You are on page 1of 15

Dalton

Transactions
View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

PERSPECTIVE

Cite this: Dalton Trans., 2013, 42, 12554

View Journal | View Issue

Zinc oxide nanoparticles: chemical mechanisms and


classical and non-classical crystallization
Bettina Ludi and Markus Niederberger*
Among all the functional materials, ZnO plays an outstanding role in terms of chemical and physical
properties, but also in terms of morphological variety and the number of reported synthesis approaches.
Complex shapes and hierarchical architectures make ZnO a perfect example to study chemical and crystal-

Received 5th March 2013,


Accepted 15th April 2013

lization mechanisms. In this review article, we will discuss the nucleation and growth of ZnO nanostructures in liquid media by classical and non-classical (i.e., particle-based) crystallization pathways. We

DOI: 10.1039/c3dt50610j

elaborate the chemical conditions and parameters that are responsible for the occurrence of one or the

www.rsc.org/dalton

other pathway.

Introduction
Zinc oxide is a IIVI semiconductor with a direct band gap
around 3.4 eV and a large exciton binding energy of about
60 meV which makes it thermally stable at room temperature.1
Therefore, ZnO is a promising material for application in
lasers and UV light-emitting-diodes.2,3 Values of the exciton
Bohr radius that gives an estimate of the onset of quantumLaboratory for Multifunctional Materials, Department of Materials, ETH Zurich,
Wolfgang-Pauli-Strasse 10, 8093 Zurich, Switzerland.
E-mail: Markus.Niederberger@mat.ethz.ch

Bettina Ludi

Bettina Ludi studied chemistry at


the University of Ulm, Germany.
In 2006 she received her
Diploma degree writing her
thesis at the Joint Research
Centre of the European Commission in Ispra, Italy. In 2012 she
completed her PhD studies on
molecular and crystallization
mechanisms of zinc oxide nanoparticles in nonaqueous media
in the Laboratory for Multifunctional Materials at ETH Zurich.
Currently she works at Novartis
Pharma AG in Basel, Switzerland.

12554 | Dalton Trans., 2013, 42, 1255412568

confinement eects in ZnO are mainly found between 1.8 and


2.8 nm.1,4,5 Thus, a decrease in band gap energy and a shift of
the absorption edge to longer wavelengths can be found for
zinc oxide nanoparticles up to a size of 46 nm.57 ZnO occurs
naturally in the hexagonal wurtzite-type structure together
with certain amounts of manganese or iron and is named
zincite.8 Although the mineral displays a yellow to red color,
pure ZnO without the presence of natural contaminants is colorless due to its large band gap.
The group of IIVI semiconductors is known to crystallize
predominantly either in cubic zinc blende-type or hexagonal
wurtzite-type structure in which the ions of one type are

Markus Niederberger is Chair


of the Laboratory for Multifunctional Materials at ETH
Zurich. He studied chemistry at
ETH Zurich, where he also
received his PhD. After a postdoctoral stay at the University of
California at Santa Barbara, he
became group leader at the Max
Planck Institute of Colloids
and Interfaces in Potsdam. In
2007 he was appointed as Assistant Professor at ETH Zurich and
Markus Niederberger
in 2012 he was promoted to
Associate Professor. The Niederberger group works on the development of rational synthesis concepts for the production of new
types of functional materials and their application in gas sensor
technology, electronics, as well as in energy storage and
conversion.

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Dalton Transactions
surrounded by four counter ions situated at the corners of a
tetrahedron. This tetrahedral coordination is characteristic of
covalent bonding between sp3-hybridized atoms and thus can
be attributed to the influence of the covalent bonding contribution of the commonly iono-covalent materials.1,4 With
increasing the fraction of the ionic bonding character
the cubic rock salt structure gains in importance1 and thus all
the three crystal structures can be found for ZnO. Whereas the
rock salt structure can only be obtained at relatively high pressures, the cubic zinc blende structure can be stabilized by epitaxial growth on cubic substrates.4 In contrast to other IIVI
semiconductors like ZnS which appear in both zinc blende
and wurtzite structures, for ZnO the second structure type is
the most stable and most common one.
The wurtzite-type structure belongs to the hexagonal space
group P63mc with the lattice parameters a = 3.2499 and c =
5.2066 9 and contains two formula units of ZnO per unit cell
(Fig. 1). The ratio
pc/a
= 1.602 deviates slightly from the ideal
value of c/a = 8=3 = 1.633.9 Within ( partial) ionic materials
crystallizing in a non-centro symmetric structure, a dipole
moment increasing with the crystal thickness and geometrically related to the unit cell can be found. In the case of zincite
this dipole moment is directed along the c-axis of the crystal
lattice, resulting in two polar {0001} surfaces with the (0001)
) being terminated by
being zinc-terminated and the (0001
10
oxygen.
From an electrostatic point of view, these polar surfaces are
unstable and the main stabilization mechanisms proposed are
either a geometrical rearrangement of the surface atoms or a
metallization of the surface by charge transfer.10 Besides the
obvious influence on crystal growth due to an increase in
surface energy for the polar surfaces, material properties like
spontaneous polarization, plasticity and piezoelectricity can be
directly related to this intrinsic polarity of the crystal structure.4 The three most commonly found crystal faces determining the growth habit and thus the shape of the zinc oxide
0}, {112
0} and {0001}. With
nanoparticles generated are {101

the {1010} surfaces being the most stable ones and the {0001}
surfaces exhibiting the highest surface energy of the low-index

Fig. 1

Crystal structure of hexagonal ZnO.

This journal is The Royal Society of Chemistry 2013

Perspective
planes,11 anisotropic crystal growth along the c-axis is obtained
frequently, leading to the formation of ZnO nanorods with a
minimum in surface area for the {0001} facets and maximizing
0} surfaces.
the area of the {101
Zinc oxide is used in a large variety of dierent fields
ranging from large-scale products to more advanced applications. A great amount of ZnO is used in the concrete and
rubber industry to improve the processing time and water
resistivity or to influence the vulcanization process.1,12 Furthermore, ZnO microparticles with their ability to absorb UV light
are used in sunscreens, and due to their antiseptic properties,
in pharmaceutical products like pastes and plasters for
wound-treatment.8,12,13 One of the most obvious applications
is their use as a white pigment in paintings, although today it
is mostly substituted by titania which possesses a higher index
of refraction.8 In the chemical industry, zinc oxide is a widely
applied heterogeneous catalyst for the production of many
chemicals such as methanol for which Cu/ZnO/Al2O3 catalysts
are used.12 In addition, ZnO is utilized as a transparent conducting oxide layer to substitute the more expensive indium
tin oxide in solar cell applications, gas sensors or varistors.12
A more comprehensive insight into the properties and
possible applications of zinc oxide materials can be found
elsewhere.1,4,6,12,14
Due to its importance in current and future technological
applications it is not surprising to find a wealth of dierent
synthesis routes for the generation of zinc oxide nanoparticles
with various shapes and dimensions in the literature. To
review all of these approaches would go beyond the scope of
the present work; thus the following discussion will be
focused on liquid phase routes. Furthermore, only those
studies giving a deeper insight into the nucleation and growth
behavior from a kinetics as well as morphological point of
view will be reviewed and these are divided into the results on
classical and non-classical crystallization phenomena.

Classical nucleation, growth and ripening


behavior
UV-vis spectroscopy constitutes a widely used tool to study the
growth of ZnO nanoparticles. Several research groups investigated the change in nanoparticle radius with reaction time by
the analysis of UV-vis absorption spectra in dependence of
dierent synthesis parameters such as temperature, solvent,
precursor species, the type of base, capping agents, concentration of reagents and water. For many of the investigated
systems, the underlying reaction conditions are basically the
same, which allows a comparison of the results.
One of the first investigations on the change in particle
radius with reaction time using zinc acetate dihydrate
(Zn(OAc)22H2O) as a precursor and sodium hydroxide in isopropanol were published by the group of Searson in 1998.15 In
their work, they were able to detect a linear increase of the
cube of the particle radius with reaction time for all temperatures and in the whole time interval investigated. This linear

Dalton Trans., 2013, 42, 1255412568 | 12555

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Perspective
dependence of the cube of the particle radius on reaction time
agrees well with the rate law deduced by Lifshitz, Slyozov16 and
Wagner (known as the LSW theory),17 pointing to diusionlimited Ostwald ripening (OR). In addition, an increase in the
ripening rates with reaction temperature due to a decrease in
solvent viscosity was found.15
Around the same time, Meulenkamp published a work
about the aging of zinc oxide sols over a maximum time interval
of more than 500 days.18 He compared the change in particle
radii of washed colloidal particles, i.e., without reaction products and possibly unreacted precursor, with untreated sols.
Whereas the increase in particle size of untreated sols was
found to proceed faster in the initial time interval of several
hundred days, a crossover point for the particle radii between
the untreated and the washed particle sols was found. The
growth of unwashed particles slowed down and finally resulted
in a constant particle radius. On the other hand, the growth of
washed particles did not stop within the whole time interval
studied.18 One could interpret this observation in such a way
that ripening processes could seriously be hampered due to the
presence of surface-adsorbing species in the unwashed sols.
Moreover, Meulenkamp found a logarithmic dependence of the
particle radius on the aging time for unwashed sols and a linear
behavior for the washed particles.18 Assuming that the growth
of the washed particles is solely determined by coarsening processes, then the growth behavior does not follow the LSW
theory for diusion-controlled OR.
Comparing the results of Meulenkamp and Searson, an
increase of coarsening rates with aging temperature could be
detected in both studies.15,1820 However, Meulenkamp
detected a logarithmic dependence of the particle radius on
the aging time for the unwashed sols,18 whereas Searson and
co-workers detected a diusion-controlled ripening process.19
Because both reaction systems are quite similar involving the
same precursor and solvent, this discrepancy might be related
to the completely dierent time intervals studied. Similar deviations from diusion-limited OR processes for the growth
kinetics of ZnO particles could be detected by the group of
Sarma. Their investigation of the temperature dependence of
particle growth behavior was performed in a similar system,
but with increasing the molar ratio between Zn(OAc)2 and
NaOH.21 A good fit for the experimentally observed change in
particle radii with reaction time was obtained by using the
same expression as that for the LSW theory, dx dx0 = Kt (d is
the mean particle diameter at time t, d0 is the mean initial particle diameter and K is a constant), but with varying values of x.
The best fits were obtained for x ranging from 48. Therefore,
the ripening rates are much slower than expected for
diusion-limited processes based on the LSW theory for which
a value of x = 3 should apply.
A general problem, which makes a comparison of the
results from dierent studies somewhat delicate, is how the
particle radii were determined from UV-vis absorption analysis. Typically, the size of nanocrystals is related to the change
in band gap energy by the Eective Mass Approximation developed by Brus.2224 However, the Eective Mass Approximation

12556 | Dalton Trans., 2013, 42, 1255412568

Dalton Transactions
was found to overestimate the particle size of semiconductors
in general25 and of ZnO in particular.5 Therefore, Sarma and
co-workers developed another approach to calculate nanoparticle size and size distribution by UV-vis absorption analysis.5
Its validity was proven by an extensive comparison with data
obtained from small angle X-ray scattering (SAXS).21
The influence of solvents and anions on particle growth behavior was studied by varying the carbon chain length of the
alcohol and the composition of the zinc oxide precursors.19,20
Changing the solvent from hexanol to ethanol revealed an
increase of the time period during which nucleation and
growth occurred.19 According to classical crystallization theory,
the nucleation rate depends on the supersaturation, i.e., the
larger the supersaturation, the faster the nucleation rate. With
the increase in the dielectric constant of the solvent from
hexanol to ethanol, also the solubility of ZnO becomes
higher.26 Consequently, the supersaturation decreases, slowing
down the nucleation rate. At longer reaction times, nanoparticle growth is determined by diusion-limited OR in all alcohols, but enhanced ripening rates were detected with
increasing carbon chain length. It is dicult to explain this
observation, because the viscosity, solubility and surface
energy of the particles are all altered by varying the solvent.
Nevertheless, it is reasonable to assign a major part of this
eect to an increase in surface energy due to the weaker interaction of the solvent with the particle surface with decreasing
dielectric constant from ethanol to hexanol.19
The use of zinc oxide precursors with dierent compositions, i.e., with dierent counter ions, also revealed a
diusion-limited OR mechanism.20 The ripening kinetics were
found to be inversely proportional to the adsorption strength
of the anion. Stronger interaction between the anion and the
ZnO particles led to a decrease in surface energy. In addition,
particularly weakly adsorbing perchlorate anions resulted in
oriented attachment (OA) of ZnO nanoparticles at higher
temperatures and radii above 2 nm.20
The impact of the water concentration on particle ripening
behavior was studied by two dierent approaches. In the first
one, dierent amounts of water were mixed with a standard
reaction solution of Zn(OAc)22H2O and NaOH in isopropanol.27 Only if the water concentration exceeded a specific
threshold, ZnO nanoparticles could be detected, but then the
particle radii increased with higher water concentrations. On
the other hand, the ripening rates remained unaected by the
amount of water.27 The second approach involved the substitution of sodium hydroxide by water. Typically, particle nucleation and growth periods were shortened and the particle sizes
became larger with higher water concentrations.7 Furthermore,
a diusion-limited OR process was postulated for the increase
in particle size at longer reaction times, and an enhancement
in ripening rates was found by adding a higher amount of
water.7 However, compared to systems in which a base was
applied, the overall reaction rates were detected to be slowed
down.7 Essentially the same system was reinvestigated by
Sarma and co-workers a few years later and, again, a linear
dependence of d3 on reaction time was found, but for longer

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Dalton Transactions
times only.28 Based on this result, they questioned whether the
validity of the LSW theory solely based on the criterion of a
linear dependence of the cubed particle diameter on the reaction time can be proven, especially if only present in the
asymptotic limit.28 A closer look at the variation of the rate
constant K with temperature and the respective activation
energy with water concentration revealed that the increase in
particle diameter cannot be described by the relationship predicted by the LSW theory for diusion-controlled processes.
Instead, it could be shown that the influence of the surface
integration reaction has to be taken into account, too, to
provide good agreement with the experimental data in the
entire time interval investigated.28
The amount of the base also plays a role in the particle generation and size evolution. This parameter was studied by
varying the amount of NaOH while keeping the concentration
of Zn(OAc)2 constant. But also here, the results are contradicting within dierent investigations. Whereas in one case particle nucleation, growth and ripening was constant for small
NaOH concentrations and diusion-controlled ripening was
enhanced for an excess of NaOH,27 another study over a larger
Zn-to-OH ratio revealed similar problems in fitting the data
with the expression for diusion-limited OR as for the study
on temperature dependence.21 Good agreement between calculated and experimental values could only be obtained by enabling
a variation of the value of x in the expression dx dx0 = Kt.
The best fits were received for x = 5.7 7.3,21 which corresponds to a much slower ripening rate compared to the expectations for diusion-controlled processes.
The last parameter investigated was the influence of the
overall precursor concentration while keeping the ratio
between Zn2+ and OH constant. In the Zn(OAc)22H2Oisopropanol system, neither a change in the initial particle size nor
in the ripening kinetics could be detected in the concentration
range explored.27 In comparison to that, ZnCl2 in ethanol
behaved dierently. A shortening of nanoparticle nucleation
and growth period could be detected with increasing the precursor concentration.26 Furthermore, the ripening rates
remained unaected only at low concentrations, whereas at
higher concentrations the rate was significantly enhanced.
This observation could be explained on the basis of smaller
interparticle distances as a result of the hundredfold increase
in overall precursor concentration.26 A closer look at the UV
spectrum in the initial stage of the reaction revealed an
increase in absorbance maximum (however without a shift)
due to particle nucleation.29 The particle size distribution
became narrower during this initial time period, in which particle generation is limited by the molecular transformation of
the precursor species.29 Finally, particle growth based on
ripening processes could be eectively reduced by the addition
of capping agents such as thiols, phosphonic acids or polyvinylpyrrolidone (PVP),3032 whereby the particles were found
to be smaller at any given reaction time for increasing
amounts of PVP.30
The discrepancy between the results of studies on similar
or comparable systems makes it rather dicult to draw general

This journal is The Royal Society of Chemistry 2013

Perspective
conclusions and to develop concepts valid for dierent synthesis approaches. A major limitation of these investigations
might be that they are mainly based on monitoring the change
in particle radii by a shift in UV absorbance, often neglecting
any information about the composition of the final solid
phase, particle aggregation behavior, reaction progress or side
products.
One way to overcome this drawback is the combination of
UV-vis spectroscopy with other characterization tools. The reaction between Zn(OAc)22H2O and LiOH in ethanol was investigated by extended X-ray absorption fine structure (EXAFS), UVvis absorption and X-ray diraction measurements.33 A significant quantity of the precursor and an acetate-intercalated zinc
hydroxy double salt (Zn5(OH)8(OAc)22H2O) were found at the
end of the investigated time period.33 The presence of precursor species may question the interpretations of particle growth
based on ripening processes. The low concentration of the
base in this particular study (Zn/OH ratio of 10) might be the
reason for slowing down the precursor conversion and promoting the formation of the zinc hydroxy double salt as the side
product. Increasing the concentration of the base to a Zn-toOH ratio of 2 and replacing LiOH by KOH, the same authors
prove the presence of both the zinc acetate precursor and ZnO
nanocrystals.34 It was shown that the whole process of particle
formation was much more complex than anticipated from the
evaluation of UV-vis spectroscopic data only. Four stages of
particle evolution with reaction time could be detected (Fig. 2):
(i) nucleation of ZnO with precursor conversion (monitored by
X-ray absorption fine structure (XAFS), UV-vis absorbance, and
SAXS measurements); (ii) particle growth due to the generation
of compact aggregates detected by SAXS as well as by UV-vis
absorbance spectra. The increasing dierence between both
values indicated the formation of a larger fraction of random
aggregates. In this stage, only a small amount of precursor was

Fig. 2 Schematic of the process of formation, aggregation, and growth of ZnO


nanoparticles by the addition of KOH to a zinc oxy acetate ethanolic solution as
examined by in situ and simultaneous time-resolved monitoring of UV-vis
absorption spectra combined with small-angle X-ray scattering and with X-ray
absorption ne structure measurements. (Reprinted with permission from
B. L. Caetano, C. V. Santilli, F. Meneau, V. Briois, S. H. Pulcinelli, J. Phys. Chem. C,
2011, 115, 44044412. Copyright 2011 American Chemical Society.)34

Dalton Trans., 2013, 42, 1255412568 | 12557

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Perspective
converted; (iii) fractal aggregate formation detected by SAXS
measurements with invariant precursor concentration and UVvis absorbance. In the final stage (iv), a second nucleation
event was found accompanied by further conversion of the precursor and a shift in UV-vis absorbance. Considering the simplicity of the system (zinc acetate hydrate and KOH in
ethanol), the complexity of the chemical formation and crystallization mechanism is amazing.
In comparison with investigations focused on the study of
the ripening behavior, nucleation and growth kinetics are even
less addressed, probably due to the short time scale within
which these processes take place. A spectroscopic tool, which
oers the resolution of such fast processes and the determination of nucleation and growth rates in dependence of the
reaction time, is hyper Rayleigh scattering.35 Using this technique, it was found that nucleation and growth were terminated within 5 s after the addition of a LiOH solution to an
ethanolic solution of Zn(OAc)22H2O. This short period of
nucleation and growth is, however, opposed to the results
obtained in the same solvent, but with lower concentrations of
zinc acetate dihydrate19 or for zinc chloride as a precursor.26,29
In the case of alkylzincalkoxides, the influence of the synthesis parameters on nuclei formation and their crystallographic structure was elucidated by EXAFS.36 The critical
nuclei already had the wurtzite structure, although a change in
interatomic distances and a distortion of the tetrahedral
coordination were observed due to the large number of atoms
located at the surface.36
Altering the overall reactant concentration, while keeping
the ratio between zinc oxide precursor and water constant,
revealed the presence of three regimes for particle generation.
Initially, the resulting particle size increased monotonically
with the precursor concentration. In regime two, a further
increase of the amount of precursor resulted in a large supersaturation, which led to the generation of a higher number of
nuclei and finally smaller particles. In the third regime, the
final particle size grows with precursor concentration, since
the amount of residual precursor after nucleation is sucient
to contribute significantly to particle growth. In addition, the
size of the resulting particles could be inversely related to the
dielectric constant of the solvent. This correlation is based on
a reduction in nucleation rates with decreasing dielectric constant due to an increase in interfacial tension. Interestingly,
the influence of surface tension dominates over the increase in
supersaturation (lower solubility with decreasing dielectric
constant) in the present case,36 unlike the behavior of nucleation rates observed in dierent n-alkanols.19
Up to here, we have only discussed the more frequently
studied hydrolytic reaction systems. However, a small number
of investigations were dedicated to nonaqueous media, too.
One example, in which the influence of the system reactivity
on the particle size and size distribution was investigated,
involved ZnO generation via an esterification reaction in
organic solvents. On the addition of dierent amounts of paratoluenesulfonic acid monohydrate ( p-TSA), acting as a catalyst
and possibly as a coordinating agent, it was detected that the

12558 | Dalton Trans., 2013, 42, 1255412568

Dalton Transactions
mean particle size as well as the particle size distribution of
the ellipsoidal ZnO nanoparticles decreased with increasing
the amount of p-TSA at short reaction times.37 The catalytic
eect of p-TSA accelerated the nucleation rate and generated
smaller particles with narrower size distribution. On the other
hand, the decrease in pH with the addition of p-TSA improved
the solubility of ZnO, resulting in lower reaction yields and
size distribution broadening due to enhanced ripening rates at
longer reaction times.37 Zinc oxide particles, synthesized in
dierent alkylamines, became smaller if the molar ratio
between the metal oxide precursor and tert-butylphosophonic
acid (TBPA), which served as capping agent, decreased.38
The formation of ZnO from zinc acetate in benzyl alcohol
after microwave heating also occurred via esterification.39 The
growth of the ZnO nanoparticles followed the LSW model for
coarsening, pointing to a diusion-limited process. Comparison of the microwave-mediated particle growth with conventional heating showed that microwave irradiation increased
the rate constant for crystal growth by a factor of about 4.39
At the end of this section addressing the dependence of
nucleation, growth and ripening behavior on chemical and
physical parameters, we try to summarize the major findings.
As expected from classical crystallization theory, an increase in
supersaturation leads to an increase in nucleation rates generating a higher number of smaller nuclei. Dierent strategies
were developed to increase supersaturation, e.g., precursors
with a higher reactivity or a decrease in product solubility. On
the other hand, an increase in surface tension decreases the
nucleation rates.
Although many factors aect the ripening rates, the two
most important ones include (i) product stability under the
given conditions, i.e., solubility and surface tension and (ii)
solute mobility, influenced by the viscosity of the solvent, in
the case of diusion-controlled ripening.
In nonaqueous systems, the change in particle size is often
accompanied by a change of particle shape due to variation of
the synthesis parameters like solvents and capping agents.
These eects will be discussed in the following section.

Nanoparticle morphology
ZnO nanoparticles generated under thermodynamically controlled crystal growth conditions tend to exhibit a faceted
spherical shape. Based on the energies of the dierent crystal
facets in the hexagonal wurtzite lattice, a hexagonal prismatic
shape is expected to be the equilibrium crystal habit. Similar
to the equilibrium shape for hexagonal CdSe particles, the
0} surfaces at the side. The top
crystal habit is enclosed by {101
and the bottom of the prism are terminated by the (0001) and
) planes, respectively, with some degree of rounding
the (0001
between these two families of planes (Fig. 3a).40 However, due
to the intrinsic anisotropy of the crystal lattice together with
the resulting significant dierences in surface energies
0} and the {0001} faces, anisotropic crystal
between the {101
growth along the c-axis is facilitated in kinetically controlled

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Dalton Transactions

Fig. 3 Equilibrium shape of CdSe particles (a)40 and growth habit for ZnO
under hydrothermal conditions (b).41

growth regimes. Thus, the idealized growth habit under hydrothermal conditions found by Laudise and Ballman corresponds to a faceted hexagonal shape, elongated along the
) being the basal plane and a
[0001] direction with the (0001
1} planes as side surfaces
hexagonal pyramid tip with the {101
41
(Fig. 3b). In fact, the formation of ZnO nanoparticles with a
rod-like morphology is often observed.4255
In principle, nanorods can be generated either based on
classical or non-classical crystallization mechanisms (e.g.,
oriented attachment). In the following section, we will discuss
nanorod formation based on the classical crystal growth and
ripening mechanism, whereas the non-classical processes will
be presented in the next chapter.
Rod formation by the hydrolysis of zinc oxide precursors by
a base in alcoholic media could partially be explained by particle ripening processes.50,56 Compared to the conditions typically applied in studies of ZnO ripening behavior discussed in
the previous section (classical nucleation, growth and ripening
behavior), the ratio between the base and the zinc oxide precursor is approximately 4 to 16 times larger.50 The growth rate
along the c-direction is enhanced by increasing the carbon
chain length of the alcohol used as a solvent from methanol to
butanol.50,56 Because nanorod formation is mainly based on a
particle ripening mechanism, this finding goes along with the
enhancement of ripening rates with increasing the carbon
chain length of the solvent.19
The anisotropic growth can eectively be suppressed by
stronger interfacesolvent interactions, which are provided by
alcohols with shorter carbon chain lengths and thus larger
dielectric constant.50,56 A similar correlation between nanorod
and carbon chain length was reported for amines for the
oxidation of dicyclohexylzinc(II) by moisture at room
temperature.43,49
Aging experiments at elevated temperature in 1,4-butanediol showed a logarithmic dependence of the nanorod volume
on the aging time.47 In comparison to ethanol, the ripening
rates in 1,4-butanediol were significantly smaller, which was
attributed to the increase in viscosity from ethanol to 1,4-butanediol47 but might also be due to a hindrance of anisotropic
growth by increased interfacesolvent interactions.
In nonaqueous approaches, e.g., the solvothermal reaction
of zinc acetate in dierent aliphatic alcohols, it was also found

This journal is The Royal Society of Chemistry 2013

Perspective
that longer nanorods were formed with an increase in carbon
chain length of the respective alcohol.46 The use of glycols,
however, eciently suppressed the preferential growth along
the c-direction.46 On the other hand, enhanced growth along
the c-direction leading to nanowire formation with lengths up
to several m was observed for inert solvents like n-alkanes
without a reactive functional group. This was attributed to the
lack of interfacesolvent interactions.46
Zn(OAc)2 in oleylamine, dodecylamine or hexadecylamine
reacted into ZnO nanorods, following the same trend that
longer carbon chains resulted in longer rods.42 However, a
decrease in aspect ratio was observed at a higher amine-to-zinc
oxide precursor molar ratio.42 For example, tetrahedral prisms
rather than nanorods were generated with an amine-to-zinc
acetate ratio of 4 : 1. In this particular case, the ZnO was
formed via an amidation reaction, and therefore a higher
amount of amine enhanced the reactivity of the system and
led to the generation of a higher number of nuclei. The
remaining monomer concentration after particle nucleation
was low and anisotropic growth along the c-direction was eectively suppressed.42 Interestingly, by adding oleic acid to the
reaction system, anisotropic growth can be restored.54
Another example, where the reactivity of the system could
be tuned, was reported for cobalt and manganese doped ZnO
particles.51 The reaction between Zn(OAc)2 and benzyl alcohol
resulted in ZnO by an esterification between the acetate ligand
and benzyl alcohol.51 Reducing the amount of benzyl alcohol
by partly replacing it with anisole as an inert co-solvent led to
an increase in rod length, finally resulting in the formation of
ZnO nanowires if only traces of benzyl alcohol were present.51
Obviously, nanorods with higher aspect ratio tend to form in
systems with decreased reactivity. However, because in this
case the ZnO nanorods are doped, any influence from the
doping process, i.e., the presence of additional precursors,
cannot be completely excluded.
Facet-selective adsorption of capping agents is a popular
tool to control the shape of nanoparticles, and thus can also
be used for the enhancement of the growth rate along the
c-direction to produce nanorods with higher aspect ratios. In the
following paragraph, we will review approaches leading to
nanorod generation by the usage of capping agents. The
addition of polyvinylpyrrolidone (PVP) to a synthesis mixture
for ZnO nanorod growth under hydrothermal conditions
resulted in larger aspect ratios.48 Although nanorod formation
was independent of the presence of a capping agent, the
increase in aspect ratio, both by reducing the width and
increasing the length of the final structures, resulted from a
0} facets.48 The growth
selective adsorption of PVP on the {101
of the capped nanorods was found to be both diusion- and
surface integration-controlled, because the surface adsorbed
PVP established a barrier.48 In contrast, the change in nanorod
length for the uncapped particles was detected to exhibit the
cubed dependence on reaction time as predicted by the LSW
theory for pure diusion-controlled ripening.48
To summarize, ZnO nanorods are often obtained at elevated
temperatures, high remaining monomer concentration after

Dalton Trans., 2013, 42, 1255412568 | 12559

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Perspective
nucleation, favorable interactions of solvents or capping
agents with the surface and increased particle ripening. But it
is obvious that these statements are not very precise and
should therefore only be regarded as general trends. Nevertheless, they can give useful hints for the prediction of whether a
given reaction system might lead to the generation of ZnO
nanoparticles with a rod-like morphology or not.
In addition to the formation of particles with a rod-like
morphology, many other shapes for zinc oxide can be found in
the literature. Due to the fact that the number of synthesis
approaches used for the preparation of these nanostructures is
as broad as the variety of their shapes, it is rather dicult to
establish general concepts or overarching rules that would
make it possible to understand and predict the formation of
such complex shapes. In the following sections we will
describe a selection of ZnO nanoparticle morphologies, but
without going into details about their possible growth
mechanism.
Urchin-like structures are, from a morphological point of
view, closely related to nanorods, because they consist of
several rods that are connected at a central point (Fig. 4). In
general, the individual spines display a hexagonal rod-like

Fig. 4 Scanning electron microscopy (SEM) image of urchin-like ZnO nanostructures generated by a CTAB assisted hydrothermal process (a) and schematic
representation of particle morphology specifying the growth direction of the
spines (b). (Micrograph reprinted with permission from H. Zhang, D. Yang, Y. J. Ji,
X. Y. Ma, J. Xu, D. L. Que, J. Phys. Chem. B, 2004, 108, 39553958. Copyright
2004 American Chemical Society.)57

Dalton Transactions
morphology elongated along the c-direction, enclosed by six
0} surfaces at the side and a pointed end.5760 ZnO par{101
ticles with such morphological characteristics can be generated by aqueous approaches using Zn(OAc)22H2O as the metal
oxide precursor and NaOH or sodium peroxide.58,59 The same
particle shape can be obtained by substituting sodium hydroxide with ammonia and applying microwave irradiation.60
However, whereas the widths of the single spines in all these
examples (with the exception of the microwave irradiated
sample) are predominantly situated in the hundred nanometers range, their lengths are in the 110 micrometer
range.5760
Keeping the hexagonal shape of the basal plane and the
pointed geometry of the tip, but reducing the elongation along
the c-direction, hexagonal pyramids were obtained.6163 The
reaction between zinc stearate in 1-octadecene with 1-octadecanol or anhydrous Zn(OAc)2 in 1-hexadecylamine with
1,12-dodecanediol resulted in hexagonal pyramids in a size
range of tens of nanometer (Fig. 5a and b).61,62 These struc1} facets and the (0001
) as the
tures were enclosed by six {101
61,62
basal plane (Fig. 5b and c).
Using a zinc acetatetrioctylphosphine oxide (TOPO) complex in dioctyl ether together
with 1,12-dodecanediol and oleic acid generated a similar
nanocone-like morphology (Fig. 6a). Again, the growth direction was found to be the c-axis, but this time the side planes
) rather
seemed to be established by step edges on the (0001
62
than by defined crystal planes (Fig. 6b and c). If the same
synthesis conditions were applied, but without oleic acid, the
zinc oxide nanocones assembled into spheres with diameters
of about 900 nm due to their high surface energy.62
Further suppression of the growth along the c-axis with still
suciently high growth rates in the a- and b-direction inverts
the shape anisotropy and particles with a hexagonal disk-like
morphology were obtained (Fig. 7). Taking advantage of the
strong coordination ability of -hydroxy carboxylic acids like
sodium malate or potassium citrate, the synthesis using
Zn(OAc)22H2O and ammonia in the presence of these ligands
under hydrothermal conditions generated stacked hexagonal
ZnO disks.64,65 Single microdisks with relatively smooth

Fig. 5 Transmission electron microscopy (TEM) (a) and high resolution transmission electron microscopy (HRTEM) image (b) of ZnO nanocrystals exhibiting a hexagonal pyramid morphology (inset b: the corresponding fast Fourier transform (FFT) pattern) with a schematic representation of the corresponding crystal habit (c).
(Micrographs reprinted with permission from Y. F. Yang, Y. Z. Jin, H. P. He, Q. L. Wang, Y. Tu, H. M. Lu, Z. Z. Ye, J. Am. Chem. Soc., 2010, 132, 1338113394. Copyright
2010 American Chemical Society.)61

12560 | Dalton Trans., 2013, 42, 1255412568

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Dalton Transactions

Perspective

1
0] zone
Fig. 6 TEM (a) image and selected area electron diraction (SAED) (b) pattern of a single oleic acid stabilized cone-shaped ZnO nanocrystal along the [21
axis (inset b: the corresponding TEM image). Scheme of the crystal morphology growing along the [0001] direction (c). (Images (ab) reprinted from J. Joo,
S. G. Kwon, J. H. Yu, T. Hyeon, Adv. Mater., 2005, 17, 18731877, with permission from Wiley & Sons.)62

Fig. 7 (a) TEM image of a hexagonal disk synthesized by microwave irradiation


with HMTA and triethyl citrate. (b) HRTEM image taken at the center of the red
hexagon in (a) and (c) SAED pattern. (d) The corresponding crystal habit.
(Images (ac) reprinted with permission from S. H. Cho, S.-H. Jung, K.-H. Lee, J.
Phys. Chem. C, 2008, 112, 1276912776. Copyright 2008 American Chemical
Society.)60

surfaces can be obtained with increasing the amount of the


capping agent.64 Using tripotassium citrate and higher pH
values resulted in very thin ZnO sheets which further assemble
to form sand rose-like particle shapes.66 A similar decrease of
the growth rate along the c-direction as observed for -hydroxy
carboxylic acids was found for poly(sodium 4-styrenesulfonate)
(PSS) as a capping agent, finally leading to the formation of
zinc oxide microdiscs which self-assemble to form spherical
superstructures.67 The same hexagonal disk-like morphology,
but with smaller particle sizes, was prepared by an aqueous
approach with Zn(NO3)26H2O, hexamethylenetetramine
(HMTA) and triethyl citrate using microwave irradiation
(Fig. 7).60 Electron diraction (ED) and HRTEM analysis60
0} facets at the side
revealed the disk to be enclosed by six {101
and two {0001} facets at the top and the bottom, leading to the
crystal habit displayed in Fig. 7d. ZnO microparticles with a
hexagonal disk-like morphology were also observed by reacting
only Zn(OAc)22H2O with HMTA in a waterethanol mixture
without any capping agent.68
The morphological changes of the ZnO nanoparticles in the
examples mentioned so far were induced by an alteration of
the relative growth rates of the dierent crystal facets either by
specific reaction conditions or by the addition of capping
agents. However, alteration of the crystal phase of the nuclei

This journal is The Royal Society of Chemistry 2013

may also lead to the evolution of nanoparticles with dierent


morphological characteristics. The formation of IIVI semiconductor tetrapods, for example, was attributed to the epitaxial
growth of wurtzite arms, elongated along the c-direction, on
the {111} facets of a zinc blende core.6971 In contrast to these
compounds, polytypism seems to be less probable for the
strongly ionic zinc oxide due to the high energy dierence
between both crystal structures. Nevertheless, the synthesis of
tetrapod shaped nanoparticles has also been reported for
doped as well as for undoped ZnO.61,72 The detection of a zinc
blende core by HRTEM studies, performed on zinc oxide tetrapods obtained by thermal vapor deposition and on manganese
doped particles prepared by a nonaqueous wet chemical
approach, suggested that a similar mechanism might also be
valid for ZnO (Fig. 8).61,73
The broad variety of shapes of ZnO nanoparticles and the
related diversity of synthesis protocols make it extremely challenging to predict the conditions required for a specific particle shape. On the other hand, completely dierent synthesis
routes produced nanoparticles with similar morphologies,
which complicates the elaboration of the main factors determining the shape. Clearly, the chemistry involved in ZnO
formation can be complex and also the crystallization
mechanism is often hardly understood. But in addition to
that, there are many other serious issues hampering the introduction of rationality into ZnO synthesis: (i) time-dependent
studies on ZnO nanoparticle evolution are still relatively
scarce; (ii) ligand exchange reactions after precursor dissolution with coordinating solvents or capping agents may lead
to the formation of monomeric species that strongly dier
from the starting precursor, possibly changing the whole molecular mechanisms; (iii) side reactions of the organic compounds present in the precursor, as a solvent and/or as a
capping agent, may occur, and these in situ generated organic
species might interact with the surface and alter the environment for particle nucleation and growth. These factors
together with experimental parameters like surfactant-to-precursor ratios, the amount of base, i.e., pH value, for hydrolytic
processes, solvent viscosities and dielectric constants at the
reaction temperature and homogeneity of the reaction

Dalton Trans., 2013, 42, 1255412568 | 12561

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Perspective

Dalton Transactions

Fig. 8 TEM image of the tetrapod-shaped ZnO nanoparticles generated by using a molar ratio of 95 : 5% zinc stearate to manganese stearate (a). The corresponding HRTEM image of the center of a bipod (b) and two-dimensional atomic simulation (c). Inset b: FFT pattern of the arm region (left) and the core (right). (Adapted
with permission from Y. F. Yang, Y. Z. Jin, H. P. He, Q. L. Wang, Y. Tu, H. M. Lu, Z. Z. Ye, J. Am. Chem. Soc., 2010, 132, 1338113394. Copyright 2010 American Chemical
Society.)61

Fig. 9 TEM image of the pearl chain intermediates after reuxing the concentrated particle solution for 2 h (a) and HRTEM images of these structures (be). TEM
and HRTEM (right) image of the nanorods obtained after one day of reux (f ). (Micrographs reprinted from C. Pacholski, A. Kornowski, H. Weller, Angew. Chem., Int.
Ed., 2002, 41, 11881191, with permission from John Wiley & Sons.)75

solutions represent the main reasons for the diculties


observed in establishing general mechanistic know-how.
Therefore, the main goal of this section was to provide an
insight into the various ZnO morphologies reported in the literature, rather than to establish rules for the generation of a
specific particle shape.

Oriented attachment
In the previous sections, examples of zinc oxide nanoparticle
growth and morphological evolution with a focus on classical
crystallization mechanisms were presented and discussed.
However, growth and shape evolution might also involve nonclassical, particle-based growth mechanisms. These processes
lead to the formation of basically single crystalline products
and should be clearly distinguished from particle self-assembly, where the particles are not crystallographically oriented
with respect to each other.
Experimental evidence for non-classical crystallization of
ZnO particles has already been found more than twenty years
ago for the generation of microrods, although the novelty or
peculiarity of the crystallization mechanism was not recognized at that time.74 A first proof of particle-based concepts
being valid for the generation of single crystalline ZnO particles was given by Pacholski et al. more than ten years later.75

12562 | Dalton Trans., 2013, 42, 1255412568

They reported the formation of single crystalline zinc oxide


nanorods from spherical particles by oriented attachment (OA)
resulting in pearl chain-like intermediate structures (Fig. 9).
HRTEM analysis revealed the almost perfect alignment of the
lattice planes of two adjacent primary particles within these
intermediate chains (Fig. 9be). Time-dependent experiments
gave strong evidence that the major reaction path for the formation of single crystalline nanorods is the OA of quasi spherical particles. Furthermore, the authors showed that the
concentration of primary particles played a crucial role for OA
to occur. Rod formation was only obtained for refluxing a previously concentrated sol. If, however, a more diluted sol was
refluxed, only an increase in particle diameter from initially 3
to 5 nm was observed, clearly indicating a ripening mechanism.75 In the final product with the nanorods, no spherical
building blocks were detected anymore, because dissolution
and recrystallization at the bottlenecks consumed them and
led to nanorods with smooth surfaces.75
A follow-up study underlined the importance of reflux time
and stabilization of the primary particles within the sol.75,76
Prolongation of the reflux time led to longer nanorods, and
weaker stabilization of the primary particles resulted in the
direct formation of nanorods during particle synthesis.75,76
The reaction of 0.68 M Zn(OAc)22H2O with KOH in methanol under reflux, which basically resembles the synthesis conditions applied for the highest concentration values by

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Dalton Transactions
Pacholski et al.,76 was reinvestigated by the group of Peukert.77
By a time-dependent investigation of the nanoparticle morphology, they were able to confirm the hypothesis that OA is
the dominant mechanism within the initial time period. In
agreement with earlier studies,75,76 this period of nanorod
growth by OA is followed by an increase in rod length and
width based on ripening processes.77 At higher reactant concentration, the anisotropic OR in the second stage of nanorod
generation was found to be a diusion-limited process,
whereas at lower precursor concentration the ripening was
found to be limited by the surface integration reaction.77
In addition to unidirectional OA, also reports on the generation of ZnO particles based on a 3D oriented attachment
process exist. Ellipsoidal particles were obtained by a controlled 3D OA of nearly spherical and preformed particles in a
mixed solvent system (Fig. 10a).78 Careful adjustment of the
solution properties led to a decrease in OA rate along the
c-direction, but an increase perpendicular to it, resulting in
the generation of porous ZnO ellipsoids.78 Single crystalline
bicone-like zinc oxide particles were prepared by a particle
mediated process within an aqueous system of Zn(NO3)26H2O
as the metal oxide precursor and ethanolamine.79 The angularity of the particles could be tuned from hexagonal bipyramidal
to quasi spherical (two hemispheres) by simply altering the
reaction conditions.79 Although the authors did not present any
time-dependent mechanistic studies, the presence of rough
surface structures points to particle generation involving OA.
Further increase in three dimensionality of the resulting structures starting from similar bicone-like particles and finally
resulting in the generation of ZnO flowers was achieved in a
sodium ascorbate-assisted aqueous solgel process (Fig. 10b).80
The primary particles were in a size range of 20 to 30 nm and
secondary attachment of spherical building blocks specifically at
the contact area between the two cones induced the formation
of petals, so that flower-like morphologies were formed. The
single crystalline nature of the cones irrespective of whether
they were present in the bicone structure or served as petal
within the ZnO flowers could be proven by means of SAED.80

Perspective
3D OA of primary nanoparticles can also lead to spherical
morphologies.8183 The resulting products of the solvothermal
reaction between Zn(NO3)26H2O as precursor species and N,Ndimethylformamide in the presence of PVP as a capping agent
exhibited a twinned spherical morphology (Fig. 10c).81 The
size of the spherical superstructures is adjustable by the
experimental conditions in a range from 100 nm to several
micrometers.81 Both of the hemispheres were found to consist
of small PVP capped primary crystallites which were, from a
crystallographical point of view, perfectly aligned with respect
to each other. However, orientational mismatch between the
two hemispheres was frequently detected leading to a second
set of spots in the SAED pattern. The combination of OA
together with OR of the small primary particles in a later stage
of the reaction produced a cavity inside the spherical mesocrystals.81 But hollow structures may also be generated solely
by OA of spherical primary particles without the combination
of a consecutive OR process. Hexagonal bullet-like particles
with an interior void were obtained under hydrothermal conditions by a stepwise OA process.84 After the particle mediated
generation of the hexagonal base, the architecture further grew
by an upward addition of primary particles. Because this
process was more pronounced at the rim of the hexagonal
base, sloped side walls were formed. Finally, closure of the top
resulted in a hollow bullet-like particle morphology.84
Spherical agglomerates with a diameter of 200 nm can also
be generated by the OA of 46 nm ZnO nanoparticles in a
mixed waterisopropanol system using triethanolamine
without a capping agent.82 Considering that ethanolamine79
and triethanolamine lead to rather dierent particle assemblies underlines the strong influence of these compounds on
the primary particle generation as well as on their OA behavior, which is most probably due to dierent surface adsorbing
properties.
In addition to these relatively simple shapes, 3D OA can
also yield more complex architectures. Hexagonally shaped
ZnO, comparable to the one shown in Fig. 7, however with
clearly distinguishable nanoparticulate subunits, was obtained

Fig. 10 TEM overview and higher magnication (inset; scale bar: 20 nm) image of ellipsoids assembled from 5 nm ZnO particles (a). Field-emission scanning electron microscopy (FESEM) images of nanoowers generated by an ascorbate (b) and microspheres obtained via a PVP assisted oriented attachment process (c).
(Image (a) reprinted with permission from Y. X. Liu, D. S. Wang, Q. Peng, D. R. Chu, X. W. Liu, Y. D. Li, Inorg. Chem., 2011, 50, 58415847. Copyright 2011 American
Chemical Society.78 Image (b) reprinted with permission from M. Raula, M. H. Rashid, T. K. Paira, E. Dinda, T. K. Mandal, Langmuir, 2010, 26, 87698782. Copyright
2010 American Chemical Society.80 Image (c) reprinted with permission from K. X. Yao, H. C. Zeng, J. Phys. Chem. C, 2007, 111, 1330113308. Copyright 2007 American Chemical Society.81)

This journal is The Royal Society of Chemistry 2013

Dalton Trans., 2013, 42, 1255412568 | 12563

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Perspective
by a nonaqueous process.85 Nevertheless, the SAED pattern
showed spots typical of single crystalline particles.
Aging of a colloidal solution of 20 nm Zn/ZnO coreshell
spheres resulted in tree branch-like particle assemblies.86
SAED together with HRTEM analysis proved that they grew by
an OA process. However, some misorientations between the
single spherical particles were detected giving the reason for
the branching events. Fine tuning of the stability of the colloidal solution was essential as insucient stability led to
random aggregation of the particles. Reduced aging times
resulted in a higher number of free particles not yet attached
to the branched structures.86
Primary crystallites with a dierent morphology than
spherical are also capable of forming secondary particles and
mesocrystals. Three-dimensional OA of pyramidal primary
nanoparticles was observed in the case of ZnO nanoflowers
synthesized by hot injection under limited ligand protection.
Of course the attachment process was inhibited if the amount
of capping agent was high enough and only individual pyramidal particles were detected.87
Another class of particle morphology for which an
increased tendency for OA is expected due to the presence of a
large fraction of the polar, high energy facets is the platelet
shape. Indeed, stacking of primary nano- and microplates
leading to more compact secondary structures such as spheres
or doughnut-like shapes is frequently observed.67,88 A closer
look at the mechanism leading to spherical, micron sized
superstructures composed of highly oriented 1020 nm thick
nanoplates (Fig. 11) revealed a combination of both OA of the
primary building blocks and dipole field directed secondary
crystallization. First, a hexagonal, twinned mesocrystal was
generated as the core of the structure by the oriented arrangement of PSS capped hexagonal nanoplates, leading to the
establishment of a macroscopic electric dipole field, which
then attracted polymer-stabilized amorphous intermediates.
These intermediates crystallized in the preferred nanoplatelet
shape and, in a subsequent step under the synergistic eect
of the electric field generated by the core and their own
dipoledipole interactions, the nanoplatelets rearranged into a
highly oriented dense shell of platelets around the central
core.88

Dalton Transactions
Examples such as the generation of spherical superstructures from thin nanoplatelets,88 hollow spheres from primary
nanodots81 or the increase in nano-rod size after the oriented
attachment process7577 indicate that a clear separation
between classical and non-classical crystallization mechanisms
for nanoparticle generation is not always straightforward.
Additionally, the general belief of OA occurring only in parallel
to coarsening processes89,90 might not be true for all systems
as discussed above.88

A case study: zinc oxide nanoparticle


formation in benzyl alcohol
Discrimination between classical and non-classical crystallization pathways is often complicated by the fact that several processes occur in parallel. In such cases, carefully conducted
time dependent studies combining dierent characterization
tools are required. An example along these lines is the formation of fan- and bouquet-like ZnO nanostructures from zinc
acetylacetonate hydrate in benzyl alcohol.91 The inorganic processes, i.e., the morphological evolution, the assembly behavior of the nanoparticles as well as their crystallinity, were
monitored by electron microscopy techniques, and the progress of the chemical reaction ( precursor conversion and thus
the generation of the monomeric species) was followed by
quantification of benzyl acetate as the product of the reaction
between the acetylacetonate ligand and the solvent
(Scheme 1). The zinc hydroxy species represents the monomer
for nanoparticle formation, building up ZnOZn units under
release of water.
Fig. 12 shows TEM images after dierent reaction times.
They nicely display the step-wise transition from relatively
undefined agglomerates (Fig. 12a) to fir tree-like morphologies, which were proposed to form by OA (Fig. 12b), and
to the final fan/bouquet-like architectures consisting of angled
nanorods that are connected at one end (Fig. 12f ). Clearly,
such a complex morphological evolution of the particle
arrangement is not a series of distinct transformations from
one shape into the next level at a specific reaction time, but a
rather continuous process during which several events happen

Fig. 11 FESEM image of ZnO mesocrystal microspheres at the end of the synthesis (10 h) (a and b) and detailed inner structure of a cracked microsphere after a
reaction time of 5 h (c) (scale bar: 200 nm). Inset c: magnication of the area marked with the arrows. (Adapted with permission from Z. Liu, X. D. Wen, X. L. Wu,
Y. J. Gao, H. T. Chen, J. Zhu, P. K. Chu, J. Am. Chem. Soc., 2009, 131, 94059412. Copyright 2009 American Chemical Society.)88

12564 | Dalton Trans., 2013, 42, 1255412568

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Dalton Transactions

Perspective

Scheme 1 Reaction mechanism leading to the formation of a zinc hydroxo species (4) involving a nucleophilic attack of benzyl alcohol on one of the carbonyl
groups of the acetylacetonate ligand (1), alcoholysis leading to the formation of benzyl acetate and acetone in its enolate form, and release of benzyl acetate (2)
and, with an H+ ion from the coordinated water molecule, acetone in its enol form (3).91

Fig. 12 TEM images of the evolution process of ZnO nanostructures after dierent reaction times of (a) 1 min, (b) 3 min, (c) 7 min, (d) 10 min, (e) 20 min and (f )
120 min.

Fig. 13 Schematic illustration of the formation mechanism proposed for zinc


oxide fan- and bouquet like structures.91

in parallel. Nevertheless, the final ZnO nanostructures are very


uniform in size and shape.
Fig. 13 summarizes the proposed formation mechanism
for the zinc oxide fans and bouquets, involving crystallization
of primary particles, their oriented attachment and surface
reconstruction inside the agglomerates. We will start our
discussion by focusing on the generation of the primary
nanocrystallites by the classical theories for nucleation and

This journal is The Royal Society of Chemistry 2013

growth. The TEM data gave evidence that crystalline zinc


oxide nanostructures already formed after just a few minutes
of reaction. This is remarkable, because at that point only
about 0.30% of the precursor was consumed according to gas
chromatography (GC) analysis. Clearly, the formation of ZnO
was limited by the reaction generating the monomeric species
(Scheme 1), excluding a purely diusion-controlled growth.
The slow generation of the monomeric species has the consequence that several sequential nucleation events took place,
continuously producing primary particles. Although nucleation had not yet finished, the first nanoparticle assemblies
already evolved. The aggregates displayed a remarkably
uniform outer shape as well as an exceptional degree of
organization with respect to the crystallographic orientation
of connected primary particles. According to HRTEM investigations, already at the initial stage of reaction ZnO agglomerates with aligned domains of primary nanoparticles were
formed, supporting an oriented attachment process. Ongoing
nucleation and growth of primary particles steadily increased
the concentration of ZnO particles, leading to the situation in
which the oriented attachment process was governed by particleparticle interactions. The dipole moment of ZnO
together with the low dielectric constant of benzyl alcohol

Dalton Trans., 2013, 42, 1255412568 | 12565

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Perspective
governed the assembly of primary particles into the fir tree
structures, and these fir tree assemblies represent the starting
point for the generation of fan and bouquet like structures by
further attachment of primary particles. Newly produced
primary particles anisotropically attached to the fir tree structures in an oriented manner along the c-axis, thereby increasing the dipole moment and thus increasing the attractive
force along this direction.
Comparing the results from GC with electron microscopy at
longer reaction times, it is interesting to note that the morphology did not change anymore although only 42% of the
yield of ZnO was reached. The rest of the ZnO, which still
formed with continuous reaction time, did not contribute to
any changes in the morphology, but to an increase of the
lengths and widths of the nanorods building up the final fan/
bouquet-like structures. Accordingly, the continuing reaction
turnover led to the growth of the nanorods as part of the ZnO
nanostructures rather than to the formation of new fan and
bouquet structures. The absence of primary particles in the
final reaction solution and the presence of only smooth
nanorod surfaces suggest that this growth involved classical
incorporation of monomer species at the surfaces of the rods
rather than oriented attachment.

Classical versus non-classical crystallization


of zinc oxide nanoparticles
In the following, we will undertake an attempt to reveal the
conditions responsible for the occurrence of classical or nonclassical crystallization pathways by a comparison of the
results presented in the previous sections.
If interference between classical crystallization mechanisms
and OA can be excluded, because the reactivity of the system is
suciently high to generate the primary particles by nucleation and classical crystal growth basically instantaneously,
then the resulting particle concentration is expected to play a
crucial role for subsequent OA to occur.
Comparison of the systems used for studying the influence
on ripening kinetics of spherical ZnO particles7,15,1921,27,28,92
with similar systems that however led to the generation of
nanorods based on an OA mechanism7577 indicates that
higher precursor concentrations favor the non-classical
pathway. But also here, there are systems that behave dierently. Oskam and Poot obtained particles with a diameter of
4 nm after reacting a particle sol with a starting precursor concentration of 100 mM in ethanol,26 whereas OA led to the formation of nanorods with the same precursor concentration at
slightly higher temperatures in methanol.75 Segets et al.
studied the nucleation and growth kinetics by hyper-Rayleigh
scattering.35 The particles synthesized with an initial precursor
concentration of also 100 mM did not undergo any rod formation by OA at 20 C in ethanol.35 The same group reported,
on the one hand, an OA mechanism to be active under reflux
conditions in methanol for a precursor concentration of

12566 | Dalton Trans., 2013, 42, 1255412568

Dalton Transactions
460 mM but, on the other hand, no rod formation could be
detected at a concentration of 70 mM,77 suggesting a concentration of 100 mM to be the limit between both mechanisms.
Furthermore, the fact that rod formation occurred in methanol
but not in ethanol and that the temperature as well as the
aging time influenced the formation of rod-like particles
clearly underlines that also other parameters play an important role during these processes.
The dependence of oriented attachment on the particle
concentration can be understood on the basis of a promoting
eect of interparticle forces such as dipoledipole interactions.
In addition, the magnitude of interparticle forces scales with
the size of the nanoparticle cores. With an enhancement of
the dipole moment with increasing particle size, also the area
of crystal facets with high surface energy increases. Because
sucient faceting is suggested to be a prerequisite for OA,
especially in the lower particle concentration limit or at higher
colloidal stabilities,20,93 classical mechanism might be preferred for very small particles. The colloidal stability is a sensitive parameter, because on the one hand it has to be high
enough to enable the crystallographically oriented alignment
of the primary particles before the attachment process86 and,
on the other hand, weak enough to allow mutual attraction of
the particles for OA. In this regard, surface capping
agents, their amount and adsorption strength play a dominant
role. It is reasonable to assume that a high amount of strongly
surface adsorbing species may prevent the oriented attachment process.20,87,94,95 However, if ligands or polymers interact
with the surface in a facet selective way, then they can positively support the OA process,88,96 or they themselves predetermine the shape of the oriented aggregate as observed for
PVP.81,83 Finally, surface capping agents can promote OA in a
more passive way by, e.g., suppressing classical ripening
processes.31
Similarly controversial as the function of the capping
agents in OA processes are other parameters such as solvent
viscosity and reaction temperature. The viscosity exhibits a
comparable eect on classical ripening and on OA kinetics,
which makes it dicult to discuss any trends. A change in
temperature may influence the OA process in several ways. An
increase in temperature enhances the rates for OA, if the OA
is controlled by a high activation energy for particle
fusion.97,98 But at the same time, the increased thermal
energy may dominate over interparticle forces such as dipole
dipole interactions.99 Decreasing the dipole interactions
between the nanoparticulate building blocks by using solvents with high dielectric constants simultaneously hinders
OA and enhances the ripening process due to increased solubility of ZnO.
Clearly, many factors influence the crystallization mechanism of ZnO nanoparticles. Whereas specific conditions in one
system promote a non-classical crystallization pathway, in
another system they may lead to opposed trends. Even if the
synthesis parameters are adjusted in a very subtle way, it
remains a challenge to achieve complete separation of classical
and non-classical mechanisms.

This journal is The Royal Society of Chemistry 2013

View Article Online

Dalton Transactions

Acknowledgements
Financial support from the Swiss National Science Foundation
(200021 119741) and ETH Zrich is gratefully acknowledged.

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

References
1 C. Klingshirn, ChemPhysChem, 2007, 8, 782.
2 D. C. Look and B. Claftin, Phys. Status Solidi B, 2004, 241,
624.
3 D. Vanmaekelbergh and L. K. van Vugt, Nanoscale, 2011, 3,
2783.
4 . zgr, Y. I. Alivov, C. Liu, A. Teke, M. A. Reshchikov,
S. Doan, V. Avrutin, S.-J. Cho and H. Morko, J. Appl. Phys.,
2005, 98, 041301/1.
5 R. Viswanatha, S. Sapra, B. Satpati, P. V. Satyam, B. N. Dev
and D. D. Sarma, J. Mater. Chem., 2004, 14, 661.
6 L. Spanhel, J. Sol-Gel Sci. Technol., 2006, 39, 7.
7 Z. S. Hu, D. J. Escamilla Ramrez, B. E. Heredia Cervera,
G. Oskam and P. C. Searson, J. Phys. Chem. B, 2005, 109,
11209.
8 A. F. Holleman, E. Wiberg and N. Wiberg, Lehrbuch der
Anorganischen Chemie, Walter de Gruyter and Co., Berlin,
34th edn, 1995.
9 S. C. Abrahams and J. L. Bernstein, Acta Crystallogr., Sect.
B: Struct. Crystallogr. Cryst. Chem., 1969, 25, 1233.
10 C. Wll, Prog. Surf. Sci., 2007, 82, 55.
11 H. J. Fan, A. S. Barnard and M. Zacharias, Appl. Phys. Lett.,
2007, 90, 143116/1.
12 C. F. Klingshirn, B. K. Meyer, A. Waag, A. Homann and
J. Geurts, Zinc Oxide: From Fundamental Properties Towards
Novel Applications, Springer, Heidelberg, 2010.
13 N. Jones, B. Ray, K. T. Ranjit and A. C. Manna, FEMS Microbiol. Lett., 2008, 279, 71.
14 Z. L. Wang, J. Phys.: Condens. Matter, 2004, 16, R829.
15 E. M. Wong, J. E. Bonevich and P. C. Searson, J. Phys.
Chem. B, 1998, 102, 7770.
16 I. M. Lifshitz and V. V. Slyozov, J. Phys. Chem. Solids, 1961,
19, 35.
17 C. Wagner, Z. Elektrochem., 1961, 65, 581.
18 E. A. Meulenkamp, J. Phys. Chem. B, 1998, 102, 5566.
19 Z. H. Hu, G. Oskam and P. C. Searson, J. Colloid Interface
Sci., 2003, 263, 454.
20 Z. S. Hu, G. Oskam, R. L. Penn, N. Pesika and
P. C. Searson, J. Phys. Chem. B, 2003, 107, 3124.
21 R. Viswanatha, H. Amenitsch and D. D. Sarma, J. Am.
Chem. Soc., 2007, 129, 4470.
22 L. Brus, J. Phys. Chem., 1986, 90, 2555.
23 L. E. Brus, J. Chem. Phys., 1983, 79, 5566.
24 L. E. Brus, J. Chem. Phys., 1984, 80, 4403.
25 R. Viswanatha, S. Sapra, T. Saha-Dasgupta and
D. D. Sarma, Phys. Rev. B: Condens. Matter, 2005, 72,
045333/1.
26 G. Oskam and F. Poot, J. Sol-Gel Sci. Technol., 2006, 37, 157.

This journal is The Royal Society of Chemistry 2013

Perspective
27 Z. S. Hu, J. F. Herrera Santos, G. Oskam and P. C. Searson,
J. Colloid Interface Sci., 2005, 288, 313.
28 R. Viswanatha, P. K. Santra, C. Dasgupta and D. D. Sarma,
Phys. Rev. Lett., 2007, 98, 255501/1.
29 A. G. Vega-Poot, G. Rodrguez-Gattorno, O. E. SoberanisDomnguez, R. T. Patio-Daz, M. Espinosa-Pesqueira and
G. Oskam, Nanoscale, 2010, 2, 2710.
30 R. Viswanatha and D. D. Sarma, Chem.Eur. J., 2006, 12,
180.
31 E. M. Wong, P. G. Hoertz, C. J. Liang, B.-M. Shi, G. J. Meyer
and P. C. Searson, Langmuir, 2001, 17, 8362.
32 N. S. Pesika, Z. S. Hu, K. J. Stebe and P. C. Searson, J. Phys.
Chem. B, 2002, 106, 6985.
33 V. Briois, C. Giorgetti, F. Baudelet, S. Blanchandin,
M. S. Tokumoto, S. H. Pulcinelli and C. V. Santilli, J. Phys.
Chem. C, 2007, 111, 3253.
34 B. L. Caetano, C. V. Santilli, F. Meneau, V. Briois and
S. H. Pulcinelli, J. Phys. Chem. C, 2011, 115, 4404.
35 D. Segets, L. M. Tomalino, J. Gradl and W. Peukert, J. Phys.
Chem. C, 2009, 113, 11995.
36 C. Lizandara-Pueyo, M. W. E. van den Berg, A. De Toni,
T. Goes and S. Polarz, J. Am. Chem. Soc., 2008, 130, 16601.
37 M. M. Demir, R. Muoz-Esp, I. Lieberwirth and
G. Wegner, J. Mater. Chem., 2006, 16, 2940.
38 P. D. Cozzoli, M. L. Curri, A. Agostiano, G. Leo and
M. Lomascolo, J. Phys. Chem. B, 2003, 107, 4756.
39 I. Bilecka, P. Elser and M. Niederberger, ACS Nano, 2009, 3,
467.
40 J. J. Shiang, A. V. Kadavanich, R. K. Grubbs and
A. P. Alivisatos, J. Phys. Chem., 1995, 99, 17417.
41 R. A. Laudise and A. A. Ballman, J. Phys. Chem., 1960, 64,
688.
42 Z. H. Zhang, M. H. Lu, H. R. Xu and W.-S. Chin, Chem.
Eur. J., 2007, 13, 632.
43 M. L. Kahn, M. Monge, V. Collire, F. Senocq, A. Maisonnat
and B. Chaudret, Adv. Funct. Mater., 2005, 15, 458.
44 E. De la Rosa, S. Seplveda-Guzman, B. Reeja-Jayan,
A. Torres, P. Salas, N. Elizondo and M. J. Yacaman, J. Phys.
Chem. C, 2007, 111, 8489.
45 J. Zhang, L. D. Sun, J. L. Yin, H. L. Su, C. S. Liao and
C. H. Yan, Chem. Mater., 2002, 14, 4172.
46 S. Kunjara Na Ayudhya, P. Tonto, O. Mekasuwandumrong,
V. Pavarajarn and P. Praserthdam, Cryst. Growth Des., 2006,
6, 2446.
47 N. S. Bell and D. R. Tallant, J. Sol-Gel Sci. Technol., 2009, 51,
158.
48 K. Biswas, B. Das and C. N. R. Rao, J. Phys. Chem. C, 2008,
112, 2404.
49 M. Monge, M. L. Kahn, A. Maisonnat and B. Chaudret,
Angew. Chem., Int. Ed., 2003, 42, 5321.
50 H. L. Cao, X. F. Qian, Q. Gong, W. M. Du, X. D. Ma and
Z. K. Zhu, Nanotechnology, 2006, 17, 3632.
51 G. Clavel, M.-G. Willinger, D. Zitoun and N. Pinna, Funct.
Mater., 2007, 17, 3159.
52 P. Rai, J.-N. Jo, I.-H. Lee and Y.-T. Yu, J. Mater. Sci.: Mater.
Electron., 2011, 22, 1053.

Dalton Trans., 2013, 42, 1255412568 | 12567

View Article Online

Published on 15 April 2013. Downloaded by Aligarh Muslim University on 07/05/2014 06:34:07.

Perspective
53 T. Andelman, Y. Y. Gong, M. Polking, M. Yin, I. Kuskovsky,
G. Neumark and S. OBrien, J. Phys. Chem. B, 2005, 109,
14314.
54 M. Yin, Y. Gu, I. L. Kuskovsky, T. Andelman, Y. Zhu,
G. F. Neumark and S. OBrien, J. Am. Chem. Soc., 2004, 126,
6206.
55 Z. H. Zhang, S. H. Liu, S. Y. Chow and M.-Y. Han, Langmuir,
2006, 22, 6335.
56 B. Cheng and E. T. Samulski, Chem. Commun., 2004, 986.
57 H. Zhang, D. Yang, Y. J. Ji, X. Y. Ma, J. Xu and D. L. Que, J.
Phys. Chem. B, 2004, 108, 3955.
58 Z. Fang, K. B. Tang, G. Z. Shen, D. Chen, R. Kong and
S. J. Lei, Mater. Lett., 2006, 60, 2530.
59 S. H. Cho, J.-W. Jang, J. S. Lee and K.-H. Lee, Langmuir,
2010, 26, 14255.
60 S. H. Cho, S.-H. Jung and K.-H. Lee, J. Phys. Chem. C, 2008,
112, 12769.
61 Y. F. Yang, Y. Z. Jin, H. P. He, Q. L. Wang, Y. Tu, H. M. Lu
and Z. Z. Ye, J. Am. Chem. Soc., 2010, 132, 13381.
62 J. Joo, S. G. Kwon, J. H. Yu and T. Hyeon, Adv. Mater., 2005,
17, 1873.
63 M. Yang, K. Sun and N. A. Kotov, J. Am. Chem. Soc., 2010,
132, 1860.
64 J. B. Liang, J. W. Liu, Q. Xie, S. Bai, W. C. Yu and Y. T. Qian,
J. Phys. Chem. B, 2005, 109, 9463.
65 J. B. Liang, S. Bai, Y. S. Zhang, M. Li, W. C. Yu and
Y. T. Qian, J. Phys. Chem. C, 2007, 111, 1113.
66 W. Zhao, X. Y. Song, Z. L. Ying, C. H. Fan, G. Z. Chen and
S. X. Sun, Mater. Res. Bull., 2008, 43, 3171.
67 J. G. Yu, C. Li and S. W. Liu, J. Colloid Interface Sci., 2008,
326, 433.
68 M. Wang, S. H. Hahn, J. S. Kim, J. S. Chung, E. J. Kim and
K.-K. Koo, J. Cryst. Growth, 2008, 310, 1213.
69 S. Asokan, K. M. Krueger, V. L. Colvin and M. S. Wong,
Small, 2007, 3, 1164.
70 L. Manna, E. C. Scher and A. P. Alivisatos, J. Am. Chem.
Soc., 2000, 122, 12700.
71 L. Manna, D. J. Milliron, A. Meisel, E. C. Scher and
A. P. Alivisatos, Nat. Mater., 2003, 2, 382.
72 X. H. Zhong, Y. Y. Feng, Y. L. Zhang, I. Lieberwirth and
W. Knoll, Small, 2007, 3, 1194.
73 Y. Ding, Z. L. Wang, T. J. Sun and J. S. Qiu, Appl. Phys. Lett.,
2007, 90, 153510/1.
74 M. Andrs Vergs, A. Mifsud and C. J. Serna, J. Chem. Soc.,
Faraday Trans., 1990, 86, 959.
75 C. Pacholski, A. Kornowski and H. Weller, Angew. Chem.,
Int. Ed., 2002, 41, 1188.
76 C. Pacholski, A. Kornowski and H. Weller, Proc. SPIE-Int.
Soc. Opt. Eng., 2004, 5513, 232.

12568 | Dalton Trans., 2013, 42, 1255412568

Dalton Transactions
77 M. Voigt, M. Klaumnzer, H. Thiem and W. Peukert, J.
Phys. Chem. C, 2010, 114, 6243.
78 Y. X. Liu, D. S. Wang, Q. Peng, D. R. Chu, X. W. Liu and
Y. D. Li, Inorg. Chem., 2011, 50, 5841.
79 Q. J. Yu, C. L. Yu, W. Y. Fu, M. X. Yuan, J. Guo, M. H. Li,
S. K. Liu, G. T. Zou and H. B. Yang, J. Phys. Chem. C, 2009,
113, 12016.
80 M. Raula, M. H. Rashid, T. K. Paira, E. Dinda and
T. K. Mandal, Langmuir, 2010, 26, 8769.
81 K. X. Yao and H. C. Zeng, J. Phys. Chem. C, 2007, 111,
13301.
82 Y. S. Fu, Y. F. Song, S. A. Kulinich, J. Sun, J. Liu and
X. W. Du, J. Phys. Chem. Solids, 2008, 69, 880.
83 M. Distaso, R. N. Klupp Taylor, N. Taccardi,
P. Wasserscheid and W. Peukert, Chem.Eur. J., 2011, 17,
2923.
84 K. X. Yao and H. C. Zeng, J. Phys. Chem. B, 2006, 110,
14736.
85 J. Buha, I. Djerdj and M. Niederberger, Cryst. Growth Des.,
2007, 7, 113.
86 H. B. Zeng, P. S. Liu, W. P. Cai, X. L. Cao and S. K. Yang,
Cryst. Growth Des., 2007, 7, 1092.
87 A. Naravanaswamy, H. F. Xu, N. Pradhan and X. G. Peng,
Angew. Chem., Int. Ed., 2006, 45, 5361.
88 Z. Liu, X. D. Wen, X. L. Wu, Y. J. Gao, H. T. Chen, J. Zhu
and P. K. Chu, J. Am. Chem. Soc., 2009, 131, 9405.
89 H. Clfen and M. Antonietti, Mesocrystals and Nonclassical
Crystallization, John Wiley and Sons, Ltd, Chichester, 1st
edn, 2008.
90 J. Zhang, F. Huang and Z. Lin, Nanoscale, 2010, 2, 18.
91 B. Ludi, M. J. Sess, I. A. Werner and M. Niederberger,
Nanoscale, 2012, 4, 1982.
92 P. K. Santra, S. Mukherjee and D. D. Sarma, J. Phys. Chem.
C, 2010, 114, 22113.
93 D. G. Stroppa, L. A. Montoro, A. Beltrn, T. G. Conti,
R. O. da Silva, J. Andrs, E. Longo, E. R. Leite and
A. J. Ramirez, J. Am. Chem. Soc., 2009, 131, 14544.
94 Z. Y. Tang, N. A. Kotov and M. Giersig, Science, 2002, 297,
237.
95 E. J. H. Lee, C. Ribeiro, E. Longo and E. R. Leite, J. Phys.
Chem. B, 2005, 109, 20842.
96 J. Polleux, N. Pinna, M. Antonietti, C. Hess, U. Wild,
R. Schlgl and M. Niederberger, Chem.Eur. J., 2005, 11,
3541.
97 R. L. Penn, J. Phys. Chem. B, 2004, 108, 12707.
98 F. Huang, H. Z. Zhang and J. F. Banfield, Nano Lett., 2003,
3, 373.
99 J. H. Yu, J. Joo, H. M. Park, S.-I. Baik, Y. W. Kim, S. C. Kim
and T. Hyeon, J. Am. Chem. Soc., 2005, 127, 5662.

This journal is The Royal Society of Chemistry 2013

You might also like