Sei sulla pagina 1di 36

CHAPTER 4

ZnO Nanoparticles: Growth,


Properties, and Applications
Mohammad Vaseem1 , Ahmad Umar2 , Yoon-Bong Hahn1
1

School of Semiconductor and Chemical Engineering and BK21 Centre for Future
Energy, Materials and Devices, Chonbuk National University, Chonju 561-756,
South Korea
2
Department of Chemistry, Faculty of Science, Advanced Materials and
Nano-Engineering Laboratory (AMNEL), Najran University, P. O. Box 1988, Najran
11001, Kingdom of Saudi Arabia
CONTENTS
1.
2.
3.
4.

5.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Crystal Structure of ZnO . . . . . . . . . . . . . . . . . . . . . . . . . . .
Nanoparticles of ZnO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Application of ZnO Nanoparticles . . . . . . . . . . . . . . . . . . .
4.1. ZnO Nanoparticles: Bio-Friendly Approach . . . . . . . .
4.2. Solar Cells, Photocatalytic, Photoluminescence, and
Sensor Application of ZnO Nanoparticles . . . . . . . . . .
4.3. Cosmetic Application of ZnO Nanoparticles . . . . . . .
Summary and Future Directions . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1
2
2
19
19
23
33
34
34

1. INTRODUCTION
Today, nanotechnology (NT) is operating in various elds of science via its operation
for materials and devices using different techniques at nanometer scale. Nanoparticles
are a part of nanomaterials that are dened as a single particles 1100 nm in diameter.
From last few years, nanoparticles have been a common material for the development of
new cutting-edge applications in communications, energy storage, sensing, data storage,
optics, transmission, environmental protection, cosmetics, biology, and medicine due to
their important optical, electrical, and magnetic properties. In particular, the unique
properties and utility of nanoparticles also arise from a variety of attributes, including the similar size of nanoparticles and biomolecules such as proteins and polynucleic acids. [1] Additionally, nanoparticles can be fashioned with a wide range of metals
ISBN: 1-58883-170-1
Copyright 2010 by American Scientic Publishers
All rights of reproduction in any form reserved.

Metal Oxide Nanostructures and Their Applications


Edited by Ahmad Umar and Yoon-Bong Hahn
Volume 5: Pages 136

ZnO Nanoparticles: Growth, Properties, and Applications

and semiconductor core materials that impart useful properties such as uorescence
and magnetic behavior [2]. Moreover, unlike their bulk counterparts, nanoparticles have
reduced size associated with high surface/volume ratios that increase as the nanoparticle
size decreases. As the particle size decreases to some extent, a large number of constituting atoms can be found around the surface of the particles, which makes the particles
highly reactive with prominent physical properties. Nanoparticles of particular materials
show unique material properties, hence, manipulation and control of the material properties via mechanistic means is needed. In addition, synthesis of nanoparticles having
uniform shape and size via easy synthetic routes is the main issue in nanoparticle growth.
For the past decade, scientists have been involved in the development of new synthetic
routes enabling the precise control of the morphology and size of the nanoparticles. In
addition, nanoparticle synthesis can be possible via liquid (chemical method), solid, and
gaseous media [315], but due to several advantages over the other methods, chemical
methods are the most popular methods due to their low cost, reliability, and environmentally friendly synthetic routes, and this method provides rigorous control of the size
and shape of the nanoparticles. In general, nanoparticles with high surface-to-volume
ratio are needed, but the agglomeration of small particles precipitated in the solution is
the main concern in the absence of any stabilizer. In this regard, preparations of stable
colloids are important for nanoparticle growth. In addition, nanoparticles are generally
stabilized by steric repulsion between particles due to the presence of surfactant, polymer
molecules, or any organic molecules bound to the surface of nanoparticles. Sometimes
van der Waals repulsion (electrostatic repulsion) also plays important role in nanoparticles stabilization.
With all the issues related to nanoparticle synthesis, there are various types of nanoparticles reported in the literature, e.g., metal nanoparticles, metal oxide nanoparticles, and
polymer nanoparticles. Among all these, metal oxide nanoparticles stand out as one of
the most versatile materials, due to their diverse properties and functionalities. Most
preferentially, among different metal oxide nanoparticles, zinc oxide (ZnO) nanoparticles have their own importance due to their vast area of applications, e.g., gas sensor,
chemical sensor, bio-sensor, cosmetics, storage, optical and electrical devices, window
materials for displays, solar cells, and drug-delivery [1620]. ZnO is an attractive material for short-wavelength optoelectronic applications owing to its wide band gap 3.37 eV,
large bond strength, and large exciton binding energy (60 meV) at room temperature.
As a wide band gap material, ZnO is used in solid state blue to ultraviolet (UV) optoelectronics, including laser developments. In addition, due to its non-centrosymmetric
crystallographic phase, ZnO shows the piezoelectric property, which is highly useful for
the fabrication of devices, such as electromagnetic coupled sensors and actuators [21].

2. CRYSTAL STRUCTURE OF ZnO


Crystalline ZnO has a wurtzite (B4) crystal structure at ambient conditions. The ZnO
wurtzite structure has a hexagonal unit cell with two lattice parameters, a and c, and
belongs to the space group of C46V or P63 mc. Figure 1 clearly shows that the structure is
composed of two interpenetrating hexagonal closed packed (hcp) sublattices, in which
each consist of one type of atom (Zn or O) displaced with respect to each other along
the threefold c-axis. It can be simply explained schematically as a number of alternating
planes stacked layer-by-layer along the c-axis direction and composed of tetrahedrally
coordinated Zn2+ and O2 . The tetrahedral coordination of ZnO gives rise to the noncentrosymmetric structure. In wurtzite hexagonal ZnO, each anion is surrounded by four
cations at the corners of the tetrahedron, which shows the tetrahedral coordination and
hence exhibits the sp3 covalent-bonding. The detailed properties of ZnO are presented in
Table 1.

3. NANOPARTICLES OF ZnO
Due to its vast areas of application, various synthetic methods have been employed to
grow a variety of ZnO nanostructures, including nanoparticles, nanowires, nanorods,

ZnO Nanoparticles: Growth, Properties, and Applications

Figure 1. The hexagonal wurtzite structure model of ZnO. The tetrahedral coordination of Zn-O is shown.
O atoms are shown as larger white spheres while the Zn atoms are smaller brown spheres.

nanotubes, nanobelts, and other complex morphologies [2235]. In the present chapter,
we mainly focus on ZnO nanoparticles synthesized by either the solgel method (solution method) or the hydrothermal method. As the solution method presents a low cost
and environmentally friendly synthetic route, most of the literature for ZnO nanoparticles is based on the solution method. In addition, synthesis of ZnO nanoparticles in the
solution requires a well dened shape and size of ZnO nanoparticles. In this regards,
Monge et al. [36] reported room-temperature organometallic synthesis of ZnO nanoparticles of controlled shape and size in solution. The principle of this experiment was based
on the decomposition of organometallic precursor to the oxidized material in air. It was
reported [37] that when a solution of dicyclohexylzinc(II) compound [Zn(c-C6 H11 )2 ] in
tetrahydrofuron (THF) was left standing at room temperature in open air, the solvent
evaporated slowly and left a white luminescent residue, which was further characterized
by X-ray diffraction (XRD) and transmission electron microscopy (TEM) and conrmed
Table 1. Physical properties of ZnO.
Properties
Lattice parameters at 300 K
a0 (nm)
c0 (nm)
c0 /a0
Density (g/cm3 
Stable phase at 300 K
Melting point ( C)
Thermal conductivity (Wcm1 C1 )
Linear expansion coefcient ( C)
Static dielectric constant
Refractive index
Band gap (RT)
Band gap (4 K)
Exciton binding energy (meV)
Electron effective mass
Electron Hall mobility at 300 K (cm2 /Vs)
Hole effective mass
Hole Hall mobility at 300 K (cm2 /Vs)

Value for an ideal hexagonal structures.

ZnO
0.32495
0.52069
1.602(1.633 )
5.606
Wurtzite
1975
0.6, 1-1.2
a0 : 6.5 cm3 106
c0 : 3.0 cm3 106
8.656
2.008
3.370 eV
3.437 eV
60
0.24
200
0.59
550

ZnO Nanoparticles: Growth, Properties, and Applications

as agglomerated ZnO nanoparticles with a zincite structure having lack of dened shape
and size. Monge et al. used a modied experimental condition using a ligand of long
chain amine, i.e., hexadecylamine (HDA) under an argon atmosphere in addition to the
above-mentioned solution, which resulted in well dened ZnO nanoparticles. It was
observed that shape, size, and homogeneity of the as-synthesized products depend upon
various reactions conditions, i.e., the nature of the ligand, the relative concentration of
reagents, the solvent, the overall concentration of reagents, the reaction time, the evaporation time, and the reaction/evaporation temperature. In addition, when a similar reaction is carried out in dry air, it leads to agglomerated ZnO nanoparticles displaying no
dened shape or size. In an elaborative manner, they analyzed that if the concentration
of reagents in solution increases from 0.042 to 0.125 mol L1 nano-objects of higher aspect
ratio will be formed. Exchanging THF for toluene or heptane produces nanoparticles of
isotropic morphology with mean diameters of 4.6 for toluene and 2.4 nm for heptane.
A slow oxidation/evaporation process in THF (2 weeks) produces only very homogenous nanodisks having size 4.1 nm (Fig. 2(b)). Reducing the reaction time under argon to
5 min prior to oxidation leads to shorter nanorods 58 27 nm in size. Increasing the
reaction temperature leads to isotropic disk-shaped nanoparticles. Exchanging HDA for
dodecylamine (DDA) or octylamine (OA) also leads to disks with mean diameters of 3.0
for DDA and 4.0 nm for OA (Figs. 2(c and d)). In addition, nuclear magnetic resonance
(NMR) studies (Fig. 3) conrmed that throughout the oxidation process, the amine ligand
remains coordinated to zinc and suggested that this coordination participates in controlling the growth of ZnO nanoparticles. Kahn et al. [38] reported the detailed experimental
procedure based on the same synthetic route with different experimental parameters, i.e.,
the effects of solvent, ligand, concentration, time, and temperature. They explained that
the reaction of organometallic complexes with oxygen or moisture leads exothermally
to a hydroxide material, but in this case they did not observe any traces of hydroxide,

(a)

(c)

(b)

(d)

Figure 2. TEM micrographs of ZnO nanoparticles. (a) ZnO nanorods grown under standard conditions. (b) ZnO
nanodisks following a slow oxidation/evaporation process in THF (2 weeks), (c) ZnO nanodisks using DDA
instead of HDA as the stabilizing ligand under standard conditions. (d) ZnO nanodisks using OA instead of
HDA under standard conditions. Reprinted with permission from [36], M. Monge et al., Angew. Chem. Int. Ed.
42, 5321 (2003). 2003, Wiley-VCH Verlag GmbH & Co.

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

(b)

Figure 3. 13 C{1 H} NMR spectra of (a) the free HDA ligand and (b) ZnO nanoparticles coated with HDA.
Reprinted with permission from [36], M. Monge et al., Angew. Chem. Int. Ed. 42, 5321 (2003). 2003, Wiley-VCH
Verlag GmbH & Co.

indicating that both hydrolysis and condensation take place at room temperature. This
can be due to either exothermic oxidation of the organometallic precursor or to the presence of amines, which are bases in solution medium. However, the observation of forming oxide even without amines conrmed that the oxidation reaction of organometallic
precursor is exothermic enough to lead the oxide, and during this process the ligands
must control the shape of the nanoparticles by kinetic control of the oxidation reaction.
In general, the mechanism of nanoparticle synthesis involves three steps, namely nucleation, growth, and ripening. In this case, water molecules could be responsible for the
nucleation step by reacting with the molecular precursor and forming nuclei. In the process, most of the precursor remains intact after this step, and the growth of the particles
can occur when the solution is exposed to moisture and air. Moreover, as-synthesized
ZnO nano-objects dissolved in most of the common organic solvents are luminescent
solutions that can be deposited on various surfaces as a monolayer or as thick layers.
This luminescent solution shows two emission bands: one near-band edge UV emission
at 370 nm and one deep green emission at 585 nm. Interestingly, these two emission
bands are not quenched by the solvents and can be observed at room temperature, both
in solution and in the solid state.
As from the above report, it is conrmed that the solvent has an important effect
on the morphology of ZnO nano-objects. Andelman et al. [39] further elaborated the
solvent effect using different solvents, i.e., trioctylamine (TOA), 1-hexadecanol (HD), and
1-octadecene (OD). It was found that during synthetic process using TOA solvent yields
nanorods, HD solvent yields nanotriangles, and OD solvent yields spherical nanoparticles. Figure 4 shows the typical XRD spectra for nanotriangles, spherical nanoparticles,
and nanorods. The relative intensity of the peaks of nanotriangles and spherical nanoparticles matches the bulk, signifying no preferred orientation. Spherical nanoparticles prepared from octadecene have diameters of 1214 nm. Figure 5 shows the TEM images of
ZnO nanotriangles at various degrees of tilt. The degree of tilt is indicated in the top
left-hand corner. At all angles, the shape remains triangular. As the different capping
agents have varying ability to stabilize certain planes, which leads to different particle morphologies, the case observed here with varying solvents also plays a signicant
role in stabilizing specic crystallographic planes of the growing nanocrystal. The use of
TOA as a solvent leads to rod growth, but when the solvent changed from TOA to OD,
the formation of spherical particles occurred because OD is not a coordinating solvent,
and no crystal favored any growth direction, so the particles grew in a spherical shape.
In addition, one possible reason for the formation of nanotriangles using hexadecanol
as a solvent is due to its moderate coordinating capacity and its relatively weak ligand

ZnO Nanoparticles: Growth, Properties, and Applications

Figure 4. XRD spectra of zinc oxide (a) nanotriangles, (b) spherical nanoparticles, and (c) nanorods. Reprinted
with permission from [39], T. Andelman et al., J. Phys. Chem. B 109, 14314 (2005). 2005, American Chemical
Society.

capacity. Moreover, as-synthesized ZnO particles analyzed by room temperature photoluminescence (PL) measurement indicated that the green band emission is associated with
surface defects and shows a strong dependence of morphology, with suppression of the
green band emission in the case of spherical nanoparticles and nanotriangles (prepared
in TOA/hexadecanol).

Figure 5. TEM images of ZnO nanotriangles at various degrees of tilt. The degree of tilt is indicated in the top
left-hand corner. At all angles, the shape remains triangular. Reprinted with permission from [39], T. Andelman
et al., J. Phys. Chem. B 109, 14314 (2005). 2005, American Chemical Society.

ZnO Nanoparticles: Growth, Properties, and Applications

Another approach was performed by Ayudhya et al. [40] to show the effect of solvent
over the morphology of as-synthesized ZnO products. In their work, single crystalline
ZnO nanoparticles in different aspect ratios were synthesized by a solvothermal method
using various organic solvents. In a typical synthetic process, zinc acetate as a precursor suspended in four various types of organic solvents was heated in an autoclave in the range of 250300 C, depending upon the solvent, used for a 2 h reaction
process. The solvents used in the experiment were alcohols (i.e., 1-butanol, 1-hexanol,
1-octanol, and 1-decanol), glycols (i.e., 1,3-propanediol, 1,4-butanediol, 1,5-pentanediol,
and 1,6-hexanediol), alkanes (i.e., n-hexane, n-octane, and n-decane), and aromatic
solvents (i.e., benzene, toluene, o-xylene, and ethylbenzene). The as-synthesized ZnO
products were characterized by XRD, SEM, and TEM. The typical XRD pattern synthesized in various groups of organic solvents (Fig. 6) conrmed that the crystalline phase
of ZnO was hexagonal without any impurities. The ZnO crystals grow along the same
lattice direction, regardless of the solvent used. SEM micrographs of ZnO nanoparticles
synthesized in glycols are shown in Figure 6. From the SEM images, it is clearly observed
that the products synthesized in glycols produced polyhedral crystals with the lowest
aspect ratios, whereas those synthesized in alcohols produced moderate aspect ratios.
The products obtained using n-alkanes or aromatic compounds as solvents produced
high aspect ratio ZnO nanorods. The morphology of ZnO nanoparticles synthesized
in alcohols strongly depends upon the chain length of the alcohol molecules, whereas
a lesser effect is shown with chain length of glycols, and for n-alkanes and aromatic
solvents, chain length effect is unnoticeable. As for the growth of ZnO nanoparticles,
there is concern that the anhydrous zinc acetate precursor can undergo decomposition
and form ZnO nuclei. The thermal stability of zinc acetate has been reported [4142] to
depend on its interaction with the solvent. Moreover, the dielectric constant of the used
solvent is attributed to the high temperature requirement in the case of n-alkanes and
aromatic compounds having low-dielectric constants compared to glycols and alcohols
having high-dielectric constant required low temperature (250 C). In addition, negatively
charged molecules adsorbed over the positively charged Zn surface of the (0001) facet
of the crystal could retard the growth of crystals in the (0001) direction, which leads to
nonpreferential growth of the crystals. The same phenomenon occurred when glycols as
solvents, having two hydroxyl groups at both ends, could adsorb onto the (0001) surface
of the ZnO crystal, which nally led to the formation of ZnO nanoparticles instead of
ZnO nanorods. On the other hand, alcohols having long chains (i.e., octanol, and decanol)

Figure 6. XRD patterns of ZnO powders synthesized in (a) 1-hexanol, (b) 1,6-hexanediol, (c) n-hexane, and
(d) benzene. Reprinted with permission from [40], S. K. N. Ayudhya et al., Crystal Growth & Design 6, 2446
(2006). 2006, American Chemical Society.

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

(b)

(c)

(d)
(e)

Figure 7. SEM micrographs of ZnO particles synthesized via the solvothermal process in (a) 1,3-propanediol,
(b) 1,4-butanediol, (c) 1,5-pentanediol, and (d) 1,6-hexanediol. Insets in the images are the corresponding TEM
micrographs. (e) Sample of the SAED pattern of the synthesized ZnO. Reprinted with permission from [40],
S. K. N. Ayudhya et al., Crystal Growth & Design 6, 2446 (2006). 2006, American Chemical Society.

show less polarity, which leads to the formation of high-aspect ratio ZnO nanoparticles. Although the dielectric constant of the solvent is the prime reason for the different
morphology of ZnO nanoparticles in this solvothermal synthesis, more detailed characterization and actual mechanism are needed.
To show the effect of acidic and basic solution routes on the morphology of ZnO,
Kawano et al. [43] synthesized ZnO nanoparticles with various aspect ratios. In a typical
synthetic process, ZnO grains and ZnO rods were obtained with various aspect ratios
at 60 C with 2 h reaction in aqueous solution of ZnSO4 via an acidic route (pH 5.6) with
addition of NaOH and in a basic solution of NaOH via a basic route (pH 13.6) with addition of ZnSO4 , respectively. The observed aspect ratios were changed by this synthetic
route, although the nal pH of the solution was the same. The detailed morphological
characterizations were performed by XRD, eld emission scanning electron microscopy
(FESEM), and eld emission transmission electron microscopy (FETEM). XRD analysis
conrmed the wurtzite ZnO type structures with peak broadening in the case of the acidic
route compared to the basic route, which further conrmed the formation of smaller particles via the acidic route. FESEM images also conrmed the formation of ZnO particles
and rods via acidic and basic routes, respectively. Further cumulative undersize distribution of precipitated ZnO particles conrmed that the particle shapes were spherical or
ellipsoidal with diameters of 32 and 44 nm, respectively, via the acidic route at pH 12.8,
which were consistent with the crystallite size calculated by Scherrers formula using
the (100) and (002) diffraction peaks observed in XRD spectra. Although the value of
[OH ]/[Zn2+ ] and the nal pH were the same in the acidic and basic routes, the number
of ZnO nuclei formed via the acidic route was deduced to be much higher than that
obtained via the basic route because the degree of saturation at the initial stage of the
acidic route was extremely high due to the low solubility of ZnO. Thus, most of the
precursor species steeply precipitated as nanograins. On the other hand, ZnO nanorods
formed in the basic route due to limitation of formed ZnO nuclei at the initial stage, and
thus particle size increased via subsequent growth in the progressive stage.
To check the effect of water addition in the precursor-methanol solution for the morphological evolution of ZnO particles, Wang et al. [44] performed reactions based on
hydrolysis of zinc acetate in methanol solvent at 60 C for 24 h and deposited over

ZnO Nanoparticles: Growth, Properties, and Applications

Al2 O3 ceramic plate via the chemical deposition method. As the water/methanol volume
ratio increased, the shape of the ZnO particles changed from irregular particles to
plates and then from plates to regular cones, including the size change from nanoscale to micro-scale. In addition, if the volume of added water increased, the height of
the cones decreased. Addition of water controlled the hydrolysis of zinc acetate and
affected the nucleation process of ZnO signicantly. Moreover, addition of water can

impede the [0001] growth and accelerate the [1100]


growth if the volume ratio of added
water/methanol is equal to or greater than 2:15. In this way, the shape and size of ZnO
can be tailored by adjusting the volume ratio.
Du et al. [45] have given a new reaction to synthesized ZnO nanoparticles with nearly
uniform, spherical morphologies and controlled the size range from 25100 nm via esterication of zinc acetate and ethanol under solvothermal reaction conditions. The reaction temperatures were adjusted from 100200 C for 2448 h in an autocontrolled oven.
In terms of characterization, XRD and TEM analysis conrmed the high crystallinity
and uniform nonagglomerated sphericity of as synthesized ZnO nanoparticles, respectively. By the several reaction conditions, it was conrmed that by changing the reaction
temperature and time, the nanoparticle size can be easily controlled. As for the reaction mechanism, Fourier transform infrared (FT-IR) analysis conrmed the existence of
ethyl acetate during the esterication reaction. In addition, it may be possible that rst
OH anions were produced by esterication reaction between CH3 COO and ethanol
and then zinc cation reacted with as-produced OH to form ZnO under solvothermal
conditions. The presence of ethanol and ester could help to improve the dispersibility of
the as-synthesized ZnO nanoparticles.
Cheng et al. [46] demonstrated the synthesis of ZnO colloidal spheres by the sol
gel method. In a typical synthetic process, they used two types of reaction processes.
In the rst reaction, 0.01 M zinc acetate dihydrate was added to 100 ml diethylene glycol (DEG), and then the reaction solution was heated at 160 C and maintained for 1 h,
which resulted in white colloidal ZnO, treated as the primary solution. In a second reaction process, 0.01 M zinc acetate and various amount of primary supernatant (520 ml)
was added to 100 ml DEG and heated at 160 C for 1 h aging. The resulting ZnO white
colloid produced 50300 nm ZnO nanoparticles, depending upon the amount of primary
supernatant. To check the structural and optical properties, optimal size with 185 nm
ZnO nanoparticles were used. As-synthesized ZnO nanoparticles were characterized by
various analytical tools, i.e., XRD, TEM, FESEM, energy dispersive spectroscopy (EDAX),
Raman spectroscopy, and UV photoluminescence measurement. TEM observation conrmed that spherical 185 nm-diameter ZnO clusters consisted of primary single crystallites ranging from 612 nm. XRD analysis conrmed the hexagonal wurtzite crystallites
of as-grown zinc oxide colloidal spheres and sample post-annealed at 350 and 500 C in
air for 1 h. Raman spectra of as-grown zinc oxide colloidal spheres and post-annealed
samples further conrmed the crystallinity of the products. Moreover, highly efcient
near-band edge UV luminescence was attributed to defect-bound excitons with high
density of states, which was conrmed by using room-temperature PL analyses. This
assumption was further proved by the observation of peak broadening and unchanged
position in low-temperature PL spectra, which is similar to the behavior observed in
the case of ZnO quantum dots. In addition, broad yellow emission and green emission
were observed in room-temperature PL and low-temperature PL, respectively. Further, in
temperature-dependent PL, defects such as oxygen interstitials Oi and oxygen vacancies
V0 dominate the visible emissions of ZnO spheres.
Cheng et al. [47] further reported the enhanced resonant Raman scattering and electronphonon coupling from self-assembled secondary ZnO nanoparticles synthesized by the
same procedure described in the above report. Figure 8 shows the typical TEM images
of ZnO nanoparticles. Figs. 8(a) and (b) show the mean particle size of 185 nm with
spherical shape of ZnO nanoparticles, which consisted of agglomerated primary single
crystallite ranging from 612 nm. The selected area electron diffraction (SAED) spectra
shown in inset of Figure 8(a) conrmed the polycrystalline nature of several secondary
ZnO nanoparticles, while the SAED spectra shown in Figure 8(b) conrmed the single

10

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

(b)

(c)

(d)

Figure 8. TEM images of secondary ZnO nanoparticles recognized of crystalline subcrystals. (a) A typical lowmagnication TEM image and SAED pattern of several uniform ZnO nanoparticles. (b) High-magnication
TEM image of one individual ZnO nanoparticle and its corresponding single-crystal-like SAED spots. (c) and (d)
HRTEM images of the central area and boundary, respectively of one individual ZnO nanoparticle. Reprinted
with permission from [47], H.-M. Cheng et al., J. Phys. Chem. B 109, 18385 (2005). 2005, American Chemical
Society.

crystalline pattern of only one ZnO nanoparticle. This means that the secondary ZnO
nanoparticles are polycrystalline, consisting of much smaller subcrystals of the same
crystal orientation. Figures 8(c and d) further provide much evidence in high resolution TEM (HRTEM) images. It may be possible that van der Waals interaction between
the surface molecules of the nanocrystallites forms the driving force for self-assembly,
and then colloidal nanocrystal can be assembled to form solids. In addition, due to the
block of diethylene glycol, the solvent may behave as a microemulsion system, causing
the individual ZnO subcrystals to grow up separately and nally assemble to form secondary ZnO nanoparticles under the driven force of van der Waals interaction. Figure 8
shows the SEM images of as-synthesized ZnO nanoparticles and samples collected after
post-annealing at 350 and 500 C in air for 1 h. SEM images clearly indicate that during
the heating process, ZnO subcrystals fused with neighboring crystals and grain size grew
accordingly, which was further conrmed by XRD analysis. Moreover, as-grown ZnO
nanoparticles exhibited a phonon red shift in a resonant Raman scattering, compared
with the samples after post-annealing at 350 and 500 C. In addition, the electron-phonon
coupling parameter is clearly extracted from resonant Raman scattering, and an interesting phenomenon of increasing electron-LO phonon coupling was also discovered when
the crystal size of ZnO enlarged after heating treatment. In addition, the Frhlich interaction may certainly play the main role in the coupling of ZnO particles. Finally, blue
shift of UV PL and visible emission induced by interstitial oxygen were also investigated
from as-grown and post-annealed ZnO samples, respectively.

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

11

(b)

(c)

Figure 9. SEM micrographs of secondary ZnO nanoparticles (a) as-grown, (b) annealed at 350 C for 1 h, and
(c) annealed at 500 C for 1 h. Reprinted with permission from [47], H.-M. Cheng et al., J. Phys. Chem. B 109,
18385 (2005). 2005, American Chemical Society.

During the synthesis of ZnO nanoparticles, the inuences of the reactant concentration
were reported by Hu et al. [48]. In a typical process, ZnO nanoparticles were synthesized
by using zinc acetate and NaOH in 2-propanol solution. As the nucleation and growth
were fast in this synthetic process, at longer times the particle size was controlled by
coarsening. In addition, coarsening kinetics were independent of the zinc acetate concentration from 0.51.25 mM at a xed [zinc acetate:NaOH] ratio of 0.625. The width
of the size distribution increased slightly with aging time. Moreover, if the zinc acetate
concentration was xed at 1 mM, the kinetics were independent of variation in the
[zinc acetate:NaOH] ratio from 0.4760.625. The presence of water in the reaction mixture was checked, and it was found that at low water concentration, the nucleation and
growth of ZnO were very slow, which only slightly affected the coarsening kinetics for
water content above 20 mM. Thus, by this synthesis method, it is conrmed that ZnO
nanoparticles are insensitive to the reactant concentration and presence of water.
In another report, Hu et al. [49] explained the inuence of anions on the coarsening kinetics of ZnO nanoparticles. Solution phase synthesis of nanoparticles possesses
coarsening (also known as Ostwald ripening) and epitaxial attachments (or aggregation),
which can compete with nucleation and growth. As a result, particle size distribution can
be modied in the system. If nucleation and growth are fast, coarsening and aggregation
can dominate the time evolution of the particle size distribution. In addition, random
aggregation usually leads to the formation of porous clusters of particles, whereas epitaxial attachment of particles leads to the formation of secondary particles with complex
shapes and unique morphologies. In this report, ZnO nanoparticles were synthesized
from Zn(CH3 COO)2 , ZnBr2 , and Zn(ClO4 )2 in 2-propanol. ZnO nanoparticles synthesized by Zn(CH3 COO)2 , and ZnBr2 in 2-propanol at 55 C for 8.5 h show particles size
65 12 nm and 49 08 nm, respectively, whereas ZnO nanoparticles synthesized by
Zn(ClO4 )2 in 2-propanol at 55 C for 40 min show elongated and irregularly shaped particles via epitaxial attachment of several smaller particles. The rate constant for coarsening

12

ZnO Nanoparticles: Growth, Properties, and Applications

at constant temperature increases in the order Br < CH3 COO < ClO4 indicating that the
rate is dependent on anion adsorption. On the other hand, the temperature dependent
rate constant for coarsening is due to the temperature dependence of the solvent viscosity
and the temperature dependence of the bulk solubility of ZnO.
Vafaee et al. [50] reported the preparation and characterization of ZnO nanoparticles
based on a novel solgel route. As-synthesized ZnO nanoparticle morphology was conrmed by TEM analysis, which shows spherical particles 34 nm in diameter. In a typical
synthesis process, zinc acetate (ZnAc) was used as a precursor, and triethanolamine (TEA)
was used as a surfactant to produce ZnO nanoparticles at 5060 C. With the help of FT-IR
analysis, they proposed that synthesis of ZnO nanoparticles occurred via an intermediate
product called zinc monoacetate, which further assisted the formation of a new complex
and then, via a polycondensation process, produced ZnO nanoparticles. In addition, three
different ratios of both ZnAc and TEA were chosen to determine the best sol, considering
their optical properties. The best sol (0.75 M ZnAc) based on its optical properties was
subjected to analysis by PL spectroscopy. Different shapes of UV (broad peak at 360 nm
with one shoulder at 330 nm) and green peaks (sharp peak at 520 nm) in the PL spectra
of ZnO nanoparticles, synthesized using 0.75 M zinc acetate, suggest the possible use in
monochromatic excitation applications.
To conrm the optimization parameter for the synthesis of zinc oxide nanoparticles,
Kim et al. [51] presented the modied solgel route using the Taguchi robust design
method. In a typical synthetic process, zinc acetate dehydrate, lithium hydroxide monohydrate (LiOH), hydroxypropylcellulose (HPC), and absolute ethanol were used for the
synthesis of ZnO nanoparticles. In this presented work, the molar concentration ratio of
[LiOH]/[Zn(Ac)2 ] was varied in the range of 15, and the concentration of zinc acetate
was xed at 0.05 M. Also, the concentration of HPC dispersant and feed rate of LiOH and
HPC solution were changed in the range of 0.10.4 g and 0.337.0 ml/min, respectively.
After implementing the Taguchi robust design method with an L9 orthogonal array to
optimize experimental condition for the preparation of ZnO nanoparticles, it was found
that the [LiOH]/[Zn(Ac)2 ] molar ratio was the main parameter, showing a prominent
effect on particle size and size distribution of the ZnO nanoparticles. By optimizing the
conditions, the observed size of ZnO nanoparticles was 30 nm with narrow particles
size distribution, conrmed by TEM analysis.
Uthirakumar et al. [52] reported the low temperature solution approach to synthesis
nanocrystalline ZnO nanoparticles from a single molecular precursor without using any
base, surfactant, template etc. via a single step process. In a typical synthetic process,
zinc acetate dihydrate was used as a precursor and methanol was used as a solvent for
synthesizing ZnO nanoparticles at 60 C in 10 h. In addition, similar experiments were
also preformed by using a mixture solvent i.e., dimethylformamide (DMF), toluene, and
THF with methanol, to check the effect of the solvent polarity and water miscibility on
the growth of ZnO nanoparticles. The growth rate was greatly controlled by the presence of a water-immiscible non-polar solvent, which led to the formation of almost pure
ZnO nanoparticles with near UV emission. On the other hand, the water-miscible polar
solvent generates fully defected deep-level emissive ZnO nanoparticles, which agglomerate on standing due to the solvent homogeneity in the reaction mixture. As for the
growth mechanism, the zinc acetate precursor underwent four stages: it was rst solvated
in methanol to form [Zn(MeOH)6 ]+ , then hydrolysis after removal of the intercalated
acetate ions produced [Zn(OH)n2n ], which further polymerized into ZnOZn bridges,
and nally transformed into ZnO. Moreover, it was observed that formation of water
molecules during decomposition of zinc acetate could be responsible for the growth rate
of ZnO nanoparticles. Finally, it was concluded that ZnO crystal growth is more sensitive
to the mixture of solvents, which depends on the miscibility, polarity, and homogeneity
of the precursor in the reaction medium.
Ge et al. [53] reported a simple method to prepare monodispersed ZnO nanoparticles with average size of 52 03 nm at low temperature by ultrasonic treatment. In a
typical synthetic process, 0.88 gm zinc acetate dihydrate was mixed with 80 ml of absolute ethanol in a beaker under magnetic stirring at 70 C. In another beaker, 0.23 gm of

ZnO Nanoparticles: Growth, Properties, and Applications

13

LiOH was dissolved in 80 ml absolute ethanol under magnetic stirring for 20 min. After
this step, the LiOH-ethanol solution was added dropwise into a Zn2+ -containing solution
at 0 C under strong stirring for 1 h, which was further ultrasonically treated for 5 min.
XRD and HRTEM images conrmed the crystallinity and structural morphology, respectively, of the as-synthesized ZnO nanoparticles. In addition, it was reported that with
varying reux time, ZnO nanoparticles can be converted to various aspect ratio ZnO
nanorods via oriented attachment mechanism that were conrmed by the BFDH model
(suggested by Bravasis, Freidel, Donnary, and Harker) and the HP model (proposed by
Hartman and Predok).
Uekawa et al. [54] reported synthesis of ZnO nanoparticles by heating Zn(OH)2 in a
diol solution. ZnO nanoparticles were obtained when Zn(OH)2 was dispersed in ethylene
glycol, 1,3-propanediol, and 1,4-butanediol, which were further treated at temperatures
above 308 K. In particular, if ethylene glycol was used as a solution for Zn(OH)2 dispersion, the synthesized ZnO nanoparticles had average particles size less than 20 nm. Moreover, if the reaction temperature was set at 308 K, the spherical secondary particles with
ZnO primary nanoparticles were obtained. When Zn(OH)2 was heated in 1,3-propanediol
at 308 K for 24 h, the spherical aggregated morphology of the ZnO primary nanoparticles
with average diameter of 9 nm was obtained and if heated in 1,4-butanediol at 308 K for
24 h, the same morphology with average primary ZnO nanoparticle size of 11 nm was
obtained, having interparticle pores in both cases. By measuring N2 adsorption isotherm
at 77 K, it was concluded that ZnO nanoparticles prepared in ethylene glycol at 308 K
contain many interparticle pores with less densely packed spherical aggregated morphology, whereas ZnO nanoparticles prepared in 1,3-propanediol and 1,4-butanediol show
more densely packed primary ZnO nanoparticles. Thus, the formation of ZnO nanoparticles depends greatly not only on the heating temperature but also on the diol solutions
used for preparation.
Lee et al. [55] synthesized ZnO nanoparticles with controlled shapes and sizes by using
a simple polyol method. It was reported that the amount of water and the method of
addition played an important role in determining the characteristics of the synthesized
particles. In a polyol synthetic method, water can induce hydrolysis and condensation
reactions of the Zn precursor when injected into a hot precursor solution maintained
at 180 C, which induces a short burst of homogenous nucleation and leads to growth
of aggregated equiaxial ZnO nanoparticles with average diameter of 24 nm. If a higher
amount of polyvinyl pyrrolidone (PVP)a water-soluble polymeris used, it will lead
to aggregation of free ZnO nanoparticles. In addition, increasing the amount of water
added to the precursor solution enlarges the aspect ratio of the rod-shaped particles
and increases the particle size of the equiaxial particles due to enhanced hydrolysis and
condensation of the Zn ion complex. Moreover, zinc acetate concentration also slightly
inuences the particles size and aspect ratio when water is injected into the hot precursor
solution. Furthermore, the effect of the hydration ratio (ratio of molar concentration of
total water, DI water + hydrated water, to zinc acetate) on the particles characteristics
via the water injection method were also discussed. Varying the hydration ratio from
4 to 8 did not change the particle morphology to a great extent. The particle diameter
increased from 24 to 32 nm, and showed a slight deviation from equiaxial growth with
increasing hydration ratio. Thus, it was concluded that method of water addition, concentration of zinc acetate, and the hydration ratio play important roles in determining
the characteristics of ZnO particles.
Ning et al. [56] reported the synthesis of mesoporous ZnO particles using octadecylamine (ODA) and DDA as templates via the solgel method. Particle size calculated
using Scherrers formula with XRD analysis was 32 nm when processed with ODA and
40 nm with DDA. The densities of ZnO processed with ODA, with DDA, and without
a template were reported as 5.31, 5.37, and 5.42 cm2 /g; respectively. In addition, it was
reported that surface analysis conrmed the porosity of the ZnO particles when processed with ODA and DDA. Moreover, hugely enhanced electroluminescence (EL) was
observed from porous ZnO particles when direct current electric eld from 24.66 V/m
was used. Furthermore, emission intensities of the ZnO sample processed with DDA

14

ZnO Nanoparticles: Growth, Properties, and Applications

and ODA were enhanced 12 times and 20 times, respectively, at a voltage of 4.66 V/m.
The observed EL spectrum shows mainly broad emission peak at 556 nm. The reported
threshold voltage is just 2 V/m. Based on the above analysis, it was conrmed that
porous ZnO particles can enhance EL intensity.
Cozzoli et al. [57] reported the non-hydrolytic route for the synthesis of nearly spherical ZnO nanocrystals with diameter less than 9 nm via a sequential reduction-oxidation
reaction. In a typical synthetic process, ZnO nanocrystals were synthesized in a surfactant mixture of hexadecylamine and oleic acid (OLEA) via a two-step chemical process:
rst hot reduction (at 180250 C) of zinc halide by superhydride (LiBEt3 H) and then oxidation of the resulting product. The reported results conrmed that controlled growth of
ZnO nanocrystal was dependent on OLEA-assisted generation of intermediate metallic
nanoparticles as well as adjustment of oxidation of the metallic nanoparticles using a
mild oxidant, triethylamine-N -oxide, rather than molecular oxygen. Furthermore, the
reported synthetic approach demonstrates that organic-soluble ZnO nanocrystals of low
size dispersion and of stable size can be useful for optoelectronic, catalytic, and sensing
purposes.
Xie et al. [58] reported the low temperature synthesis of uniform ZnO particles with
controllable morphologies. In addition, characteristic luminescence patterns were also
presented. In a typical synthesis process, uniform ZnO particles were synthesized in an
aqueous solution with the presence of TEA below 80 C assisted via sonication. It was
reported that with increasing TEA concentration, one can systematically control the morphology of elongated rugby ball-like ellipsoidal to half-ellipsoidal ZnO particles. FESEM
analysis of many rugby ball-like ZnO particles shows that particles have an average
length of about 620 nm and mean diameter of about 400 nm. By systematic investigation, it was conrmed that formation of rugby ball-like ZnO particles resulted from the
rst growth of a half-ellipsoidal particle followed by the germination and growth of a
second half at its base. Moreover, it was studied with close relationship between particle
characteristics and optical properties with a high spatial resolution cathodoluminescence
(CL) and shows that the ellipsoidal particles are intrinsically encoded with characteristic
barcode-like UV luminescence patterns. Additionally, luminescence spectra can be tuned
by heat treatments at elevated temperatures. By this extensive proof, the authors believe
that well-dened uniform ellipsoidal ZnO particles embedded with unique luminescence
characteristic can hold great potential for use in bioengineering and photonics, such as
biological labeling, multiplexed bioassays, and optical probes inside photonic crystals.
Buha et al. [59] reported the nonaqueous synthesis of nanocrystalline zinc oxide
nanoparticles. In a typical synthesis process, zinc(II) acetylacetonate, as a precursor was
dissolved in the oxygen-free solvent acetonitrile, which was transferred into a Teon
autoclave and then heated at 100 C for 2 days. The resulted products were characterized
by TEM, SEM, and XRD analysis. The TEM micrograph shows the particles size in the
range of 1585 nm, sometimes with well faceted hexagonal morphology. It is interesting
to note that in such a simple reaction, systems like zinc acetylacetonate and acetonitrile
are able to induce the formation of complex structures without any additional structuredirecting agent. Even the large number of organic species detected in nal products
conrmed the complex reaction pathways during the reaction, and these organic components during nanoparticle formation are prerequisite to understanding and controlling
the nonaqueous synthesis of metal oxide materials.
Glaria et al. [60] reported synthesis of ZnO nanoparticles via an organometallic route
and explained that lithium ions act as growth-controlling agent. For the synthesis of
ZnO nanoparticles, solid Zn(c-C6 H11 )2 was dissolved in a THF solution of lithium precursor and OA used as stabilizer. Two different lithium precursors, i.e., Li[N(CH3 )2 ] and
Li[N(Si(CH3 )3 )2 ], and one sodium precursor, namely, Na[N(Si(CH3 )3 )2 ], were used with
the proportion varied from 1 to 10 mol% compared to Zn. It was observed that Li precursors induced the synthesis of ZnO nanoparticles; otherwise, without Li or with the
use of Na precursor the synthesis of ZnO nanorods was induced. Figure 10 shows the
TEM micrograph of ZnO nanoparticles synthesized using the Li[N(CH3 )2 ] precursor with
nanoparticle size varied from 37 07 nm to 25 04 nm [series 1]. Figure 11 shows the

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

(b)

(c)

(d)

15

Figure 10. TEM images of series 1 nanoparticles: (a) 1%, (b) 2%, (c) 5%, and (d) 10% Li. Reprinted with permission from [59], A. Glaria et al., New J. Chem. 32, 662 (2008). 2008, The Royal Society of Chemistry.

TEM micrograph of ZnO nanoparticles synthesized using the Li[N(Si(CH3 )3 )2 ] precursor


with nanoparticle size varied from 43 10 nm to 31 08 nm [series 2]. Figures 10(ad)
and Figures 11(ad) show that as the Li amount increases, the size of the nanoparticles decreases, whatever the Li precursor. The insets of Figures 11(c and d) show the
HRTEM image and conrm the monocrystalline nature of the ZnO nanoparticles. XRD
analysis conrmed the presence of the hexagonal zincite phase, space group P63 mc in
all samples. In addition, the optical properties of these nanoparticles were measured by
dissolving solid samples in distilled THF. The absorption spectrum for all the samples
shows a strong absorption between 300 and 350 nm followed by a sharp decrease. Furthermore, the luminescence properties of these samples were also investigated, which
shows one broad emission band in the visible range for an excitation wavelength of
320 nm. This shows that presence of Li ions leads to a blue shift of the emission band
of ZnO nanoparticles. The observed emission maxima vary from 582 to 535 nm for the
Li[N(CH3 )2 ] precursor and from 581 to 534 nm for the Li[N(Si(CH3 )3 )2 ] precursor. This
blue shift increases as the concentration of precursor increases, and consequently, as the
size of the nanoparticles decreases. Moreover, the observed emission intensity is very
strong, which can be clearly seen by the human eye as illustrated in Figure 12, which
opens the perspective for the preparation of LEDs.
Bardhan et al. [61] synthesized sub-micrometer ZnO particles with controlled morphology, i.e., rings, bowls, hemispheres, and disks, via a simple wet-chemistry approach using
zinc acetate as a precursor, ammonium hydroxide as a base, and ethanol as a solvent. The
reported morphologies were varied with the concentration of zinc acetate, i.e., at 0.05 M

16

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

(b)

(c)

(d)

Figure 11. TEM images of series 2 nanoparticles: (a) 1%, (b) 2%, (c) 5%, and (d) 10% Li. Reprinted with permission from [59], A. Glaria et al., New J. Chem. 32, 662 (2008). 2008, The Royal Society of Chemistry.

(rings), 0.01 M (bowls), 0.02 M (hemispheres), and 0.025 M (disks). Moreover, reaction
temperature, pH, and concentration of ammonium hydroxide also played an important
role for the formation of various ZnO morphologies. In addition, these synthesized ZnO
particles show strong white-light emission via UV excitation, which is due to the presence
of surface defects resulting from the method of fabrication and synthesis conditions. As a
result, the authors believe that based on the properties of these ZnO particles, it may lead
to the development of economical, white light-emitting materials for solid-state lighting
applications.
Hong et al. [62] reported the synthesis of quasi-spherical ZnO nanoparticles with diameters of 20 nm using zinc acetate as a precursor. In a typical synthetic process, 5%
PEG surfactant solution was transferred into a three-neck ask, and then zinc acetate
and (NH4 )2 CO3 solutions were added dropwise with vigorous stirring. The resulting suspension was kept for 2 h at room temperature under stirring. After the completion of
the reaction, the product was ltered and washed with ammonia solution and ethanol,
dried under vacuum for 12 h, and calcinated at 450 C for 3 h. The as-synthesized ZnO
nanoparticle surfaces were further grafted by polystyrene (PSt) in a non-aqueous suspension via free-radical polymerization to reduce the aggregation among nanoparticles and
to improve the compatibility between the nanoparticles and the organic matter, which
made a stable suspension in organic solvents. The resulting ZnO nanoparticles and PStgrafted ZnO nanoparticles were characterized by TEM, XRD, FT-IR analysis, zeta potential measurement, lipophilic degree (LD) test, photocatalytic analysis, sedimentation test,
and contact angle measurement. It is reported that bare ZnO nanoparticles have high

ZnO Nanoparticles: Growth, Properties, and Applications

17

Figure 12. Evolution of the emission at room temperature of the ZnO nanoparticles for series 2: (a) 1%, (b) 2%,
(c) 5%, and (d) 10%. Reprinted with permission from [59], A. Glaria et al., New J. Chem. 32, 662 (2008). 2008,
The Royal Society of Chemistry.

photocatalytic activity, although PSt-grafted ZnO nanoparticles have no photocatalytic


activity. Moreover, the LD of the composite particles after high temperature was stable,
and the photoluminescence of the PSt-grafted ZnO nanoparticles was observed by the
naked eye. In addition, ZnO nanoparticles can also be used to reinforce the electrical
conductivity of poly(vinylidene uoride) (PVDF) lms.
Ultrasound-assisted green synthesis of ZnO nanoparticles in room-temperature ionic
liquids (RTILs) was reported by Goharshadi et al. [63]. In a typical synthesis process, zinc
acetate dihydrate was dissolved in distilled water, and then NaOH was added to make
a transparent Zn(OH)2
4 solution, followed by the addition of an ionic liquid, 1-hexyl-3methylimidazolium bis (triuoromethylsulfonyl) imide, liquid [hmim][NTf2 ]. After this
step, the resulting solution was ultrasonically irradiated for 1 h with 40 kHz frequency of
ultrasound waves. The total acoustic power injected into the sample solution was found
to be 50 W. The as-synthesized ZnO nanoparticles were characterized by XRD, SEM,
and TEM. Various experiments were done to check the effects of RTILs and ultrasound
irradiation on the morphology of ZnO nanoparticles. On the basis of this observation,
it was found that RTIL and ultrasound have a critical role in the formation of ZnO
products. As for the growth mechanism, RTIL consists of cations [C6 mim]+ and anions
[NTf2 ] . Cationic species combined with Zn(OH)2
4 species present in the solution though
electrostatic attraction, and these cation-anion couples led to the formation of ZnO nuclei
via dehydration due to strong polarization of [C6 mim]+ . Moreover, the newly generated
ZnO surface was greatly inhibited by [C6 mim]+ ions, so the anisotropic growth of ZnO
crystals were markedly modied. As the method is very effective for the synthesis of ZnO
nanoparticles in a green media, it could be useful for synthesizing ZnO nanoparticles
with high yields.
Another approach to synthesize ZnO nanoparticles via one-step mechanochemical process was reported by Lu et al. [64]. In a typical synthesis process, matrix salts were
prepared by mixing zinc sulphate heptahydrate, potassium hydroxide, and potassium

18

ZnO Nanoparticles: Growth, Properties, and Applications

chloride. The mixing reaction was carried out in a paste state at room temperature
with short grinding time without any external energy input. The as-produced ZnO
nanoparticles had a mean diameter of 22.1 nm, which exhibited excellent UV-blocking
properties (UV absorption maxima at 358 nm) for cosmetic application, conrmed by
UV-visible (UV-vis) spectrometry. In addition, authors compared the raw material costs
with other mechanochemical processes and found that this process is more favorable
than others.
Up to now, ZnO nanoparticles synthesized by using zinc precursors with either base
or solvent for the combination of OH ions to produce zinc hydroxide moieties, which
produced ZnO nanoparticles via dehydration, have been reported. There was some
more literature related to direct conversion of inorganic or organic precursor to zinc
oxide nanoparticles, i.e., Rataboul et al. [65] reported the synthesis of ZnO nanoparticles
from thermal oxidation of Zn particles, which were produced by the decomposition of
organometallic precursor [Zn(C6 H11 )2 ] in a wet anisole. Zn particles can be prepared in
the presence or absence of polymer. In addition, whatever the synthesis and stabilization
modes, the particles display a uniform size and narrow size distribution. As-synthesized
products were characterized by TEM, XRD, and XPS, which were fully consistent with
the results.
Gattorno et al. [66] reported a novel synthesis pathway of ZnO nanoparticles with
narrow size distribution from spontaneous hydrolysis of zinc carboxylate salts in a polar
basic aprotic solvent i.e., dimethyl sulfoxide (DMSO) or DMF at room temperature. The
reproducibility of as-synthesized products depends upon the control over water content
and reaction temperature. As the hydrolysis of zinc carboxylates produced ZnO nanoparticles with different sizes, solvent basicity and the interaction of DMSO and water play an
important role in the hydrolysis mechanism. To check the stability and optical properties,
ZnO colloids were analyzed by UV-vis electronic absorption and emission spectroscopy,
and crystallinity was conrmed by powder X-ray diffraction spectroscopy. By HRTEM
analysis, it was conrmed that low concentration (2 104 M) of zinc cyclohexanebutyrate
produced 2.12 nm average size ZnO nanocrystallites, and zinc acetate produced 3 nm
average size ZnO nanocrystallites. In addition, if zinc cyclohexanebutyrate was used as
a starting material, ZnO nanocrystals with rock salt coexisting with a wurtzite structure
were produced. The presence of rock salt ZnO nanoparticles might be due to the phase
transformation induced by particle size and/or by the interaction of cyclohexanebutyrateZnO nanoparticles. Moreover, dynamic light backscattering size measurements of ZnO
nanoparticles were also performed in the DMSO colloidal dispersion to detect small individual nanoparticles and assemblies of ZnO nanoparticles. By more extensive research,
it was found that cyclohexanebutyrate acts as a more effective capping agent than acetate.
Moreover, although low colloidal (2 104 M) ZnO dispersions in DMSO did not show
any occulation or red shifts in 2 months, probably due to the concentrated dynamic
stabilizing action of carboxylate ions and solvent molecules, ZnO colloids in DMF were
not stable and readily formed precipitates, which can adhere to glass walls, and produced ZnO lms. This synthesis method is reported as a new, direct, clean, and very easy
pathway to obtain ZnO nanoparticles and can be applied to other metallic carboxylate
salts to form the corresponding nanostructures metal oxides.
A new method to produce zinc oxide nanoparticles by thermal decomposition of zinc
alginate was reported by Baskoutas et al. [67]. The reported method is based on the
preparation of zinc alginate gels by ionic gelation between zinc solution and sodium
alginate. The resulting wet beads were heated at 800 and 450 C for 24 h. The structural
morphologies and crystallinity of the as-synthesized ZnO nanoparticles were characterized by SEM, TEM, XRD, and micro-Raman spectroscopy. In more detailed observation, it
was reported that ZnO nanoparticles possessed wurtzite structures with single crystalline
hexagonal phase conrmed by XRD analysis and SAED. In addition, it was reported that
heating temperature and the kind of zinc agent (i.e., zinc nitrate or zinc acetate) inuence
the size of ZnO nanoparticles. Furthermore, Raman scattering conrmed the existence of
defects in the nanoparticles.

ZnO Nanoparticles: Growth, Properties, and Applications

19

Wahab et al. [68] reported the synthesis of ZnO nanoparticles from the conversion of
hydrozincite [Zn5 (CO3 )2 (OH)6 ]. In a typical synthetic process, hydrozincite was prepared
by reaction between zinc acetate dihydrate with urea in deionized water at 70 C for 2 h
via the solgel method. The quality and composition of the as-grown hydrozincite were
conrmed by XRD analysis with all the characteristic peaks for hydrozincite as well as by
FT-IR analysis. Furthermore, as-synthesized plate-like hydrozincite was converted to ZnO
nanoparticles with calcination at different temperatures, i.e., at 300, 500, 700, and 900 C.
The morphological characterization was done by FE-SEM and TEM analysis, which show
that as the calcination temperature increased, particles size also increased in the range of
20300 nm. As-synthesized zinc oxide nanoparticles were further characterized by HRTEM equipped with SAED, and the distance between lattice fringes was conrmed as
0.52 nm corresponding to the (0001) crystal plane. Thermogravimetric analysis (TGA) of
as-grown hydrozincite from room temperature to 700 C revealed that the primary weight
loss, which starts at 130 C was due to solvent evaporation and secondary weight loss
observed at 290 C was due to phase transformation from hydrated zinc oxide to zinc
oxide.
Niasari et al. [69] presented the synthesis of ZnO nanoparticles by thermal decomposition of [bis(acetylacetonato)zinc(II)]-oleylamine complex. First, zinc acetate-oleylamine
complex was prepared by the reaction between zinc acetate and oleylamine (C18 H37 N)
at 100 C for 90 min in a high purity oxygen atmosphere followed by injection of metalcomplex solution into triphenylphosphine (C18 H15 P) at 220 C, resulting in a black color
solution of [bis(acetylacetonato)zinc(II)]-oleylamine complex. Further, the black solution
was aged at 210 C for 45 min, resulting in a white nanoparticle product that was precipitated by adding excess ethanol to the solution. The resulted white products were
characterized by XRD, PL spectroscopy, and FT-IR spectroscopy. Morphological characterization shows zinc oxide nanoparticles with an average size of 1220 nm, which was
conrmed by SEM and TEM analysis. PL analysis shows the important strong blue-shift
emission band of ZnO nanoparticles, which was attributed to quantum size effect.

4. APPLICATION OF ZnO NANOPARTICLES


4.1. ZnO Nanoparticles: Bio-Friendly Approach
As biomolecules are very sensitive to the solution pH and temperature, there is a general
need to synthesize metal oxide semiconducting nanoparticles for possible applications
in biological sensing, biological labeling, drug and gene delivery, and nanomedicines
[7073]. In particular, due to their easy fabrication, environmentally friendly nature, and
non-toxic synthesis route, ZnO nanoparticles can provide a better option for various biological applications. However, water solubility and biocompatibility of ZnO nanoparticles
are the main requisites for biological applications. In this regard, Bauermann et al. [74]
reported the bio-friendly synthesis of ZnO nanoparticles in aqueous solution at nearneutral pH and low temperature (37 C). In a detailed synthesis process, a specic volume
of zinc nitrate hexahydrate was added into the buffer tris(hydroxymethyl)aminomethane
at pH 8 in an incubator for 4 h at 37 C. The as-synthesized ZnO nanoparticles were
characterized by SEM, TEM, XRD analysis, FT-IR spectroscopy, and thermogravimetry/mass spectrometry (TG/MS). Figure 13(a) shows the TEM image of as-synthesized
ZnO nanoparticles with a mean diameter of 20 nm, and Figure 13(b) shows the SEM
image of ZnO nanoparticles calcined at 1,000 C with a mean diameter of 300 nm.
Moreover, Figure 14 represents the XRD pattern of as-synthesized ZnO nanoparticles as well as samples calcined at different temperatures, i.e., 80, 600, and 1,000 C,
which conrmed the crystallinity and stable wurtzite phase of ZnO nanoparticles. XRD
crystal planes peaks became sharper as the samples were heated at higher temperatures with increasing crystal size. Furthermore, in terms of application, the buffer
tris(hydroxymethyl)aminomethane represented a standard nontoxic buffer that is inert to
a wide variety of chemicals and biomolecules and can be satisfactorily used for a variety
of biological reactions. In addition, this buffer has an important role for the sphericity

20

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

(b)

Figure 13. (a) Bright-eld TEM image of as-obtained ZnO nanoparticles precipitated in aqueous solution at pH
8 and 37 C and (b) SEM image of ZnO particles after heat treatment in air at 1,000 C for 2 h. Reprinted with
permission from [74], L. P. Bauermann et al., J. Phys. Chem. B 110, 5182 (2006). 2006, American Chemical
Society.

of the synthesized ZnO nanoparticles, it acts as a polydentade ligand, which adsorb


strongly on one or more surfaces of ZnO, inhibiting crystal growth, and as a result, nearly
spherical ZnO nanoparticles are produced. Moreover, the buffer also increases the rate
of hydrolysis of the zinc-water complex by consuming protons during the reaction and
produced ZnO with some trapped protons in the interstitial sites of ZnO crystals, which
after further heating at about 180 C caused a decrease in the unit cell volume of ZnO due
to the removal of interstitial protons from the crystalline structure of ZnO. The authors
believe that during crystallization, new hybrids of ZnO can be produced by introducing
biomolecules.
It is well reported that for biological applications the water solubility of a nanomaterial
is the main concern, and generally water solubility is achieved by surface modication with water-soluble ligands, silanization, or encapsulation within block-copolymer
micelles. In this regard, Wang et al. [75] reported the synthesis of a water-soluble ZnOAu nanocomposite having dual functionality, i.e., ZnO provides uorescence, and Au is
used for organic functionality for bioconjugation. In a typical synthetic process, rst, ZnO
nanocrystals were prepared using zinc acetate dihydrate in ethanol via reuxing with stirring at 80 C for 3 h. Furthermore, the as-synthesized ZnO nanocrystals were employed
as a seeding surface for the nucleation and growth of reduced gold by citrate to produce ZnO-Au nanocrystals having water-soluble characteristics. As-synthesized ZnO-Au
nanocrystals were characterized by TEM and XRD and conrmed as dumbbell-shaped
ZnO-Au nanocrystals having wurtzite ZnO and fcc Au with diameters of 4.9 and 7.1 nm,

d
c
b
a

Figure 14. X-ray powder diffraction patterns of ZnO nanoparticles precipitated in aqueous solution at pH 8
and 37 C thermally treated in air at (a) 37 C, (b) 80 C, (c) 600 C, and (d) 1,000 C. Reprinted with permission
from [74], L. P. Bauermann et al., J. Phys. Chem. B 110, 5182 (2006). 2006. American Chemical Society.

ZnO Nanoparticles: Growth, Properties, and Applications

21

respectively. It was reported that surface plasmon absorption band of ZnO-Au NCs was
broadened and red shifted relative to monometallic Au nanoparticles. In addition, the
UV emission intensity of ZnO-Au nanocrystals was 1 order of magnitude higher than
in pure ZnO nanocrystals due to the strong interfacial between ZnO and Au. Moreover,
multiphonon Raman scattering of ZnO-Au NCs was enhanced by strong localized electromagnetic of the Au surface plasmon.
Reddy et al. [76] reported the toxicity of ZnO nanoparticles to gram-negative [Escherichia
coli (E. coli)], gram-positive [Staphylococcus aureus (S. aureus)] bacterial systems, and primary human immune cells. ZnO nanoparticles were synthesized by forced hydrolysis
of zinc acetate at 160 C in diethylene glycol media. As-synthesized ZnO nanoparticles
with 13 nm diameter were characterized by TEM, XRD, and UV-vis spectrophotometery. The ZnO nanoparticles showed complete inhibition of E. coli growth at concentrations 3.4 mM, whereas growth of S. aureus was completely inhibited for concentrations
1 mM. In a more detailed observation, ow cytometry viability assays using a two color
live/dead Backlight kit, demonstrated that growth-inhibiting properties of ZnO nanoparticles corresponded to the loss of cell viability, but identical particles have minimal effects
on primary human T cell viability at concentrations toxic to gram-negative and grampositive bacteria. These observations conrmed the toxic nature of ZnO nanoparticles for
different bacterial systems, which could lead to biomedical and antibacterial applications.
Another approach regarding the use of ZnO nanoparticles in biological applications
was recently reported by Hanley et al. [77]. The authors reported the preferential killing
of cancer cells and activated human T cells using ZnO nanoparticles. For the synthesis of
813 nm ZnO nanoparticles, the authors adopted the forced hydrolysis of zinc acetate at
160 C in DEG. Then, ZnO nanoparticles were reconstituted in phosphate buffered saline
(PBS) solution. After reconstitution, nanoparticles were sonicated for 10 min and immediately vortexed before being added to cell culture. The response of normal human cells to
ZnO nanoparticles under different signaling environments was examined and compared
to the response of cancerous cells. In addition, ZnO nanoparticles exhibited a strong
ability to kill cancerous T cells (2835) compared to normal cells. Moreover, it was
observed that activation state of the cell contributes to the nanoparticle toxicity, as resting T cells display a relative resistance while cells stimulated through the T cell receptor
and CD28 costimulatory pathway show greater toxicity, which results in a direct relation to the level of activation. It was reported that appearance of toxicity was due to the
involvement of generated reactive oxygen species, as it was found that cancerous T cells
produced higher inducible levels than normal T cells. Furthermore, ZnO nanoparticles
induced apoptosis in Jurkat T cells, which is shown in Figure 15. To analyze the induced
apoptosis, two types of samples were prepared: (1) untreated cells and (2) cells treated
with 0.3 Mm nanoparticles for 20 h or treated with 100 nM okadaic acid for 20 h (positive
control) and then stained with green uorescent annexin V antibody to detect apoptotic
membrane and stained with red uorescent dye PI to detect permeable membranes using
the Vybrant apoptosis assay kit #2 (Molecular Probes). In terms of characterization, cells
were visualized by confocal microscopy which is shown in Figures 15(AC) for control
cells not treated with nanoparticles. Figure 15(A) shows control differential interference
contrast (DIC), Figure 15(B) shows a control DIC image with green and red uorescence
overlay, and Figure 15(C) shows a control green and red uorescence image. Cells treated
with ZnO nanoparticles are shown in Figures 15(DG), in which Figure 15(D) shows
treated nanoparticles in the DIC image, Figure 15(E) shows a DIC image with green and
red urorescence overly, Figure 15(F) shows a green and red urorescence image, and
Figure 15(G) shows an additional green and red urorescence image of nanoparticletreated cells of lower magnication. To further clarify naonparticle-induced apoptosis,
cells were left untreated (Fig. 16(A)), cells were treated with 100 nM okadaic acid for 20 h
as a positive control for apoptosis (Fig. 16(B)), and cells were treated with 0.3 mM ZnO
NP for 20 h (Fig. 16(C)) and then stained with DNA dye, acridine orange, and visualized
by uorescent microscopy. In Figures 16(B and C), arrows indicate the typical apoptotic
cells characterized by shrunken appearance and condensed or fragmented nuclei. Collectively, these results indicate that ZnO NPs induce apoptosis in Jurkat T cells. These

22

ZnO Nanoparticles: Growth, Properties, and Applications

(A)

(D)

(B)

(E)

(C)

(F)

(G)

Figure 15. ZnO nanoparticle-induced apoptosis in Jurkat T cells. Cells were left untreated, treated with 0.3 mM
ZnO NP for 20 h, or treated with 100 nM okadaic acid for 20 h (positive control) and stained with a green
uorescent annexin V antibody to detect apoptotic membranes and with the red uorescent dye PI to detect
permeable membranes using the Vybrant apoptosis assay kit #2 (Molecular Probes). Cells were visualized by
confocal microscopy and representative images are shown. (A)(C) control cells not treated with nanoparticles, (A) control DIC image, (B) control DIC image with green and red uorescence overlay, (C) control
green and red uorescence image. (D)(G) cells treated with nanoparticles; (D) nanoparticle-treated DIC image,
(E) nanoparticle-treated DIC image with green and red uorescence overly, (F) nanoparticle-treated green and
red uorescence image, (G) an additional green and red uorescence image of nanoparticle-treated cells of
lower magnication. Reprinted with permission from [77], C. Hanley et al., Nanotechnology 19, 295103 (2008).
2008, Institute of Physics Publishing Ltd.

observations may provide the basis for the development of new rational strategies to
protect against NP toxicity or enhance the destruction of disease-causing cell types such
as cancer cells.
Padmavathy et al. [78] reported the synthesis of ZnO nanoparticles with various sizes
and then investigated the antibacterial activity of the as-synthesized ZnO nanoparticles using a standard microbial method. In a typical synthetic process, two methods
were employed. In the rst method, zinc nitrate and sodium hydroxide were mixed
at room temperature with stirring for 2 h, and the resulting zinc hydroxide precipitate
was washed with distilled water until pH became neutral, followed by dropwise addition of H2 O2 to produce zinc peroxide translucent solution, which was then heated at
350 C to produce ZnO with active surface oxygen species. In another method, surfacemodied ZnO nanocrystals were prepared by dissolving zinc acetate in 2-propanol at
80 C with stirring, followed by the addition of 2-mercaptoethanol, which was continuously stirred for 2 h. The resulting mixture was then hydrolyzed by adding NaOH in
2-propanol, followed by ultrasonic agitation for 2 h, and then the synthesized products
were washed and dried. The as-synthesized products were characterized by XRD, TEM,
and PL spectroscopy. Figure 17 shows the TEM micrograph and SAED pattern of ZnO
nanoparticles formed by the precipitation method (Fig. 17(a)) and formed by base hydrolysis in propanol medium (Fig. 17(b)). It was observed that when capping molecules
were used, the kinetics of nucleation and accumulation were affected in such a way

ZnO Nanoparticles: Growth, Properties, and Applications

(A)

23

(B)

(C)

Figure 16. Detection of apoptotic morphological changes in Jurkat cells treated with ZnO NP. Cells were left
untreated (A), or treated with 100 nM okadaic acid for 20 h as a positive control for apoptosis (B), or treated
with 0.3 mM ZnO nanoparticles for 20 h (C) and stained with acridine orange and visualized by uorescent
microscopy. Arrows indicate typical apoptotic cells characterized by a shrunken appearance and condensed or
fragmented nuclei. Reprinted with permission from [77], C. Hanley et al., Nanotechnology, 19, 295103 (2008).
2008, Institute of Physics Publishing Ltd.

that the growth rate of large particles decreased while that of small particles remained
the same, which in turn produced particles with narrow size distribution as compared
to particles synthesized without a capping molecule. Furthermore, as-synthesized ZnO
nanoparticles with different concentrations underwent bacteriological tests by standard
microbial method in terms of minimum inhibitory concentration (MIC) and disk diffusion, which were performed in Luria-Bertani and nutrient agar media on solid agar
plates and in liquid. Figure 18 shows the bacterial efcacies with ZnO suspension with
thee different nanoparticle sizes after 24 h incubation of aliquots in the lowest concentration range (0.011 mM) and the highest concentration range (5100 mM). As a result,
enhanced bioactivity of smaller particles was recorded. For smaller ZnO nanoparticles,
more particles were needed to cover a bacterial colony (2 m), and more particles resulted
in the generation of a large number of active oxygen species, which kill bacteria more
effectively. As a result, it was observed that ZnO nanoparticles were more abrasive than
bulk ZnO and thus contributed greater mechanical damage of the cell membrane and
enhanced the anti-bacterial effect of ZnO nanoparticles. Reported observations and results
conrmed that ZnO nanoparticles may be applicable to medical devices that are coated
with nanoparticles against microbes.

4.2. Solar Cells, Photocatalytic, Photoluminescence, and Sensor


Application of ZnO Nanoparticles
Regarding ZnO nanoparticle application in solar cells, Suliman et al. [79] reported the
synthesis of ZnO nanoparticles with average diameter of 30 nm by using zinc chloride
as a precursor and NaOH as a base in a PVP solution of water at 160 C for 8 h via
the hydrothermal method. The as-synthesized structures were characterized by TEM,
SEM, and XRD analyses. Absorption spectrum was measured using a UV-vis spectrophotometer. To make a ZnO lm over transparent conducting glass (TCO), ZnO nanoparticles were dissolved in ethanol and then applied over the TCO surface using the doctor
blade technique, which resulted in a 6 m thick lm of ZnO nanoparticles over the
TCO, and nally it was annealed for 30 min at 450 C. To make dye-sensitized ZnO
thin lms, the lm was soaked in 0.5 mM ethanol solution of ruthenium complex,
cis bis(isothiocyanato)bis(2,2-bipyridyl-4,4-dicarboxylato)ruthenium (II) (N3 dye). The
TCO acted as a counter electrode on which 340 nm thick layer of Pt was deposited by
sputtering. Electrolyte was made by 0.03 M I2 /0.3 M LiI in propylene carbonate (PC)
which was attracted into the interelectrode space by capillary forces, and then the resulting lms of 0.4 cm2 were illuminated through the conducting glass support with an Oriel

24

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

(b)

Figure 17. (a) TEM micrograph and SAED pattern (inset) of ZnO formed by precipitation method. (b) TEM
micrograph and SAED pattern (inset) of ZnO formed by base hydrolysis in propanol medium. Reprinted with
permission from [78], N. Padmavathy et al., Sci. Technol. Adv. Mater. 9, 35004 (2008). 2008, Institute of Physics
Publishing Ltd.

91192 AM 1.5 solar simulator as the light source. ZnO nanoparticles lms were then
measured by photocurrentvoltage (IV ) which gives a ll factor of 0.513, short-circuit
current of 1.2 mA/cm2 , open-circuit voltage of 573 mV, and an overall light-to-electricity
conversion efciency of 0.75%.
Recently, Zhang et al. [80] reported the synthesis of polydisperse aggregated ZnO
nanocrystals and their application in dye-sensitized solar cells. In a typical synthetic process, ZnO aggregates were synthesized via polyol-mediated precipitation by using zinc
acetate in diethylene glycol at 160 C with reuxing. It was reported that with adjustment in zinc acetate concentration, rate of heating, and the amount of stock solution
that is added, one can readily control the size of individual ZnO aggregates. To make a
photoelectrode lms, ZnO aggregates were deposited by drop-cast method on a uorinedoped tin oxide (FTO) glass substrate, and thickness of lms depended on the number
of drops. Finally, the ZnO lms were heated at 350 C in air for 1 h to remove residual
organic chemicals. The as-synthesized product morphologies were characterized by SEM

25

ZnO Nanoparticles: Growth, Properties, and Applications


(a)

(b)

Figure 18. (a) and (b) Bactericidal efciency of samples 1 and 2 and bulk ZnO suspensions at different concentrations. Reprinted with permission from [78], N. Padmavathy et al., Sci. Technol. Adv. Mater. 9, 35004 (2008).
2008, Institute of Physics Publishing Ltd.

and XRD. Figure 19 shows the SEM images of hierarchically-structured ZnO lms with
polydisperse aggregates of ZnO with a size distribution of 120360 nm (Fig. 19(a)) and
120310 nm (Fig. 19(b)), lms consisting of monodisperse aggregates having average
sizes of 350, 300, 250, and 210 nm (Figs. 19(cg)), magnied SEM image (Fig. 19(h)),
and a schematic illustration to show the structure of the ZnO aggregates formed by
closely packed nanocrystallites 12 nm in size. To make the ZnO lm sensitized, it was
immersed into N3 dye for 20 min and then rinsed with ethanol to get rid of the excess
dye. Then, cells were constructed using a platinum-coated silicon wafer as the counter
electrode and the ZnO lm as a working electrode. These two electrodes were placed
side by side with 20 m separating space, where the I/I
3 electrolyte was injected with
capillarity. The solar cells performance was measured by the irradiation of air mass
(AM) 1.5 simulated sunlight at 100 mW cm2 . Figure 20 shows the typical IV curve

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)

(i)

Figure 19. SEM images of hierarchically-structured ZnO lms with submicrometer-sized aggregates. SEM
images of the lms consisting of polydisperse aggregates with a size distribution of (a) 120360 nm and
(b) 120310 nm. SEM images of the lms consisting of monodisperse aggregates with average sizes of (c) 350 nm,
(d) 300 nm, (e) 250 nm, and (f) 210 nm. (h) is a magnied SEM image, and (i) is a schematic illustration to
show the structure of ZnO aggregates formed by closely packed nanocrystallites. Reprinted with permission
from [80], Q. Zhang et al., Adv. Funct. Mater. 18, 1 (2008). 2008, Wiley-VCH Verlag GmbH & Co.

26

ZnO Nanoparticles: Growth, Properties, and Applications

Figure 20. An example of an IV curve for dye-sensitized ZnO solar cells under AM 1.5 irradiation. The
square shadow is plotted to illustrate the determination of the maximal power output of the solar cells.
Reprinted with permission from [80], Q. Zhang et al., Adv. Funct. Mater. 18, 1 (2008). 2008, Wiley-VCH Verlag
GmbH & Co.

for the polydisperse aggregated sample, which demonstrated the derivation of the opencircuit voltage Voc , the short-circuit current density Isc , and the maximum output power
density Pmax . The overall energy conversion efciency  and ll factor FF can be calculated sequentially by  = Pmax /Pin and F F = Pmax /Voc Isc ). As a result, it was found that
overall energy-conversion efciency of the cells could be affected by either average size or
size distribution of the ZnO aggregates. The highest overall energy-conversion efciency
of 4.4% was achieved by using lms formed by polydisperse ZnO aggregates with broad
size distribution of 120360 nm. In addition, variation in solar cell efciency was observed
due to light scattering, which was generated by submicrometer sized aggregates with a
size distribution comparable to the wavelength of incident light, which could extend the
travelling distance of light within the photoelectrode lm. Moreover, the observed high
efciency with polydisperse aggregates lms also due to its ability to provide the lm
with a closely packed structure, which was benecial to the transport of electrons in the
photoelectrode lm.
Regarding the application of ZnO nanoparticles in photocatalytic activity, Houskova
et al. [81] reported the synthesis of zinc sulde (ZnS) nanoparticles by homogenous
hydrolysis of zinc sulfate and thioacetadmide (TAA) at 80 C and then its conversion to
ZnO nanoparticles with annealing at temperature above 400 C in an oxygen atmosphere.
The as-synthesized ZnO nanoparticles were characterized by XRD and SEM, HRTEM and
SAED. The Brunauer-Emmett-Teller (BET) and Barrett-Joyner-Halenda (BJH) methods
were used to determine surface area and porosity, which showed pore sizes in the range
of 26 nm. Further, photocatalytic activities of as-synthesized ZnO nanoparticles were
determined by decomposition of Orange II dye in aqueous solution under UV irradiation
of 365 nm wavelength. Samples heated at 700 C exhibited a good photocatalytic activity,
k = 00379 min1 (k for P25 Degussa is 0.0222 min1 . Moreover, it was reported that assynthesized ZnO nanoparticles were evaluated for their non-photochemical degradation
of chemical warfare agents to non-toxic products, which established a good decomposition of the mustard gas.
Xu et al. [82] reported the synthesis of hierarchically assembled porous ZnO nanoparticles through a self-assembled pathway using surface-modied colloidal ZnO nanocrystals
as building blocks and P-123 copolymers as the template in aqueous solution. In a typical
synthetic process, rst, colloidal ZnO nanocrystallites were prepared by hydrolysis of
zinc acetate in a LiOH-ethanol solution, and then ZnO colloids were mixed with taurine
in deionized (DI) water with taurine/ZnO molar ratio as 1:1.4. The solution pH was
adjusted to 5.0 by adding 1 M HCL followed by vigorous stirring at 25 C for 24 h. The
resulting mixture was marked as sample A, and then in another experiment P-123 was
mixed with DI water at pH 5 and stirred at 25 C for 24 h (sample B). Finally, a solution

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

27

(b)

Figure 21. (a) Low-magnication cross-sectional TEM image and (b) HRTEM image of the as-obtained porous
ZnO nanoparticles. The inset in (b) clearly shows the lattice fringes. Reprinted with permission from [82], F. Xu
et al., Chem. Mater. 19, 5680 (2007). 2007, American Chemical Society.

was added dropwise into sample B with stirring, and after 3 h stirring, the resulting
mixture was heated in an autoclave at 70 C for 3 days, followed by washing and drying.
and then the product was calcined at 400 C for 57 h. The as-synthesized ZnO products
were then characterized by XRD, SEM, TEM, HRTEM, and PL spectroscopy. Figure 21
shows the TEM image of larger ZnO particles, which conrmed that these larger particles
may contain many smaller ZnO nanoparticles with uniform size. The HRTEM (Fig. 21(b))
image clearly indicates the contrast difference in each individual nanoparticle having
pores of 3 nm with average overall particle size 17 nm. Figure 21(b) (inset) shows
another HRTEM image of the lattice fringes of the nanocrystal with a spacing of 0.26 nm,
correspond to the interplanar distance of (002) plane of hexagonal ZnO. Figure 22 illustrates the detailed self-assembly processes involving the functionalization of individual
ZnO nanoparticles. Furthermore, on the basis of calorimetric measurements, the surface
enthalpy  of the hydrated porous ZnO is 142 021 J/m2 , which is in good agreement
with that of ZnO nanoparticles, which again supported the presence of self-assembled
ZnO nanocrystals in nanoporous ZnO. Finally, the photocatalytic activity of porous ZnO
nanoparticles was tested on the photodegradation of phenol under ambient conditions.
Figure 23 shows the emission spectra of residual phenol in aqueous solution under exposure to UV light for various times in the presence of porous ZnO nanoparticles, TiO2
nanoparticles (PC-500), commercial ZnO powder, and ZnO nanopowder. The porous ZnO

Figure 22. Schematic illustration of the self-assembly process, involving the functionalization of individual ZnO
nanoparticles. Reprinted with permission from [82], F. Xu et al., Chem. Mater. 19, 5680 (2007). 2007, American
Chemical Society.

28

ZnO Nanoparticles: Growth, Properties, and Applications

(a)

(b)

(c)

(d)

(e)

Figure 23. PL emission spectra of the residual phenol under exposure to UV light in the presence of (a) porous
ZnO nanoparticles, (b) TiO2 nanoparticles (PC-500), (c) commercial ZnO powder, and (d) ZnO nanopowder.
(e) Curves of the residual fraction of the phenol as a function of UV irradiation time when using () porous
ZnO nanoparticles, () TiO2 nanoparticles (PC-500), () commercial ZnO powder, and () ZnO nanopowder.
Reprinted with permission from [82], F. Xu et al., Chem. Mater. 19, 5680 (2007). 2007, American Chemical
Society.

nanoparticles show superior activity to TiO2 nanoparticles, because ZnO absorbs over
a large fraction of UV light and the corresponding threshold of ZnO is 425 nm, while
other ZnO (commercial and nanopowder ZnO) had less activity than the porous ZnO
but higher activity than TiO2 due to the unique surface features and higher surface area.
These results indicate that porous ZnO nanoparticles had good photoreactivity in the
decomposition of phenol in waste water.
Functionalized ZnO nanoparticles that show liquid-like behavior were synthesized
and their PL properties were reported by Bourlinos et al. [83]. First, ZnO nanocrystals (37 nm) were prepared by alkaline hydrolysis of zinc acetate in the presence
of LiOH H2 O in absolute ethanol for 45 days. As-prepared ZnO colloid was precipitated by adding excess of heptane, followed by centrifugation and drying at

ZnO Nanoparticles: Growth, Properties, and Applications

29

room temperature. Surface modication of ZnO nanoparticles with charged organosilane [(CH3 O)3 Si(CH2 )3 N+ (CH3 )(C10 H21 )Cl ] was carried out in an alkaline environment.
The Cl counter anions in the nanosalts could be readily exchanged by C9 H19 -C6 H4 (OCH2 CH2 )20 O(CH2 )3 SO
3 ions, yielding the corresponding sulfonate nanosalt as a waxy
solid that melts at 30 C, resulting in a uid with considerably higher viscosity than the
corresponding potassium sulfonate salt. The as-synthesized products were characterized
by XRD, TEM, FT-IR, and TGA/DTA analysis. Furthermore, the PL quantum yield of the
ZnO sulfonate nanosalt (0.065 mgmL1 in acetonitrile), corresponding to the green-yellow
emission band, was measured relative to that rhodamine 6G (R6G, 30 m in methanol).
Moreover, it was suggested by the authors that this property prole could lead to new
innovative applications in the areas of optics and photonics, and tuning the emission
towards the UV (by doping or different chemical processing) may activate lasing in the
ZnO nanoparticles, thus leading to the development of uid/exible laser sources.
Masuda et al. [84] reported PL from ZnO nanoparticles embedded in an amorphous
matrix. In a typical synthetic process, Zn(NO3 )2 [10 x M], Al(NO3 )3 (x M), and urea
(3.3 M) were dissolved in distilled water and kept at 90 C for 2 days to precipitate out.
Then the precipitate was washed with distilled water and further calcinated at 200900 C
for 3 h in air to synthesize ZnO nanoparticles dispersed in an amorphous matrix. The
morphologies of the as-synthesized products were characterized by SEM, chemical composition was determined by inductively coupled plasma, ZnO nanoparticles in an amorphous matrix were observed by TEM, crystallinity was observed by XRD, PL images
of ZnO nanoparticles were excited by visible light, UV light at 312 nm, or UV light at
254 nm, PL spectra were evaluated by a uorescence spectrometer using excitation light
at 287 nm, and the PL at low temperature was evaluated with a cryostat using liquid
helium. Figure 24 shows the TEM images of ZnO nanoparticles in the amorphous matrix
prepared with Al addition at 42% (Figs. 24(a1 and a2)) or 23% (Figs. 24(b1 and b2)) after
calcination at 250 C for 3 h in air. It was conrmed that ZnO nanoparticles 47 nm in
diameter were prepared by addition of 23% Al (Fig. 24(b)), and crystallization of ZnO
was not suppressed compared to 23 nm diameter with 42% Al addition (Fig. 24(a)).
Furthermore, control of ZnO nanoparticle size by addition of Al was conrmed by XRD
measurement. Figure 25 shows the PL images of ZnO nanoparticles (14, 5.5, 4, or 2.5 nm
in diameter) with Al addition excited by visible light, UV light at 312 nm, or UV light at
254 nm. It was conrmed that ZnO with no Al addition (0%) emitted only slightly under
UV light 312 nm or UV light 254 nm. As a result, the PL intensity of the ZnO nanoparticles drastically increased with Al addition, and the 2.5 nm ZnO nanoparticles (42% Al)
glowed brightly under both UV 312 nm and UV 254 nm. However, excessive Al addition
decreased the PL intensity due to the suppressed crystallization of ZnO nanoparticles,
which is essential for PL. It is shown in Figure 25 that the same ZnO nanoparticles can
show different color emissions with different excitation light in this system. In addition,
improvement of PL intensity of ZnO nanoparticles was evaluated by uorescence spectrometer and is shown in Figure 26(a), which conrmed that as the nanoparticle size
decreases, a blue shift is observed with increasing intensity due to the quantum size
effect. Figure 26(b) shows the absorption spectra of various sizes of ZnO nanoparticles.
The absorption spectrum shows the absorption edge as 3.15 eV of ZnO nanoparticles
without Al addition, whereas the adsorption edge shifted to a higher energy of 3.45
eV as the size of nanoparticles decreased to 2.5 nm due to quantum size effect. Figures 26(c and d) show PL spectra at low temperature for 2.5 nm ZnO nanoparticles
and conrm that ZnO nanoparticles were too small to show temperature dependence
of intensity and center wavelength of PL spectra. In conclusion, factors needed for controlling the crystallization of ZnO nanoparticleswhich is useful for increasing the PL
propertiesare Al concentration, calcination temperature, calcination period, and precursor, i.e., hydrotalcite. The authors believed that this novel process can be useful for
an advanced technology with strong growth potential for PL devices.
Recently, Jin et al. [85] reported the solution-processed UV photodetectors based on
colloidal ZnO nanoparticles. Solution-processed UV photodetectors were conveniently
fabricated by using lms of ZnO nanoparticles. The devices show low dark currents with

30

ZnO Nanoparticles: Growth, Properties, and Applications


(a) Zn : Al = 58 : 42
a1

a2

(b) Zn : Al = 77 :23
b1

b2

Figure 24. TEM images of ZnO nanoparticles in the amorphous matrix prepared with Al addition at (a1) and
(a2) 42% or (b1) and (b2) 23% after calcination at 250 C for 3 h in air. Reprinted with permission from [84],
Y. Masuda et al., Crystal Growth & Design 5, 1503 (2008). 2008, American Chemical Society.

a resistance > 1 T at room temperature as a consequence of low free carrier density


in the lms in the absence of illumination. In addition, at wavelength below 400 nm,
a strong photocurrent was seen with a responsivity of 61 A/W at an average intensity
of 1.06 mW/cm2 illumination at 370 nm. The characteristic times for rise and fall of the
photocurrent are <01 s and 1 s, respectively. Moreover, the devices work reproducibly
in air. Indeed the adsorption and desorption of oxygen from the nanoparticle surfaces
is essential to their operation. The photocurrent of the device is associated with a lightinduced desorption of oxygen from the nanoparticle surfaces, removing electron traps
and increasing the free carrier density, which reduces the Schottky barrier between the
contacts and the ZnO nanoparticles for electron injection. It is authors suggestion that
because of the very large surface-to-volume ratio, ZnO nanoparticle lms are ideal materials for use in UV photodetectors that rely on a mechanism related to gas adsorption
and desorption. Due to advantages of solution-processable fabrication, these devices have
potential for use in large-area UV photodetector applications.

31

ZnO Nanoparticles: Growth, Properties, and Applications

(a) 14 nm

(b) 5.5 nm

(c) 4 nm

(d) 2.5 nm

Vis

312 nm

254 nm

Figure 25. PL images of ZnO nanoparticles (a) 14, (b) 5.5, (c) 4, or (d) 2.5 nm in diameter in the amorphous
matrix excited by visible light, UV light 312 nm, or UV light 254 nm. Reprinted with permission from [84],
Y. Masuda et al., Crystal Growth & Design 5, 1503 (2008). 2008, American Chemical Society.

Regarding the use of ZnO nanoparticles for sensor applications, Baruwati et al. [86]
reported the hydrothermal synthesis of high-crystalline ZnO nanoparticles by using
zinc nitrate as a precursor and ammonium hydroxide as a base at 120 C for 624 h,
which were further used in sensing liqueed petroleum gas (LPG) and ethanol (EtOH).
As-synthesized ZnO nanoparticles were characterized by XRD, TEM, and TG/DTA, and
optical properties were measured by UV-DRS. Moreover, electrical properties were studied by AC impedance and DC conductivity measurement. For the gas sensing measurements, ZnO nanoparticles were coated over a cylindrical alumina tube of length 15 mm
and diameter 5 mm. Then, a heated coil was inserted inside the alumina tube and a pair
of electrodes was xed at two ends of the tube for electrical contacts. The gas sensing
properties toward reducing gases like LPG, ammonia, hydrogen, and EtOH were studied, and it was observed that nanoparticles showed high sensitivity to LPG and ethanol
at low operating temperatures. Moreover, it was also observed that if Pd was incorporated into ZnO nanoparticles, operating temperature decreased by more than 100 C, and
sensing characteristics improved in terms of response time and recovery times.
Recently, Cao et al. [87] reported methanal and xylene sensors at relatively low working temperatures. ZnO nanoparticles were prepared by a solid-state chemical reaction between zinc chloride and NaOH under ambient conditions. As-synthesized ZnO
nanoparticles with 4070 nm diameters were characterized by XRD, TEM, and HRTEM
equipped with SAED. As HCHO and C6 H4 (CH3 ) are the main sources for house pollution
and harm the health of human beings, the sensor must detect and monitor using suitable
techniques. In this regard, the authors reported gas sensitivity using ZnO nanoparticles.
It was observed that the optimal working temperatures are 300 C for HCHO sensors and
150 C for C6 H4 (CH3 ). The observed responses based on ZnO nanoparticles were 6.0 for
HCHO and 4.4 for C6 H4 (CH3 ). Based on the concentration of HCHO and C6 H4 (CH3 ) as
2100 ppm, the sensor showed response time of 6 s and recovery time of 14 s for HCHO
and response time of 8 s and recovery time of 13 s for C6 H4 (CH3 ). The authors also
reported the synthesis of ZnO nanorods, which show high sensitivity compared to synthesized ZnO nanoparticles. This may be because ZnO products prepared by the solidstate chemical methods under different conditions have different surface defects, which
may be active sites to adsorb the testing gas. The nanorods have different responses
owing to their different surface-to-volume ratios and surface states. The long narrow

32

ZnO Nanoparticles: Growth, Properties, and Applications


(a)

(b)

(c)

Figure 26. (a). PL spectra and (b) absorption spectra of ZnO nanoparticles (14, 5.5, 4, or 2.5 nm in diameter) in
the amorphous matrix. (c) PL spectra and (d) temperature dependence of ZnO nanoparticles at low temperature.
Reprinted with permission from [84], Y. Masuda et al., Crystal Growth & Design 5, 1503 (2008). 2008, American
Chemical Society.

ZnO Nanoparticles: Growth, Properties, and Applications

33

nanorods may be a favorable factor for the response signal caused by the chemical interaction of testing gases with the ZnO surface. Thus we conclude that not only morphology
but also size is very important for the materials applied to gas sensors.

4.3. Cosmetic Application of ZnO Nanoparticles


In addition to the above applications, i.e., gas sensors, chemical and biosensors, light
emitting diodes, photo-detectors, and photocatalytic application, ZnO nanoparticles also
exhibit tremendous UV-blocking properties. Generally, sunlight consists of three types
of UV radiation, i.e., UV-A (320400 nm), UV-B (290320 nm), and UV-C (250290 nm).
UV-A radiation is the main concern as it contributes 95% of the total sunlight radiation. UV-B radiation contributes 5%, and UV-C radiation has no prominent effect, as
it is absorbed by ozone at the surface of the Earth [8889]. Moreover, UV-A radiation is
considered more dangerous than UV-B, because it is 100 times more intense than UV-B
and can penetrate deeper into the dermis area of the skin. In the view of the abovementioned UV radiation values, it is important to block such types of harmful radiation,
as exposure causes skin cancer in humans. Generally, to protect the skin, materials having UV-blocking properties are added to cosmetic formulations. For the protection of
skin from UV-A radiation, ZnO nanoparticles provided an effective UV-blocking material
compared to TiO2 . Generally, ZnO nanoparticles effectively absorb UV-A radiation rather
than scatter it, but TiO2 usually scatters these wavelengths. Although ZnO absorption
for UV-A radiation is good compared to TiO2 , photocatalytic activity hinders its possible
application in cosmetic formulations. In addition, due to the high photocatalytic activity
of ZnO, reactive oxygen species are generated, which can oxidize ingredients involved
in the cosmetic formulation.
In this regard, some reports regarding the solution-based synthesis of ZnO nanoparticles and silica-coatings over the surface of as-synthesized ZnO nanoparticles to avoid
the undesired photoactivity are presented. Siddiquey et al. [90] reported the synthesis of silica-coated ZnO nanoparticles via a microwave-assisted route, and its photocatalytic activities were demonstrated. Silica-coated ZnO nanoparticles were characterized
by XPS, FT-IR, HRTEM, CHN elemental analysis, and zeta potential measurements. FT-IR
and XPS analysis clearly conrmed the silica-coating on the ZnO nanoparticles. HRTEM
micrographs revealed the continuous and dense silica coating of 3 nm thickness on the
surface of ZnO nanoparticles. In addition, silica coatings over ZnO nanoparticles were
conrmed by measurement of the zeta potential of silica-coated ZnO nanoparticles and
SiO2 . In both cases the zeta potentials were in good agreement. The photocatalytic activities of the non-coated and silica-coated ZnO nanoparticles were evaluated by photodegradation of methylene blue (MB) aqueous solution. The non-coated ZnO nanoparticles
showed high photocatalytic activity, and the concentration of MB decreased rapidly with
increasing UV irradiation time. On the other hand, when high Si-content (22.7 at%) was
used to coat the surface of ZnO nanoparticles, these silica layers effectively inhibited the
photocatalytic activity of ZnO nanoparticles. In addition, silica-coated ZnO nanoparticles
also showed excellent UV shielding ability and visible light transparency.
Regarding the surface coating of ZnO nanoparticles to reduce photocatalytic activity,
Fangli et al. [91] reported the preparation and properties of zinc nanoparticles coated
with zinc aluminate. Zinc oxide nanoparticles coated with zinc aluminate were prepared
by performing Al2 O3 precipitation on the precursor basic carbonate of zinc (BCZ) of zinc
oxide. As-synthesized coated ZnO nanoparticles were conrmed by TEM analysis, which
showed that a homogenous layer was formed on the surface of the zinc oxide nanoparticles, and the coated particle size was about 5060 nm. HRTEM analysis clearly indicated
that the outer thick alumina layer was 45 nm. In addition, the existence of the alumina
coating layer was conrmed by XPS and showed the presence of Zn, O, and Al. Moreover, XRD and lattice fringe data showed that the coating was of the ZnAl2 O4 phase
formed by the reaction of basic carbonate of zinc and basic carbonate of aluminium.
The catalytic activities of the zinc aluminate-coated ZnO nanoparticles were measured
by conductometric determination method (CDM) using castor oil as the oxidizing material. During the measurement of photocatalytic activities, the samples with castor were

34

ZnO Nanoparticles: Growth, Properties, and Applications

irradiated by a 250 W UV lamp. As a result, it was conrmed that zinc aluminate coating
is effective in reducing the catalytic activity of zinc oxide nanoparticles. Moreover, zinc
oxide nanoparticles coated with zinc aluminate not only retain a high UV absorptivity,
but also increase their reectance of visible light.

5. SUMMARY AND FUTURE DIRECTIONS


ZnO nanoparticles were synthesized using various routes. The solgel method proved a
successful synthetic route in terms of its cost, ease of handling, reliability, repeatability,
and environmental friendliness. Moreover, with detailed review, it is conrmed that the
various applications of ZnO nanoparticles depend upon the control of both physical and
chemical properties such as size, size dispersity, shape, surface state, crystal structure,
organization onto a support, and dispensability. In addition, these factors mainly depend
upon the synthetic method. Therefore, shape, size, and dispersity can be controlled by
tuning different parameters during the synthesis process, e.g., the precursor type and
concentration, types of capping molecule, types of solvent, reaction time, and reaction
temperature.
Regarding future aspect of ZnO nanoparticles, we conclude that although there are
various reports regarding ZnO nanoparticles with various shapes and sizes, there is still
a lack of quality synthesis in terms of crystallinity, size, and sphericity of the particles and the non-use of organic solvents (which may hinder its possible application in
biomedical sciences). ZnO nanoparticles used in cosmetic applications also require nonphotocatalytic activity, which can be hindered by the uniform surface coating of ZnO
nanoparticles by silica or other molecules, and toxicity to the human skin is also a main
concern. Application of ZnO nanoparticles in the biological realm requires high quality ZnO nanoparticles in aqueous solution at neutral pH and physiological temperature,
because biomolecules are very sensitive to changes in temperature and pH. In addition,
relatively small size, ease of transport within tissues/organs, ability to cross plasma membranes, and potential targeting of biologically active molecules will facilitate biomedical
applications of nanoparticles in the eld of medicine. At a conceptual level, we need a
better understanding of the relationship between size, shape, and structure of zinc oxide
nanoparticles, and how one can tune its capability for electronic and chemical interaction
with biological molecules and its sensing (biological and chemical) properties.

REFERENCES
1. M. Zheng, F. Davidson, and X. Huang, J. Am. Chem. Soc. 125, 7790 (2003).
2. M. Ferrari, Nat. Rev. Cancer 5, 161 (2005).
3. D. C. Reynolds, D. C. Look, B. Jogai, J. E. Hoelscher, R. E. Sheriff, M. T. Harris, and M. J. Callahan, J. Appl.
Phys. 88, 2152 (2000).
4. J. H. Lim, C. K. Kong, K. K. Kim, I. K. Park, D. K. Hwang, and S. J. Park, Adv. Mater. 18, 2720 (2006).
5. M. T. Mohammad, A. A. Hashim, and M. H. Al-Maamory, Mater. Chem. Phys. 99, 382 (2006).
6. A. B. G. Lansdown and A. Taylor, Int. J. Cosmet. Sci. 19, 167 (1997).
7. J. H. Kim, Y. C. Hong, and H. S. Uhm, Surf. Coat. Technol. 201, 5114 (2007).
8. M. S. Tokumoto, V. Briois, and C. V. Santilli, J. SolGel Sci. Technol. 26, 547 (2003).
9. J. Zhang, L. D. Sun, J. L. Yin, H. L. Su, C. S. Liao, and C. H. Yan, Chem. Mater. 14, 4172 (2002).
10. M. M. Demir, R. Munoz-Espi, I. Lieberwirth, and G. Wegner, J. Mater. Chem. 16, 2940 (2006).
11. S. Park, K. R. Lee, C. H. Jung, S. J. Kim, and H. C. Shin, Jpn. J. Appl. Phys. 35, 996 (1996).
12. L. C. Damonte, L. A. M. Zelis, B. M. Soucase, and M. A. H. Fenollosa, Powder Technol. 148, 15 (2004).
13. R. Radoi, P. Fernandez, J. Piqueras, M. S. Wiggins, and J. Solis, Nanotechnology 14, 794 (2003).
14. X. Zhao, B. Zheng, and C. Li, Powder Technol. 100, 20 (1998).
15. J. S. Lee, K. Park, S. Nahm, M. IL. Kang, I. L. W. Park, S. W. Kim, W. K. Cho, H. S. Han, and S. Kim,
J. Cryst. Growth 254, 423 (2003).
16. J. B. Baxter and E. S. Aydil, Appl. Phys. Lett. 86, 53114 (2005).
17. M. H. Huang, S. Mao, H. Feick, H. Q. Yan, Y. Wu, H. Kind, E. Weber, R. Russo, and P. Yang, Science 292,
1897 (2001).
18. J. Song, J. Zhou, and Z. L. Wang, Nano Lett. 6, 1656 (2006).
19. Z. L. Wang, Annu. Rev. Phys. Chem. 55, 159 (2004).
20. J. Sawai, H. Igarashi, A. Hashimoto, T. Kokugan, and M. Shimizu, J. Chem. Eng. Japan 29, 556 (1996).
21. S. C. Minne, S. R. Manalis, and C. F. Quate, Appl. Phys. Lett. 67, 3918 (1995).

ZnO Nanoparticles: Growth, Properties, and Applications


22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.

35

R. Q. Song, A. W. Xu, B. Deng, Q. Li, and G. Y. Chen, Adv. Funct. Mater. 17, 296 (2007).
P. X. Gao, Y. Ding, W. Mai, W. L. Hughes, C. Lao, and Z. L. Wang, Science 309, 1700 (2005).
P. Hu, Y. Liu, X. Wang, L. Fu, and D. Zhu, Chem. Commun. 1304 (2003).
J. Zhang, L. Sun, C. Liao, and C. Yan, Chem. Commun. 262 (2002).
Q. Tang, W. Zhou, J. Shen, W. Zhang, L. Kong, and Y. Qian, Chem. Commun. 712 (2004).
L. E. Greene, M. Law, J. Goldberger, F. Kim, J. C. Johnson, Y. Zhang, R. J. Saykally, and P. Yang, Angew.
Chem. 42, 3031 (2003).
X. Y. Kong, Y. Ding, R. Yang, and Z. L. Wang, Science 303, 1348 (2004).
D. Qian, J. Z. Jiang, and P. L. Hansen, Chem. Commun. 1078 (2003).
X. Zhong and W. Knoll, Chem. Commun. 1158 (2005).
F. Li, Y. Ding, P. Gao, X. Xin, and Z. L. Wang, Angew. Chem. 43, 5238 (2004).
G. Q. Ding, W. Z. Shen, M. J. Zheng, and D. H. Fan, Appl. Phys. Lett. 88, 103106 (2006).
B. G. Shpeizer, V. I. Bakhmutov, and A. Cleareld, Micro. Meso. Mater. 90, 81 (2006).
X. Wang, C. J. Summers, and Z. L. Wang, Adv. Mater. 16, 1215 (2004).
S. Polarz, A. V. Orlov, F. Schth, and A. H. Lu, Chem. Eur. J. 13, 592 (2007).
M. Monge, M. L. Kahn, A. Maisonnat, and B. Chaudret, Angew. Chem. Int. Ed. 42, 5321 (2003).
C. L. Carnes and K. J. Klabunde, Langmuir 16, 3764 (2000).
M. L. Kahn, M. Monge, V. Colliere, F. Senocq, A. Maisonnat, and B. Chaudret, Adv. Funct. Mater. 15, 458
(2005).
T. Andelman, Y. Gong, M. Polking, M. Yin, I. Kuskovsky, G. Neumark, and S. OBrien, J. Phys. Chem. B
109, 14314 (2005).
S. K. N. Ayudhya, P. Tonto, O. Mekasuwandumrong, V. Pavarajarn, and P. Praserthdam, Crystal Growth &
Design 6, 2446 (2006).
Y. Yang, H. Chen, B. Zhao, and X. Bao, J. Cryst. Growth 263, 447 (2004).
Y. Zhang, K. Yu, D. Jiang, Z. Zhu, H. Geng, and L. Luo, Appl. Surf. Sci. 242, 212 (2005)
T. Kawano and H. Imai, Crystal Growth & Design 6, 1054 (2006).
H. Wang, C. Xie, and D. Zeng, J. Cryst. Growth 277, 372 (2005)
H. Du, F. Yuan, S. Huang, J. Li, and Y. Zhu, Chem. Lett. 33, 770 (2004).
H.-M. Cheng, H.-C. Hsu, S.-L. Chen, W.-T. Wu, C.-C. Kao, L.-J. Lin, and W.-F. Hsieh, J. Cryst. Growth 277,
192 (2005).
H.-M. Cheng, K.-F. Lin, H.-C. Hsu, C.-J. Lin, L.-J. Lin, and W.-F. Hsieh, J. Phys. Chem. B 109, 18385 (2005).
Z. Hu, J. F. H. Santos, G. Oskam, and P. C. Searson, J. Colloid and Inter. Sci. 288, 313 (2005).
Z. Hu, G. Oskam, R. L. Penn, N. Pesika, and P. C. Searson, J. Phys. Chem. B 107, 3124 (2003).
M. Vafaee and M. S. Ghamsari, Matt. Lett. 61, 3265 (2007).
K. D. Kim, D. W. Choi, Y.-H. Choa, and H. T. Kim, Colloids and Surf. A: Physiochem. Eng. Aspects 311, 170
(2007).
P. Uthirakumar, B. Karunagaran, S. Nagarajan, E.-K. Suh, and C.-H. Hong, J. Cryst. Growth 304, 150 (2007).
M. Y. Ge, H. P. Wu, L. Niu, J. F. Liu, S. Y. Chen, P. Y. Shen, Y. W. Zeng, Y. W. Wang, G. Q. Zhang, and
J. Z. Jiang, J. Cryst. Growth 305, 162 (2007).
N. Uekawa, S. Iahii, T. Kojima, and K. Kakegawa, Matt. Lett. 61, 1729 (2007).
S. Lee, S. Jeong, D. Kim, S. Hwang, M. Jeon, and J. Moon, Superlatt. Microstruct. 43, 330 (2008).
G. Ning, X. Zhao, J. Li, and C. Zhang, Opt. Mats. 28, 386 (2006).
R. Xie, D. Li, H. Zhang, D. Yang, M. Jiang, T. Sekiguchi, B. Liu, and Y. Bando, J. Phys. Chem. B 110, 19147
(2006).
J. Buha, I. Djerdj, and M. Niederberger, Crystal Growth & Design 7, 113 (2007).
A. Glaria, M. L. Kahn, T. Cardinal, F. Senocq, V. Jubera, and B. Chaudret, New J. Chem. 32, 662 (2008).
R. Bardhan, H. Wang, F. Tam, and N. J. Halas, Langmuir 23, 5843 (2007).
E. K. Goharshadi, Y. Ding, M. N. Jorabchi, and P. Nancarrow, Ultra. Sonochem. 16, 120 (2009).
J. Lu, K. M. Ng, and S. Yang, Ind. Eng. Chem. Res. 47, 1095 (2008).
F. Rataboul, C. Nayral, M. Casanove, A. Maisonnat, and B. Chaudret, J. Orgmett. Chem. 643, 307 (2002).
G. R. Gattorno, P. S. Jacinto, L. R. Vzquez, J. Nmeth, I. Dkny, and D. Dz, J. Phys. Chem. B 107, 12597
(2003).
P. D. Cozzoli, A. Kornowski, and H. Weller, J. Phys. Chem. B 109, 2638 (2005).
R. Y. Hong, L. L. Chen, J. H. Li, H. Z. Li, Y. Zheng, and J. Ding, Polym. Adv. Technol. 18, 901 (2007).
S. Baskoutas, P. Giabouranis, S. N. Yannopoulos, V. Dracopoulos, L. Toth, A. Chissanthopoulos, and
N. Bouropoulos, Thin Sol. Films 515, 8461 (2007).
R. Wahab, S. G. Ansari, Y. S. Kim, M. A. Dar, and H.-S. Shin, J. Alloys and Compounds 461, 66 (2008).
M. S. Niasari, F. Davar, and M. Mazahari, Matt. Lett. 62, 1890 (2008).
S. E. McNeil, J. Leukoc Biol. 78, 585 (2005).
J. K. Jaiswal and S. M. Simon, Trends Cell Biol. 14, 497 (2004).
T. K. Jain, M. A. Morales, S. K. Sahoo, D. L. Leslie-Pelecky, and V. Labhasetwar, Molecular Pharmaceutics
2, 194 (2005).
R. K. Visaria, R. J. Grifn, B. W. Williams, E. S. Ebbini, G. F. Paciotti, C. W. Song, and J. C. Bischof, Mol.
Cancer Ther. 5, 1014 (2006).
L. P. Bauermann, J. Bill, and F. Aldinger, J. Phys. Chem. B 110, 5182 (2006).
X. Wang, X. Kong, Y. Yu, and H. Zhang, J. Phys. Chem. C 111, 3836 (2007).

36

ZnO Nanoparticles: Growth, Properties, and Applications

76. K. M. Reddy, K. Feris, J. Bell, D. G. Wingett, C. Hanley, and A. Punnoose, Appl. Phys. Lett. 90, 213902
(2007).
77. C. Hanley, J. Layne, A. Punnoose, K. M. Reddy, I. Coombs, A. Coombs, K. Feris, and D.Wingett, Nanotechnology 19, 295103 (2008).
78. N. Padmavathy and R. Vijayaraghavan, Sci. Technol. Adv. Mater. 9, 35004 (2008).
79. A. E. Suliman, Y. Tang, and L. Xu, Sol. Eng. Mat. & Sol. Cells 91, 1658 (2007).
80. Q. Zhang, T. P. Chou, B. Russo, S. A. Jenekhe, and G. Cao, Adv. Func. Mater. 18, 1 (2008).
81. V. Houskova, V. Stengl, S. Bakardjieva, N. Murafa, A. Kalendova, and F. Oplustil, J. Phys. Chem. A 111,
4215 (2007).
82. F. Xu, P. Zhang, A. Navrotsky, Z.-Y. Yuan, T.-Z. Ren, M. Halasa, and B.-L. Su, Chem. Mater. 19, 5680 (2007).
83. A. B. Bourlinos, A. Stassinopoulos, D. Anglos, R. Herrera, S. H. Anastasiadis, D. Petridis, and
E. P. Giannelis, Small 2, 513 (2006).
84. Y. Masuda, M. Yamagishi, W. S. Seo, and K. Koumoto, Crystal Growth & Design 8, 1503 (2008).
85. Y. Jin, J. Wang, B. Sun, J. C. Blakesley, and N. C. Greenham, Nano Lett. 8, 1649, (2008).
86. B. Baruwati, D. K. Kumar, and S. V. Manorama, Sens. Actuators B Chem. 119, 676 (2006).
87. Y. Cao, P. Hu, W. Pan, Y. Huang, and D. Jia, Sens. Actuators B Chem. 134, 462 (2008).
88. D. Fairhurst and M. Mitchnick, Sunscreens, Development, Evaluation, and Regulatory Aspects,
2nd Edition. Marcel Dekker, New York, 1997.
89. D. B. Brown, A. E. Peritz, D. L. Mitchell, S. Chiarello, J. Uitto, F. P. Gasparro, Photochem. Photobiol. 72, 340
(2000).
90. I. A. Siddiquey, T. Furusawa, M. Sato, and N. Suzuki, Mater. Res. Bult. (2007), in press.
91. Y. Fangli, H. Peng, Y. Chunlei, H. Shulan, and L. Jinlin, J. Mater. Chem. 13, 634 (2003).

Potrebbero piacerti anche