You are on page 1of 17

ARTICLE

www.rsc.org/materials
Chemistry
Materials
Journal of
Understanding the factors that govern the deposition and
morphology of thin films of ZnO from aqueous solution{
Kuveshni Govender, David S. Boyle, Peter B. Kenway and Paul O’Brien*

The Manchester Materials Science Centre and Department of Chemistry, University of


Manchester, Oxford Road, Manchester, UK M13 9PL. E-mail: paul.obrien@man.ac.uk;
Fax: 0161-2754616; Tel: 0161-2754652

Received 31st March 2004, Accepted 19th May 2004


First published as an Advance Article on the web 15th June 2004

The influence of the choice of complexing ligand, zinc counter-ion, pH, ionic strength, supersaturation, deposition
time and substrate on the nature of ZnO films grown from chemical baths (CBD) are discussed. There are
significant differences between CBD and similar routes such as hydrothermal methods for ZnO films. Modelling of
speciation and experimental results suggest that acicular ZnO morphologies are best obtained by limiting the
concentration of one of either Zn21 or OH2 in the presence of a large excess of the other. The presence of a prior
ZnO layer can facilitate nucleation at lower levels of supersaturation and enable size tailoring of ZnO columns.
The point at which the substrate is introduced into the bath is crucial and can lead to a significant difference in
both the width of the rods and optical transparency of the films. HR-TEM has yielded important structural
information and a growth mechanism for single crystalline ZnO rods by CBD is described for the first time.

1. Introduction ultimately oxoanions are formed. In order to form the poly-


nuclear species, which subsequently develop into metal oxide
The deposition from aqueous solution or CBD (chemical bath particles, reactions involving condensation reactions must occur.
deposition) of metal oxide thin films involves controlled pre- Two important processes have been recognized, olation is the
cipitation on a substrate via hydrolysis and/or condensation formation of an ‘‘ol’’ bridge by reaction of a hydroxo- and aquo-
reactions of metal ions and/or complexes from aqueous solu- species as follows:
tion.1 In CBD, the crystal morphology is strongly influenced by
experimental conditions including chemical speciation in the M–OH 1 M–OH2 A M–OH–M 1 H2O (1)
solution (ligand, pH, metal counter-ion, ionic strength are all
important), the level of supersaturation, the temperature and Oxolation leads to an ‘‘oxo’’ bridge by the dehydration of
the nature of the substrate. Control of the size, shape and hydroxo-species:
orientation of ZnO crystallites on the substrates is a pre-
requisite for creation of high surface area materials for use in M2–(OH)2 A M–O–M 1 H2O (2)
many types of devices including photovoltaic and optoelec-
tronic devices.2 The fundamental optical physics of ZnO Zinc hydroxide is amphoteric and complexation by OH2
nanostructures as a function of crystallite dimensionality have can lead to soluble species such as ‘‘[Zn(OH)3]2’’ and
only recently been experimentally verified.3 In this paper, we ‘‘[Zn(OH)4]22’’ and hence ‘‘zinc hydroxide’’ is more soluble
have modelled the initial composition of deposition baths and in basic solution than a simple consideration of the solubility
developed strategies for morphological control during CBD of product (y10216) would suggest.4
ZnO, leading to reproducible routes to polycrystalline and In the CBD of ZnO, ligands are employed to keep the free
single crystal array thin films. The mechanisms of growth of zinc ion concentration low. Raising the bath temperature pro-
thin films of ZnO rod arrays have been studied using HR-TEM motes some dissociation of the zinc complex, leading to
with FIB sample preparation techniques. controlled supersaturation of the free metal ion. Zinc is a labile
In order to understand the important physiochemical processes metal ion in aqueous solution and equilibria within stirred
involved in the CBD of ZnO, it is useful to summarise its solution solutions are generally attained quickly. Thermodynamic model-
chemistry. In an aqueous solution, metal cations Mz1 are solvated ling of solutions representing the initial states of deposition baths
by water giving rise to aquo-ions, typically [M(OH2)n]z1. The is a useful aid to understanding these systems.5,6
M–OH2 bond is polarised which facilitatesdeprotonation of the
coordinated water. In dilute solutions, a range of monomeric Morphology of ZnO crystallites
species exist such as ([M(OH2)n2p(OH)p](z2p)1 and other hydroxy The thermodynamically stable phase for ZnO is zincite with
species [M(OH)n] (it is often customary to omit the water); the wurtzite (hcp) structure although two cubic phases, a
metastable sphaleritic (zincblende)7 and high pressure rocksalt
{ Electronic supplementary information (ESI) available: Fig. A,B: film type8 have been reported. The elongated hexagonal zinc oxide
thickness profiles for Zn–en and Zn–TEA systems; Fig. C: SEM images crystal has both polar and non-polar faces. The former are
DOI: 10.1039/b404784b

of ZnO films on ZnO template layers from HMT baths; Fig. D: XRD either Zn-terminated (the (0001) face) or O-terminated (the
patterns of ZnO microcolumns grown on ZnO template layers on (0001̄) face). The latter include (112̄0) and (101̄0) faces. The
TO(F) glass, and on Au/TO(F) glass; Fig. E: SEM images showing morphology of a particular crystal is determined by the slowest
effect of increasing ionic strength on ZnO film growth; Fig. F: SEM
images and grain size distributions of ZnO films from HMT baths;
growing faces. Polar faces with surface dipoles are thermody-
Fig. G: micrographs of ZnO thin film samples (Methods 1 and 2); namically less stable than non-polar faces, often undergo
Table A: thermodynamic data; Table B: rod dimensions for different rearrangement to minimise their surface energy and also tend to
deposition times. See http://www.rsc.org/suppdata/jm/b4/b404784b/ grow more rapidly. Therefore it is necessary to distinguish between

This journal is ß The Royal Society of Chemistry 2004 J. Mater. Chem., 2004, 14, 2575–2591 2575
equilibrium and kinetic growth morphologies. The former corres- both globular and acicular. Formation of ZnO in acidified
ponds to a minimum in the surface free energy. However, ther- baths containing ZnCl2 and HMT was suggested via two
modynamic models that consider morphological development as a routes; direct formation or via a Zn5Cl2(OH)8 intermediate.34
function of internal structure are poor at describing the solution The globular form was predominant at high concentrations of
growth of ZnO at moderate temperatures.9 HMT. Andrés-Vergés and co-workers also used aqueous baths
The morphology of films or powders needs to be optimised containing zinc nitrate and HMT to form ‘‘rod-like’’ pre-
for each application. The morphology has been shown to affect cipitates.20 Individual microcrystals were formed within a
catalytic and photocatalytic activity.10 In ZnO powders, the narrow concentration range of zinc salt and HMT depending
non-polar faces typically account for y80% of the total surface on the salt used, whereas aggregates were formed when the
and the (101̄0) faces can determine the adsorption of com- solution pH was not lowered. A decrease in the pH of the
plexes.11 A wide variety of ZnO crystallite morphologies are solution resulted in the formation of individual needles or
observed for both precipitates and thin films including prismatic microcrystals. These workers first suggested the idea
columnar grains,12–14 rods,2,15 stars5 and spherical habits.2,16 of orientated attachment to explain the acicular growth of ZnO
Despite numerous studies, there is little understanding of the crystallites from solution. Spherical particles formed in solu-
mechanisms and factors that govern the observed morphology. tion tended to aggregate along their polarized anisotropic
Two growth mechanisms are usually discussed for the growth c-axes to yield primary ‘‘rod-like’’ zinc oxide crystallites. Similar
of single crystals: growth-dissolution-recrystallization pheno- observations have been made more recently for quasi-spherical
mena or nanoparticle aggregation.17 ZnO nanoparticles formed in basic methanolic solutions.35
The morphology of ZnO is greatly complicated by its pro- The deposition of thin films of rod-like ZnO arrays on
pensity to twinning. Twinned ZnO crystals have been investi- fluorine doped tin oxide (TO(F)) glass substrates has been
gated and shown to have a fourling structure, in which three reported by Vayssieres, using zinc nitrate–HMT solutions
‘‘legs’’ are related to the fourth (spine) by twinning. The spines similar to the work of Andrés-Vergés’ but in closed vessels.15
are related by twinning along the (112̄2) planes with a slight The success of the method depends on separating homo-
distortion at the twinned faces, such that two angles between geneous and heterogeneous nucleation by controlling the inter-
pairs of spines are 97.6u and four are 116.1u.18,19 In the second facial tension. We have grown ZnO rods on ZnO template
phase of fourling crystal growth, three sheets lying in the three layers on TO(F)-glass using carboxylate precursors in open
planes determined by the leg and the spine grow out from the baths2 and demonstrated room temperature lasing from such
spine. Many deviations from this idealised structure can occur. arrays.36 There is an important distinction to be made between
Fourlings of ZnO have been reported during forced hydrolysis of aqueous solution deposition of ZnO (thin films or precipitates)
zinc nitrate using HMT (hexamethylenetetramine),20 en (ethyl- from open and closed baths; the latter conditions may appro-
enediamine)21,22 and NH4OH23 using hydrothermal methods ach hydrothermal growth. In general, the solubility of ZnO is
and by the decomposition of zinc hydroxide carbonate.24 Under greater under hydrothermal conditions and growth is typically
hydrothermal conditions, bipyramidal twinning occurs via the slower. Imai has also demonstrated the importance of substrate
(0001̄) plane whereas dumbbell-like examples are formed via the and shown that upright ZnO rods could be grown only on
(0001) plane.25 The twinned relations of the crystallites were crystalline ZnO templates.37
found to be influenced by additives in the reaction medium. HMT is commonly used in solution routes to acicular ZnO,
however it is not essential and its role is often misunderstood.
Solution growth of ZnO rod array films The primary role of HMT in aqueous deposition baths is to act
as a pH buffer. The kinetics of decomposition of HMT have
There is interest in highly structured metal oxide thin films as been investigated.38 The conversion of HMT to formaldehyde
high surface area substrates for solar cells (e.g. in the Grätzel and (and ammonia) was found to be pH dependent in buffers of
eta configurations).26,27 One problem can be a need to anneal constant ionic strength, with the reaction half-life decreasing
as-deposited films in order to effect more complete crystal- from 13.8 h at pH 5.8 to 1.6 h at pH 2.0 In a typical CBD
lization. A low-temperature synthetic procedure that requires no experiment, HMT provides a slow controlled supply of OH2
post-deposition annealing step is potentially very useful.28 Few and is a poor ligand for zinc. Significant formation of the
systematic studies have been reported although recently Imai has Zn(OH)2 phase cannot be discounted, therefore a growth
shown that using different substrates with identical baths yields mechanism based on dissolution–reprecipitation of preliminary
different morphologies.29 We have investigated solution-based hydroxy-zinc clusters and zinc oxide is possible in heated
approaches with a defined range of experimental conditions for deposition baths containing HMT.31
growing single crystal arrays of ZnO by CBD.2
The literature describing ZnO precipitates is widely dis-
persed. The formation of high-aspect ratio ZnO crystallites has 2. Results and discussion
attracted interest.30 Earlier studies of ZnO single crystal rods Forced hydrolysis of zinc carboxylates
have not led to a universally accepted mechanism for the crystal
growth. There is some evidence from ageing experiments that The forced hydrolysis described by Matijevic is the simplest
the formation of ‘‘Zn(OH)2‘‘ is a prerequisite for the controlled precipitative process for metal oxides and involves ageing
growth of needle-like ZnO crystallites.31 This observation is of aqueous metal salt solutions at elevated temperatures (80–
consistent with a dissolution–reprecipitation mechanism in 100 uC).23 The pH decreases during the course of the reaction
which Zn(OH)2 acts as a reservoir of zinc. The Zn21 concen- due to the deprotonation of the hydrated cation.
tration is held below the level where undesirable secondary DH
nucleation processes, which can lead to twinning, occur. ½M(H2 O)x nz DA ½M(H2 O)x{1 (OH)(n{1)z
(3)
The use of hexamethylenetetramine (HMT) is reported ??M(OH)n or MOy
widely in the growth of acicular ZnO precipitates and thin
films. An early report in the patent literature describes heating Studies of the speciation of Zn21 in chemical baths allow
aqueous solutions of zinc chloride (ZnCl2) and an organic determination of the initial degree of supersaturation within
compound that form a base (such as HMT) or an acid (such as the system. Speciation calculations were carried out for baths
ethylene chlorohydrin) by hydrolysis.32 The control of ZnO containing only zinc acetate in order to determine the point at
crystallite morphology using HMT has been described by which the solution becomes supersaturated with respect to the
Fujita et al.,33 where precipitates derived from zinc nitrate were Zn(OH)2 (Fig. 1a). At 25 uC this critical value of pH is 6.68, in
globular or rod-shaped whereas those from zinc chloride were general agreement with experimental results.

2576 J. Mater. Chem., 2004, 14, 2575–2591


Fig. 2 Relationship between supersaturation in deposition baths, rate
of crystallite growth and morphology.

Development of crystal morphology


The production of a supersaturated (SS) solution is a pre-
requisite for crystallization and the level of SS is important in
determining the final crystal morphology, as exemplified by
CBD ZnO and shown in Fig. 2. We can broadly identify three
main growth regions, defined by the level of supersaturation.
. for relatively low levels of supersaturation in the region
below SS*, heterogeneous nucleation dominates. Crystal faces
grow via the outward displacement of a growth spiral originat-
Fig. 1 a. Speciation diagram illustrating the hydrolysis of zinc acetate ing from screw dislocations (the so-called BCF mechanism)42 in
(0.025 mol dm23) at 25 uC. Zn(OH)2 (s), represented by AH-2(s), the central region of the face. Well-formed polyhedral crystals
is predicted to precipitate at pH 6.68. A ~ Zn21; B ~ CH3COO2. are generally produced under these conditions.
b. SEM image of ZnO precipitates on TO(F) glass formed from forced . in the intermediate supersaturation range SS* v SS v
hydrolysis of zinc acetate (0.025 mol dm23) at 90 uC. SS**, growth occurs by a mixture of slow spiral growth and
2-dimensional nucleation of clusters on to crystal faces from
Preliminary experiments involved attempts to deposit solution (polynuclear processes). In this region, 2D growth
ZnO films using CBD methods (open baths) on TO(F) glass generally predominates. Since the probability of occurrence of
by forced hydrolysis of zinc salts. Due to the volatility of zinc 2D nucleation is far greater near to edges and corners of faces,
carboxylates as compared to nitrate, chloride and sulfate salts, growth steps will advance from them to yield a crenellated face
deposition of material only occurred for the former. Zinc and often ‘‘hollowed’’ rod morphology.
acetate and zinc formate solutions (0.025 mol dm23) containing . for high supersaturation SS w SS**, homogeneous nuclea-
the TO(F) substrates were heated at 90 uC for 2 hours. tion is more important. The crystal faces become rough and
There are some unique characteristics of ZnO–carboxylate continuous linear growth occurs. In this region, ZnO crystals
systems, for example, Li et al. have demonstrated that forma- adopt a dendritic to spherulitic morphology and with
tion of ZnO from basic solutions of zinc acetate by forced increasing supersaturation, growth kinetics are controlled by
hydrolysis is only possible under conditions which allow the bulk diffusion as the solution becomes more viscous.
release of volatile by-products and is hence facilitated by use of The acicular rod morphologies, typically observed for CBD-
open baths or the periodic release of pressure in hydrothermal ZnO films using baths containing HMT and a zinc salt at
systems.9 Carboxylates are often selectively adsorbed on the moderately acidic pH 5,2 can be explained in terms of a
surface sites on ZnO. Surface defects, on nominally Zn-free kinetically-controlled reaction involving relatively high con-
O-terminated faces, in the form of steps with {101̄0} orientated centrations of one (i.e. Zn21) over the other (i.e. OH2 ) com-
faces that expose zinc cations, have been evidenced by ponent. Similar arguments apply to growth of columnar ZnO
exposure to formic acid with subsequent formation of films using basic baths and ethylenediamine ligand, where the
Zn-bound formate species.39 There is indirect evidence from high–low concentration relationship is reversed.
work involving etching of ZnO single crystals that indicates
carboxylates inhibit the rates of dissolution one hundred fold
and tend to aid ageing and growth of crystallites.40,41 Dis- Influence of ligand on crystallite morphology
solution processes are controlled by Zn–O hydrolysis reactions For cations that are not easily hydrolysed, such as Zn21,
at Zn21 kink sites and catalysed by both H1 and ligands for precipitation of the metal hydroxide can be promoted by
zinc. raising the pH with a base. Lewis bases, such as ethylenedia-
Precipitates were visible in the deposition bath with zinc mine (en) and triethanolamine (TEA), can also form complexes
acetate, but not zinc formate. For the former, the homogeneous with the metal. The complexes can also act as a reservoir for
process was dominant and precipitation occurred almost metal, by buffering the free concentration to below the pre-
immediately upon heating of the solution. The films obtained cipitation point.
were powdery and non-adherent and scanning electron micro- In a ZnO CBD system where both ZnO and ‘‘Zn(OH)2’’ may
scopy of the films revealed poor surface coverage with rod- be formed, the less stable phase in contact with the media will in
like particles, often with dumbbell morphology, randomly general precipitate first. Under conditions such that the
orientated on the substrate with little evidence for perpendi- hydroxide precipitates first, changing the temperature or pH
cular growth (Fig. 1b). The particle size distribution was large can result in the formation of the oxide. The oxide can thereby
and there is also evidence of twinning. form via a dissolution/reprecipitation process or by a phase

J. Mater. Chem., 2004, 14, 2575–2591 2577


transformation. A lattice rearrangement, however, is more recorded to determine whether material such as zinc hydroxide
complicated and depends on the degree of rearrangement and (crystalline or amorphous) was present. No peaks indicative of
dissolution of hydroxide to provide Zn21. The uniformity of zinc hydroxide were observed.
the initial hydroxide particles may be key in controlling Scanning electron micrographs of ZnO films grown from
uniformity of the oxide particles. baths containing en and zinc (as the acetate, formate, chloride
There are several reports in the literature on the effect of or sulfate) provided useful information and, with the exception
ligands in the deposition bath on the morphology of of zinc sulfate, films comprised twinned crystallites that formed
ZnO.20,21,30 The hydrolysis of the cation can be controlled by as starlike particles with needle-like spines (Fig. 4a: i, ii, iii).
a slow release of hydroxide into the metal salt solution. At McBride has postulated that these star-like morphologies grow
elevated temperatures both HMT and urea undergo slow by a modified LaMer mechanism.31 Stars are formed from
chemical decomposition to generate OH2 ions. twinning along the (112̄2) plane of the hexagonal lattice. Cross-
sectional SEM indicated that the film thickness in all cases was
Deposition of ZnO films on TO(F) glass using en as complexing of the order y1 mm. Films formed from sulfate-containing
agent baths were less crystalline and appeared to be a mixture of
small flowerlike particles and larger flat platelets (Fig. 4a: iv).
We have previously demonstrated that good quality ZnO thin More detailed examination of films deposited from baths con-
films may be deposited from baths containing zinc acetate taining en and zinc acetate appeared to show that the indivi-
(0.018 mol dm23) and en (0.042 mol dm23, at a final pH of 11).5 dual arms of the starlike ZnO crystallites possessed some
Further experiments were conducted using similar bath helical structure (Fig. 4b). Similar morphology has been
conditions but with different zinc salts. Films were formed reported for ZnO nanorods grown on nanocrystalline ZnO
from baths containing en and either zinc acetate, formate, templates, using baths containing HMT and zinc nitrate
sulfate or chloride. No films could be deposited from baths modified by addition of sodium citrate.43 It has been suggested
containing zinc nitrate. Speciation calculations were performed that the presence of citrate generates the helical morphology.
for baths containing zinc acetate and en (at 25 uC), the Moreover, the structure closely resembles that observed for
distribution plots indicated that baths are supersaturated with nacreous calcium carbonate in red abalone (gastropod Haliotis
respect to the hydroxide by pH ~ 6.92 (Fig. 3a and ESI:{ Table rufescens) and thus may offer insights into biomineralisation
A for binding constants). Small decreases in stability constant processes. It appears that in the presence of acetate and en,
values at higher temperatures did not produce significant ZnO may also be deposited with similar helical structures.
changes to speciation diagrams. It was found that good quality Growth kinetics were monitored using the QCM technique,
films were produced at pH 11 (at 70 uC, 2 h), such baths are typically an induction period was followed by a growth and a
supersaturated and the concentration of free zinc is very low, short termination phase. The QCM technique is not directly
orders of magnitude lower than [OH2].
All ZnO films were of the zincite phase (JCPDS 36-1451) as
determined by XRD (Fig. 3b). The crystallinity was slightly
improved upon annealing in air at 400 uC. In general, films were
thin (y1 mm) and peaks originating from the TO(F) substrate
were evident in the XRD. Infrared spectra (ATR-FTIR) were

Fig. 3 a. Speciation diagram for Zn–en system at 25 uC. Dashed


line represents zinc hydroxide precipitation point. A ~ [Zn21] ~
0.006 mol dm23 and B ~ [en] ~ 0.042 mol dm23, [CH3COO2] ~ Fig. 4 a. SEM images (top) of ZnO films from en-baths (70 uC, pH 11)
0.012 mol dm23. AH-x represents the soluble zinc-hydroxy species. containing: i. zinc acetate; ii. zinc chloride; iii. zinc formate and iv. zinc
b. X-Ray diffractograms obtained for as-deposited ZnO films from sulfate. [Zn] ~ 0.018 mol dm23, [en] ~ 0.042 mol dm23. b. SEM image
i. en-baths (70 uC). ii. TEA-baths and iii. HMT-baths on TO(F) glass. (below) of ZnO film deposited from zinc acetate–en bath showing indivi-
* Reflections assigned to SnO2–glass substrate (JCPDS 41-1445). dual crystallite arms possessing helical structures as described elsewhere.43

2578 J. Mater. Chem., 2004, 14, 2575–2591


comparable to growth on TO(F)-glass substrates as the micro- in air. Again no evidence for zinc hydroxide in the films was
balance uses an Au-coated crystal. However useful com- obtained from IR spectroscopy. Results obtained from QCM
parisons can be made for different baths, assuming all other measurements were different from those with en-containing
factors (temperature, stirring rate, etc.) are equal. For the baths (ESI:{ Figure B). The induction period (0–500 s)
en-system (ESI:{ Figure A) two distinct linear growth phases was characterised by oscillations, followed sequentially by a
following the induction period could be identified; a short and relatively linear period of rapid growth (y500–1500 s; average
rapid growth period over y250 s (rate ~ 0.12 nm s21) and a rate ~ 0.09 nm s21) and then slower deposition (y1500–2500 s;
slower growth phase from y1500–3000 s (rate ~ 0.01 nm s21). average rate ~ 0.02 nm s21). In comparison to the previously
described en-system, in the TEA system the transition from the
Deposition of ZnO films on TO(F) glass using TEA as ‘‘fast’’ to ‘‘slow’’ growth was relatively abrupt.
complexing agent
Deposition of ZnO films on TO(F) glass using HMT
In order to investigate the sensitivity of particle morphologies
to changing bath chemistry, our existing literature preparation Further studies of the effect of the chelating ligand on ZnO
was followed with minor modifications (pH 12 vs. pH 11 and particle morphology were carried out with HMT, which
TEA : metal ratio 5 : 1 vs. 4 : 1 and a similar metal ion con- decomposes slowly in heated aqueous solutions to yield
centration). Experiments were conducted at 70 uC with TEA ammonia and formaldehyde. The concentration of HMT has
(0.072 mol dm23) using different zinc salts (0.018 mol dm23) on been reported to influence the rate of ZnO formation.34
TO(F) glass substrates as for the Zn–en experiments. The final Speciation calculations for the HMT system indicate that the
pH of the bath was 11. In initial experiments films were only pH at which the bath becomes supersaturated at 25 uC with
obtained from baths containing zinc nitrate. respect to the hydroxide is y6.8 (Fig. 6a). All other factors
Speciation calculations were carried out on baths containing being equal (i.e. ligand and metal concentration), at values of
zinc nitrate and TEA. Supersaturation with respect to the pH w 6, supersaturation in terms of ‘‘zinc hydroxide’’ will
hydroxide occurred at pH 6.44 similar to the en system (Fig. 5a). be greater for HMT than en. Thin film formation requires
Moreover, the best quality films were produced at pH 11, where production of a supersaturated system hence we would expect
the bath is again supersaturated and the concentration of the free ZnO film formation to occur at lower pH values for HMT than
zinc is orders of magnitude lower than [OH2], similar to the en. HMT decomposes under CBD conditions, which com-
en-system. However in contrast to the en-system, in which the promises any simple thermodynamic speciation calculations.
zinc is mainly in the form of Zn–en complexes at pH 11, for TEA However, under typical experimental conditions, ZnO thin
the zinc speciation is defined by zinc-hydroxy complexes. films will be deposited from supersaturated bath solutions
Comparison of micrographs of thin films from the present containing relatively high concentrations of free zinc and low
study (Fig. 5b) with those obtained for ZnO powders,21 under
similar bath chemistry, leads to the following conclusions.
Firstly, the particle morphologies are virtually identical in both
cases. However, there is a substrate effect, as the widths of
particles comprising films are narrower than those formed in
solution (Fig. 5b: i and iii). The morphology of the films can be
described as spherical aggregates.
For both ‘‘as-deposited’’ and annealed films, the formation
of crystalline zincite was evident from XRD (Fig. 3b: ii).
However, unlike films deposited from en-containing baths, no
changes in the diffraction pattern were observed upon annealing

Fig. 6 a. Speciation diagram (top) of Zn–HMT system at 25 uC.


Dashed line represents ‘‘zinc hydroxide’’ precipitation point. A ~
[Zn21] ~ 0.0083 mol dm23 and B ~ [CH3COO2] ~ 0.016 mol dm23
and C ~ [HMT] ~ 0.025 mol dm23. b. SEM images (below) of
Fig. 5 a. Speciation diagram of Zn–TEA system (above) at 25 uC with ZnO films deposited from HMT-baths (0.1 mol dm23 Zn(NO3)2 and
A ~ [Zn21] ~ 0.018 mol dm23 and B ~ [TEA] ~ 0.072 mol dm23. 0.1 mol dm23 HMT) in stoppered flasks in oven: i. pH ~ 6.8, TO(F)
Dashed line represents zinc hydroxide precipitation point. b. (below) glass, precipitation initiated at 50 uC; ii. pH ~ 6.8, ZnO template
i. Top view and ii. cross-section SEM images of annealed ZnO films layers, precipitation initiated at 50 uC; iii. pH ~ 5.0, TO(F) glass,
from TEA-baths (70 uC, pH 11). [Zn] ~ 0.018 mol dm23; [TEA] ~ precipitation initiated at 95 uC; iv. pH ~ 5.0, ZnO template layers,
0.072 mol dm23. iii. Precipitates obtained from solution. precipitation initiated at 95 uC.

J. Mater. Chem., 2004, 14, 2575–2591 2579


concentrations of hydroxide. In essence this represents a dependent effects. With increasing concentration of a given
reversal of the situation encountered with the en-system. specifically adsorbed ion (SAI), the point of zero charge (PZC)
Several workers have reported the homogeneous precipita- may shift and in addition, the sign of the surface charge as given
tion of ZnO in HMT baths but Andrés-Vergés and co-workers by the zeta (f)-potential may reverse. The first effect occurs
demonstrated that a decrease in solution pH resulted in the because the SAI may interfere with the adsorption of potential-
formation of individual needles or prismatic microcrystals from determining ions (i.e. H1 and OH2 for ZnO in contact with
baths containing zinc nitrate and HMT.20 We have reasoned aqueous solution). The second effect arises from adsorption of
that the presence of a pre-existing template layer (deliberate or the SAI into the inner layer of the solid–solution interface, which
otherwise) would promote heterogeneous growth of ZnO films. in turn can reverse the charge of the outer diffuse double-layer.
Moreover, as precipitation occurred rapidly in baths in which In the present study, the influence of zinc salt counter-ion on
no pH adjustment was made, by using both ZnO templates and crystallite morphology has been investigated. As rod-like ZnO
lower solution pH, good quality films can be deposited. SEM morphologies were of particular interest in the present study,
images were recorded for films deposited on ZnO template initial experiments were conducted using TO(F) glass substrates
layers at an unadjusted pH of 6.8 (Fig. 6b: ii) and the lower pH and baths containing HMT and zinc nitrate, chloride, per-
of 5 (Fig. 6b: iv), using bath conditions described above. It is chlorate, acetate or sulfate. Growth morphology is highly
clearly seen that films deposited at both pH values are dense, sensitive to the level of supersaturation. Speciation calculations
importantly however, those films deposited at lower values of indicate that there are small, but significant, differences in levels
pH have well defined faces. The phenomenon has been of ‘‘zinc hydroxide’’ supersaturation and ZnO precipitation
rationalised in terms of slower growth kinetics.20 points for baths containing different counter-ions. In terms of
The above observations have led to a new two-step approach thermodynamic stability of complexes, the calculations suggest
for the solution growth of highly orientated ZnO microcolumn that precipitation should follow the order sulfate, nitrate y
arrays. The procedure involves deposition of ZnO template chloride y formate and acetate. Empirical observations followed
layers of the desired morphology and subsequent overgrowth the order sulfate, perchlorate, chloride, nitrate, formate and
of ZnO microcolumns on the templates.2 Similar strategies acetate anion. The induction period for baths containing acetate
have been employed subsequently by others.44–46 was significantly longer than for baths containing the other salts.
Acicular morphologies encountered for solution grown ZnO
Baths containing sulfate ion did not produce films on TO(F) glass
are generally assumed to result from one of two mechanisms
in this study. Images obtained by SEM of films deposited using
based on dissolution–recrystallization phenomena or alter-
zinc chloride (Fig. 7b), perchlorate (Fig. 7c) and nitrate (Fig. 7d)
natively nanoparticle-oriented aggregation. Results obtained by
on TO(F) glass indicated that films were not dense and the
Andrés-Vergés et al. are more consistent with the latter process.
acicular crystallites poorly orientated. Although films deposited
However, the processes by which the particles couple at a certain
using zinc acetate were dense the rod-like crystallites were not
size are not clear. The spherical particle size was found to be
well orientated (Fig. 7a). The particle dimensions also varied
critical; these fuse to form embryonic rod-like microcrystals. It
substantially for each counter-ion used.
was postulated that the oriented coupling through the particle
In subsequent experiments, ZnO template layers were used as
c-axis may be related to the net polarisation that occurs in this
direction in ZnO microcrystals. The particles subsequently substrates (prepared using TEA baths) deposited on TO(F)
develop by a ripening mechanism. In this latter step, the particles coated glass. It was observed that good quality films, comprising
increase in length but not in width until they reach the required discrete upright ZnO rods, on template layers, could only be
dimensions. The formation of the well-developed faces is the final formed when substrates were immersed just before or as visible
step that takes place under slower kinetic conditions.20 Recently precipitation occurred. For comparison attempts were also made
the group of Weller have reported similar observations for ZnO to deposit films using zinc formate. Scanning electron micro-
nanoparticle assemblies.35 graphs (ESI:{ Figure C) revealed that films deposited on ZnO
The overall pattern of results indicates the sensitivity of the templates were well orientated. The dimensions of crystallites
morphology of generated ZnO particles to the nature of the varied significantly for each precursor used. In agreement with
complexing ligand. The presence of en in the baths typically the earlier results, no films were obtained using zinc sulfate.
leads to star-like morphologies, those with TEA produce spheri- From the results obtained in the present study, it is clear
cal aggregates and HMT leads to formation of rods/needles. that use of different counter-ions in baths containing growing
crystallites has a noticeable effect on final film morphology.
Influence of counter-ion on growth of ZnO crystallites Differences in kinetics (growth rates, adsorption of counter-
ions on growing crystal faces, etc.) are the likely origin. The
All other factors being equal, changing the counter-ion for zinc
often yields different crystallite morphology. Changes in
morphology may derive from the effects of species that act
as promoters or inhibitors for nucleation and growth processes.
For example, by impeding growth on one or more faces, the
crystal develops only in certain directions and hence the final
morphology is altered. The process is often termed ‘‘crystal
growth inhibition’’. For example, studies of the interaction of
diblock copolymers with ZnO crystal faces have shown that
these substances influence both morphology and particle size
distribution of growing crystallites by adsorbing onto specific
faces and retarding growth perpendicular to these faces.47 An
additional factor associated with the counter-ion of the zinc
salt is volatility of the component or derived species (especially
for use of zinc carboxylates).
Adsorption of ions from solution to substrate may arise simply
as a consequence of simple electrostatic attraction of species on
the charged surface of ZnO (as a consequence of the solution pH Fig. 7 SEM images of ZnO films on TO(F) glass from HMT baths
and amphoteric surface hydroxyl groups). However specific (0.025 mol dm23 zinc salt and 0.025 mol dm23 HMT, pH 5, 90 uC ,
adsorption effects may occur and lead to concentration 1 hour); A. Zn(CH3COO)2; B. ZnCl2; C. Zn(ClO4)2; D. Zn(NO3)2.

2580 J. Mater. Chem., 2004, 14, 2575–2591


effect is more pronounced in precipitates removed from the crystallites comprising films do not exhibit significant difference
baths used to deposit thin films and characterised by SEM. in width. The platelets arising from sulfate baths (no films could
Material obtained from baths containing acetate (Fig. 8a: i), be deposited) are flat and are y30 mm wide, much larger than any
formate (Fig. 8a: ii) and chloride baths (Fig. 8a: iii) were mainly of the particles formed from other systems. The most likely
rod-like whilst crystallites obtained from baths containing explanation for the formation of these platelets is considerable
nitrate (Fig. 8a: iv) and perchlorate (Fig. 8a: v) anions were crystal growth inhibition of the polar c faces by sorption of the
acicular. Those using sulfate were composed of flat hexagonal sulfate anions.48,49 Films obtained from chloride baths were
platelets (Fig. 8a: vi). Twinning was observed in ZnO particles homogeneous and again show the effect of the template layer on
obtained from perchlorate baths. morphology. The powders, however, are composed of both
A comparison of the dimensions of the crystalline pre- agglomerated rods and much larger flakes. Nitrate- and
cipitates from the bath with crystals comprising films (ESI:{ perchlorate-containing baths produced less homogeneous films.
Figure C) gives an indication of the effect of the template layer. The driving force behind the directional and ordered aggre-
The ZnO precipitates obtained from acetate and formate baths gation has been studied by Ocaña et al., who observed similar
differ significantly in their typical dimensions. By contrast, the growth patterns using TEM for the production of ellipsoidal
a-Fe2O3 particles with phosphate anions.50 Matijevic et al. also
studied the effect of anionic surfactants on the size and shape of
ZnO particles. They found that the addition of a fluorinated
polyether carboxylate to a system, which typically produced
intertwined ZnO crystallites, results in the formation of separ-
ated rod-like particles.30
The clear differences in CBD of the ZnO crystallite mor-
phology as a function of counter-ion are interesting, as they
contrast with previous results obtained by our group using
chemically similar systems and hydrothermal conditions.21 In
the first instance, it is clear that the behaviour of different
counter-ions cannot be adequately described as simple adsorp-
tion of different ions on the ZnO surfaces. As described earlier, it
is known that carboxylate species are often selectively adsorbed
on a variety of surface sites on ZnO. Studies in the vapour phase
have shown that formate species adsorb on the ZnO (0001) polar
(Zn terminated) and ZnO (101̄0) non-polar faces while the
pristine (0001̄) surface was unreactive.51 The difference in reac-
tivity was accounted for by the lack of acid–base pairs on the
oxygen polar face. Interaction of formate with the (0001̄) face can
occur at defect sites, i.e. at exposed Zn21 at steps or oxygen
vacancies. Three structures have been proposed for the formate
anion sorbing onto metal oxide surfaces, viz. bridging, bidentate
and unidentate.52 Formate was found to coordinate to the metal
in a bidentate fashion on the (101̄0) face and in a monodentate
manner on the (0001) face or (0001̄) face at defect sites.
TEM images and SAED of crystallites removed from ZnO
films deposited from HMT/acetate-containing baths onto
template layers are shown in Fig. 8b: i and ii. Precipitates
obtained from these baths were similarly characterised (Fig. 8b:
iii and iv). Although the sample preparation resulted in some
contamination of ZnO rods by organic material, it was clear
that many of the rods were composed of discrete subunits,
orientated and connected along the c-axis of the growing
crystallites. The subunits are of the order y1 mm in length, for
both ZnO rods formed in solution and as thin films. The
mechanism of growth of ZnO rods is clearly complex and may
involve a number of processes. More detailed studies were
conducted using TEM and focussed ion beam (FIB) sample
preparation techniques and are discussed later.

Influence of substrate
The effect of the substrate on the nature of the ZnO films formed
was studied. Deposition experiments were performed using
identical bath conditions (0.025 mol dm23 Zn(CH3COO)2 and
0.025 mol dm23 HMT, pH 5, 90 uC, 1 h) but different
substrates: gold-coated TO(F) glass, ZnO template layers on
TO(F) glass, TO(F) glass and single crystalline (0001) sapphire.
Micrographs for the first three are shown in Fig. 9a
(identical magnification). There are marked differences in
average diameter of the crystallites deposited on the three
Fig. 8 a. SEM images (top) of ZnO precipitates obtained from HMT
baths (0.025 mol dm23 zinc salt and 0.025 mol dm23 HMT, pH 5, 90 uC, substrates. By contrast, the lengths of rods (as-determined by
1 hour); i. Zn(CH3COO)2; ii. Zn(HCOO)2; iii. ZnCl2; iv. Zn(NO3)2; cross-sectional SEM) were similar, in the range 2–2.5 mm. The
v. Zn(ClO4)2; vi. ZnSO4. b. TEM images and SAED (below) of ZnO films deposited on both Au-coated TO(F) substrates and the ZnO
precipitates obtained from zinc acetate–HMT baths. template layers were well-aligned. For the microcolumns grown

J. Mater. Chem., 2004, 14, 2575–2591 2581


relationships between (112̄0) ZnO and (011̄2) Al2O3 and (0001)
ZnO and (01̄11) Al2O3.56 Deposition experiments were also
conducted on substrates that were used to produce the films
shown in Fig. 10a, after rinsing in nitric acid to remove deposits
and then using the standard cleaning procedure. Arrays of ZnO
rods (rather than the previously observed rotational twins)
were found to grow in troughs or scratches, formed presumably
through the cleaning process (Fig. 10b).

Influence of solution pH
In the present study, the influence of bath pH on as-deposited
films of ZnO, grown on ZnO template layers, using the zinc
acetate/HMT system was investigated. As for zinc nitrate/HMT
systems, homogeneous precipitation occurred more rapidly in
zinc acetate/HMT systems where no pH adjustment was made as
compared to those where the pH was lowered. The SEM
micrographs (Fig. 11) reveal that films from such baths (pH ~
6.8) comprised crystallites of different sizes with poorly
developed crystal faces. In subsequent experiments, baths were
adjusted to pH ~ 5 using HCl or acetic acid. In these cases, films
were composed of rods with well-defined faces and in addition,
the individual particles had a smaller average diameter than those
obtained from the former. The non-chloride containing baths
also showed the slowest homogeneous precipitation, i.e.
possessed the longest induction time before reaction occurred.

Effect of changing ionic strength


For sparingly soluble metal compounds, such as ZnO or Zn(OH)2
in the present study, increasing the ionic strength tends to increase
the solubility of the solid phase because the activity of the ions
decreases. In order to investigate the effects of ionic strength on
Fig. 9 a. SEM images (top) of ZnO films from HMT baths (0.025 mol dm23 the nature of ZnO films, experiments were conducted using the Zn
Zn(CH3COO)2 and 0.025 mol dm23 HMT, pH 5, 90 uC , 1 hour) on acetate/HMT system on ZnO template layers.
i. Au-coated TO(F) glass; ii. ZnO template layers on TO(F) glass; This work involved the variation of the ionic strength in the
iii. TO(F) glass. b. Crystallite size distribution (below) determined from Zn–HMT system using KCl as a background electrolyte. The
SEM images of ZnO nanocolumns deposited on Au-TO(F) glass. Radii ionic strength of the system was varied in the range [KCl] ~
of ZnO microcolumns reach terminal values within few minutes of
growth, leading to development of a stable size range that could be
0.025–0.4 mol dm23. The precise ionic strength of the bath
fitted to a normal Gaussian distribution (400 data sets, R ~ 0.984). could not be determined as solution pH was adjusted by
dropwise addition of glacial acetic acid. However, useful trends
directly on Au/TO(F) glass there are no signals arising from could be identified. It was observed that as [KCl] increased,
polycrystalline ZnO template layers (ESI:{ Figure D). Crystallite
size distributions were determined from micrographs of ZnO
nanocolumns deposited on Au-TO(F) glass (Fig. 9b) in order to
quantify the polydispersity and homogeneity of films. Using an
image analysis routine,53 average (mean) radii values of 266 nm
(standard deviation 54 nm) were determined.
Growth of ZnO films on single crystal (0001) sapphire from
HMT-containing baths was attempted. Preliminary experi-
ments using similar material from co-workers appeared to
indicate that unusual 60u rotational twinned structures could
be deposited (Fig. 10a). Previously a great deal of research
has focussed on ZnO films grown on (0001) sapphire
substrates.54,55 Other workers have demonstrated epitaxial

Fig. 10 SEM images of ZnO films deposited on single crystal sapphire


(0001) from baths containing zinc acetate (0.025 mol dm23) and HMT Fig. 11 SEM images of ZnO films from HMT baths (0.025 mol dm23
(0.025 mol dm23) at 90 uC. a. Starlike crystallites obtained, some 60u Zn(CH3COO)2 and 0.025 mol dm23 HMT) deposited on ZnO template
rotational twins are observed; b. rod arrays grown in troughs on layers at A. pH 6.8; B. pH 5 (adjusted with HCl); C. pH 5 (adjusted with
cleaned substrate previously used to deposit films shown in a. CH3COOH).

2582 J. Mater. Chem., 2004, 14, 2575–2591


homogeneous precipitation occurred more rapidly. It was Effect of varying growth times
apparent that no ZnO rods were deposited from baths with
Andrés-Vergés et al. found that for ZnO grown under hydro-
[KCl] w 0.2 mol dm23. The effect (if any) of zinc-chloro
thermal conditions, an increase in bath temperature for
complex formation, which would tend to increase stability to
solutions containing ZnCl2 and HMT resulted in a change in
precipitation with increasing [KCl], appears to be insignificant. morphology from rods to needle morphology.20 In these
As shown in Fig. 12, SEM images reveal that within the studies, the length and final shape of the rod were shown to be
range [KCl] ~ 0.025–0.15 mol dm23 the width of individual very dependent on the growth time. It is also reported that
crystallites was decreased from those formed from baths ageing time also has a major influence on final crystal
containing no KCl. The observation could be rationalised in morphology, ZnO habit changing from flaky aggregates to
terms of specific ion adsorption e.g. on kink sites on the a-faces flower-like morphology with increasing growth time.58
of the ZnO crystal, which could indirectly promote growth in The development of ZnO rod array films (using zinc acetate)
the c direction. However, no measurements were done to on ZnO template layers as a function of deposition time was
determine if particles also decreased in length. Ionic strength investigated. Films were removed from baths after different
effects have been demonstrated for solution growth of calcite periods of time and characterised by SEM. The micrographs
crystals, at low ionic strength, aggregation and alignment (Fig. 13) show that crystallites possess an average diameter of
phenomena are present, no such mechanisms operate at higher y300 nm after 2 minutes and y1 mm after 1 h. The cross-
ionic strength.57 For the range [KCl] w 0.15 mol dm23 parti- sectional SEM images allow the length of the rods to be
cles changed morphology from rods to discs. Similar results determined. However it was only possible to determine the
were obtained using zinc formate rather than acetate as a average rod length for those grown for times greater than
precursor (see ESI:{ Figure E). 10 min. At shorter growth times, it was difficult to distinguish
The influence of ionic strength on ZnO film growth is clearly the template layer from the ZnO rod growth (ESI:{ Table B).
complex. From CBD experiments employing zinc acetate or Evidence for oriented attachment processes was obtained for
formate and HMT, it appears that there is a limiting ionic strength homogeneous precipitates lying on films (Fig. 13D).
whereby film growth is possible, beyond which only homogeneous
precipitates may be formed. Changing ionic strength also has
morphological effects, as evidenced from decreasing ZnO rod Summary of preceding survey of factors that influence CBD of ZnO
crystallite width with increasing ionic strength.
In general, the most likely crystal morphology for ZnO under
kinetically controlled growth conditions and at low super-
saturation is as acicular rods, whereas under thermodynamic
control, prismatic structures are expected.17 Hence by limiting
the supply of one reactant employing a large excess of the
other, anisotropic growth is facilitated along the polar c-axis of
the zincite structure. Several factors exert a profound effect on
the morphology of ZnO thin films grown by CBD and there are
important differences between CBD and the chemically similar
hydrothermal routes to ZnO powders or films. Under hydro-
thermal conditions, Trindade et al. have reported that neither

Fig. 12 SEM images showing the effect of ionic strength on ZnO films
(on template layers) obtained from HMT baths (0.025 mol dm23 Fig. 13 SEM images of ZnO films from HMT baths (0.025 mol dm23
Zn(CH3COO)2 and 0.025 mol dm23 HMT, pH 5, 1 h). The ionic strength Zn(CH3COO)2 and 0.025 mol dm23 HMT) deposited on ZnO template
was varied using different concentrations of KCl: A. 0 mol dm23; layers deposited for: A. 2 min; B. 5 min; C. 10 min; D. 20 min; E. 30 min;
B. 0.025 mol dm23; C. 0.05 mol dm23; D. 0.075 mol dm23; E. 0.1 mol dm23; F. 1 h. For D, a large dumbbell precipitate formed homogeneously in the
F. 0.125 mol dm23; G. 0.3 mol dm23; H. 0.4 mol dm23. bath appears to indicate formation by orientated attachment.

J. Mater. Chem., 2004, 14, 2575–2591 2583


solution pH, choice of counter ion and ionic strength, bath such films are y90 nm wide and y600 nm long. The aspect ratio
temperature or reaction time exert any significant influence on of the latter (y6.7), grown from time TB, is clearly better than the
the characteristics of resultant ZnO powders. The choice of former (y4.8) grown from time TA.
ligand and nature of hydrothermal treatment were found to be
important,21 results that are clearly different from those
Use of nanocrystalline template layers for overgrowth of ZnO
obtained for CBD films.
films by forced hydrolysis of zinc carboxylates and using different
The results obtained in the present study suggest that nuclea-
chelating ligands
tion and growth promotion/inhibition are significant and vary
considerably with small changes in bath chemistry. For example, The pattern of results in the preceding sections indicated that
only large homogeneous precipitates can be obtained from the simplest and most promising synthetic route towards
deposition baths in the presence of sulfate, which appears to growth of ZnO nanorod array films would involve deposition
completely impede nucleation and growth of ZnO thin films, on on to nanocrystalline ZnO template layers. Preliminary
both TO(F) glass and ZnO template layers. Moreover, the experiments involved the forced hydrolysis of zinc acetate
morphology of these precipitates (flat hexagonal y30 mm wide and zinc formate using nanocrystalline ZnO templates on glass
platelets) is very different to those obtained using other zinc salts substrates. It was very apparent that very low levels of
(i.e. acetate, formate, nitrate, chloride and perchlorate). Flattened homogeneous precipitation occurred using baths containing
crystallites were also observed in experiments involving increasing zinc acetate, moreover no visible homogeneous precipitation
ionic strength (using KCl) and HMT/zinc acetate systems. could be detected using zinc formate solutions. Transparent
In the aqueous CBD of ZnO, the influence of counter-ion on films were obtained with both precursors. Subsequent experi-
crystallite morphology may be through weak electrostatic ments involved the use of these template layers for the
forces (e.g. specific adsorption of ions resulting in change in overgrowth of ZnO films deposited using different chelating
p.z.c. or f-potential, double layer compression, etc.) or more ligands viz. en, TEA and HMT. Reactions were similar to those
significant physi- or chemisorption behaviour (e.g. at attractive reported earlier for each ligand on TO(F) glass substrates. The
kink sites on crystal faces). However, increasing ionic strength transparent films were characterised by XRD and SEM.
also tends to increase the solubility of the solid phase being X-Ray diffractograms obtained for nanocrystalline ZnO tem-
deposited but decrease the values of thermodynamic stability plates and overlying films are shown in Fig. 15. For films pro-
constants of the metal–ligand complexes in deposition baths. duced by forced hydrolysis of zinc acetate, using en and HMT,
These facts highlight the complexity of CBD systems and there is strong enhancement of the (002) peak. For films deposited
illustrate the difficulty in unambiguously assigning direct from baths containing TEA ligand the effect is not as pronounced.
causal relationships between the many interrelated parameters As indicated earlier, attempts to deposit ZnO films by forced
(e.g. crystal morphology and choice of counter-ion). hydrolysis of zinc carboxylates with no added ligand were
unsuccessful using TO(F) glass substrates. However, SEM
Dependence of film morphology on point of introduction of images provide evidence that films are produced using identical
substrate in deposition bath baths (0.025 mol dm23 zinc acetate or zinc formate) on
nanocrystalline ZnO templates (Fig. 16A,B and C,D respec-
Empirically it has been found that the size of the crystallites tively). The films were homogeneous and composed of rod-like
comprising films depends largely on when the ZnO template array structures (ca. 140 nm wide and 650 nm long) when zinc
substrate is immersed in the bath. If the substrate is immersed in acetate was used (Fig. 16B), in contrast to the more nodular
the bath from the point at which the pH of the solution is adjusted particles (ca. 140 nm wide and 300 nm long) obtained when zinc
to 5 at ambient temperature, the nanocrystalline template formate was used (Fig. 16D).
dissolves as the bath is heated to 90 uC and no film is deposited. Films obtained from baths using en as a ligand for zinc were
If the substrate coated with a nanocrystalline ZnO template is composed of fused columnar crystals (ca. 110 nm wide and
inserted in the bath about 15 min after the desired temperature of 300 nm long), with well-defined hexagonal end facets (c-faces).
90 uC is reached (TA), the bath produces no homogeneous Hence the presence of a template layer appears to direct growth
precipitation during the 2 h reaction period and a dense white film towards acicular, rather than starlike, morphologies. The
is obtained. Images obtained by SEM reveal that the particles are crystallite formed homogeneously in the bath and laying on top
y 250 nm wide and y1.2 mm long (Fig. 14C,D). However, if the of the film possesses a star-like morphology (Fig. 16E, inset)
substrate is placed in the heated deposition bath at the point of and resembles those comprising films deposited over TO(F)
onset of visible turbidity (ca. 30 min after reaching 90 uC) (TB), a glass (Fig. 4a). Films deposited using the TEA ligand have a
transparent film is obtained and some homogeneous precipita-
tion is present in the bath (Fig. 14A,B). The ZnO rods comprising

Fig. 15 XRD patterns of ZnO films from: a. HMT-system ([ZnAc] ~


0.025 mol dm23, [HMT] ~ 0.025 mol dm23, pH 5, 90 uC, 2 h); b. en-
system ([ZnAc] ~ 0.018 mol dm23, [en] ~ 0.042 mol dm23, pH 11,
70 uC, 2 h); c. forced hydrolysis of zinc acetate (0.025 mol dm23, 90 uC, 2 h);
d. TEA-system ([Zn(NO3)2] ~ 0.018 mol dm23, [TEA] ~ 0.042 mol dm23,
Fig. 14 SEM images of ZnO films deposited on nanocrystalline ZnO pH 11, 70 uC, 2 h) deposited on nanocrystalline ZnO templates.
templates introduced in baths from time: A,B TB and C,D TA. Nanocrystalline ZnO template shown in e. ZnAc ~ zinc acetate.

2584 J. Mater. Chem., 2004, 14, 2575–2591


Fig. 16 SEM images of ZnO films deposited on nanocrystalline ZnO templates from: A,B. forced hydrolysis of zinc acetate; C,D. forced hydrolysis
of zinc formate; E,F. en-system; G,H. TEA-system.

more globular appearance and are less defined than others in obtained from films (Fig. 18) suggest that films grown from
these investigations. baths containing zinc nitrate and HMT possess crystallites with
Based on the success of deposition baths containing HMT the largest aspect ratio (ca. 10).
and using microcrystalline ZnO template layers, further work
was conducted on the influence of the nanocrystalline template
layer on the subsequent growth of films, using HMT as the Growth of ZnO rods on nanocrystalline ZnO templates from
ligand. The results were unambiguous, micrographs obtained baths containing the products of decomposition of HMT
from films provide clear evidence that all other factors being (formaldehyde and ammonia)
equal, the width of ZnO rods grown on nanocrystalline The amine HMT decomposes slowly in heated aqueous solu-
(Fig. 17D, E and F) rather than microcrystalline (Fig. 17A, B tions to yield ammonia and formaldehyde as initial reaction
and C) template layers is far smaller. products. Attempts were made to grow ZnO rods on nano-
crystalline ZnO templates directly from aqueous baths (90 uC)
containing zinc acetate, ammonia and formaldehyde. The
Influence of choice of counter-ion on growth of ZnO rods on
concentrations of the latter components were chosen assuming
nanocrystalline ZnO template layers
stoichiometric decomposition of HMT. Reactions were carried
The influence of the zinc counter-ion on the ZnO rod dimen- out at the initial solution pH of 6.9 and at pH 5, which was the
sions was investigated using similar strategies to those pH employed earlier in this study for the successful CBD of
employed for microcrystalline template layers. Micrographs ordered arrays of ZnO rods. No precipitation occurred from

J. Mater. Chem., 2004, 14, 2575–2591 2585


Fig. 17 A. CBD-ZnO microcrystalline templates on TO(F) glass comprising nodular grains (deposited from TEA-systems); B,C. ZnO microcolumns
grown on CBD-ZnO templates from HMT-system ([Zn] ~ 0.025 mol dm23, [HMT] ~ 0.025 mol dm23, pH 5, 90 uC); D. nanocrystalline ZnO templates
on glass microscope slides (prepared by sol gel (SG) method); E,F. ZnO nanocolumns grown on SG-ZnO templates from HMT-system.

baths held at pH 5 in which dissolution of the nanocrystalline are small differences in levels of ‘‘zinc hydroxide’’ super-
ZnO template layer occurred. Experiments were also conducted saturation and precipitation points (Fig. 19b). The hydroxide
without formaldehyde i.e. only zinc acetate and ammonia pre- precipitation points for the baths containing Zn:HMT in the
sent. Speciation calculations were performed in order to assess ratios, 2:1, 1:1, 1:2, 1:4, 1:8 occur at pH values of 6.8, 6.8, 6.9.
the initial degree of supersaturation and identify the important 7.04 and 7.16 respectively. The calculated pH, at which the bath
components in deposition baths (Fig. 19a). Baths containing is supersaturated with respect to the hydroxide containing
either ammonia or ammonia and formaldehyde were found be Zn:HMT in a 1:1 ratio, is exactly the same as the bath con-
supersaturated with respect to zinc hydroxide at pH 6.8. These taining the stoichiometric amount of ammonia and formalde-
results are in general agreement with experimental observations. hyde (or ammonia alone). However, these small differences in
Similar thermodynamic calculations were performed for baths supersaturation at a given pH do appear to be associated with
containing different concentrations of HMT. Speciation calcula- morphological changes in films, as SEM images for the films
tions indicate that with an increase in HMT concentrations, there deposited using different HMT concentrations show that as the
HMT concentration increases, the width of the rods decreases.
The aspect ratio for the rods, however, does not change

Fig. 19 a. Speciation diagram of Zn–NH3–HCHO system at 25 uC with


A ~ [Zn21] ~ 0.0083 mol dm23, B ~ [CH3COO2] ~ 0.016 mol dm23,
Fig. 18 SEM images of ZnO films on nanocrystalline ZnO templates C ~ [NH3] ~ 0.0083 mol dm23 and [HCHO] ~ 0.016 mol dm23. Dashed
from HMT baths (0.025 mol dm23 zinc salt and 0.025 mol dm23 HMT, line represents zinc hydroxide precipitation point. b. Speciation diagram
pH 5, 90 uC, 2 hours); A,B. Zn(CH3COO)2; C,D. Zn(HCOO)2; of Zn–HMT system at 25 uC showing [Zn21]free and zinc hydroxide
E,F. ZnCl2; G,H Zn(NO3)2. precipitation points for increasing [HMT].

2586 J. Mater. Chem., 2004, 14, 2575–2591


significantly i.e. narrower rods tend to reach smaller terminal
lengths.
For films deposited from aqueous baths (90 uC) at initial
pH ~ 6.9, containing zinc acetate, ammonia and formaldehyde
on nanocrystalline ZnO template layers, XRD and SEM
measurements were recorded. X-Ray diffraction patterns were
very similar to those obtained for ZnO rod films deposited from
baths containing HMT and zinc acetate on ZnO template layers
and exhibited enhancement of the (002) reflection. In agreement
with previous results, no evidence for incorporation of hydroxide
was obtained from ATR-FTIR analysis. SEM micrographs were
also very similar to those obtained for analogous HMT systems,
as films comprised ordered rod arrays of perpendicularly
orientated ZnO rods on nanocrystalline template layers
Fig. 21 XRD patterns of ZnO films deposited on ZnO nanocrystalline
(Fig. 20A,B). The dimensions of the rods were y85 nm in template layers from HMT baths (0.025 mol dm23 Zn(CH3COO)2
width and y600 nm long, which were smaller than those and 0.025 mol dm23 HMT, pH 5, 90 uC) for different periods of time.
obtained for films deposited from HMT/zinc acetate systems and a. 2 min. b. 5 min. c. 10 min. d. 20 min. e. 30 min. f. 1 h. g. 2 h.
grown on nanocrystalline ZnO templates (Fig. 18A). For baths
containing only zinc acetate and ammonia, the films comprised In attempts to improve the aspect ratio of ZnO rods, as-
rods with globular caps (Fig. 20D) in contrast to the rods formed deposited ZnO films were immersed into fresh baths. An
when formaldehyde was present (Fig. 20A,B). additional motivation for this work was to gain further insights
into the actual growth mechanism of ZnO single crystals. In
Effect of growth time preliminary experiments, a direct comparison was made of films
prepared using baths (HMT–zinc acetate system) identical to
The XRD patterns for the films, deposited for different times,
those employed previously but using three different reaction
are shown in Fig. 21. All patterns are consistent with zincite
phase. As the growth time increases, the (002) reflection sequences. In the first group, ZnO rods were deposited on
increases in relative intensity, consistent with formation of ZnO nanocrystalline ZnO templates for 2 h (Method 1). In the second
rod crystallites. The (100) and (101) reflections remain broad group, ZnO rods were deposited on nanocrystalline ZnO for 2 h
during deposition in contrast to the (002) peak, which becomes and subsequently transferred into fresh heated baths, at the point
noticeably sharper with increasing deposition time. This where visible turbidity became apparent, for a period of 1 h
observation is consistent with the growth of ZnO rods (Method 2). The final experiments followed the second procedure
crystallites, along the c-axis. and then films were removed and placed in another fresh bath at
Scanning electron micrographs were recorded and grain size the critical point for an additional 1 h (Method 3).
distributions determined for films deposited for different growth Films were characterised by SEM and it was apparent that
periods (ESI:{ Figure F). Grain size profiles were unimodal, although the length of the ZnO rods increases upon exposure to
generally asymmetric and typical for abnormal grain growth fresh CBD solution, the width of rods also increases and hence
where grain size distributions are directed by different substrate– no improvement in aspect ratio is observed (Table 1 and Fig. 22).
grain relationships. The average grain diameter increased linearly In addition, crystallites comprising the films become very densely
from y82 nm after 5 min to y136 nm after 2 h, corresponding to packed, which is not desirable for applications that require post-
an average growth rate of y0.23 nm min21. From cross sectional deposition processing such as infiltration of dye molecules onto
SEM images it was clear that growing ZnO rods tend to broaden the rods. No conclusive evidence for film growth via orientated
slightly with deposition time, hence the definition of ‘‘average attachment processes could be gained. The resolution of the
grain size’’ is not precise and a direct comparison is only sensible SEM is insufficient for detailed interpretation of growth of
for rods of similar length. nanostructured films. However, it was clear that in common with
the microcrystalline systems described earlier (see Fig. 13D),
growth of homogeneous precipitates may occur through
orientated attachment processes (Fig. 22G).
Attempts were made to gain more detailed structural informa-
tion, by use of HR-TEM and focussed ion beam (FIB) sample
preparation techniques. Samples were prepared from ZnO thin
films deposited by Method 1 (2 h deposition) and Method 2 (2 1
1 h deposition) as described previously (ESI:{ Figure G). The FIB
technique allows thin (y50 nm) specimens for HR-TEM to be

Table 1 Rod dimensions and aspect ratio of the films deposited for
the various times

time diameter/nm length/nm aspect ratio


a
2 min y77 ¡ 4 —
a
5 min y83 ¡ 6 —
a
10 min y85 ¡ 10 —
20 min y102 ¡ 9 220 ¡ 10 2.1
30 min y108 ¡ 11 530 ¡ 25 4.9
1h y122 ¡ 10 1000 ¡ 50 8.2
2h y137 ¡ 11 1500 ¡ 75 10.9
Fig. 20 SEM images of ZnO rod arrays deposited on nanocrystalline 2h11h y320 ¡ 11 2000 ¡ 100 6.3
ZnO templates layers from baths containing: A,B. zinc acetate 2h11h11h y560 ¡ 12 2500 ¡ 125 4.5
(0.025 mol dm23), HCHO (0.016 mol dm23) and NH3 (0.0083 mol dm23) a
The presence of the template layer compromised accurate length
(pH 6.9, 90 uC, 2 h) and C,D. zinc acetate (0.025 mol dm23) and NH3 measurements.
(0.0083 mol dm23) (pH 6.9, 90 uC, 2 h).

J. Mater. Chem., 2004, 14, 2575–2591 2587


Fig. 23 HR-TEM of cross-section of ZnO thin film sample prepared by
Method 1 (mag. 627.5k) and SAED recorded for single crystalline ZnO
rod along the (100) zone axis. For reference the first series of reflections
are indexed. Double diffraction phenomena occurs leading to the
presence of forbidden reflections (e.g. (001)) and are denoted by X.

conjunction with an active interface. Under typical dynamic


bath conditions, it is more likely that these external aggregates
of nanoparticles will be subject to dissolution–reprecipitation
processes. The predominance of one process over the other will
Fig. 22 SEM images of ZnO rod arrays deposited on nanocrystalline determine the velocity of growth in a given direction. Sorption
ZnO template layers for: A,B. 2 h (Method 1); C,D. 2 h 1 1 h of ions or anisotropic nanoparticles with favourable orienta-
(Method 2) and E,F,G. 2 h 1 1 h 1 1 h (Method 3). tion approaching the polar faces of wurtzite ZnO (in contrast
to the non-polar faces), is expected to be effective and prohibit
obtained with minimal damage in preparation. The integrity of desorption and rearrangement of growth units.
the original structure is maintained and each interface is clearly Evidence for orientated attachment of large ZnO crystallites
represented (Fig. 23). Moreover, the single crystalline nature of to existing ZnO rods grown on templates has been obtained,
individual ZnO rods is clear from SAED. for both films grown by Method 1 (2 h deposition on ZnO
The crystalline rods of ZnO formed comprise oriented nano- templates) and Method 2 (2 h deposition followed by 1 h
crystallites, in which the lattice planes of individual particles are growth in fresh baths). The pattern of results obtained from the
perfectly aligned to form a single crystalline entity (Fig. 24). It is TEM studies appears to show that the process is not as
also evident from TEM that the outer layers of ZnO rods (ca. significant as the mechanism involving ZnO nanocrystal
y5 nm) are composed of randomly orientated nanoparticles. aggregation. An example of the phenomena for films grown
The atomic structure is clearly resolved in Fig. 25. Nucleation by Method 1 is shown in Fig. 26, where the rod–rod interface of
will involve conditions of low to moderate supersaturation in a composite unit is shown. Although visually there appears to

Fig. 24 HR-TEM of single ZnO rod (mag. 310k) grown on nanocrystalline ZnO template layer (Method 1). a. The outer porous layer is composed
of randomly orientated nanoparticles. b. Area indicated in a is shown in greater detail. The extended lattice structure, comprising individual ZnO
nanocrystals, is clearly evident.

2588 J. Mater. Chem., 2004, 14, 2575–2591


in ZnO rods could be obtained, from HR-TEM or SAED. The
pattern of results obtained in this study for the deposition of
ZnO rods on nanocrystalline ZnO template layers do not
support a mechanism based exclusively on a growth spiral.
However, our observations do not preclude the possibility that
such a mechanism may operate in early stages of growth.
A likely sequence of events would be 2D-nucleation on the
template layers, via the orientated aggregation of polar nanopar-
ticles and the subsequent development of single crystalline material
by dissolution-reprecipitation phenomena. As supersaturation in
baths decreases due to consumption of reagents, 2D-nucleation
and growth processes on surface terraces decrease rapidly in
significance. Crystal faces develop a smoother appearance through
aging via dissolution-recrystallisation processes, whereas adsorp-
tion and incorporation of growth units is limited by diminishing
concentrations of growth sites such as kinks on crystal faces.

3. Conclusions
The influence of reaction conditions including ligand, counter-
ion, pH, ionic strength, supersaturation, deposition time and
Fig. 25 HR FEG-TEM image of ZnO rod showing lattice structure.
substrate on the nature of ZnO films grown by CBD has been
examined and discussed, the most important of which are
be considerable misalignment of the rods, at higher magnifica- summarised in Fig. 27.
tion (Fig. 26B) the lattice configuration between the two Use of different ligands leads to deposition of ZnO films with
separate rods is evidently good. Similar structures were found different morphologies on TO(F) glass, e.g. ethylenediamine
for ZnO rod arrays grown by Method 2. Hence it appears that usually gave rise to starlike crystals, TEA produced nodules
under the appropriate conditions, ZnO rods can develop by a and HMT usually produced rods. Similarly, using identical
combination of different growth processes. ligands but different zinc salts often gives rise to different
The primary aim of the HR-TEM investigation was to better morphologies, with sulfate usually producing the most drama-
understand the growth mechanism of ZnO rods on nano- tic change. Speciation studies and empirical evidence suggest
crystalline templates. If we assume that deposition occurs from that growth of acicular ZnO morphologies by CBD are best
solutions of low to intermediate levels of supersaturation, obtained by limiting the concentration of one reactant (i.e.
growth would be expected to proceed via dislocation-based either Zn21 or OH2) in the presence of an excess of the second
growth spiral or 2D-nucleation based mechanisms. For all component, promoting the kinetically controlled form of the
samples in the present study (both in preliminary work and the final crystal. The presence of an existing ZnO substrate can
later samples) no evidence for the presence of significant influence the morphology of subsequent ZnO growth. A
concentrations of defects such as dislocations or stacking faults striking example is that ZnO films could not be deposited on

Fig. 26 (Left) HR-TEM of ZnO rods grown by Method 1 showing orientated attachment of two smaller rods. (Right) Inset is shown at higher
magnification to demonstrate the good lattice alignment of the two ZnO single crystals.

J. Mater. Chem., 2004, 14, 2575–2591 2589


Fig. 27 Schematic overview of main results obtained in this study.

TO(F) glass by the forced hydrolysis of zinc carboxylates, but methods. Substrates (microscope slides, TO(F) glass, micro-
rod-like arrays of ZnO were produced from chemically or nanocrystalline ZnO templates, Au-coated TO(F) glass,
equivalent baths on nanocrystalline ZnO templates. Templates single crystal sapphire (0001)) were cleaned by a standard
of ZnO can direct the deposited film e.g. in the en system, which procedure. A Mettler Toledo MA 235 pH/ion analyser and
typically produced starlike crystallites, columnar structures InLab 413 electrode were used to record solution pH.
were deposited. High surface area arrays of upright ZnO Modelling of solutions was performed using SPECIES
crystallites on nanocrystalline ZnO templates have been (AcadSoft Ltd.). The program calculates and displays specia-
deposited using baths containing HMT. The point at which tion curves as a function of pH. The input parameters for
the substrate is introduced into the bath is crucial and leads to a SPECIES are the stability constants for the homogeneous
significant difference in the width of the rods and optical equilibria being modelled and their stoichiometric coefficients.
transparency of the films. Thinner rods are produced when the Film thickness measurements were performed using the QCM
substrate is immersed in the bath at the point of visible technique, using a Maxtek PM-700 Series Plating Monitor and
turbidity i.e. at the onset of precipitation. Zinc nitrate pre- probe employing a quartz crystal oscillator (5 MHz, AT-cut,
cursors tended to lead to formation of rod arrays on gold crystal, unpolished). The QCM apparatus was controlled
nanocrystalline templates that possessed the highest aspect and experimental data recorded by use of a PC.
ratios. It was also found that deposition baths containing zinc Forced hydrolysis was performed using solutions of zinc
acetate and the decomposition products of HMT (initial pH ~ acetate (0.025 mol dm23) or zinc formate (0.025 mol dm23) on
6.9) produced similar rods arrays of ZnO to the standard HMT TO(F) glass. Final pH values for the zinc acetate and formate
system (initial pH ~ 5). It was found that increasing HMT baths were 6.9 and 6.4 respectively. Solutions were heated at
leads to slightly thinner rods, however, there was no change in 95 uC for a period of two hours. For experiments involving
the aspect ratio as crystallites were shorter in length. nanocrystalline ZnO template layers on glass microscope slides,
Attempts have been made to improve the aspect ratio of ZnO
substrates were immersed in solutions at ambient temperature
rods, by growing ZnO rods using fresh solutions on preformed
prior to heating of baths using the temperature-controlled
ZnO rods arrays. The procedure was unsuccessful. The final
waterbath.
ZnO films comprised dense rod arrays that were increased in
Deposition of ZnO thin films by CBD methods using ethyl-
length and in width with some fused columnar structures.
enediamine (en) as ligand was achieved using aqueous solutions
HR-TEM has yielded important structural information and
of zinc acetate (0.018 mol dm23) and en (0.042 mol dm23),
an insight into the growth of ZnO rods by CBD. The most
adjusted to a pH of 11 with aqueous sodium hydroxide
likely sequence of events involves 2D-nucleation on ZnO
(5 mol dm23). Films were grown at the desired temperature
template layers, growth via orientated aggregation of polar
(50–80 uC) for 1 hour. For TEA systems, solutions contain-
nanoparticles and subsequent development of single crystalline
material by dissolution–reprecipitation. ing zinc nitrate (0.018 mol dm23) and triethanolamine
(0.072 mol dm23), adjusted to a pH of 11 with aqueous
sodium hydroxide (5 mol dm23), were used. Films were grown
Experimental over a range of temperatures (50–80 uC) for 1 hour. Unless
otherwise stated, experiments employing HMT were conducted
Film growth
using bath solutions containing zinc acetate (0.025 mol dm23)
Films of ZnO were grown on substrates immersed in aqueous and HMT (0.025 mol dm23), adjusted to pH 5 with aqueous
deposition solutions using forced hydrolysis and CBD acetic acid (5 mol dm23), for deposition times of 1 h.

2590 J. Mater. Chem., 2004, 14, 2575–2591


Microcrystalline template layers were deposited using the 11 D. Scarano, G. Spoto, S. Bordiga, A. Zecchina and C. Lamberti,
TEA method described above. Nanocrystalline templates were Surf. Sci., 1992, 276, 281.
12 M. Izaki and T. Ohmi, J. Electrochem. Soc., 1997, 144, L3.
grown by sol–gel methods,59 involving hydrolysis and con- 13 M. Izaki and J. Katayama, J. Electrochem. Soc., 2000, 147, 210.
densation of an alcoholic Zn-precursor under basic conditions, 14 K. Ito and K. Nakamura, Thin Solid Films, 1996, 286, 35.
in order to produce a homogeneous sol for subsequent dip- 15 L. Vayssieres, K. Keis, S. T. Lindquist and A. Hagfeldt, J. Phys.
coating on glass microscope slides. A suspension of zinc acetate Chem., 2001, 105, 3350.
dihydrate (17.8 g, 0.08 mol) in n-propanol (120 cm3) was heated 16 N. Saito, H. Haneda, W. S. Seo and K. Koumoto, Langmuir, 2001,
under reflux conditions (130 uC, 20 min). The mixture was 17, 1461.
17 A. P. A. Oliveira, J. F. Hochepied, F. Grillon and M. H. Berger,
allowed to cool to ambient temperature before rapid addition Chem. Mater., 2003, 15, 3207.
of tetramethylammonium hydroxide (TMAH; 25% in MeOH, 18 M. L. Fuller, J. Appl. Phys., 1944, 15, 164.
36 cm3) to yield a transparent nanoparticulate ZnO coating sol 19 J. M. Cowley, A. L. G. Rees and J. A. Spink, Proc. Phys. Soc,
(y0.5 mol dm23), which could be concentrated further by 1951, 64, 639.
removal of solvent using a rotary evaporator. Films were 20 M. Andrés-Vergés, A. Mifsud and C. J. Serna, J. Chem. Soc.,
prepared by dip-coating substrates in the sol and sintering in a Faraday Trans., 1990, 86, 959.
21 T. Trindade, J. D. Pedrosa de Jesus and P. O’Brien, J. Mater.
furnace at 400 uC for 2 min.
Chem., 1994, 4, 1611.
22 E. Matijevic, Faraday Discuss., 1991, 92, 229.
Characterisation studies 23 E. Matijevic, Chem. Mater., 1993, 5, 412.
24 F. A. Sigoli, M. R. Davolos and M. Jafelicci, Jr., J. Alloys Compd.,
X-Ray diffraction studies were performed using secondary 1997, 262–263, 292.
graphite monochromated Cu Ka radiation (40 kV) on either a 25 B. G. Wang, E. W. Shi and W. Z. Zhong, Cryst. Res. Technol.,
Philips X’Pert Materials Diffractometer (APD) or Bruker AXS 1998, 33, 937.
D8 diffractometer. Measurements on the former were taken 26 B. O’Regan and M. Grätzel, Nature, 1991, 353, 737.
27 C. Rost, I. Sieber, K. Ernst, S. Siebentritt, M. C. Lux-Steiner and
using a glancing angle incidence detector at an angle of 3u, for R. Könenkamp, Appl. Phys. Lett., 1999, 75, 692.
2h values over 10–95u in steps of 0.04u with a count time of 2 s. 28 F. F. Lange, Science, 1996, 273, 903.
For the latter, scans were done over 2h values of 5u–90u and a 29 S. Yamabi and H. Imai, J. Mater. Chem., 2002, 12, 3773.
step size 0.01u or 0.02u. Scanning electron microscopy (SEM) 30 A. Chittofrati and E. Matijevic, Colloids Surf., 1990, 48, 65.
and energy dispersive X-ray analysis (EDAX) on carbon- 31 R. A. McBride, J. M. Kelly and D. E. McCormack, J. Mater.
coated films was performed using a Philips Excel 30 FEG SEM Chem., 2003, 13, 1196.
32 F. Kazumi and T. Daigaku, Jpn. Pat. JP 57205320, 1982.
instrument or Philips 525 SEM instrument with an EDAX 33 K. Fujita, K. Murata, T. Nakazawa and I. Kayama, Yogyo
DX4 EDS unit. Samples were carbon-coated (using an Kyokaishi, 1984, 92, 227.
Edwards Coating System E306A) for EDAX or otherwise 34 K. Fugita and K. Matsuda, Bull. Chem. Soc. Jpn., 1992, 65, 2270.
gold coated (Edwards Sputter Coater S150B). The focused ion 35 C. Pacholski, A. Kornowski and H. Weller, Angew. Chem., Int.
beam (FIB) thinning technique was used to prepare cross- Ed., 2002, 41, 1188.
sectional ZnO thin film samples for characterisation by high 36 K. Govender, D. S. Boyle, P. O’Brien, D. Binks, D. West and
D. Coleman, Adv. Mater., 2002, 14, 1221.
resolution transmission electron microscope (HR-TEM). This 37 S. Yamabi and H. Imai, J. Mater. Chem., 2002, 12, 3773.
work was conducted at FEI Bristol using a FIB 200 instrument 38 J. G. Strom, Jr. and H. W. Jun, J. Pharm. Sci., 1980, 69, 1261.
(Ga1 beam). In order to minimise ion beam induced artefacts 39 R. Lindsay, E. Michelangeli, B. G. Daniels, T. V. Ashworth,
(i.e. amorphous layer and Ga implantation), a Pt buffer layer A. J. Limb, G. Thornton, A. Gutiérrez-Sosa, A. Baraldi,
was deposited on the area of interest on the specimen by ion- R. Larciprete and S. Lizzit, J. Am. Chem. Soc., 2002, 124, 7117.
beam assisted CVD, prior to thinning. TEM electron micro- 40 H. Gerischer and N. Sorg, Electrochim. Acta, 1992, 37, 827.
41 E. A. Meulenkamp, J. Phys. Chem., 1998, 102, 5566.
scopy was accomplished using either a Philips CM200 (200 kV) 42 W. K. Burton, N. Cabrera and F. C. Frank, Philos. Trans. R. Soc.
microscope or a FEI Tecnai F30 FEGTEM (300 kV) instru- London, 1951, A243, 299.
ment. TEM specimens were mounted on a carbon-coated 43 Z. R. Tian, J. A. Voigt, B. McKenzie and M. J. McDermott, J. Am.
copper TEM grid. Electronic absorption spectra were obtained Chem. Soc., 2002, 124, 12954.
using a Helios Beta Thermospectronic spectrophotometer. 44 L. E. Greene, M. Law, J. Goldberger, F. Kim, J. C. Johnson,
Infrared spectra were recorded using a Specac single reflectance Y. Zhang, R. J. Saykally and P. Yang, Angew. Chem., Int. Ed.,
2003, 42, 3031.
ATR instrument (4000–400 cm21, resolution 4 cm21). 45 M. Guo, P. Diao, W. Yan-Jie, B. Wang and S. M. Cai, Wuli
Huaxue Xuebao, 2003, 19, 478.
46 C. H. Hung and W. T. Whang, Mater. Chem. Phys., 2003, 82, 705.
Acknowledgements 47 M. Öner, J. Norwig, W. H. Meyer and G. Wegner, Chem. Mater.,
1998, 10, 460.
POB, DSB and KG thank the EPSRC and Royal Society 48 A. Degen and M. Kosec, J. Eur. Ceram. Soc., 2000, 20, 667.
(United Kingdom) and NRF (South Africa) for financial 49 M. Castellano and E. Matijevic, Chem. Mater., 1988, 1, 78.
support. The authors thank Dr Chengee Jiao from FEI 50 M. Ocana, M. P. Morales and C. J. Serna, J. Colloid Interface Sci.,
Company for FIB-TEM sample preparation. 1994, 171, 85.
51 A. Gutiérrez-Sosa, T. M. Evans, A. P. Woodhead, R. Lindsay,
C. A. Muryn, G. Thornton, J. Yoshihara, S. C. Parker,
References C. T. Campbell and R. J. Oldman, Surf. Sci., 2001, 477, 1.
52 S. Crook, H. Dhariwal and G. Thornton, Surf. Sci., 1997, 382, 19.
1 P. O’Brien and J. McAleese, J. Mater. Chem., 1998, 8, 2309. 53 M. A. Cousins and K. Durose, Thin Solid Films, 2000, 361–362, 253.
2 D. S. Boyle, K. Govender and P. O’Brien, Chem. Commun., 2002, 80. 54 D. M. Bagnall, Y. F. Chen, Z. Zhu, T. Yao, S. Koyama,
3 J. C. Johnson, H. Yan, P. Yang and R. J. Saykally, J. Phys. Chem. M. Y. Shen and T. Goto, Appl. Phys. Lett., 1997, 70, 2230.
B, 2003, 107, 8816. 55 P. Zhu, Z. K. Tang, G. K. L. Wong, M. Kawasaki, A. Ohtomo,
4 G. Bohnsack, Ber. Bunsen-Ges. Phys. Chem., 1988, 92, 803. H. Koinuma and Y. Segawa, J. Solid State Commun., 1997, 103, 459.
5 P. O’Brien, T. Saeed and J. Knowles, J. Mater. Chem., 1996, 6, 1135. 56 H. Shen, M. Wraback, J. Pamulapati, S. Liang, C. Gorla and
6 A. Degen and M. Kosec, J. Eur. Ceram. Soc., 2000, 20, 667. Y. Lu, MRS Internet J. Nitride Semicond. Res., 1999, 4S1, G3.60.
7 W. L. Bragg and J. A. Darbyshire, Trans. Faraday. Soc., 1932, 28, 522. 57 A. P. Collier, C. J. D. Hetherington and M. J. Hounslow, J. Cryst.
8 Structure Reports, vol. 27, eds. W. B. Pearson, I. D. Brown and Growth, 2000, 208, 513.
A. Mc. L. Mathieson, D. Reidel Publishing Co., 1962, p. 475. 58 J. Zhang, L. Sun, J. Yin, H. Su, C. Liao and C. Yan, Chem.
9 W. J. Li, E. W. Shi, W. Z. Zhong and Z. W. Yin, J. Cryst. Growth, Mater., 2002, 14, 4172.
1999, 203, 186. 59 M. Hilgendorff, L. Spanhel, Ch. Rothenhausler and G. Müller,
10 D. Li and H. Haneda, Chemosphere, 2003, 51, 129. J. Electrochem. Soc., 1998, 145, 3632.

J. Mater. Chem., 2004, 14, 2575–2591 2591

You might also like