors, nanocom-
posites are comprehensively investigated as promising electrode materials for pseudocapacitors by
integrating nano CuO into different carbon nanostructures. The carbon nanostructures have been
demonstrated not only to serve as the physical support of CuO but also to provide the channels for
charge transport in such CuO/carbon nanocomposite electrode [365].
Recently, Zhang et al. [44] mixed porous CuO nanobelts prepared by a wet chemical process with
single-wall CNTs (SWCNTs) to fabricate pseudocapacitor electrodes. They compared the electrochem-
ical properties of different electrode congurations, including CuO nanobelts mixed with SWCNTs
(named as CuO/CNT electrode), pure CuO nanobelts (named as CuO electrode), and CuO nanobelts
mixed with carbon black and the PVDF binder (named as CuO/CB electrode). The CuO/CNT electrodes
exhibited the best cycling stability among the three types of electrodes tested (Fig. 80). The specic
capacitance of the electrodes after 1000 charge/discharge cycles was shown to follow CuO/CNT
electrode (128 F/g) > CuO/CB electrode (101 F/g) > pure CuO electrode (73 F/g). The best properties
Fig. 78. Electrochemical properties of the CuO/C blend electrode and the CuO/CNT nanocomposite electrode. (a) Voltage
proles of the CuO/C blend electrode (1st, 5th, 10th, 30th, 60th, 80th, and 100th cycles) at a rate of 0.1 C. (b) Voltage proles of
the CuO/CNT nanocomposite electrode (1st, 10th, 60th, 80th, and 100th cycles) at a rate of 0.1 C. (c) Cycling performance of the
CuO/C blend electrode (black) and the CuO/CNT (gray) electrode at 0.1 C. (d) Rate capabilities of the CuO/C blend electrode
(black) and the CuO/CNT nanocomposite electrode (red) at 0.230 C [341]. Copyright 2012 Wiley.
276 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Table 8
Electrochemical properties of CuO nanostructures used as active materials of LIB anodes, 1 C = 600 mA/g (EC: ethylene carbonate,
DEC: diethylene carbonate, DMC: dimethyl carbonate, PC: propylene carbonate, DME: 1,2-dimethoxyethane).
Morphology Synthesis method Electrolyte
solution
Discharge
capacity in
rst cycle
(mA h/g)
Reversible
capacity in n
cycle (mA h/g)
Current
density/rate
Ref.
Sheaf-like Hydrothermal process 1 M LiPF
6
-
EC/DMC/
EMC
965 580 (41) C/2 [111]
Nanoower-like Thermal oxidation 1 M LiPF6-
EC/DME
1000 530 (50) 0.02 mA/
cm
2
[342]
Nanoplatelets Hydrothermal process 1 M LiPF
6
-
EC/DEC
788 300 (50) 60 mA/g [229]
Leaike Hydrothermal process 1 M LiPF
6
-
EC/DMC
1028 440 (30) 0.1 C [99]
Nanorods Chemical precipitation
method
1 M LiPF
6
-
EC/DMC/
PC
1100 550 (100) 50 mA/g [79]
Nanoribbons Solution-based method 1 M LiPF6-
EC/DMC/
DEC
866 608 (275) 175 mA/g [343]
Ultrane nanowire Chemical precipitation
method
1 M LiPF6-
EC/DEC
1200 660 (70) 0.3 C [344]
Nanotube Electrochemical method 1 M LiPF6-
EC/DME
911 417 (30) 0.02 mA/
cm
2
[345]
Polycrystalline
nanowires
Chemical precipitation
method
1 M LiPF6-
EC/DMC/
DEC
1040 650 (100) 0.5 mA/cm
2
[331]
Flowerlike Solid-state thermal
conversion of precursors
1 M LiPF6-
EC/DEC
850 550 (50) 0.15 mA/
cm
2
[346]
Networklike Solid-state thermal
conversion of precursors
1 M LiPF
6
-
EC/DMC
970 520 (50) 0.2 mA/cm
2
[347]
Nanosheet Composite-hydroxide-
mediated approach
1 M LiPF6-
EC/DEC
1198 420 (40) 0.1 C [348]
Pillowshaped Solid-state thermal
conversion of precursors
1 M LiPF
6
-
EC/DMC
770.3 320 (50) 0.1 C [156]
Hierarchical
nanoplates
Hydrothermal process 1 M LiPF
6
-
EC/DMC
1009 600 (20) 0.1 C [349]
Interlaced nanodiscs Chemical precipitation
method
1 M LiPF
6
-
EC/DMC
971 290 (20) C/5 [350]
Leaike Electrochemical route 1 M LiPF6-
EC/DEC
1063 >500 (30) 0.1 C [171]
Urchin-like particles Sequential dissolution
precipitation process
1 M LiPF6-
EC/DMC
800 >560 (>50) 150 mA/g [62]
Cog-shaped
microparticles
Microemulsion-
mediated process
1 M LiPF6-
EC/DMC
1052 810 (10) 0.05 C [351]
Bird nest-like
spheres
Chemical precipitation
method
1 M LiPF6-
EC/DMC
958 392 (50) 0.1 C [352]
Needle-like spheres Chemical precipitation
method
1 M LiPF6-
EC/DMC
1047 441 (50) 0.1 C [352]
Bundlelike Solid-state thermal
conversion of precursors
1 M LiPF6-
EC/DMC
1179 666 (50) 0.3 C [158]
Pineneedle-like
arrays
Anodic polarization
route
1 M LiPF6-
EC/DMC/
EMC
1015.6 583.1 (100) 2 C [353]
Mesoporous
nanosheet cluster
array
Ammonia vapor-phase
corrosion route
1 M LiPF6-
EC/DMC/
EMC
965.7 639.8 1 C [165]
Hierarchically
nanosheet-
assembled
gearlike pillar
arrays
Ammonia vapor-phase
corrosion route
1 M LiPF6-
EC/DMC/
EMC
892 651.6 (100) 0.5 C [164]
(continued on next page)
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 277
of CuO/CNT electrode were due to its open networks with large surface area formed by the unique
entangled 1D CuO nanostructures. They further fabricated exible electrodes by depositing a mixture
of CuO nanobelts and SWCNTs as network lms onto pure SWCNT lms without the addition of any
binders (Fig. 81). Under the same electrochemical testing condition, these CuO-on-SWCNT exible
electrodes also showed stable cycling performance upon repeated charging/discharging process and
had much higher specic capacitances than those of pure SWCNT electrodes.
Zhao et al. [370] fabricated pseudocapacitor electrode based on hydrothermally synthesized leaf-like
porous CuOgraphene nanocomposite. The leaf-like porous CuOgraphene nanocomposite device exhib-
ited superior performance to leaf-like porous CuOwithout graphene in terms of specic capacitance and
cyclability. The superior electrochemical performance of CuOgraphene nanocomposite can be ascribed
to the higher conductivity of the nanocomposite, which promotes fast and effective hydroxyl-ion diffu-
sioninthehost material andgreatlyimproves volumechangecushionandstructurestabilityuponcycling
because of the robust matrix and their layer-by-layer structures. Moreover, high capacity retention up to
95.1% can be maintained after 1000 continuous chargedischarge (Fig. 82). This result is due to the im-
proved electrical contact by graphene and mechanical stability of the layer-by-layer structure.
In addition to the aforementioned examples, other CuO nanostructures [371,372] and their com-
posites, such as CuO-doped activated carbon [373], CNT/CuO composites [374], and CuOPAA hybrid
lms [375], have also been reported as promising electrode materials for supercapacitors.
Table 8 (continued)
Morphology Synthesis method Electrolyte
solution
Discharge
capacity in
rst cycle
(mA h/g)
Reversible
capacity in n
cycle (mA h/g)
Current
density/rate
Ref.
Porous CuO
nanorods
Solid-state thermal
conversion of precursors
1 M LiPF6-
EC/DMC
>1200 654 (200) 300 mA/g [354]
CuO nanoparticles/
graphene
composite
In situ chemical
synthesis approach
1 M LiPF6-
EC/DMC
817 423 (50) 0.1 C [355]
Pinenut-like CuO
nanoparticles/
graphene
composite
Solution-based method
combined with
electrochemical
exfoliation route
1 M LiPF6-
EC/DMC
1400 >500 (30) 0.2 C [356]
CuO/C lms with
sisal-like nano/
micro binary
structures
Solution immersion and
subsequent heat
treatment
1 M LiPF6-
EC/DMC
1084 671 (50) 0.6 C [357]
CuO-interlaced
nanodiscs/
MWCNT
composite
Chemical precipitation
method
1 M LiPF6-
EC/DMC
1025 440 (20) 0.2 C [350]
CuO NWsCNTs Solid-state thermal
conversion of precursors
1 M LiPF6-
EC/DMC
1100 576 (50) 0.1 C [151]
CuO/C microspheres Heat
treatment + oxidation
process
1 M LiPF6-
EC/DMC
1150 470 (50) 100 mA/g [358]
Sheet-like CuO/
graphene
composite
Microwave irradiation
method
1 M LiPF6-
EC/DMC
1192 801 (40) 70 mA/g [359]
Leaf-shaped
nanoparticle/
ordered
mesoporous
carbon CMK-3
Precipitation method 1 M LiPF6-
EC/DEC
1590 801.3 (180) 0.1 C [360]
Nanoribbon-like
CuO/graphene
Converting CO
2
-based
precursor method
1 M LiPF6-
EC/DMC
>1050 600 (>5) 100 mA/g [361]
CuO/C coreshell Carbonization 1 M LiPF6-
EC/DMC
730 500 (10) 0.2 C [362]
278 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
5.3. Application in sensors
5.3.1. Application in gas sensors
Semiconducting MO gas sensors are advanced sensor types. Since the pioneering reports on their
gas sensitivity at elevated temperatures in the 1950s [376], gas sensors based on these MOs for detec-
tion of hazardous, ammable, and toxic gases has been increasingly developed because of their small
dimension, lowcost, and high compatibility with microelectronic processing [38,377,378]. The physi-
cal and chemical properties of nanoscale materials are strongly inuenced by their dimensional con-
straints and morphologies. Gas sensors based on nanostructured MOs are expected to exhibit better
performance than their bulk or microcounterparts. In this regard, MO nanostructures, such as 1D
Fig. 79. Schematic diagram of the pseudocapacitor [1]. Copyright 2012 Elsevier.
Fig. 80. (a) Galvanostatic charge/discharge curves measured with a current density of 1 A/g for different electrodes. (b) Cycling
performance for different electrodes at a current density of 1 A/g in 1.0 M LiPF6/EC:DEC [44]. Copyright 2011 American
Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 279
nanowires, 1D nanorods, 2D nanosheets, 3D nanoowers, and complex hierarchical architectures,
have attracted considerable interest for their potential as building blocks for fabricating gas sensors
because of their high effective surface areas [38,376379]. Among various MOs, CuO has the unique
property of being intrinsically p-type and is advantageous because of its lowcost, high stability, non-
toxicity, and capability for electron transfer [380]. Therefore, CuO nanostructures have been exten-
sively investigated as good candidates for sensing applications. Gas sensors based on CuO
nanostructures were operated by measuring the resistance changes when they are exposed to reduc-
ing or oxidizing gases [293]. A suitable operation temperature is crucial to the gas sensing capability of
CuO because it enhances the sensitivity and reduces the response and recovery time [381]. Coating of
noble metal catalysts, such as Ag, Au, Pd, and Pt, on the surface of CuO nanostructures will enhance
their sensing properties because they can act as adsorption sites for analytes or as surface catalyst
[381,382].
The sensing mechanism of gas sensors based on p-type CuO is contrary to n-type MOs
[293,381,382]. In the air, oxygen molecules are adsorbed onto the CuO surface forming oxygen ions
O

; O

2
; O
2
, charging it negatively. If a reducing species, such as NH
3
is present, it oxidizes and re-
leases a negative charge (electron). In this regard, the electron transfer from oxygen ions
O

; O

2
; O
2
to CuO results in decreased hole density and thus decreased conductance. Meanwhile,
the number of oxygen ions O

; O

2
; O
2
absorbed on the surface of CuO is also reduced, which leads
to the decrease in the magnitude of the negative quasi-gate resulting in further decrease inconductiv-
ity. Exposure to oxidizing gas, such as NO
2
, likely results in the occurrence of charge transfer from the
CuO to the NO
2
as governed by NO
2
e

!NO

2
. The electron transfer from the CuO to the NO
2
results
in the lowering of the quasi-Fermi level and thus increases free hole density and enhances conduc-
tance. Additionally, the negatively charged chemisorbed NO
2
molecule at the CuO surface may act
Fig. 81. (a and b) Optical images of CuO nanobelt-based two-layered SWCNTs and CuO nanobelts mixed with SWCNT (9:1
weight ratio) electrode. (c) Edge of SWCNTs and CuO nanobelts. (d) Galvanostatic charge/discharge curves measured with a
current density of 5 A/g for different electrodes. (e) Cycling performance for SWCNTs and CuO nanobelts mixed with SWCNT
electrode at a current density of 5 A/g in 1.0 M LiPF6/EC: DEC [44]. Copyright 2011 American Chemical Society.
280 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
as a negative gate bias and increase CuO conductance. The proposed sensing mechanisms of NH
3
and
NO
2
are shown in Fig. 83.
A number of studies have focused on the development of novel CuO nanostructures for the
detection of a wide range of gases, such as organic gases [381], hydrogen sulde (H
2
S)
Fig. 82. (A) Schematic illustration of the formation of leaf-like CuOgraphene nanostructures. (B) TEM images of (a) CuO and (b
and c) leaf-like CuOgrapheme and (d and e) HRTEM images of CuOgraphene. (C) Rate capacitance of CuO and CuOgraphene
with increasing current densities. (D) Cycle life of CuOgraphene at 2 A/g in a 6 M KOH solution [370]. Copyright 2013 Royal
Society of Chemistry.
Fig. 83. Schematic of the proposed sensing mechanism of NH
3
and NO
2
[293]. Copyright 2010 American Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 281
[133,197,230,82,384,385,387], CO [285,302,383,385], NO
2
[383,389], ethanol gas [183,244], NH
3
[386], HCN [144,390], and humidity [168,388].
Zhu et al. [381] reported gas sensors based on 3D hierarchically porous CuO microspheres synthe-
sized via a facile CTAB-assisted hydrothermal route followed by calcination. Gas sensing tests demon-
strated that the sensors based on these hierarchically porous CuO microspheres showed higher
responses to ethanol, propanol, and acetone compared with that of commercial CuO powder. The re-
sponse and recovery times were below 20 s for 200 ppm of the tested target gases. The excellent gas
sensing performance is mainly caused by the novel 3D architectures with high surface area that facil-
itate mass transport. Their sensing properties can be further enhanced by Ag nanoparticle loading on
the surface of CuO microspheres. These results highlight the potential applications of the as-prepared
CuO microspheres in monitoring ammable and toxic organic gases.
Yang et al. [82] investigated the gas sensing performance of CuO nanorods with 15 nm to 20 nm
diameter and 60 nm to 80 nm long synthesized by a microwave-assisted hydrothermal method and
observed a higher sensing response to ethanol and ethyl-acetate than to other target gases (Fig. 84).
This high response was attributed to the small size of the CuO nanorods. Liu et al. [160] reported sim-
ilar good performance toward ethanol detection based on novel worm-like CuO microstructures.
Qin et al. [230] indicated that porous CuO hollow spheres with hierarchical pores synthesized by a
simple one-pot template-free method exhibit excellent sensing performance toward H
2
S, including
low detection limit, high sensitivity, and rapid response and recovery time. This excellent sensing per-
formance is attributed to the special structure with open macropores for gas molecule buffering, mes-
opores and micropores for fast H
2
S molecule diffusion and effective sensing, and good conductivity of
the quasi-single-crystal nanosheets for fast carrier transportation.
Kim et al. [383] fabricated CuO nanowire-based gas sensors providing a novel and simple approach
to detect pollutant gases. The CuO nanowires were rst synthesized on a Cu foil by thermal oxidation.
Then, they were separated from the surface by ultrasonic treatment and immersed in a solution con-
taining isopropyl alcohol. The solution droplet was placed onto the alumina substrate and dried. The
substrate had a microheater (beneath the CuO nanowires) and two Au electrodes. The resistance of
CuO nanowire-based gas sensor increase when exposed to 10100 ppm reducing CO and 65 ppm oxi-
dizing NO
2
. This study provides a novel and simple method for detecting pollutant gases emitted from
gasoline and diesel engines.
Fig. 84. Schematic illustration of fabrication of the gas sensor using CuO nanorods as sensing material [82]. Copyright 2011
Elsevier.
282 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Liao et al. [302] tested the sensing properties of a single CuO nanowire synthesized via thermal
oxidation. The CuO nanowires were removed from the substrate by sonication and then dispersed
in ethanol. A single CuO nanowire sensor exhibited the selectivity to CO rather than other tested gases
(Fig. 85). The sensor has highest sensitivity at 200 C with a response time less than 10 s when exposed
to CO gas. Aslani and Oroojpour [285] reported morphology-dependent CO gas sensors based on
different CuO nanostructures. Their gas sensitivity tests suggested that the CuO nanostructures, espe-
cially cloud-like morphology, are highly sensitive to CO. The enhanced sensing performance of CuO
with cloud-like morphology is due to a synergic effect between small crystallite size/high surface area
and potential barrier modication. This result indicates that the performance of resistive CuO-based
Fig. 85. Response of a single CuO nanowire sensor to different gases with a xed concentration of 500 ppm [302]. Copyright
2009 Institute of Physics.
Fig. 86. (a) Schematic diagram of the setup for the assembly of two vertically aligned CuO nanowire arrays. Two pieces of
nanowire array were attached to a Cu plate and the micromanipulator tip. The distance between the two arrays can be adjusted
by activating the manipulator. (b) In situ observation of the assembly process when pushing one nanowire array to the other. (c)
IV curves measured at different stages of the assembly process [384]. Copyright 2008 American Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 283
sensors to CO gas can be controlled by tuning the morphology of CuO nanostructures. Raksa et al.
[244] fabricated ethanol gas sensors based on vertically aligned CuO nanowires synthesized by
annealing Cu foils. The sensor was tested at 200280 C in ethanol gas with a concentration of
1001000 ppm. The response and recovery times were 100 and 120 s, respectively, with optimum
sensitivity of 1.5 under 1000 ppm ethanol vapor at 240 C. Hsueh et al. [183] reported the fabrication
of crabwise CuO nanowires on glass substrate for ethanol gas sensing. A gate between two Cu thin lm
pads was created by patterning. After thermal oxidation process, the CuO nanowires on the two pads
both grew vertically and laterally connected to each other. The sensing response was 1.27, which was
measured at 300 C under 1000 ppm ethanol gas.
Chen et al. [384] fabricated vertically aligned CuO nanowire arrays for the detection of H
2
S. The
CuO nanowires were synthesized by directly heating Cu foils in air at 500 C for 5 h. Two pieces of
CuO nanowire arrays were assembled by a micromanipulator monitored through an SEM, as shown
in Fig. 86a and b. As one nanowire array approached to other, more and more nanowires were con-
nected, and the resistance of the whole sensor was reduced as implied by the increased slope in IV
curves, as illustrated in Fig. 86c. Normalized response curves of sensor to H
2
S at 160 C are shown
in Fig. 87. This device showed a detection limit of 0.5 ppm. When the nanowires are exposed to reduc-
ing H
2
S gases at low concentrations, the device conductivity decreases. By contrast, when the sensors
are exposed to H
2
S with a concentration higher than 5 ppm, surface reaction begins and thus a layer of
CuS is formed. This phenomenon increases the effective contact area between the adjacent nanowires,
resulting in an increase in conductivity. This sensor also demonstrated good selectivity to H
2
S com-
pared with the response to H
2
, CO, and NH
3
(see Figs. 88 and 89).
Steinhauer et al. [385] reported novel gas sensing devices based on CuO nanowires synthesized on-
chip by thermal oxidation of electroplated Cu microstructures. They observed that the as-fabricated
gas sensing devices could detect CO as low as 10 ppm and exhibited extraordinary sensitivity to
H
2
S with detected concentrations as low as 10 ppb, even in the presence of humidity. The enhanced
sensitivity and improved performance are attributed to the high surface-to-volume ratio of CuO nano-
wires that are entirely surrounded by the gas atmosphere, which is a highly favorable gas sensor
conguration.
A sensor device featuring massive aligned CuO nanowires synthesized by dielectrophoresis and
subsequent thermal oxidation was fabricated for H
2
S detection by Li et al. [197]. The as-fabricated
Fig. 87. (a) Sensing response of a CuO nanowire array sensor to air-diluted H
2
S at 160 C. The device can detect H
2
S at
concentrations higher than 500 ppb. The device conductivity dropped at low concentrations and increased at concentrations
higher than 5 ppm. The initial drops of the response curves are given in the insets. The concentration dependence of the
normalized sensor responses (I/I
0
) is plotted in the inset [384]. Copyright 2008 American Chemical Society.
284 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
CuO nanowire sensor showed good response and repeatability upon H
2
S exposure with a detection
limit of 2.5 ppb and a linear response ranging from 10 ppb to 100 ppb.
Mashock et al. [386] reported a signicant enhancement in CuO nanowire sensing performance at
room temperature through surface functionalization with SnO
2
nanocrystals to form a pn junction. A
very high sensitivity enhancement up to 300% can be realized for detecting 1% NH
3
diluted in air. The
sensor based on SnO
2
NC-functionalized CuO nanowires displaces mostly repeatable sensing signal
and excellent dynamic properties (Fig. 90). The improved sensitivity was due to the fact that the
creation of pn junctions between the host p-type CuO nanowire and the deposited n-type SnO
2
nanocrystals facilitates gas detection through the change in the electrical conductivity of the hybrid
NC-nanowire structure. Moreover, the operating temperature for the SnO
2
NC-CuO nanowire sensing
system was as low as the room temperature compared with the required high temperature (typically
above 200 C) for SnO
2
nanowire-based sensors. This study has potential for engineering the selectiv-
ity of gas sensors at low temperature.
Zhang et al. [133] presented a highly sensitive and selective CuO nanosheet-based H
2
S sensor. This
sensor was reversible, with a rapid response time of 4 s, and a short recovery time of 9 s. The CuO
nanosheets exhibited a detection limit of 2 ppb and a broad linear range from 30 ppb to 1.2 ppm.
Fig. 88. (a) Gas sensing device consisting of eight oxidized Cu lines (thickness = 2.5 lm) and CuO nanowires bridging the gap in
between. Device resistance is measured between the Ti/Au contact pads and is dominated by suspended CuO nanowires. The
inset shows a magnied view of the CuO nanowire bridges. (b) Comparison of the sensor response of a typical CuO nanowire
device toward H
2
S in dry and humid air (relative humidity = 50%) [385]. Copyright 2012 Elsevier.
Fig. 89. (a) A diagram of sensor chip fabrication procedure. (b) A sensor chip before oxidation. (c) Same sensor chip after
oxidation. (d) Surface morphology of CuO nanowires after oxidation [197]. Copyright 2012 Royal Society of Chemistry.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 285
The gas sensor also exhibited high selectivity for H
2
S determination relative to 100-fold concentra-
tions of other gases that may exist in the atmosphere including O
2
, CO, NO, H
2
, and NO
2
. Remarkably,
the CuO nanosheet-based gas sensor showed strong recovery ability and long-term stability. These
promising results facilitated the development of gas sensors based on 2D nanomaterials, such as nano-
plates and nanodiscs. Ramgir et al. [387] reported a similar highly sensitive and selective on CuO thin
lm-based H
2
S sensor. To enhance the H
2
S gas sensing properties of the CuO, Kim et al. [382] fabri-
cated a H
2
S gas sensor from multiple networked Pd-functionalized CuO nanorods. Their results dem-
onstrated that a response of 31,243% at 100 ppmH
2
S at 300 C was achieved for Pd-functionalized CuO
nanorod sensors, whereas with bare-CuO nanorod sensors only a response of 400% was obtained un-
der similar conditions. The recovery and response times of the Pd-functionalized nanorod sensor were
also found to be shorter and longer than that of bare CuO nanorod sensor, respectively. These results
show the potential application in H
2
S gas sensing.
Xu et al. [168] realized a humidity sensor based on honeycomb-like CuO synthesized via a two-step
electrochemical deposition. Their results demonstrated that the honeycomb-like CuO exhibited good
humidity-sensitivity at room temperature (Fig. 91). These characteristics suggest potential application
as an effective and high performance humidity sensor.
Hsueh et al. [388] was developed by humidity sensor based on CuO nanowires grown on a glass
substrate. The fabrication processes are shown in Fig. 92. Humidity sensor based on samples with a
larger initial Cu lm thickness and a longer average CuO nanowire length exhibit a larger sensor re-
sponse (Fig. 93).
Doping CuO nanostructures has been proven to be a promising strategy for enhancing gas sensing
performance of CuO. Kim et al. [389] reported a gas sensor based on Cr-doped CuO nanosheets and
Cr-doped CuO nanorods prepared by heating the slurry containing cuhydroxide/crhydroxide. They
demonstrated that the doped CuO nanostructures with 2.2 at% Cr can enhance the NO
2
response to
24.1 times and signicantly increase the NO
2
selectivity. This enhancement is attributed to the
increase in surface area and pore volume as well as the electronic sensitization caused by the
incorporation of Cr into the CuO lattice and the catalytic function.
Fig. 90. (a) Photo of a silicon chip (4 5 mm
2
) with Au electrodes (boxed areas). (b) SEM image of an SnO
2
NC-functionalized
CuO nanowire bridging a pair of Au ngers. (c) Dynamic responses of devices to 1% NH
3
at room temperature. (d) Eighteen
cycles of 1% NH
3
detection using D#1-nanocrystals and nanowires. Each cycle consists of (sequentially) 2 min of clean air ow,
2 min of 1% NH
3
ow (diluted in air), and 2 min of clean air for sensor recovery [386]. Copyright 2012 American Chemical
Society.
286 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Recently, Yang et al. [86,144] developed quartz crystal microbalance (QCM) sensors for the
detection of trace hydrogen cyanide (HCN) gas based on different CuO nanostructures (ower-,
boat-, ellipsoid-, and plate-like structures). The specic surface area and morphology of these CuO
nanostructures have an important role in the sensitivity of the sensors. The plate-like CuO with the
highest specic surface area of 9.3 m
2
/g has the highest sensitivity of 2.26 Hz/lg. The response curves
Fig. 91. (a) Relative humidity vs. dc resistance plots at room temperature of CuO honeycombs. The insets show schematic
illustration (upperright) and the corresponding IV curves (lowerleft) of a CuO honeycomb-based humidity sensor. (b)
Resistance variations with time for the CuO sample [168]. Copyright 2009 Institute of Physics.
Fig. 92. Schematic diagram of the fabricated CuO nanowire humidity sensor [388]. Copyright 2011 Elsevier.
Fig. 93. Dynamic responses measured from the three fabricated humidity sensors with 5 V applied bias at 25 C. The
thicknesses of the Cu lm for samples A, B, and C are 0.5, 1, and 2 lm, respectively [388]. Copyright 2011 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 287
shown in Fig. 94 are similar for three continuous cycles with nearly no changes in response, response
time, and recovery time, which indicates very good reproducibility and stability.
Yang et al. [390] developed an excellent selectivity and sensitivity of hydrogen cyanide detection
using CuO nanoparticle-functionalized QCM resonator. This method provides insights on CuO as a sen-
sor material in public security and environmental applications.
In addition to the applications in gas sensors of CuO nanostructures, Akhavan and Ghaderi [380]
demonstrated the application of CuO nanoakes in high sensitive sensing of bacteria, where the
CuO nanoakes were synthesized on a Cu foil through surface oxidation. Zaman et al. [126] also re-
ported the application in pH sensor of well-crystallized ower-shaped CuO nanostructures composed
of thin leaves synthesized by a simple low-temperature chemical bath method.
5.3.2. Application in enzyme-free glucose electrochemical sensors
Glucose detection is important in several areas not only in clinical diagnostics of blood glucose
sensing, but also in environmental monitoring and food industry [39,391,392,394]. To avoid diabetic
emergencies, frequent testing of physiological blood glucose levels is important to conrm the effec-
tivity of treatment [39,392,394]. Thus, considerable efforts have been invested in highly sensitive, reli-
able, selective, and low-cost approaches for the precise monitoring of the glucose level. Among the
numerous detection methods developed to measure glucose concentration [39,391394], the electro-
chemical detection technique is one of the most promising technology. This method is not only simple,
sensitive, and selective, but also has a lower detection limit, faster response time, and better long-term
stability and is inexpensive [39,392394]. A number of electrochemical approaches of glucose detec-
tion are based on the use of glucose oxidase (GOD). The GOD-based glucose sensors catalyze the oxi-
dation of glucose to glucolactone and simultaneously produce H
2
O
2
. Glucose level is determined by
estimation from electrochemical response to the liberated H
2
O
2
[39,391,392]. However, certain limi-
tations of GOD-based glucose sensors still exist, such as high cost and lack of enzyme stability, com-
plicated immobilization procedures of enzyme, and the relatively high over potential for the H
2
O
2
to
introduce coexisting interferences in the biological uids together with critical operating conditions
[39].
To date, studies have focused on the development of enzyme-free electrochemical sensors based on
nanostructured materials (Fig. 95) with high sensitivity, fast response, and long-term stability
[39,393]. In particular, electrochemically active nanomaterials consisting of metals, metallic CNTs,
transition MOs, or conducting polymers have been extensively examined for non-enzymatic electro-
chemical sensors because of their large surface-to-volume ratio and increased catalytic activities
[39,391,392]. Among these nanomaterials, CuO nanostructures are promising candidates for active
Fig. 94. Proles of the frequency shift of a plate-like CuO-functionalized QCM resonator upon exposure to air containing
50 ppm HCN at 25 C in three continuous cycles [144]. Copyright 2011 American Chemical Society.
288 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
electrode materials of non-enzymatic electrochemical sensors because of their high sensitivity, low
detection limit, less interference, and low-cost fabrication for sensors [39].
In Ref. [395], CuO nanoowers and 1D nanorods were synthesized for the rst time using compos-
ite-hydroxide-mediated and composite-molten-salt methods, respectively. Both nanostructures were
applied to modify the graphite substrates for non-enzymatic glucose detection. The authors observed
that CuO nanostructure modied-graphite electrodes display greatly improved performances and in-
crease the electrocatalytic ability toward glucose oxidation compared with single graphite electrode,
which may be attributed to their large surface area, high surface energy, and enhanced electron trans-
fer ability. The CuO nanostructure-modied electrodes also have good selectivity of glucose in the
presence of dopamine and ascorbic acid.
2D CuO NLs prepared by a simple room temperature method were used in Naon/GCE to construct
an enzyme-free glucose sensor for detecting the electrocatalytic activity toward the oxidation of glu-
cose in an alkaline solution [149]. The peak current of Naon/CuO-NLs/GCE is about 15 times that at
Naon/commercial-CuO/GCE (Fig. 96). This result reveals good electrocatalytic activity of the CuO-NL
involved electrode. Well-dened, stable, and fast amperometric response observed in Fig. 96b indi-
cates that Naon/CuO-NLs/GCE sensor shows good sensitivity for glucose detection. The tested results
in Fig. 96c show that no obvious current change occurs when AA and UA are added to the system
(points 6 and 7 on the curve), but the addition of glucose leads to an abrupt current change (point
5). These results suggest that the sensor containing CuO-NLs exhibits good selectivity to glucose
and shows good interference-free ability.
Zhuang et al. [396] synthesized CuO nanowires by dehydration of Cu(OH)
2
nanowires on a Cu rod
for non-enzymatic glucose sensing. The glucose sensor has fast response (<1 s) as well as good sensi-
tivity and selectivity to glucose. To improve the performance of enzyme-free glucose electrochemical
sensors, chemically modied CuO nanostructures with carbon-based materials are investigated as
attractive materials for sensor electrodes. Yang et al. [397] reported MWCNT-coated CuO NL compos-
ites prepared by a facile chemical precipitation method using the special structure and extraordinary
mechanical and unique electronic properties of the MWCNTs. The sensors based on these composites
exhibited high sensitivity of 664.3 lA mM
1
cm
2
and excellent detection limit of 5.7 lM. This excel-
lent performance as a glucose sensor was due to the unique structure and large surface area of the
prepared electrode, which is composed of CuO NLs and MWCNTs.
Luo et al. [398] developed a non-enzymatic amperometric glucose sensor based on CuOnanocubes
graphene nanocomposite-modied glassy carbon electrode (CuOGGCE) by a simple electrochemical
Fig. 95. Nanomaterials can be incorporated into these sensors to increase surface area, improve catalytic action, modify
operating parameters, and improve electron transfer from the enzyme to the electrode. This incorporation can be accomplished
through the use of single types of nanomaterials, such as (a) CNTs or (b) nanocomposites, which consist of multiple
nanomaterials working together [392]. Copyright 2010 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 289
deposition method. Graphene can increase the surface area and promote the electron transfer
reactions of glucose oxidation because of its unique nanostructure and extraordinary properties
(e.g., excellent electronic transport property and high electrocatalytic activity). Compared with the
unmodied graphene or CuO electrode, the CuO-modied graphene electrode displays substantially
higher electrocatalytic activity and faster response to glucose oxidation with a higher current re-
sponse. The CuO-modied graphene electrode also shows long-term stability, good reproducibility,
and excellent specicity to glucose in the presence of common interferents.
Notably, various nanostructured CuO and their composites have been extensively explored as po-
tential building blocks for application in enzyme-free glucose electrochemical sensors during the last
two decades. Other novel CuO nanostructures and their composites applicable to glucose sensors are
summarized in Table 9.
5.4. Application in solar cells
CuO, as a p-type semiconductor with the bandgap energy in the 1.22.1 eV range [61,279], has
been broadly examined for photovoltaic applications because of its low cost, high solar absorbance,
low thermal emittance, non-toxicity, and simple manufacturing process. Moreover, CuO is very prom-
ising for use in solar photovoltaics because of its excellent stability, good electrical properties, and
high carrier concentration of CuO [33,304].
Anandan et al. rst demonstrated vertically aligned CuO nanorod arrays grown on a Cu electrode
[123] for application as a cathode in dye-sensitized solar cells (DSSCs) based on n-TiO
2
nanoparticles.
The as-prepared CuO nanorod arrays that were synthesized at different preparation conditions serving
as a cathode in a solar cell were subjected to IV characteristic measurement in the dark and under
illumination (15 mW/cm
2
). An open circuit voltage (V
oc
) ranging from 230 mV to 564 mV, a short-cir-
cuit current (I
sc
) ranging from 0.332 mA to 0.521 mA, a ll factor (FF) ranging from 0.16 to 0.23, and a
power conversion efciency ranging from 0.12% to 0.29% were obtained. For comparison, the power
conversion efciency of regular solar cells using a Pt cathode (ITO/TiO
2
/dye/electrolyte/Pt/ITO) under
Fig. 96. (a) CVs of a Naon/commercial-CuO/GCE before (1) and after (2) 2.0 mM glucose addition; CVs of a Naon/CuO-NLs/
GCE before (3) and after (4) 2.0 mM glucose addition. (b) Amperometric current vs. time for the detection of glucose using the
Naon/CuO-NLs/GCE. (c) Amperometric responses to the injection of 1.0 mM glucose (5), 0.10 mM AA (6), and 0.1 mM UA (7) at
the Naon/CuO-NLs/GCE [149]. Copyright 2011 Institute of Physics.
290 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
the same illumination conditions was 1.23%, which is not much higher than that of the CuO nanorods/
Cu cathode. This result indicates thatthe preparation of the CuO nanorods/Cu cathode should still be
improved by reducing the resistance and surface states of the CuO nanorod lms. Liu et al. [412] re-
ported better results for CuO nanoneedle array prepared on a pure Cu NC line layer by thermal oxida-
tion in the air for application as a cathode in TiO
2
-basedDSSCs. A full sun efciency of 1.12% with an FF
of 0.37 was obtained. This efciency is much higher than that (0.120.29%) in Anandans study [123]
and even higher than the 0.8% efciency with Cu
2
O lm as the cathode. The authors attributed the en-
hanced efciency to the higher surface area of CuO nanoneedle array, which can provide an enhanced
electron diffusion length and dye adsorption.
In addition to its use as a cathode, CuO can also act as p-type metal-oxide layer with n-type ZnO
layer to form a barrier layer to control charge recombination dynamics and improve the perfor-
mance of ZnO DSSCs. Raksa et al. [413] investigated the effects of CuO layer as a barrier layer on
power conversion characteristics of ZnO DSSCs with different photoelectrodes. CuO powder,
nanowire, and thin lm were used as a layer on the top of ZnO layer to form a blocking layer. A
schematic of the structure of a fabricated solar cell is depicted in Fig. 97. The authors found that
ZnO DSSCs with CuO thin lm exhibit higher current density and higher power conversion efciency
than those without CuO thin lm. These characteristics indicate that CuO thin lm can prevent elec-
trons from backing or reverse transferring to dye and electrolyte resulting in low back or reverse
current. Consequently, less energy-wasting interfacial recombination and effective electron injection
process to ZnO could be achieved. However, DSSCs with CuO powder and nanowire exhibit lower
efciency than those of CuO thin lm and comparable with DSSCs without the CuO layer. This nd-
ing is attributed to the thicker layer of CuO powder and nanowire resulting in a retardation of the
interfacial recombination dynamics of CuO blocking layer, which is comparable to the dye excited-
state decay. Sahay et al. [414] reported a similar study using CuO nanobers as a barrier layer for
ZnO-based DSSCs. A 25% increase in the current density was observed with the use of CuO as a
blocking layer.
Table 9
CuO nanostructures and their composites applicable to glucose sensors.
Electrode matrix Detection
techniques
Linear range Sensitivity/detection limit
(lM)
Response time
(s)/applied
potential (V)
Ref.
CuO micro-/
nanostructures
Amperometric 0.916 mM 9.02 lA mM
1
/20 /0.6 [399]
CuO nanobers Amperometric 6 lM to 2.5 mM 431.3 lA mM
1
cm
2
/80 1/0.4 [400]
CuO nanowires-
naon
Amperometric 0.5488 lM 64.1 lM mM
1
/4.5 10
2
<2/0.6 [401]
0.9885.488 mM
Mesoporous CuO Amperometric 170310 lM 0.04 lA Mm
1
/ <5/0.38 [402]
CuO nanoparticles Amperometric 5 lM to 2.3 mM 1397 lA mM
1
cm
2
/0.5 <2/0.55 [403]
porous CuO Amperometric 1 lM to 2.5 mM 2.9 lA cm
2
mM
1
/0.14 3/0.65 [404]
CuCuO nanowire
(NW) composite
Amperometric 0.112 mM /50 5/0.3 [405]
CuO microbers
composed of CuO
nanoparticles
Amperometric 0.2 lM to 0.6 mM 2321 lA mM
1
cm
2
/
2.2 10
3
/0.4 [406]
CuO nanobers-ITO Amperometric 0.2 lM to 1.3 mM 873 lA mM
1
cm
2
/
4 10
2
<1/0.48 [407]
CuO nanoparticles-
MWCNTs
Amperometric 0.4 lM to 1.2 mM 2596 lA mM
1
cm
2
/0.2 <1/0.4 [408]
CuO nanospheres Amperometric 50 lM to 2.55 mM 404.53 lA cm
2
mM
1
/1 /0.6 [409]
CuO nanowalls-Cu Amperometric 0.0510 lM 0.5563 lA lM
1
10/0.1 [410]
CuO nanosheets Amperometric 5.00 100
1
to
1.00 10 mM
5.20 10
2
lA cm
2
mM
1
/ /0.5 [411]
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 291
Given its simple production method and relative good carrier mobility, CuO is used to act as a
p-type semiconductor active layer combined with an organic electron acceptor, such as C
60
and PCBM,
to form an inorganic and organic hybrid solar cell. A full sun power conversion efciency of 0.04% of a
bi-layer CuO and PCBM hybrid solar cell was achieved [415], which is four times higher than that of
the PCBM-only cell. This result indicates the potential of these CuO nanocrystals for light-harvesting
applications. However, the conversion efciency of hybrid solar cells based on CuO with C
60
is very
low (1.8 10
6
%) compared with that of CuO-based DSSCs (1.12%) [142]. Further study to synthesize
higher quality CuO thin lms may improve the CuO cell efciency [416]. Wang et al. [417] also
achieved nanowire-based all-oxide pn heterojunction solar cells consisting of vertically oriented
p-type CuO nanowires that were surrounded by a lm constructed from n-type ZnO nanoparticles.
The growth of p-CuO nanowire arrays with a simple thermal oxidation method and then the fabrica-
tion of nanowire-based heterojunctions by coating the p-CuO nanowire arrays in an n-ZnO layer
through a thermal decomposition method are shown in Fig. 98. Their optoelectronic properties and
photovoltaic performance were investigated in detail. A 154% increase in photoconductance was
obtained under white-light illumination of 100 mW/cm
2
compared with the conductance in the dark.
The time-related photocurrent of the fabricated device to the periodic light irradiation of 141 mW/cm
2
at a bias voltage of 0.5 V is shown in Fig. 99a. The heterojunctions exhibit a photo-generated current of
0.264 mA after three cycles. Under the white-light illumination of 100 mW/cm
2
, the heterojunctions
show a larger open-circuit voltage of 0.37 V and anFF of 36.9% after annealing at 100 C for 25 min
(Fig. 99b). The overall power conversion efciency of the solar cell is 0.1%, which is twice than that
of all Cu
2
O-based oxide solar cell [418]. These results indicate that the heterojunctions could be useful
for durable and non-toxic photovoltaic components. However, they also introduce other complicating
factors, which must be addressed separately to fully realize effective CuO nanowire-based photovol-
taic devices.
In summary, CuO nanostructures can not only serve as a p-type hole conductor but also act as a
barrier layer in DSSCs along with a p-type active layer in hybrid solar cells. Further study of synthe-
sizing high surface area, good quality, and well-dened CuO nanostructures would probably enhance
the conversion efciency of CuO-based solar cells.
5.5. Application in photodetectors
Photoconductivity is a well-known property of semiconductors in which the electrical conductivity
changes (usually increases) because of incident radiation [419]. These changes involve several succes-
sive or simultaneous mechanisms including absorption of the incident light, carrier photogeneration,
Fig. 97. Schematic diagram of DSSC structures with different photoelectrodes for ZnO/CuO layer [413]. Copyright 2009 Elsevier.
292 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
and carrier transport (including carrier trapping, detrapping, and recombination) [419]. A mechanism
using a ZnO nanowire as the model system has been comprehensively outlined [419422] (Fig. 100).
With large surface-to-volume ratios and Debye length comparable to their small sizes, MO nanom-
aterials have already displayed superior sensitivity to light in experimental devices. Thus, the fabrica-
tion and characterization of novel nanostructures for applications in photodetectors based on different
MO nanomaterials has been extensively explored [422]. CuO is intrinsically a p-type semiconductor
with a low band gap (1.22.0 eV), indicating its potential in photodetection and optical switching
applications in the visible range in which other MOs with their larger band gaps fail to perform [293].
Hansen et al. [293] reported the fabrication of large-scale high-density CuO nanowire arrays via di-
rect oxidation of Cu. These nanowires with diameters of 200 nm and length of several micrometers
were rst transferred to polydimethylsi-loxane (PDMS) by gently pressing a thin piece of PDMS on top
Fig. 98. Schematic diagram of the fabrication of ZnO-coated CuO nanowire arrays [417]. Copyright 2011 Optical Society of
America.
Fig. 99. (a) Time-related photocurrent of the ZNO-coated CuO nanowire arrays layer to the periodic light irradiation
(J
light
= 141 mW/cm
2
, t
On/Off
= 30 s) at a bias voltage of 0.5 V. (b) Current densityvoltage curve (before and after annealing) in
the standard AM 1.5 global solar spectrum illumination condition for the photovoltaic performance [417]. Copyright 2011
Optical Society of America.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 293
of the nanowires and then removed. Next, this PDMS with high-density nanowires on the surface was
placed in a small amount of ethanol, and the nanowires were ultrasonically removed from the PDMS.
Finally, the nanowires dispersed in ethanol were deposited onto the electrode grids to form a single
nanowire photodetector device.
The IV characteristics of the single CuO nanowire photodetector device, measured in dark room
and under white illumination with different light intensities are shown in Fig. 101a. The conductance
exhibits a relative increase of 2587% for light intensities from1.62 mW/cm
2
to 45 mW/cm
2
, respec-
tively. The photocurrent is found to be power law-dependent on the light intensity, namely Ip = P
0.42
,
where P stands for the light power density as illustrated in Fig. 101b. This result ts well with the
theory of photoresponse that gives power law dependence with an exponent of 0.5 for the high-
injection case. Moreover, the device responsivity to white light approaches 8 and 1 A/W for power
densities of 1 and 45 mW/cm
2
, respectively. The strongly enhanced responsivity of the CuO nano-
wire-based optical sensors is likely to stem from combined effects of photocarrier multiplication
inside the nanowire as well as carrier injection from the contacts and surface states for low incident
intensities.
Time-resolved on/off measurements of the photoresponse to white light were also conducted, and
the results are shown in Fig. 102a. A relatively slow recovery time of 36 s was observed, indicating
Fig. 100. Photoconduction in a nanowire photodetector: (a) schematic of a nanowire photodetector. Trapping (b) and
photoconduction (c) mechanism in ZnO nanowires [419]. Copyright 2010 John Wiley & Sons, Inc.
Fig. 101. (a) IV curves recorded for the CuO nanowire device in dark and various light intensities. (b) Photocurrent vs. incident
light intensity with power law t. (Inset) Estimate of the photoconductive gain vs. estimated photon ux of the irradiated
nanowire [293]. Copyright 2010 American Chemical Society.
294 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
slowmolecular desorption of the photo induced and chemically absorbed oxygen molecules. However,
a time constant of s = 240 ms is obtained from the inset of Fig. 102b, suggesting that this CuO nano-
wire-based device holds promise for fast photo sensing applications.
Remarkably, an IR photodetector based on CuO nanowires has been successfully fabricated and
investigated by Wang et al. [41]. The CuO nanowires with an average length of 1.2 lm and an average
diameter of 50 nm were synthesized via direct heating of a Cu wire in the air. A schematic diagram of
the fabricated photodetector is shown in Fig. 103a. Time-resolved measurements of the photore-
sponse to an 808 nm light with different light intensities were conducted, and the results are shown
in Fig. 103b. The current ofthe device remarkably increases along with exposure to illumination under
an IR lamp. When the light was turned off, the current suddenly reverts to its original value. The rise-
and fall-times of the fabricated CuO IR photodetector were 15 and 17 s, respectively. Furthermore,
measured photocurrent increased with increasing excitation power. A plot of photo-generated current
Ip (Ip = I
ill
I
dark
), where I
dark
is the current measured in dark, and I
ill
is the current measured under
illumination) vs. light intensity (P) indicated that Ip displayed a power law dependence on the light
intensity, namely, Ip P
1.1
. This condition suggests that the generation efciency of charged carriers
was approximately proportional to the number of photons illuminated onto the sample. The authors
also investigated the response of the on/off ratio with or without illumination in oxygen ambient and
vacuum. Dark and photocurrents measured in oxygen ambient were both smaller than those mea-
sured in vacuum. This behavior was ascribed to the desorptionadsorption of O
2
from the surface
of the CuO nanowire that was induced by IR light and has a remarkable effect on the photoresponse
of CuO nanowire-based devices.
Fig. 102. Time-resolved photoresponse of CuO nanowire device (biased at 5 V) (a) and rst leg of the time-resolved
photoresponse of CuO nanowire device (biased at 2 V) (b) with curve t of the rst regime (inset) [293]. Copyright 2010
American Chemical Society.
Fig. 103. (a) Schematic diagram of the fabricated photodetector. (b) Dynamic photoresponse measured in vacuum [41].
Copyright 2011 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 295
Manna et al. [423] reported a visible light photodetector made of mesoporous CuO dandelion struc-
tures synthesized by a simple template-free hydrothermal route. For comparison, a visible light pho-
todetector made of CuO NC prepared by a chemical precipitation method was also fabricated. A
schematic diagram for the photocurrent measurements and IV curves are illustrated in Fig. 104a. Oh-
mic contact between Au and CuO was observed, indicating that the as-prepared CuO exhibited p-type
semiconducting behavior. Photo generated carriers could signicantly increase the conductivity when
semiconductor materials are illuminated by high-energy photons. A larger surface area (BET surface
area = 325 m
2
/g) of the mesoporous CuO dandelion structures than that of CuO NC (BET surface
area = 5.5 m
2
/g) can further enhance the sensitivity to white light and might even lead to the real-
ization of single photon detection. On white-light illumination, the photocurrent in the CuO dandelion
structures reached a maximum value of 238 10
6
A, leading to the photo (I
ph
)-to- dark current (I
d
)
ratio I
ph
/I
d
of 397, which was remarkably higher compared to the CuO nanocrystals (I
ph
/I
d
= 92). When
the white light was switched off, the photocurrent dropped very fast as shown in Fig. 104b. The
authors suggested that the notable photoresponse of the CuO dandelion structures might be attributed
Fig. 104. (a) Dark IV and (b) growth and decay curves of the photocurrent under and upon removal of 15 min white-light
illumination at 3 V bias voltage of the CuO dandelion structures and CuO nanocrystals; the inset of (a) is the schematic diagram
of the photocurrent measurement device. (c) Experimental and tted curves of the fast response region of photocurrent growth
of the CuO dandelion structures. (d) Time-resolved measurement of photoresponse to white light of the CuO dandelion
structures under the period of 3 min at 3 V bias voltage [423]. Copyright 2010 American Chemical Society.
296 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
to the O
2
adsorption on the surface of CuO dandelion structures. Upon illumination, the photo-
generated and free holes of the p-type CuO discharge surface chemisorb O
2
through the surface
electronhole recombination [h + O
2
(adsorbed) ?O
2
(gas)], while the excess holes abruptly increase
the conductivity of the CuO dandelions. When illumination is switched off, the electrons recombine
with the holes and are also captured by the re-adsorbed O
2
molecules increasing the depletion width,
which hastens the current decay. Some current remains after the sudden fall in the current (Fig. 104b),
which may have resulted from the presence of the excess photogenerated holes in p-type CuO
dandelions. Time-resolved measurement of photoresponse to white light was conducted, and the
result is shown in Fig. 104d. The on and off current for each of the three cycles remain the same
within the noise envelope, indicating the reversibility and stability of the CuO dandelion structures
optical switches. The photoconductivity response time is 3 min. These results demonstrate that CuO
dandelion structures are promising materials for optoelectronic devices such as optical switches,
photo detectors, and solar cells.
More interestingly, IR and visible wavelength-dependent photovoltaic effects of the devices fabri-
cated by coating double-walled CNT (DWCNT) lms on CuO nanowire arrays to form hetero-dimen-
sional contacts by irradiating with 405, 532, and 1064 nm lasers were observed by Xu et al. [424].
Photo-induced electrons in the formed hetero-dimensional contacts were proved to be transported
from the lower dimensional materials to the neighboring higher-dimensional materials. This behavior
is known as the hetero-dimensionality (HD) effect. When the devices were irradiated by light with
energies above the bandgaps of CuO as shown in Fig. 105b and c, a negative photocurrent was ob-
served, which indicates p-type semiconducting nature of the CuO nanowire. Given that the photoex-
citations occurred within the CuO nanowire, the PV effects of the device under these conditions were
attributed to well-known heterojunction (HJ) effects. For illumination by the photo energy lower than
the bandgap of CuO (Fig. 105), a reverse photocurrent direction relative to the 405 and 532 nm cases
was observed.
Fig. 105. DWCNT lm/CuO nanowire array hetero-dimensional contacts. (a) Schematic diagram. (b), (c), and (d) are
photoresponses under illumination with 405, 532, and 1064 nm lasers [424]. Copyright 2012 Institute of Physics.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 297
The excitation of non-equilibrium carriers within DWCNTs can occur together with the heat in-
duced by the radiation of the IR laser, which is the possible excitation source for the DWCNTs. There-
fore, the photo- or thermal-induced electrons illuminated by a 1064 nm laser can be transported from
the CNT lms to the CuO nanowire arrays because of the excitation within DWCNTs and HD effect,
resulting in reverse photocurrent directions relative to the 405 and 532 nm cases. These phenomena
were further conrmed by control experiments conducted on CNT lm/CuO granular lm hetero-
dimensional contacts. These interesting results shed new light on the use of hetero-dimensional con-
tacts for wavelength-dependent photosensors.
5.6. Application in catalysis
CuO is one of the most prominent catalysts and is extensively used in environmental catalysis. CuO
nanostructures typically have higher catalytic activity than their bulk or micro counterparts caused by
their large surface area [52]. CuO nanostructures are potential catalysts for oxidation reactions of CO
and for substituting noble metal catalysts because of their high catalytic activity, nontoxicity,
low-cost, and availability [53]. The performance of nanocrystals is usually strongly related to the
surface structure of facets enclosing the crystals together with the specic surface area [19]. CuO
nanostructures with different shapes, including irregular nanoparticles, nanobelts, and nanoplatelets,
were synthesized by controlling several critical synthesis parameters to explore their catalytic prop-
erties [52]. The catalytic performances of these CuO samples for CO oxidation are shown in Fig. 106.
The observed specic reaction rates on the nanoplatelets exposing the (01) planes were over six times
higher than that on the nanoparticles with the close-packed (111) planes and roughly thrice as much
as that on nanobelts with (001) planes. These results demonstrate the important role of the shape and
the exposed crystal planes of the CuO during the CO catalytic process.
Huang et al. [425] developed a controllable synthesis of various morphologies of CuO nanostruc-
tures by hetero-metal cations in aqueous solution at room temperature. They observed that the
as-synthesized mesoporous CuO quasi-monocrystalline nanosheets with a uniform exposed crystal
plane (002) demonstrated superior catalytic activity for CO oxidation with high CO conversion ef-
ciency (47:77 mmol
CO
g
1
CuO
h
1
at 200 C), further suggesting that the catalytic activity of CO oxidation
depends strongly on the surface structure of facets. The exposed crystal plane (002) terminates with a
Cu atomic layer, which results in insufcient oxygen atoms coordinating with Cu atoms on the surface.
Moreover, this exposed plane has more dangling bonds that could absorb more CO gas.
Qiu et al. [195] investigated the catalytic activity for CO oxidation of CuO with different shapes.
They observed a fair correlation between specic surface area and performance, indicating that the
catalytic activity for complete oxidation is strongly depended on specic surface area of the obtained
Fig. 106. Specic rate of CO conversion over CuO nanoparticles, nanobelts, and nanoplatelets at 110 C, the predominantly
exposed crystal planes of each nanostructure are also indexed [52]. Copyright 2006 Institute of Physics.
298 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
samples. Spheroidal CuO exhibited higher catalytic activities and capacitance values than those of CuO
nanorods because the former has a larger specic surface area.
1D nanostructured materials have improved catalytic activity because of their large surface-to-vol-
ume ratios. They can provide more grain boundaries and expose active crystal planes [426]. The load-
ing of 1D nanostructured materials can be simply controlled without aggregation. Previous studies
have shown that the nature of the MO surface plays a crucial role in their catalytic activity [19,53].
Feng and Zheng [426] reported the rst experimental study of catalytic CO oxidation over 1D CuO
nanowires grown directly on Cu meshes by a thermal oxidation method. The CO oxidation over the
as-grown CuO nanowires, Ar and H
2
plasma-treated CuO nanowires, and the mechanism responsible
for the plasma effect were systematically investigated. The O oxidation percentage increased from 4%
over the as-grown CuO nanowires to 29% and 85% over the Ar and H
2
RF plasma-treated nanowires,
respectively, when the tests were performed at 140 C under the fuel-lean condition. The similar level
of improvement for the fuel-rich condition was also observed. According to the combined character-
izations, the mechanism responsible for the plasma enhancement effect was revealed by the authors.
The enhancement effect is mainly due to the generation of grain boundaries and the reduction of Cu
(II) to the more active oxidation state of Cu (I).
Zhong et al. [427] demonstrated that porous CuO synthesized by the hydrothermal method exhib-
ited good activity for CO oxidation, and its temperature for the 100% CO conversion was 145 C. This
good catalytic activity was attributed to the reversible oxygen storage and release of CuO via the
chemical reaction of CuO to Cu
2
O. Moreover, the catalytic activity of CuO can be further improved after
integration with Au particles and 100% conversion of CO was achieved at 100 C. Zhou et al. [428]
indicated that the catalytic activity for CO oxidation was affected by the structure and morphology
of CuO. They also observed that the specic reaction rate for CO oxidation on the porous microstruc-
tures was over seven times higher than that on the microowers and over four times higher than that
on microspheres (Fig. 107). The higher catalytic activity of the nanoporous CuO was ascribed to its
large surface area and the pores with uniform sizes, which favor more effective greater exposure of
the surfaces to the gas molecules.
Zhang et al. [429] studied the oxygen adsorption capacity and catalytic performance of CO oxida-
tion of CuO and observed that the adsorbed oxygen molecules played crucial roles in CO catalytic oxi-
dation on CuO catalysts. The oxygen adsorption behavior was detected by XPS and O
2
-TPD
experiments as shown in Fig. 108. The CuO with higher oxygen adsorption capacity exhibited much
higher activity than that of the sample with low oxygen adsorption capacity at low temperature, sug-
gesting that the catalytic activity for complete oxidation is directly related to the amount of adsorbed
oxygen.
Fig. 107. Specic rates of CO conversion compared with CuO microowers, microspheres, and porous microstructures at 110 C
[428]. Copyright 2010 Wiley.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 299
Similar oxygen adsorption behavior was further conrmed by Zhong et al. [430]. They prepared
poorlycrystallized CuO nanorods directly on Cu foil via electrochemical corrosion and oriented-attach-
ment in a KOH liquid membrane. The authors found that the CuO nanorods exhibit strong oxygen
adsorption capacity and high oxidation states (close to 3+), especially at high temperatures and in a
high O
2
concentration environment. The catalytic activity of the CuO nanorods for CO oxidation
occurred on the adsorbed oxygen rather than on the lattice oxygen and the Cu oxidation states did
not drop to 1+ during catalytic reactions. Moreover, the rate-controlled step of CO oxidation was
the surface oxidizability of the CuO nanorods, which was enhanced by high temperature and an
oxidizing environment, such as an O
2
-rich atmosphere with higher oxygen adsorption capacity.
Pillai and Deevi [431] reported room temperature oxidation of CO to CO
2
over unsupported CuO
catalyst prepared by a controlled heating of precipitated Cu(OH)
2
after activation of the catalyst in
an oxygenCO atmosphere. The catalyst efciency for CO oxidation was 117 mL h
1
gcat
1
at 50 C.
They demonstrated that drying and calcination of the precipitated hydroxide are crucial parameters
for obtaining an active catalyst. The active phase was found to be a metastable non-stoichiometric
form of CuO formed during the treatment of the oxide in a reducingoxidizing environment. Yu
et al. [432] obtained a unique 3D walnut-like CuO nanostructure assembled from many 8 nm grains.
The CuO nanostructure was synthesized through dehydration of Cu(OH)
2
nanostrands in aqueous
solution at room temperature. The authors noted that the walnut-like CuO nanostructure demon-
strated high surface area of 61.24 m
2
g
1
and also high catalytic activity of 160 mL h
1
gcat
1
for
CO oxidation at 50 C. The catalytic activity is 23 times higher than that of 40 nm commercial CuO
nanoparticles, which is the highest value reported to date for CO catalytic oxidation on CuO catalysts
without any support at 50 C. Such behavior could be ascribed to the ne grain size of the CuO nano-
walnuts contributing to the high surface area and high catalytic activity.
In addition, CuO on various supports have also been extensively investigated and are known to be
highly active for CO oxidation [433436]. For instance, Derekaya et al. [433] examined the inuence of
the preparation, composition, and pretreatment conditions of catalysts on the activity/selectivity of
the selective CO oxidation reaction over ceria-supported CuO catalysts. Their results showed that a
catalyst-support composition containing 5% CuO exhibit the highest and most stable CO conversion
and selectivity. Ratnasamy et al. [434] demonstrated that the CO oxidation activity/selectivity of
CuO on different supports increases in the order CuOZrO
2
< CuOCeO
2
ZrO
2
6 CuOCeO
2
, and a cat-
alyst-support composition containing 5% CuO exhibited stable activity in long-term experiments.
Interestingly, Horns et al. [437] reported an inverse conguration in which large-sized CuO particles
act as the support for ceria instead of traditional direct catalyst congurations, where CuO is dispersed
Fig. 108. (a) O 1s XPS spectra. (b) O
2
-TPD proles of the synthesized CuO samples and the commercial CuO sample [429].
Copyright 2009 Royal Society of Chemistry.
300 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
onto ceria. A wider full CO conversion window and higher CO
2
selectivity was achieved over this in-
verse catalyst (Fig. 109). Such behavior was ascribed to the limited reducibility of large-sized CuO par-
ticles in this catalyst.
In addition to their function as catalysts for CO oxidation, CuO nanostructures can also be em-
ployed as low-cost catalysts for organic synthesis, such as CN cross-coupling and reactions of benz-
imidazole with various arylhalides [438,439], oxidation of cyclohexene to 2-cyclohexene-1-one [440],
the arylation of the CH bond of heterocycles to form the corresponding heterocycle compounds [441]
and the synthesis of propargylamine [83] and other organic compounds [195,442445].
5.7. Application in photocatalysis
Organic dyes, which are widely used in various industrial processes, form an integral part of indus-
trial wastewater. Dye removal and degradation have been a subject of concern worldwide because of
the potential toxicity and visibility of dye in surface water [446]. Numerous processes and methodol-
ogies have been developed for dye removal from wastewater [446,447,450]. The utilization of solar UV
and visible light for the degradation of various organic pollutants using semiconductor catalysts (usu-
ally MOs) Fig. 110) has drawn particular attention because it has been successfully employed to treat a
wide range of toxic and nondegradable organic pollutants into readily biodegradable compounds in
the wastewater without involving complex technologies [446450].
A variety of semiconductor catalysts, such as TiO
2
[448,450], ZnO [447,449], Fe
2
O
3
[448], and CuO
[47], are used as photocatalysts to degrade organic pollutants. CuO is a promising candidate because of
Fig. 109. Catalytic activity under 1% CO, 1.25% O
2
, and 50% H
2
(Ar balance) for the indicated catalysts. Top: CO conversion.
Bottom: selectivity to CO
2
[437]. Copyright 2010 American Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 301
its relatively low-cost. To enhance the photocatalytic ability of CuO for degradation of organic pollu-
tants, H
2
O
2
is often added because its electronic structure, characterized by a lled VB and an empty
CB, plays an important role in the photocatalytic process [449,451]. Miyauchi et al. [452] proved that,
in the absence of H
2
O
2
, CuO is not an effective photocatalyst for degrading organic pollutants because
of its inability to produce a good amount of OH

radicals, which have high oxidative ability and cause


destructive oxidation of the organic dye. This result may be attributed to a fact that the VBs of CuO
(Fig. 111) are more negative than the redox potential necessary for the generation of OH

radicals.
Thus, under illumination, CuO cannot create OH

radicals and has a smaller amount of oxidative activ-


ities for degradation of organic pollutants.
Yu et al. [453] also found that the as-prepared CuO, Cu
2
O, and CuO/Cu
2
O composite hollow spheres
show no photocatalytic activity under visible-light irradiation, unless H
2
O
2
was added as a source of
OH

radicals. Yang and He [85] developed a simple hydrothermal method to synthesize a series of CuO
nanostructures (ower-, boat-, plate-, and ellipsoid-like) for the degradation of methylene blue (MB).
Their results also showed that nearly no MB degradation was observed after 15 h light irradiation in
the absence of H
2
O
2
as shown in Fig. 112. The photocatalytic activities of plate-like, ower-like, and
boat-like CuO nanostructures are higher than those of ellipsoid-like CuO nanostructure and commer-
cial CuO in the presence of H
2
O
2
because of their relatively higher specic surface areas (see Fig. 113).
Fig. 110. Schematic of the principle of photocatalysis [449,450]. Copyright 2010 Elsevier.
Fig. 111. Diagram depicting the redox potentials of VB and CB as well as the bandgap energies for various MOs estimated at
pH 7. The redox potential positions of H
+
/H
2
and OH

/OH

at pH 7 are also illustrated [452]. Copyright 2002 American Chemical


Society.
302 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Sun et al. [142] further conrmed the morphology-dependent photocatalytic ability of CuO. The
authors observed that 3D mesostructural CuO spindles exhibit a higher photocatalytic degradation
of RhB than those of 2D mesostructural CuO plates in the presence of H
2
O
2
under the same measure-
ment condition. After 1 h visible light irradiation, the degradation efciency of MB was found to follow
CuO spindles (72.5%) > CuO plates (51.9%) > commercial powders (36.6%). Moreover, compared with
the commercial CuO powders with and without the addition of H
2
O
2
, the addition of H
2
O
2
noticeably
enhanced the photocatalytic activity, further suggesting that H
2
O
2
has a crucial function in the
photocatalytic degradation of organic dyes. In addition, 95% of RhB was photodegraded after ve suc-
cessive cycles with almost unchanged photocatalytic activity of 3D mesostructural CuO44 spindles,
indicating a stable photocatalytic activity. The superior photocatalytic performance of 3D CuO
spindles is related to their large surface area.
A number of studies have focused on the similar type of morphology-dependent photocatalytic
ability of CuO in the presence of H
2
O
2
[83,399,454456]. This result further indicates that the surface
area together with the presence of H
2
O
2
has a critical function in the degradation efciency and rate of
organic dyes. The photocatalytic degradation of dyes by CuO in the presence of H
2
O
2
under light
Fig. 112. Time proles of MB degradation: (a) MB + plate-like CuO. (b) MB + H
2
O
2
. (c) MB + H
2
O
2
+ ower-like CuO. (d)
MB + H
2
O
2
+ boat-like CuO. (e) MB + H
2
O
2
+ plate-like CuO. (f) MB + H
2
O
2
+ ellipsoid-like CuO. (g) MB + H
2
O
2
+ commercial CuO.
Error bars were made from three repeated trials [85]. Copyright 2011 Elsevier.
Fig. 113. (a) Photodegradation plot of RhB by different catalysts. (A) Without any catalyst and H
2
O
2
; (B) commercial CuO and
without H
2
O
2
; (C) commercial CuO and H
2
O
2
; (D) 2D mesoplates and H
2
O
2
; (E) 3D mesospindles and H
2
O
2
. (b)
Photodegradation performance of 3D CuO spindles toward the mixture of RhB and H
2
O
2
solution within ve cycles under
visible light irradiation [142]. Copyright 2013 Royal Society of Chemistry.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 303
irradiation is supposed to occur by the following mechanism [142,457]. When CuO photocatalyst ab-
sorbs a photon with energy equal or higher than the bandgap energy (E
g
), electrons (e

) in the VB can
be excited to the CB, with simultaneous generation of the same amount of holes (h
+
) in the VB
[142,449,451,457]. The formed e

and h
+
pairs can be captured by H
2
O
2
molecules, leading to oxidant
formation (

OH,

OOH, and

O

2
). These oxidant species favorably react with a dye (R) or other organic
compound, producing wide ranges of intermediates. These intermediates include radical and radical
cations to realize complete mineralization with the formation of CO
2
, H
2
O, or other inorganic ion as
illustrated in Fig. 114.
Various physical parameters, such as particle size, shape, composition, and structure, can inuence
the properties of CuO nanostructures. Hence, tailoring the photocatalytic properties of CuO nanostruc-
tures is possible by controlling these parameters. Thus, CuO nanostructures with various morpholo-
gies and differences have recently been used as photocatalysts to degrade organic pollutants
without additional H
2
O
2
.
Wang et al. [458] compared the photocatalytic degradation of RhB between commercial CuO and
CuO nanowires. They found that 95.5% of RhB is degraded using CuO nanowires as catalyst, whereas
only 39.6% of commercial CuO is degraded under the same 9 h UV duration (Fig. 115). The photo-
graphic images of RhB solutions under different irradiation times as shown in Fig. 115 clearly demon-
strate that the color of the suspension nally changed into colorless after irradiation for 11 h using
CuO nanowires as the catalyst. The higher activity of CuO nanowires was attributed to the 1D nano-
structure with an average diameter of less than 15 nm. This nding is benecial for improving the ef-
ciency of the electronhole separation, transferring to the surface, and thereby enhancing the activity
of the photocatalytic reaction.
The function of CuO nanowires for photocatalytic degradation of organic dyes has also been paid
much attention [125,459]. Li et al. [124] studied the photocatalytic degradation of RhB using CuO
catalysts with different morphologies. Compared with commercial CuO powder, solution-phase-
synthesized ultralong CuO nanowires have the best photocatalytic activity. This characteristic is
attributed to the highest exposure of (002) crystallographic surfaces, polycrystalline structure, and
small size of particles of the CuO nanowires. The relationship between the photocatalytic activity
and the crystallographic direction ratio of [002]/[200] shown in Fig. 115 clearly demonstrates that
the higher the [002]/[200] crystallographic direction ratio, the better the photocatalytic performance
of the CuO catalysts.
CuO nanostructures with various morphologies (1D nanoseeds and nanoribbons, 2D nanoleaves,
and 3D shuttle- and shrimp-like structure and nanoowers) are hydrothermally synthesized with
PEG as structure directing reagent by Liu et al. [47]. The effects of morphology and surface crystal
planes on the photocatalytic degradation of RhB in water were investigated under sunlight irradiation.
To avoid the inuence on the photocatalytic activity of the CuO nanostructures by PEG 200, IR spec-
troscopy was conducted to determine the existence of organic species in the nal products. No C@O
Fig. 114. Schematic illustration of photo-induced formation of electronhole pairs in a semiconductor CuO with H
2
O
2
in the
presence of water pollutant (R).
304 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
vibration band was found in the IR spectroscopy of CuO, indicating the absence of PEG200 molecules
oxidized into carboxylic acids and modied on the surface of the CuO nanostructures (Fig. 116).
These photocatalytic results showed that different morphologies of CuO structures demonstrate
different photocatalytic activities on the photodecomposition of RhB aqueous solution. The structures
with the same exposed {100} facets (nanoowers, NLs, and nanosheets) were reported to have a high-
er photocatalytic activity than nanoseeds and shuttle- and shrimp-like structures because of their lar-
ger surface area. However, CuO nanoribbons with relatively smaller surface area than the CuO
nanoowers and NLs were found to exhibit the best photocatalytic activity. TEM analyses indicated
that only the CuO nanoribbons have high Miller index facets of the type {121}, whereas the other
CuO structures all have {100} facets. Thus, the high photocatalytic activity of the CuO nanoribbons
may be related to the exposed high surface-energy facet, which has a higher chemical activity than
that of the low surface-energy facet. Two possible as-cleaved terminations are found for a {100} ter-
mination, namely, a Cu terminating layer and an oxygen terminating layer (Fig. 116c), whereas the
termination for the {121} plane is more complex. This complicated atomic stacking of {121} facets
allows the O atoms at the surface be more highly polarized than the ones at the {100} facets. This phe-
nomenon leads to a higher chemical activity of the {121} facets and enhanced photocatalytic activity
of the CuO nanoribbons. These experimental results demonstrate that photocatalytic activity is not
Fig. 115. (a) Photocatalytic degradation efciency of RhB on different catalysts under UV light irradiation [458]. (b)
Photographic images of RhB degraded under UV light irradiation by CuO nanowires [458]. Copyright 2012 Elsevier. (c)
Photocatalytic degradation of RhB on different catalysts on UV light irradiation [124]. (d) Relationship between the
photocatalytic degradation activity of RhB on CuO nanowire bundles at 8 h and the [002]/[200] crystallographic direction ratio
[124]. Copyright 2010 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 305
only inuenced by the morphology but also by the exposed high energy surface of the CuO
nanostructures.
Although photocatalytic activity can be improved by synthesizing novel CuO nanostructures,
preparations of unique CuO/carbon-based or CuO/other MOs composites have been investigated as
alternative materials to further enhance the photocatalytic activity of CuO.
Liu et al. [460] synthesized CuO nanoower-decorated reduced graphene oxide (CuONF/rGO)
nanocomposite by heating a mixture of Cu salts and GO in the presence of PQ11 by a one-pot route
(Fig. 117a) and used it as a photocatalyst for RhB degradation in the presence of H
2
O
2
. Results show
that the photocatalytic activity of CuONF/rGO nanocomposite is much higher than that of single CuO
nanoower or rGO under the same testing condition as shown in Fig. 117b. The increase was attrib-
uted to the fast adsorption of RhB by pp stacking interactions between RhB and the p-conjugation
system of CuONF/rGO together with the novel 2D planar structure of CuONF/rGO, leading to rapid
RhBdegradation.
Shi and Chopra [461] investigated the photocatalytic activity of CuO/Co
3
O
4
nanowire heterostruc-
tures prepared by sputtering and subsequent air annealing for phenol degradation. They observed that
a thick polycrystalline Co
3
O
4
shell (50 nm) on CuO nanowires shows the best photodegradation
performance (efciency: 5090%) in a low-powered (8 W) UV or visible light illumination in the
presence of H
2
O
2
. More interestingly, CuO nanowires coated with a thin (<10 nm) Co
3
O
4
shell with
embedded nanoparticle show an anomalously high efciency (67.5%) under visible light without
H
2
O
2
. The high photocatalytic activity of CuO nanowires after combining with Co
3
O
4
was related to
Fig. 116. A: IR spectra of the PEG200 and the CuO nanostructures:(a) PEG200, (b) nanoseeds, (c) shuttle-like nanostructures, (d)
shrimp-like structures, (e) nanoowers constructed by nanorods, (f) nanosheets, (g) NLs, (h) nanoowers, and (i) nanoribbons.
B: Photodecomposition activity of the catalysts with or without the catalysts. (a) Without CuO nanostructures, (b) nanoseeds,
(c) shuttle-like nanostructures, (d) shrimp-like structures, (e) nanoowers constructed by nanorods, (f) nanosheets, (g) NLs, (h)
nanoowers, and (i) nanoribbons. C: (a) Crystal structure of the monoclinic CuO with the unit cell outlined, (b) sideview of the
atomic arrangement of the {100} facet, and (c) sideview of the atom arrangement of the {121} facet. Deep red circles are O
atoms, and pink circles are Cu atoms [47]. Copyright 2012 Elsevier.
306 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
the novel morphology and interfaces of CuOCo
3
O
4
nanowire heterostructures, which facilitate charge
transfer and separation.
5.8. Application in enhancement of thermal conductivity of nanouid
Nanouid refers to a new class of heat transfer uid containing nanosized particles, bers, or tubes
that are stably suspended in a carrier liquid [462]. Compared with the traditional uid, a nanouid
exhibits signicant enhancement of thermal conductivity because of size effect and Brownian motion
of the nanoparticles in liquids. A nanouid can be applied in various important elds, including micro-
electronics, aerospace, transportation, and medicine [463]. Nonmetallic solid CuO, with a relatively
high thermal conductivity of 76.5 W/m k, has been considered as additives in conventional uids to
enhance thermal conductivity [463].
Lee et al. [464] demonstrated that a nanouid consisting of CuO nanoparticles in ethylene glycol
exhibits enhanced thermal conductivity. A maximum increase of 20% was observed for 4 vol% CuO
nanoparticles with an average diameter of 35 nm dispersed in ethylene glycol. A similar study
[465] also achieved a maximum of 22.4% enhancement of the thermal conductivity of ethylene glycol
in the presence of 5 vol% CuO nanoparticles. This result was attributed to the higher thermal conduc-
tivities of CuO solid materials along with the larger specic surface area of CuO nanoparticles, which is
important to enhance thermal conductivity. Zhu et al. [463] synthesized CuO nanouid with high vol-
ume fractions by transforming an unstable Cu(OH)
2
precursor to CuO in H
2
O under an ultrasonic
vibration, followed by microwave irradiation. The thermal conductivity of CuO nanouid increases
with increasing nanoparticle volume fraction. The enhancement of thermal conductivity of nanouid
obtained by the authors was higher than that of the reported results (Fig. 118). The enhancement is
mainly caused by the good dispersion and shuttle-like shape (aspect ratio is 5) of the CuO
nanoparticles.
Moghadassi et al. [466] recently investigated the effect of CuO nanoparticles on the thermal con-
ductivities of parafn and MEG. The authors found that the effective thermal conductivity of the uids
increases with an increase in the volume fraction of nanoparticles in the suspension. The addition of
more nanoparticles or the reduction of nanoparticle size in the suspension further enhances the ther-
mal conductivity of the uids. Large-scale stable CuO nanouid was also developed by Zhu et al. [467]
via a wet chemical method. The thermal conductivity of CuO nanouid was observed to increase with
an increase in particle loading. The development of novel methods for obtaining nanouid with well-
dispersed small-sized nanoparticles, high thermal conductivity, long-term stability, and high yield in
large-scale may have great potential application into advanced vehicles.
Fig. 117. (a) A scheme (not to scale) to illustrate the formation of CuONF/rGO nanocomposite from GO and Cu salts by a one-pot
heat treatment method. (b) Photodegradation of RhB (a) over CuONF/rGO (j), without catalyst (d), over CuO nanoparticles (N),
and rGO (.) [460]. Copyright 2012 Royal Society of Chemistry.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 307
Temperature and morphology of nanoparticles also remarkably affect the thermal conductivity of a
nanouid [151,468,469]. Therefore, the conductivity of CuO nanoparticle-based nanouid is strongly
dependent on the test temperature and synthesis conditions. For example, Das et al. [468] indicated
that a two to fourfold increase in thermal conductivity enhancement of nanouids is feasible over a
temperature range from 21 C to 51 C. Smaller CuO particles show enhanced conductivity with
temperature. Hojjat et al. [469] reported that the thermal conductivity of nanouids and the base uid
increases with temperature, and the effect of temperature are more signicant at high nanoparticle
concentrations. Dey et al. [152] presented temperature dependence of the thermal conductivity
enhancement of CuOH
2
O nanouids prepared fromdifferent CuOnanoparticles (Fig. 119). They found
that the particles with seed-like shape are most effective in enhancing the thermal conductivity of the
base uid. These results make nanouids more attractive as cooling uid for high-energy density
devices where the cooling uid is likely to work at a temperature higher than the room temperature.
5.9. Application in nEMs
EMs, including explosives, pyrotechnics, and propellants, are widely used in airbag igniters, mining,
de-construction, fuses, joining, soldering, brazing, and also in numerous defense-related areas. NEMs
Fig. 118. Comparison between the experimental data and the reported values of CuO nanouid. K
e
and K
f
represent the thermal
conductivity of nanouid and base uid, respectively [463]. Copyright 2008 American Chemical Society.
Fig. 119. Temperature dependence of the thermal conductivity enhancement of CuOwater nanouids prepared from different
CuO nanoparticles. The pH synthesis and the corresponding shapes of the particles are provided in the inset [152]. Copyright
2012 Royal Society of Chemistry.
308 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
have remarkably improved performance in reaction rate, propagation rate, and ignition compared
with conventional EMs. Therefore, nEMs can be potentially applied in actuation, ignition, propulsion,
power, welding, uidic, and electro-explosive devices at the microscale and nanoscale [42,470472].
One important category of nEMs is nanoscale thermite, which is a composition of a nanometal and a
nano-MO that produces an exothermic oxidationreduction reaction known as a thermite reaction.
Among the different nano-MOs, nanoscale CuO is commonly combined with nanometal (mainly Al)
to obtain high-performance nEMs [215,473487]. Various approaches are used to combine nano-
CuO with Al to achieve nEMs. Dreizin et al. used arrested reactive milling to mix CuO nanoparticles
with Al to obtain CuO/Al nEMs (nanocomposite) [474476]. Sanders et al. used physical mixing to
combine CuO nanoparticles with Al to obtain CuO/Al nanoscale energetic composites [477]. CuO thin
lms with nanoscale thickness were also combined with Al or Ti by sputtering to realize reactive mul-
tilayer foils [478480].
Jian et al. [215] recently synthesized novel hollow CuO spheres by a droplet-to-particle aerosol
spray pyrolysis method with the introduction of sucrose and H
2
O
2
(gas-blowing agent) into the
Cu(NO
3
)
2
3H
2
O precursor solution. CuO/Al-based nEMs was obtained by integrating Al nanoparticles
with hollow CuO spheres via ultrasonic mixing. The nEMs reached a pressure rise of 0.896 MPa in
1.2 ls with the potential of rapid oxygen release during reaction, indicating promising applications
in nanoenergetic gas generators.
The latest study of Sverac et al. [486] presented DNA-directed assembly of CuO and Al nanopar-
ticles to obtain CuO/Al-based nEMs. The CuO and Al nanoparticles were rst suspended and stabilized
in an aqueous solution for hours. The CuO nanoparticles were functionalized with thiol-modied
oligonucleotides, and these oligonucleotides were added after the adsorption of neutravidin to Al
nanoparticles. Then, the CuO and Al nanoparticles were coated with non-complementary DNA strands.
The aggregation growth rate of CuO and Al became linear time dependence by adding an excess linker
that is complementary to both the strands coated on CuO and Al. The synthetic process is shown in
Fig. 120. The resulting CuO/Al nEMs showed a heat of reaction of 1.800 J/g, which is among the best
value ever achieved.
1D CuO nanowires and nanorods are used as templates to achieve nEMs by integrating with metal
Al through self-assembly [481485]. This synthesis method can obtain nEMs with nanoscale mixing
Fig. 120. Schematic of the DNA-directed assembly of Al/CuO nEMs [486]. Copyright 2012 Wiley.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 309
uniformity and improved performance. Apperson et al. [481] used an adhesive polymer to assemble Al
nanoparticles along CuO nanorods to obtain CuO/Al-based nEMs. In this approach, a large number of
Al nanoparticles can be integrated onto the sidewalls of CuO nanorods. However, the attachment of Al
nanoparticles to the nanorods is not uniform along the axial direction of the nanorods. Self-assembly
is used to synthesize CuO/Al-based nEMs by integrating nano Al (by thermal evaporation or magne-
tron sputtering) with CuO nanowires grown from Cu thin lm to form core/shell structures with
CuO nanowires as the core and Al as the shell [482485]. This approach is advantageous because of
its improved mixing uniformity, enhanced contact, reduced impurities, and lower activation energy.
Zhang et al. [482] synthesized CuO/Al-based nEMs by integrating Al with aligned CuO nanowires
on a silicon substrate. Vertically aligned CuO nanowires were grown by thermal oxidation of a Cu lm
deposited on silicon. Al was then deposited onto the CuO nanowires by thermal evaporation, forming
the core/shell nEMs. The CuO nanowires and the core/shell CuO/Al nEMs are shown in Fig. 121a and b,
respectively. The tested heat of reaction of the CuO/Al nEMs was 2950 J/g by thermal analysis. Ohkura
et al. [483] also achieved CuO/Al core/shell nEMs by integrating nano Al with CuO nanowires. The
authors found that the onset temperature of nEMs ranges from 545 C to 569 C with different Al
thicknesses of 100425 nm. This result indicates the insensitivity of ignition temperature to the equiv-
alence ratio. However, the onset temperature of nEMs is much lower than that of the micron-sized
CuO and Al EMs.
A promising direction of nEM research is to integrate nEMs with silicon-based microelectrome-
chanical systems (MEMS) to achieve functional nanoenergetics-on-a-chip (NoC). NoC is the key for
various nEM applications, such as actuation, ignition, propulsion, power, uidic, and electro-explosive
devices, at the microscale and nanoscale [42]. To achieve NoC, Zhang et al. [484] synthesized Al/CuO
nanowire-based nEMs onto silicon, a basic material for microelectronics and MEMS. Then, they devel-
oped a functional nanoinitiator by integrating the Al/CuO-based nEMs with an Au/Pt/Cr microheater
fabricated onto a substrate, where the nEMs were successfully ignited by the microheater through
joule heating. The CuO/Al nEMs integrated on the microheater are shown in Fig. 122a. The initiator
has high output energy, and the ejected high temperature products (Fig. 122b) can ensure the success-
ful ignition of the attached reactive material even without contact. The ignition power, ignition delay,
and ignition energy of the initiators were 1.16 W, 0.10.6 ms, and 0.120.70 mJ, respectively. The area
and mass of the CuO/Al nEM lm in Fig. 122a are 1.2 1.2 mm
2
and 0.02 mg, respectively, which
produces the combustion ame shown in Fig. 122b. The results from Ref. [484] contributed to the
achievement of functional NoC devices.
Yang et al. [485] synthesized CuOx/Al-based micro and nano EMs by integrating Al with CuO
microstructures and nanostructures fabricated onto silicon substrates. The heat released before Al
melting (with an onset temperature of 500 C) correlates with the amount of the CuO nanowires.
The experiments also indicated that the nanostructures (CuO nanowires and nano Al) can help reduce
the ignition delay and ignition energy of the CuOx/Al-based EMs. A possible key factor that contributes
to the enhanced thermite reactions and ignition properties of the nEMs is the higher surface energy
associated with the CuO nanowires and the solidsolid diffusion process promoted by the nano Al.
Fig. 121. (a) CuO nanowires grown from Cu lm. (b) CuO/Al core/shell nEMs [482]. Copyright 2007 American Institute of
Physics.
310 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
Thiruvengadathan et al. [487] comprehensively studied the combustion characteristics of CuO/Al
nEMs synthesized by self-assembly of CuO nanorods and Al nanoparticles. They rst coated a PVP
monolayer around CuO nanorods, followed by integration with Al nanoparticles via self-assembly.
The CuO/Al nEMs exhibited enhanced pressurization rate and combustion wave speed compared with
those of random mixture between CuO nanorods and Al nanoparticles. The authors also achieved
hybrid nanoenergetic composites by integrating the self-assembled CuO/Al nEMs with explosives
(NH
4
NO
3
, RDX, and CL-20). The observed best combustion characteristics were 2600 50 m/s of
combustion wave speed and 11 1 MPa/ls of pressurization rate. The presence of OH groups obtained
during the wet-chemical synthesis of CuO nanorods contributed to the increased gas generation,
which accelerated ame propagation.
5.10. Application in eld emission displays (FEDs)
FED has high brightness, good color rendition, short response time, and low power consumption
[488]. The use of 1D nanostructures in FEDs has been extensively studied. Among the various 1D nano-
structure materials, CuO nanowires are widely chosen as eld emitters because of the low turn-on
eld, high current density, and low fabrication cost.
Zhu et al. [43] synthesized CuO nanowires by thermal oxidation method and studied their eld
emission properties. The synthesized nanowires are vertically aligned with high density and sharp
tips, which is benecial to eld emission. The results showed that the CuO nanowires have turn-on
eld of 3.54.5 V/lm and high emission current density of 450 lA/cm
2
at 7 V/lm. To enhance the eld
emission properties of CuO nanowires, Zhu et al. [489] adopted the plasma treatment after CuO nano-
wire synthesis. The CuO nanowires were prepared by thermal oxidation method. Then, the CuO nano-
wires were exposed to the plasma generated from oxygen or CF
4
gas. Comparative results of the CuO
nanowires before and after the plasma treatment showed that the nanowires have sharpened tips and
reduced average diameter, even bent and slightly shortened. These changes in morphology resulted in
lower turn-on eld and higher current density. The current density under 5.3 V/lm increased from
0.13 mA/cm
2
to 0.34 and 0.54 mA/cm
2
after 5 and 10 min of oxygen plasma treatment, respectively,
whereas the turn-on eld decreased from 3.6 V/lm to 3.2 and 3.0 V/lm after 5 and 10 min of treat-
ment (Fig. 123a). Similarly, the maximum current density shown in Fig. 123b increased from
0.04 mA/cm
2
to 1 mA/cm
2
and turn-on eld decreased from 4.2 V/lm to 3.7 V/lm after 10 min expo-
sure to plasma produced by CF
4
. These results indicated that the eld emission properties of CuO
nanowires are signicantly enhanced after additional plasma treatment.
Wang and Li [248] improved the eld emission of CuO nanowire arrays by coating a ZnO layer on a
Cu substrate before thermal oxidation. The ZnO layer was deposited by immersing a 0.1 mm Cu foil
into an aqueous solution containing zinc nitrate hexahydrate and hexamethylenetetramine at 95 C
for several hours, followed by washing and drying. Then, the ZnO-coated Cu foil was heated at
400 C for 4 h in ambient air. Compared with the directly heated Cu foil without ZnO coatings, the
Fig. 122. (a) CuO/Al nEM-based nano initiator. (b) Optic image of the combustion ame of the nano initiator [484].
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 311
synthesized ZnO-coated CuO nanowires exhibit higher density and average length, lower average
diameter, and higher aspect ratio. The turn-on eld of the ZnO-coated CuO nanowire array was
0.85 V/lm at a current density of 10 lA/cm
2
. The CuO nanowires without ZnO coatings exhibited
the turn-on eld of 6.5 V/lm at the same current density. Moreover, the ZnO-coated sample can reach
a maximum current density of 230 lA/cm
2
at a eld of 4 V/lm. The authors reported that the
enhanced eld emission properties may have resulted from the enlarged nanowire density, aspect
ratio, and crack elimination.
Liu et al. [490] reported the enhanced eld emission properties by increasing the density of CuO
nanowires, which were synthesized in water vapor environment. A 1-lm-thick Cu lm was deposited
onto an ITO substrate by sputtering, followed by heat treatment at 430 C for 3 h. Although the results
proved that the CuO nanowires prepared under wet environment have slightly improvement in eld
emission properties, the measured lowest turn-on eld (7.5 MV/m) corresponding to the current den-
sity of 10 lA/cm
2
is still relatively high, compared with other reports. Hsueh et al. [491] fabricated
CuO nanowire eld emitter arrays on a glass substrate. The Cu lm was deposited onto the substrate
after the deposition of Ti and CuO layers to prevent peeling off. Then, the Cu lm was heated at 450 C
for 5 h in air. The turn-on eld of the as-fabricated eld emitter was 4.5 V/lm. Zhan et al. [492] re-
ported the fabrication of gated CuO nanowire eld emitter arrays on a glass substrate. Cr and Al thin
lm layers were rst deposited by sputtering and patterned as cathode electrode on the substrate, fol-
lowed by the deposition of isolating thin lm layers of SiO
x
SiN
x
SiO
x
. Subsequently, another 500-nm-
thick Cr and Al thin lm was deposited and patterned onto the isolating layer as gated electrode. Then,
the isolating layer without the protection of Cr/Al layer was etched using the reactive ion etching (RIE)
process. The authors also introduced a cleaning and lift-off process after RIE process to further remove
the etch residues. A 50 nm Cr thin lm and a 1000 nm Cu thin lm were then sputtered and patterned
on the exposed cathode electrode. Finally, the sample was annealed at 400 C for 3 h for the growth of
CuO nanowires. A prototype comprised CuO nanowire eld emitter arrays was presented in Ref. [492].
Densely packed CuO nanowires grew on the patterned Cu thin lms (Fig. 124a and b). The pictures of
the prototype and display images of Chinese characters and English letters are shown in Fig. 124c.
Although the characters and letters are identiable, the uniformity of the images still needs to be
improved.
5.11. Application in superhydrophobic surfaces
Superhydrophobic surfaces commonly refer to surfaces with a water contact angle (CA) of 150 or
larger. When water droplets land on these surfaces, they can move freely and roll off with dust
Fig. 123. (a) Field emission JE curves and FowlerNordheim (FN) plots (inset) of as-grown and O
2
plasma-treated CuO
nanowires. (b) Field emission JE curves and FN plots (inset) of as-grown and CF4 plasma-treated CuO nanowires [489].
Copyright 2006 Elsevier.
312 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
particles and surface contaminants [493,494]. In recent years, great interest has been focused on the
design and fabrication of superhydrophobic surfaces because of their promising applications in many
elds, such as Lotus effect self-cleaning coatings (anti-biofouling), surface protection, textiles, water
movement, microuidics, and oilwater separation [146,147,242,493502]. To achieve a surface with
superhydrophobic properties, a low-energy surface must be combined with high surface roughness.
Generally, two different processes are widely used by researchers to form superhydrophobic surfaces.
The rst process creates hierarchical micro-/nanostructures on hydrophobic substrates, whereas the
other chemically modies a micro-/nanostructured surface with low-surface energy materials [499].
In 2006, Liu et al. [146] studied the water CAs of CuO ower-like structures that comprised hierar-
chical 2D nanosheets and CuO spherical architectures constructed by ultra-thin nanowalls as shown in
Fig. 16. These CuO nanostructures were synthesized on the Cu substrate surface by Cu oxidation under
hot alkaline conditions. Prior to the water CA measurement, the obtained samples were chemically
modied with a wax layer. For comparison purposes, a at Cu foil coated with a wax layer was also
fabricated. The water CAs of CuO ower-like structures and spherical architectures were measured
to be 151 and 147, respectively. However, the CAs of modied at Cu surface was determined to
be only 93. This result is due tothe fact that hierarchical nanostructures grown on the Cu surface
can strongly enhance hydrophobicity.
Liu et al. [147] obtained novel cabbage-like CuO hierarchical microstructures synthesized on a Cu
foil by a similar synthetic process to that used in Ref. [146]. They observed that the cabbage-like CuO
hierarchical microstructures on the Cu foil can also lead to superhydrophobic surface with a high
water CA of 155 after modication with a wax as shown in Fig. 125b, and the CA of a at Cu foil
coated with wax layer is only 102 as shown in Fig. 125e. However, the CA of the as-prepared cab-
bage-like CuO without wax modication is lower than 10 (Fig. 125f). This result suggests that mod-
ication of the rough surface with low surface energy materials is also crucial to the superhydrophobic
property. Interestingly, the observed superhydrophobicity was found to have long-term stability, and
the treated surface can resist the attack of acid and alkali at ambient temperature. Similar promising
results are also obtained by several other groups [495497].
Fig. 124. (a) and (b) Top and cross-sectional view of SEM images of the gated CuO nanowire emitter arrays, respectively. (c)
Picture of a vacuum-packaged 3.5 in.-gated FED using a CuO nanowire emitter [492]. Copyright 2010 American Vacuum Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 313
The superhydrophobic surfaces are mainly characterized by low-energy surface and high surface
roughness. Therefore, the use of low-energy surface materials modied on CuO surfaces, as well as
the change in the surface chemical composition, morphology, and surface roughness, can directly af-
fect the wetting behavior of CuO. Basu et al. [498] investigated the wetting behavior of CuO thin lm
synthesized on a smooth glass substrate that was modied with various long chain amine or thiols.
They found that the as-prepared CuO without modication shows CAs lower than 5, whereas the
CAs of amine- and thiol-modied samples were 130 and 150, respectively. These results reveal that
thiol modication is better than amine modication for obtaining greater CAs. In addition, the re-
quired time for different structures of thiols and amine to change a hydrophilic CuO lm into a hydro-
phobic CuO surface was found to be different. This behavior is attributed to the long chain thiols that
are more efcient surface functionalizing agent than amine. Consequently, thiols can encapsulate the
CuO thin lms better because of their stronger afnity to the CuO surface.
Li et al. [499] obtained superhydrophobic CuO surfaces by combining both simple solution-
immersion process and self-assembly of uoroalkylsilane. The as-prepared superhydrophobic surfaces
showed tunable water adhesion that ranged from extremely low to very high. The superhydrophobic
surface adhesion is strongly dependent on the microstructure congurations of CuO on the superhy-
drophobic surfaces. The remarkable results as observed by the authors provide a new strategy to
prepare superhydrophobic surfaces with tunable water adhesion. Recently, Zhang et al. [500] obtained
controllable superhydrophobic micro- and nanostructured tier surfaces, and each of these surfaces is
made of self-assembled CuO. The wetting behavior of the micro- and nanostructured CuO tier surfaces
is closely related to morphological information, such as pitch, height, and shape.
Chaudhary and Barshilia [501] prepared superhydrophobic CuO/Cu(OH)
2
surfaces by a simple solu-
tion-immersion process at room temperature without low surface energy material. The authors
proved that the wetting behavior of the CuO/Cu(OH)
2
surfaces can be changed by changing the
morphology of the CuO nanostructures. Furthermore, they observed the wettability transition of the
CuO/Cu(OH)
2
surfaces from superhydrophobic to superhydrophilic by the alteration of oxygen plasma
treatment and dark storage. The reversible transition between superhydrophobicity and superhydro-
philicity is connected to the modication of the surface roughness and the surface chemical compo-
sition under the inuence of oxygen plasma. The transition between superhydrophobicity and
superhydrophilicity can also be achieved on the CuO nanowire/lm surface synthesized by simply
heating the Cu substrates in air at 400 C for different durations [242]. The superhydrophobicity is
attributed to the partial deoxidation of the upmost layer of CuO surfaces into Cu
2
O-like hydrophobic
Fig. 125. (a) Optical image of many water droplets on the modied Cu substrate. (bf) Shapes of a water droplet on different
surfaces and the corresponding CA values. (b) Modied CuO microcabbage surface. (c and d) Modied underdeveloped CuO
microcabbage surfaces. (e) Modied at Cu surface. (f) As-prepared CuO microcabbage surface [147]. Copyright 2008 Elsevier.
314 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
surfaces by the authors. Nonetheless, superhydrophilicity is recovered if the superhydrophobic CuO
lm is subjected to high temperature annealing.
Guo et al. [502] rst synthesized CuO nanowire array by thermal oxidation of Cu foil at 500 C for
4 h in the air. Then, ZnO nanorods were integrated around the CuO nanowires by a simple hydrother-
mal approach to obtain ZnO/CuO heterohierarchical nanotree array on the Cu substrate surface. The
SEM images of a heterohierarchical nanotree array are shown in Fig. 126a, where the inset is the sche-
matic of the heterohierarchical structure. The interesting ZnO/CuO heterohierarchical nanotree array
after silanization shows remarkable superhydrophobic performance. The CA on a surface of CuO nano-
wires array was 161.3, and the samples with ZnO/CuO hierarchical nanostructure presented higher
static CAs (Fig. 126b). The static CA of ZnO/CuO hierarchical nanostructure increased with an increase
in density of hierarchical nanorods and reached a maximum value of 168.4 for sample c. Moreover,
their wettability can be manipulated by the morphologies of hierarchical ZnO nanorods. The authors
attributed the improved superhydrophobic performance to the trapped air and hierarchical roughness.
The self-cleaning behavior of the ZnO/CuO heterohierarchical nanotree array surface was also dem-
onstrated in Fig. 127 [502]. Dust particles (black real dots) were placed on the nanotree array surface
along the red dash line. When the sample surface was lifted to tightly contact a droplet suspended on a
Fig. 126. (a) Cross-sectional view of the ZnO/CuO heterohierarchical nanotree array on a Cu substrate. (b) Statistical
distribution of the static contact angle measured in 10 different locations for each sample [502]. Copyright 2011 American
Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 315
microsyringe and then moved from one end (Fig. 127a) to the other end, the water droplet remained
intact without any apparent deformation. Interestingly, the dust particles entered the water droplet
and accompanied the sliding of the droplet without any dust particle residue. These results prove that
the ZnO/CuO heterohierarchical nanotree array surface has excellent self-cleaning performance.
In summary, different factors, such as surface chemical composition, morphology, and surface
roughness along with the use of low-energy surface materials for modication, can affect the wetting
behavior of nanostructured CuO synthesized on at substrates. The synthesis of novel hierarchical
micro-/nanostructured CuO with large surface roughness and carefully designed surfaces are still
required for practical applications in self-cleaning surfaces and various other industrial processes.
5.12. Application in removal of arsenic (As) and organic pollutants from waste water
In recent years, the awareness of As contamination in natural water has increased remarkably be-
cause of its several side effects and lethality [45,46,503505]. Therefore, great efforts have been ex-
erted in developing effective and affordable technologies for As removal from water. Adsorption
process adsorbs pollutants on the adsorbent surface and is reported to be a cost-effective and facile
method [45,46,503,504]. Conventional adsorbents, including aluminum, iron, titanium, zirconium,
and manganese, have been investigated extensively to remove As from water [503]. However, conven-
tional adsorbents have multiple disadvantages that restrict the capability and effectiveness of As re-
moval, especially for the more toxic As(III) compared with As(V). Therefore, the development of
new effective adsorbents for As removal is a research priority of many scientists.
In 2005, Reddy and Attili proved that CuO is an effective adsorbent for As removal because no
pre-treatment steps are needed to change the pH or oxidation of As(III) to As(V). The presence of
interfering ions, namely silicate and phosphate, does not interfere in the removal process [506]. The
effective removal of As by CuO was attributed to its high point of zero charge of 9.4 0.4 [506,507].
Subsequently, Roth and Reddy et al. [508] tested the effectiveness of CuO for As removal using a
Fig. 127. Schematic of self-cleaning behavior for ZnO/CuO heterohierarchical nanotree array. To guarantee the controllability of
the experiment, a water (not viscous) droplet slides (not rolls off) the surfaces [502]. Copyright 2011 American Chemical
Society.
316 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
ow-through system. Their results showed that CuO can effectively remove both As(III) and As(V)
from the groundwater without affecting the pH or other ions.
Cao et al. [509] investigated the effectiveness in As removal using different hierarchical nanostruc-
tured CuO. Compared with commercial nanoscale CuO, doughnut-like CuO structures were found to
have a better removal capacity for As(III). These novel doughnut-like CuO can also be easily separated
and recycled during the water treatment process. This nding was attributed to its large specic sur-
face and unique 3D hierarchical porous structure with micrometer-sized holes on the surface of the
particle, which allows easy access on inner nanometer-sized crystals and pores. Pillewan et al.
[507] examined the efciency of CuO-incorporated mesoporous alumina (COIMA) for the removal
of As(III) and As(V) from water. Their results showed that the adsorption capacity of COIMA is signif-
icantly higher than those of unmodied alumina. Simultaneous removal of As(III) and As(V) with re-
moval efciencies of more than 95% for both As(III) and As(V) was achieved without affecting the
water quality.
Yu et al. [46] recently reported novel cottoncandy-like CuO 3D hierarchical microspheres synthe-
sized by a facile surfactant-free solvothermal route and subsequent calcination process and used them
as the adsorbent for As(III) removal. SEM and TEM results (Fig. 128) demonstrated that the cottoncan-
dy-like CuO hierarchical microspheres were highly porous and composed of interwoven nanobers
with a diameter of 13 lm. As(III) adsorption experiments showed that these novel CuO hierarchical
nanostructures exhibited fast adsorption rates and higher removal capacity than other reported nano-
structured CuO as summarized in Table 10. The enhancement was due to their novel 3D highly porous
Fig. 128. SEM images of the CuO precursor with (a) low and (b) high magnications. TEM images of the CuO precursor with (c)
low and (d) high magnications [46]. Copyright 2012 American Chemical Society.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 317
structure and large surface area. Moreover, the presence of most of the competing anions in water and
pH values have no inuence on the adsorption capability of As(III) onto the CuO microspheres as
shown in Fig. 129. XPS results showed that the whole adsorption process involves As(III) adsorption
onto CuO and As(III) oxidation to As(V), followed by As(V) adsorption onto CuO sample. These prom-
ising results suggest that the cottoncandy-like CuO microspheres are effective for As(III) adsorption
and can be used to develop a simple and efcient As(III) removal method.
5.13. Toxicity of CuO nanoparticles
CuO nanoparticles are highly toxic to humans [510]. Compared with other MO nanoparticles and
CNTs, CuO nanoparticles show the highest cytotoxic potential and most DNA damage and oxidative
DNA lesions [510]. Karlsson et al. [511] compared the toxicity of nano- and micrometer particles of
CuO as well as of Fe
2
O
3
, Fe
3
O
4
, and TiO
2
. These MOs cause cell death, DNA damage, mitochondrial
damage, and oxidative DNA lesions (Fig. 130). CuO nanoparticles are much more toxic compared with
CuO micrometer-sized particles and the most toxic among the various MOs.
However, the mechanism of toxicity induced by the CuO nanoparticles remains controversial. In
Ref. [512], the authors demonstrated that the toxicity of CuO is attributed to soluble metal ions orig-
inating from the MO particles. Rousk et al. [513] showed a direct acute toxicity of nano-CuO acting on
soil bacteria, and the nontoxicity of the macroparticulate (bulk) form. CuSO
4
was proved to be more
toxic than either oxide form. This result suggests that the engineered nanoparticle (ENP; at least
one dimension between 1 nm and 100 nm) form can be more toxic than the non-ENP form, but only
when the metal dissolution is higher in this form. Horie et al. [514] further conrmed that Cu ion re-
lease is the most important factor of cellular inuence induced by CuO nanoparticles. The primary par-
ticle size and specic surface area of CuO nanoparticles do not directly affect cellular inuences. Risco
et al. [515] argued that the toxicity of nanoparticles can only be partially explained by the dissolution
of CuO nanoparticles to Cu ions. Particle size of CuO should have a direct function in the toxicity of
Table 10
Adsorption capacity comparisons of cottoncandy-like CuO in [46] with other structured CuO adsorbents for As(III) removal.
Absorbents Absorption capability (mg/g) Ref.
Cottoncandy-like CuO 3D hierarchical microspheres >12.9 [46]
Doughnut-like CuO 4.7 [509]
CuO nanoparticles 1.4 [509]
CuO-incorporated mesoporous alumina 2.16 [507]
Fig. 129. Effects of (a) competing anions and (b) pH value on As(III) adsorption onto the cottoncandy-like CuO [46]. Copyright
2012 American Chemical Society.
318 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
these compounds. Karlsson et al. [511] believed that the toxicity of the CuO nanoparticles is likely ex-
plained by the combination of high surface reactivity and high surface area. Dey et al. [152] noted that
the two commonly accepted paradigms for toxicity caused by CuO nanoparticles are (1) the generation
of reactive oxygen species (ROS) and (2) induction of oxidative stress in the cell. As particle dimen-
sions are reduced towards the nanoscale, the surface-to-volume ratio proportionally increases, and
the small size effects associated with nanoparticles become more pronounced. These paradigms result
in the availability of chemically reactive functional groups on the surface that could play a role in the
adverse biological effects. Therefore, large surface areas of the small-sized CuO nanoparticles cause
higher biological activity than their bulk or microcounterparts. Toxic oxidative stress can cause a wide
range of mitochondrial function, including disruption of electron ow in the inner membrane, dissi-
pation of the mitochondrial membrane potential (Dym), and mitochondrial Ca
2+
uptake.
Although the decreasing particle size generally produces more toxic effects, Dey et al. [152] indi-
cated that the morphology of the CuO nanoparticles also has a very important function in the toxicity.
Given the unique disposal of the surface features of the leaf-shaped particles, they were able to induce
more surface-related phenomena. These particles generate more ROS on their surfaces, facilitating the
higher toxicity of leaf-shaped particles compared with the other shaped particles [153]. Applerot et al.
[516] further demonstrated that the potent antibacterial activity of CuO nanoparticles is due to ROS
generation by the nanoparticles attached to the bacterial cells, which in turn provokes an enhance-
ment of the intracellular oxidative stress. This paradigm was conrmed by several assays including
lipid peroxidation and reporter strains of oxidative stress. The established mechanistic route underly-
ing the antibacterial activity of CuO is depicted in Fig. 131, which clearly indicates that oxy radicals,
namely superoxide anions, are generated in CuO water suspensions.
However, the particle size effect of MOs on toxicity should be considered cautiously because the
bulk ZnO is more toxic than the nano-ZnO. The micrometer particles of TiO
2
cause more DNA damage
Fig. 130. (a) Cell viability of A549 cells after 18 h exposure to 40 lg/cm
2
nano- or micrometer particles of different
compositions. (b) Mitochondrial depolarization in A549 cells after 16 h exposure to 40 lg/cm
2
nano- or micrometer particles of
different compositions. (c) DNA damage in cultured A549 cells after 4 h exposure to 40 lg/cm
2
nano- or micrometer particles of
different compositions. (d) Oxidative DNA damage in cultured A549 cells after 4 h exposure to 40 lg/cm
2
nano- or micrometer
particles of different compositions [511]. Copyright 2009 Elsevier.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 319
compared with TiO
2
nanoparticles, and the iron oxides show low toxicity and no clear difference be-
tween the different particle sizes [511,513]. The tiny nanoparticles can easily transfer to the deep area
of the lung or penetrate many body barriers, such as the skin and bloodbrain barrier, and translocate
to other organs [517520]. However, these nanoparticles are not very efciently phagocytized by the
alveolar macrophages, which is the main mechanism for particle clearance in human body. The depo-
sition and aggregation of the CuO nanoparticles cause lung irritation, chronic lung inammations,
asthma exacerbation, and granulomaformation consisting of macrophage-like multinucleated cells.
Therefore, necessary measures should be undertaken to protect the researchers from direct exposure
to CuO nanoparticles. Work or lab coat, rubber gloves, respirator, and safety goggles are the minimum
appropriate level of protection. If potential aerosol exposure threat exists, chemical fume hood or air-
purifying respirator should be used for the processes. After the investigations, the hands should be
thoroughly cleaned.
6. Conclusion and outlook
In summary, CuO nanostructures have been widely studied and are receiving increasing interest
because of their interesting properties and promising applications in various areas. In this study,
we present a comprehensive review of the state-of-the-art research activities of different CuO nano-
structures. We focus on the main synthetic strategies along with associated formation mechanisms,
their interesting fundamental properties, and promising applications. Investigation of the synthetic
strategies for the control and manipulation of CuO nanostructures with manageable precise dimen-
sions and ideal hierarchy of unique structures is crucial for obtaining corresponding unique properties,
which will undoubtedly enable a variety of promising applications that would not be possible for bulk
Fig. 131. Schematic illustration of the antibacterial mechanism of CuO nanoparticles and the relative cellular structure of (a)
Escherichia coli (Gram-negative) and (b) Staphylococcus aureus (Gram-positive). The major structural differences between the
two cells are the thickness of the rigid peptidoglycan layer and the presence of an outer membrane in Gram-negative cells.
Gram-negative cells have a very thin peptidoglycan layer with only a few molecules thick, whereas Gram-positive cells have a
very thick layer. The carotenoid pigments of S. aureus provide integrity to its cell membrane and increase its protection against
oxidative stress. The bacterial cell damage is mediated by the harmful superoxide anions formed by the cells attached/
internalized CuO nanoparticles [516]. Copyright 2012 Wiley.
320 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
materials. In this regard, we present a variety of synthesis techniques for producing diverse types of
nanostructured CuO with various morphologies, including 0D nanoparticles, 1D nanowires, 1D nano-
rods, 1D nanotubes, 2D nanodiscs, 2D nanoplates, 2D nanoleaves, 2D nanosheets, and 3D hierarchical
micro/nanostructures. The effects of various synthesis parameters on manipulating the nanoscale fea-
tures along with the associated growth mechanisms on how these unique morphologies developed are
also discussed. A better understanding of the fundamental physical and chemical properties of CuO
nanostructures is essential for their application as building blocks for future micro/nanoscale func-
tional devices. The up-to-date research progresses in the physical and chemical properties of CuO
nanostructures are systemically introduced and summarized. A wide range of promising applications
based on these CuO nanostructures are highlighted.
Although encouraging developments and fascinating achievements in CuO nanostructures have
been achieved as reviewed in this article, numerous problems and challenges still exist that need to
be addressed and solved.
(1) Challenges still exist in designing more versatile and reliable but simple synthetic schemes for
the large-scale production of tailored high-quality CuO nanostructures. The extremely simple
thermal oxidation of Cu substrates or Cu lms deposited on other substrates in ambient air
can be adopted for large-scale synthesis of CuO 1D nanostructures with good crystal quality.
However, the cracking problem during the synthetic process using the thermal oxidation
method needs to be solved to achieve more reliable CuO 1D nanostructure-based functional
devices, especially the devices that need to be used under liberating, rubbing, and affecting
environments.
(2) Although nanoscale CuO has relatively high carrier concentration and low resistivity, the hole
mobility and electrical conduction of CuO are poor, which limit more applications of this mate-
rial. Further efforts are required to improve these characteristics by developing new growth
process (e.g., doping) and techniques for device fabrication (e.g., use of various metal contacts).
Moreover, different device architectures are necessary to better explore the electrical properties
of this interesting material.
(3) For the gas sensor and optoelectronic device applications, CuO nanostructures are normally dis-
persed in solutions and deposited on the substrates to fabricate nanodevices. Improving the
yield and repeatability of these solution-based methods is an issue that needs to be addressed.
New techniques are required to directly grow CuO nanostructures into aligned arrays, onto ex-
ible substrates, and into self-assembled structures with functionality. These techniques are cru-
cial for nanosystem integration and design of functional nanodevices.
(4) Given its high theoretical capacity, high safety, environmental benignity, and lowcost, CuO
nanostructures as electrode materials for next-generation rechargeable LIBs with both high
energy and high power densities have attracted signicant attention. However, poor electronic
conductivities and large volume variation during the lithium insertion/release process limit
their practical applications. Optimization of the electrode nanostructure and interfacial chem-
istry in combination with novel 3D micro/nanostructured conductive frameworks along with
the robustness of system design are required.
(5) For CuO nanostructures in nEMs and devices, developing CuO and fuel (Al or Mg)-based novel
coreshell nanoenergetic arrays with fuel as the core and CuO as the shell is very promising
because the nanoenergetic arrays possess long-term storage stability caused by the stable
CuO shell. Another interesting direction is to combine the exothermic property of CuO/fuel with
the superhydrophobic property of CuO to achieve CuO-based superhydrophobic (waterproof)
nEMs and devices, which have promising underwater applications in both civilian and defense
areas.
(6) Although CuO nanostructures with various morphologies and sizes have been achieved by dif-
ferent synthetic methods, the growth mechanisms responsible for the formation of CuO nano-
structures are still not clearly elucidated. Precise detection of the development from precursors
to nal morphology formation, in situ characterization techniques for real-time observation,
and theory development to explain the formation processes are necessary.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 321
(7) Crystallographic data on the lattice structure and positions of Cu and O atoms together with the
bond lengths of CuO are available in the literature. However, the electronic properties (band
structure, bandgap energies, and effective masses) and surface structure of nanoscale CuO are
still not well known. More advanced theoretical investigations on the electronic properties
and surface structure of CuO nanostructures should be conducted.
(8) To further remarkably enhance the performance of CuO-based devices, controlled synthesis of
CuO nanostructures with optimized dimension, shape, and orientation, doping of CuO nano-
structures, functionalization of CuO nanostructures with certain functional groups, combination
with other functional nanomaterials to formhybrid structures, and surface coating of CuO nano-
structures are important.
(9) To date, studies on the CuO nanostructures mainly focus on the fabrication and characterization
of single or small-scale device structures. Although these proof-of-concept structures were
found to be effective for investigating the intrinsic properties of the devices, they are not pos-
sible for practical or commercial applications. Novel assembly and integration technologies that
can fabricate CuO nanostructures in a scalable approach while maintaining good uniformity in
performance among different devices are required.
(10) With regard to the CuO nanoparticles, the lack of low-cost, uniform-size, well-dispersed, and
high-quality nanoparticles still remains an obstacle. Therefore, attention should be focused
on the exploration of new synthetic routes to obtain desirable CuO nanoparticles for potential
applications, such as thermal, optical, magnetic, and biomedical areas.
(11) Lastly, CuO nanoparticles are proven to be highly toxic to humans. Thus, great caution should be
taken by researchers when handling these nanoparticles. Conducting a systematical evaluation
on how these CuO nanoparticles will affect human health and the environment is warranted.
Given the continuous development of the growth and fabrication techniques, more novel CuO
nanostructures with interesting properties and promising applications will be available in the near
future.
Acknowledgments
This work was supported by Hong Kong Research Grants Council (Project No. CityU 125412) and
NSAF (Grant No. U1330132).
References
[1] Tiwari JN, Tiwari RN, Kim KS. Zero-dimensional, one-dimensional, two-dimensional and three-dimensional
nanostructured materials for advanced electrochemical energy devices. Prog Mater Sci 2012;57:724803.
[2] Spencer MJS. Gas sensing applications of 1D-nanostructured zinc oxide: insights from density functional theory
calculations. Prog Mater Sci 2012;57:43786.
[3] Barth S, Hernandez-Ramirez F, Holmes JD, Romano-Rodriguez A. Synthesis and applications of one-dimensional
semiconductors. Prog Mater Sci 2010;55:563627.
[4] Dutta S, Chattopadhyay S, Sarkar A, Chakrabarti M, Sanyal D, Jana D. Role of defects in tailoring structural, electrical and
optical properties of ZnO. Prog Mater Sci 2009;54:89136.
[5] Comini E, Baratto C, Faglia G, Ferroni M, Vomiero A, Sberveglieri G. Quasi-one dimensional metal oxide semiconductors:
preparation, characterization and application as chemical sensors. Prog Mater Sci 2009;54:167.
[6] Chen Z, Jiao Z, Pan D, Li Z, Wu M, Shek C, et al. Recent advances in manganese oxide nanocrystals: fabrication,
characterization, and microstructure. Chem Rev 2012;112:383355.
[7] Chen X, Mao S. Titanium dioxide nanomaterials: synthesis, properties, modications, and applications. Chem Rev
2007;107:2891959.
[8] Garci M, Arias A, Hanson J, Rodriguez J. Nanostructured oxides in chemistry: characterization and properties. Chem Rev
2004;104:406374.
[9] Laurent S, Forge D, Port M, Roch A, Robic C, Elst L, et al. Magnetic iron oxide nanoparticles: synthesis, stabilization,
vectorization, physicochemical characterizations, and biological applications. Chem Rev 2008;108:2064110.
[10] Hochbaum AI, Yang P. Semiconductor nanowires for energy conversion. Chem Rev 2010;110:52746.
[11] Sui R, Charpentier P, Dmling A, Wang W, Wang K, Gonzlez-Gallardo S, et al. Synthesis of metal oxide nanostructures by
direct solgel chemistry in supercritical uids. Chem Rev 2012;112:305782.
[12] Xia Y, Yang P, Sun Y, Wu Y, Mayers B, Gates B, et al. One-dimensional nanostructures: synthesis, characterization, and
applications. Adv Mater 2003;15:35389.
[13] Vayssieres L. On the design of advanced metal oxide nanomaterials. Int J Nanotechnol 2004;1:141.
322 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
[14] Wang ZL. Zinc oxide nanostructures: growth, properties and applications. J Phys: Condens Matter 2004;16:R82958.
[15] Shankar K, Basham JI, Allam NK, Varghese OK, Mor GK, Feng X, et al. Recent advances in the use of TiO
2
nanotube and
nanowire arrays for oxidative. J Phys Chem C 2009;113:632759.
[16] Teja AS, Koh P-Y. Synthesis, properties, and applications of magnetic iron oxide nanoparticles. Prog Cryst Growth Charact
Mater 2009;55:2245.
[17] Pan J, Shen H, Mathur S. One-dimensional SnO
2
nanostructures: synthesis and applications. J Nanotechnol
2012;2012:112.
[18] Jun Y-W, Choi J-S, Cheon J. Shape control of semiconductor and metal oxide nanocrystals through nonhydrolytic colloidal
routes. Angew Chem Int Ed (English) 2006;45:341439.
[19] Zhou Z, Tian N, Li J, Broadwell I, Sun S. Nanomaterials of high surface energy with exceptional properties in catalysis and
energy storage. Chem Soc Rev 2011;40:416785.
[20] Park J, Joo J, Kwon SG, Jang Y, Hyeon T. Synthesis of monodisperse spherical nanocrystals. Angew Chem Int Ed
2007;46:463060.
[21] Zheng H, Ou JZ, Strano MS, Kaner RB, Mitchell A, Kalantar-zadeh K. Nanostructured tungsten oxide properties, synthesis,
and applications. Adv Funct Mater 2011;21:217596.
[22] Devan R, Patil R, Lin J, Ma Y. One-dimensional metal-oxide nanostructures: recent developments in synthesis,
characterization, and applications. Adv Funct Mater 2012;22:332670.
[23] Hu J, Chen M, Fang X, Wu L. Fabrication and application of inorganic hollow spheres. Chem Soc Rev 2011;40:547291.
[24] Li Y, Somorjai Ga. Nanoscale advances in catalysis and energy applications. Nano Lett 2010;10:228995.
[25] Singh DP, Ali N. Synthesis of TiO
2
and CuO nanotubes and nanowires. Sci Adv Mater 2010;2:295335.
[26] Nguyen T-D, Dinh C-T, Do T-O. A general procedure to synthesize highly crystalline metal oxide and mixed oxide
nanocrystals in aqueous medium and photocatalytic activity of metal/oxide nanohybrids. Nanoscale 2011;3:186173.
[27] Lou XWD, Archer La, Yang Z. Hollow micro-/nanostructures: synthesis and applications. Adv Mater 2008;20:39874019.
[28] Lignier P, Bellabarba R, Tooze RPR. Scalable strategies for the synthesis of well-dened copper metal and oxide
nanocrystals. Chem Soc Rev 2012;41:170820.
[29] Anandan S, Yang S. Emergent methods to synthesize and characterize semiconductor CuO nanoparticles with various
morphologiesan overview. J Exp Nanosci 2007;2:2356.
[30] Li Y, Yang XY, Feng Y, Yuan ZY, Su BL. One-dimensional metal oxide nanotubes, nanowires, nanoribbons, and nanorods:
synthesis, characterizations, properties and applications. Crit Rev Solid State Mater Sci 2012;37:174.
[31] Filipic G, Cvelbar U. Copper oxide nanowires: a review of growth. Nanotechnology 2012;23:194001.
[32] Liu Y, Chu Y, Zhuo Y, Li M, Li L, Dong L. Anion-controlled construction of CuO honeycombs and owerlike assemblies on
copper foils. Cryst Growth Des 2007;7:46770.
[33] Vaseem M, Umar A, Kim SH, Hahn Y-B. Low-temperature synthesis of ower-shaped CuO nanostructures by solution
process: formation mechanism and structural properties. J Phys Chem C 2008;112:572935.
[34] Zheng X, Xu C, Tomokiyo Y, Tanaka E, Yamada H, Soejima Y. Observation of charge stripes in cupric oxide. Phys Rev Lett
2000;85:51703.
[35] MacDonald aH. Copper oxides get charged up. Nature 2001;414:40910.
[36] Song M-K, Park S, Alamgir FM, Cho J, Liu M. Nanostructured electrodes for lithium-ion and lithium-air batteries: the latest
developments, challenges, and perspectives. Mater Sci Eng R Rep 2011;72:20352.
[37] Kislyuk VV, Dimitriev OP. Nanorods and nanotubes for solar cells. J Nanosci Nanotechnol 2008;8:13148.
[38] Choi KJ, Jang HW. One-dimensional oxide nanostructures as gas-sensing materials: review and issues. Sensors (Basel,
Switzerland) 2010;10:408399.
[39] Rahman MM, Saleh Ahammad aJ, Jin J-H, Ahn SJ, Lee J-J. A comprehensive review of glucose biosensors based on
nanostructured metal-oxides. Sensors (Basel, Switzerland) 2010;10:485586.
[40] Zhou LP, Wang BX, Peng XF, Du XZ, Yang YP. On the specic heat capacity of CuO nanouid. Adv Mech Eng 2010;2010:14.
[41] Wang SBB, Hsiao CHH, Chang SJJ, Lam KTT, Wen KHH, Hung SCC, et al. A CuO nanowire infrared photodetector. Sensor
Actuat A: Phys 2011;171:20711.
[42] Rossi C, Zhang K, Esteve D, Alphonse P, Tailhades P, Vahlas C. Nanoenergetic materials for MEMS: a review. J
Microelectromech Syst 2007;16:91931.
[43] Zhu YW, Yu T, Cheong FC, Xu XJ, Lim CT, Tan VBC, et al. Large-scale synthesis and eld emission properties of vertically
oriented CuO nanowire lms. Nanotechnology 2005;16:8892.
[44] Zhang X, Shi W, Zhu J, Kharistal D, Zhao W, Lalia B, et al. High-power and high-energy-density exible pseudocapacitor
electrodes made from porous CuO nanobelts and single-walled carbon nanotubes. ACS Nano 2011;5:20139.
[45] Ali I. New generation adsorbents for water treatment. Chem Rev 2012;112:507391.
[46] Yu X-Y, Xu R-X, Gao C, Luo T, Jia Y, Liu J-H, et al. Novel 3D hierarchical cotton-candy-like CuO: surfactant-free
solvothermal synthesis and application in As(III) removal. ACS Appl Mater Interfaces 2012;4:195462.
[47] Liu J, Jin J, Deng Z, Huang S-Z, Hu Z-Y, Wang L, et al. Tailoring CuO nanostructures for enhanced photocatalytic property. J
Colloid Interface Sci 2012;384:19.
[48] Kumar R, Diamant Y, Gedanken A. Sonochemical synthesis and characterization of nanometer-size transition metal
oxides from metal acetates. Chem Mater 2000;12:23015.
[49] Zhang X, Wang G, Liu X, Wu J, Li M, Gu J, et al. Different CuO nanostructures: synthesis, characterization, and applications
for glucose sensors. J Phys Chem C 2008;112:168459.
[50] Kumar R, Elgamiel R, Diamant Y, Gedanken A, Norwig J. Sonochemical preparation and characterization of nanocrystalline
copper oxide embedded in poly (vinyl alcohol) and its effect on crystal growth of copper oxide. Langmuir
2001;17:140610.
[51] Xu L, Sithambaram S, Zhang Y, Chen C-h, Jin L, Joesten R, et al. Novel urchin-like CuO synthesized by a facile reux method
with efcient olen epoxidation catalytic performance. Chem Mater 2009;21:12539.
[52] Zhou K, Wang R, Xu B, Li Y. Synthesis, characterization and catalytic properties of CuO nanocrystals with various shapes.
Nanotechnology 2006;17:393943.
[53] Zhou K, Li Y. Catalysis based on nanocrystals with well-dened facets. Angew Chem Int Ed (English) 2012;51:60213.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 323
[54] Rout L, Sen TK, Punniyamurthy T. Efcient CuO-nanoparticle-catalyzed CS cross-coupling of thiols with iodobenzene.
Angew Chem Int Ed (English) 2007;46:55836.
[55] Rout L, Jammi S, Punniyamurthy T. Novel CuO nanoparticle catalyzed CN cross coupling of amines with iodobenzene.
Org Lett 2007;9:33979.
[56] Mumm F, Helvoort ATJV, Sikorski P. Easy route to superhydrophobic copper-based wire-guided droplet microuidic
systems. ACS Nano 2009;3:264752.
[57] Singh DP, Neti NR, Sinha A, Srivastava ON. Growth of different nanostructures of Cu
2
O (nanothreads, nanowires, and
nanocubes) by simple electrolysis based oxidation of copper. J Phys Chem C 2007;111:163845.
[58] Ching W, Xu Y, Wong K. Ground-state and optical properties of Cu
2
O and CuO crystals. Phys Rev B 1989;40:768495.
[59] Ito T, Yamaguchi H, Okabe K, Masumi T. Single-crystal growth and characterization of Cu
2
O and CuO. J Mater Sci
1998;33:355566.
[60] Ghisen J. Electronic-structure of Cu
2
O and CuO. Phys Rev B 1988;38:1132230.
[61] Jayatissa AH, Guo K, Jayasuriya AC. Fabrication of cuprous and cupric oxide thin lms by heat treatment. Appl Surf Sci
2009;255:94749.
[62] Park JC, Kim J, Kwon H, Song H. Gram-scale synthesis of Cu
2
O nanocubes and subsequent oxidation to CuO hollow
nanostructures for lithium-ion battery anode materials. Adv Mater 2009;21:8037.
[63] Debbichi L, Marco de Lucas M, Pierson J, Kruger P. Vibrational properties of CuO and Cu
4
O
3
from rst-principles
calculations, and Raman and infrared spectroscopy. J Phys Chem C 2012;116:102327.
[64] Kuo CH, Huang MH. Morphologically controlled synthesis of Cu
2
O nanocrystals and their properties. Nano Today
2010;5:10616.
[65] Abdu Y, Musa AO. Copper (I) oxide (Cu
2
O) based solar cells a review. Beyero J Pure Appl Sci 2009;2:812.
[66] Korzhavyi P, Johansson B. Literature review on the properties of cuprous oxide Cu
2
O and the process of copper oxidation,
2011. http://www.skb.se/upload/publications/pdf/TR-11-08.
[67] Deng D, Kim MG, Lee JY, Cho J. Green energy storage materials: nanostructured TiO
2
and Sn-based anodes for lithium-ion
batteries. Energy Environ Sci 2009;2:81837.
[68] Su X, Wu QL, Zhan X, Wu J, Wei S, Guo Z. Advanced titania nanostructures and composites for lithium ion battery. J Mater
Sci 2012;47:251934.
[69] Zhuang Z, Peng Q, Li Y. Controlled synthesis of semiconductor nanostructures in the liquid phase. Chem Soc Rev
2011;40:5492513.
[70] Wang X, Li Y. Solution-based routes to transition-metal oxide one-dimensional nanostructures. Pure Appl Chem
2006;78:4564.
[71] Wang D, Xie T, Li Y. Nanocrystals: solution-based synthesis and applications as nanocatalysts. Nano Res 2009;2:3046.
[72] Cheng C, Fan HJ. Branched nanowires: synthesis and energy applications. Nano Today 2012;7:32743.
[73] Neupane MP, Kim YK, Park IS, Kim Ka, Lee MH, Bae TS. Temperature driven morphological changes of hydrothermally
prepared copper oxide nanoparticles. Surf Interface Anal 2009;41:25963.
[74] Chakraborty S, Das A, Begum MR, Dhara S, Tyagi AK, Garg AB, et al. Vibrational properties of CuO nanoparticles
synthesized by hydrothermal technique. AIP Conf Proc 2011;1349:8412.
[75] Sue K, Kawasaki S-i, Suzuki M, Hakuta Y, Hayashi H, Arai K, et al. Continuous hydrothermal synthesis of Fe
2
O
3
, NiO, and
CuO nanoparticles by superrapid heating using a T-type micro mixer at 673 K and 30 MPa. Chem Eng J 2011;166:94753.
[76] Outokesh M, Hosseinpour M, Ahmadi SJ, Mousavand T, Sadjadi S, Soltanian W. Hydrothermal synthesis of CuO
nanoparticles: study on effects of operational conditions on yield, purity, and size of the nanoparticles. Ind Eng Chem Res
2011;50:354054.
[77] Cao M, Hu C, Wang Y, Guo Y, Guo C, Wang E. A controllable synthetic route to Cu, Cu
2
O, and CuO nanotubes and nanorods.
Chem Eng (Cambridge) 2003;1:18845.
[78] Cheng G. Synthesis and characterisation of CuO nanorods via a hydrothermal method. Micro Nano Lett 2011;6:774.
[79] Gao X, Bao J, Pan G, Zhu H, Huang P, Wu F, et al. Preparation and electrochemical performance of polycrystalline and
single crystalline CuO nanorods as anode materials for Li ion battery. J Phys Chem B 2004;108:554751.
[80] Shrestha KM, Sorensen CM, Klabunde KJ. Synthesis of CuO nanorods, reduction of CuO into Cu nanorods, and diffuse
reectance measurements of CuO and Cu nanomaterials in the near Infrared region. J Phys Chem C 2010;114:1436876.
[81] Dar Ma, Kim YS, Kim WB, Sohn JM, Shin HS. Structural and magnetic properties of CuO nanoneedles synthesized by
hydrothermal method. Appl Surf Sci 2008;254:747781.
[82] Yang C, Su X, Xiao F, Jian J, Wang J. Gas sensing properties of CuO nanorods synthesized by a microwave-assisted
hydrothermal method. Sensor Actuat B: Chem 2011;158:299303.
[83] Srivastava R, Anu Prathap MU, Kore R. Morphologically controlled synthesis of copper oxides and their catalytic
applications in the synthesis of propargylamine and oxidative degradation of methylene blue. Colloids Surf Physicochem
Eng Aspects 2011;392:27182.
[84] Xu X, Yang H, Liu Y. Self-assembled structures of CuO primary crystals synthesized from Cu(CH
3
COO)
2
NaOH aqueous
systems. CrystEngComm 2012;14:528998.
[85] Yang M, He J. Fine tuning of the morphology of copper oxide nanostructures and their application in ambient degradation
of methylene blue. J Colloid Interface Sci 2011;355:1522.
[86] Yang M, He J, Hu X, Yan C, Cheng Z. CuO nanostructures as quartz crystal microbalance sensing layers for detection of
trace hydrogen cyanide gas. Environ Sci Technol 2011;45:608894.
[87] Guan X, Li L, Li G, Fu Z, Zheng J, Yan T. Hierarchical CuO hollow microspheres: controlled synthesis for enhanced lithium
storage performance. J Alloys Compd 2011;509:336774.
[88] Zhang Y, Wang S, Li X, Chen L, Qian Y, Zhang Z. CuO shuttle-like nanocrystals synthesized by oriented attachment. J Cryst
Growth 2006;291:196201.
[89] Chen H, Shin D-W, Lee J-H, Park S-M, Kwon K-W, Yoo J-B. Three-dimensional CuO nanobundles consisted of nanorods:
hydrothermal synthesis, characterization, and formation mechanism. J Nanosci Nanotechnol 2010;10:51218.
[90] Zhang Y, Or SW, Wang X, Cui T, Cui W, Zhang Y, et al. Hydrothermal synthesis of three-dimensional hierarchical CuO
buttery-like architectures. Eur J Inorg Chem 2009;2009:16873.
324 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
[91] Zou Y, Li Y, Zhang N, Liu X. Flower-like CuO synthesized by CTAB-assisted hydrothermal method. Bull Mater Sci
2011;34:96771.
[92] Cudennec Y, Lecerf A. The transformation of Cu(OH)
2
into CuO, revisited. Solid State Sci 2003;5:14714.
[93] Volanti DP, Orlandi MO, Andrs J, Longo E. Efcient microwave-assisted hydrothermal synthesis of CuO sea urchin-like
architectures via a mesoscale self-assembly. CrystEngComm 2010;12:1696.
[94] Volanti DP, Keyson D, Cavalcante LS, Simes aZ, Joya MR, Longo E, et al. Synthesis and characterization of CuO ower-
nanostructure processing by a domestic hydrothermal microwave. J Alloys Compd 2008;459:53742.
[95] Keyson D, Volanti DP, Cavalcante LS, Simes aZ, Varela Ja, Longo E. CuO urchin-nanostructures synthesized from a
domestic hydrothermal microwave method. Mater Res Bull 2008;43:7715.
[96] Zarate R, Hevia F, Fuentes S, Fuenzalida V, Zuniga A. Novel route to synthesize CuO nanoplatelets. J Solid State Chem
2007;18:14649.
[97] Wang J, Zhang W-D. Fabrication of CuO nanoplatelets for highly sensitive enzyme-free determination of glucose.
Electrochim Acta 2011;56:75106.
[98] Liu Q, Liu H, Liang Y, Xu Z, Yin G. Large-scale synthesis of single-crystalline CuO nanoplatelets by a hydrothermal process.
Mater Res Bull 2006;41:697702.
[99] Dar Ma, Nam SH, Kim YS, Kim WB. Synthesis, characterization, and electrochemical properties of self-assembled leaf-like
CuO nanostructures. J Solid State Electrochem 2010;14:171926.
[100] Zou YL, Li Y, Li JG, Xie WJ. Hydrothermal synthesis of momordica-like CuO nanostructures using egg white and their
characterisation. Chem Pap 2012;66:27883.
[101] Zhang H, Li S, Ma X, Yang D. Controllable growth of dendrite-like CuO nanostructures by ethylene glycol assisted
hydrothermal process. Mater Res Bull 2008;43:12916.
[102] Xiao H-M, Fu S-Y, Zhu L-P, Li Y-Q, Yang G. Controlled synthesis and characterization of CuO nanostructures through a
facile hydrothermal route in the presence of sodium citrate. Eur J Inorg Chem 2007;2007:196671.
[103] Jia W, Reitz E, Shimpi P, Rodriguez EG, Gao P-X, Lei Y. Spherical CuO synthesized by a simple hydrothermal reaction:
concentration-dependent size and its electrocatalytic application. Mater Res Bull 2009;44:16816.
[104] Hong J, Li J, Ni Y. Urchin-like CuO microspheres: synthesis, characterization, and properties. J Alloys Compd
2009;481:6105.
[105] Jiang Z, Niu Q, Deng W. Hydrothermal synthesis of CuO nanostructures with novel shapes. Nanoscience 2007;12:404.
[106] Teng F, Yao W, Zheng Y, Ma Y, Teng Y, Xu T, et al. Synthesis of ower-like CuO nanostructures as a sensitive sensor for
catalysis. Sensor Actuat B: Chem 2008;134:7618.
[107] Cheng Z, Xu J, Zhong H, Chu X, Song J. Hydrogen peroxide-assisted hydrothermal synthesis of hierarchical CuO ower-like
nanostructures. Mater Lett 2011;65:204750.
[108] Dar Ma, Ahsanulhaq Q, Kim YS, Sohn JM, Kim WB, Shin HS. Versatile synthesis of rectangular shaped nanobat-like CuO
nanostructures by hydrothermal method; structural properties and growth mechanism. Appl Surf Sci 2009;255:627984.
[109] Xia J, Li H, Luo Z, Wang K, Yin S, Yan Y. Ionic liquid-assisted hydrothermal synthesis of three-dimensional hierarchical
CuO peachstone-like architectures. Appl Surf Sci 2010;256:18717.
[110] Gao S, Yang S, Shu J, Zhang S, Li Z, Jiang K. Green fabrication of hierarchical CuO hollow micro/nanostructures and
enhanced performance as electrode materials for lithium-ion batteries. J Phys Chem C 2008;112:193248.
[111] Pan Q, Huang K, Ni S, Yang F, Lin S, He D. Synthesis of sheaf-like CuO from aqueous solution and their application in
lithium-ion batteries. J Alloys Compd 2009;484:3226.
[112] Abaker M, Umar A, Baskoutas S, Kim SH, Hwang SW. Structural and optical properties of CuO layered hexagonal discs
synthesized by a low-temperature hydrothermal process. J Phys D: Appl Phys 2011;44:155405.
[113] Zhang Z, Che H, Wang Y, Song L, Zhong Z, Su F. Preparation of hierarchical dandelion-like CuO microspheres with
enhanced catalytic performance for dimethyldichlorosilane synthesis. Catal Sci Technol 2012;2:195360.
[114] Wang C, Ye Y, Tao B, Geng B. Hydrothermal route to twinned-hemisphere-like CuO architectures with selective
adsorption performance. CrystEngComm 2012;14:3677.
[115] Zhang C, Chen J, Zeng Y, Rui X, Zhu J, Zhang W, et al. A facile approach toward transition metal oxide hierarchical
structures and their lithium storage properties. Nanoscale 2012;4:371824.
[116] Zhu J, Bi H, Wang Y, Wang X, Yang X, Lu L. CuO nanocrystals with controllable shapes grown from solution without any
surfactants. Mater Chem Phys 2008;109:348.
[117] Zhu J, Li D, Chen H, Yang X, Lu L, Wang X. Highly dispersed CuO nanoparticles prepared by a novel quick-precipitation
method. Mater Lett 2004;58:33247.
[118] Mahapatra O, Bhagat M, Gopalakrishnan C, Arunachalam KD. Ultrane dispersed CuO nanoparticles and their
antibacterial activity. J Exp Nanosci 2008;3:18593.
[119] Bera D, Qian L, Tseng T-K, Holloway PH. Quantum dots and their multimodal applications: areview. Materials
2010;3:2260345.
[120] Wang W, Liu Z, Liu Y, Xu C, Zheng C, Wang G. A simple wet-chemical synthesis and characterization of CuO nanorods.
Appl Phys A: Mater Sci Process 2003;76:41720.
[121] Lu C, Qi L, Yang J, Zhang D, Wu N, Ma J. Simple template-free solution route for the controlled synthesis of Cu(OH)
2
and
CuO nanostructures. J Phys Chem B 2004;108:1782531.
[122] Ethiraj AS, Kang DJ. Synthesis and characterization of CuO nanowires by a simple wet chemical method. Nanoscale Res
Lett 2012;7:70.
[123] Anandan S, Wen X, Yang S. Room temperature growth of CuO nanorod arrays on copper and their application as a cathode
in dye-sensitized solar cells. Mater Chem Phys 2005;93:3540.
[124] Li Y, Yang X-Y, Rooke J, Van Tendeloo G, Su B-L. Ultralong Cu(OH)
2
and CuO nanowire bundles: PEG200-directed crystal
growth for enhanced photocatalytic performance. J Colloid Interface Sci 2010;348:30312.
[125] Wang W, Wang L, Shi H, Liang Y. A room temperature chemical route for large scale synthesis of sub-15 nm ultralong CuO
nanowires with strong size effect and enhanced photocatalytic activity. CrystEngComm 2012;14:591422.
[126] Zaman S, Asif MH, Zainelabdin a, Amin G, Nur O, Willander M. CuO nanoowers as an electrochemical pH sensor and the
effect of pH on the growth. J Electroanal Chem 2011;662:4215.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 325
[127] Xiang JYY, Tu JPP, Zhang L, Zhou Y, Wang XLL, Shi SJJ. Self-assembled synthesis of hierarchical nanostructured CuO with
various morphologies and their application as anodes for lithium ion batteries. J Power Sources 2010;195:3139.
[128] Liu J, Huang X, Li Y, Sulieman KM, He X, Sun F. Self-assembled CuO monocrystalline nanoarchitectures with controlled
dimensionality and morphology. Cryst Growth Des 2006;6:16906.
[129] Zhang W, Wen X, Yang S. Controlled reactions on a copper surface: synthesis and characterization of nanostructured
copper compound lms. Inorg Chem 2003;42:500514.
[130] Wang W, Zhou Q, Fei X, He Y, Zhang P, Zhang G, et al. Synthesis of CuO nano- and micro-structures and their Raman
spectroscopic studies. CrystEngComm 2010;12:2232.
[131] XuR, XieT, ZhaoY, Li Y. Single-crystal metal nanoplatelets: cobalt, nickel, copper, andsilver. Cryst GrowthDes 2007;7:190411.
[132] Singh DP, Ojha AK, Srivastava ON. Synthesis of different Cu(OH)
2
and CuO (nanowires, rectangles, seed-, belt-, and
sheetlike) nanostructures by simple wet chemical route. J Phys Chem C 2009;113:340918.
[133] Zhang F, Zhu A, Luo Y, Tian Y, Yang J, Qin Y. CuO Nanosheets for sensitive and selective determination of H
2
S with high
recovery ability. J Phys Chem C 2010;114:192149.
[134] Yang Z, Xu J, Zhang W, Liu A, Tang S. Controlled synthesis of CuO nanostructures by a simple solution route. J Solid State
Chem 2007;180:13906.
[135] Li JP, Sun FQ. Synthesis of carambola-like CuO via a chemical bath method. Adv Mater Res 2010;148149:116770.
[136] Wang J, He S, Li Z, Jing X, Zhang M, Jiang Z. Self-assembled CuO nanoarchitectures and their catalytic activity in the
thermal decomposition of ammonium perchlorate. Colloid Polym Sci 2009;287:8538.
[137] Wang X, Xi G, Xiong S, Liu Y, Xi B, Yu W, et al. Solution-phase synthesis of single-crystal CuO nanoribbons and nanorings.
Cryst Growth Des 2007;7:9304.
[138] Xu H, Wang W, Zhu W, Zhou L, Ruan M. Hierarchical-oriented attachment: from one-dimensional Cu(OH)
2
nanowires to
two-dimensional CuO nanoleaves. Cryst Growth Des 2007;7:27204.
[139] Xia C, Xiaolan C, Ning W, Lin G. Hierarchical CuO nanochains: synthesis and their electrocatalytic determination of nitrite.
Anal Chim Acta 2011;691:437.
[140] Zheng L, Liu X. Solution-phase synthesis of CuO hierarchical nanosheets at near-neutral pH and near-room temperature.
Mater Lett 2007;61:22226.
[141] Yuan Z, Wang Y, Qian Y. A facile room-temperature route to ower-like CuO microspheres with greatly enhanced lithium
storage capability. RSC Adv 2012;2:86025.
[142] Sun S, Zhang X, Zhang J, Wang L, Song X, Yang Z. Surfactant-free CuO mesocrystals with controllable dimensions: green
ordered-aggregation-driven synthesis, formation mechanism and their photochemical performances. CrystEngComm
2013. http://dx.doi.org/10.1039/C2CE26216A.
[143] Huang J, Tang F, Gu F, Shi C, Zhai M. Flower-like CuO hierarchical nanostructures: synthesis, characterization, and
property. Optoelectron 2012. http://dx.doi.org/10.1007/s12200-012-0293-7.
[144] Jana S, Das S, Das NS, Chattopadhyay KK. CuO nanostructures on copper foil by a simple wet chemical route at room
temperature. Mater Res Bull 2010;45:6938.
[145] Zhang X, Guo Y-G, Liu W-M, Hao J-C. CuO three-dimensional owerlike nanostructures: controlled synthesis and
characterization. J Appl Phys 2008;103:114304.
[146] Liu J, Huang X, Li Y, Sulieman KM, He X, Sun F. Hierarchical nanostructures of cupric oxide on a copper substrate:
controllable morphology and wettability. J Mater Chem 2006;16:4427.
[147] Liu J, Huang X, Li Y, Li Z, Chi Q, Li G. Formation of hierarchical CuO microcabbages as stable bionicsuperhydrophobic
materials via a room-temperature solution-immersion process. Solid State Sci 2008;10:156876.
[148] Jia W, Reitz E, Sun H, Li B, Zhang H, Lei Y. From Cu
2
(OH)
3
Cl to nanostructured sisal-like Cu(OH)
2
and CuO: synthesis and
characterization. J Appl Phys 2009;105:064917.
[149] Zhao Y, Zhao J, Li Y, Ma D, Hou S, Li L, et al. Room temperature synthesis of 2D CuO nanoleaves in aqueous solution.
Nanotechnology 2011;22:115604.
[150] Zhang L, Yu JC, Xu A-W, Li Q, Kwong KW, Yu S-H. Peanut-shaped nanoribbon bundle superstructures of malachite and
copper oxide. J Cryst Growth 2004;266:54551.
[151] Huang H, Yu Q, Ye Y, Wang P, Zhang L, Gao M, et al. Thin copper oxide nanowires/carbon nanotubes interpenetrating
networks for lithium ion batteries. CrystEngComm 2012;14:7294300.
[152] Dey KK, Kumar A, Shanker R, Dhawan A, Wan M, Yadav RR, et al. Growth morphologies, phase formation, optical &
biological responses of nanostructures of CuO and their application as cooling uid in high energy density devices. RSC
Adv 2012;2:1387.
[153] Zhu C, Chen C, Hao L, Hu Y, Chen Z. Template-free synthesis of Cu
2
Cl(OH)
3
nanoribbons and use as sacricial template for
CuO nanoribbon. J Cryst Growth 2004;263:4739.
[154] Xu J, Xue D. Fabrication of malachite with a hierarchical sphere-like architecture. J Phys Chem B 2005;109:1715761.
[155] Jia W, Reitz E, Sun H, Zhang H, Lei Y. Synthesis and characterization of novel nanostructured shbone-like Cu(OH)
2
and
CuO from Cu
4
SO
4
(OH)
6
. Mater Lett 2009;63:51922.
[156] Wan M, Jin D, Feng R, Si L, Gao M, Yue L. Pillow-shaped porous CuO as anode material for lithium-ion batteries. Inorg
Chem Commun 2011;14:3841.
[157] Li J-Y, Xiong S, Xi B, Li X-G, Qian Y-T. Synthesis of CuO perpendicularly cross-bedded microstructure via a precursor-based
route. Cryst Growth Des 2009;9:410815.
[158] Wang L, Cheng W, Gong H, Wang C, Wang D, Tang K, et al. Facile synthesis of nanocrystalline-assembled bundle-like CuO
nanostructure with high rate capacities and enhanced cycling stability as an anode material for lithium-ion batteries. J
Mater Chem 2012;22:11297302.
[159] Fan Y, Liu R, Du W, Lu Q, Pang H, Gao F. Synthesis of copper(II) coordination polymers and conversion into CuO
nanostructures with good photocatalytic, antibacterial and lithium ion battery performances. J Mater Chem
2012;22:1260917.
[160] Liu X, Zhang J, Kang Y, Wu S, Wang S. Brochantite tabular microspindles and their conversion to wormlike CuO structures
for gas sensing. CrystEngComm 2012;14:6205.
326 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
[161] Wen X, Zhang W, Yang S. Synthesis of Cu(OH)
2
and CuO nanoribbon arrays on a copper surface. Langmuir
2003;19:5898903.
[162] Zhang W, Ding S, Yang Z, Liu A, Qian Y, Tang S, et al. Growth of novel nanostructured copper oxide (CuO) lms on copper
foil. J Cryst Growth 2006;291:47984.
[163] Hou H, Xie Y, Li Q. Large-scale synthesis of single-crystalline quasi-aligned submicrometer CuO ribbons. Cryst Growth Des
2005;5:2015.
[164] Chen X, Zhang N, Sun K. A vapor-phase corrosion strategy to hierarchically mesoporous nanosheet-assembled gearlike
pillar arrays for super-performance lithium storage. J Phys Chem C 2012;116:2122431.
[165] Chen X, Zhang N, Sun K. Facile fabrication of CuO mesoporous nanosheet cluster array electrodes with super lithium-
storage properties. J Mater Chem 2012;22:1363752.
[166] Yuan G-Q, Jiang H-F, Lin C, Liao S-J. Shape- and size-controlled electrochemical synthesis of cupric oxide nanocrystals. J
Cryst Growth 2007;303:4006.
[167] Toboonsung B, Singjai P. Formation of CuO nanorods and their bundles by an electrochemical dissolution and deposition
process. J Alloys Compd 2011;509:41327.
[168] Xu J, Yu K, Wu J, Shang D, Li L, Xu Ye, et al. Synthesis, eld emission and humidity sensing characteristics of honeycomb-
like CuO. J Phys D: Appl Phys 2009;42:075417.
[169] Lu L, Huang X. Room temperature electrochemical synthesis of CuO ower-like microspheres and their electrooxidative
activity towards hydrogen peroxide. Microchim Acta 2011;175:1517.
[170] Mukherjee N, Show B, Maji SK, Madhu U, Bhar SK, Mitra BC, et al. CuO nano-whiskers: electrodeposition, Raman analysis,
photoluminescence study and photocatalytic activity. Mater Lett 2011;65:324850.
[171] Xu M, Wang F, Ding B, Song X, Fang J. Electrochemical synthesis of leaf-like CuO mesocrystals and their lithium storage
properties. RSC Adv 2012;2:22403.
[172] Jiang X, Herricks T, Xia Y. CuO nanowires can be synthesized by heating copper substrates in air. Nano Lett
2002;2:13338.
[173] Chen JTJ, Zhang F, Wang J, Zhang GaG, Miao BBB, Fan XYX, et al. CuO nanowires synthesized by thermal oxidation route. J
Alloys Compd 2008;454:26873.
[174] Kumar A, Srivastava AK, Tiwari P, Nandedkar RV. The effect of growth parameters on the aspect ratio and number density
of CuO nanorods. J Phys: Condens Matter 2004;16:853143.
[175] Mema R, Yuan L, Du Q, Wang Y, Zhou G. Effect of surface stresses on CuO nanowire growth in the thermal oxidation of
copper. Chem Phys Lett 2011;512:8791.
[176] Yuan L, Zhou G. Enhanced CuO nanowire formation by thermal oxidation of roughened copper. J Electrochem Soc
2012;159:C205.
[177] Zhang K, Rossi C, Tenailleau C, Alphonse P, Chane-Ching J-Y. Synthesis of large-area and aligned copper oxide nanowires
from copper thin lm on silicon substrate. Nanotechnology 2007;18:275607.
[178] Zhang K, Rossi C, Tenailleau C, Conedera V. CuO nanowires grown from Cu lm heated under a N
2
/O
2
ow. J Nanosci
Nanotechnol 2009;9:141822.
[179] Zhang K, Yang Y, Pun E, Shen R. Local and CMOS-compatible synthesis of CuO nanowires on a suspended microheater on a
silicon substrate. Nanotechnology 2010;21:235602.
[180] Cheng S-L, M-f Chen. Fabrication, characterization, and kinetic study of vertical single-crystalline CuO nanowires on Si
substrates. Nanoscale Res Lett 2012;7:119.
[181] Tsai C-M, Chen G-D, Tseng T-C, Lee C-Y, Huang C-T, Tsai W-Y, et al. CuO nanowire synthesis catalyzed by a CoWP
nanolter. Acta Mater 2009;57:15706.
[182] Park YY-W, Seong N-JN, Jung H-JH, Chanda A, Yoon SS-G. Growth mechanism of the copper oxide nanowires from copper
thin lms deposited on CuO-buffered silicon substrate. J Electrochem Soc 2010;157:K11924.
[183] Hsueh HT, Chang SJ, Hung FY, Weng WY, Hsu CL, Hsueh TJ, et al. Ethanol gas sensor of crabwise CuO nanowires prepared
on glass substrate. J Electrochem Soc 2011;158:J106.
[184] Chang S, Yang T. Sensing performance of EGFET pH sensors with CuO nanowires fabricated on glass substrate. Int J
Electrochem Sci 2012;7:50207.
[185] Anandan S, Lee G, Wu J. Sonochemical synthesis of CuO nanostructures with different morphology. Ultrason Sonochem
2012;19:6826.
[186] Qiu M, Zhu L, Zhang T, Li H, Sun Y, Liu K. Ultrasound assisted quick synthesis of square-brick-like porous CuO and optical
properties. Mater Res Bull 2012;47:243741.
[187] Zou Y, Li Y, Guo Y, Zhou Q, An D. Ultrasound-assisted synthesis of CuO nanostructures templated by cotton bers. Mater
Res Bull 2012;47:313540.
[188] Xiao G, Gao P, Wang L, Chen Y, Wang Y, Zhang G. Ultrasonochemical-assisted synthesis of CuO nanorods with high
hydrogen storage ability. J Nanomater 2011;2011:16.
[189] Deng C, Hu H, Ge X, Han C, Zhao D, Shao G. One-pot sonochemical fabrication of hierarchical hollow CuO
submicrospheres. Ultrason Sonochem 2011;18:9327.
[190] Jung A, Cho S, Cho W, Lee K. Morphology-controlled synthesis of CuO nano- and microparticles using microwave
irradiation. Kor J Chem Eng 2012;29:2438.
[191] Xu L, Xu H, Wang F, Zhang F, Meng Z, Zhao W, et al. Microwave-assisted synthesis of ower-like and plate-like CuO
nanopowder and their photocatalytic activity for polluted lake water. J Kor Ceram Soc 2012;49:1514.
[192] H-c Song, S-h Park, Y-d Huh. Fabrication of hierarchical CuO microspheres. Bull Kor Chem Soc 2007;28:47780.
[193] Cai PJ, Shi M. Large scale synthesis of shuttle like CuO nanocrystals by microwave irradiation. Adv Mater Res
2010;92:11723.
[194] Wang H, Xu J-Z, Zhu J-J, Chen H-Y. Preparation of CuO nanoparticles by microwave irradiation. J Cryst Growth
2002;244:8894.
[195] Qiu G, Dharmarathna S, Zhang Y, Opembe N, Huang H, Suib SL. Facile microwave-assisted hydrothermal synthesis of CuO
nanomaterials and their catalytic and electrochemical properties. J Phys Chem C 2012;116:46877.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 327
[196] Zhu J, Qian X. From 2-D CuO nanosheets to 3-D hollow nanospheres: interface-assisted synthesis, surface photovoltage
properties and photocatalytic activity. J Solid State Chem 2010;183:16329.
[197] Li X, Wang Y, Lei Y, Gu Z. Highly sensitive H
2
S sensor based on template-synthesized CuO nanowires. RSC Adv
2012;2:2302.
[198] Hoa ND, Van Quy N, Jung H, Kim D, Kim H, Hong S-K. Synthesis of porous CuO nanowires and its application to hydrogen
detection. Sensor Actuat B: Chem 2010;146:26672.
[199] Zhu Y, Zhou G, Lin Y, Liu L. Controllable synthesis of well-aligned CuO nanotube arrays using porous alumina templates.
Cryst Res Technol 2012;47:65862.
[200] Wu H, Wei X, Shao M, Gu J, Qu M. Synthesis of copper oxide nanoparticles using carbon nanotubes as templates. Chem
Phys Lett 2002;364:1526.
[201] Mu C, He J. Conned conversion of CuS nanowires to CuO nanotubes by annealing-induced diffusion in nanochannels.
Nanoscale Res Lett 2011;6:150.
[202] Malandrino G, Finocchiaro S, Nigro R, Bongiorno C, Spinella C, Fragala I. Free-standing copper(II) oxide nanotube arrays
through an MOCVD template process. Chem Mater 2004;16:555961.
[203] Hsieh C-T, Chen J-M, Lin H-H, Shih H-C. Synthesis of well-ordered CuO nanobers by a self-catalytic growth mechanism.
Appl Phys Lett 2003;82:3316.
[204] Zhou H, Wong S. A facile and mild synthesis of 1-D ZnO, CuO, and a-Fe
2
O
3
nanostructures and nanostructured arrays.
ACSNano 2008;2:94458.
[205] Wang S, Huang Q, Wen X, Li X-y, Yang S. Thermal oxidation of Cu
2
S nanowires: a template method for the fabrication of
mesoscopic Cu
x
O (x = 1,2) wires. PCCP 2002;4:34259.
[206] Mallick P, Sahu S. Structure, microstructure and optical absorption analysis of CuO nanoparticles synthesized by solgel
route. Nanosci Nanotechnol 2012;2:714.
[207] Ahmad T, Chopra R, Ramanujachary KV, Loand SE, Ganguli aK. Canted antiferromagnetism in copper oxide nanoparticles
synthesized by the reverse-micellar route. Solid State Sci 2005;7:8915.
[208] Song X, Yu H, Sun S. Single-crystalline CuO nanobelts fabricated by a convenient route. J Colloid Interface Sci
2005;289:58891.
[209] Han D, Yang H, Zhu C, Wang F. Controlled synthesis of CuO nanoparticles using TritonX-100-based water-in-oil reverse
micelles. Powder Technol 2008;185:28690.
[210] Li C, Yin Y, Hou H, Fan N, Yuan F, Shi Y, et al. Preparation and characterization of Cu(OH)
2
and CuO nanowires by the
coupling route of microemulsion with homogenous precipitation. Solid State Commun 2010;150:5859.
[211] Hennemann J, Sauerwald T, Kohl C, Wagner T, Bognitzki M, Greiner A. Electrospun copper oxide nanobers for H
2
S
dosimetry. Phys Status Solidi (a) 2012;209:9116.
[212] Xiang H, Long Y, Yu X, Zhang X, Zhao N, Xu J. A novel and facile method to prepare porous hollow CuO and Cu nanobers
based on electrospinning. CrystEngComm 2011;13:485660.
[213] Sahay R, Suresh Kumar P, Aravindan V, Sundaramurthy J, Chui Ling W, Mhaisalkar SG, et al. High aspect ratio electrospun
CuO nanobers as anode material for lithium-ion batteries with superior cycleability. J Phys Chem C 2012;116:1808792.
[214] Morales J, Snchez L, Martn F, Ramos-Barrado JR, Snchez M. Nanostructured CuO thin lm electrodes prepared by spray
pyrolysis: a simple method for enhancing the electrochemical performance of CuO in lithium cells. Electrochim Acta
2004;49:458997.
[215] Jian G, Liu L, Zachariah M. Facile aerosol route to hollow CuO spheres and its superior performance as an oxidizer in
nanoenergetic gas generators. Adv Funct Mater 2013;23:13416.
[216] Chen U, Chueh Y, Lai S, Chou L, Shih H. Synthesis and characterization of self-catalyzed CuO nanorods on Cu/TaN/Si
assembly using vacuum-arc Cu deposition and vaporsolid reaction. J Vac Sci Technol B 2006;24:13942.
[217] Hai Z, Zhu C, Huang J, Liu H, Chen J. Controllable synthesis of CuO nanowires and Cu
2
O crystals with shape evolution via
gamma-irradiation. Inorg Chem 2010;49:72179.
[218] Zhang Q, Liu S-J, Yu S-H. Recent advances in oriented attachment growth and synthesis of functional materials: concept,
evidence, mechanism, and future. J Mater Chem 2009;19:191207.
[219] Zeng HC. Synthetic architecture of interior space for inorganic nanostructures. J Mater Chem 2006;16:649.
[220] Zhang Z, Sun H, Shao X, Li D, Yu H, Han M, et al. Three-dimensionally oriented aggregation of a few hundred nanoparticles
into monocrystalline architectures. Adv Mater 2005;17:427.
[221] Liu B, Zeng HC. Mesoscale organization of CuO nanoribbons: formation of dandelions. J Am Chem Soc
2004;126:81245.
[222] Li Y, Tan B, Wu Y. Ammonia-evaporation-induced synthetic method for metal (Cu, Zn, Cd, Ni) hydroxide/oxide
nanostructures. Chem Mater 2008;20:56776.
[223] Dong TY, Chen CN, Cheng HY, Chen CP, Jheng NY. Controlled morphologies of copper oxide single crystalline
nanostructures by wet chemistry and thermal decomposition processes. Inorg Chim Acta 2011;367:15865.
[224] Xu Y, Chen D, Jiao X, Xue K. CuO microowers composed of nanosheets: synthesis, characterization, and formation
mechanism. Mater Res Bull 2007;42:172331.
[225] Lin XZ, Liu P, Yu JM, Yang GW. Synthesis of CuO nanocrystals and sequential assembly of nanostructures with shape-
dependent optical absorption upon laser ablation in liquid. J Phys Chem C 2009;113:175437.
[226] Physics S, August R. The theory of Ostwald ripening. Journal of Statistical Physics 1985;38:23152.
[227] Zeng HC. Ostwald ripening: a synthetic approach for hollow nanomaterials. Curr Nanosci 2007;3:17781.
[228] Liu B, Zeng HC. Symmetric and asymmetric Ostwald ripening in the fabrication of homogeneous coreshell
semiconductors. Small (Weinheim an der Bergstrasse, Germany) 2005;1:56671.
[229] Zou G, Li H, Zhang D, Xiong K, Dong C, Qian Y. Well-aligned arrays of CuO nanoplatelets. J Phys Chem B 2006;110:16327.
[230] Qin Y, Zhang F, Chen Y, Zhou Y, Li J, Zhu A, et al. Hierarchically porous CuO hollow spheres fabricated via a one-pot
template-free method for high-performance gas sensors. J Phys Chem C 2012;116:119942000.
[231] Teo JJ, Chang Y, Zeng HC. Fabrications of hollow nanocubes of Cu
2
O and Cu via reductive self-assembly of CuO
nanocrystals. Langmuir 2006;22:736977.
328 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
[232] Zhang W, Wen X, Yang S, Berta Y, Wang ZL. Single-crystalline scroll-type nanotube arrays of copper hydroxide
synthesized at room temperature. Adv Mater 2003;15:8225.
[233] Wen X, Zhang W, Yang S, Dai ZR, Wang ZL. Solution phase synthesis of Cu(OH)
2
nanoribbons by coordination self-
assembly using Cu
2
S nanowires as precursors. Nano Lett 2002;2:1397401.
[234] Kaur M, Muthe KP, Despande SK, Choudhury S, Singh JB, Verma N, et al. Growth and branching of CuO nanowires by
thermal oxidation of copper. J Cryst Growth 2006;289:6705.
[235] Shao P, Deng S, Chen J, Xu N. Large-scale fabrication of ordered arrays of microcontainers and the restraint effect on
growth of CuO nanowires. Nanoscale Res Lett 2011;6:86.
[236] Gonalves AMB, Campos LC, Ferlauto aS, Lacerda RG, Goncalves aMB. On the growth and electrical characterization of
CuO nanowires by thermal oxidation. J Appl Phys 2009;106:034303.
[237] Zhong ML, Zeng DC, Liu ZW, Yu HY, Zhong XC, Qiu WQ. Synthesis, growth mechanism and gas-sensing properties of large-
scale CuO nanowires. Acta Mater 2010;58:592632.
[238] Li X, Zhang J, Yuan Y, Liao L, Pan C. Effect of electric eld on CuO nanoneedle growth during thermal oxidation and its
growth mechanism. J Appl Phys 2010;108:024308.
[239] Yuan L, Wang Y, Mema R, Zhou G. Driving force and growth mechanism for spontaneous oxide nanowire formation
during the thermal oxidation of metals. Acta Mater 2011;59:2491500.
[240] Zhu Y, Mimura K, Isshiki M. Inuence of oxide grain morphology on formation of the CuO scale during oxidation of copper
at 6001000 C. Corros Sci 2005;47:53744.
[241] Liang J, Kishi N, Soga T, Jimbo T. The synthesis of highly aligned cupric oxide nanowires by heating copper foil. J
Nanomater 2011;2011:18.
[242] Chang FM, Cheng SL, Hong SJ, Sheng YJ, Tsao HK. Superhydrophilicity to superhydrophobicity transition of CuO nanowire
lms. Appl Phys Lett 2010;96:114101.
[243] Jia W, Guo M, Zheng Z, Yu T, Wang Y, Rodriguez EG, et al. Vertically aligned CuO nanowires based electrode for
amperometric detection of hydrogen peroxide. Electroanalysis 2008;20:21537.
[244] Raksa P, Gardchareon A, Chairuangsri T, Mangkorntong P, Mangkorntong N, Choopun S. Ethanol sensing properties of CuO
nanowires prepared by an oxidation reaction. Ceram Int 2009;35:64952.
[245] Liang J, Kishi N, Soga T, Jimbo T. Cross-sectional characterization of cupric oxide nanowires grown by thermal oxidation of
copper foils. Appl Surf Sci 2010;257:626.
[246] Hansen B, Chan H, Lu J, Lu G, Chen J. Short-circuit diffusion growth of long bi-crystal CuO nanowires. Chem Phys Lett
2011;504:415.
[247] Kim DK, Bae JH, Kim HJ. Thermite reaction between CuO nanowires and Al for the crystallization of a-Si. Trans Electr
Electron Mater 2010;11:2347.
[248] Wang R-C, Li C-H. Improved morphologies and enhanced eld emissions of CuO nanoneedle arrays by heating ZnO coated
copper foils. CrystGrowthDes 2009;9:222934.
[249] Mumm F, Sikorski P. Oxidative fabrication of patterned, large, non-aking CuO nanowire arrays. Nanotechnology
2011;22:105605.
[250] Mumm F, Sikorski P. Oxidative fabrication of patterned, large, defect-free CuO nanowire arrays. Nanotechnology
2011;22:105605 [supporting material].
[251] Amin G. ZnO and CuO nanostructures: low temperature growth, characterization, their optoelectronic and sensing
applications. 1st ed. LiU-Tryck: Norrkping (Sweden); 2012 [SE-601 74].
[252] Asbrink S, Norrby LJ. A renement of the crystal structure of copper (II) oxide with a discussion of some exceptional
e.s.d.s. Acta Crystallogr Sect A: Found Crystallogr 1970;26:815.
[253] Forsyth J, Hull S. The effect of hydrostatic pressure on the ambient temperature structure of CuO. J Phys: Condens Matter
1991;3:525761.
[254] Yang B, Thurston T, Tranquada J, Shirane G. Magnetic neutron scattering study of single-crystal cupric oxide. Phys Rev B
1989;39:43439.
[255] Poizot P, Hung C-J, Nikiforov MP, Bohannan EW, Switzer Ja. An electrochemical method for CuO thin lm deposition from
aqueous solution. Electrochem Solid-State Lett 2003;6:C215.
[256] Meyer B, Polity A, Reppin D, Becker M, Hering P, Klar P, et al. Binary copper oxide semiconductors: from materials towards
devices. Phys Status Solidi (b) 2012;249:1487509.
[257] Bourne L, Yu P, Zettl A, Cohen M. High-pressure electrical conductivity measurements in the copper oxides. Phys Rev B
1989;40:109736.
[258] Azam A, Ahmed AS, Oves M, Khan MS, Memic A. Size-dependent antimicrobial properties of CuO nanoparticles against
Gram-positive and -negative bacterial strains. Int J Nanomed 2012;7:352735.
[259] Vidyasagar C, Arthoba Naik Y, Venkatesha T, Viswanatha R. Solid-state synthesis and effect of temperature on optical
properties of CuO nanoparticles. Nano-Micro Lett 2012;4:737.
[260] Siemons W, Koster G, Blank DH, Hammond RH, Geballe TH, Beasley MR. Tetragonal CuO: end member of the 3d transition
metal monoxides. Phys Rev B 2009;79:195122.
[261] Himmetoglu B, Wentzcovitch R, Cococcioni M. First-principles study of electronic and structural properties of CuO. Phys
Rev B 2011;84:18.
[262] Palkar V, Ayyub P, Chattopadhyay S, Multani M. Size-induced structural transitions in the CuO and CeO systems. Phys
Rev B: Condens Matter 1996;53:216770.
[263] Wu D, Zhang Q, Tao M. LSDA+U study of cupric oxide: electronic structure and native point defects. Phys Rev B
2006;73:235206.
[264] Grioni M, Czyzyk M, Groot FD, Fuggle J, Watts B. Unoccupied electronic states of CuO: an oxygen 1s X-ray-absorption
spectroscopy investigation. Phys Rev B 1989;39:488690.
[265] Anisimov VI, Zaanen J, Andersen OK. Band theory and Mott insulators: Hubbard U instead of Stoner I. Phys Rev B
1991;44:94354.
[266] Anisimov V, Aryasetiawan F, Lichtenstein A. First-principles calculations of the electronic structure and spectra of
strongly correlated systems: the LDA+ U method. J Phys: Condens Matter 1997;9:767808.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 329
[267] Hu J, Li D, Lu JG, Wu R. Effects on electronic properties of molecule adsorption on CuO surfaces and nanowires. J Phys
Chem C 2010;114:171206.
[268] Zaanen J, Sawatzky G, Allen J. Band gaps and electronic structure of transition-metal compounds. Phys Rev Lett
1985;55:41821.
[269] Tahir D, Tougaard S. Electronic and optical properties of Cu, CuO and Cu
2
O studied by electron spectroscopy. J Phys:
Condens Matter 2012;24:175002.
[270] Benndorf C, Caus H, Egert B, Seidel H, Thieme F. Identication of Cu (I) and Cu (II) oxides by electron spectroscopic
methods: AES, ELS and UPS investigations. J Electron Spectrosc Relat Phenom 1980;19:7790.
[271] Koffyberg FP. A photoelectrochemical determination of the position of the conduction and valence band edges of p-type
CuO. J Appl Phys 1982;53:1173.
[272] Nakaoka K, Ueyama J, Ogura K. Photoelectrochemical behavior of electrodeposited CuO and Cu
2
O thin lms on
conducting substrates. J Electrochem Soc 2004;151:C661.
[273] Chiang C-Y, Aroh K, Franson N, Satsangi VR, Dass S, Ehrman S. Copper oxide nanoparticle made by ame spray pyrolysis
for photoelectrochemical water splitting Part II. Photoelectrochemical study. Int J Hydrogen Energy 2011;36:1551926.
[274] Chiang CY, Chang MH, Liu HS, Tai CY, Ehrman S. Process intensication in the production of photocatalysts for solar
hydrogen generation. Ind Eng Chem Res 2012;51:520715.
[275] Ruhle S, Shalom M, Zaban A. Quantum-dot-sensitized solar cells. ChemPhysChem 2010;11:2290304.
[276] Alonso M, Marcus I, Garriga M, Goi a, Jedrzejewski J, Balberg I. Evidence of quantum connement effects on interband
optical transitions in Si nanocrystals. Phys Rev B 2010;82:18.
[277] Buhro W, Colvin V. Semiconductor nanocrystals shape matters. Nat Mater 2003;2:1389.
[278] Rehman S, Mumtaz a, Hasanain SK. Size effects on the magnetic and optical properties of CuO nanoparticles. J Nanopart
Res 2010;13:2497507.
[279] Ogwu A, Darma T, Bouquerel E. Electrical resistivity of copper oxide thin lms prepared by reactive magnetron
sputtering. J Achiev Mater Manufact Eng 2007;24:1727.
[280] Borgohain K, Mahamuni S. Formation of single-phase CuO quantum particles. J Mater Res 2002;17:12203.
[281] Marabelli F, Parravicini G, Salghetti-Drioli F. Optical gap of CuO. Phys Rev B 1995;52:14336.
[282] Al-Gaashani R, Radiman S, Tabet N, Razak Daud a. Synthesis and optical properties of CuO nanostructures obtained via a
novel thermal decomposition method. J Alloys Compd 2011;509:87619.
[283] Jin C, Baek K, Park S, Kim HW, Lee WI, Lee C. Inuence of coating and thermal annealing on the structure and
luminescence properties of CuO nanorods. Solid State Commun 2010;150:18127.
[284] Lin H-H. Characterizing well-ordered CuO nanobrils synthesized through gassolid reactions. J Appl Phys
2004;95:588995.
[285] Aslani A, Oroojpour V. CO gas sensing of CuO nanostructures, synthesized by an assisted solvothermal wet chemical
route. Physica B 2011;406:1449.
[286] Chang S-S, Lee H-J, Park HJ. Photoluminescence properties of spark-processed CuO. Ceram Int 2005;31:4115.
[287] Vila M, Daz-Guerra C, Piqueras J. Optical and magnetic properties of CuO nanowires grown by thermal oxidation. J Phys D
Appl Phys 2010;43:1354038.
[288] Huang C-Y, Chatterjee a, Liu SB, Wu SY, Cheng CL. Photoluminescence properties of a single tapered CuO nanowire. Appl
Surf Sci 2010;256:368892.
[289] Xu JF, Ji W, Shen ZX, Li WS, Tang SH, Ye XR, et al. Raman spectra of CuO nanocrystals. J Raman Spectrosc 1999;30:4135.
[290] Singh I, Bedi RK. Studies and correlation among the structural, electrical and gas response properties of aerosol spray
deposited self assembled nanocrystalline CuO. Appl Surf Sci 2011;257:75929.
[291] Banerjee aN, Kundoo S, Chattopadhyay KK. Synthesis and characterization of p-type transparent conducting CuAlO
2
thin
lm by DC sputtering. Thin Solid Films 2003;440:510.
[292] Ohya Y, Ito S, Ban T, Takahashi Y. Preparation of CuO thin lms and their electrical conductivity. Key Eng Mater
2000;181182:1136.
[293] Hansen BJ, Kouklin N, Lu G, Lin I-K, Chen J, Zhang X. Transport, analyte detection, and opto-electronic response of p-type
CuO nanowires. J Phys Chem C 2010;114:24407.
[294] Hsieh JH, Kuo PW, Peng KC, Liu SJ, Hsueh JD, Chang SC. Opto-electronic properties of sputter-deposited Cu
2
O lms treated
with rapid thermal annealing. Thin Solid Films 2008;516:544953.
[295] Serin T, Yildiz a, Horzum Sahin S, Serin N. Extraction of important electrical parameters of CuO. Phys B: Condens Matter
2011;406:5758.
[296] Serin T, Yildiz a, Sahin SH, Serin N. Multiphonon hopping of carriers in CuO thin lms. Physica B 2011;406:35515.
[297] Nair MTS, Guerrero L, Arenas OL, Nair PK. Chemically deposited copper oxide thin lms: structural, optical and electrical
characteristics. Appl Surf Sci 1999;150:14351.
[298] Jundale D, Joshi P, Sen S, Patil VB. Nanocrystalline CuO thin lms: synthesis, microstructural and optoelectronic
properties. J Mater Sci: Mater Electron 2012;23:14929.
[299] Boschloo G, Edvinsson T, Hagfeldt A. Dye-sensitized nanostructured ZnO electrodes for solar cell applications. In: Soga T,
editor. Nanostructured materials for solar energy conversion. Nagoya: Elsevier Science; 2007. p. 231 [chapter 8].
[300] Wu H, Lin D, Pan W. Fabrication, assembly, and electrical characterization of CuO nanobers. Appl Phys Lett
2006;89:133125.
[301] Shao P, Deng S, Chen J, Chen J, Xu N. Study of eld emission, electrical transport, and their correlation of individual single
CuO nanowires. J Appl Phys 2011;109:023710.
[302] Liao L, Zhang Z, Yan B, Zheng Z, Bao QL, Wu T, et al. Multifunctional CuO nanowire devices: p-type eld effect transistors
and CO gas sensors. Nanotechnology 2009;20:085203.
[303] Li D, Hu J, Wu R, Lu JG. Conductometric chemical sensor based on individual CuO nanowires. Nanotechnology
2010;21:485502.
[304] Chauhan D, Satsangi V, Dass S, Shrivastav R. Preparation and characterization of nanostructured CuO thin lms for
photoelectrochemical splitting of water. Bull Mater Sci 2006;29:70916.
330 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
[305] Kamat PV, Tvrdy K, Baker DR, Radich JG. Beyond photovoltaics: semiconductor nanoarchitectures for liquid-junction solar
cells. Chem Rev 2010;110:666488.
[306] Chen X, Shen S, Guo L, Mao SS. Semiconductor-based photocatalytic hydrogen generation. Chem Rev 2010;110:650370.
[307] Xu Y, Schoonen M. The absolute energy positions of conduction and valence bands of selected semiconducting minerals.
Am Mineral 2000;85:54356.
[308] Gonzalez-Valls I, Lira-Cantu M. Vertically-aligned nanostructures of ZnO for excitonic solar cells: a review. Energy
Environ Sci 2009;2:1934.
[309] Chiang CY, Shin Y, Ehrman S. Li doped CuO lm electrodes for photoelectrochemical cells. J Electrochem Soc
2011;159:B22731.
[310] Chiang CY, Shin Y, Aroh K, Ehrman S. Copper oxide photocathodes prepared by a solution based process. Int J Hydrogen
Energy 2012;37:82329.
[311] Yang B, Tranquada J, Shirane G. Neutron scattering studies of the magnetic structure of cupric oxide. Phys Rev B
1988;38:1748.
[312] Brown P, Chattopadhyay T, Forsyth J, Nunez V, Tasset F. Antiferromagnetism i n CuO studied by neutron polarimetry. J
Phys: Condens Matter 1991;4281:42817.
[313] Azzoni C, Paleari A, Parravicini G. On the low-temperature magnetic properties of CuO single crystals. J Phys: Condens
Matter 1992;4:135966.
[314] Punnoose a, Magnone H, Seehra M, Bonevich J. Bulk to nanoscale magnetism and exchange bias in CuO nanoparticles.
Phys Rev B 2001;64:18.
[315] Rao GN, Yao YDY, Chen JW, Narsinga Rao G. Evolution of size, morphology, and magnetic properties of CuO nanoparticles
by thermal annealing. J Appl Phys 2009;105:093901.
[316] Seehra M, Punnoose A. Particle size dependence of exchange-bias and coercivity in CuO nanoparticles. Solid State
Commun 2003;128:299302.
[317] Zheng X, Mori T, Nishiyama K, Higemoto W, Xu C. Dramatic suppression of antiferromagnetic coupling in nanoparticle
CuO. Solid State Commun 2004;132:4936.
[318] Xiao H-M, Zhu L-P, Liu X-M, Fu S-Y. Anomalous ferromagnetic behavior of CuO nanorods synthesized via hydrothermal
method. Solid State Commun 2007;141:4315.
[319] Dietl T. Zener Model description of ferromagnetism in zinc-blende magnetic semiconductors. Science 2000;287:101922.
[320] Xu Q, Schmidt H, Zhou S, Potzger K, Helm M, Hochmuth H, et al. Room temperature ferromagnetism in ZnO lms due to
defects. Appl Phys Lett 2008;92:082508.
[321] Thakur P, Cezar J, Brookes N, Choudhary R, Prakash R, Phase D, et al. Direct observation of oxygen induced room
temperature ferromagnetism in MoO
2
thin lms by X-ray magnetic circular dichroism characterizations. Appl Phys Lett
2009;94:062501.
[322] Gao D, Li J, Li Z, Zhang Z, Zhang J, Shi H, et al. Defect-mediated magnetism in pure CaO nanopowders. J Phys Chem C
2010;114:117037.
[323] Shang D, Yu K, Zhang Y, Xu J, Wu J, Xu Ye, et al. Magnetic and eld emission properties of straw-like CuO nanostructures.
Appl Surf Sci 2009;255:40936.
[324] Gao D, Yang G, Li J, Zhang J, Zhang J, Xue D. Room-temperature ferromagnetism of owerlike CuO nanostructures. J Phys
Chem C 2010;114:1834751.
[325] Gao D, Zhang J, Zhu J, Qi J, Zhang Z, Sui W, et al. Vacancy-mediated magnetism in pure copper oxide nanoparticles.
Nanoscale Res Lett 2010;5:76972.
[326] Li X, Wang C. Engineering nanostructured anodes via electrostatic spray deposition for high performance lithium ion
battery application. J Mater Chem A 2013;1:16582.
[327] Venkatachalam S, Zhu H, Masarapu C, Hung K, Liu Z, Suenage K, et al. In-situ formation of sandwiched structures of
nanotube/Cu
x
O
y
/Cu composites for lithium battery applications. ACS Nano 2009;3:217784.
[328] Poizot P, Laruelle S, Grugeon S, Dupont L, Tarascon JM. Nano-sized transition-metal oxides as negative-electrode
materials for lithium-ion batteries. Nature 2000;407:4969.
[329] Zheng S-F, Hu J-S, Zhong L-S, Song W-G, Wan L-J, Guo Y-G. Introducing dual functional CNT networks into CuO
nanomicrospheres toward superior electrode materials for lithium-ion batteries. Chem Mater 2008;20:361722.
[330] Zhang X, Zhang D, Ni X, Song J, Zheng H. Synthesis and electrochemical properties of different sizes of the CuO particles. J
Nanopart Res 2007;10:83944.
[331] Chen LB, Lu N, Xu CM, Yu HC, Wang TH. Electrochemical performance of polycrystalline CuO nanowires as anode material
for Li ion batteries. Electrochim Acta 2009;54:4198201.
[332] Ju J-H, Ryu K-S. Synthesis and performance of CuO with complex hollow structure as anode material for lithium
secondary batteries. J Electrochem Soc 2011;158:A814.
[333] Hu Y, Huang X, Wang K, Liu J, Jiang J, Ding R, et al. Kirkendall-effect-based growth of dendrite-shaped CuO hollow micro/
nanostructures for lithium-ion battery anodes. J Solid State Chem 2010;183:6627.
[334] Wang SQ, Zhang JY, Chen CH. Dandelion-like hollow microspheres of CuO as anode material for lithium-ion batteries.
Scripta Mater 2007;57:33740.
[335] Kong M, Zhang W, Yang Z, Weng S, Chen Z. Facile synthesis of CuO hollow nanospheres assembled by nanoparticles and
their electrochemical performance. Appl Surf Sci 2011;258:131721.
[336] Wang X, Tang D-M, Li H, Yi W, Zhai T, Bando Y, et al. Revealing the conversion mechanism of CuO nanowires during
lithiationdelithiation by in situ transmission electron microscopy. Chem Commun 2012;48:48124.
[337] Ji L, Lin Z, Alcoutlabi M, Zhang X. Recent developments in nanostructured anode materials for rechargeable lithium-ion
batteries. Energy Environ Sci 2011;4:2682.
[338] Dbart a, Dupont L, Poizot P, Leriche J-B, Tarascon JM. A transmission electron microscopy study of the reactivity
mechanism of tailor-made CuO particles toward lithium. J Electrochem Soc 2001;148:A126674.
[339] Grugeon S, Laruelle S, Herrera-Urbina R, Dupont L, Poizot P, Tarascon J-M. Particle size effects on the electrochemical
performance of copper oxides toward lithium. J Electrochem Soc 2001;148:A28592.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 331
[340] Wang B, Wu XL, Shu CY, Guo YG, Wang CR. Synthesis of CuO/graphene nanocomposite as a high-performance anode
material for lithium-ion batteries. J Mater Chem 2010;20:106614.
[341] Ko S, Lee J-I, Yang HS, Park S, Jeong U. Mesoporous CuO particles threaded with CNTs for high-performance lithium-ion
battery anodes. Adv Mater 2012;24:44516.
[342] Xiang JY, Tu JP, Yuan YF, Wang XL, Huang XH, Zeng ZY. Electrochemical investigation on nanoower-like CuO/Ni
composite lm as anode for lithium ion batteries. Electrochim Acta 2009;54:11605.
[343] Ke F-S, Huang L, Wei G-Z, Xue L-J, Li J-T, Zhang B, et al. One-step fabrication of CuO nanoribbons array electrode and its
excellent lithium storage performance. Electrochim Acta 2009;54:58259.
[344] Wang F, Tao W, Zhao M, Xu M, Yang S, Sun Z, et al. Controlled synthesis of uniform ultrane CuO nanowires as anode
material for lithium-ion batteries. J Alloys Compd 2011;509:9798803.
[345] Xiang JY, Tu JP, Huang XH, Yang YZ. A comparison of anodically grown CuO nanotube lm and Cu
2
O lm as anodes for
lithium ion batteries. J Solid State Electrochem 2007;12:9415.
[346] Pan Q, Jin H, Wang H, Yin G. Flower-like CuO lm-electrode for lithium ion batteries and the effect of surface morphology
on electrochemical performance. Electrochim Acta 2007;53:9516.
[347] Wang H, Pan Q, Zhao J, Yin G, Zuo P. Fabrication of CuO lm with network-like architectures through solution-immersion
and their application in lithium ion batteries. J Power Sources 2007;167:20611.
[348] Xu M, Wang F, Zhao M, Yang S, Sun Z, Song X. Synthesis of copper oxide nanostructures via a composite-hydroxide-
mediated approach: morphology control and the electrochemical performances as anode material for lithium ion
batteries. Physica E 2011;44:50610.
[349] Xiang JY, Tu JP, Qiao YQ, Wang XL, Zhong J, Zhang D, et al. Electrochemical impedance analysis of a hierarchical CuO
electrode composed of self-assembled nanoplates. J Phys Chem C 2011;115:250513.
[350] Seo S-D, Jin Y-H, Lee S-H, Shim H-W, Kim D-W. Low-temperature synthesis of CuO-interlaced nanodiscs for lithium ion
battery electrodes. Nanoscale Res Lett 2011;6:397.
[351] Zhang W, Li M, Wang Q, Chen G, Kong M, Yang Z, et al. Hierarchical self-assembly of microscale cog-like superstructures
for enhanced performance in lithium-ion batteries. Adv Funct Mater 2011;21:351623.
[352] Xiang JY, Tu JP, Zhang L, Zhou Y, Wang XL, Shi SJ. Simple synthesis of surface-modied hierarchical copper oxide spheres
with needle-like morphology as anode for lithium ion batteries. Electrochim Acta 2010;55:18204.
[353] Chen X, Zhang N, Sun K. Facile fabrication of CuO 1D pine-needle-like arrays for super-rate lithium storage. J Mater Chem
2012;22:150804.
[354] Wang L, Gong H, Wang C, Wang D, Tang K, Qian Y. Facile synthesis of novel tunable highly porous CuO nanorods for high
rate lithium battery anodes with realized long cycle life and high reversible capacity. Nanoscale 2012;4:68505.
[355] Mai YJ, Wang XL, Xiang JY, Qiao YQ, Zhang D, Gu CD, et al. CuO/graphene composite as anode materials for lithium-ion
batteries. Electrochim Acta 2011;56:230611.
[356] Seo S-D, Lee D-H, Kim J-C, Lee G-H, Kim D-W. Room-temperature synthesis of CuO/graphene nanocomposite electrodes
for high lithium storage capacity. Ceram Int 2013;39:174955.
[357] Wang H, Pan Q, Zhao J, Chen W. Fabrication of CuO/C lms with sisal-like hierarchical microstructures and its application
in lithium ion batteries. J Alloys Compd 2009;476:40813.
[358] Huang XH, Wang CB, Zhang SY, Zhou F. CuO/C microspheres as anode materials for lithium ion batteries. Electrochim Acta
2011;56:67526.
[359] Lu LQ, Wang Y. Sheet-like and fusiform CuO nanostructures grown on graphene by rapid microwave heating for high Li-
ion storage capacities. J Mater Chem 2011;21:1791621.
[360] Yang M, Gao Q. Copper oxide and ordered mesoporous carbon composite with high performance using as anode material
for lithium-ion battery. Microporous Mesoporous Mater 2011;143:2305.
[361] Ko J, Kim S, Hong J, Ryu J, Kang K, Park C. Synthesis of graphene-wrapped CuO hybrid materials by CO2 mineralization.
Green Chem 2012;14:23914.
[362] Liu X, Li Z, Zhang Q, Li F, Kong T. Preparation of CuO/C coreshell nanowires and its application in lithium ion batteries.
Mater Lett 2012;80:379.
[363] Wang G, Zhang L, Zhang J. A review of electrode materials for electrochemical supercapacitors. Chem Soc Rev
2012;41:797828.
[364] Zhang L, Zhao X. Carbon-based materials as supercapacitor electrodes. Chem Soc Rev 2009;38:252031.
[365] Zhi M, Xiang C, Li J, Li M, Wu N. Nanostructured carbonmetal oxide composite electrodes for supercapacitors: a review.
Nanoscale 2012;5:7288.
[366] Zhang H, Feng J, Zhang M. Preparation of ower-like CuO by a simple chemical precipitation method and their application
as electrode materials for capacitor. Mater Res Bull 2008;43:32216.
[367] Zhang H, Zhang M. Synthesis of CuO nanocrystalline and their application as electrode materials for capacitors. Mater
Chem Phys 2008;108:1847.
[368] Wang G, Huang J, Chen S, Gao Y, Cao D. Preparation and supercapacitance of CuO nanosheet arrays grown on nickel foam.
J Power Sources 2011;196:575660.
[369] Huang J, Wu H, Cao D, Wang G. Electrochimica acta inuence of Ag doped CuO nanosheet arrays on electrochemical
behaviors for supercapacitors. Electrochim Acta 2012;75:20812.
[370] Zhao B, Liu P, Zhuang H, Jiao Z, Fang T, Xu W, et al. Hierarchical self-assembly of microscale leaf-like CuO on graphene
sheets for high-performance electrochemical capacitors. J Mater Chem A 2013;1:36773.
[371] Dubal DP, Dhawale DS, Salunkhe RR, Jamdade VS, Lokhande CD. Fabrication of copper oxide multilayer nanosheets for
supercapacitor application. J Alloys Compd 2010;492:2630.
[372] Zhang XYL, Wang LL, Ji R, Wang G, Geng B. High electrochemical performance based on ultrathin porous CuO nanobelts
grown on Cu substrate as integrated electrode. Phys Chem Chem Phys 2013;15:5215.
[373] Sun G, Li K, Sun C, Liu Y, He H. Physical and electrochemical characterization of CuO-doped activated carbon in ionic
liquid. Electrochim Acta 2010;55:266772.
[374] Kim D-W, Rhee K-Y, Park S-J. Synthesis of activated carbon nanotube/copper oxide composites and their electrochemical
performance. J Alloys Compd 2012;530:610.
332 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
[375] Shaikh JS, Pawar RC, Moholkar aV, Kim JH, Patil PS. CuOPAA hybrid lms: chemical synthesis and supercapacitor
behavior. Appl Surf Sci 2011;257:438997.
[376] Potyrailo Ra, Surman C, Nagraj N, Burns A. Materials and transducers toward selective wireless gas sensing. Chem Rev
2011;111:731554.
[377] Wang C, Yin L, Zhang L, Xiang D, Gao R. Metal oxide gas sensors: sensitivity and inuencing factors. Sensors (Basel,
Switzerland) 2010;10:2088106.
[378] Jimnez-Cadena G, Riu J, Rius FX. Gas sensors based on nanostructured materials. The Analyst 2007;132:108399.
[379] Lee J-H. Gas sensors using hierarchical and hollow oxide nanostructures: overview. Sensor Actuat B: Chem
2009;140:31936.
[380] Akhavan O, Ghaderi E. Copper oxide nanoakes as highly sensitive and fast response self-sterilizing biosensors. J Mater
Chem 2011;21:12935.
[381] Zhu G, Xu H, Xiao Y, Liu Y, Yuan A, Shen X. Facile fabrication and enhanced sensing properties of hierarchically porous
CuO architectures. ACS Appl Mater Interfaces 2012;4:74451.
[382] Kim H, Jin C, Park S, Kim S, Lee C. H
2
S gas sensing properties of bare and Pd-functionalized CuO nanorods. Sensor Actuat B:
Chem 2012;161:5949.
[383] Kim Y-S, Hwang I-S, Kim S-J, Lee C-Y, Lee J-H. CuO nanowire gas sensors for air quality control in automotive cabin. Sensor
Actuat B: Chem 2008;135:298303.
[384] Chen J, Wang K, Hartman L, Zhou W. H
2
S detection by vertically aligned CuO nanowire array sensors. J Phys Chem C
2008;112:1601721.
[385] Steinhauera S, Brunet E, Maier T, Mutinati G, Kcka A, Freudenberg O, et al. Gas sensing properties of novel CuO nanowire
devices. Sensor Actuat B: Chem 2012. http://dx.doi.org/10.1016/j.snb.2012.09.034.
[386] Mashock M, Yu K, Cui S, Mao S, Lu G, Chen J. Modulating gas sensing properties of CuO nanowires through creation of
discrete nanosized pn junctions on their surfaces. ACS Appl Mater Interfaces 2012;4:41929.
[387] Ramgir NS, Ganapathi SK, Kaur M, Datta N, Muthe KP, Aswal DK, et al. Sub-ppm H
2
S sensing at room temperature using
CuO thin lms. Sensor Actuat B: Chem 2010;151:906.
[388] Hsueh HT, Hsueh TJ, Chang SJ, Hung FY, Tsai TY, Weng WY, et al. CuO nanowire-based humidity sensors prepared on glass
substrate. Sensor Actuat B: Chem 2011;156:90611.
[389] Kim K, Jeong H, Kim H, Choi K, Kim H, Lee J. Selective detection of NO
2
using Cr-doped CuO nanorods. Sensors
2012;12:801325.
[390] Yang M, He J, Hua X, Yan C, Cheng Z, Zhao Y, et al. Copper oxide nanoparticle sensors for hydrogen cyanide detection:
unprecedented selectivity and sensitivity. Sensor Actuat B: Chem 2011;155:6928.
[391] Zhu Z, Garcia-Gancedo L, Flewitt AJ, Xie H, Moussy F, Milne WI. A critical review of glucose biosensors based on carbon
nanomaterials: carbon nanotubes and graphene. Sensors (Basel, Switzerland) 2012;12:59966022.
[392] Cash KJ, Clark Ha. Nanosensors and nanomaterials for monitoring glucose in diabetes. Trends Mol Med 2010;16:58493.
[393] Hahn Y, Ahmad R, Tripathy N. Chemical and biological sensors based on metal oxides nanostructures. Chem Commun
2012;48:1036985.
[394] Vaddiraju S, Tomazos I, Burgess D, Jain F, Papadimitrakopoulos F. Emerging synergy between nanotechnology and
implantable biosensors: a review. Biosens Bioelectron 2010;25:155365.
[395] Wang X, Hu C, Liu H, Du G, He X, Xi Y. Synthesis of CuO nanostructures and their application for nonenzymatic glucose
sensing. Sensor Actuat B: Chem 2010;144:2205.
[396] Zhuang Z, Su X, Yuan H, Sun Q, Xiao D, Choi MMF. An improved sensitivity non-enzymatic glucose sensor based on a CuO
nanowire modied Cu electrode. Analyst 2008;133:12632.
[397] Yang Z, Feng J, Qiao J, Yan Y, Yu Q, Sun K. Copper oxide nanoleaves decorated multi-walled carbon nanotube as platform
for glucose sensing. Anal Methods 2012;4:1924.
[398] Luo L, Zhu L, Wang Z. Nonenzymatic amperometric determination of glucose by CuO nanocubesgraphene
nanocomposite modied electrode. Bioelectrochemistry (Amsterdam, Netherlands) 2012;88:15663.
[399] Prathap MUA, Kaur B, Srivastava R. Hydrothermal synthesis of CuO micro-/nanostructures and their applications in the
oxidative degradation of methylene blue and non-enzymatic sensing of glucose/H
2
O
2
. J Colloid Interface Sci
2012;370:14454.
[400] Wang W, Zhang L, Tong S, Li X, Song W. Three-dimensional network lms of electrospun copper oxide nanobers for
glucose determination. Biosens Bioelectron 2009;25:70814.
[401] Zhang P, Zhang L, Zhao G, Feng F. A highly sensitive nonenzymatic glucose sensor based on CuO nanowires. Microchim
Acta 2011;176:4117.
[402] Wang M-F, Huang Q-A, Li X-Z, Wei Y. Mesoporous CuO: alternative enzyme-free glucose sensing structure with excellent
kinetics of electrode process. Anal Methods 2012;4:3174.
[403] Liu S, Tian J, Wang L, Qin X, Zhang Y, Luo Y, et al. A simple route for preparation of highly stable CuO nanoparticles for
nonenzymatic glucose detection. Catal Sci Technol 2012;2:8137.
[404] Cherevko S, Chung C-H. The porous CuO electrode fabricated by hydrogen bubble evolution and its application to highly
sensitive non-enzymatic glucose detection. Talanta 2010;80:13717.
[405] Wang G, Wei Y, Zhang W, Zhang X, Fang B, Wang L. Enzyme-free amperometric sensing of glucose using CuCuO
nanowire composites. Microchim Acta 2009;168:8792.
[406] Cao F, Gong J. Nonenzymatic glucose sensor based on CuO microbers composed of CuO nanoparticles. Anal Chim Acta
2012;723:3944.
[407] Liu G, Zheng B, Jiang Y, Cai Y, Du J, Yuan H, et al. Improvement of sensitive CuO NFsITO nonenzymatic glucose sensor
based on in situ electrospun ber. Talanta 2012;101:2431.
[408] Jiang L-C, Zhang W-D. A highly sensitive nonenzymatic glucose sensor based on CuO nanoparticles-modied carbon
nanotube electrode. Biosens Bioelectron 2010;25:14027.
[409] Reitz E, Jia W, Gentile M, Wang Y, Lei Y. CuO nanospheres based nonenzymatic glucose sensor. Electroanalysis
2008;20:24826.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 333
[410] Zhang X, Gu A, Wang G, Wei Y, Wang W, Wu H, et al. Fabrication of CuO nanowalls on Cu substrate for a high performance
enzyme-free glucose sensor. CrystEngComm 2010;12:1120.
[411] HussainIbupoto Z, Khun K, BeniV, Liu X, Willander M. Synthesis of novel CuO nanosheets and their non-enzymatic
glucose sensing applications. Sensors 2013;13:792638.
[412] Liu Y, Liao L, Li J, Pan C. From copper nanocrystalline to CuO nanoneedle array: synthesis, growth mechanism, and
properties. J Phys Chem C 2007;111:50506.
[413] Raksa P, Nilphai S, Gardchareon A, Choopun S. Copper oxide thin lm and nanowire as a barrier in ZnO dye-sensitized
solar cells. Thin Solid Films 2009;517:47414.
[414] Sahay R, Sundaramurthy J, Suresh Kumar P, Thavasi V, Mhaisalkar SG, Ramakrishna S. Synthesis and characterization of
CuO nanobers, and investigation for its suitability as blocking layer in ZnO NPs based dye sensitized solar cell and as
photocatalyst in organic dye degradation. J Solid State Chem 2012;186:2617.
[415] Lim Y-F, Choi JJ, Hanrath T. Facile synthesis of colloidal CuO nanocrystals for light-harvesting applications. J Nanomater
2012;2012:16.
[416] Kidowaki H, Oku T, Akiyama T, Suzuki A, Jeyadevan B, Cuya J. Fabrication and characterization of CuO-based solar cells. J
Mater Sci Res 2011;1:13843.
[417] Wang P, Zhao X, Li B. ZnO-coated CuO nanowire arrays: fabrications, optoelectronic properties, and photovoltaic
applications. Opt Express 2011;19:112719.
[418] Yuhas BD, Yang P. Nanowire-based all-oxide solar cells. J Am Chem Soc 2009;131:375661.
[419] Zhai T, Li L, Wang X, Fang X, Bando Y, Golberg D. Recent developments in one-dimensional inorganic nanostructures for
photodetectors. Adv Funct Mater 2010;20:423348.
[420] Shen G, Chen D. One-dimensional nanostructures for photodetectors. Recent Patents Nanotechnol 2010;4:2031.
[421] Soci C, Zhang A, Bao X-Y, Kim H, Lo Y, Wang D. Nanowire photodetectors. J Nanosci Nanotechnol 2010;10:143049.
[422] Zhai T, Fang X, Liao M, Xu X, Zeng H, Yoshio B, et al. A comprehensive review of one-dimensional metal-oxide
nanostructure photodetectors. Sensors (Basel, Switzerland) 2009;9:650429.
[423] Manna S, Das K, De SK. Template-free synthesis of mesoporous CuO dandelion structures for optoelectronic applications.
ACS Appl Mater Interfaces 2010;2:153642.
[424] Xu J, Sun J, Wei J, Xu J. The wavelength dependent photovoltaic effects caused by two different mechanisms in carbon
nanotube lm/CuO nanowire array hetero-dimensional contacts. Appl Phys Lett 2012;100:251113.
[425] Huang H, Zhang L, Wu K, Yu Q, Chen R, Yang H, et al. Hetero-metal cation control of CuO nanostructures and their high
catalytic performance for CO oxidation. Nanoscale 2012;4:783241.
[426] Feng Y, Zheng X. Plasma-enhanced catalytic CuO nanowires for CO oxidation. Nano Lett 2010;10:47626.
[427] Zhong Z, Ng V, Luo J, Teh SP, Teo J, Gedanken A. Manipulating the self-assembling process to obtain control over the
morphologies of copper oxide in hydrothermal synthesis and creating pores in the oxide architecture. Langmuir
2007;23:59717.
[428] Zhou M, Gao Y, Wang B, Rozynek Z, Fossum JO. Carbonate-assisted hydrothermal synthesis of nanoporous CuO
microstructures and their application in catalysis. Eur J Inorg Chem 2009;2010:72934.
[429] Zhang H, Cao JL, Shao GS, Yuan ZY. Synthesis of transition metal oxide nanoparticles with ultrahigh oxygen adsorption
capacity and efcient catalytic oxidation performance. J Mater Chem 2009;19:60979.
[430] Zhong K, Xue J, Mao Y, Wang C, Zhai T, Liu P, et al. Facile synthesis of CuO nanorods with abundant adsorbed oxygen
concomitant with high surface oxidation states for CO oxidation. RSC Adv 2012;2:115208.
[431] Pillai UR, Deevi S. Room temperature oxidation of carbon monoxide over copper oxide catalyst. Appl Catal B
2006;64:14651.
[432] Yu Q, Huang H, Chen R, Wang P, Yang H, Gao M, et al. Synthesis of CuO nanowalnuts and nanoribbons from aqueous
solution and their catalytic and electrochemical properties. Nanoscale 2012;4:261320.
[433] Balk Derekaya F, Kutar C, Gldr . Selective CO oxidation over ceria supported CuO catalysts. Mater Chem Phys
2009;115:496501.
[434] Ratnasamy P, Srinivas D, Satyanarayana C, Manikandan P, Senthil Kumaran R, Sachin M, et al. Inuence of the support on
the preferential oxidation of CO in hydrogen-rich steam reformates over the CuOCeO
2
ZrO
2
system. J Catal
2004;221:45565.
[435] Sun C, Zhu J, Lv Y, Qi L, Liu B, Gao F, et al. Dispersion, reduction and catalytic performance of CuO supported on ZrO
2
-
doped TiO
2
for NO removal by CO. Appl Catal B 2011;103:20620.
[436] Ling P, Li D, Wang X. Supported CuO/c-Al
2
O
3
as heterogeneous catalyst for synthesis of diaryl ether under ligand-free
conditions. J Mol Catal A: Chem 2012;357:1126.
[437] Horns a, Hungra aB, Bera P, Lpez Cmara a, Fernndez-Garca M, Martnez-Arias a, et al. Inverse CeO
2
/CuO catalyst as
an alternative to classical direct congurations for preferential oxidation of CO in hydrogen-rich stream. J Am Chem Soc
2010;132:345.
[438] Pan K, Ming H, Yu H, Liu Y, Kang Z, Zhang H, et al. Different copper oxide nanostructures: synthesis, characterization, and
application for CN cross-coupling catalysis. Cryst Res Technol 2011;46:116774.
[439] Ganesh Babu S, Karvembu R. CuO nanoparticles: a simple, effective, ligand free, and reusable heterogeneous catalyst for
N-arylation of benzimidazole. Ind Eng Chem Res 2011;50:9594600.
[440] Zhu M, Diao G. High catalytic activity of CuO nanorods for oxidation of cyclohexene to 2-cyclohexene-1-one. Catal Sci
Technol 2012;2:824.
[441] Zhang W, Zeng Q, Zhang X, Tian Y, Yue Y, Guo Y, et al. Ligand-free CuO nanospindle catalyzed arylation of heterocycle CH
bonds. J Org Chem 2011;76:47415.
[442] Suramwar NV, Thakare SR, Karade NN, Khaty NT. Green synthesis of predominant (111) facet CuO nanoparticles:
heterogeneous and recyclable catalyst for N-arylation of indoles. J Mol Catal A: Chem 2012;359:2834.
[443] Kim JY, Park JC, Kang H, Song H, Park KH. CuO hollow nanostructures catalyze [3 + 2] cycloaddition of azides with
terminal alkynes. Chem Commun 2010;46:43941.
[444] Akkilagunta V, Kakulapati R. Synthesis of unsymmetrical suldes using ethyl potassium xanthogenate and recyclable
copper catalyst under ligand-free conditions. J Org Chem 2011;76:681924.
334 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
[445] Jammi S, Sakthivel S, Rout L, Mukherjee T, Mandal S, Mitra R, et al. CuO nanoparticles catalyzed CN, CO, and CS cross-
coupling reactions: scope and mechanism. J Org Chem 2009;74:19716.
[446] Lam S-M, Sin J-C, Abdullah AZ, Mohamed AR. Degradation of wastewaters containing organic dyes photocatalysed by zinc
oxide: a review. Desalin Water Treat 2012;41:13169.
[447] Chong MN, Jin B, Chow CWK, Saint C. Recent developments in photocatalytic water treatment technology: a review.
Water Res 2010;44:29973027.
[448] Tan YN, Wong CL, Mohamed AR. An overview on the photocatalytic activity of nano-doped-TiO
2
in the degradation of
organic pollutants. ISRN Mater Sci 2011;2011:261219.
[449] Zhang Y, Ram MK, Stefanakos EK, Goswami DY. Synthesis, characterization, and applications of ZnO nanowires. J
Nanomater 2012;2012:624520.
[450] Ahmed S, Rasul M, Martens WN, Brown R, Hashib M. Heterogeneous photocatalytic degradation of phenols in
wastewater: a review on current status and developments. Desalination 2010;261:318.
[451] Wang Z, Zhao S, Zhu S, Sun Y, Fang M. Photocatalytic synthesis of M/Cu
2
O (M = Ag, Au) heterogeneous nanocrystals and
their photocatalytic properties. CrystEngComm 2011;13:22627.
[452] Miyauchi M, Nakajima A, Watanabe T, Hashimoto K. Photocatalysis and photoinduced hydrophilicity of various metal
oxide thin lms. Chem Mater 2002;14:28126.
[453] Yu H, Yu J, Liu S, Mann S. Template-free hydrothermal synthesis of CuO/Cu
2
O composite hollow microspheres. Chem
Mater 2007;19:432734.
[454] Xu H, Zhu G, Zheng D, Xi C, Xu X, Shen X. Porous CuO superstructure: precursor-mediated fabrication, gas sensing and
photocatalytic properties. J Colloid Interface Sci 2012;383:7581.
[455] Zaman S, Zainelabdin A, Amin G, Nur O, Willander M. Efcient catalytic effect of CuO nanostructures on the degradation
of organic dyes. J Phys Chem Solids 2012;73:13205.
[456] Gao F, Pang H, Xu S, Lu Q. Copper-based nanostructures: promising antibacterial agents and photocatalysts. Chem Eng
(Cambridge) 2009. http://dx.doi.org/10.1039/b904801d:3571-3.
[457] Natarajan TS, Thomas M, Natarajan K, Bajaj HC, Tayade RJ. Study on UV-LED/TiO
2
process for degradation of Rhodamine B
dye. Chem Eng J 2011;169:12634.
[458] Wang L, Zhou Q, Zhang G, Liang Y, Wang B, Zhang W, et al. A facile room temperature solution-phase route to synthesize
CuO nanowires with enhanced photocatalytic performance. Mater Lett 2012;74:2179.
[459] Liu X, Li Z, Zhang Q, Li F, Kong T. CuO nanowires prepared via a facile solution route and their photocatalytic property.
Mater Lett 2012;72:4952.
[460] Liu S, Tian J, Wang L, Luo Y, Sun X. One-pot synthesis of CuO nanoower-decorated reduced graphene oxide and its
application to photocatalytic degradation of dyes. Catal Sci Technol 2012;2:33944.
[461] Shi W, Chopra N. Controlled fabrication of photoactive copper oxidecobalt oxide nanowire heterostructures for efcient
phenol photodegradation. ACS Appl Mater Interfaces 2012;4:5590607.
[462] Li Y, Zhou Je, Tung S, Schneider E, Xi S. A review on development of nanouid preparation and characterization. Powder
Technol 2009;196:89101.
[463] Zhu HT, Zhang CY, Tang YM, Wang JX. Novel synthesis and thermal conductivity of CuO nanouid. J Phys Chem C
2007;111:164650.
[464] Lee S, Choi SUS, Li S, Eastman JA. Measuring thermal conductivity of uids containing oxide nanoparticles. J Heat Transfer
1999;121:2809.
[465] Liu M-S, Lin MC-C, Huang I-T, Wang C-C. Enhancement of thermal conductivity with CuO for nanouids. Chem Eng
Technol 2006;29:727.
[466] Moghadassi AR, Hosseini SM, Henneke DE. Effect of CuO nanoparticles in enhancing the thermal conductivities of
monoethylene glycol and parafn uids. Ind Eng Chem Res 2010;49:19004.
[467] Zhu H, Han D, Meng Z, Wu D, Zhang C. Preparation and thermal conductivity of CuO nanouid via a wet chemical method.
Nanoscale Res Lett 2011;6:181.
[468] Das SK, Putra N, Thiesen P, Roetzel W. Temperature dependence of thermal conductivity enhancement for nanouids. J
Heat Transfer 2003;125:567.
[469] Hojjat M, Etemad SG, Bagheri R, Thibault J. Thermal conductivity of non-Newtonian nanouids: experimental data and
modeling using neural network. Int J Heat Mass Transfer 2011;54:101723.
[470] Dreizin EL. Metal-based reactive nanomaterials. Prog Energy Combust Sci 2009;35:14167.
[471] Chou SK, Yang WM, Chua KJ, Li J, Zhang K. Development of micro power generators a review. Appl Energy
2011;88:116.
[472] Xu D, Yang Y, Cheng H, Li Y, Zhang K. Integration of nano-Al with Co
3
O
4
nanorods to realize high-exothermic coreshell
nanoenergetic materials on a silicon substrate. Combust Flame 2012;159:22029.
[473] Zhou X, Xu D, Zhang Q, Lu J, Zhang K. Facile green in situ synthesis of Mg/CuO core/shell nanoenergeticarrays with a
superior heat-release property and long-termstorage stability. ACS Appl Mater Interfaces 2013;5:76416.
[474] Stamatis D, Jiang Z, Hoffmann VK, Schoenitz M, Dreizin EL. Fully dense, aluminum-rich AlCuO nanocomposite powders
for energetic formulations. Combust Sci Technol 2008;181:97116.
[475] Ermoline A, Schoenitz M, Dreizin EL. Reactions leading to ignition in fully dense nanocomposite Al-oxide systems.
Combust Flame 2011;158:107683.
[476] Ermoline A, Stamatis D, Dreizin E. Low-temperature exothermic reactions in fully dense AlCuO nanocomposite powders.
Thermochim Acta 2011;527:528.
[477] Sanders VE, Asay BW, Foley TJ, Tappan BC, Pacheco AN, Son SF. Reaction propagation of four nanoscale energetic
composites (Al/MoO
3
, Al/WO
3
, Al/CuO, and Bi
2
O
3
). J Propul Power 2007;23:70714.
[478] Blobaum KJ, Reiss ME, Plitzko JM, Weihs TP. Deposition and characterization of a self-propagating CuOx/Al thermite
reaction in a multilayer foil geometry. J Appl Phys 2003;94:2915.
[479] Blobaum KJ, Wagner aJ, Plitzko JM, Van Heerden D, Fairbrother DH, Weihs TP. Investigating the reaction path and growth
kinetics in CuOx/Al multilayer foils. J Appl Phys 2003;94:2923.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 335
[480] Zhu P, Shen R, Fiadosenka N, Ye Y, Hu Y. Dielectric structure pyrotechnic initiator realized by integrating Ti/CuO-based
reactive multilayer lms. J Appl Phys 2011;109:084523.
[481] Apperson S, Shende RV, Subramanian S, Tappmeyer D, Gangopadhyay S, Chen Z, et al. Generation of fast propagating
combustion and shock waves with copper oxide/aluminum nanothermite composites. Appl Phys Lett 2007;91:243109.
[482] Zhang K, Rossi C, Ardila Rodriguez Ga, Tenailleau C, Alphonse P. Development of a nano-Al/CuO based energetic material
on silicon substrate. Appl Phys Lett 2007;91:113117.
[483] Ohkura Y, Liu S-Y, Rao PM, Zheng X. Synthesis and ignition of energetic CuO/Al core/shell nanowires. Proc Combust Inst
2011;33:190915.
[484] Zhang K, Rossi C, Petrantoni M, Mauran N. A nano initiator realized by integrating Al/CuO-based nanoenergetic materials
with a Au/Pt/Cr microheater. J Microelectromech Syst 2008;17:8326.
[485] Yang Y, Xu D, Zhang K. Effect of nanostructures on the exothermic reaction and ignition of Al/CuOx based energetic
materials. J Mater Sci 2012;47:1296305.
[486] Sverac F, Alphonse P, Estve A, Bancaud A, Rossi C. High-energy Al/CuO nanocomposites obtained by DNA-directed
assembly. Adv Funct Mater 2012;22:3239.
[487] Thiruvengadathan R, Bezmelnitsyn A, Apperson S, Staley C, Redner P, Balas W, et al. Combustion characteristics of novel
hybrid nanoenergetic formulations. Combust Flame 2011;158:96478.
[488] Zhan RZ, Chen J, Deng SZ, Xu NS. Study of techniques for improving emission uniformity of gated CuO nanowire eld
emitter arrays. J Vac Sci Technol B 2010;28:C2C45.
[489] Zhu YW, Moo aM, Yu T, Xu XJ, Gao XY, Liu YJ, et al. Enhanced eld emission from O
2
and CF
4
plasma-treated CuO
nanowires. Chem Phys Lett 2006;419:45863.
[490] Liu GX, Chen J, Deng SZ, Xu NS. Preparation of CuO nanowire eld emitter arrays by thermal oxidation in water vapor
containing environments and their eld emission properties. In: 24th, International vacuum nanoelectronics conference
(IVNC), 2011. p. 378.
[491] Hsueh H, Hsueh T, Chang S, Tsai T, Hung F, Chang S, et al. CuO-nanowire eld emitter prepared on glass substrate. IEEE
Trans Nanotechnol 2011;10:11615.
[492] Zhan RZ, Chen J, Deng SZ, Xu NS. Fabrication of gated CuO nanowire eld emitter arrays for application in eld emission
display. J Vac Sci Technol B 2010;28:558.
[493] Crick CR, Parkin IP. Preparation and characterisation of super-hydrophobic surfaces. Chemistry (Weinheim an der
Bergstrasse, Germany) 2010;16:356888.
[494] Li J, Guo Z, Liu J-H, Huang X-J. Copper nanowires array: controllable construction and tunable wettability. J Phys Chem C
2011;115:1693440.
[495] Chen X, Kong L, Dong D, Yang G, Yu L, Chen J, et al. Fabrication of functionalized copper compound hierarchical structure
with bionic superhydrophobic properties. J Phys Chem C 2009;113:5396401.
[496] Liu X, Jiang Z, Li J, Zhang Z, Ren L. Super-hydrophobic property of nano-sized cupric oxide lms. Surf Coat Technol
2010;204:32004.
[497] Zhang Q, Xu D, Hung T, Zhang K. Facile synthesis, growth mechanism and reversible superhydrophobic and
superhydrophilic properties of non-aking CuO nanowires grown from porous copper substrates. Nanotechnology
2013;24:065602.
[498] Basu M, Sinha AK, Pradhan M, Sarkar S, Negishi Y, Pal T. Fabrication and functionalization of CuO for tuning
superhydrophobic thin lm and cotton wool. J Phys Chem C 2011;115:2095363.
[499] Li J, Liu X, Ye Y, Zhou H, Chen J. Fabrication of superhydrophobic CuO surfaces with tunable water adhesion. J Phys Chem C
2011;115:47269.
[500] Zhang BJ, Park J, Kim KJ, Yoon H. Biologically inspired tunable hydrophilic/hydrophobic surfaces: a copper oxide self-
assembly multitier approach. Bioinspiration Biomimetics 2012;7:036011.
[501] Chaudhary A, Barshilia HC. Nanometric multiscale rough CuO/Cu(OH)
2
superhydrophobic surfaces prepared by a facile
one-step solution-immersion process: transition to superhydrophilicity with oxygen plasma treatment. J Phys Chem C
2011;115:1821320.
[502] Guo Z, Chen X, Li J, Liu J-H, Huang X-J. ZnO/CuO hetero-hierarchical nanotrees array: hydrothermal preparation and self-
cleaning properties. Langmuir 2011;27:6193200.
[503] Reddy KJ, Roth TR. Arsenic removal from natural groundwater using cupric oxide. Ground Water 2013;51:8391.
[504] Chiban M, Zerbet M, Carja G, Sinan F. Application of low-cost adsorbents for arsenic removal: a review. J Environ Chem
Ecotoxicol 2012;4:91102.
[505] Duarte AALS, Cardoso SJA, Alada AJ. Emerging and innovative techniques for arsenic removal applied to a small water
supply system. Sustainability 2009;1:1288304.
[506] Martinson Ca, Reddy KJ. Adsorption of arsenic(III) and arsenic(V) by cupric oxide nanoparticles. J Colloid Interface Sci
2009;336:40611.
[507] Pillewan P, Mukherjee S, Roychowdhury T, Das S, Bansiwal A, Rayalu S. Removal of As(III) and As(V) from water by copper
oxide incorporated mesoporous alumina. J Hazard Mater 2011;186:36775.
[508] Roth TR, Reddy KJ. In: Bundschuh J, editor. Natural arsenic in groundwaters of Latin America. London: Taylor & Francis;
2008. p. 60514.
[509] Cao A, Monnell JD, Matranga C, Wu J, Cao L, Gao D. Hierarchical nanostructured copper oxide and its application in arsenic
removal. J Phys Chem C 2007;111:186248.
[510] Karlsson HL, Cronholm P, Gustafsson J, Mller L. Copper oxide nanoparticles are highly toxic: a comparison between
metal oxide nanoparticles and carbon nanotubes. Chem Res Toxicol 2008;21:172632.
[511] Karlsson HL, Gustafsson J, Cronholm P, Mller L. Size-dependent toxicity of metal oxide particles-a comparison between
nano- and micrometer size. Toxicol Lett 2009;188:1128.
[512] Aruoja V, Dubourguier H-C, Kasemets K, Kahru A. Toxicity of nanoparticles of CuO, ZnO and TiO
2
to microalgae
Pseudokirchneriella subcapitata. Sci Total Environ 2009;407:14618.
[513] Rousk J, Ackermann K, Curling SF, Jones DL. Comparative toxicity of nanoparticulate CuO and ZnO to soil bacterial
communities. PloS One 2012;7:e34197.
336 Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337
[514] Horie M, Fujita K, Kato H, Endoh S, Nishio K, Komaba LK, et al. Association of the physical and chemical properties and the
cytotoxicity of metal oxide nanoparticles: metal ion release, adsorption ability and specic surface area. Metallomics
2012;4:35060.
[515] Luna-delRisco M, Orupld K, Dubourguier H-C. Particle-size effect of CuO and ZnO on biogas and methane production
during anaerobic digestion. J Hazard Mater 2011;189:6038.
[516] Applerot G, Lellouche J, Lipovsky A, Nitzan Y, Lubart R, Gedanken A, et al. Understanding the antibacterial mechanism of
CuO nanoparticles: revealing the route of induced oxidative stress. Small 2012;8:332637.
[517] Yang Z, Liu ZW, Allaker RP, Reip P, Oxford J, Ahmad Z, et al. A review of nanoparticle functionality and toxicity on the
central nervous system. J Roy Soc Interface 2010;7:S41122.
[518] Brayner R. The toxicological impact of nanoparticles. Nano Today 2008;3:4855.
[519] De Jong WH, Borm PJa. Drug delivery and nanoparticles: applications and hazards. Int J Nanomed 2008;3:13349.
[520] Soenen SJ, Rivera-Gil P, Montenegro J-M, Parak WJ, De Smedt SC, Braeckmans K. Cellular toxicity of inorganic
nanoparticles: common aspects and guidelines for improved nanotoxicity evaluation. Nano Today 2011;6:44665.
Q. Zhang et al. / Progress in Materials Science 60 (2014) 208337 337

Vous aimerez peut-être aussi