Sie sind auf Seite 1von 39

CHAPTER 2

Growth, Properties, and


Applications of Copper Oxide
and Nickel Oxide/Hydroxide
Nanostructures
Ahmad Umar, Mohammad Vaseem, Yoon-Bong Hahn
School of Semiconductor and Chemical Engineering, Centre for Future Energy
Materials and Devices, Chonbuk National University, Jeonju 561-756, South Korea
CONTENTS
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2. Growth and Properties of Copper Oxide Nanostructures . . . . 3
2.1. One-Dimensional Nanostructures of Copper Oxide . . . 3
2.2. Complex Nanostructures of Copper Oxide . . . . . . . . . 9
2.3. Spherical, Urchin, and Flower-Shaped
Nanostructures of Copper Oxide . . . . . . . . . . . . . . . . 13
3. Applications of Copper Oxide Nanostructures . . . . . . . . . . 23
3.1. Photocatalytic Properties of Copper
Oxide Nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2. Field-Emission Properties of
Copper Oxide Nanostructures . . . . . . . . . . . . . . . . . . 27
3.3. Sensor Applications of Copper
Oxide Nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . 29
4. Growth and Properties of Nickel Oxide and
Hydroxide Nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . 30
5. Concluding Remarks and Future Directions . . . . . . . . . . . . 37
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
1. INTRODUCTION
The science of nanotechnology is based on using the smallest units of matter to
design/process new materials/devicesatom by atomin order to obtain superior per-
formance on the basis of atomic-scale architecture. Thus, fabrication of nanomaterials
ISBN: 1-58883-170-1
Copyright 2010 by American Scientic Publishers
All rights of reproduction in any form reserved.
1
Metal Oxide Nanostructures and Their Applications
Edited by Ahmad Umar and Yoon-Bong Hahn
Volume 2: Pages 139
2 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
with well-dened structures and precisely controlled sizes is crucial to the develop-
ment of nanotechnology. The investigation of nanostructures can provide unprecedented
understanding of materials and devices; nanostructures exhibit novel and signicantly
improved physical/chemical/biological properties, phenomena, and processes compared
to their bulk counterparts [1, 2]. Previous work has been done with various materials to
systematically investigate the experimental conditions for fabricating nanostructures suit-
able for desired applications. Synthesis of inorganic nanostructures controlled in terms
of size and shape has been strongly motivated by the desired practical applications that
depend on the size and shape of such structures [172]. Therefore, over the last few
decades, signicant effort has been made to synthesize inorganic nanostructures with
the desired physical/chemical properties for possible applications in the fabrication of
efcient nanodevices [428]. Metal oxide semiconducting nanostructures are one of the
most versatile classes of semiconducting materials due to their diverse properties and
functionalities. Metal oxide nanostructures exhibit unique properties that can be used in
a variety of applications for the fabrication of novel and efcient nanodevices. Therefore,
rapid research developments have been made in the eld of metal oxide nanostructures
in terms of their growth and applications. The study of metal oxides has attracted a
great deal of interest due to the importance of their size- and shape-dependent properties
in electronic/optoelectronic applications. Hence, investigations have attempted to con-
trol the morphology of metal oxide nanomaterials. Among the variety of nanostructured
morphologies, the so-called 1-D nanostructured materials (e.g., nanotubes, nanorods,
nanowires, nanobelts, nanoribbons) have been intensively studied due to their substantial
importance and potential applications. However, for novel technologies based on nano-
scale machines/devices, not only 1-D nanomaterials are needed but also other complex
nanostructures are desirable. Nanomaterials with higher degrees of engineering and more
complex architectures, such as 2-D and 3-D nanostructures, have potential applications
in light emission/detection, eld-emission applications, biomedical devices, selective gas
sensors, electrode materials in batteries, and so forth. Numerous 2-D or 3-D nanostruc-
tures of various materials have already been synthesized and reported in the literature;
these complex nanostructures exhibit excellent physical properties and hence provide
opportunities for scientists to use these nanostructures for the fabrication of highly ef-
cient nanodevices. Therefore, the development of rational, general strategies for fabri-
cation of multidimensional, interconnected, patterned assemblies of nanoscale building
blocks will be signicant accomplishments. These developments are key to the success of
bottom-up approaches toward future nanodevices able to exploit these building blocks.
Among different semiconducting metal oxides, copper oxide (CuO) has been stud-
ied as a unique and attractive mono-oxide material for both fundamental investigations
and practical applications. CuO is an important p-type/positive-doped, transition-metal
oxide semiconductor, having a narrow bandgap (bandgap energy |
g
= 1.2 eV). CuO
semiconductors exhibit a versatile range of applications such as fabrication of electrical,
optical, and photovoltaic devices; selective gas sensing devices; heterogeneous catalysts;
magnetic storage media; eld-emission devices (e.g., eld-emission gun); solar cells; and
Li-ion electrode materials. CuO also possesses complex magnetic phases and forms the
basis for several high-T
C
superconductors (i.e., superconductors for which the Curie tem-
perature T
C
is high) and materials with high magnetoresistance. Moreover, CuO can be
used to prepare a variety of organicinorganic nanostructured composites with unique
characteristics, which include high thermal and electrical conductivity, high mechanical
strength, and high-temperature durability. Due to their versatile properties and wide
applications, various CuO nanostructures have already been synthesized using a variety
of fabrication techniques; these techniques have included hydrothermal methods, solgel
techniques, gas-phase oxidation, and microemulsion. Solution processes present an easy,
low-energy, low-temperature, and cost-effective approach to obtain CuO products with
good yields. Therefore, up to now, a variety of procedures for the preparation of CuO
nanostructures by solution methods have been reported in literature; such procedures
yield nanorods, nanowires, nanotubes, nanosheets, and nanoribbons. Other than these
so-called 1-D structures, some complex nanostructures of CuO have also been reported
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 3
in the literature. In addition to these, nanostructures of nickel hydroxide (Ni(OH)
2
)
one of the most important transition-metal hydroxideshave received increasing atten-
tion due to their exotic properties and extensive applications. Notably, applications of
Ni(OH)
2
include use as an active material in positive electrodes and use in Ni-based,
alkaline rechargeable batteries with the characteristics of high power density, excellent
cyclability, high specic energy, and low toxicity. As for nickel oxide (NiO), it forms
p-type, wide-bandgap material, which can be used as a transparent, semiconducting
layer. As an important material, NiO can be used in various applications, such as mag-
netic materials, catalysts, electrochromic lms, fuel-cell electrodes, selective gas sensors,
active optical bers, and battery electrodes. It has been observed that the performance
of Ni-based, alkaline rechargeable batteries and other devices depends on the struc-
tural/morphological features of Ni(OH)
2
and NiO; hence, considerable work has been
done to prepare and investigate nanocrystalline Ni(OH)
2
and NiO structures.
In this chapter, we assemble information about the growth, properties, and applica-
tions of various kinds of copper oxide, nickel oxide, and nickel hydroxide nanostructures
grown using fabrication techniques reported in the literature to date.
2. GROWTH AND PROPERTIES OF COPPER OXIDE
NANOSTRUCTURES
Among the various semiconducting metal oxides, copper oxide (CuO) has been stud-
ied as a unique mono-oxide material that is attractive for both fundamental investiga-
tions and practical applications. Due to its versatile properties, CuO is highly useful
for fabrication of various highly efcient devices. Nanostructures of CuO have been
used in various applications such as high-T
C
superconductors, heterogeneous catalysts,
selective gas sensors, Li-ion electrode materials, eld-emission devices, magnetic stor-
age media, and solar energy transformation [128]. Due to the excellent properties and
versatile applications of CuO, a wide range of CuO nanostructures have been syn-
thesized using various fabrication techniques. The CuO nanostructures reported in the
literature include nanowires, nanorods, nanobelts, nanotubes, nanoplatelets, nanorib-
bons, nanobers, nanoparticles, nanosheets, ower-shaped structures, and sea urchin-like
nanostructures. Here, we present growth procedures and properties of many CuO nanos-
tructures fabricated via a variety of techniques.
2.1. One-Dimensional Nanostructures of Copper Oxide
The so-called 1-D nanostructures (i.e., nanorods, nanowires, nanotubes) have gener-
ated great interest due to the prospects of these materials. For example, the study of
transport processes in 1-D systems has provided insight into the transport processes of
low-dimensional systems. These 1-D nanostructures have versatile applications such as
fabrication of nanoelectronics, where these nanostructures can be used as single-electron
transistors. Hitherto, a variety of 1-D CuO nanostructures have been reported in the
literature. In this section, we review growth procedures and properties of 1-D CuO nano-
structures fabricated using various methodologies.
Hsieh and colleagues demonstrated the synthesis and structural characterizations of
well-ordered CuO nanobers [29]. These well-ordered CuO nanobers were synthesized
via a self-catalytic growth process using a polycarbonate membrane as a template and
Cu nuclei sites (Cu(111)). The Cu nuclei sites were uniformly deposited on a Cu sub-
strate via high-voltage input (electric eld = 15 Vcm
1
) in a copper sulfate solution. Two
sizes of Cu nuclei were obtained depending upon the pore diameter of the polycarbonate
membranes used as template. Upon heat treatment in an oxygen (O
2
) atmosphere, the
electrodeposited Cu nuclei were transformed into CuO nanober arrays. X-ray diffraction
(XRD) and transmission electron microscopy (TEM) analyses conrmed the nanocrys-
talline nature of the as-grown CuO nanobers. In the synthesis of CuO nanobers, Hsieh
and colleagues observed that the self-catalytic reaction involved proceeds through the
thermal oxidation of Cu nuclei on the (111) plane [29]. By this process, well-ordered
CuO nanobers with different diameters were formed; interestingly, the diameter of an
4 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
individual nanober approximated the size to its Cu nucleus. This indicates that the
size of the Cu nucleus determines the diameter of the as-grown nanober. The possible
growth mechanism for the formation of CuO nanobers was also demonstrated by Hsieh
and colleagues; they conrmed that the nanobers were grown via a self-catalytic growth
process using Cu nuclei as basic units [29].
Cao and colleagues reported the synthesis of Cu, Cu
2
O, and CuO nanotubes/nanorods
in the presence of a structure-directing surfactant, cetyl trimethylammonium bromide
(CTAB) by varying reaction conditions and using Cu(OH)
2
4
as an inorganic precursor
[30]. The authors claimed that by using this method, not only the products but also the
morphologies could be controlled. The Cu and Cu
2
O nanostructures were obtained by
reducing Cu(OH)
2
4
with hydrazine hydrate (NH
2
NH
2
H
2
O) and glucose at room tem-
perature (i.e., 300 K), respectively. Facile hydrothermal treatment of Cu(OH)
2
4
results the
formation of CuO nanostructures; results of structural observations conrmed that the
CuO nanostructures grown were single-crystal structures. In this synthesis, the investi-
gators observed that at a lower concentration of the Cu(OH)
2
4
precursor tubular struc-
tures were formed, while at a higher concentration rod-like structures were formed.
Regarding the growth process of the CuO nanostructures grown (i.e., nanorods, nano-
tubes), the authors speculated that, due to electrostatic interaction, the inorganic pre-
cursor Cu(OH)
2
4
and the cationic surfactant CTAB could form different conformational
surfactantinorganic composites under different reaction conditions; the different com-
posites could then serve as templates. Thus, by manipulating surfactantinorganic aggre-
gates in solution, various kinds of CuO nanostructures can be obtained [30].
In another report, Xu and colleagues presented the synthesis of CuO nanowires grown
on Cu foil by thermal oxidation of the Cu foils at 400

C and 500

C under various gaseous


environments (e.g., air, N
2
, O
2
, with/without water vapor) [31]. These authors observed
that nanowires were formed exclusively from monoclinic CuO crystals under a gaseous
atmosphere with a sufciently high oxygen partial pressure; no nanowires were found in
samples oxidized in N
2
, with/without water vapor. Moreover, uniform nanowires were
formed at high density in wet air; nanowires were formed in a only small amounts in
dry air. The CuO nanowires formed in pure oxygen (O
2
) had the highest density. Finally,
the authors concluded that the high oxygen partial pressure enhanced both the nucleation
probability and the growth rate of the nanowires, while the effect of water vapor was
mainly to assist nucleation. Results of detailed structural observations conrmed that the
nanowires grown possessed monoclinic structures. A vaporsolid mechanism was pro-
posed by the authors for the growth of these CuO nanowires presented in this study [31].
In work by Mei and colleagues, CuO nanowire arrays on Si-based (i.e., SiO
2
) nano-
scale islands were fabricated via nanochannels of Si-based, porous anodic alumina (PAA)
as templates at room temperature under pulse voltage in a conventional solution for
electrodeposition of Cu [32]. Detailed structural characterizations by XRD and X-ray pho-
toelectron spectrometry (XPS) revealed that the main component of the nanowires was
Cu
2
O and that the nanowires had a preferential growth direction along the (111) planes
that were connected with the nanoscale SiO
2
islands (as conrmed by TEM results).
Regarding the growth of the nanowires obtained, it was proposed that the formation of
Cu
2
O was due to the alkalinity of the solution of anodized material [32].
By using single-crystal Cu
2
(OH)
2
CO
3
nanoribbons as precursors, various kinds of 1-D
CuO nanostructures/morphologies (e.g., nanoribbons, scroll-like structures, rod-like
structures, arrays of CuO nanoparticles) have been prepared via heat treatment by Zhu
and colleagues [33]. These authors observed that the morphologies of CuO nanostructures
are strongly dependant on heat-treatment conditions. For instance, at relatively low-heat
treatment and low heating rates, the morphological features of the precursor can be pre-
served; arrays of CuO nanoparticles can be obtained at high heating rates; CuO rod-like
structures can be prepared with increased heat treatment. Moreover, the authors suggested
that the formation of scroll-like structures of CuO in a transition stage of Cu
2
(OH)
2
CO
3
decomposition may be understood according to the position/arrangement of Cu atoms in
the crystal structure of a Cu
2
(OH)
2
CO
3
cell. Detailed structural characterization indicated
that the CuO nanostructures obtained were monoclinic, nanocrystalline structures [33].
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 5
Xu and colleagues reported the synthesis and detailed structural characterization of
CuO nanowires grown onto Cu foils heated in wet air at controlled temperatures [34].
Two different morphologies of nanowires (i.e., straight and curved) were obtained from
two temperature zones. Regarding the growth of the nanowires obtained, the authors
claimed that growth behavior of nanowires can be understood in terms of the kinetics
(that is, a short circuit of diffusion of atoms/ions during the reaction), the strength of
the nanowires, and the thickness ratio of the oxide scale/coating versus the metal foil.
Thus, increasing the oxidation temperature may produce a reduction in the density of
the growing nanostructures and an increase in their diameter and strength, causing them
to adopt the straight morphology of nanowires. Moreover, deformation of the thin oxide
scale/coating under further thermal stress may contribute to the formation of curved
nanowires [34].
Large-scale synthesis of CuO nanowires has been achieved by thermal evaporation
of Cu foils in ambient oxygen (O
2
) at 300

C900

C by Huang and colleagues [35].


These authors observed that the evaporation conditions can affect the formation of the
nanowires grown by this method. Thus, in the lower end of the temperature range exam-
ined, nanowires with small diameters were formed. The amount of nanowires obtained
also varied with the evaporation temperature. When the evaporation temperature was as
high as 800

C, only a small amount of CuO nanobers formed; however, when the evap-
oration temperature was 400

C750

C, large amounts of CuO nanowires were obtained.


Additionally, the authors observed that with increases in evaporation time, the length
of nanowires obtained increases; however, the growth rate decreases. Results of exten-
sive studies of the as-grown CuO nanowires conrmed the nanocrystalline nature for the
nanostructures grown by this large-scale method. A vaporsolid growth mechanism was
proposed for the growth of nanowires; the authors claimed that by heating the Cu foil
in an oxygen (O
2
) atmosphere, CuO lms were easily obtained on the substrate. Thus,
when Cu foil was heated in an oxygen (O
2
) atmosphere, the foil was oxidized to form
Cu
2
O in a rst step, and CuO was formed in a second step (i.e., oxidation of Cu
2
O) [35].
Additionally, growth of CuO nanowires depended on supersaturation.
Lu and colleagues demonstrated the high-yield, low-cost, controlled synthesis of
Cu(OH)
2
nanowires and nanoribbons in solution-phase reactions by simply dripping
KOH and ammonia (NH
3
) solutions into an aqueous solution of CuSO
4
at ambient
temperature [36]. Well-dened CuO nanostructures (e.g., nanoplatelets, nanoleaets,
nanowires) were then produced by thermal dehydration of the as-prepared Cu(OH)
2
nanostructures in solution or in the solid state. These investigators observed that the
conversion from Cu(OH)
2
to CuO in solution occurred mainly through a reconstructive
transformation involving a dissolution process followed by CuO crystallization. The ther-
mal dehydration of 1-D Cu(OH)
2
nanostructures in the solid state normally resulted in
morphology-reserved, 1-D CuO nanostructures [36]. Results of detailed structural studies
conrmed that the CuO nanostructures obtained were monoclinic, nanocrystalline struc-
tures (see Fig. 1 for studies on nanowires). Regarding the conversion of Cu(OH)
2
nanos-
tructures into CuO nanostructures, the authors stated that, when Cu(OH)
2
nanowires
(obtained by the KOHNH
3
route) were heated in the original solution immediately after
the addition of the two alkaline solutions, a complete transformation to CuO occurred at
temperatures as low as 50

C [36]. The authors also suggested that, during the solution-


phase synthesis of CuO nanostructures, the composition and morphology of the nal
product are largely dependent on the synthesis conditions (e.g., basicity, solvent); further,
the solid precursors (e.g., Cu(OH)
2
) may exist only momentarily if at all.
Du and Van Tendeloo synthesized CuO nanowires and nanobelts using copper nitrate
(Cu(NO
3
)
2
), NH
3
, and NaOH in a Teon-lined stainless autoclave at 130

C for 10 h [37].
The products were characterized by various techniques such as XRD, HRTEM, and elec-
tron energy-loss spectrometry (EELS). By XRD, the nanostructures obtained were pure
CuO with a monoclinic structure having the lattice constants of u = 4.685 , | = 3.425 ,
c = 5 , p = 99.549; these constants were very similar to the reported data (JCPDS: 45-
0937). Regarding the growth process of CuO nanostructures from Cu(OH)
2
nanowires,
the authors suggested that during the heating process the Cu(OH)
2
nanowires lost H
2
O
6 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
(A)
(C)
(B)
(D)
Figure 1. Results of SEM (A), TEM (B), ED (C), and XRD (D) studies of CuO nanowires obtained by thermal
dehydration of as-prepared Cu(OH)
2
nanowires in the solid state at 120

C for 2 h. Scale bars: 500 nm (A);


200 nm (B). Reprinted with permission from [36], C. Lu et al., J. Phys. Chem. B 108, 17825 (2004).
2004, American Chemical Society. ED = electron diffraction, SEM = scanning electron microscopy, TEM =
transmission electron microscopy, XRD = X-ray diffraction.
and transformed into CuO while the basic morphology of Cu(OH)
2
nanowires remained
[37]. According to Cudennec and Lecerf, the transformation from Cu(OH)
2
to CuO can
be understood by the simple chemical reaction [38] described in reaction (I).
Cu(OH)
2(s)
CuO
(s)
+H
2
O
(g)
(I)
It was suggested that the loss of water was performed by an oxolation mechanism,
which would involve a dehydration process and the formation of OCuO bridges [38].
These bridges would be formed after the loss of water and would be followed by con-
traction of the structure along the [010] direction. Simultaneously, shifts of CuO
4
groups
or Cu atoms along the [001] direction would take place to promote evolution toward
crystallized CuO; this would lead to the formation of CuO nanowires.
Regarding the growth of CuO nanobelts, Du and Van Tendeloo suggested that the
growth mechanism was different from that of CuO nanowires and was due to a process of
thermal dehydration and then recrystallization as the Cu(OH)
2
nanobelts [37]. In a typical
process, divalent copper ions (Cu
2+
) were rst dissolved in the form of complex anions
Cu(OH)
2
4
, which result in the square planar surrounding. These anions can be considered
the precursors for the formation of CuO. Here, a condensation phenomenon combines
with a loss of two hydroxyl ions and one water molecule leading to the formation of
chains of square planar CuO
4
groups and then to the formation of CuO in solid form.
This transformation can be described according to reactions (II) and (III).
Cu(OH)
2(S)
+2OH

(aq)
Cu(OH)
2
4 (aq)
(II)
CuO
(s)
+2OH

(aq)
+H
2
O (III)
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 7
Therefore, it is predicted that the presence of water facilitates the morphological trans-
formation from nanowire to nanobelt.
Yu and colleagues reported the growth of rod-shaped CuO nanostructures and inves-
tigated the polarized micro-Raman scattering for an individual CuO nanorod [39]. The
CuO nanorods were synthesized by a one-step annealing process using commercially
available Cu plates as starting material. For the synthesis, a Cu plate was annealed at
400

C for 24 h in air. After the desired reaction time, a black ash-like top layer formed
on the Cu plate; this layer contained CuO nanorods. The authors investigated individual
CuO nanorods with different aspect ratios using polarized micro-Raman scattering. An
obvious anisotropy in the intensity of the Raman modes was observed when the electric
eld vector of the incident laser beam was parallel/perpendicular to the long axis of
a nanorod. The mechanism responsible for the polarized Raman spectra observed was
attributed to the polarization effect produced by the large length:diameter ratio of the
CuO nanorods and the large contrast in the dielectric properties of these nanorods versus
the surrounding environment [39].
CuO nanorods of various crystalline structures/morphologies were synthesized in an
NaOH solution by a simple, efcient method by Gao and colleagues [40]. These investi-
gators observed that the crystalline structures/morphologies of the products were highly
dependent upon the temperature of the hydrothermal treatment in the synthesis proce-
dure. Results of structural characterization by XRD conrmed the monoclinic, nanocrys-
talline state of the as-grown CuO nanostructures. When the as-grown CuO nanomaterials
were used as anode materials for Li-ion batteries, ne polycrystalline nanorods exhibited
high electrochemical capacity (766 mAhg
1
) as compared to single-crystal bulk nanorods
(416 mAhg
1
); this was due to the large surface area and numerous structural defects of
the polycrystalline nanorods. However, the capacity retention of polycrystalline nanorods
was not as good as that of single-crystal nanorods due to the Li-driven, irreversible
morphological/crystalline changes during the electrochemical reaction cycles to which
anode materials are exposed [40].
Chang and Zeng reported several wet chemical methods for synthesis of 1-D CuO
nanostructures in which waterethanol solutions were used as solvents at 77

C82

C
and 1 atm [41]. Various 1-D CuO nanostructures (e.g., nanorods, nanowires, nanorib-
bons, nanoplatelets, nanosheets) were obtained by such methods. It was observed that
at low reaction temperatures and normal atmospheric pressure, monodisperse, single-
crystal CuO nanorods (with a selected breadth of 515 nm) could be prepared by simply
changing the starting Cu ion concentration. For nanorods, preferential growth is along
the [010] direction. By using a two-step, continuous process in which seeding and length
growth could be controlled under a pseudo-steady-state operation, these investigators
demonstrated the synthesis of rigid or exible nanorods/nanoribbons with lengths up
to 1 jm. The morphologies of the pristine nanorods obtained were further modied
by an aging treatment. These authors also reported the self-assembly of small nanorods
into 2-D netted structures. These netted structures were formed after prolonged heat-
ing, and the nanorods were attached to each other using their {001}, {100}, and {110}
crystal planes. Nanoplatelets or nanosheets of CuO were also synthesized using a high
concentration of NaOH; in this procedure, growth along the [100] direction becomes
pronounced. Results of detailed structural characterization using XRD and selected area
electron diffraction (SAED) conrmed that the CuO nanostructures grown were mono-
clinic, crystal structures [41].
Yao and colleagues demonstrated the synthesis/characterization of uniform, monodis-
perse CuO nanorods using a process in which size could be controlled [42]. This pro-
cess involved spontaneous aggregation/crystallization of CuO nanoparticles generated
from a solidliquid, arc-discharge process under ambient conditions in the absence of
surfactants/additives. These investigators found that newly formed Cu nanoclusters gen-
erated by this method rapidly oxidized into CuO nanoparticles, and these CuO nanopar-
ticles spontaneously aggregated/self-organized and then crystallized into uniform CuO
nanorods via a prolonged aging time under ambient conditions. Therefore, by choosing
a suitable reducing agent to prevent oxidation of the Cu nanoclusters, selective synthesis
8 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
of CuO, Cu
2
O, and Cu nanostructures was achieved. Results from XRD conrmed that
the CuO nanorods obtained were monoclinic, crystalline structures [42].
The growth and properties of CuO nanowires grown by thermal oxidation of Cu
sheets in an oxygen (O
2
) atmosphere were investigated by Kaur and colleagues [43].
This procedure was carried out in a resistively heated furnace at various tempera-
tures (400

C800

C) and various times under a ow of oxygen (O


2
). These investigators
observed that branched CuO nanowires were obtained by long oxidation times. The
length of the nanowires obtained increased with annealing time; at 22 h, nanowires
of >20 jm length were obtained. Results of detailed structural studies conrmed that
the nanowires and their branches had grown along specic crystallographic directions;
that is, nanowires grew along the [010] direction and the branches grew along the [210]
direction. The bandgap of as-grown nanowires was determined using UV-visible absorp-
tion spectrometry; the bandgap of the as-grown CuO nanowires was larger than that of
bulk material [43].
Wang and colleagues reported the synthesis and sensing applications of hydrother-
mally grown CuO nanorods [44]. The growth of CuO nanorods was performed in a
Teon-lined stainless steel autoclave (at 120

C150

C for 12 h) using the cationic sur-


factant CTAB, copper acetate, and NaOH as source materials. The authors reported that
during the reaction, the formation of CuO occurred according to the following reaction:
Cu
2+
+2OH

CuO+H
2
O. The XRD pattern of the as-prepared product was indexed
to the tenorite structure of CuO with lattice constants of u = 4.96 , | = 3.43 , and
c = 5.13 (JCPDS: 41-0254). Results of detailed structural characterization using high-
resolution TEM (HRTEM) conrmed that the nanorods synthesized were grown parallel
to the [001] direction [44].
A solution-phase route for the synthesis of single-crystal CuO nanoribbons was devel-
oped by Wang and colleagues [45]. The nanoribbons grown had widths and thicknesses
of 1080 nm and 520 nm, respectively; their lengths ranged from several hundred nano-
meters to several micrometers. In addition to nanoribbons, CuO nanorings (diameter =
100300 nm) were fabricated by the same procedure [45]. These CuO nanostructures were
fabricated by the reaction of CuCl
2
and NaOH in the presence of sodium dodecylben-
zenesulfonate (NaDBS). The as-prepared products were characterized by powder XRD,
TEM, and HRTEM; the CuO nanoribbons and nanorings both had monoclinic, single-
crystal structures. The nanorings were closed but not by the simple superposition of the
two ends of nanoribbons. On the basis of TEM observations, the formation of nanorib-
bons and nanorings was interpreted to be a multistage process in which initial nanoakes
would split into nanoribbons due to Brownian motion of the surfactant molecules; then
nanoribbons, which would have polar surfaces, would coil into nanorings to reduce elec-
trostatic energy. Moreover, it was proposed that single-crystal nanorings were formed via
a spontaneous, self-coiling process during the growth of polar nanoribbons that would
have an alternating stack of Cu
2+
and O
2
ions along the [002] axis [45].
Wen and colleagues synthesized aligned, CuO nanoribbon arrays (that were approxi-
mately perpendicular to the Cu substrate surface) by a solution-treatment process with a
subsequent heat-treatment process [46]. For the synthesis of CuO nanoribbons, the initial
Cu(OH)
2
nanoribbons were fabricated by a simple, coordinated self-assembly method
in which Cu
2+
ions were oxidized from the surface of Cu foil in an alkaline solution.
Heat treatment in a horizontal, quartz-tube furnace was used to remove water from the
Cu(OH)
2
nanoribbons to yield CuO nanoribbons. The CuO nanoribbons synthesized were
50 nm in width and several nanometers in thickness; nanoribbon length was controlled
by varying reaction temperature and time [46]. The CuO nanoribbons obtained were char-
acterized using scanning electron microscopy (SEM), TEM, HRTEM, and XRD (Fig. 2). By
low-magnication TEM, the average diameter of the nanoribbons obtained was 50 nm
(Fig. 2(A)). The SAED pattern from an ensemble of nanoribbons (Fig. 2(A), inset) was
typical of monoclinic CuO with three diffraction rings of (

110), (

111), and (111). The cor-


responding XRD pattern again conrmed the monoclinic nature of the CuO nanoribbons
obtained (Fig. 2(B)). SEM images (as in Fig. 2(C)) showed that the CuO nanoribbons
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 9
(A)
(C)
(D)
(B)
Figure 2. TEM (A), XRD (B), SEM (C), and HRTEM (D) analyses of CuO nanoribbons obtained from Cu(OH)
2
nanoribbons by heat treatment. Inset of panel A represents the SAED pattern of an ensemble of CuO nanorib-
bons. Inset of panel D represents the SAED pattern of single CuO nanoribbons. Reprinted with permission
from [46], X. Wen et al., Langmuir 19, 5898 (2003). 2003, American Chemical Society. HRTEM = high-
resolution transmission electron microscopy, SAED=selected area electron diffraction, SEM=scanning electron
microscopy, TEM= transmission electron microscopy, XRD = X-ray diffraction.
obtained after heat treatment were still aligned along the surface of the Cu foil sub-
strate where the initial Cu(OH)
2
nanoribbons had been fabricated by oxidation. HRTEM
images (as in Fig. 2(D)) revealed that the CuO nanoribbons obtained were crystals, but
the crystallinity was not perfect; fringes were often wavy or discontinuous. Nevertheless,
the CuO crystallites had a preferential orientation in the nanoribbons; that is, the [

110]
direction is along the longest dimension of the nanoribbon. This agreed with the SAED
pattern (Fig. 2(D), inset). Crystallographic analysis by both XRD and SAED indicated
that the nanoribbon axis was along the [

110] direction. It is plausible that the (

110) and
(002) planes of CuO crystallites (with d
hkl
values of 2.75 and 2.52 , respectively) were
converted from the (100) and (002) planes of Cu(OH)
2
crystallites (with d
hkl
values of
2.95 and 2.63 , respectively) by release of H
2
O [46].
2.2. Complex Nanostructures of Copper Oxide
In addition to 1-D CuO nanostructures, a variety of complex nanostructures of CuO have
been reported in the literature. In this section, we summarize the growth procedures
and properties of complex CuO nanostructures synthesized using various methodologies
reported in the literature.
Complex dendrite-like CuO nanostructures, consisting of a rod-like main stem and
rod-like subbranches, have been synthesized via a simple, ethylene glycol (EG)assisted
hydrothermal method by Zhang and colleagues [47]. Detailed structural characterization
10 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
by XRD and SAED indicated that the dendrite-like CuO nanostructures were in a mon-
oclinic state and an individual branch had a single-crystal nature. The authors observed
that the structures/morphologies of the products strongly depend on growth conditions
such as pH and temperature. The function of EG in the hydrothermal synthesis of CuO
nanostructures was investigated; EG not only promoted growth of branches but also had
reducing ability within a certain range of alkaline conditions. These investigators pro-
posed that in the EG-assisted hydrothermal synthesis of CuO nanostructures, pH had
three functions: providing a source of OH

ions, facilitating the reducing ability of EG,


and stimulating branched structural growth. The quantity of OH

ions, the reducing


ability of EG, and the branched growth of CuO nanostructures all increased as the pH
increased. When the pH of the solution was -3, the concentration of OH

ions was too


low to form CuO nanostructures; with increasing pH (i.e., pH 510), sufcient quantities
of OH

ions promoted the formation of branched CuO nanostructures. Thus, a variety


of factors affected the structure/morphology of evolving CuO nanostructures [47].
Liu and colleagues reported the synthesis and characterization of 2-D CuO layered,
oval nanosheets and 3-D CuO nanoellipsoids grown by facile, large-scale, template-free
methods at 65

C [48]. The investigators observed that the shape and dimensionality of


the CuO nanostructures fabricated could be controlled by tuning the reaction pH. Well-
dened, single-crystal CuO nanostructures were produced under these conditions. At
pH 8.5, CuO nanosheets were formed; at pH 7.5, CuO nanoellipsoids were formed. The
morphological control of the synthetic architecture was based initially on primary build-
ing units and subsequently on arrangement of the units along the preferential crystal
faces. SEM and TEM were used to evaluate the precise morphologies of the as-prepared
structures. These results revealed that both the 2-D layered, oval nanosheets and the
3-D nanoellipsoids were elongated along the [010] direction. XRD results conrmed that
the products obtained were monoclinic, nanocrystalline structures. Room-temperature,
UV-visible absorption spectra (\ = 200800 nm) were used to estimate the bandgap ener-
gies of the nano-sized semiconductors. These CuO nanoarchitectures exhibited blueshifts
in UV-visible spectra and large bandgaps (i.e., larger than those of bulk CuO crystals).
These investigators also presented additional controlled experiments in which changes in
growth temperature and alkaline reactants were used to prepare other ultrane nanoar-
chitectures. Finally, the investigators demonstrated the growth of single-crystal CuO
architectures built from 0-D and 1-D nanocrystals through a one-step, solution-phase
chemical route under controlled conditions [48].
Zhu and colleagues reported the synthesis and characterization of CuO nanouids
with high volume fractions created by transforming an unstable Cu(OH)
2
precursor to
CuO with a combination of ultrasonic vibration and microwave irradiation in water [49].
The inuences of type of precursors, ultrasonic vibration, microwave irradiation, and
dispersant were also studied. These investigators proposed that the unstable Cu(OH)
2
precursor was disrupted to form small particles under ultrasonic vibration in water.
The precursor was completely transformed into CuO nanoparticles under microwave
irradiation in water, and the dispersant (e.g., ammonium citrate) prevented aggregation
of nanoparticles in the preparation. These factors resulted in the formation of a stable,
aqueous CuO nanouid [49].
CuO nanostructures with various complex morphologies were synthesized by hydrol-
ysis of copper acetate in the presence of urea (NH
2
CONH
2
) under mild hydrother-
mal conditions by Zhong and colleagues [50]. In this method of synthesis, various
monoamines (e.g., isobutylamine, octylamine, dodecylamine, octadecylamine) and
diamines (e.g., ethylenediamine dihydrochloride, hexamethylenediamine) were used as
structure-directing agents. Use of a monoamine led to formation of 1-D aggregates of CuO
precursor particles (Pre-CuO), while use of a diamine led to formation of 2-D aggregates
of Pre-CuO. In both cases, the shorter carbon-chain amine molecules showed a stronger
structure-directing function than the longer carbon-chain amine molecules. When the
rate of hydrolysis was slowed by coupling the hydrolysis reaction to an esterication
reaction, 1-D aggregates of Pre-CuO were formed; when the rate of hydrolysis was in
a middle range, spherical Pre-CuO architectures (composed of small linear aggregates)
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 11
were formed. However, under high rates of hydrolysis (achieved by increasing the con-
centration of the precipitation agent (i.e., urea) or by conducting the reaction at high
temperatures (i.e., 120

C)), only Pre-CuO nanoparticles with featureless morphology


were formed. In one set of experiments, spherical Pre-CuO architectures were formed
and converted to a porous structure (CuO
x
) after removal of octylamine via calcination.
Compared to 1-D and 2-D aggregates, the porous architecture obtained was highly ther-
mostable and did not collapse even after calcination at 500

C [50].
Chen and colleagues demonstrated the growth and properties of shuttle-like CuO
nanostructures synthesized using the process known as pulsed-laser-induced liquidsolid
interfacial reaction (PLIIR) [51]. In a typical reaction process, deionized water was used
as the reactive liquid, and the solid target was Cu bulk material (purity = 99.7%). The
Cu target was rst xed on the bottom of a reaction chamber. Subsequently, deion-
ized water was poured slowly into the chamber until the target was covered by 5 mm.
Finally, the pulsed laser was focused on the target surface, and during the laser abla-
tion, the target and reaction uid were maintained at room temperature while the target
was rotated at a slow speed (i.e., 10 rotations min
1
). After the pulsed laser interacted
with the target for 45 min, the PLIIR chamber was heated to 120

C for 30 min. The


powders synthesized were collected from the reaction uid and analyzed in terms of
their structural and optical properties. SEM, XRD, and TEM (in conjunction with an
energy-dispersive X-ray spectrometer) were used to determine the morphology, structure,
and composition of the as-prepared products. The XRD pattern of the as-prepared CuO
nanostructures conrmed the monoclinic crystal nature of the nanostructures grown by
this procedure. HRTEM images indicated that the interplanar spacing of 0.25 nm and
0.27 nm corresponded to the {110} and {200} crystallographic planes of CuO nanostruc-
tures, respectively. An anomalous peak (\
max
= 258 nm) was observed in the UV absorp-
tion spectrum of the shuttle-like CuO nanoparticles; the authors proposed from their
experimental analyses that the anomalous UV absorption peak was induced by the shape
of the CuO nanoparticles. Moreover, these investigators employed theoretical calculations
based on discrete dipole approximation to pursue the physical origin of the anomalous
UV absorption peak. Thus, both experimental and theoretical evidence were used to
show that the shape of nanoparticles has a great inuence on the optical properties of
nanomaterials [51].
Xiao and colleagues demonstrated the controlled synthesis of CuO nanostructures with
various morphologies [52]. The 1-D, 2-D, and 3-D CuO nanostructures obtained were
synthesized in the presence of sodium citrate by a hydrothermal process under con-
trolled the reaction conditions. These investigators observed that, when the molar ratio of
sodium citrate to CuSO
4
(represented as the SC:Cu
2+
ratio) was -1.0, 1-D CuO nanorods
with diameters of 3040 nm and lengths of 100200 nm were formed. In the absence of
sodium citrate, 2-D ake-like CuO nanostructures with widths of 150200 nm and lengths
of 300400 nm were formed. Moreover, when SC:Cu
2+
ratios were >1.0, 3-D branch-like
CuO nanostructures with lengths of hundreds of nanometers and diameters of 20100 nm
were formed. UV-visible studies were performed to assess the optical properties of the as-
prepared CuO nanostructures. By UV absorption measurements, the CuO nanostructures
obtained displayed a blueshift in the bandgap relative to bulk CuO material. Moreover,
the optical bandgap energy (|
g
) of CuO nanostructures obtained could be tuned through
morphological control of the structures. UV absorption measurements revealed that the
estimated |
g
of the 1-D rod-like, 2-D ake-like, and 3-D branch-like CuO nanostructures
was 2.36 eV, 1.60 eV, and 1.40 eV, respectively; all of these values were larger than the
reported values (|
g
= 1.2 eV) of bulk CuO materials [52].
Wang and colleagues demonstrated the synthesis and characterization of CuO
nanowhiskers by a one-step, solid-state reaction in the presence of a nonionic surfactant,
polyethylene glycol (PEG) 400 [53]. For the growth of CuO nanowhiskers, CuCl
2
2H
2
O,
NaOH, and PEG were used as source materials. In detailed structural observations, XRD
data conrmed that the nanowhiskers obtained were monoclinic, nanocrystalline struc-
tures with a space group of C
6
2l
. TEM observations conrmed that the nanowhiskers
12 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
grown were shaped as rods with smooth, clean surfaces; also, the nanowhiskers had uni-
form diameters of 8 nm and lengths > 100 nm. By HRTEM, the interplanar spacing of
the nanowhiskers grown was 0.248 nm, which corresponds to the {002} crystal planes
of CuO nanostructures [53].
Zhang and colleagues reported the large-scale growth of CuO nanostructured lms
including nanotube arrays, ordered nanotube arrays with a special nanoplate wall struc-
ture, and nanoower lms [54]. These investigators fabricated such novel nanostructures
by manipulating the conditions of the chemical reactions. The CuO nanotube arrays were
successfully synthesized by heating CuO nanotubes under a nitrogen (N
2
) atmosphere
(Cu(OH)
2
nanotubes were initially prepared under N
2
and used as a precursor for the
growth of CuO nanotubes). These investigators observed that CuO nanotubes (even after
heating) retain morphology similar to that of initially prepared Cu(OH)
2
nanotube pre-
cursor. In addition, ordered CuO nanotube arrays with a special nanoplate wall structure
were obtained by heating Cu(OH)
2
nanotubes in an autoclave with benzene (C
6
H
6
) as the
solvent. The CuO nanotube arrays formed still kept the initial nanotube structure, but
exhibited features different from other CuO nanotubes. Two complex nanostructures
CuO nanotube arrays (prepared under a nitrogen (N
2
) atmosphere) and ordered CuO
nanotube arrays with a special nanoplate wall structure (prepared in benzene)were dis-
tinguished by two aspects. (1) The latter CuO tubes had average diameters in the range
of 300500 nm, which were thicker than those of the CuO tubes formed under a nitro-
gen (N
2
) atmosphere. (2) The special nanoplate wall structure (formed in benzene) had
a maize-like morphology, which was made up of well-aligned nanoplates (with uniform
thicknesses of 15 nm) stacked along the direction of the tube axis; the walls of the CuO
nanotubes formed under a nitrogen (N
2
) atmosphere were smooth. Also, some of the
nanotube tips grew closed under the solvothermal condition. Films of CuO nanoowers
were prepared by oxidizing Cu foil hydrothermally in an autoclave. By various analytic
methods, all the CuO nanostructures reported were conrmed to be well crystallized
with monoclinic structure [54].
Liu and colleagues reported the large-scale synthesis and the characterization of single-
crystal CuO nanoplatelets grown via a hydrothermal process (120

C for 12 h) using cupric


dodecylsulfate (Cu(DS)
2
) and NaOH as source materials [55]. In this synthetic process,
the Cu(DS)
2
functions both as structure-directing agent and source material for the pro-
duction of Cu
2+
ions. The CuO nanoplatelets synthesized are monoclinic, single-crystal
structures (lattice constant u =0.468 nm) as conrmed by XRD results. Detailed structural
analyses by SEM and TEM revealed that the CuO nanoplatelets obtained were of thick-
ness 50 nm, width 120300 nm, and length 480700 nm. The width-to-thickness ratios
were 26. XPS and other analytic methods indicated that the CuO nanoplatelets grown
were highly pure. In addition, these investigators conducted several time-dependant
reactions to suggest the possible growth mechanism involved. In UV-visible absorption
spectral analyses, the as-prepared CuO nanoplatelets exhibited an absorption peak at
287 nm. From absorption edge analyses, the optical bandgap of the as-grown CuO
nanoplatelets was calculated at 2.60 eV, much larger than the reported value for bulk
CuO materials (|
g
= 1.85 eV) [55].
Zhao and colleagues demonstrated the synthesis and characterization of a variety of
CuO nanostructures including radially arrayed nanowhiskers, cubic nanoparticles, and
spherical nanoparticles [56]. In a typical reaction process, an aqueous solution of copper
acetate and ethylene glycol (EG) were place into a microwave reux system under ambi-
ent conditions for 15 min at a power of 365 W. Various concentrations of EG in aqueous
solutions were used to control the morphologies of the CuO nanostructures obtained
during the microwave heating/irradiation. Reaction temperature, irradiation time, and
copper acetate concentration were also found to be critical factors affecting the morphol-
ogy of the products. Detailed structural observations conrmed that all the products
synthesized were monoclinic, nanocrystalline structures [56].
Interestingly, 2-D hierarchical CuO nanosheets were synthesized by Zheng and Liu via
a facile, solution-phase method at near-neutral pH (7.5) and near room temperature (i.e.,
30

C) without the use of any additives or templates [57]. From low-magnication SEM
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 13
images, it was clearly seen that the as-prepared CuO products were grown on a large
scale. It was also observed that most of the nanosheets were ellipsoidal with average
horizontal-axis and longitudinal-axis sizes of 400 nm and 200 nm, respectively. Results
of high-resolution SEM conrmed the 2-D sheet-like conguration and indicated that the
thickness of the nanosheets obtained was 25 nm. Moreover, the 2-D sheet-like structures
were actually composed of small CuO nanocrystals; hence, these structures were termed
hierarchical 2-D nanosheet structures. The XRD pattern conrmed that the as-prepared
hierarchical 2-D CuO nanosheets were monoclinic, nanocrystalline CuO structures with
the lattice constants u
0
= 4.684 , |
0
= 3.425 , c
0
= 5.129 . The optical properties of
the as-grown hierarchical 2-D CuO nanosheets were observed by UV-visible studies, and
|
g
was estimated at 1.94 eV [57].
Well-aligned arrays of CuO nanoplatelets synthesized by a hydrothermal route with-
out the assistance of any kind of template were reported by Zou and colleagues [58].
In a typical reaction process, cuprous chloride (CuCl) was dissolved in a concentrated
NH
3
solution, and the resultant solution was added to deionized water with continuous
stirring. The mixture obtained was then transferred to a 60-mL Teon-lined stainless steel
autoclave, which was then sealed and maintained at 140

C for 40 h. After terminating the


reaction, the autoclave was cooled to room-temperature, and the products synthesized
were scratched, centrifuged, washed with distilled water and ethanol, and dried at 40

C
for 2 h in air. The CuO nanostructures synthesized were characterized in detail in terms
of their structural and optical properties using various analytic tools. XRD peaks for the
as-prepared CuO nanostructures could be clearly indexed to the monoclinic structure of
(space group C2,c). Moreover, compared with the standard diffraction peaks for crys-
talline CuO structures, no other peaks were observed (i.e., peaks belonging to impurities
such as Cu(OH)
2
or Cu
2
O) conrming the high purity of as-prepared CuO nanostruc-
tures. Field-emission SEM (FESEM) observations provided detailed information about
the morphologies of the as-grown CuO nanostructures. It was observed that the CuO
nanocrystals self-organized into wall-like structures and that most of the nanocrystals
with uniform morphology were upright, outward, densely packed, and well aligned.
From FESEM observations, it was also found that these nanocrystals possessed platelet-
like structures (i.e., nanoplatelets). It was observed that nanoplatelets were 5080 nm
in thickness, 150250 nm in width, and 0.81.5 jm in length. Most of the as-prepared
nanoplatelets had four clear edges, resembling prisms; however, the nanoplatelets did not
obtain uniform tops. In addition, a few patch-like nanocrystals were also found among
the products. Electron microscopic analysis showed that the nanoplatelets grew along
the [010] direction. The Ostwald ripening mechanism was used to describe the growth
of CuO nanoplatelets. The optical/electrochemical properties of the as-prepared prod-
ucts were also investigated. The arrays of CuO nanoplatelets exhibited a blueshift in the
UV-visible spectra, a slow capacity fading rate, and a relatively high coulombic efciency
in the chargedischarge process.
2.3. Spherical, Urchin, and Flower-Shaped Nanostructures of Copper Oxide
There are a few reports in the literature that present the synthesis and characterization of
spherical and sea urchin-shaped CuO nanostructures. In this section, we summarize the
growth procedures and properties of such CuO nanostructures synthesized using various
methodologies.
Keyson and colleagues demonstrated the synthesis and characterization of urchin-
shaped CuO nanostructures [59]. The synthesis was carried out by a simple, novel,
hydrothermal microwave method using PEG, CuCO
3
Cu(OH)
2
, and NH
4
OH at 120

C in
1 h. Various techniques were used for detailed structural characterization of the urchin-
shaped CuO nanostructures obtained. All diffraction peaks of the structures synthesized
could be indexed to CuO monoclinic structures. The lattice parameters were calculated
by least-squares renement using the UnitCell-97 program (Department of Earth Sci-
ences, University of Cambridge); the estimated parameters were u = 4.692 , | = 3.428 ,
and c = 5.136 with volume cell of 81.52
3
. These values were consistent with those
14 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
reported in the literature (JCPDS: 45-0937). The FESEM images revealed that the urchin-
shaped particles grown were uniform spheres with diameters of 0.71.9 jm. The specic
surface area of the CuO nanostructured microspheres was 170.5 m
2
g
1
. In the pro-
posed growth mechanism of urchin-shaped CuO nanostructures, PEG acted as a template
of sorts for the formation of the nanostructures obtained. During the rst stage of the
growth process, the aqueous solution of Cu
2+
plus PEG and NH
4
OH gave rise to the rst
nucleation seeds, which acted as initial nuclei for particle growth. When the particles
reached a critical dimension, PEG absorbed the small particles by exposed OH bonds;
these PEG-absorbed particles then acted as templates for the formation of CuO nanos-
tructures. In the nal stage of the growth process, urchin-shaped CuO nanostructures
were formed.
Vaseem and colleagues also reported sea urchin-like and sheet-like CuO nanostruc-
tures synthesized in large quantities via simple, solution processes [60]. Urchin-like CuO
nanostructures were obtained from Cu powder present in a strong alkali solution of cop-
per nitrate [Cu(NO
3
)
2
]; in this process, small CuO nanosheets were nucleated and grown
on the outer surfaces of Cu powder particles. Sheet-like structures, some arranged in
ower-shaped morphologies, were also found in the same reaction solution. Figure 3
shows typical morphologies of the urchin-like CuO nanostructures synthesized. Metal-
lic Cu powder was used to synthesize urchin-like CuO nanostructures that were made
of thin CuO nanosheets. Aggregates of almost-spherical Cu particles with diameters of
12 jm were observed in samples of the Cu powder (Figs. 3(A and B)). These Cu par-
ticles were oxidized and converted into CuO in the presence of a strong alkali solution
of Cu(NO
3
)
2
. Upon examination of the general morphology of the as-prepared urchin-
like CuO nanostructures, it was revealed that these structures were created by the ne
arrangement of thin CuO nanosheets. The authors proposed that the Cu particles were
(A) (B)
(C) (D)
Figure 3. Typical low-magnication (A) and high-magnication (B) FESEM images of Cu powder prior to
processing and low-resolution (C) and high-resolution (D) FESEM images of as-grown, sea urchin-like CuO
nanostructures grown on the Cu powder by a simple, solution process. Reprinted with permission from [60],
M. Vaseem et al., Mater. Lett. 62, 1659 (2008). 2008, Elsevier B. V., Amsterdam, The Netherlands. FESEM =
eld-emission scanning electron microscopy.
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 15
oxidized and converted into CuO particles in the presence of the strong alkali solution
of Cu(NO
3
)
2
; CuO nanosheets were then arranged on these CuO particles as the reac-
tion proceeded for longer time. Interestingly, these investigators found that the sizes of
the Cu particles were smaller than the urchin-like CuO nanostructures formed; thus,
they believed that some converted CuO particles adhered to each other and formed
the larger, urchin-like CuO nanostructures. The typical diameter of urchin-like morpho-
logies was 24 jm (Fig. 3(C)), while some smaller, urchin-like morphologies were also
seen in micrographs (Fig. 3(D)). In addition to urchin-like morphologies, 2-D sheet-like
structures were also observed in samples of the reactant solution (Fig. 4). Figure 4(A)
presents a low-magnication image of the as-grown nanosheets and reveals that the
sheets were grown in large quantity. Interestingly, it is also seen that some nanosheets
were arranged in such a fashion that they made ower-like morphologies with an average
diameter of 25 jm (Fig. 4(B)) while some nanosheets were arranged in irregular man-
ner (Fig. 2(C)). As revealed by high-resolution FESEM, the thicknesses of the as-grown
nanosheets were in the range of 4060 nm (Fig. 4(D)); the as-grown nanosheets were
24 jm wide. TEM images (Figs. 5(A and B)) of the as-grown urchin-like CuO nano-
structures are fully consistent in terms of morphology and dimensionality with FESEM
images (Figs. 3(C and D)). The XRD pattern of the as-prepared products agreed well with
those of nanocrystalline CuO structures, and the peaks could be indexed to a monoclinic,
crystalline CuO structure (JCPDS: 48-1548) (Fig. 5(C)). Moreover, in the XRD pattern,
two dominant peaks (located at 20 values of 35.6

and 38.8

and indexed as (

111)(002)
and (111)(200) planes, respectively) were characteristic for pure, monoclinic CuO crys-
tallites. Several bands appeared in the Fourier transform infrared (FTIR) spectrum of an
as-grown sample (Fig. 5(D)). The presence of weak absorption at 3334 cm
1
is due to the
stretching vibration of adsorbed water and surface hydroxyl groups. Several other bands
in the FTIR spectrum (at 598, 525, and 430 cm
1
) are characteristic of monoclinic CuO
structures and conrm the monoclinic structure of the as-grown nanostructures [60].
(A) (B)
(C) (D)
Figure 4. Typical low-magnication (A, B) and high-magnication (C, D) FESEM images of ower-like nanos-
tructures composed of thin nanosheets and irregularly arranged sheet-like nanostructures grown in the reactant
solution. Reprinted with permission from [60], M. Vaseem et al., Mater. Lett. 62, 1659 (2008). 2008, Elsevier
B. V., Amsterdam, The Netherlands. FESEM= eld-emission scanning electron microscopy.
16 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
(A)
(C)
(D)
(B)
Figure 5. Typical TEM images (A, B), XRD pattern (C), and FTIR spectrum (D) of the as-grown, sea urchin-like
CuO nanostructures grown by a simple, solution process. Reprinted with permission from [60], M. Vaseem
et al., Mater. Lett. 62, 1659 (2008). 2008, Elsevier B. V., Amsterdam, The Netherlands. FTIR = Fourier transform
infrared, TEM= transmission electron microscopy, XRD = X-ray diffraction.
Synthesis and characterization of CuO nanostructured microspheres using a novel,
solid-stabilized emulsion method was reported by He [61]. For the growth of CuO
microspheres, Cu(NO
3
)
2
, dimethyl oxalate, 1-hexanol, and acetone were used as source
materials; the synthesis was performed at 65

C for 5 h. The synthesized structures were


characterized in detail using various analytic tools such as SEM, TEM, XRD, size anal-
ysis, and measurements based on BET theory (so-called from the originators, Brunauer,
Emmett, and Teller). The average diameter of the CuO microspheres obtained was
2.8 jm. From the SEM structural characterization, the investigator observed that the
surfaces of the CuO microspheres were made of pin-like nanostructures (with a pin
diameter of 95 nm and a pin length of at least 600 nm); from these ndings, the author
coined the term nanostructured microspheres. The XRD analysis indicated that the CuO
nanostructured microspheres were monoclinic lattices. The BET measurements showed
that the specic surface area of the CuO nanostructured microspheres was 56.8 m
2
g
1
.
All peaks in the XRD pattern obtained for the products synthesized could be indexed
to monoclinic CuO lattices (JCPDS: No. 5-661), and no peaks from other structures or
impurities were observed. Thus, XRD results conrmed that the CuO nanostructured
microspheres synthesized were high purity. The author proposed that the CuO nanos-
tructured microspheres obtained were formed via interfacial precipitation with droplets
of a solid-stabilized emulsion serving as templates [61].
A simple method for the synthesis of hierarchical CuO microspheres was reported by
Song and colleagues [62]. The synthesis was done under microwave irradiation using
CuCl
2
2H
2
O, N,N,N

,N

-tetramethylethylenediamine (TMEDA), and NaOH as source


materials., A microwave oven (at 100% output power for 2 h) was used in this synthesis.
The XRD pattern conrmed that the products obtained were monoclinic CuO structures
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 17
with lattice constants of u = 4.6837 , | = 3.4226 , c = 5.1288 , and p = 99.54. No
other XRD peaks were detected for the CuO structures synthesized, indicating that the
CuO products obtained were free of impurities. From SEM and TEM observations, the
CuO microspheres synthesized had an average diameter of 1.52.5 nm. The authors con-
cluded that the microspheres were formed via self-assembly of nanoakes (width of
250 nm, length of up to 1 jm); each nanoake was formed via self-assembly of dozens
of smaller nanoplates (width 25 nm). These investigators found that, as the concentra-
tion of TMEDA was increased and the concentration of NaOH was decreased, the mor-
phology of the CuO products changed from simple aggregates of nanoakes to spherical
assemblies (with the external shape of the hierarchical CuO microspheres resembling a
chestnut bur) [62].
Dandelion-shaped CuO nanostructures were synthesized and characterized by Liu and
Zeng [4] and have been reported in the literature by others. In a typical reaction process,
an ethanolic solution of Cu(NO
3
)
2
was mixed with the NH
3
, NaOH, and NaNO
3
solu-
tions, transferred to a Teon-lined stainless steel autoclave, and heated at 100

C180

C
for 224 h. XRD patterns conrmed the nanocrystalline nature and monoclinic symme-
try of the as-prepared CuO dandelion-like nanostructures. From detailed experiments,
the authors observed that the CuO nanostructures synthesized seemed to be built from
small crystalline strips, which were, in turn, build from even smaller 1-D nanoribbons.
The crystalline strips were aligned perpendicular to the spherical surface (i.e., pointing
toward a common center). Thus, the structures seemed like dandelions that were coreless
(i.e., with a hollow cavity); the thickness of the shell wall was about one-third to one-
quarter of the sphere diameter. Regarding the formation of the spherical structures, the
authors suggested that the geometric shapes of the reactants (i.e., building blocks) played
a key role, since no surfactants/emulsions had been used. Further, a simple array of
rhombic CuO crystal strips could easily generate curvature, and the lateral engagement
of these building units could naturally lead to a shell-like structure [4].
The synthesis and characterization of hollow CuO microspheres grown by a simple,
hydrothermal route, in which neither sophisticated techniques nor catalysts/surfactants
were required, was reported by Zhang and colleagues [63]. The synthesis was performed
using copper acetate as a Cu source and hexamethylenetetramine ((CH
2
)
6
N
4
, HMTA)
as a complexing reagent. The authors observed that, under hydrothermal treatment,
the copper acetate (as precursor/reactant) and (CH
2
)
6
N
4
(as indirect template) coopera-
tively controlled the synthesis of hollow CuO microspheres. Regarding the formation of
such structures, the authors predicted that hollow microspheres could be formed with
the assistance of a soft template of gas bubbles of NH
3
produced by decomposition of
(CH
2
)
6
N
4
. Detailed structural observations were performed using various analytic tech-
niques. Results from XRD conrmed that the CuO microspheres grown were mono-
clinic, CuO nanocrystalline structures. As no peaks from other structures or impurities
were observed; hence, the CuO structures formed were high purity. FESEM observa-
tions conrmed that the products grown were spherical microstructures with diameters
of 11.5 jm. The authors also observed that the hollow CuO microspheres were formed
from smaller CuO nanoparticles [63].
Zhang and colleagues demonstrated aggregation-based formation of nanostructured
CuO particles achieved through a simple oxidation of Cu metal in formamide and sub-
sequent hydrolysis of the resulting Cuformamide complexes in aqueous solution [64].
The XRD pattern conrmed the monoclinic, crystalline nature of the CuO nanostruc-
tures obtained. By TEM, the structures appeared similar to ellipsoids, having the average
lengths and widths of 130 nm and 45 nm, respectively; hence, the products were
called nanoellipsoidal particles. By high-resolution TEM, the nanostructures synthesized
appeared to be accumulations of several thousand small (67 nm) CuO nanoparticles;
the small primary nanoparticles appeared to be connected to one another to form the
larger, secondary ellipsoidal architecture with recognizable boundaries/voids between
the component subunits. The anisotropic growth rates of the ellipsoidal particles were
ordered as | > u > c or [010] > [100] > [001]. This analysis included different particle
aggregation potentials; aggregation/attachment occurred preferentially between the (010)
18 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
lattice planes of the nanoparticles. Importantly, the anisotropic growth of the nanostruc-
tures obtained [64] was different from the classical crystal growth of CuO nanowires
along the [111] direction [65, 66]. Additionally, the authors observed that the selective
adsorption of formamide molecules on various crystallographic planes could play an
important role in the anisotropic growth of ellipsoidal nanoarchitectures [64].
There are several reports in the literature on the growth and properties of ower-
shaped CuO nanostructures [4, 6769]. In this section, we assemble information on the
ower-shaped CuO nanostructures reported to date.
Flower-shaped CuO nanostructures were synthesized and characterized by Li and col-
leagues [67]. In a typical reaction process, equimolar solutions of Cu(NO
3
)
2
and HMTA
were transferred to a vial and heated at 90

C120

C for 3 h. The substrate (i.e., Si, indium


tin oxide [ITO]coated glass) was placed at the bottom of the vial. Reactions were per-
formed over a range of pH; to adjust the pH of the reaction solution, nitric acid (HNO
3
)
or NaOH was added during the reaction. The investigators studied the inuence of
reaction parameters on the morphologies of the CuO nanostructures produced. Reactant
concentration, pH, synthesis temperature, and presence of seed layers on the substrate
signicantly affected the morphologies obtained. Additionally, adhesion of the products
(i.e., CuO nanostructures) onto the substrate was also affected by reaction parameters.
In the absence of a seed layer or at low concentrations of reactants, spherical assemblies
of CuO nanoplatelets were obtained. The shape of the nanoplatelets and the density of
their packing were strongly affected by pH, synthesis temperature, and concentrations
of reactants. Synthesis from solutions with high concentrations enabled fabrication of
nanostructured lms on the substrate; the morphology and adherence of such lms to
the substrate was dependent on pH and synthesis temperature. Results of detailed struc-
tural analyses conrmed the monoclinic, crystalline nature of the CuO nanostructures
synthesized [67].
Zhu and colleagues demonstrated the synthesis and characterization of ower-like
CuO nanostructures by hydrolyzing copper acetate in an aqueous solution in the absence
of any surfactants/additives [68]. All diffraction peaks in the XRD pattern of the as-
prepared CuO nanostructures were well consistent with monoclinic, crystalline CuO
structures. No peaks due to for other structures or impurities were detected in the XRD
patterns; these results conrm that the CuO products synthesized were pure. By SEM
and TEM analyses, the products synthesized were owery spheres with diameters of
0.40.8 jm. Moreover, the ower-shaped structures were composed of many intercon-
nected, needle-like crystallites with diameters of 1015 nm [68]. Regarding the spherical
appearance of the products synthesized, it was proposed that the geometric shape of
the reactants (i.e., building blocks) played a key role, since no surfactants/emulsions
were used [4, 68]. A simple array of needle-shaped CuO crystallites could easily gener-
ate curvature, and the lateral engagement of these building units could naturally lead to
spherical structures [4, 68].
A simple, solution route to prepare ower-shaped CuO nanostructures was developed
by Yang and colleagues [69]. For the growth of ower-shaped nanostructures, Cu(NO
3
)
2
and NaOH were used as starting materials. By controlling the molar ratio of NaOH to
Cu(NO
3
)
2
, reaction temperature, and concentration of the starting NaOH solution, CuO
nanoribbons or nanorods and their assemblies into hierarchical structures were synthe-
sized. For example, 1-D CuO nanorods/nanoribbons were prepared with a molar ratio
of NaOH to Cu(NO
3
)
2
of 40; when the molar ratio of NaOH to Cu(NO
3
)
2
was increased
to 200, the products were spherical or ower-like hierarchical assemblies, which were
built up from 1-D CuO nanorods/nanoribbons. In addition, ethanol greatly inuenced
the morphologies of CuO nanostructures and enhanced the dehydration rate of Cu(OH)
2
by reducing the surface tension of the reaction solution [69].
Vaseem and colleagues also demonstrated the growth and characterization of ower-
shaped CuO nanostructures prepared by a simple, solution process [69]. These inves-
tigators heated Cu(NO
3
)
2
, NaOH, and HMTA at 100

C for 3 h without the use of any


complex apparatus or reagents [70]. Results of morphological investigations by FESEM
revealed that the ower-shaped nanostructures were monodisperse and high density.
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 19
FESEM images also revealed that the owers were built up from many triangular petals.
The diameters of the petals varied from the bases to the tips; that is, petals exhibited
sharpened tips with wider bases. These petals were connected to each other through
their wider bases and were rooted in one center to form the ower-like morphol-
ogy. The typical length of one petal was 600800 nm; the diameter at the base and
tip were 150 nm (range 50 nm) and 50 nm (range 20 nm), respectively. The full
array of a single ower-shaped structure was 23 jm. Interestingly, it appeared that
these nanostructures were formed by the accumulation of many layers, with each layer
containing several petals. In addition, the sizes of the petals differed from the upper
portion to the lower portion of the nanostructures; this arrangement created beautiful
ower-like structures (Figs. 6(A and B)). Detailed structural characterization by TEM and
HRTEM conrmed the monoclinic, crystalline nature of the as-prepared ower-shaped
CuO nanostructures (Figs. 6(C and D)). The researchers found that the ower-shaped
morphologies obtained were strongly dependent on several parameters; the concentra-
tion of HMTA, presence/absence of NaOH or HMTA, and reaction time were the most
inuential parameters. The authors conducted extensive studies on these topics [70].
To optimize reaction time for creation of CuO ower-like morphologies, Vaseem and
colleagues performed various time-dependant experiments with reaction times in the
ranging of 0.517 h [70]. Figure 7 presents FESEM images of the time-dependent mor-
phology of CuO nanostructures. During the early stage of the reaction (0.5 h), shuttle-like
morphologies were formed (Fig. 7(A)). But with increasing reaction time (12 h), these
shuttle-like morphologies began to adhere/assemble as somewhat ower-like morpho-
logies (Fig. 7(B)). The authors observed that 3 h was the optimal reaction time to obtain
monodisperse, ower-like morphologies (Fig. 7(C)). Further increases in reaction time
caused the ower-shaped structures to blur. As the best ower-like structures formed
(A) (B)
(C) (D)
Figure 6. Low-magnication (A) and high-magnication (B) FESEM images; low-magnication (C) and high-
magnication (D) TEM images of the ower-shaped CuO nanostructures grown by a solution process over
3 h. Reprinted with permission from [70], M. Vaseem et al., J. Phys. Chem. C. 112, 5729 (2008). 2008,
American Chemical Society. FESEM = eld-emission scanning electron microscopy, TEM = transmission elec-
tron microscopy.
20 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
(A) (B) (C)
(D) (E) (F)
(G) (H) (I)
Figure 7. FESEM images of time-dependant assembly of ower-shaped CuO nanostructures grown via a solu-
tion process using 0.1 M Cu(NO
3
)
2
and 0.05 M HMTA at 100

C (pH = 6.0). (A) 0.5 h; (B) 1 h; (C) 3 h; (D) 5 h;


(E) 7 h; (F) 9 h; (G) 11 h; (H) 14 h; and (I) 17 h reaction times. Reprinted with permission from [70], M. Vaseem
et al., J. Phys. Chem. C 112, 5729 (2008). 2008, American Chemical Society. FESEM = eld-emission scanning
electron microscopy, HMTA = hexamethylenetetramine.
in only 3 h, the authors suggested that, with prolonged reaction time, petals precipi-
tated over already-formed owers and lled gaps between two adjacent petals (Fig. 7(D),
5 h reaction time). By further extending the reaction time (717 h), the authors found
that precipitation at interstices of two petals was increased, and only blurred ower-like
morphologies were obtained (Fig. 7(EI)) [70].
Figure 8 presents FESEM images of ower-shaped CuO nanostructures as a function of
HMTA concentration. During these experiments, Vaseem and colleagues varied HMTA
concentration in the range of 0.0250.2 M, but held constant other variables, such as
temperature (100

C), pH (6.0), time (3 h), and Cu(NO


3
)
2
concentration (0.1 M) [70]. In
general, as the concentration of HMTA increased, blurring/lling (i.e., lling of the spaces
between two adjacent petals) in the ower-shaped nanostructures was observed. The
optimal HMTA concentration for the monodisperse ower-shaped morphologies was
0.05 M. In these experiments, an aqueous solution of 0.1 M Cu(NO
3
)
2
was mixed with
various concentrations of HMTA (0.0250.2 M). Importantly, in the course of mixing
Cu(NO
3
)
2
and HMTA, no immediate precipitation was apparent, but the clear light-blue
solution of Cu(NO
3
)
2
turned turbid with the addition of HMTA. Moreover, to maintain
the pH 6.0 for all reactions, a few drops of 1 M NaOH were included in the solution,
and interestingly, a blue-colored precipitate of Cu(OH)
2
appeared immediate. From this
observation, the authors suggested that Cu(NO
3
)
2
initially reacts with NaOH to forms the
blue precipitate of Cu(OH)
2
via a simple chemical reaction (expressed in reaction (IV)).
Cu(NO
3
)
2
3H
2
O+2NaOH Cu(OH)
2
+2NaNO
3
+3H
2
O (IV)
The authors further suggested that the formation of Cu(OH)
2
was important for growth
of the CuO crystallites, which initially served as building blocks for the formation of
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 21
(A)
(B)
(E) (F)
(C) (D)
Figure 8. FESEM images of HMTA concentration-dependant morphologies of ower-shaped CuO nanostruc-
tures while keeping other reaction parameters constant (0.1 M Cu(NO
3
)
2
, 100

C, 3 h, pH 6.0). (A) 0.025 M,


(B) 0.05 M, (C) 0.075 M, (D) 0.1 M, (E) 0.15 M, and (F) 0.2 M HTMA. Reprinted with permission from [70],
M. Vaseem et al., J. Phys. Chem. C 112, 5729 (2008). 2008, American Chemical Society. FESEM= eld-emission
scanning electron microscopy, HMTA = hexamethylenetetramine.
the nal products; hence, at appropriate heating, Cu(OH)
2
led the formation of CuO
crystallites according to reaction (V).
Cu(OH)
2
A
CuO+H
2
O (V)
At the early stages of the synthesis reaction, only a few drops of NaOH were added;
hence, there would not be enough OH

ions to produce sufcient Cu(OH)


2
. There-
fore, Vaseem and colleagues proposed that, during the reaction, HMTA provided two
things: rst, HTMA was hydrolyzed and OH

ions were produced by reactions (VI) and


(VII) [70]. (It had previously been reported that, at elevated temperature, HMTA could
hydrolyze in the distilled water and slowly generate OH

ions [71].)
(CH
2
)
6
N
4
+6H
2
O 6HCHO+4NH
3
(VI)
NH
3
+H
2
O NH
4+
+OH

(VII)
22 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
Hence, from the reactions (VI) and (VII), an increase in Cu(OH)
2
could be stimu-
lated by HMTA-generated hydroxyl ions; this would be a second step for the generation
of Cu(OH)
2
with reaction (IV) being the rst step. Therefore, a continuous supply of
Cu
2+
ions and OH

ions from the Cu precursor and HMTA, respectively, could lead


to the continuous generation of Cu(OH)
2
, which is nally converted into CuO crystal-
lites and forms the ower-shaped CuO nanostructures [70]. Additionally, it had pre-
viously been observed that the concentration of OH

ions can signicantly affect the


nucleation and growth behaviors (e.g., number of nuclei, concentration of growth units)
of CuO crystals [48]. Therefore, to obtain a specic nanostructure, there should be an
optimal concentration of OH

ions in the solution. In this synthesis, the concentration


of HMTA was varied while keeping other reaction parameters constant, and from the
experimental results, Vaseem and colleagues found that the concentration of HMTA was
important for the growth of the best quality, monodisperse ower-shaped nanostruc-
tures [70]. The concentration of HMTA for the best results from these experiments was
determined to be 0.05 M (Fig. 8(B)). At a lower concentration of HMTA (0.025 M), the
ower-shaped structures obtained were not well developed, and it seemed as though they
were in initial stages of growth (Fig. 8(A)). Moreover, it had previously been reported
that higher OH

ion concentrations could create diffusion layers on certain surfaces of


CuO nanostructures; these layers could produce additional growth anisotropy, allow-
ing only energetically favorable crystallographic planes to grow [41]. Vaseem and col-
leagues also observed this phenomena by increasing the concentration of HMTA; the
quantity of the OH

ions was increased, and overgrowth over the previously formed


ower-shaped structures was observed [70]. By increasing the concentration of HMTA
above 0.05 M, overgrowth increased and lled the vacant spaces between adjacent petals
(Fig. 8(CE)); nally, blunt shaped owers were obtained at highest HMTA concentration
used (Fig. 8(F)). From these results, one can conclude that the HMTA concentration was
(A) (B)
(C) (D)
Figure 9. Low-magnication (A) and high-magnication (B) FESEM images of the CuO structures grown in the
presence of NaOH alone (i.e., without HMTA) using 0.1 M Cu(NO
3
)
2
at 100

C for 3 h. Low-magnication (C)


and high-magnication (D) FESEM images of the CuO structures grown in presence of HMTA alone (i.e.,
without NaOH) by using 0.1 M Cu(NO
3
)
2
and 0.05 M HMTA at 100

C for 3 h. Reprinted with permission from


[70], M. Vaseem et al., J. Phys. Chem. C 112, 5729 (2008). 2008, American Chemical Society. FESEM = eld-
emission scanning electron microscopy, HMTA = hexamethylenetetramine.
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 23
one of the most important/prominent factors affecting the ower-shaped morphology of
CuO nanostructures.
Second (and in addition to providing the OH

ions to the solution), HMTA acted as


an additive that controlled the shape of the nanostructures [70, 72]. To clarify the role
of HMTA, Vaseem and colleagues conducted additional experiments in the absence or
presence of HMTA (keeping constant other reaction parameters such as temperature,
pH, and time) [70]. Figures 9(A) and (B) presents low- and high-magnication images of
CuO structures prepared without HMTA (but with NaOH); the reaction was maintained
at pH 6. It can be seen from these images that blunt-shaped nanopetals were formed
without HMTA; some petals were connected to each other as if in the preliminary stages
of forming ower-like morphologies. On the other hand, in the experiments done in the
presence of HMTA (but without NaOH), the nanopetals formed were well dened with
sharp, clean tips; ower-shaped nanostructures were also obtained (Figs. 9(C and D)).
Hence, from these observations, one can conclude that, in this case, HMTA was acting as
an effective shape-directing agent and helped control the shape of the petals and nally
the ower-shaped structures [70].
3. APPLICATIONS OF COPPER OXIDE NANOSTRUCTURES
3.1. Photocatalytic Properties of Copper Oxide Nanostructures
Vaseem and colleagues reported the growth, photocatalytic, and X-ray absorption near
edge structure (XANES) properties of ower-shaped CuO nanostructures apparently
assembled from triangular leaves with sharp tips and wide bases [73]. Detailed struc-
tural observations conrmed the monoclinic, nanocrystalline structure of the as-prepared
nanostructures. In Figure 10, it is clearly shown that ower-shaped structures were grown
in very large quantity (Fig. 10(A)). A clear view of the single ower-shaped structure
(A) (B)
(C)
Figure 10. Low-magnication (A), high-magnication (B), and very high-resolution (C) FESEM images of
ower-shaped CuO nanostructures grown by a simple, solution process. Reprinted with permission from [73],
M. Vaseem et al., Catal. Commun. 10, 11 (2008). 2008, Elsevier B. V., Amsterdam, The Netherlands. FESEM=
eld-emission scanning electron microscopy.
24 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
is shown in Figure 10(B); this ower consists of many triangular petals. The diameters
of the petals are varied from the base to the tips; that is, petals have wide/bases and
thin/sharp tips. The wide bases of the petals are connected to each other, rooted in one
center, and (nally) constructed into beautiful, ower-like morphologies. Each petal of the
ower-shaped structures appears to be created by accumulation of a large number (per-
haps thousands) of CuO nanoparticles; this is shown in Figure 10(C), a high-resolution
FESEM image. A low-resolution TEM image of the ower-shaped CuO structure is con-
sistency with FESEM observations (Fig. 11(A)). Here, the straight/parallel lattice fringes
of the structures reveal that the petals are single-crystal in nature. Moreover, the spacing
between two neighboring fringes are 0.27 nm, which corresponds to the distance of the
(110) plane of monoclinic CuO (Fig. 11(B)). All the observed peaks in the XRD pattern
can be ascribed as monoclinic CuO structures and yield data close to the reported data
for such structures (JCPDS: 05-0661; u = 4.684 , | = 3.425 , c = 5.129 , p = 99.47

;
space group C2,c). Moreover, the major peaks located at 20 values of 35.6

and 38.8

that indexed as (

111)(002) and (111)(200) planes, respectively, are characteristic of pure,


monoclinic, crystallite CuO structures (Fig. 11(C)). Figure 11(D) shows the FTIR spec-
trum of the as-prepared ower-shaped CuO nanostructures. The sample exhibited weak
absorption bands at 3365 and 1634 cm
1
and strong absorptions bands at 432, 526, and
596 cm
1
. These weak absorption bands could be attributed to the stretching and bending
vibrations of absorbed water and surface hydroxyls, respectively [74]. Moreover, these
strong absorption bands are characteristic of monoclinic CuO structures [75]. Therefore,
the FTIR spectral data conrmed that the products synthesized were pure CuO with
monoclinic structures.
(A)
(C) (D)
(B)
Figure 11. Low-magnication (A) and high-resolution (B) TEM images of ower-shaped CuO nanostructures.
The distance between two adjacent fringes is 0.27 nm; these nanostructures exhibit growth along the [010]
direction for petals grown in ower-shaped nanostructures. Typical XRD pattern (C) and FTIR spectrum (D)
of ower-shaped CuO nanostructures. Reprinted with permission from [73], M. Vaseem et al., Catal. Commun.
10, 11 (2008). 2008, Elsevier. FTIR = Fourier transform infrared, TEM = transmission electron microscopy,
XRD = X-ray diffraction.
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 25
(A) (B)
Figure 12. (A) UV-DRS and (B) plots of o
1,2
versus photon energy l for ower-shaped CuO nanostructures.
Reprinted with permission from [73], M. Vaseem et al., Catal. Commun. 10, 11 (2008). 2008, Elsevier B. V.,
Amsterdam, The Netherlands. DRS = diffuse reectance spectrometry, UV = ultraviolet.
For the determination of bandgap, UV-DRS (diffuse reectance spectrometry) of ower-
shaped CuO nanostructures were carried out and are presented in Figure 12 [73]. From
the observed UV-DRS graph, the bandgap of the ower-like CuO nanostructures was
calculated according to the Eq. (1) and was estimated to be 1.78 eV.
o()l = (l |
g
)
m,2
(1)
In Eq. (1), o is the absorption coefcient, l is Plancks constant, is the frequency of
photons, is a proportionality constant, and m = 1 for direct transitions. Interestingly,
the authors observed that the bandgap of the ower-like CuO nanostructures was larger
than the reported bandgap of bulk CuO materials (1.4 eV) [76]; this difference may be
caused by the quantum size effect in the synthesized nanostructures (i.e., the blueshift
as reported in the literature) [77, 78]. The petals of the synthesized CuO nanostructures
have diameters in the range of 5070 nm; the authors observed that these petals appeared
to be formed by the accumulation of several thousand small CuO nanoparticles [73].
Because of their sharp tips and because they are formed by the accumulation of sev-
eral thousand small nanoparticles, the petals of ower-shaped CuO nanostructures could
exhibit the quantum connement effect [77]. It was also indicated that the ower-like
CuO nanostructure could catalyze the photocatalytic decomposition of an organic pollu-
tant by the formation of an excess of either superoxide radicals (O

2
) or hydroxyl radicals
(OH) or both at the CuO surface [79]. Therefore, the authors examined the degrada-
tion of methylene blue (MB) to conrm the photocatalytic ability of the ower-like CuO
nanostructures [73]. For decomposition of MB, 100 mL of 50 jM MB and 0.1 g of photo-
catalysts were stirred for 30 min in a glass reactor prior to light illumination. A 300 W
xenon lamp (Oriel) was used as a light source, and the light was passed through either
an infrared water lter or no lter. The ltered light was focused onto the reactor. Sam-
ple aliquots were withdrawn by a 1 mL syringe intermittently during the illumination
and then ltered through a 0.45 jm poly(tetrauoroethylene) (PTFE) lter (Millipore,
Billerica, MA). Using UV-visible spectrophotometry, the degradation of MB was mon-
itored (\
max
665 nm) as a function of irradiation time (Fig. 13). The results showed
that, although the photocatalytic ability of CuO nanostructures was 1.5 times lower than
other photocatalysts (e.g., TiO
2
, CuOTiO
2
composite), the ower-like CuO nanostruc-
tures had the same potential for photocatalytic applications as inactive bulk CuO particles
[80]. The reason for the relatively low photocatalytic ability of CuO nanostructures may
be due to recycling of Cu
1+
ions under light on the CuO surface [79] combined with
the large surface area of ower-like CuO nanostructures. Additionally, these investiga-
tors studied the photocatalytic degradation of MB under UV irradiation to gauge the
oxidation capability of CuO nanostructures as well as UV irradiation (i.e., no catalyst)
(Fig. 14) [73]. Figure 14 (inset) illustrates typical time-dependent UV-visible spectra of
26 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
Figure 13. Absorption spectrum of a methylene blue solution in the presence of ower-shape CuO nanostruc-
tures. Reprinted with permission from [73], M. Vaseem et al., Catal. Commun. 10, 11 (2008). 2008, Elsevier
B. V., Amsterdam, The Netherlands.
MB solution during photocatalytic degradation. The spectra of MB in the visible region
exhibits a main peak with \
max
665 nm. Even though the rate of decrease was not fast,
the absorption peaks of MB gradually decreased during the photocatalytic reaction. MB
was also degraded by UV irradiation without the CuO catalyst. The electronic/geometric
structure around Cu in the as-prepared material was characterized with X-ray absorption
ne-structure spectrometry (XAFS). X-ray absorption spectrometry (XAS) measurements
were conducted at beam-line 3C1 of PAL (2.5 GeV; stored current of 130180 mA). The
radiation was monochromatized using a Si(111) double-crystal monochromator, and the
incident beam was detuned by 20% using a piezoelectric translator in order to minimize
contamination from higher harmonics, in particular, the 3rd-order reection of Si crys-
tals. The energy was calibrated by measuring the XAS spectrum of Cu metal foil and
by assigning the rst inection point in the rising portion of the absorption spectra at
8979 eV. The obtained data were analyzed using the IFEFFIT suite of software programs
(University of Chicago) [81].
(A)
(B)
Figure 14. Photocatalytic decomposition of methylene blue (MB; 0.1 g. MB solution, 100 mL of 50 jM MB)
under UV light: (A) no catalyst, (B) CuO catalyst. Light source: 300 W xenon lamp (Oriel, Newport Corporation,
Irvine, CA) equipped with an infrared liquid lter. A is the absorbance of MB (\
max
= 515 nm) and A
0
is the
initial absorbance. Inset indicates a typical spectral change with irradiation time. Reprinted with permission
from [73], M. Vaseem et al., Catal. Commun. 10, 11 (2008). 2008, Elsevier B. V., Amsterdam, The Netherlands.
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 27
(A)
(a)
(d)
(c)
(b) (b)
(c)
(a)
(d)
(B)
Figure 15. Cu K-edge (A) and pre-edge (B) XANES spectra of (a) Cu foil, (b) Cu
2
O, (c) CuO, and (d) as-prepared
ower-shaped CuO nanostructures. Reprinted with permission from [73], M. Vaseem et al., Catal. Commun. 10,
11 (2008). 2008, Elsevier B. V., Amsterdam, The Netherlands. XANES = X-ray absorption near edge structure.
Figure 15 shows the XANES spectra of as-prepared material and copper references (e.g.,
Cu foil/metal, Cu
2
O, CuO). The authors found that the XANES features of as-prepared
material were closer to that of CuO, rather than to those of Cu metal and Cu
2
O [73].
The absorption edge of Cu K-edge XANES was assigned to the main 1s 4p transition.
Cu(0) and Cu(+1) with a d
0
conguration have no hole/vacancy in the 3d orbital, and
Cu(+2) is in a d
9
conguration. Thus, Cu(2+) represents a weak pre-edge peak, meaning
the quadruple allows the 1s 3d transition, and it serves as a characteristic feature for
Cu(+2) because there is no 3d hole/vacancy in Cu(0) or Cu(+1). As shown in Figure 15,
the Cu K-edge XANES spectrum of the as-prepared sample showed a weak pre-edge
peak 89768978 eV comparable that of a Cu(+2)O reference; these data indicated that
the oxidation state of Cu in the as-prepared material was divalent. The position of the
XANES peaks (obtained as the maximum point of the rst derivative function) were
used to determine the oxidation state of CuO; the position shifts toward higher energy
as the oxidation state of the material increases. The positions of the XANES peaks for the
as-prepared sample and copper references were determined to be 8979.0, 8980.4, 8983.6,
and 8983.7 eV for Cu foil, Cu
2
O, CuO, and the as-prepared sample, respectively. These
data provided clear evidence of the divalent oxidation state of copper in the as-prepared
sample [8285].
3.2. Field-Emission Properties of Copper Oxide Nanostructures
Field emission, one of the most fascinating properties of nanostructured materials for
the practical application in vacuum microelectronic devices such as eld-emission dis-
plays, X-ray sources, and microwave devices, has been studied extensively in the past
few decades [8688]. During this time, carbon-based materials, especially carbon nano-
tubes, were studied as promising materials for eld emitters due to their high mechan-
ical stability, good conductivity, low turn-on eld, and large emission currents [8688].
Importantly, it appeared that metal oxide nanostructures emitters, as compared to car-
bon nanotubes emitters, are more stable in harsh environments and have controllable
electrical properties [88].
Regarding eld-emission properties of CuO nanostructures, Hsieh and colleagues
reported the eld-emission properties of various CuO nanostructures grown by a two-step
28 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
fabrication process [29]. The rst step was the rapid formation of Cu nuclei by elec-
trodeposition; the second step was a self-catalytic growth mechanism on the copper
element/polycarbonate substrate [29]. These investigators produced high-purity CuO
nanobers at 1 atm without the use of complicated/sophisticated instruments. A major
advantage of this process was that the vertically oriented CuO nanobers did not required
any adhesive technology on the interface between the bers and the substrate/template
because of the self-growth mechanism. To study eld-emission properties, Hsieh and col-
leagues used three types of CuO nanostructures, i.e., nanorods, nanobers, and nanoparti-
cles (grown using self-catalytic growth processes at 400

C, 500

C, and 600

C, respectively).
Their results indicated that the eld-emission current was signicantly affected by the
morphologies of the CuO nanomaterials. Typical turn-on voltage for CuO nanober arrays
was assessed at 67 Vjm
1
with an emission area of 1 mm
2
. Based on Fowler-Nordheim
plots, the values of work function for the nanober arrays were estimated at 0.562.62 V
and 0.301.39 eV from a two-stage linearity plot. In addition, XPS analysis showed no
obvious changes in the chemical composition of the nanober arrays before and after
eld-emission tests. According to these analyses, the highly ordered CuO nanober arrays
could be promising candidates for eld-emission devices [29].
Chen and colleagues demonstrated the growth and temperature-dependant, eld-
emission properties of CuO nanobelt lms [89]. The aligned CuO nanobelt lms were
synthesized on Cu foils under ambient conditions at room temperature. In a typical
reaction process, rst, an aqueous solution was prepared in a 100-mL glass bottle by
mixing 7.5 mL NaOH solution, 3.0 mL (NH
4
)
2
S
2
O
8
solution, and 4.5 mL deionized water.
Second, a piece of Cu foil (ultrasonically cleaned in acetone and deionized water) was
immersed into the solution, and a lm began to form on the foil. When the color of the
lm became completely deep black, the Cu foil was taken out of the solution, rinsed, and
dried. The typical morphologies of the aligned CuO nanobelt lms were characterized
by SEM; these data revealed that most of the CuO nanobelts grew nearly vertical to the
substrate/foil. The XRD pattern conrmed the crystalline nature of the as-grown CuO
nanobelts. The eld-emission characteristics, including the emission currentapplied eld
plot and emission-site distribution, were studied using the transparent-anode technique.
In addition, the investigators studied the temperature dependence (from room temper-
ature to 750 K) of the eld-emission characteristics. The threshold eld for obtaining
a current density of 10 jAcm
2
was 11 MVm
1
; this eld decreased as temperature
increased and at 700 K was 6 MVm
1
. At a xed eld of 10 MVm
1
, a very large
(i.e., 10
3
) increase in the emission current was observed [89].
Lin and colleagues synthesized and characterized well-ordered CuO nanobrils [76]. In
a typical reaction process, cleaned high-purity Cu foil was electrodeposited by immersing
a two-electrode test cell in an electrolyte solution of 1 M CuSO
4
with additives such
as citric acid, Na(PO
3
)
x
, and gelatin. Cu nuclei were formed, and the growth of CuO
nanobrils from these Cu nuclei occurred in a specially prepared reactor. The growth
reaction was carried out at 500

C under the ow of oxygen (O


2
). The heating rate was
set at 3

Cmin
1
, and the reaction time was set at 1 h. After this gassolid reaction, the
surface of the Cu foil became black, indicating the formation of dense CuO nanobril
arrays. Results of detailed analyses revealed the nanocrystalline CuO nanobrils with
a mean length of 8 jm and an average density of 10
7
10
8
cm
2
. Advanced PL studies
demonstrated that the bandgap energy of the CuO nanobrils was 1.67 eV; this value
corresponded to the narrow range of bandgap energy associated with semiconducting
materials. Furthermore, a redshift occurred when the aspect ratio of the CuO nanobrils
changed. Investigations of electron eld emission revealed that the CuO nanobril lms
were stable and decay resistant during cyclic tests. In addition, the work function of the
CuO nanobrils lm (in the range of 4.14.3 eV) was estimated from Fowler-Nordheim
plots, and light-dot density (bril density) of each lm could reach 10
7
10
8
cm
2
. The
typical turn-on voltage was detected at 6 Vjm
1
with an emission area of 1 mm
2
.
Sung and colleagues reported the synthesis, characterization, and eld-emission prop-
erties of CuO nanowires on Cu-coated Si substrates by a wet chemical process at 70

C
[24]. Results of detailed structural characterization revealed the crystalline nature of the
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 29
CuO nanowires grown. The effect of hydrogen plasma treatment on the eld-emission
characteristics of the CuO nanowires was investigated. The results showed that hydro-
gen plasma treatment enhanced the eld-emission characteristics of the CuO nanowires
showing a decrease in turn-on voltage and an increase in the eld-enhancement factor.
Sung and colleagues demonstrated that hydrogen plasma treatment played an important
role in the improvement of eld-emission characteristics of CuO emitters. These authors
observed that with increasing hydrogen plasma processing time, the eld-enhancement
factor was improved, and this was caused mainly due to reduction of CuO to Cu and
surface cleaning by the hydrogen plasma.
Da Rocha and colleagues demonstrated for the rst time the statistical analysis of data
to evaluate the eld-emission properties of numerous samples of CuO nanostructured
eld emitters [90]. This analysis was done largely in terms of Seppen-Katamuki charts,
eld strength, and emission current. Physical and mathematical models were also derived
to describe the effect of small, electric-eld perturbations in the Fowler-Nordheim equa-
tion and then to explain trends in the data represented in the Seppen-Katamuki charts.
For most of the samples included in this analysis, the eld enhancement factor and emis-
sion area parameters appeared to be very sensitive to variations in the electric eld. These
investigators found that the anodecathode distance was critical in the eld-emission
characterization of samples having a nonrigid nanostructure [90].
3.3. Sensor Applications of Copper Oxide Nanostructures
Zhang and colleagues demonstrated the synthesis and gas-sensor performance of near-
monodisperse Cu
2
O and CuO nanospheres [17]. These fabricated Cu
2
O and CuO
nanosphere sensors exhibited high sensitivity toward alcohol or gasoline. The Cu
2
O sen-
sor was fabricated by the dip-coating of as-prepared Cu
2
Oalcohol colloids onto the
ceramic tube of the sensor body without an additional annealing process (except the
aging of the gas-sensor system). The Cu
2
O and CuO sensors were cycled by increasing
alcohol/gasoline concentrations (10800 ppm) in ambient air at a working temperature
of 210

C. These investigators observed that the as-prepared Cu


2
O nanosphere sensor was
much more sensitive than sensors using CuO, Cu
2
O octahedral microparticles, or Cu
2
O
nanoparticles. The response time of the Cu
2
O sensor was only 15 s to alcohol and 25 s
to gasoline. The resume time of the Cu
2
O sensor also was only 30 s for alcohol and
45 s for gasoline. The detection limit of the as-prepared Cu
2
O sensor was as low as
several parts per million when detecting certain kinds of gas/fumes (e.g., alcohol, gaso-
line). Perhaps the sensitivity difference between Cu
2
O and CuO nanosphere lms was
attributable to the reaction mechanisms of Cu
2
O versus CuO to detected gases. When in
contact with a reducing gas (i.e., electron donator), such as alcohol or gasoline, the neg-
atively charged oxygen (i.e., O

, O
2
) adsorbed on the Cu
2
O or CuO nanosphere surface
reacted. The reaction between the reducing gas and O

/O
2
led to a decrease in the car-
rier hole density in the surface-charge layer and an increase in the Cu
2
O/CuO resistance.
Specically, below the working temperature of 210

C, the as-prepared Cu
2
O nanospheres
adsorbed ambient oxygen (O
2
) more easily than the as-prepared CuO nanospheres; thus,
the surface of the Cu
2
O nanospheres was covered with a higher density of active (i.e.,
negatively charged) oxygen (i.e., O

, O
2
). Unlike the Cu
2
O formation in the body of the
nanosphere, the surface Cu
2
O formation was easily changed to become Cu
2
O
2x
(where
0 - x - 1), an active transient state more favorable for interaction with reducing gases.
Li and colleagues reported the synthesis and gas-sensing properties of CuO nanoparti-
cles and nanoplates [91]. CuO nanoparticles were prepared by a thermal decomposition
method; CuO nanoplates were prepared by a hydrothermal method. In a typical syn-
thesis of CuO nanoparticles, Cu(NO
3
)
2
3H
2
O, glucose, and polyvinylpyrrolidone (PVP)
were dissolved in distilled water and stirred for several minutes; the reaction mixture
was heated to 90

C; and then 1 M NaOH was added until the color of the reaction mix-
ture changed to reddish brown. After the reaction cooled to room temperature, the solid
was collected; the resulting products were transferred to a mufe furnace and heated at
500

C for 4 h to yield a black powder. To prepare CuO nanoplates, Cu(NO


3
)
2
and HMTA
30 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
were dissolved in distilled water. The resulting mixture was transferred to a Teon-lined,
stainless steel autoclave and heated at 180

C for 12 h. After the reaction mixture cooled,


the black precipitate was washed and then dried at 60

C for 10 h. The sensitivity of the


as-prepared CuO nanostructures to nitrogen dioxide (NO
2
) gas and alcohol at varied tem-
peratures was studied systematically. The results revealed that the as-prepared products
had high sensitivity and quick response/recovery times. These investigators also found
that the suitable operating temperatures for CuO nanoplates to detect NO
2
or alcohol
were 200

C and 350

C, respectively. Moreover, the sensitivity of CuO nanoplates was


higher than that of CuO nanoparticles [91].
Zhang and colleagues reported a simple, liquidsolid approach for growing large-
scale, uniform, epitaxially oriented, single-crystal CuO nanowires on the surface of Cu
nanostructures (i.e., Cu nanodendrites) at room temperature [92]. FESEM morphologi-
cal studies of the products grown revealed that the nanowires were 5080 nm wide at
the root and 300400 nm long. Detailed structural studies revealed the crystalline nature
of the nanowires; each structure was a single-crystal nanowire with a growth orienta-
tion in the [110] direction. XRD and EDX analyses revealed that the CuO nanowires
only cover the surface of the dendritic Cu nanostructures; that is, the Cu nanodendrites
formed the centers of the CuCuO nanocomposites. Electrochemical studies revealed that
the CuCuO nanocomposites had a stronger ability to promote electron transfer than
CuO nanowires/nanoparticles alone because of the electrical conductivity of Cu, which
was still in the center of the composite. As a result, the CuCuO nanocomposites had
not only a larger surface for electrochemical probe molecules to reach the active sites but
also supplied more efcient transport/passage for electrical signaling. A glassy carbon
electrode modied with the CuCuO nanocomposite and Naon (a sulfonated tetrauo-
roethylene copolymer) responded more sensitively than an electrode modied with CuO
nanowires (or CuO nanoparticles). The CuCuO nanocomposite electrode exhibited a
linear response to H
2
O
2
with a linear range up to 500 nmol L
1
and a detection limit
of -50 nmol L
1
. An optimized glucose sensor (with immobilized glucose oxidase) dis-
played a linear range up to 500 nmol L
1
and a detection limit of -50 nmol L
1
.
Cao and colleagues demonstrated the formation of hierarchical CuO nanostructures
and their application in As removal [93]. In a typical process, copper acetate was added
to ethylene glycol to form a cloudy solution that was stirred and heated to 150

C.
The cloudy mixture turned clear and then became opaque again, indicating the for-
mation of an organo-copper precursor/intermediate. The product was collected after
centrifugationredispersion cycles with alcohol and calcined at 400

C for 2 h to obtain
crystalline CuO structures. The structures synthesized consisted of nanometer-sized CuO
crystals self-organized into micro-sized monoliths with hierarchical architectures. The
authors observed that by changing the reactant concentration, the structure and mor-
phology of the product were readily tunable, varying from simple microplates to well-
organized, doughnut-like structures and to multilayered microspheres. The as-grown
doughnut-like CuO structures were used for arsenic(III) removal with high capacity.
In addition, these structures can be easily separated/recovered and recycled for water
treatment processes. Moreover, the high specic surface area of the hierarchical CuO
structures provided an effective way to prevent aggregation and greatly facilitated
separation/recovery recycling during practical applications/operations [93].
4. GROWTH AND PROPERTIES OF NICKEL OXIDE AND
HYDROXIDE NANOSTRUCTURES
Nickel hydroxide (Ni(OH)
2
), as one of the most important transition-metal hydroxides,
has received increasing attention due to its exotic properties and extensive applications;
examples of applications include as an active material in positive electrodes and in Ni-
based, alkaline rechargeable batteries (especially because Ni(OH)
2
has the characteristics
of high power density, excellent cyclability, high specic energy, and low toxicity) [94].
As far as nickel oxide (NiO) is concerned, it is p-type, wide-bandgap material that can be
used as a transparent, p-type semiconducting layer [95]. As an important material, NiO
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 31
can be used in various applications, such as magnetic materials, catalysts, electrochromic
lms, fuel-cell electrodes, selective gas sensors, active optical bers, and battery elec-
trodes [9698]. It has been observed that the performance of Ni-based, alkaline recharge-
able batteries and other devices depends on the structural and morphological features of
Ni(OH)
2
and NiO; hence, considerable work has been done to prepare and investigate
nanocrystalline Ni(OH)
2
and NiO structures [94100].
A variety of 1-D NiO and Ni(OH)
2
nanostructures (e.g., nanowires, nanorods, nano-
tubes) have been synthesized and reported in the literature [101103].
Wu and colleagues reported the synthesis of NiO nanowires (modied by PVP)
by a simple, solution-phase process [101]. These authors reported the growth of NiO
nanowires with diameters of 40100 nm and the expected ratio (i.e., length vs. diameter)
in the range of 5490. The NiO nanowires grown exhibited unique photoluminescence
features and displayed a signicant UV luminescence. XANES analysis was used to
characterize the local Ni environment and identify the electronic structure. Comparing
experimental and theoretical spectra at Ni and O K-edges, these investigators detected
lattice distortion via the analysis of the characteristic pre-edge features and the multiple
scattering structures detected in XANES spectra. These authors also explained the cor-
relation between experimental features and the disordered/distorted local structures of
the nanowires grown.
Xu and colleagues reported the synthesis and characterization of NiO nanorods by
the thermal decomposition of a NiC
2
O
4
precursor [102]. For the synthesis process, nickel
acetate (Ni(CH
3
COO)
2
2H
2
O), oxalic acid (H
2
C
2
O
4
2H
2
O), nonyl phenyl ether (9)/(5)
(NP-9/5), and NaCl were used. In a typical reaction process to prepare the precursor, the
nickel acetate and oxalic acid were mixed in 1:1 molar ratio with 5 mL of NP-9/5 in a
mortar, ground for several minutes, and kept in a thermostatic oven at 50

C60

C for 6 h;
the intermediary product was washed several times. This intermediary product was the
NiC
2
O
4
precursor for the fabrication of NiO nanorods. NiC
2
O
4
was simultaneously mixed
with NaCl powder and ground for several minutes. The ground mixture was annealed
at 900

C for 2 h in an alumina tube. After heat treatment, the as-prepared product was
washed and characterized. Detailed structural characterization conrmed the crystalline
nature of the NiO nanorods grown. The NiO nanorods grown had diameters of 1080 nm
and lengths of several micrometers. The as-prepared NiO nanorods were structurally
uniform and single crystals. The growth mechanism of the NiO nanorods was most likely
controlled by a vaporsolid growth mechanism. The authors observed that the surfactant
NP-9/5 was of critically importance for the formation of NiO nanorods [102].
Homogeneous, ordered arrays of NiO nanotubes were fabricated by metallorganic,
chemical vapor deposition (MO-CVD) using a template and Ni(tta)
2
tmeda (tta =
2-thenoyltriuoroacetone, tmeda =TMEDA) as the source material [103]. Malandrino and
colleagues observed from the XRD pattern that the NiO nanotubes formed had the cubic
structures. SEM images of the NiO nanotubes (after removing the template) indicated
the formation of well-ordered nanotube arrays. Using TEM, the authors also observed
that the nanotubes grown were open on both ends. Moreover, the optical properties
of the NiO nanotubes, determined using spectroscopic ellipsometry, strongly depended
on the polarization direction of the light electric eld; these results indicated optical
anisotropy related to the vertical alignment of the nanotubes. The authors suggested that
their approach to preparing the nanotubes had the advantage of being a simple, easy
method for the production of nanostructures such as freestanding, NiO nanotube arrays
on a large scale/area. Moreover, the nanotubes synthesized could be used either directly
embedded in the template or in a freestanding mode since the template could be easily
removed using selective chemical etching [103].
Wang and colleagues reported a new, large-scale method to fabricate single-crystal NiO
nanoplatelets with a unique porous structure [104]. NiO nanostructures were synthesized
by thermal treatment of the precursor, Ni(OH)
2
nanoplatelets, at 400

C for 2 h in air. By
TEM analysis, the shape of the precursor Ni(OH)
2
nanoplatelets was sustained during
the thermal-decomposition process to form the NiO nanoplatelets. The surface of the
NiO nanoplatelets was the {111} crystal plane of the cubic NiO nanostructure, and the
32 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
adjacent edges (with 120

angles) should correspond to the {110} and {101} crystal planes.


The electron diffraction (ED) patterns obtained for different individual nanoplatelets
conrmed the single-crystal nature of the as-grown NiO nanoplatelets. Results of elec-
trochemical tests revealed that these NiO nanoplatelets exhibit improved cyclicity and
reversible capacity in contrast to other NiO nanoparticles. These results suggest that the
unique, platelet-like porous structure is benecial to the electrochemical performance of
NiO nanostructures [104].
Powders of single-crystal p-nickel hydroxide (p-Ni(OH)
2
) nanosheets with hexago-
nal structure were synthesized by Liang and colleagues using a hydrothermal method
at 200

C with nickel acetate (Ni(CH


3
COO)
2
) as the Ni source and aqueous NH
3
as
both an alkaline and complexing reagent [105]. The authors observed that yields
of p-Ni(OH)
2
nanosheet powders were >92.4%. From XRD patterns, the as-grown
p-Ni(OH)
2
nanosheets were well crystallized with hexagonal structures. By TEM, the
authors observed that the p-Ni(OH)
2
nanosheets had irregular morphologies with sizes
in the range of 25160 nm; many nanosheets were irregularly hexagonal with 120

angles between adjacent edges. The authors suggested that the surfaces of the nanosheets
would correspond to the {00-01} crystal planes of the hexagonal p-Ni(OH)
2
structures,
that the 120

angles would correspond to those of the {10-10} and {01-10} crystal planes,
and that the edges would correspond to those of the {10-10} and {01-10} crystal planes.
In addition, the authors also synthesized single-crystal NiO nanosheets by a thermal
decomposition method at 400

C for 2 h using single-crystal p-Ni(OH)


2
nanosheets as
the precursor. The sheet morphology of p-Ni(OH)
2
nanostructures was sustained after
thermal decomposition to NiO nanostructures [105].
One-dimensional NiO strips, composed of self-assembled NiO nanoparticles, were pre-
pared by Ni and colleagues using a thermal decomposition method with nickel dimethyl-
glyoximate as a starting reagent [106]. The morphology of the precursor was maintained
during the heating process; the size of the strips (i.e., the number of subunit nanoparti-
cles assembled) was tailored by adjusting the calcination temperature. Infrared spectral
analysis indicated that few organic molecules remained after calcination. Raman spectra
revealed that many vacancies existed in the product, which was prepared at a relatively
low temperature.
Al-Hajry and colleagues reported the preparation of ower-shaped p-Ni(OH)
2
nano-
structures and ower-shaped, cubic NiO nanostructures; both types of nanostructures
were composed of thin, nanosheet networks and were prepared via a simple, aqueous
route using nickel chloride (NiCl
2
) and ammonium hydroxide (NH
4
OH) [107]. To pre-
pare the ower-shaped p-Ni(OH)
2
structures, 0.025 M NiCl
2
was prepared in 50 mL of
deionized water; this solution was then heated to 65

C for 4 h with continuous stirring.


Subsequently, a 7 M aqueous solution of NH
4
OH was added dropwise into the NiCl
2
solution at 65

C with vigorous stirring. The pH of the solution was kept at 8.6. The solu-
tion temperature was controlled using a manually adjustable thermocouple. The reaction
was maintained for another 4 h. After the reaction was complete, the greenish precipitate
was collected and washed several times with methanol and deionized water and then
left to dry at room temperature. This as-grown product was calcined at 600

C for 2 h
to prepare the ower-shaped, cubic NiO nanostructures. The two similar products (the
so-called as-grown and calcined products) were structurally characterized using various
analytic methods [107].
The crystallinity and crystal structures of the as-grown and calcined products were
examined by XRD (Fig. 16) [107]. For the rst sample, all the diffraction peaks could be
indexed as pure hexagonal p-Ni(OH)
2
structures with cell constants of u = 3.126 and
c = 4.662 . No other peaks that might indicate impurities (such as o-Ni(OH)
2
) or other
structures were observed in the pattern; these results conrmed the as-prepared specimen
as a pure, single-crystal p-Ni(OH)
2
structure (Fig. 16(A)). For the second sample, all
peaks in the XRD pattern were fully matched with pure, cubic-structured crystalline
NiO nanostructures (Fig. 16(B)). Several peaks were observed in the XRD pattern at
20 = 37.1

, 43.3

, 62.8

, 75.2

, and 79.3

; these were assigned to the {111}, {200}, {220},


{311}, and {222} crystal planes, respectively. The indexed peaks are fully consistent with
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 33
(A)
(B)
Figure 16. Typical XRD patterns of as-grown ower-shaped p-Ni(OH)
2
nanostructures (A) and as-grown NiO
nanostructures (B). Reprinted with permission from [107], A. Al-Hajry et al., Superlattice Microst. 44, 216 (2008).
2008, Elsevier. XRD = X-ray diffraction.
cubic-structured crystalline NiO nanostructures (JCPDS: 47-1049). In addition, no peaks
for any impurities (e.g., o-Ni(OH)
2
, p-Ni(OH)
2
) or other structures were observed in the
XRD pattern; these results further conrm the pure, crystalline structure of the cubic NiO
nanostructures [107].
Figure 17, panels AC, depict the general morphologies of the as-grown p-Ni(OH)
2
nanostructures. Large-quantities of ower-like structures composed of thin nanosheets
were obtained (Figs. 17(A and B)) [107]. The high-magnication FESEM image revealed
that the nanosheets were connected to each other in ways that formed network-like
structures (Fig. 17(C)). The average dimensions of the nanosheets were 1.52.0 jm, and
the typical thickness was 5070 nm. The nanosheets were joined to each other to form
ower-like structures that exhibited spherical morphologies. Typically, the full array of
a single ower-shaped structure was 35 jm with a width of 56 jm. Figure 17(D)
presents typical TEM images of thin sheets of the p-Ni(OH)
2
nanostructures. As can
be seen from the low-magnication TEM image, the nanosheets were transparent and
constructed with Ni(OH)
2
nanobers 15 nm in diameter and 400 nm in length. The
authors also observed that some of the nanosheets appeared to be rolled to become
needle-like structures. HRTEM images clearly exhibited the distance between two lattice
fringes (i.e., 0.27 nm), which corresponded to the {01-

10} crystal planes of p-Ni(OH)


2
(Fig. 17(D), inset) [107].
Figure 18, panels A and B, present low-magnication FESEM images of ower-shaped,
cubic NiO nanostructures obtained by calcination of as-grown p-Ni(OH)
2
nanostruc-
tures at 600

C in air. The general morphologies and dimensions of the NiO products


obtained were similar to the initially grown ower-shaped p-Ni(OH)
2
nanostructures
[107]. Figure 18(C) presents an high-resolution FESEM image of the ower-shaped, cubic
NiO nanostructures. Some breakage in the nanosheet networks was observed at the
outer surfaces of the ower-shaped, cubic NiO structures [107]; this breakage was most
probably due to the higher calcination temperature (600

C). Figure 19(A) shows a typ-


ical FTIR spectrum of as-prepared, ower-shaped p-Ni(OH)
2
nanostructures. A sharp,
narrow peak was observed at 3637 cm
1
[107]; this peak was characteristic for OH
stretching vibration and conrmed the brucite-like (Mg(OH)
2
) structure (i.e., hexagonal
lamellar structure) of the p-Ni(OH)
2
nanostructures [108]. The other peaks observed at
3450 and 1633 cm
1
[107] were assigned to stretching vibration and bending vibration
of adsorbed water molecules, respectively. The strong peak at 520 cm
1
and the small
peak at 460 cm
1
corresponded to the oOH of hydroxyl groups and NiO stretching
34 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
(A) (B)
(C) (D)
Figure 17. Typical low-magnication (A, B) and high-magnication (C) FESEM images plus low-resolution (D)
and high-resolution (D, inset) TEM images of ower-shaped p-Ni(OH)
2
nanostructures composed of nanosheet
networks prepared by a simple, solution process. Reprinted with permission from [107], A. Al-Hajry et al.,
Superlattice Microst. 44, 216 (2008). 2008, Elsevier B. V., Amsterdam, The Netherlands. FESEM= eld-emission
scanning electron microscopy, TEM= transmission electron microscopy.
modes, respectively [108, 109]. The typical FTIR spectrum of the ower-shaped, cubic
NiO nanostructure (Fig. 19(B)) exhibited a composition different from the parent Ni(OH)
2
sample [107]. A sharp, strong peak at 3637 cm
1
was observed corresponding to OH
stretching vibration. Moreover, the two peaks observed at 536 and 460 cm
1
were related
to the pure NiO stretching vibration.
The thermal behavior of ower-shaped p-Ni(OH)
2
nanostructures was examined
by thermogravimetric analysis (TGA) and differential thermal analysis (DTA) [107];
these measurements are depicted in Figure 20. By TGA, the authors observed that the
p-Ni(OH)
2
started to decompose (i.e., lose weight) at 285

C. The major weight loss


occurred rapidly between 300

C and 340

C. The total weight loss was measured as


20.1%, which was in a good agreement with the theoretical value (19.4%) calculated
according to the following endothermic equation: Ni(OH)
2
NiO+H
2
O [95]. Moreover,
the TGA curve revealed that there was no obvious weight loss when the temperature
was >580

C; hence, the authors concluded that p-Ni(OH)


2
nanostructures could be con-
verted to NiO nanostructures at 580

C [107]. This was in agreement with XRD obser-


vations that conrmed the complete transformation of p-Ni(OH)
2
nanostructures into
crystalline, cubic NiO nanostructures (Fig. 16(B)) [107].
For the synthesis of Ni(OH)
2
nanostructures, NH
4
OH was slowly added to an aque-
ous solution of NiCl
2
with continuous stirring at 65

C. The typical chemical reaction


for the formation of Ni(OH)
2
can be explained by a simple chemical equation: NiCl
2
+
2NH
4
OH Ni(OH)
2
+2NH
4
Cl. After the reaction, the by-product, ammonium chloride
(NH
4
Cl), was washed away with deionized water and organic solvents. It is known that
p-Ni(OH)
2
has a brucite-like (Mg(OH)
2
) structure (i.e., hexagonal lamellar structure) and
hence has a tendency to form 2-D lamellar structure [108]. Moreover, under normal pre-
cipitation conditions, these structures present as thin, sheet-like morphologies with the
sheet surfaces perpendicular to the c axis, as sheet-like structures are thermodynamically
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 35
(A) (B)
(C)
Figure 18. Low-magnication (A, B) and high-resolution (C) FESEM images of ower-shaped, cubic NiO struc-
tures obtained by calcination of as-grown p-Ni(OH)
2
nanostructures at 600

C in air. Reprinted with permis-


sion from [107], A. Al-Hajry et al., Superlattice Microst. 44, 216 (2008). 2008, Elsevier B. V., Amsterdam,
The Netherlands. FESEM= eld-emission scanning electron microscopy.
stable along the c-axis direction [110]. Therefore, with prolonged reaction time, many
Ni(OH)
2
nanosheets aggregated and formed the nanosheet networks in the ower-shaped
Ni(OH)
2
nanostructures. Further, NiO nanostructures were formed by the simple decom-
position of Ni(OH)
2
(in nanostructures) into NiO according to the following equation:
Ni(OH)
2
(endothermic) NiO+H
2
O [110].
(B)
(A)
Figure 19. Typical FTIR spectra for as-prepared ower-shaped p-Ni(OH)
2
nanostructures (A) and calcined,
ower-shaped, cubic NiO nanostructures (B). Reprinted with permission from [107], A. Al-Hajry et al., Super-
lattice Microst. 44, 216 (2008). 2008, Elsevier B. V., Amsterdam, The Netherlands. FTIR = Fourier transform
infrared.
36 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
Figure 20. TGA curve of the as-prepared ower-shaped p-Ni(OH)
2
nanostructures. The DTA curve for the
same material is also shown, conrming the transition from p-Ni(OH)
2
nanostructures to NiO nanostructures.
Reprinted with permission from [107], A. Al-Hajry et al., Superlattice Microst. 44, 216 (2008). 2008, Elsevier
B. V., Amsterdam, The Netherlands. DTA = differential thermal analysis, TGA = thermogravimetric analysis.
Zhu and colleagues reported the high-yield synthesis of single-crystal Ni(OH)
2
nanocolumns at the relatively low temperature of 60

C [111]. The authors observed that


each nanocolumn was formed as ordered stacking of oriented, single-crystal Ni(OH)
2
nanosheets. In a typical reaction process, Ni(OH)
2
nanocolumns were synthesized by
adding an aqueous solution of hydrazine (NH
2
NH
2
) to an aqueous solution of NiCl
2
and reacting at 60

C for 24 h. The green precipitate was isolated by centrifugation,


washed using distilled water and absolute ethanol, and dried at 50

C for 4 h. The authors


proposed that the hydrazine used here played a key role in the formation of nanocol-
umn morphology. The authors observed that the sizes and morphologies of the Ni(OH)
2
stacking sheets can be tailored by changing the experimental conditions. In further pro-
cessing, porous NiO nanosheets and nanocolumns were obtained by annealing Ni(OH)
2
nanocolumns at 500

C for 3 h in a mufe furnace in ambient air. The NiO nanocolumns


were unstable and collapsed on further processing; only NiO nanosheets were obtained
upon further annealing at 700

C for another 3 h. The authors observed that the dis-


persed individual porous NiO nanosheets exhibited a higher catalytic activity for CO
oxidation [111].
Bai and colleagues reported the synthesis of sea urchin-like NiO nanoarchitectures by
thermal decomposition of the as-prepared precursors deposited from an aqueous solu-
tion of NiCl
2
in the presence of urea (NH
2
CONH
2
) without using templates or surfac-
tants [112]. Morphological studies revealed that the product was microspheres with a
uniform diameter of 1 jm. The authors observed that the microspheres were com-
posed of numerous densely aligned 1-D nanowires extending radially from the centers
of the nanostructures. The nanowires were not uniform along the lengths and became
sharper toward the tips. Detailed studies revealed that the samples obtained were of high
purity and that the microspheres were polycrystalline, cubic NiO nanostructures; the NiO
nanoarchitectures were composed of numerous nanowires with diameters of 1030 nm.
The authors discussed the possible formation mechanism [112].
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 37
Chen and Gao developed a facile route to synthesize ultrane nanowires, superthin
nanoakes, and uniform nanodisks of p-Ni(OH)
2
using ethanol as the growth media
under hydrothermal conditions [100]. The results showed that the p-Ni(OH)
2
nanowires/
nanoakes/nanodisks grown were single-crystal nanostructures.
Urchin-like NiO nanostructures were prepared by thermal decomposition of the pre-
cursor obtained via a hydrothermal process using urea as a hydrolysis-controlling agent
and PEG as a surfactant [109]. Structural characterization indicated that the urchin-
like NiO nanostructures were high purity. Electrochemical properties of urchin-like NiO
nanostructures were examined by cyclic voltammetry and galvanostatic chargedischarge
measurements. The results revealed that NiO nanostructures calcined at 300

C had a
higher specic capacitance and better capacitive behavior than NiO nanostructures cal-
cined at 500

C.
5. CONCLUDING REMARKS AND FUTURE DIRECTIONS
This chapter presents an up-to-date review on the various morphologies of copper oxide,
nickel oxide, and nickel hydroxide nanostructures synthesized by a variety of fabrica-
tion techniques using various source materials reported in the literature. Even though
a number of studies on the synthesis and characterization of CuO, NiO, and Ni(OH)
2
nanostructures have been reported in the literature, there are still many avenues which
deserve to be explored to gain a better understanding of the these metal oxide nanostruc-
tures. In considering the synthetic point of view, exploration is needed to nd additional
methods by which high-quality metal oxide nanostructures can be obtained at low tem-
perature and low cost. In addition, successful device applications of these metal oxide
nanostructures are still few; hence, in the near future, more effort is needed in the devel-
opment of CuO, NiO, and Ni(OH)
2
nanostructures for the fabrication of efcient devices.
REFERENCES
1. S. Iijima, Nature 354, 56 (1991).
2. X.-M. Lin and A. C. S. Samia, J. Magn. Magn. Mater. 305, 100 (2006).
3. W.-T. Liu, J. Biosci. Bioeng. 102, 1 (2006).
4. B. Liu and H. C. Zeng, J. Am. Chem. Soc. 126, 8124 (2004).
5. H. G. Yang and H. C. Zeng, Angew. Chem. Int. Ed. 43, 5930 (2004).
6. A. Umar and Y. B. Hahn, Appl. Phys. Lett. 88, 173120 (2006).
7. A. Umar and Y. B. Hahn, Nanotechnology 17, 2174 (2006).
8. A. Umar, S. Lee, Y. H. Im, and Y. B. Hahn, Nanotechnology 16, 2462 (2005).
9. A. O. Musa, T. Akomolafe, and M. J. Carter, Sol. Energy Mater. Sol. Cells 51, 305 (1998).
10. J. B. Reitz and E. I. Solomon, J. Am. Chem. Soc. 120, 11467 (1998).
11. H. Wang, J.-Z. Xu, J.-J. Zhu, and H.-Y. Chen, J. Cryst. Growth 244, 88 (2002).
12. M. Goto, A. Kasahara, T. Oishi, Y. Konishi, and M. Tosa, J. Appl. Phys. 94, 2110 (2003).
13. R. V. Kumar, Y. Diamant, and A. Gedanken, Chem. Mater. 12, 2301 (2000).
14. T. L. Hsiung, H. P. Wang, Y.-M. Lu, and M. C. Hsiao, Rad. Phys. Chem. 75, 2054 (2006).
15. P. Poizot, S. Laruelle, S. Grugeon, L. Dupont, and J. M. Tarascon, Nature 407, 496 (2000).
16. J. Morales, L. Snchez, S. Bijani, L. Martnez, M. Gabs, and J. R. Ramos-Barrado, Electrochem. Solid-State
Lett. 8, A159 (2005).
17. J. Zhang, J. Liu, Q. Peng, X. Wang, and Y. Li, Chem. Mater. 18, 867 (2006).
18. T. D. Ewers, A. K. Sra, B. C. Norris, R. E. Cable, C.-H. Cheng, D. F. Shantz, and R. E. Schaak, Chem. Mater.
17, 514 (2005).
19. S. C. Ray, Sol. Energy Mater. Sol. Cells 68, 307 (2001).
20. H. Yang, J. Ouyang, A. Tang, Y. Xiao, X. Li, X. Dong, and Y. Yu, Mater. Res. Bull. 41, 1310 (2006).
21. F. Lanza, R. Feduzi, and J. Fuger, J. Mater. Res. 5, 1739 (1990).
22. F. Favier, E. C. Walter, M. P. Zach, T. Benter, and R. M. Penner, Science 293, 2227 (2001).
23. L. Armelao, D. Barreca, M. Bertapelle, G. Bottaro, C. Sada, and E. Tondello, Thin Solid Films 442, 48 (2003).
24. W.-Y. Sung, W.-J. Kim, S.-M. Lee, H.-Y. Lee, Y.-H. Kim, K.-H. Park, and S. Lee, Vacuum 81, 851 (2007).
25. Y. W. Zhu, T. Yu, F. C. Cheong, X. J. Xu, C. T. Lim, V. B. C. Tan, J. T. L. Thong, and C. H. Sow, Nanotechnology
16, 88 (2005).
26. H. Zhang, X. Zhang, H. Li, Z. Qu, S. Fan, and M. Ji, Cryst. Growth Des. 7, 820 (2007).
27. G. Malandrino, S. T. Finocchiaro, L. Nigro, C. Bongiorno, C. Spinella, and I. L. Fragal, Chem. Mater. 16,
5559 (2004).
28. W. Zhang, X. Wen, S. Yang, Y. Berta, and Z. L. Wang, Adv. Mater. 15, 822 (2003).
38 Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures
29. C.-T. Hsieh, J.-M. Chen, H.-H. Lin, and H.-C. Shih, Appl. Phys. Lett. 82, 3316 (2003).
30. M. Cao, C. Hu, Y. Wang, Y. Guo, C. Guo, and E. Wang, Chem. Commun. 1884 (2003).
31. C. H. Xu, C. H. Woo, and S. Q. Shi, Superlattice Microst. 36, 31 (2004).
32. Y. F. Mei, G. G. Siu, Y. Yang, R. K. Y. Fu, T. F. Hung, P. K. Chu, and X. L. Wu, Acta Mater. 52, 5051 (2004).
33. C. L. Zhu, C. N. Chen, L. Y. Hao, Y. Hu, and Z. Y. Chen, Solid State Commun. 130, 681 (2004).
34. C. H. Xu, C. H. Woo, and S. Q. Shi, Chem. Phys. Lett. 399, 62 (2004).
35. L. S. Huang, S. G. Yang, T. Li, B. X. Gu, Y. W. Du, Y. N. Lu, and S. Z. Shi, J. Cryst. Growth 260, 130 (2004).
36. C. Lu, L. Qi, J. Yang, D. Zhang, N. Wu, and J. Ma, J. Phys. Chem. B 108, 17825 (2004).
37. G. H. Du and G. Van Tendeloo, Chem. Phys. Lett. 393, 64 (2004).
38. Y. Cudennec and A. Lecerf, Solid State Sci. 5, 1471 (2003).
39. T. Yu, X. Zhao, Z. X. Shen, Y. H. Wu, and W. H. Su, J. Cryst. Growth 268, 590 (2004).
40. X. P. Gao, J. L. Bao, G. L. Pan, H. Y. Zhu, P. X. Huang, F. Wu, and D. Y. Song, J. Phys. Chem. B 108, 5547
(2004).
41. Y. Chang and H. C. Zeng, Cryst. Growth Des. 4, 397 (2004).
42. W.-T. Yao, S.-H. Yu, Y. Zhou, J. Jiang, Q.-S. Wu, L. Zhang, and J. Jiang, J. Phys. Chem. B 109, 14011 (2005).
43. M. Kaur, K. P. Muthe, S. K. Despande, S. Choudhury, J. B. Singh, N. Verma, S. K. Gupta, and J. V. Yakhmi,
J. Cryst. Growth 289, 670 (2006).
44. C. Wang, X. Q. Fu, X. Y. Xue, Y. G. Wang, and T. H. Wang, Nanotechnology 18, 145506 (2007).
45. X. Wang, G. Xi, S. Xiong, Y. Liu, B. Xi, W. Yu, and Y. Qian, Cryst. Growth Des. 7, 930 (2007).
46. X. Wen, W. Zhang, and S. Yang, Langmuir 19, 5898 (2003).
47. H. Zhang, S. Li, X. Ma, and D. Yang, Mater. Res. Bull. 43, 1291 (2007).
48. J. Liu, X. Huang, Y. Li, K. M. Sulieman, X. He, and F. Sun, Cryst. Growth Des. 6, 1690 (2006).
49. H. T. Zhu, C. Y. Zhang, Y. M. Tang, and J. X. Wang, J. Phys. Chem. C 111, 1646 (2007).
50. Z. Zhong, V. Ng, J. Luo, S.-P. Teh, J. Teo, and A. Gedanken, Langmuir 23, 5971 (2007).
51. X. Y. Chen, H. Cui, P. Liu, and G. W. Yang, Appl. Phys. Lett. 90, 183118 (2007).
52. H.-M. Xiao, S.-Y. Fu, L.-P. Zhu, Y.-Q. Li, and G. Yang, Eur. J. Inorg. Chem. 2007, 1966 (2007).
53. W. Wang, Y. Zhan, X. Wang, Y. Liu, C. Zheng, and G. Wang, Mater. Res. Bull. 37, 1093 (2002).
54. W. Zhang, S. Ding, Z. Yang, A. Liu, Y. Qian, S. Tang, and S. Yang, J. Cryst. Growth 291, 479 (2006).
55. Q. Liu, H. Liu, Y. Liang, Z. Xu, and G. Yin, Mater. Res. Bull. 41, 697 (2006).
56. Y. Zhao, J.-J. Zhu, J.-M. Hong, N. Bian, and H.-Y. Chen, Eur. J. Inorg. Chem. 2004, 4072 (2004).
57. L. Zheng and X. Liu, Mater. Lett. 61, 2222 (2007).
58. G. Zou, H. Li, D. Zhang, K. Xiong, C. Dong, and Y. Qian, J. Phys. Chem. B 110, 1632 (2006).
59. D. Keyson, D. P. Volanti, L. S. Cavalcante, A. Z. Simes, J. A. Varela, and E. Longo, Mater. Res. Bull. 43,
771 (2008).
60. M. Vaseem, A. Umar, S. H. Kim, A. Al-Hajry, and Y. B. Hahn, Mater. Lett. 62, 1659 (2008).
61. Y. He, Mater. Res. Bull. 42, 190 (2007).
62. H.-C. Song, S.-H. Park, and Y.-D. Huh, Bull. Korean Chem. Soc. 28, 477 (2007).
63. Y. Zhang, S. Wang, Y. Qian, and Z. Zhang, Solid State Sci. 8, 462 (2006).
64. Z. Zhang, H. Sun, X. Shao, D. Li, H. Yu, and M. Han, Adv. Mater. 17, 42 (2005).
65. X. Jiang, T. Herricks, and Y. Xia, Nano Lett. 2, 1333 (2002).
66. W. Wang, Z. Liu, Y. Liu, C. Xu, C. Zheng, and G. Wang, Appl. Phys. A 76, 417 (2003).
67. D. Li, Y. H. Leung, A. B. Djurii c, Z. T. Liu, M. H. Xie, J. Gao, and W. K. Chan, J. Cryst. Growth 282, 105
(2005).
68. J. Zhu, H. Bi, Y. Wang, X. Wang, X. Yang, and L. Lu, Mater. Lett. 61, 5236 (2007).
69. Z. Yang, J. Xu, W. Zhang, A. Liu, and S. Tang, J. Solid State Chem. 180, 1390 (2007).
70. M. Vaseem, A. Umar, S. H. Kim, and Y.-B. Hahn, J. Phys. Chem. C 112, 5729 (2008).
71. K. Govender, D. S. Boyle, P. B. Kenway, and P. OBrien, J. Mater. Chem. 14, 2575 (2004).
72. H. Wang, C. Xie, D. Zeng, and Z. Yang, J. Colloid Interface Sci. 297, 570 (2006).
73. M. Vaseem, A. Umar, Y. B. Hahn, D. H. Kim, K. S. Lee, J. S. Jang, and J. S. Lee, Catal. Commun. 10, 11
(2008).
74. Y. Xu, D. Chen, X. Jiao, and K. Xue, Mater. Res. Bull. 42, 1723 (2007).
75. R. A. Nyquist and R. O. Kagel, Infrared Spectra of Inorganic Compounds, p. 220. Academic Press,
New York, 1971.
76. H.-H. Lin, C.-Y. Wang, H. C. Shih, J.-M. Chen, and C.-T. Hsieh, J. Appl. Phys. 95, 5889 (2004).
77. M. H. Chou, S. B. Liu, C. Y. Huang, S. Y. Wu, and C.-L. Cheng, Appl. Surf. Sci. 254, 7539 (2008).
78. H. Fan, L. Yang, W. Hua, X. Wu, Z. Wu, S. Xie, and B. Zou, Nanotechnology 15, 37 (2004).
79. J. Bandara, I. Guasaquillo, P. Bowen, L. Soare, W. F. Jardim, and J. Kiwi, Langmuir 21, 8554 (2005).
80. Z. Jin, X. Zhang, Y. Li, S. Li, and G. Lu, Catal. Commun. 8, 1267 (2007).
81. M. Newville, J. Synchrotron Rad. 8, 322 (2001).
82. L. S. Kau, D. J. Spira-Solomon, J. E. Penner-Hahn, K. O. Hodgson, and E. I. Solomon, J. Am. Chem. Soc.
109, 6433 (1987).
83. A. Moen, D. G. Nicholson, and M. Rnning, J. Chem. Soc. Faraday Trans. 91, 3189 (1995).
84. J. M. Zuo, M. Kim, M. OKeeffe, and J. C. H. Spence, Nature 401, 49 (1999).
85. J. L. Fulton, M. M. Hoffmann, J. G. Darab, B. J. Palmer, and E. A. Stern, J. Phys. Chem. A 104, 11651 (2000).
86. N. de. Jonge, Y. Lamy, K. Schoots, and T. H. Oostercamp, Nature 420, 393 (2002).
87. S. H. Jo, Y. Tu, Z. P. Huang, D. L. Carnahan, J. Y. Huang, D. Z. Wang, and Z. F. Ren, Appl. Phys. Lett. 84,
413 (2004).
Growth, Properties, and Applications of Copper Oxide and NiO/Hydroxide Nanostructures 39
88. C. J. Lee, T. J. Lee, S. C. Lyu, Y. Zhang, H. Ruh, and H. J. Lee, Appl. Phys. Lett. 81, 3648 (2002).
89. J. Chen, S. Z. Deng, N. S. Xu, W. Zhang, X. Wen, and S. Yang, Appl. Phys. Lett. 83, 746 (2003).
90. M. S. F. da Rocha, T. E. A. Santos, A. C. de Paulo, V. R. Hering, D. den Engelsen, J. H. Vuolo, S. S.
Mammana, and V. P. Mammana, Appl. Surf. Sci. 254, 1859 (2008).
91. Y. Li, J. Liang, Z. Tao, and J. Chen, Mater. Res. Bull. 43, 2380 (2008).
92. X. Zhang, G. Wang, W. Zhang, N. Hu, H. Wu, and B. Fang, J. Phys. Chem. C 112, 8856 (2008).
93. A.-M. Cao, J. D. Monnell, C. Matranga, J.-M. Wu, L.-L. Cao, and D. Gao, J. Phys. Chem. C 111, 18624 (2007).
94. J. Chen, D. H. Bradhurst, S. X. Dou, and H. K. Liu, J. Electrochem. Soc. 146, 3606 (1999).
95. X. Li, X. Zhang, Z. Li, and Y. Qian, Solid State Commun. 137, 581 (2006).
96. M. M. Natile and A. Glisenti, Chem. Mater. 15, 2502 (2003).
97. T. Seto, H. Akinaga, F. Takano, K. Koga, T. Orii, and M. Hirasawa, J. Phys. Chem. B 109, 13403 (2005).
98. P. Lunkenheimer, A. Loidl, C. R. Ottermann, and H. Bange, Phys. Rev. B 44, 5927 (1991).
99. M. Cao, X. He, J. Chen, and C. Hu, Cryst. Growth Des. 7, 170 (2007).
100. D. Chen and L. Gao, Chem. Phys. Lett. 405, 159 (2005).
101. Z. Y. Wu, C. M. Liu, L. Guo, R. Hu, M. I. Abbas, T. D. Hu, and H. B. Xu, J. Phys. Chem. B 109, 2512 (2005).
102. C. Xu, G. Xu, and G. Wang, J. Mater. Sci. 38, 779 (2002).
103. G. Malandrino, L. M. S. Perdicaro, I. L. Fragal, R. L. Nigro, M. Losurdo, and G. Bruno, J. Phys. Chem. C
111, 3211 (2007).
104. X. Wang, L. Li, Y. Zhang, S. Wang, Z. Zhang, L. Fei, and Y. Qian, Cryst. Growth Des. 6, 2163 (2006).
105. Z.-H. Liang, Y.-J. Zhu, and X.-L. Hu, J. Phys. Chem. B 108, 3488 (2004).
106. X. Ni, Q. Zhao, F. Zhou, H. Zheng, J. Cheng, and B. Li, J. Cryst. Growth 289, 299 (2006).
107. A. Al-Hajry, A. Umar, M. Vaseem, M. S. Al-Assiri, F. El-Tantawy, M. Bououdina, S. Al-Heniti, and Y.-B.
Hahn, Superlattice Microst. 44, 216 (2008).
108. L.-X. Yang, Y.-J. Zhu, H. Tong, Z.-H. Liang, L. Li, and L. Zhang, J. Solid State Chem. 180, 2095 (2007).
109. X.-M. Liu, X.-G. Zhang, and S.-Y. Fu, Mater. Res. Bull. 41, 620 (2006).
110. D. Wang, C. Song, Z. Hu, and X. Fu, J. Phys. Chem. B 109, 1125 (2005).
111. J. Zhu, Z. Gui, Y. Ding, Z. Wang, Y. Hu, and M. Zou, J. Phys. Chem. C 111, 5622 (2007).
112. L. Bai, F. Yuan, P. Hu, S. Yan, X. Wang, and S. Li, Mater. Lett. 61, 1698 (2007).

Das könnte Ihnen auch gefallen