You are on page 1of 163

GROWTH OF ANODIC SELF-ORGANIZED

TITANIUM DIOXIDE NANOTUBE LAYERS


(Wachstum anodischer selbst-organisierter Titandioxid Nanorhren Schichten)

Dissertation

Der Technischen Fakultt der


Universitt Erlangen-Nrnberg
zur Erlangung des Grades

DOKTORINGENIEUR

vorgelegt von

von Herrn Dipl.- Ing. Jan Mack

Erlangen 2008

Als Dissertation genehmigt von


der Technischen Fakultt der
Universitt Erlangen-Nrnberg

Tag der Einreichung: 29.2.2008


Tag der Promotion:

15.5.2008

Dekan: Prof. Dr. - Ing. J. Huber


Berichterstatter: Prof. Dr. P. Schmuki, Prof. Dr. M. Graham

Abstract
Synthesis of one dimensional nanostructures has been attracting significant and
continually increasing research interest over the past years, as these structures provide a high
potential for technological applications due to their intriguing properties. Of particular interest
are approaches that rely on a self-organization of the matter, as they provide a high degree of
order on the nanoscale and lead to exciting new arrangements. Within electrochemical
materials science, anodization approaches that lead to the growth of the self-organized porous
structures, such as alumina and silicon, have been widely developed during the past decades.
Due to a highly ordered structure and specific properties of these materials, they have found
use in various applications. In contrast to these two classical examples, no such ordered
nanostructures on titanium could be obtained until recently.
This thesis describes the synthesis of self-organized titanium dioxide nanotube layers
by an electrochemical anodization of Ti in various electrolytes that contain fluoride anions.
These layers offer a unique combination of wide band gap semiconductor properties with a
high surface area and precisely controlled morphology making them promising candidates for
direct applications in photocatalysis and solar energy conversion, for example. In order to
determine the conditions for the self-organization, a whole range of parameters was
investigated. The conditions that need to be fulfilled to grow ordered layers have been
identified within the present work. Further, by aid of modern surface analytical techniques,
the mechanistic aspects of the tube growth have also been investigated and discussed. Several
functional applications of the nanotube layers are presented and some others are outlined.

Kurzfassung
Die Synthese eindimensionaler Nanostrukturen stand in den letzten Jahren im Zentrum
vieler Forschungsinteressen, da diese Strukturen aufgrund ihrer faszinierenden Eigenschaften
ein hohes Potential fr technologische Anwendungen bieten. Bei der Herstellung sind
Methoden von besonderem Interesse, die auf der Selbstorganisation von Materie beruhen, da
so ein hohes Ma an Ordnung im Nanometerbereich erreicht wird und faszinierende neue
Anordnungen entstehen. Im Bereich der elektrochemischen Werkstoffwissenschaften wurden
Anodisierungsmethoden, die zum Wachstum von selbst-organisierten porsen Strukturen, wie
die des porsen Aluminium- und Siliziumoxids fhren, whrend der letzten Jahrzehnte weit
reichend entwickelt. Aufgrund der hchstgeordneten Struktur und der spezifischen

Eigenschaften dieser Materialien, haben sie in zahlreichen Anwendungen Nutzen gefunden.


Im Gegensatz zu diesen zwei klassischen Beispielen konnten bis vor kurzem keine
solchermaen geordneten Nanostrukturen auf Titan erzeugt werden.
Diese Arbeit beschreibt die Synthese von selbst-organisierten Schichten von
Titandioxid Nanorhren durch elektrochemische Anodisation von Titan in verschiedenen
Elektrolyten, die Fluoridanionen enthalten. Diese Schichten bieten eine einmalige
Kombination der Eigenschaften eines Halbleiters mit groer Bandlcke mit einer groen
Oberflche und einer genau kontrollierten Morphologie, was sie zu viel versprechenden
Kandidaten

fr

direkte

Anwendungen

im

Bereich

der

Photokatalyse

und

der

Solarenergieumwandlung macht. Zur Bestimmung der genauen Konditionen fr das selbstorganisierte Wachstum wurde ein sehr groer Bereich von Parametern untersucht. Die
Bedingungen fr das kontrollierte Wachstum geordneter Schichten wurden in dieser Arbeit
bestimmt. Des Weiteren wurden mechanistische Aspekte des Rhrchenwachstums untersucht
und diskutiert. Verschiedene funktionelle Eigenschaften der Nanorhrenschichten sind hier
lnger und einige andere kurz dargestellt.

Contents
1

Introduction....................................................................................
1.1 General
1.2 Scope and structure of this thesis .

Literature and theoretical background.......................................


2.1 Self-organized systems the role model porous alumina
2.1.1 Growth mechanism of porous anodic alumina..
2.1.2 Perfectly ordered porous alumina..
2.1.3 Applications of porous alumina.
2.2 Oxide films on Ti ...
2.2.1 Fundamentals of the passive oxide films on Ti ....
2.2.2 TiO2 film growth
2.2.3 Breakdown of passivity.
2.2.4 Film growth efficiency ................................
2.2.5 Stress effects in the formation of TiO2 (and other valve metal oxides)
2.2.5.1 Pilling-Bedworth ratio ....
2.2.5.2 Electrosriction .
2.2.5.3 Stresses due to transport of ions..
2.3 Ti anodization in fluoride containing electrolytes .
2.3.1 Background ..
2.3.2 First ordered structures
2.4 Properties of TiO2 .
2.4.1 Structure ...
2.4.2 Ionic properties ....
2.4.3 Electronic properties
2.5 Functional applications of TiO2 ..
2.5.1 Anodic TiO2 films
2.5.2 Nanostructured TiO2
2.6 Other methods for TiO2 nanotube synthesis .

Methods
3.1 Sample preparation .
3.2 Synthesis of TiO2 films
3.3 Surface analyses ..
3.3.1 Electron microscopy ...
3.3.2 X-ray diffractometry
3.3.3 X-ray photoelectron spectroscopy
3.4 Photoelectrochemistry ..

Results
4.1 Polarization curves in the fluoride containing solutions

1
1
4
5
5
6
8
8
11
11
14
19
21
22
22
23
25
26
26
29
32
32
34
35
38
38
40
43
45
45
45
47
47
47
49
49
50
50

4.2 Growth of the self-organized nanotube layers in acidic electrolytes.


4.3 Growth of the self-organized nanotube layers in neutral aqueous
electrolytes..
4.3.1 Anodization in Na2SO4 / NaF electrolytes
4.3.2 Anodization in (NH4)2SO4 / NH4F electrolytes .
4.3.3 Stages of the tube growth in aqueous electrolytes
4.4 Tailoring the dimensions of the self-organized nanotube layers in
non-aqueous electrolytes ..
4.4.1 Anodization in glycerol electrolytes .
4.4.1.1 Influence of temperature during anodization in
glycerol electrolytes..
4.4.1.2 Influence of water content in the glycerol electrolyte .
4.4.1.3 Variation of the tube diameter influence of applied potential ..
4.4.2 Anodization in ethylene glycol electrolytes ..
4.4.2.1 Hexagonal ordering of the nanotubes ..
4.5 Structure of the nanotube layers ..
4.6 Chemical composition of the nanotube layers .
4.6.1 Fluoride removal ...
4.7 Applications of self-organized TiO2 nanotube layers .

8
9

55
56
64
70
72
73
76
78
82
86

7.1 TiO2 nanotube layers as a promising and prospective material ...


7.2 Specific features of the nanotube layers ..

90
92
97
100
102
106
106
106
112
116
117
117
123
125
131
133
134
134
135
137
139
139
140

References ..
List of symbols ...

142
148

Discussion ..
5.1 Self-organized nanotube growth ..
5.1.1 Overview ..
5.1.2 Influence of potential ...
5.1.3 Influence of fluorides ..
5.2 Other factors of the growth of the nanotube layers ...
5.2.1 pH .
5.2.2 Stress effects .
5.2.3 Diffusion ...
5.2.4 Temperature ..
5.2.5 Role of the compact oxide layer ...
5.3 Other features related to the nanotubes ..
5.3.1 Growth efficiency .
5.3.2 Structure

6
7

53

Conclusions
Outlook ..

1. INTRODUCTION
1.1 General
Over the last ten years, the formation of self-organized nanostructures and
nanopatterns has attracted a great scientific and technological interest due to its far-reaching
and innumerable applications. Apart from these facts, the popularity and significance of
these self-organized nanostructures stem from the nature of their fabrication that relies on
self-organization processes (often called self-assembly). These processes represent basically
parallel methods for surface processing on the nano-scale, compared to fairly slow, tedious
and increasingly complicated sequential writing techniques with a resolution that depends
strongly on the quality and homogeneity of the material and to a certain extent on the
operator skills. The main advantage of self-organization processes is that they represent a
``smart nano-technique, i.e. no active manipulation on the nano-scale is required. Therefore
it is not surprising that a large part of materials science nowadays targets these nano-scale
fabrication techniques. Nanotechniques are a natural consequence of the necessity of
achieving smaller and smaller electronic and photo-devices that satisfy the actual requests of
the technological evolution.
Within materials science, a highly promising approach to form self-organized
nanostructured porous oxides is essentially based on a very simple process electrochemical
anodic polarization. Some important findings in this particular field include the growth of
ordered nanoporous aluminium oxide (Al2O3, alumina) [1-13] and ordered macroporous Si
[14-18]. Synthesis of both these materials has stimulated considerable research efforts and
given rise to many other materials to be processed in a similar fashion. Figure 1.1 shows
examples of these self-ordered structures - one can see clearly their high degree of ordering
and lateral homogeneity and regularity.
Figure 1.2 shows schematically the basic approach for the fabrication of ordered
porous anodic alumina [1,6]. Key is to find / elucidate the optimal electrochemical
conditions that lead to self-ordering. At the beginning of the present thesis, there was the
question, if similar nanostructures with highly ordered features, large surface area, chemical
and mechanical stability could be grown also on other materials. An ideal candidate would
be a metal that is easily accessible and forms a stable oxide on its surface. These criteria are
fulfilled for a group of so-called valve metals [19]. To this class belong, apart from Al, also

Ti, Zr, Ta, Nb, V, Hf, W and some less common metals (Bi, Mo, etc.).

a)

b)

Figure 1.1 Scanning electron microscope (SEM) images of (a) self-organized porous
alumina (taken from Ref. [7]) and (b) porous silicon (taken from Ref. [18]) formed by an
electrochemical anodization. The images show side views of the layers, the insets shows
magnified views (from the marked areas).

native Al2O3 layer


Al

Anodization
neutral
electrolyte

barrier Al2O3 layer

acidic electrolytes
+ optimized conditions

ordered porous Al2O3 layer

Al

A
Figure 1.2 Two distinct Al2O3 morphologies resulting from the aluminium anodization [1,6].
2

In general, Ti can be taken as a straight-forward example and as an excellent


candidate, as its stable oxide titanium dioxide (TiO2) has an extremely wide range of
functional applications based on its exceptional semiconducting properties and great
biocompatibility. Some most important TiO2 applications are in photocatalysis, solar energy
conversion and biology, and will be discussed in more detail in Chapter 2.
For a long time, it appeared to be impossible to achieve growth of self-organized
porous TiO2 by electrochemical anodization of Ti. One of the reasons was that TiO2 is
chemically very stable and only few inorganic compounds, for instance HF and H2O2, can
significantly chemically dissolve it. Nevertheless, in 1999, a French group showed for the
first time that self-organized porous structures of TiO2, as shown in Fig. 1.3, with thickness
in the range of a 100 nm can be formed by anodization of Ti under specific conditions,
namely in chromic acid HF-containing electrolyte [20, 24].

Figure 1.3 SEM image of self-organized porous TiO2 structure formed by anodization of Ti in
a chromic acid/HF mixture at 5V for 20 min (taken from Ref. [20]).
It is noteworthy that in all later works, detailed analysis of the structure shown in Fig.
1.3 and related other structures revealed that they consist essentially of nanotubes, instead of
nanopores. As a consequence, a term ``nanotubes has been adopted for these nanostructures.
Since then, this topic has been a center of research in our laboratory and many other
laboratories world-wide. How seriously and rapidly this field has been evolving is
demonstrated in Fig. 1.4 showing the number of publications that have been released since
1999, with nearly 50 % of the work coming from our laboratory.

40
35

Publications

30
25
20
15
10
5
0
1999 2000 2001 2002 2003 2004 2005 2006 2007

Year

Figure 1.4 Number of publications in the field of self-organized TiO2 nanotube layers

1.2 Scope and structure of this thesis


The present thesis describes a part of the effort of our lab in the formation of selforganized TiO2 nanotube layers by electrochemical anodization of Titanium. The major tasks
of this thesis include studies of the growth of the TiO2 nanotube layers by electrochemical
anodization of Titanium, elucidation of the tube growth mechanism and tailoring the tube
morphology.
Chapter 2 gives an overview about the formation and applications of self-organized
porous layers and summarizes the literature on the formation, properties and applications of
TiO2, with major attention given to electrochemical anodization. Chapter 3 gives an
overview of methods and techniques used in the thesis. In Chapter 4, growth studies and
factors influencing the morphology and structure of the nanotubes are presented. In
particular, it also describes a few applications of nanotubular TiO2 layers that are not only of
scientific interest, but also of potential technological relevance. Chapter 5 provides
discussion of the results shown in Chapter 4 and aims at detailed explanations of the
parameters influencing the tube growth. In the last two chapters, the findings of the thesis are
summarized (Chapter 6) and some further perspectives for the nanotubular layers are
outlined (Chapter 7).

2. LITERATURE AND THEORETICAL BACKROUND


This chapter firstly emphasises the porous anodic Al2O3 (PAA) films and summarizes
some most important applications of this material. Subsequently, it deals with anodic TiO2
films and addresses important features regarding the methodology, properties, structure and
applications of not only anodic, but also of nanostructured TiO2 films.
In general, the electrochemical anodization (or simply anodization) is a method that is
widely used for various purposes: to protect metal from atmospheric corrosion, to provide a
decorative finish, to reduce the friction on sliding surfaces and to prevent galling [19,22-24].
It is a simple technique with a low cost, because it does not require the use of expensive and
extensive experimental instrumentation.
In additition to anodization, there is a whole range of other physico-chemical
techniques to form various metal oxides. These include thermal oxidation [25], plasma
oxidation [26], magnetron sputtering [27], physical vapour deposition [28], chemical vapour
deposition [29], atomic layer deposition [30], pulsed-laser deposition [31], hydrothermal
treatments [32] and rarely also electrodeposition (cathodic electrosynthesis) [33].
In the next section, the synthesis of one of the most investigated self-organized materials
- porous alumina - is described. Particular attention is given to porous alumina applications
in order to demonstrate the versatility of this material and to outline possible uses of TiO2 in
the light of the Al2O3 success and popularity.

2.1 Self-organized systems the role model porous alumina


Anodization of aluminum has been well known for many decades. The primary target
of the Al anodization was decoration and corrosion-protecting of its surfaces by a barrier
oxide layer aluminium oxide Al2O3 [1,34]. The basic model for porous alumina was
reported in 1953 by Keller [1], who described its morphology as an almost ideally
hexagonally close-packed duplex structure consisting of porous and barrier Al2O3 layer and
demonstrated the contrast between the anodization of aluminium in neutral and acidic
electrolytes. As shown in Fig. 1.2, the first case leads to a so-called barrier type of alumina
layer, while the latter case may lead to a porous alumina layer formation due to an enhanced

oxide dissolution, induced by an electric field. Example of a typical porous alumina layer is
shown in Fig. 2.1.

Figure 2.1 SEM image of porous anodic alumina (taken from Ref. [35])
Later, various morphologies of the porous layer were achieved by a tailored
anodization resulting in layers having a large surface area and very narrow size of pore
distribution [2]. In the 1980s, O` Sullivan, Thompson and Wood [3-5], with the aid of
transmission and scanning electron microscopy techniques (strongly developing at that time),
provided a significant step towards the understanding of the growth mechanisms of alumina
oxide layers. It was only in 1995 that the Japanese researchers Makuda and Fukuda [6], after
continuous effort, succeeded in growing much more ordered layers, known as perfectly
ordered porous alumina (or eventually honeycomb porous alumina).
All in all, porous alumina is probably the best investigated system for self-organized
pore formation to date, as one can deduce from a vast number of publications in this field.

2.1.1 Growth mechanism of porous anodic alumina


The pore formation is a complex process that requires a specific localized distribution
of oxide formation and dissolution rates, which are determined by the electrical field in the
oxide, stress factors and the local chemistry. Porous alumina arrays are typically obtained in
several distinct electrolytes, such as phosphoric, oxalic, chromic and sulfuric acid under fixed
voltage and temperature conditions, which depend on the type of electrolyte used [1-13].

Considering the morphology evolution and the corresponding current response in time,
it is possible to separate the pore growth into four distinct stages as shown in the pore
formation mechanism in Fig. 2.2 [36]. In the course of anodization, firstly a barrier oxide on
the entire Al area is formed (Fig. 2.2a). On distinct locations (with defects and stresses) an
enhanced dissolution, due to a non-homogenous electrical field, takes place that leads to the
development of penetrations paths, where essentially the majority of potential lines are
concentrated (Fig. 2.2b). This leads to a local increase of the field strength and a consequent
enhancement of their development through field-assisted dissolution. This is an auto catalytic
process (Fig. 2.2c), which continues until the steady state anodic film morphology is created
(Fig. 2.2d).

Figure 2.2 Growth model of porous anodic alumina (taken from Ref. [36])
In general, the structure of the porous oxide is considered two-fold (duplex),
consisting of a high purity oxide (on the interface with Al metal) and the oxide with
incorporated anions (on the interface with an electrolyte) [35-37]. Furthermore, there are three
classes of electrolytes in terms of ionic mobility in the oxide: immobile, outwardly mobile and
inwardly mobile ions. Thus, the resulting thickness of the layer is strongly influenced by the
directionality of the electrolyte species [38].

2.1.2 Perfectly ordered porous alumina


The generation of honey-comb porous alumina oxide layers with a large aspect ratio,
introduced by Masuda and Fukuda in 1995 [6], can be considered as a foundation stone in
this research area. Example of such a structure is given in Fig. 2.3. In order to grow these
layers, very highly defined anodization conditions are required, including the right potential,
temperature and electrolyte, purity and pretreatment of the Al substrate. The anodization
itself is done in two steps: a long first anodization step is performed, the porous layer is next
removed and the second anodization is performed leading to an extremely ordered porous
structure. The pores are arranged in a hexagonal lattice with a pore size of approximately 150
nm. The formation of highly ordered patterns can often be improved by a pre-sensitization of
distinct surface location by various lithographic techniques (using an appropriate mask), or
indentation techniques (using a mold with defined dimensions) that predefine the initiation
sites for etching [7,9,35].

Figure 2.3 SEM top-view (left) and side-view (right) of perfectly hexagonally ordered anodic
porous alumina membrane after etching the Al substrate (taken from Refs. [35,39]).
Nowadays, more complex shapes of pores can be grown by a choice of anodization
parameters [11-13]. Probably the most spectacular is a recent finding of so-called hard
anodization approach for porous alumina layers that is accompanied by a high electric field,
powerfully driving the dissolution processes and therefore allowing the growth of porous
layers with very high aspect ratio in comparably shorter periods [12,13].

2.1.3 Applications of porous alumina


As already mentioned, the great interest in self organized porous alumina is mostly

based on its applications. Probably the most significant is its use as a template, where a
secondary material is deposited inside the highly ordered pores, and this either adds
functionality to the template, or the template is selectively removed and the highly defined
secondary material with possibly interesting optical, magnetic and surface-related properties
is left behind.
The use of porous alumina as a template dates back several decades, when nanoporous
membranes were used for synthesis of nanofibrils [40], or for deposition of magnetic metals
to make magnetic recording media [41]. More recently, deposition of conducting polymers
[42] and different metals into the membranes has become possible [43,44].
Nowadays, perfectly ordered porous alumina templates are available [6] and, mainly
by a choice of the electrolyte and the applied potential, the pore diameters can be tuned
between some nanometers and several 100 nm [10-13, 35]. This property is excellent for
achieving the nano-size requirements necessary for e.g. magnetic devices (e.g. dots,
nanowires, where their magnetic domain size plays a key role in determining the exchange of
magnetic forces [43, 45]), but also for direct applications of PAA, in optical devices (e.g.
photonic crystals [46]), bio-compatible substrates [39] and for micro- and ultrafiltration [47].
Within the gerenal concept of using PAA as a template, the interior of the pores may
be either filled with the secondary material precursor, or be electroplated with the material in
order to generate e.g. nanowire arrays, as shown schematically in Fig. 2.4. By this approach,
essentially all classes of materials (e.g. metals, semiconductors and polymers) can be prepared
by electrochemical and chemical means [42-44, 48-50].

Selective Al2O3 removal


Deposition
Porous anodic Al2O3

Nanowires

Figure 2.4 Regular pore array of porous alumina as a template for micro- or nano-fabrication.
A thick porous oxide is grown by anodization, then the material is deposited into the pores,
finally the aluminum template is dissolved leaving nanowires of the deposited material behind.
9

Figure 2.5 shows three examples of the PAA used as a template for electrodeposition.
In the first case (Fig. 2.5a), PAA served as a host for electrodeposited Au nanofibers. In the
second case (Fig. 2.5b), Ni was filled into PAA in order to create magnetically active
structure. In the third case (Fig. 2.5c), arrays of Pt nanowires have been synthesized by (i)
electrodeposition of Pt into PAA and (ii) subsequent selective PAA dissolution (as sketched in
Fig. 2.4).

a)

b)

c)

Figure 2.5 Examples of PAA template for electrodeposition of (a) Au (taken from Ref. [44]),
(b) Ni (taken from Ref. [48]) and (c) Pt nanowire arrays (after PAA selective deposition,
taken from Ref. [50]).

In summary, it is evident that use of the porous alumina template and the tailored
filling provide an excellent tool for fabrication of various nanostructures with highly defined
dimensions and structure (with e.g. possibility to achieve single crystal materials). However,
alumina is an insulator in contrast to TiO2, which is a wide band-gap semiconductor. In
principle, very interesting functional structures could be achieved via deposition into TiO2
nanotube layers for applications in e.g. solar energy conversion.

10

2.2 Oxide films on Titanium


As for all non-noble metals, Ti is coated in most oxygen containing environments with
a native oxide layer. However, this layer is typically on the order of a few nanometers in
thickness. Typical approaches to grow thicker TiO2 layers include two techniques [22]: an
anodic oxidation of Ti, where an applied electric field controls the ion (O2-, Ti4+) transfer
during the oxide growth, and a thermal oxidation of Ti, during which the oxygen ion diffusion
controls the TiO2 layer thickness. Both processes are time-dependent: in the first case, the
applied potential sets an electric field that is proportional to a logarithm of time, whereas in
the latter case the oxygen diffusion is proportional to the square root of time [22]. One should
also mention that Ti is a typical example of a valve metal (as well as Al, Ta, etc.). Valve
metals represent a class of metals that can be anodized to grow considerably thicker oxide
layers (several 100 nm), whereas on the other common metals (e.g. Fe, Ni), only a few nm
thick oxide layers can be achieved.
On the other hand, there are a number of approaches to decrease the thickness of the
oxide layer, or to completely remove it. In addition to the mechanical and physico-chemical
methods such as polishing, or sputtering, purely chemical approaches can also be used.
Typical compounds that can chemically dissolve TiO2 are namely acidic HF and H2O2
solutions and highly alkaline hydroxide solution (via a formation of soluble complexes).
Immersion of Ti / TiO2 substrates in these agenst may lead to complete TiO2 removal, but the
Ti surface after such treatment will again immediately reoxidize and form a thin native TiO2
layer, as soon as it has an access to the atmosphere, or moisture.

2.2.1 Fundamentals of the passive oxide films on Ti


Based on thermodynamical considerations (Nernst equation, solubility products,
chemical equilibria and the resulting Gibbs free energy) the state of the metal in the solution
(with certain pH and concentration of redox species) can be predicted. The result of such
calculations is often presented as pH vs. potential diagram, known as the Pourbaix diagram
[51]. This diagram provides information, whether the metal is inert, actively dissolving,
hydrated, or covered with a stable oxide layer. The diagram for Ti is shown in Fig. 2.6. As it
can be seen, in the anodic region of the diagram (above the H2O / H2 line), TiO2 is present
over the whole pH range. Below the H2O / H2 line, Ti tends to from titanous and titanyl

11

species over the whole pH range. However, these metastable species typically immediately
react to Ti4+ if the potential rises and form a stable TiO2). This formed native TiO2 layer has
an excellent corrosion resistance. The thickness of such an oxide layer is usually in the range
of few nanometers (depending on oxygen content in the surrounding environment) and it can
be measured by surface analytical techniques. With increasing anodic potential, the thickness
of this protective layer can be further increased, as shown later in this section.
Pourbaix diagrams are available for many elements in aqueous environments and are
often a valuable tool in the preliminary assessment of the thermodynamic stability of system.
However, they have several drawbacks. They do not provide information on the kinetics of
the processes, they consider only pure elements and O, OH-, H2O species (no other
complexing agents) and low ionic concentrations of the species (< 10-6 mol / l).

Figure 2.6 Pourbaix diagram (taken from Ref. [19]) for the system Ti H2O at 25C (10-6
mol / l ionic concentration). Electronic data are added: UVB, UCB, UFB, USS match positions of
the valence band, the conduction band, the flat band potential and surface states. Eg = band
gap energy. ox = U - UFB = potential drop in the oxide film (meaning of these values will be
explained later in the section Electronic properties).
To evaluate the kinetics of the reactions by electrochemical methods, it is often
convenient to use polarization curves [22,23,52]. A typical semilogarithmic polarization curve
is shown in Figure 2.7. For simplicity, let us consider only the anodic direction of polarization,

12

i.e. towards positive potential. When sweeping the potential in the positive direction, from the
open circuit potential (Uocp), the current will increase quickly to considerably high values. It is
noteworthy that this shape of the curve is often characterized in terms of the Tafel equation
[53]. Under oxide formation conditions, it will reach a critical value of current (jcrit) and
passivation of the metal will take place (the active / passive transition). This oxide formation
is manifested by a dramatic decrease of current in the polarization curve at the corresponding
potential the passivation potential (Up). The anodic current density is lowered by several
orders of magnitude and levels off to the passive current density (jp). With a continous sweep
towards higher potential, a situation shown on the right side of the curve will occur (with an
onset depending on the material and experimental conditions). This region, where the current
rapidly increases as the potential drifts up, is the trans-passive region. It can be ascribed
essentially to one, or a combination of these processes: further oxidation or decomposition of
the film, oxygen evolution or other electrolyte decomposition reactions [22].

Active
dissolution

Trans-passive
region

Passivation

Anodic

log j

Non-valve metals

jcrit
Uocp

Valve metal s

jp
Up

Ubd

Figure 2.7 Typical anodic polarization curve for a metal electrode exhibiting passivation at a
distinct anodic potential, Up. The different regions correspond to active dissolution, passivity
and trans-passive range. The current flow in the passive range (jp) is a measure for the
resistance, or the quality of the oxide film. In the trans-passive region of the curve, at
potentials higher than Ubd, electric, or ionic breakdown of the film occurs [54].
It is noteworthy that depending on the metal, the width of this passivation region may
vary significantly (under certain experimental conditions). For all valve metals, including Ti,
this region is very wide (dozens of volts), whereas for instance for Fe it is less than 2V [55].

13

However, once the passive film is formed, it should not be considered as a rigid layer, but
instead as a system in dynamic equilibrium between film dissolution and growth. In principle,
the chemical composition and the thickness of electrochemically formed passive films depend
(apart from the base metal), on the anodic potential, time, electrolyte composition and
temperature, etc. For Titanium and other valve metals, the thickness increases linearly with
the anodic potential (up to several 10 nm), thus enhancing their passivation in the passivity
region of the curve shown in Fig. 2.7. However, trans-passivity and anodic breakdown are
usually connected with irreversible changes of the surface morphology and damage of the
material.

2.2.2 TiO2 film growth

The growth of TiO2 by anodization can be described simply by the following


reactions:
The oxidation (anodic) reaction occurs at the Ti metal/oxide interface:
Ti Ti4+ + 4 e-

(2.1)

The corresponding cathodic reaction is a reduction of H2O at the counter electrode:


4 H2O + 4 e- 2 H2 + 4 OH-

(2.2)

A schematic of the TiO2 growth is shown in Figure 2.8 [19,22,23]. The Ti metal is
connected as an anode in an electrolytic cell with an electrolyte that does not dissolve the
oxide. In this case, the presence of water in the eletrolyte is essential, as it acts as a source of
O ions. The applied current sets up an electrostatic field accros the oxide (or increases a field
that already exists) and this maintains Ti oxidation and transports O2- ions inward and Ti4+
ions outward and thus leads to a continued growth of the film (thicknening of the film).
The detailed anodic reaction, known as field-aided hydrolysis, can be written as:
Ti + 2 H2O TiO2 + 4 H+ + 4 e-,

(2.3)

thus illustrating that at the anode, H+ ions are permanently produced, leading to a local pH
decrease.

14

Counter
electrode

Ti

2-

H2
OH

Ti

H2O

H 2O

4+

e-

Ti

TiO2

Electrolyte

Figure 2.8 Growth of TiO2 on the Ti electrode in the presence of an externally applied
potential.

Oxide growth can essentially be described as a problem of ionic conduction (or mass
transfer) [23] as at least one of the ionic species has to diffuse or migrate through the oxide.
Essentially one of the ionic species (Ti4+ or O2-) may be transported faster under the field
thus has a higher transport number (to be discussed in the section on Ionic properties).
Accordingly, the layer grows at the inner, or at the outer interface (alternatively the ion
through-transport can be formulated as cation and anion vacancies movement through the
lattice). Many anodic oxides, including TiO2, are semiconductors (as discussed later in section
Electronic properties). Because of a non-equilibrated stoichometry, this defect structure may
strongly influence the ionic and electronic transport [56].
Models for the oxide growth consider different processes to be rate-determining [23,
56]. The most basic approach assumes that the ion transport through the film is controlled by
the electric field across the layer. In these models, an activation energy for an ion or vacancy
hopping is lowered by the applied field (the high-field mechanism). The relevant activation
energy for the hopping process can either be between the oxide lattice sites, or between the
oxide interfaces. In the latter case, the ion entrance at both oxide interfaces is the ratedetermining step. In addition, the mixed conduction model considers both of the above
mentioned factors to have a comparable influence.

15

These high-field models have been revealed to be relevant for the TiO2 growth, as well
as for any other valve metal oxides. The first experiments regarding the ionic conductivity of
oxide films on aluminium have been firstly performed by Gnterschultze and Betz in 1934
[57] and by Verwey in 1935 [58]. The first two researchers realized that current density j was
exponentially dependent on the field strength F and suggested following form:
j = . exp(.F),

(2.4)

where j represents the current density, and are constants and F is the field strength
across the oxide. They provided values and for a variety of valve metals based on the
thickness obtained from capacity measurements.
The metal/oxide interface was firstly taken in account soon after by Mott and Cabrera
[59, 60] and Vermilyea [61]. Dewald has made the extension to their theory by including the
effects of the space charge layer [62]. At low or moderate temperatures and in the presence of
sufficiently high fields (> 106 V / m), pure diffusion of ionic species has negligible
contribution for the film growth.
The field strength F across the oxide film can be estimated from the potential drop
across the oxide (U) and the thickness of the oxide d:

F=

U
d

(2.5)

Therefore, the thickness of the barrier layer is inversely proportional to the logarithm
of current density (j) for a given anodization potential U.
Once a constant positive potential is applied on the working electrode, the oxide film
starts to grow and this leads to an exponential decrease of the current density as the field
strength across the oxide continuously decreases [63]. According to Eqs. 2.4 and 2.5, the film
growth (at constant voltage) is self-limiting as F is lowered with increasing film thickness.
This results in an inverse logarithmic law:

1
= A B log(t ) ,
d

(2.6)

where d is thickness of the oxide, A and B are constants and t is time.

16

Thus, in principle an infinite oxide growth would take place. However, often the
limiting situation is reached, when the formation rate of the oxide has dropped to the (in many
cases very small) chemical dissolution rate in the environment, as shown schematically in Fig.
2.9. This also means, that with an extended anodization period, the oxide thickness is not
significantly altered, unless the anodic potential is changed [19, 64].

Vgrowth

high field

observed result ( Vgrowth Vdiss)


steady state thickness
steady state field
Vdiss

chemical dissolution
Time

Figure 2.9 Competition of the growth rate (Vgrowth) and the chemical dissolution rate (Vdiss) of

TiO2 during the anodization resulting in a steady-state thickness of TiO2.

Therefore the final thickness of the oxide film is proportional to the equilibrium field
and can be approximated by a linear dependence on the formation voltage. Hence, the oxide
thickness d can be expressed as:
d = k.U,

(2.7)

where U is the applied potential and k is a growth constant.


Another interesting view of this growth issue is based on the relationship of the field
strength, film thickness and the current density. From equations 2.4 and 2.5 it is evident that
an infinitesimal increase of the thickness results in (i) a decrease of the field strength (Eq. 2.5)
and (ii) an exponential decrease of the current density (Eq. 2.4) and vice versa. This might
have certain implications particularly for non-flat substrates. In reality, an average metallic
substrate has a lot of defects (scratches, dislocations, etc.) and therefore one would expect
non-homogenous current flow through these areas (and thus a non-homogenous thickness
over the entire substrate). However, experimental results of Schultze et al. [64,65] regarding

17

the influence of the field effects on the oxide film thickness have suggested that the oxide
layer thickness does not change statistically, but it is almost homogenous with local changes
of a very few %. This means, in other words, that if current flows at one moment through a
thinner oxide area (as it can more easily flow there), this area will get thicker and the current
will likely find again another (easier) path to flow and thicken the oxide film there. At the
end, the substrate is covered with an oxide layer of homogenous thickness.
In any case, the anodic formation of TiO2 layers in a wide range of electrolytes and
over a wide range of anodic potentials results in an excellent resistance of Ti to corrosion, as
indicated electrochemically by a low level of electronic conductivity, great thermodynamical
stability and a low dissolution tendency in aqueous environments [19]. However, according to
a vaste number of reports [19,24,65-69], each electrolyte demonstrates an intrinsically
specific growth constant in the range of 1.3 3.3 nm of thickness gained per applied anodic
volt. These values correspond to an equilibrium field strength of approximately 106-107 V/m
[67]. Thus, the growth rate depends on a particular electrolyte and its concentration,
temperature, potential sweep rate, agitation speed and cathode to anode surface area ratio.
This ionic influence can be ascribed to the fact that the different electrolyte anions may
compete with O2- for inward migration and thus enter the film to different degrees.
To determine the oxide thickness, optical methods, such as ellipsometry [70] (or
spectro-reflectometry [71]) were used for several decades. At the present time, the thickness
of the oxides can be measured much more accurately by modern spectroscopical techniques,
combined with a wide variety of polishing and electropolishing techniques that guarantee the
the necessary surface flatness of the substrates.
Under cathodic polarization a range of valve metal oxides show significant cation
inward diffusion, accompanied by alterations in the electronic structure of the oxide (e.g.,
incorporation of additional states within the band gap). The most likely mobile species are
protons. For example, Dyer and Leach [66] examined oxidized titanium and niobium and
found that hydrogen enters the film under cathodic bias in non-negligible amounts. In the case
of TiO2, cathodic reduction of anodically passivated electrode converts the TiO2 into TiOOH,
as concluded from observations of the change in the absorption coefficient and refractive
index. It was also shown that the reduced oxide species (Ti3+) can subsequently dissolve to
into the solution and that this reduced oxide is an intermediate state in the active dissolution
process [71,72]. Thus, cation diffusion into the passivating films can be substantial under
cathodic bias. Hydrogen ingress into Ti oxide layers and Ti base metal (forming hydrides

18

which lead to material embrittlement [73, 74]) by cathodic polarization has recently also been
studied by in situ neutron reflectometry [75].

2.2.3 Breakdown of passivity


Under certain conditions, the passive state of metal (created by an oxide layer) is prone
to localized instabilities. These can be of electronic nature (leading e.g. to dielectric
breakdown) or due to the chemical reactivity (leading to e.g. enhanced localized oxide
dissolution). The reasons for these local instabilities are manifold and have origins in a wide
variety of bulk inhomogeneities (grain boundaries, dislocations, precipitates, inclusions) and
properties of passive films (thermally induced film rupture, high resistivity, eletrostriction,
local composition and structural differences).
Breakdown, the local destruction of the passive film, occurs under specific conditions,
namely in the case of high field strength (dielectric breakdown), attack of halide ions
(localized corrosion) and by hydrogen evolution (layer lift-off) [67,76-84]. The dielectric
breakdown of an oxide layer takes place at comparably high voltages, when extra charge
carriers are generated either by tunneling of electrons (Zener mechanism), or by collision of
electrons (avalanche mechanism) [67,78]. Therefore, the resistivity of the film plays a crucial
role for the breakdown to occur and it is influenced by the the band gap energy, structure and
impurities. During the breakdown, the resistivity of the layer drops by several decades.
Ikonopisov proposed that for TiO2 the avalanche mechanism involving electron injection from
the electrolyte into the conduction band is most likely responsible for dielectric breakdown
[82]. Further, he reported that composition and concentration of the electrolyte also play an
important role, whereas the current density, temperature and roughness of the surface are less
significant.
In general, the onset of the layer breakdown is apparent in I-U curves (as shown in Fig.
2.7) by a rapid current increase (or sometimes just by current or potential fluctuations), at a
distinct breakdown voltage Ubd. It is typically accompanied by sparking, or acoustic crackling
noise.
Anodic breakdown of oxide films on a wide range of valve metals (Nb, Ta, Zr, Hf, Al,
Ti, W, Mo) has been investigated for years by many researchers [20,23,76-84]. In the case of
Ti, it has been mainly investigated in H2SO4 electrolytes [78,79,83]. However, for all valve
metals it has been shown that breakdown onset is dependent on the growth mode

19

(galvanostatic, potentiostatic) and on the rapidity of voltage rise [19,76]. Breakdown is


accompanied by oxygen evolution and in some cases also by crystallization [76]. There has
been a lot of discrepancy in the literature on the exact reasons for the breakdown phenomena.
Generally, it can be stated that it is linked with film damage, oxygen evolution and change in
the electrical characteristics. Nevertheless, it is still a subject of debate, whether the quality of
the metal substrate will play any role in the breakdown. In any case, breakdown of the passive
film will happen at high voltages on a rough surface, as well as on a perfectly polished surface.
For extended anodization at high voltages, the breakdown spots may meander over the
surface (a new breakdown site occurs in the neighborhood to an old spot, while the old spot is
being continously shut down by a repassivating oxide). This type of anodization is called
``spark anodization [79,84,85] and it leads to a surface covered with an irregular structure of
pores with diameter in the range between hundreds of nanometers up to few micrometers in
size (depending on the breakdown potential). An example of this structure is shown in Fig.
2.10. It can be seen that the layer does not have a constant thickness (or is not microscopically
flat) across the surface. This can be explained in terms of different ionic current concentration
flowing through the breakdown spots, thus leading to growth of oxide with different thickness
around these spots.

Figure 2.10 SEM image of porous TiO2 layer on Ti grown by sparking anodization at 150
V in H2SO4 (taken from Ref. [85]).

Although in electrochemistry the dielectric breakdown of valve metal oxides is the


most investigated case, there is the second reason for the local destruction of passivity
presence of aggressive ions that may penetrate into the oxide film. These ions most typically
include chloride anions, whereas effects of fluorides are less studied. However, their influence
20

should not be underestimated, as shown in the section Ti anodization in fluoride containing


electrolytes. The essence of localized corrosion lies in distinct anodic sites on the surface,
where active metal dissolution reaction occurs (Me Me2+ + 2e-) by aid of halide ions
surrounded by a passivated cathodic zone, where the reduction takes place (depending on pH,
e.g.: 2H+ + 2e- H2). As a result, an active pit in the metal is formed that continues growing
since the dissolution process is autocatalytic [86]. Although this is typically a problem for Fe
and steels, also in the case of Ti, localized corrosion may have severe impact. For example,
localized breakdown of passivity of Ti, mainly crevice corrosion, can take place at elevated
temperatures in seawater and in deaerated acids [87,88]. Pitting of Ti has mainly been studied
in the presence of Cl- [66,67,84,89] and Br- anions [88, 90-92].

2.2.4 Film growth efficiency


For the determination of the effective growth efficiency, it is essential to know the
theoretical mass of the oxide that is produced by the anodic oxidation. The mass of TiO2 (as
well as of any other oxide) can be expressed in terms of Faradays law [93]:
m = qM / Fz,

(2.8)

where m is the oxide mass, q is the charge obtained by integration of current in time, z is
number of exchanged electrons and F is Faradays constant. The result of this calculation is
then compared with the experimentally obtained mass of the oxide film. The most straight
forward method is to weigh the oxide [22]. The weight of the oxide film, as well as the weight
of the metal used for the film growth, may be determined by stripping of the films by
mechanical means, or by chemically dissolving the oxide (leaving the metal) or the metal
(leaving the oxide). To determine the thickness from the increase in weight of the electrode,
apart from the geometry, the composition and the density of the film must be known. With the
latter parameter, one may have a large calculation error, as the densities are reported typically
for bulk crystalline oxide, while the films are often amorphous. Therefore, sputtering
techniques might be more suitable to determine the thickness. Nevertheless, the growth
efficiency of an oxide film can be evaluated by using either approach.
In reality, the oxide growth efficiency can never reach 100 % due to some side
reactions (most typically oxygen evolution) causing current leakage. It is also worthy to

21

mention, that in the case of breakdown anodization, the oxide film growth efficiencies are
very small due to an excessive oxygen evolution [81].

2.2.5 Stress effects in the formation of TiO2 (and other valve metal oxide)
It has been well established that compressive or tensile stresses are generated during
anodic oxide film formation on various metal electrodes [80,83,94-100]. These stresses play
an important role in the quality of the films, namely for the shape and quantity of cracks
created in the film while growing, and may have a strong influence on the breakdown of the
films. It has been revealed that the most important factor influencing stresses in the films is
the volume expansion, or shrinkage of the anodic oxide film during its growth. In this
particular case, the direction of the stress (compressive or tensile) is determined by the
Pilling-Bedworth ratio [101]. Two other important factors have been determined for the stress
generation in the oxide films: electrostriction [100] and the transport number of mobile anions
through the film [94]. In addition to these factors, there are other, less significant effects that
may cause stresses in the film, such as crystallization of an amorphous film and dehydration
of a hydrous film [83].

2.2.5.1 Pilling-Bedworth ratio

Pilling-Bedworth ratio (R), is defined as the ratio of the molar volume of the metal
oxide, which is produced by the reaction of metal and oxygen, to the consumed metal molar
volume:
R = V MeO produced / V Me consumed = M.d / m.D,

(2.9)

where M and D are the molecular weight and density of the metal oxide, m and d are the
atomic weight and density of the metal [101].
As a consequence, if R is lower than 1, a tensile stress in the film is generated, and the
metal oxide tends to be porous and non-protective because it cannot cover the whole metal
surface. If R is higher than 1, a large compressive stress is likely to exist in the metal oxide
that may lead to buckling and spalling. The amorphous oxide generally occupies a larger
volume than the same weight of the crystalline oxide (the density of amorphous film is lower
than that of a crystalline film). Theoretically, the growth of crystalline oxide will be favoured,

22

whenever the amorphous oxide would be under compressive stress. However, this is very
often not the case, as discussed in section Structure.
Nevertheless, for the complete evaluation of the oxide quality, other factors, such as
thermal expansion coefficients and adherence between MeO and Me must be taken into
account. It is obvious that to form a protective oxide on a metal surface, R should not be
significantly higher than 1, the thermal expansion coefficient of MeO and Me should be
similar and the adhesion of MeO should be very high.
Table 2.1 Summary of R for selected valve metals in their crystalline and amorphous forms

(taken from Ref. [94]).


Metal
Al
Ta
Ti
Nb
Zr
Hf

R (crystalline MeO)
1.3
2.5
1.9
2.5
1.5
1.6

R (amorphous MeO)
1.6
2.6
2.4
2.8
-

2.2.5.2 Electrostriction

Electrostriction becomes important for the compressive stress generation in the


dielectric films [97-100,102]. The high electric field across the film exerts an electrostrictive
force normal to the plane of the film. For example, Wtrich estimated experimentally that the
contribution of electrostriction to the total compressive stress for a flat Al2O3 film would be
approximately 50 % [102]. Nelson and Oriani measured stresses generated during the anodic
oxidation of Ti and Al using a bending cantilever beam technique [94]. Although some scatter
was present in their work, they have shown that a tensile stress develops at the metal-oxide
interface, due to a volume difference between the oxidized metal and the forming oxide and
also due to the electrostriction. Slowly grown oxides had greater electrostrictive deflections
than more rapidly grown oxides. Recently, Seo and coworkers studied more accurately the
stresses in anodic oxide films on Ti by using a beam bending method, i.e. using laser
irradiation of a thin metal electrode coated onto one side of a glass plate [100]. While being
anodized, the glass plate was bending and the beam reflection angle was changing (and
recorded) and the changes of bending radius were subsequently converted quantitatively into
a change of stress in the growing film. They have shown that the contribution of the

23

electrostrictive stress to the total compressive stress of the anodic oxide film on titanium can
be evaluated both experimentally and theoretically.
el

The stress, , normal to the plane of a dielectric oxide film due to electrostrictive force
z

can be expressed by using a simple parallel-plate capacitor model, as in Eq. 2.10:

el r o ( ) 2
=
z
d
2

(2.10)

where , o , r , and d are the potential difference across the oxide film, the permittivity in
vacuum, the relative permittivity, and the thickness of the oxide film, respectively. Since the
oxide film is constrained by the metal substrate, a lateral compressive stress,

el
xy

, is

developed and given by Eq. 2.11:

el
xy

el
ox r o ( ) 2
ox
,
=
z 1 ox 2
1 ox
d

(2.11)

where ox is the Poissons ratio of the oxide film. However, the semiconductor nature of TiO2
has to be taken in account. TiO2 is an n-type semiconductor, therefore one can assume, that if
the film is not dielectic, but semiconductive, the electric field in the film is present across a
space-charge layer instead of the whole film thickness (meaning of the term ``space-charge
laye will be explained later in the section Electronic properties). The width of the spacecharge layer, W, can be expressed as in Eq. 2.12:
2 0
kT

U U FB
W=
qN d
q

0. 5

(2.12)

with o being permittivity of vacuum, dielectric constant, Nd doping concentration, UFB


flatband potential, k Boltzmann constant, q charge of an electron. It follows that W depends
on the potential difference in the space charge layer Us (between the film formation potential
U and the flatband potential UFB) and the donor density Nd (doping concentration). Using
conventinal TiO2 values of Nd = 1x1026 m-3, r = 85, UFB = 0 V, and U in the range of few V,
for Eq. 2.12 one obtains a thickness (or width) of the space charge layer that is comparable to
the entire thickness of the oxide film. Therefore, if the anodic TiO2 films are not very thick,
they behave like a dielectric, while they are formed. Thus, Eq. 2.10 and 2.11 can be used to
estimate the stresses in the film caused by electrostriction.
The compressive stress of the film due to electrostriction can be experimentally
obtained by measuring changes in the surface stress during the cathodic polarization sweep

24

from the film formation potential to the flat band potential after film formation. By taking the
calculated values of

el
xy

(using ox = 0.4) and the experimentally obtained values of the

stress (both compare very well) and dividing them by the total surface stress of the film
(measured by the beam-bending approach) it was revealed that the contribution of the
compressive stress caused by electrostriction to the total compressive stress did not exceed 24 %, depending on the applied potential. So it can be concluded that electrostriction has only a
limited importance for the stress generation in TiO2 films and according to these reports, it
can be concluded that electrostriction is more important in the case of Al2O3 films than in the
TiO2 case [100].
2.2.5.3 Stresses due to transport of ions

Nelson and Oriani proposed that in addition to the PB ratio, the transport number of
mobile ions in the film also has an influence on the stress generation [94]. They introduced a
model describing essentially two possible situations based on the type of ionic migration
during the oxide growth. When the transport through the oxide is predominantely anionic, the
oxygen ions penetrate into the metal substrate at the metal / oxide interface and compressive
stress may be created. In contrast, when the transport is mainly cationic, vacancies are
produced in the metal substrate at the metal / oxide interface, producing free volume since the
sizes of metal ions are smaller than the metal atoms. Thus, the metal / oxide interface shrinks
and a tensile stress is generated. The following criterion using the Pilling-Bedworth ratio (R)
has been suggested for the stress determination [94]:
t Oc = 1 / R,

(2.13)

Where t Oc is critical transport number of oxygen ions in the film (in the case of TiO2/Ti couple
giving ~ 0.53 for R being 1.9). If the transport number of oxygen ions, to, is lower than t Oc , the
tensile stress is generated and vice versa for the compressive stress. However, this model is
still subject to debate, because the real transport numbers in the TiO2 growth (~0.4 for Ti
cations, ~0.6 for oxygen anions are somewhat close to the critical value of t Oc (~0.53) and the
accuracy of their determination together with variation in the Pilling-Bedworth ratio may
strongly affect the direction of the generated stress and may lead to misleading conclusions.

25

2.3 Ti anodization in fluoride containing electrolytes


While anodization of Ti has been extensively investigated in various electrolytes (see
Ref. [19] for an overview), particularly acidic electrolytes containing Cl- [66,67,84,89] and
Br- ions [88,90-92], only a small number of reports regarding anodization of Ti in HF or Fcontaining electrolytes have been published until this decade [17-22]. In the next section, the
more important findings on Ti anodization in fluoride-containing electrolytes are summarized.

2.3.1 Background
In fluoride electrolytes, deviations from the typical valve metal anodization
electrochemistry occur. This is due to the fact that the formed oxide (Eq. 2.3) is now
significantly chemically attacked dissolved and the formation of soluble fluoride
complexes takes place, according to following equations:
Ti4+ + 6 F- [TiF6]2-

(2.14)

TiO2 + 6 HF [TiF6]2 + 2 H2O + 2 H+

(2.15)

These equations will be used later to explain various effects within this thesis. In the
absence of an electric field, these reactions account for the dissolution of the already formed
anodic TiO2 immersed in the fluoride-containing solution, or etching of the Ti substrate
(under dissolved TiO2) via dissolving Ti4+ species. In the presence of an electric field, an
anodic oxide is formed (Eq. 2.3) and it is simultaneously dissolved (Eq. 2.15), thus both
processes are in dynamic equlibrium. Particularly in view of the chemical dissolution of TiO2,
various considerations should be made. The rates of the reactions (TiO2 dissolution rates) of
course depend on the concentrations of both reactants [103-105]. However, it has been
reported that the kinetics of the reactions depend strongly also on the degree of hydration and
also the field-aided contributions may affect the kinetics, when taking place on anodically
polarized oxide [103].
The most important role of TiO2 (as a passive layer) is the protection of the Ti
substrate. From this point of view, investigations of the processes on Ti substrates in the
presence of fluorides has a fundamental importance for industrial Ti applications, as well as in
e.g. dentistry (Ti dental implants come in contant with fluorides from toothpastes and

26

prophylactic gels, see Ref. [106-108]). In general, flouride anions are extremely unwanted for
these applications. There is though an application in metallurgy, where use of fluoride
containing solutions is highly desired for polishing of Ti, known as pickling [109]. For this
purpose, typically HNO3 / HF aqueous mixtures are used in order to remove the uppermost
part of the Ti substrate and to release stresses originating from scratches, defects etc.
Considering the total amount of available literature on the passive behavior of Ti and
its alloys in different halide-containing electrolytes, comparably small efforts have been
undertaken in the case of fluoride containing electrolytes. Thomas and Nobe compared the
effects of fluoride ion with respect to chloride ion on titanium [110]. They showed that active
titanium corrosion in H2SO4 was largely unaffected by additions of NaC1 (up to 5.8 wt.%). In
their work, a small decrease in the critical current required for passivation was recorded along
with a decrease in the passive current density. Chloride ions therefore were shown to act as a
mild inhibitor in the passive corrosion of titanium, but fluoride anions (at much smaller
concentrations) were very strong activators of both active and passive corrosion.
All in all, fluoride additions have shown to be extremely effective in preventing
passivation by increasing the critical current and also the passive current density, therefore,
reducing the effectiveness of the protective TiO2 film, particulary in an acid environment,
where fluoride anions readily dissolve TiO2. In this regard, the passive behaviour of TiO2 and
its growth in electrolytes containing HF, or fluoride salts have been further studied in a
number of reports by Mandry et al. [111], Caprani et al. [112], Kelly [113], Hepel et al. [103],
Wilhelmsen et al. [104] and recently Cattarin et al. [114,115] by the aid of polarization
techniques and electrochemical impedance spectroscopy (EIS). Caprani et al. investigated the
pre-passive region in acidic fluoride containing electrolytes [112]. Follow-up experiments
done by Kelly were extended into a passive range, up to 10 V, in which dissolution proceeds
through oxide formation [113]. Futhermore, he used rotating disc electrodes and H2SO4
electrolytes containing different amounts of NaF and showed that stirring had an influence on
the peak heights, but not on the current plateau [113]. He further revealed the following
dependence of j = k[HF]1.8 from the dependence of the steady-state current on HF
concentration (the reason for the reaction rate of 1.8 remains, however, unclear). He denoted
that the fluoride concentration (NaF) can be expressed by the HF concentration, since in
concentrated acidic solutions, fluoride is mainly present as undissociated acid. Unfortunatelly,
throughout all afore-mentioned reports, somewhat incompatible dissolution data of Ti in HF
are found, because different backround electrolytes were used (composed of H2SO4, NaOH,
HF, NaF etc.). On the other hand, all of them show that the HF content has a strong influence
27

on the passive anodic current (the higher the HF concentration, the higher the anodic current).
Moreover, the active dissolution was shown to be enhanced by increasing the temperature
[111] and the stirring rate [113].
In general, it seems that a complete description of processes occuring on the Ti anodes
in electrolytes with fluorides seems to be a very complex issue. The smooth relationship
between growth and dissolution rates suggests a continuous modification of the structure of
the oxide depending on temperature and applied potential. Formation of an anodic film
appears to involve competition between the oxide layer formation and the oxide layer
dissolution at its interface with the electrolyte. For this competition, HF concentration, applied
potential, background electrolyte, agitation and temperature play a crucial role. For
anodization at low voltages, where no breakdown occurs, the fluorides play a similar role as
chlorides, because they may dissolve the oxide (Eq. 2.15) at weak spots (defects, cracks) and
further penetrate it leaving nanochannels behind, or even ordered nanopores at certain
conditions (as shown in Fig. 1.3). This pore growth, described in more detail in next section,
is in fact a very specific type of corrosion process, as it leads to re-structuring of the surface
by dissolution. Nevertheless, at comparably higher voltages (several dozens up to hundreds of
volts), breakdown of the oxide film occurs also in fluoride containing electrolytes. The
process of local breakdown ranges from an inconspicuous repassivation to an accelerated
dissolution or (in extreme cases) vaporisation of large parts of the sample [19]. The passive
film represents the dielectric medium of a capacitor, which is short-circuited by the local
breakdown. The electric energy stored on the capacitor heats the breakdown channel and
further accelerates the breakdown. Typically, on the very short scale, complete etching of the
Ti substrate in F-containing electrolyte will take place being enhanced and accelerated by the
high flowing currents.
During chemical TiO2 dissolution, other products are often formed, such as titaniumoxo-flourocomplexes [103]. It should also be mentioned that during the growth of valve metal
oxides in fluoride environments, incorporation of fluorides into an oxide film leading to a
formation of metal fluorides may occur. It has been reported in the case of Ta [116], that TaF5
is formed, however, it seems to be a rather rare case. In the case of Ti, in KF containing
electrolytes at neutral or alkaline pH, various potassium titanates can also be formed [117].

28

2.3.2 First ordered structures


In contrast to porous silicon and alumina, little work was available on porous TiO2
formation by anodization of Titanium until recent years. The very first indication of being
able to produce a self-ordered porous structure was given by Kelly [113], who showed a
comparably higher thickness gain per applied volt (~ 6 nm / V) during anodization of Ti in
NaF containing electrolyte. As it became clear later, this can essentially be ascribed to the fact
that the resulting TiO2 structure was, in fact, porous instead of compact. At the start of this
thesis, except for the early work of Kelly [113], there were only three reports on the growth of
self-organized TiO2 nanopores or nanotubes by anodic etching. They were the following
(chronologically ordered): In 1999, Zwilling and coworkers reported [20] on anodization of Ti
and Ti alloys in chromic acid-HF mixtures. With more clear evidence than Kelly, they
showed that the surface of their substrates was converted by anodization into an ordered
nanoporous structure, as shown in Fig. 1.3. In 2001, Gong and coworkers [118] reported on
the formation of self-organized TiO2 nanotubular layers in aqueous HF electrolytes with the
tube size dependent on the applied anodization voltage. Finally, in 2003, my colleague Radim
Beranek together with his coworkers at LKO investigated the anodization of Ti in H2SO4
electrolytes containing HF [119]. These four reports have in common that the formed
structures have thicknesses in the 200 - 500 nm range and their growth was accompanied by a
strong substrate etching. An example of the self-organized nanotube layers is shown in Fig.
2.11 that was grown according to the Beraneks paper [119].

Gap
Figure 2.11 Self-organized TiO2 nanotube layer grown by anodization of a Ti substrate at 20
V in 0.15 wt. % HF / 1 M H2SO4 for 2 hours (thickness ~ 400 nm, single tube diameter ~ 100
nm, tube wall ~ 10-20 nm).

29

At this point, the terminology currently used in the field of these ordered layers should
be clarified. The morphology of the layers shown in Fig. 1.3 and 2.11 has evidently porous
features. Nevertheless, if one looks at the image shown in Fig. 2.11 closer and more carefully,
it can be seen that there are protrusions in the layer, marked with a white arrow. Therefore,
the objects are not pores, but vertically oriented hollow cylinders with wavy-shaped walls
attached to the Ti substrate. This nanotube shape can be typically observed throughout all the
reports in the field, and that is why these structures adopted, after the first report of Zwilling
et al. [20] beginning with the 2001 Gong et al. paper [118], a new name nanotubes, more
specifically, self-organized anodic TiO2 nanotubes. This is the main difference between the
self-ordered alumina (shown in Fig. 1.1b, 2.3) and titania (shown in Fig. 2.11). The currently
most likely explanation for the nanotube shape instead of the porous shape is given in Chapter
5.
The paper by Beranek et al. [119] also reported an interesting series of experiments
using different HF concentrations yielding: (i) a compact TiO2 layer with a few nonregular
pores (in electrolytes with < 0.05 wt. % HF); (ii) a self-ordered porous layer formed in the HF
concentration window between 0.05 and 0.4 wt. % HF; (iii) uniform etching of the surface
(electropolishing) in solutions containing > 0.4 wt.% HF. Figure 2.12 shows a corresponding
set of SEM images taken from samples anodized from OCP to 20 V.

< 0.05 wt. %

0.1- 0.4 wt. %

> 0. 4 wt. %

50 m
Figure 2.12 SEM images of Ti samples anodized in three different HF concentration ranges
showing formation of a compact TiO2 layers with some minor porous features (a) and a
porous TiO2 layer (b). At higher HF content, the Ti surface is electropolished (c).

Another interesting feature of this anodization approach is that the higher the HF
concentration, the higher is the current density, in accord with the former work of Kelly [113]
in NaF electrolytes. This situation is shown in Figure 2.13 that plots current transients
recorded in 1 M H2SO4 with different additions of HF. In pure background electrolyte
(H2SO4), the typical exponential current decay is observed that is ascribed to the compact

30

oxide layer growth. As described earlier in the text, an addition of HF in the electrolyte results
in a deviation from the classical high-field behaviour in all cases. This current behavior can be
ascribed to different stages in the pore formation process as the same behaviour has been
previously reported for self-organized porous alumina formation [26]. After an initial
exponential decay, the current increases again due to the field-aided dissolution and increased
sample surface/electrolyte contact area. Then, the current reaches a quasi-steady state value
(depending on the HF content). Interestingly enough, it is not only the current density that
increases with the HF concentration, but also the magnitude of current bursts that are apparent
from the curves [119]. While for pure H2SO4 the curve is essentially smooth, for all HFcontaining electrolytes, these current spikes, or, say, current oscillations (as they are quite
periodic), can be observed.

Figure 2.13 Current transients recorded during anodization of Ti at a constant voltage of 20 V


(after a potential ramp from the OCP with sweep rate of 1 V/s) in a) pure 1 M H2SO4, and in 1
M H2SO4 with addition of different amounts of HF b) 0.05 wt. %, c) 0.15 wt. % HF, d) 0.3 wt.
% HF, e) 0.4 wt. % HF (taken from Ref. [119]).

Additionally, as can be seen from Fig. 2.11, there are some periodic wavy ring-like
features on the side walls. One may only speculate about the relation of these features (i.e.
whether the current density oscillations are responsible for the heterogenous ripple
morphology, or whether it is vice versa). Although the exact sequence remains unclear, two
considerations can be put forward that are known from the literature. First, these nanotube
features could be linked directly to an earlier model proposed for the oscillatory behavior in
macroporous silicon formation, suggesting a spatially inhomogeneous current flow [120]. The
31

current bursts are assumed to form semispherical bumps in the oxide, i.e. their lateral
extension is similar to the thickness of the anodic oxide. A second view was given by Ono et
al. [10,121] in the case of porous alumina growth. They observed regular morphological
features on the pore walls and ascribed them to the termination of the uniform film growth
due to interrupted field-distribution (caused by stresses in the films, local film thickening, or
gas evolution). As a consequence, they also observed periodic modulations of the current
density.
To summarize, the most important new finding on Ti anodization in fluoride
electrolytes is that the growth of highly organized titania nanotubular structures can be
achieved. Furthermore, it is evident that in order to grow these layers, the ``right set of
anodization conditions (particularly an appropriate HF concentration and applied potential)
must be used. Furthermore, within the optimized working conditions, it is possible to a certain
extent, as demonstrated by Beranek with the HF content [119] and by Grimes with different
applied potential [118], to modify the thickness and the tube diameter, respectively. It is also
remarkable that this anodization approach leading to highly ordered nanostructures has been
adopted meanwhile (after specific tailoring the anodization conditions) to many other valve
metals including Zr [122,123], Hf [124], Nb [125,126], W [127,128], Ta [129,130] and Al
[131] and their alloys (e.g. Ti6Al7Nb and Ti6Al4V [132], Ti29Nb13Ta4.6Zr [133], TiNb
[134,135], TiZr [136], TiAl [137].

2.4 Properties of TiO2


2.4.1 Structure

The native TiO2 layer on Ti is typically 2 10 nm thick (depending on the environment)


and amorphous (density ~ 3.8 0.1 g / cm3). Otherwise, TiO2 typically crystallizes in three
crystalline modifications: anatase (tetragonal, 3.81 g / cm3), rutile (tetragonal, 4.25 g / cm3)
and brookite (orthorhombic, 4.17 g / cm3), the first two are more common [138].
Determination of the structure can be performed by aid of several techniques, such as most
commonly by X- ray diffractometry, transmission electron microscopy (TEM) and Raman
spectroscopy (RS). The crystalline or amorphous nature of TiO2 varies with processing
techniques and parameters.

32

Typically, the structure of anodic TiO2 films is reported to be either amorphous for low
voltages (ref. [64,139], or crystalline with anatase [140], or rutile structure [139,141].
However, throughout the literature there is unfortunately a lot of discrepancy in the
experimental conditions leading to the TiO2 crystallization (e.g. [67,99,141]) and strong
variations in the result can be found, depending on how the anodization was performed
(particularly significant differences were shown for different potential sweep rates [139]).
From a thermodynamical point of view, crystallization or phase transformations take
place through nucleation and growth processes. The Gibbs free energy difference between
the amorphous and crystalline phases or between two crystalline phases has positive
contributions from surface energy and strain energy and a negative contribution from volume
energy (assuming that no residual stress is present in the material) [142]. Hence, the higher
the strain energy or surface free energy, the larger the critical nucleus size. The strain energy
associated with the nucleation (or formation) is estimated to be about 10 kJ / mol for rutile,
which is much larger than that for anatase, about 0.4 kJ / mol [143]. According to the
thermodynamic data, the chemical free-energy difference, G = GA- GR, between anatase and
rutile is about 6 kJ / mol at any temperature; that is, the difference in the chemical driving
forces between amorphous rutile and amorphous anatase transformations, GAmR GAmA is about 6 kJ / mol, which means that the rutile formation is preferred only in terms

of the chemical free energy. When these phases are formed in the amorphous TiO2 matrix,
however, the driving force including the strain energy for rutile formation becomes smaller,
by approximately 4 kJ / mol, than that for anatase formation. Thus, the rutile formation is
expected to be restrained (or the nucleation rate of rutile is lowered) and the anatase formation
is preferred [143]. Thus, the most common phase transformation in TiO2 is from the anatase
phase to the rutile phase, where rutile is the stable phase. Additionally, crystallization and
subsequent phase transformations will be impeded, if for instance the size of the particles or
thickness of the walls in a porous architecture is comparable, or lower than the critical nucleus
size [32, 144].
Structure of the TiO2 films is a crucial parameter for effectivenes of its applications, often
there is a need to convert an amorphous TiO2 structure into a crystalline one. For instance,
anatase has been preferred in dye-sensitized solar cells [145] and in photocatalysis [146,147],
whereas rutile is mostly used in the area of dielectrics and gas sensors [148]. Crystallization
or phase transformation can be achieved either by thermal annealing [142], or electron-beam
irradiaton [149]. Nevertheless, crystallization and phase transformations have adverse effects

33

on the stability of TiO2 nanoarchitectures, especially when they occur at elevated temperature.
Their high surface area makes them prone to solid-state sintering, which leads to grain growth,
densification, and eventually complete collapse of the structure [144]. Conversely, in special
cases, amorphization of a crystalline structure can be achieved by ion implantation [150], or
by bombarding with heavy particles [151].
There has been a number of reports describing the influence of various
electrochemical parameters on structural (recrystallization), optical (refractive index),
compositional (swelling or dehydration) and semiconductive properties (depletion, defects) of
anodically grown titania films [99,140,141,152-157], also in terms of chemical stability [153]
and impact on the photoelectrochemical response [152]. It is believed that these differences
are related to the different amount of the bound water in the oxide film and coverage of the
oxide film by hydrophilic OH- groups an effect called ``ageing. As a consequence of this
effect, film stability or corrosion resistance will usually be improved during long term
applications [19]. Therefore, ageing is of anodic films is of great practical relevance.

2.4.2 Ionic properties


As demonstrated in Fig. 2.8 for TiO2, all anodic oxide films grow simultaneously at
the oxide / electrolyte and metal / oxide interfaces by outward migration of cations (in this
case Ti4+) and inward migration of O2- / OH- ions, respectively [22,23,158]. In this mode of
film growth, cooperative transport processes of ions are involved. Thus, it is important to
evalute and to understand the factors that are connected with ions in the passive films.
Ionic properties are given by ionic conductivies , transport number of ions t and
defect concentration (intrinsic or impurities). Ideal, stable insulators and semiconductors have
= 0. Real (non-ideal) oxide films are always non-stoichometric due to an excess of metal
ions or a defficiency of oxygen ions ( 0). During the passive film growth, is increasing
with increasing field through the oxide. Moreover, considering the semiconductive nature of
oxide films (next section), foreign ions (such as protons, impurities) can contribute donors,
acceptors or traps in the band gap. Defect concentration (or donor concentration) of passive
films is usually very high, in the range between 1019 to 1021 cm-3 [159].
The transport numbers (sometimes also called transference number) of cations and
anions are dependent upon the film material and have been widely studied by different groups
[68,94,160-162] by aid of markers in the oxide films (radiotracing elements or immobile

34

species) and their depth distribution (Davies & Domeiji[160], Khalil & Leach [68], Nelson &
Oriani [94], Habazaki et al. [161], Wang & Hebert [162]). The transport number of cations
(the ratio of ionic current carried by cations to the total ionic current) is determined using the
ratio of the number of metal ions above the marker plane (e.g. Si [161]) to the number of
metal and marker ions in the total film thickness, based on the results from e.g. Rutherford
backscattering spectroscopy (RBS) analysis. Throughout these reports, a good agreement on
transport numbers has been found for nearly all valve metal oxides. In the case of TiO2 growth,
the transport number of the Ti4+ cation was experimentally estimated to be ~ 0.4 (thus the
transport number of O2- anion is ~ 0.6).
In general, for all valve metals, the outward migration is associated with the fieldassisted dissociation of the respective metal-oxygen bonds and formation of positively
charged species. The migration rates of these species in the anodic films are related to the
energies of the metal-oxygen bonds (relative migration rates are decreasing with an increase
in the metal-oxygen binding energy) [161]. Particularly for the TiO2 case, the observation that
the films are amorphous implies that an ion passing through the film must encounter not a
series of identical barriers, but a series of barriers with a range of activation distance and
heights [22]. These barriers include various defects: grain boundaries, linear dislocations,
vacancies, interstitials and channels.

2.4.3 Electronic properties


Three classes of materials can be defined, based on the energetic distance between
their valence bands and their conduction bands: metals, semiconductors and insulators. Most
of the anodic oxides are semiconductors [146,159]. TiO2 is a ntype semiconductor with the
distance between the bands depending on its structure. This distance is usually characterized
by the so-called band gap energy Eg. Typically, anatase possesses an Eg around 3.2 eV,
rutile possesses an Eg around 3.0 eV. Amorphous TiO2 films have a very non-stoichometric
composition and their bandgap energies may vary [159,163].
Owing to the semiconducting nature, electronic properties of TiO2 can be investigated
by photoelectrochemical methods and by capacitance measurements. For this, it is essential to
consider the energetic situation on the n-type semiconductor vs. electrolyte interface before
and after contacting each other, as shown in Fig. 2.14. This description of semiconductor vs.
electrolyte interface essentially reflects a model describing the p-n junction, the so-called

35

Schottky diode, based on assumptions of having: (i) an ideal, crystalline semiconductor with
energetically discrete, homogenous and completely ionized donor level near the conduction
band, (ii) no surface states and (iii) neglecting the Helmholtz layer in the electrolyte [159].
Upon a contact of both sides, a thermodynamic equilibrium is established between
them (Fig. 2.14b) via an adjustment of the Fermi level of the semiconductor and the red/ox
potential of the electrolyte. As a result of the electron transfer across the interface, so-called
space charge layer, W, is formed with upward bend bending typical for n-type semiconductors
(p-type semiconductors have downward bending). W can be expressed as in Eq. 2.12.
The energy difference between the conduction band edge in the bulk of the
semiconductor and the band edge at the surface is called the Schottky barrier, with an energy
difference (or ``height) equal to Us (see Eq. 2.12). Parameter U depends on whether an
anodic potential is applied on the electrode (in this case U=Uan). If there is no external
polarization, than U = Uredox. The flatband potential Ufb and the positions of the conduction
and valence band, Ucb and Uvb, are included and marked in the Pourbaix diagram in Fig. 2.6. It
is also worthy to mention that W has a strong implication on the breakdown mechanism, that
is, for small W tunneling breakdown occurs, for large W avalanche breakdown occurs [159].
vac. level

conduction
band
E
EC

E
EC

redox

redox

Energy

valence
band

(a)

(b)
solid

liquid

solid

liquid

Figure 2.14 Energetic situation at the n-type semiconductor/environment interface before (a)

and after (b) contact with the electrolyte.


For a description of the space charge layer capacitance - Csc, a parallel plate
condensator model can be used (C ~ A/d, where A is the plate surface and d is distance
between the plates), replacing d by W yields the so-called Mott-Schottky equation [159]:
36

1
C sc 2

kT
U U FB

2
q
0 qN d A
2

0.5

(2.16)

Thus, by plotting Csc-2 vs. electrode potential U, the flat band potential UFB and the
doping concentration Nd can be evaluated (assuming that is known). Usually,
electrochemical impedance spectroscopy (EIS), a technique based on an ac-perturbation of the
electrode, is used to determine the space charge layer capacitance (at sufficiently high
frequencies) by aid of different model circuits combining capacitance of the electrode,
polarization resistance and resistance of the electrolyte (sometimes also back-contact
resistance, Helmholtz layer resistance etc.)
The band gap energy of anatase TiO2 (3.2 eV) corresponds to a wavelength of
approximately 390 nm, i.e. electrons from the TiO2 valence band cannot be excited to the
conduction band by light that has a longer wavelength than 390 nm [146]. Thus, only upon
UV light irradiation, electrons will be excited from the valence band to the conduction band.
The resulting charge carriers (electron-hole pairs) are separated in the field of the Schottky
barrier and photocurrent is generated (assuming a closed electrical circuit). In line with the
Grtner model, the generated photocurrent, iph, can be separated into contributions from either
the space charge layer (depletion layer), or in the diffusion layer (deeper in the bulk
semiconductor) [164]. However, for non-ideal semiconductors, including TiO2, the charge
carriers may recombine there and thus may not contribute significantly to the whole
photocurrent. Thus the photocurrent iph will depend on the rate of generation of electron hole
pairs, as well as on their recombination and their reaction at the oxideelectrolyte interface.
In the most basic approach, the generated photocurrent (neglecting the diffusion layer,
assuming single doping species and no recombination) may be expressed as:

i ph = qW ,

(2.17)

where is the incident light intensity and is the light absorption coefficient (that is related
to the photon energy h (and thereby to the wavelength of the light) and may be expressed by
following equation:

(h E )

=A

(2.18)

where A is a constant depending on the electronic structure of the material. For allowed direct
transitions, n = 0.5, for allowed indirect transitions, n = 2 [146,159].
The electronic properties of TiO2, such as the quantum yield or doping concentration,
depend strongly on the formation conditions. Vacancies, interstitials and surface states yield

37

additional electronic levels in the band-gap as well. To additionally tailor the photoresponse
of TiO2 (to introduce new sub-bands with lower Eg), doping is often performed by e.g. N
[165,166], or C [167]. Both dopants help to extend the TiO2 photoresponse into the visible
light region.

2.5 Functional applications of TiO2


Titanium dioxide is a highly functional material that has found tremendous
technological interest during the past 20 years. In addition to its specific properties that were
mentioned in the previous section, its use is highly desirable due to high chemical and
mechanical stability, non-toxic behavior and low-cost preparation. In this section, the
applications of anodic films and nanostructured films will be outlined.

2.5.1 Anodic TiO2 films


Titanium is anodized for both decorative purposes (to provide a decorative finish of
the surfaces) and functional purposes (e.g. to protect metal from atmospheric corrosion, to
increase osseointegration ability of Ti implants with bones, to reduce the friction on sliding
surfaces, to prevent galling and to provide thermal control).
Ti and its alloys have been intensively investigated for biomedical purposes, namely
as a dental, or hip implant material, due to their superior strength and biocompatibility [168].
However, for a good implant accommodation in the body, the presence of an interface layer
between the metal implant surface and the hydroxyapatite coating of the bone is required to
osseointegrate with bones, because native TiO2 coatings have no ability to form a strong bond
with bony tissue due to their low surface energy [168]. Therefore, these substrates are
anodized to have either planar, or mesoporous TiO2 coverage (achieved by anodization below,
or above the breakdown potential, respectively, as decribed earlier in the text). Particularly,
the mesoporous structures (an example of which is shown Fig. 2.10) that have porous features
in the range of a few nm up to a few hundreds of nm have been shown to strongly enhance the
osseointegration ability of implants [84,85,168] and thus lead to a high degree of adoption of
the implant into the body.
Decorative anodizing can be used to color code tools and hardware according to size
and type and for jewellery and hobby items. Anodizing in strongly alkaline solutions can

38

produce coatings up to several microns in thickness, which can be used to prevent galling and
increase corrosion resistance [24]. On the other hand, anodizing in acidic, neutral and mildly
basic solutions produces a very thin (0 to 200 nm), nonporous oxide. The oxide is transparent,
but parts have vivid rainbow-like colors due to interference coloring [22]. White light falling
on the oxide is partially reflected and partially transmitted and refracted in the oxide film. The
light that reaches the metal / oxide surface is mostly reflected back into the oxide. Several
reflections may take place. A phase shift occurs during this process along with multiple
reflections. The degree of absorption and number of reflections depends on the thickness of
the film. The light that was initially reflected from the oxide surface interferes with the light
that has traveled through the oxide and has been reflected off the metal surface. Depending on
the thickness of the oxide, certain wavelengths (colors) will be in-phase and enhanced while
other wavelengths will be out of phase and dampened. Hence, the observed color is mainly
determined by the oxide thickness. As explained earlier in this Chapter, the oxide thickness is
primarily voltage controlled. At any given voltage the oxide film grows to a specific thickness
and then stops thickening. However, other factors such as material, pretreatment, anodizing
solution chemistry and temperature, load size, anode : cathode ratio, anodizing time, and tank
configuration affect the color of the anodized piece, making it somewhat difficult to predict
and control the resultant color [22].
The color of anodized titanium is brilliant with an iridescent quality. Examples of
coloured TiO2 film on Ti are shown in Fig. 2.15. The color will not fade, or wear off since it
is produced by the above mentioned physical phenomenon of interference at the oxide and the
metal surface. However, any coating placed on top of the oxide, such as finger prints, will
affect the color.
Table 2.2 lists the color spectrum of anodized titanium along with the applied voltage
and calculated oxide thickness (from the refractive index). The growth values (nm per V)
seem to be rather high, but they have only demonstrative character. In the case of nanoporous
or nanopatterned TiO2 surfaces, the color is not bright. Instead, it is rather diffuse in color due
to the light diffraction and scattering due to the nanostructure (as discussed by e.g. Fujishima
et al. [169]).

39

Figure 2.15 Optical images of TiO2 grown on a Ti metal substrate during 1 minute of
anodization at 20 V (a) and at 40 V (b) in 1 M H2SO4.
Table 2.2 Color as a variable of anodic TiO2 thickness (taken from Ref. [24])
Applied voltage (V)
2
6
10
15
20
25
30
35
40
45
50
55
60
65
75

Color
Silver
Light brown
Golden brown
Purple blue
Dark blue
Sky blue
Pale blue
Steel blue
Light olive
Greenish yellow
Lemon yellow
Golden
Pink
Light purple
Blue

Film thickness ()
< 100
240
360
490
580
700
820
930
1040
1150
1250
1320
1410
1560
1770

2.5.2 Nanostructured TiO2


A major part of functional TiO2 applications is related to its semiconducting properties
described earlier in the text and using nanosize TiO2 features due to their large surface area.
UV light irradiation (whose energy exceeds Eg), generates electron / hole couples and the key
use of this charge carrier generation can be either in solar cells [145], or particularly in
photocatalysis [146,147,159], due to the very strong oxidizing power and suitable band
positions of TiO2. The history of photocatalysis dates back to 1972, when Fujishima and
Honda reported for the first time on light-induced water splitting on TiO2 surfaces [147]. This
finding gave rise to novel approaches in heterogenous catalysis using the TiO2 ability to
40

decompose unwanted organic compounds (including fat, germs, etc.) applicable in waste
water treatments and self-cleaning applications [170] as well as for anti-fogging applications
[171] (e.g. coating for windshields). The principle of the photocatalytic decomposition is
demonstrated for the TiO2 nanoparticle surface in Fig. 2.16.

Backside
O 2O2

TiO2

Electrolyte

Conduction
band
e-

h >
Eg

h+

H2O
OH

CxHy
CO2 +H2O

Valence
band

Figure 2.16 Principle of the heterogenous photocatalysis on the TiO2 nanoparticle surface
(according to Ref. [146]).

As a consequence of the charge carrier generation (upon irradiation), the carriers either
recombine inside the particle, or migrate to its surface, where they can react with adsorbed
molecules. In aqueous solutions positively charged valence band holes typically form OH
radicals, while electrons in the conduction band mainly reduce dissolved molecular oxygen to
super-oxide O2- ions. Various organic molecules (CxHx) present in the solution may react with
these oxidizing agents inducing their oxidative degradation into simple inorganic compounds
including carbon dioxide and water [170]. In order to achieve a maximal decomposition
efficiency, adequate band edge positions, rapid charge separation, high quantum yield and a
large area of the catalyst are desired. Therefore, for pollution control applications, selfcleaning applications, and antifogging applications, nanoparticulate, or nanoporous films (or
nanoparticulate suspensions) are mainly used.
To form TiO2 nanoparticles, the sol-gel methods (based on mixing Ti-alkoxides with
alcohols [163,172,173], or direct oxidation of TiCl4 are mainly used [174]. An example of a
commercially available nanopowder TiO2 P25 produced by Degussa is shown in Fig. 2.17.
The widespread route to produce nanoporous TiO2 networks on surfaces is by classical
powder processing (using e.g. nanoscale TiO2 particle suspensions followed by sintering).

41

The as-formed nanoparticulate layers have shown to be highly efficient in sensing


applications, e.g. for oxygen sensing [148]. The interaction of a gas with a semiconductor
electrode is typically a surface phenomenon, and hence the use of porous structures is
desirable for these applications owing to their large surface areas.

100 nm
Figure 2.17 Degussa P25 TiO2 nano-powder

TiO2 nanoparticles have been also used in the field of solar energy conversion
namely in dye-sensitized nanocrystalline solar cells [145] (often called Grtzel cell after its
inventor). The principle of operation and energy level scheme of the solar cell is shown in
Figure 2.18. Such cells consist typically of a 5 - 10 m thick layer of TiO2 nanoparticles with
suitable organic dyes attached onto their surface. The dye molecules (often called
panchromatic sensitizers) act as acceptors of the visible-light photons. After absorption of
photon in the dye, an injection of the electron into the conduction band of the TiO2 takes place
that leads to generation of current on the charge collector (typically a thin Pt layer). The dye
molecule is regenerated by the redox system present in the electrolyte (typically a mixture of
I2 and I-), which itself is regenerated at the counter electrode.
Another application of porous TiO2 structures could be for metal mirrors, photonic
crystals, or waveguides [175], due to the high refractive index of TiO2 (2.7, i.e. about 60%
higher than porous alumina, [176]).

42

Figure 2.18 Schematic of operation of the dye-sensitized electrochemical photovoltaic cell.


The photoanode, made of a mesoporous dye-sensitized semiconductor, receives electrons
from the photo-excited dye which is thereby oxidized, and which in turn oxidizes the mediator,
a redox species dissolved in the electrolyte. The mediator is regenerated by reduction at the
cathode by the electrons circulated through the external circuit. The open-circuit voltage of
the solar cell corresponds to the difference between the redox potential of the mediator and
the Fermi level of the nanocrystalline film indicated with a dashed line. Courtesy of EPFLLPI.

2.6 Other methods for TiO2 nanotube synthesis


Since the discovery of carbon nanotubes in 1991 [177] continuously increasing
research interest in one dimensional (1D) nanomaterials has been established. Nowadays, not
only carbon materials are widely studied, but also a variety of metals and oxides, such as TiO2
[178], ZnO [179], etc. The inorganic nanotubes, in particular the TiO2 ones, are of a great
potential for various technological applications, due to their high surface to volume ratio,
enhanced electronic properties (in comparison with nanoparticles), well-defined structures
and the possibility to precisely tailor their dimensions on the nanoscale. In the case of TiO2,
several studies indicated that nanotubes have improved performance in photocatalysis
[180,181] and photovoltaics [182,183] compared to colloidal or nanoparticulate forms of TiO2.
Up to now, suspensions, bundles and arrays of rather disordered TiO2 nanotubes have
been produced by a variety of different methods including hydrothermal synthesis [178, 184]
based on mixing and heating of TiO2 nanoparticles with NaOH), deposition into porous

43

alumina templates [185-187], and the seeded-growth mechanism [188]. While each of these
methods has certain advantages (large scale production, variations of TiO2 structure), none are
capable of yielding dense arrays of uniform nanotubes aligned on substrates for easy device
implementation and for precise control of the nanotube dimensions.

44

3. METHODS
3.1 Sample preparation
All experiments done in this work were carried out on Titanium specimens in the form
of thin foils (thickness 100 m), or thick sheets (thickness 2 mm), if not denoted otherwise
with purity of 99.6% (composition Al 300, Ca 20, Cr 50, Cu 5, Fe 1500, Mn 100, Ni 50, Si
300, Sn 200, C 300, H 60, N 150, O 2000 ppm) manufactured by Goodfellow (Cambridge,
England.) The foils were used as received, without any surface pretreatment. The sheets were
used mainly for the surface analyses (XPS), which requires flat and homogenous surfaces.
Therefore, for XPS the sheets were mechanically ground under water flux with 400, 800 and
1200 SiC grinding paper (Struers, Denmark). During grinding, chemical etching for 30
seconds was carried out several times to remove a deformation layer using a mixture of 2 ml
of 40% HF and 5 ml of 30% H2O2 in 100 ml deionized water (DI water). Afterwards,
polishing to a mirror-like surface with a diamond paste of 6, 3 and 1 m was performed.
Between these steps, ultrasonic cleaning was done in ethanol to remove residues of a mixture
of Ti and the polishing media.
Prior to all the experiments, both types of Titanium were thoroughly degreased by
sonication in acetone, isopropanol and methanol, followed by rinsing with DI water and
drying in a nitrogen stream.

3.2 Synthesis of TiO2 films


The electrochemical set-up used in this thesis is shown in Fig. 3.1. It consists of a
conventional three-electrode configuration with platinum mesh as a counter electrode. The
reference electrode was a Haber-Luggin capillary with a Ag / AgCl (1 M KCl) electrode with
the glass tip (filled typically with 1 M H2SO4 to prevent leakage of Cl- into electrolytes). The
Ti working electrode was pressed against an O-ring (rubber, Viton) of an electrochemical cell
(made from polymethyl-methacrylate resistent to the fluoride etching) with 1 cm2 exposed to
the electrolyte. The Pt electrode was placed parallel to the Ti electrode (~ 2 cm) for a good
current distribution. The reference electrode was adjacent to the Ti electrode (~ 1.2 cm) to
minimize the IR drop due to the electrolyte resistance.

45

The potential was applied between the working and the reference electrode using a
high-voltage potentiostat Jaissle IMP 88 and the resulting anodization current between the
working and the counter electrode was measured by a Keithley digital multimeter. Both
devices were interfaced to a PC for data acquisition through a digital Prodis 1/16 TI interface.
The electrolytes were prepared from reagent grade chemicals, purchased from SigmaAldrich and Merck and DI water (conductivity ~ 0.05 S / cm). If not denoted otherwise, the
electrolyte volume of 200 cm3 was used for the experiments. This was large enough to avoid
significant drifts in the electrolyte composition during the experiments.

1 - PE cell
2 - Rubber ring
3 - Ti probe
4 - Cu contact
5 - Pt mesh
6 - Ag/AgCl electrode
7 - Glass tip with frit
8 - PC
9 - Potentiostat

8
6

7
5

234
1

Figure 3.1 Scheme of the electrochemical setup for Ti anodization

For growing the nanotubular layers, potentiodynamic and potentiostatic regimes were
used. The electrochemical treatment consisted of a potential sweep with different sweep rates
from either the open - circuit potential (OCP), or 0 V, to a different end potential, followed by
holding these end potentials for a certain time. After the experiments, the samples were rinsed
with DI water and dried in a nitrogen stream. If not denoted otherwise, all experiments were
performed without electrolyte stirring. Some of the experiments were performed at different

46

temperatures using a Julabo F10 thermostat (Julabo, Germany) via a rubber-coated coil
directly immersed into the electrolytes.
In order to convert the amorphous TiO2 structure into a crystalline one, selected
samples were annealed at different temperatues (typically 450 C) for 3 h in air by using a
Jetfirst Rapid Thermoannealer (RTA, Jipelec, France) with heating and cooling rates of
30C/min.

3.3 Surface analyses


3.3.1 Electron microscopy
Scanning electron microscopy (SEM) was employed for the morphological
characterization of anodized samples. No charging effects were observed, therefore no
additional coatings (e.g. Au) were required. Two microscopes were used in this work: a
Hitachi FE-SEM S4800 equipped with a cold field-emission gun and a LEO 1350 SEM
equipped with W-triode filament. The accelerating voltages used for imaging were in the
range between 10 and 20 keV and the working distance was in the range between 7 and 12
mm. All cross-sectional images in this thesis were taken from cracked layers after mechanical
bending, cutting-off, or scratching the samples (with a knife). An energy dispersive X-ray
analyser (EDX) fitted to the SEM chamber was used for chemical analyses of the samples.
Transmission electron microscopy (TEM) was employed for the morphological and
structural characterization of the nanotubular layers. A PHILIPS CM 30 T/STEM (point
resolution 2.3 ) and a high-resolution TEM PHILIPS CM 300 UT (point resolution 1.75 )
were used. The acceleration voltage of electrons was 100 and 300 keV, respectively.
Nanotubes were mechanically scrapped off, rinsed with ethanol and deposited onto a Cu grid.

3.3.2 X-ray diffractometry


X-ray diffractometry (XRD) was employed for the structural characterization of the anodized
samples using an X-ray diffractometer (Phillips Xpert-MPD PW3040) with Cu K radiation
and wavelength 1.54056 . Table 3.1 provides an overview of typical XRD reflections for
TiO2 (anatase and rutile) as well as for Ti (in agreement with International Center Diffraction
Data ICDD).

47

Table 3.1 XRD data for anatase, rutile and Ti (taken from Ref. [189])

Anatase

Rutile

I int.n

I int.n

3.5163

25.308

100.0

3.2477

27.440

100.0

2.4307

36.950

6.5

2.4875

36.078

46.0

2.3786

37.790

20.3

2.2965

39.196

7.1

2.3322

38.571

7.4

2.1873

41.239

18.0

1.8921

48.046

28.2

2.0541

44.049

6.5

53.884

18.1

1.6874

54.323

56.0

1.6662

55.071

17.8

1.6239

56.634

16.5

1.4931

62.115

3.1

1.4795

62.750

7.8

1.4808

62.690

13.9

1.4524

64.057

8.1

1.3642

68.754

6.2

1.3598

69.010

19.6

1.3379

70.302

6.8

1.3464

69.795

9.8

1.2646

75.051

10.5

1.3038

72.425

1.0

1.2505

76.049

9.8

1.2437

76.534

2.2

1.2005

79.827

1.1

1.7001

Titanium
H

I int.n

2.5548

35.096

24.8

2.3430

38.387

25.5

2.2431

40.169

100.0

1.7268

52.985

13.8

1.4750

62.963

15.5

1.3327

70.621

15.9

1.2774

74.172

2.3

1.2482

76.208

16.8

1.2324

77.367

12.0

48

3.3.3 X-ray photoelectron spectroscopy


X-ray photoelectron spectroscopy (XPS) was employed to identify the chemical
composition of the anodized samples using a PHI 5600 XPS spectrometer and Al K
monochromatic radiation (1486.6 eV; 300 W) as the exciting source. Sputter depth profiles
were were acquired using 3.2 kV Ar+ sputtering operating at 20 mA. The thickness of the
layers was evaluated at a 50% drop in the oxygen signal using calibration against a SiO2
standard layer. The following peaks were used for the detection of elements: the Ti peak at
459 eV, the O peak at 529 eV, the F peak at 685 eV, the C peak at 284.8 eV, the S peak at 165
eV, the P peak at 133 eV, the Cl peak at 199 eV and the N peak at 396 and 400 eV.

3.4 Photoelectrochemistry
The photocurrent spectra were recorded in 0.1 M Na2SO4 at 500 mV (vs. Ag/AgCl
electrode), using a Jaissle 1030T potentiostat (the same electrode setup as for anodization)
with 5 or 10 nm steps in the range of 700-250 nm using an Oriel 6356 150 W Xe-arc lamp, an
Oriel Cornerstone 7400 1/8 m monochromator, a lock-in amplifier / chopper (fchop = 30 Hz)
and an electrochemical cell with a quartz window.

49

4. RESULTS
4.1 Polarization curves in the fluoride containing solutions
In order to preliminary characterize the electrochemical behavior of Ti under various
conditions, a series of experiments were performed on mirror-like polished flat Ti samples
using anodic polarization in fluoride-free and fluoride containing electrolytes. Fig. 4.1 shows
polarization curves recorded in acidic electrolytes consisting of (a) 1 M H2SO4, H3PO4, HClO4
and (b) the same electrolytes with 0.16 M HF additions. Very different behaviour can be
observed for both types of electrolytes. In the HF-free electrolytes (Fig. 4.1a), Ti shows
spontaneous passivation, i.e. the native TiO2 oxide present on the surface remains present
after immersion of the sample. This is reflected by a comparably positive open circuit
potential of approximately 150 mV (vs. Ag / AgCl) and the absence of an active / passive
transition. With an applied anodic potential, the TiO2 layer additionally grows due to the
field-effect described in Chapter 2, resulting in thicker anodic oxide film according to Eq. 2.3.
For Ti samples immersed into HF-containing electrolytes (Fig. 4.1b), a strong
deviation in the current behavior occurs. The native oxide layers are dissolved within seconds
to minutes according to the Eq. 2.15. This results in an oxide free surface in HF-containing
electrolytes that shows an open circuit potential of approximately -800 mV. Therefore all
samples anodized in HF-containing electrolytes show around -400 mV an active / passive
transition (approximately in line with the Pourbaix diagram in Fig. 2.6 for acidic electrolyte).
Fig. 4.1c shows polarization curves recorded in 1 M H2SO4 with four different HF
concentrations. With an increase in the HF concentration a higher active peak current density
is obtained as well as higher passive current densities, in line with previous reports
[20,21,103,104,109-113]. Even a low amount of HF (0.05 M) strongly changes the situation
on the Ti / TiO2 / electrolyte interface. In other words, 0.05 M of HF in the electrolyte is able
to dissolve the native oxide layer, to attack the titanium underneath and to inhibit the oxide
formation. For lower HF concentrations (0.01 M) in the electrolyte, however, the situation is
somewhere in between passivity and active dissolution.

50

Figure 4.1 Polarization curves recorded with polished Ti substrates in acids without (a) and (b)
with 0.16 M HF and in 1 M H2SO4 with different HF concentrations (c); sweep rate 5 mV / s.

Apart from the electrochemical behavior, interesting compositional and morphological


differences can be revealed between the samples anodized in HF-free and HF-containing
electrolytes. A set of a scanning electron microscope (SEM) images shown in Fig. 4.2 were
from the substrates anodized up to 10 V in the electrolytes used in Fig. 4.1. These images
exhibit the surfaces of the samples anodized in H2SO4 (Fig. 4.2a), H3PO4 (Fig. 4.2b) and
HClO4 (Fig. 4.2c) in the absence (left side) and presence of fluorides (right side). The insets
show lower magnification images of the samples. For the pure acids, despite a relative
flatness and homogeneity of the Ti substrates on the microscale (except for clearly
distinguishable grains and their boundaries), the morphology of the resulting layers on the
nanoscale is not as smooth as one would expect, the oxide layers exhibit a flaky morphology,
but otherwise there are no significant pits or pores on the surface. The probable reason for this

51

heterogenous morphology is surface stresses that are generated due to an oxide expansion
(described in Chapter 2).
In comparison, images taken from the samples anodized in the HF-containing
electrolytes show strong morphological differences and features significantly different for
each background electrolyte. Remarkable is the presence of the nano-pillars for the perchloric
acid case (Fig. 4.2c right).
These findings indicate different adsorption characteristics of the different anions on
the Ti substrate. Later in the text, it will be shown that selection of the background electrolyte
(respectively the background anion) has a large influence on the morphology of the forming
nanostructures.

Figure 4.2 SEM images of Ti samples anodized in 1 M H2SO4 (a), H3PO4 (b) and
HClO4 (c) without (left column) and with 0.16 M HF (right column). The insets show low
magnification of the layers.

52

4.2 Growth of the self-organized nanotube layers in acidic


electrolytes
As shown previously in Chapter 2, the first generation of self-organized TiO2
nanotube layers was reported using anodization of Ti in acidic HF solutions [20,21,118,119].
Similar experiments were also carried out within the present work in H2SO4 / HF mixtures at
20 V for different anodization periods. Fig. 4.3 shows three SEM images of a sample
anodized for 2 hours in 1 M H2SO4 / 0.16 M HF in different views on the sample. By
comparing the layer thicknesses, gained from the cross-sectional views, determination of the
growth kinetics of the nanotube layers with anodization time is possible. Data from the time
vs. thickness measurements are presented in Fig. 4.3d. A linear increase of the thickness is
observed initially, and later the gain in thickness becomes smaller.

Figure 4.3 SEM images of Ti sample anodized in 1 M H2SO4 / 0.16 M HF at 20 V for 2 hours
taken from the side (a), top (b) and bottom (c) of the nanotube layer; (d) dependence of the
layer thickness on anodization time.

53

As marked in Fig. 4.3a, the top-views (Fig. 4.3b) and the bottom-views (Fig. 4.3c)
revealed other morphological features of the layers. Every single tube is open at the top (the
inner tube diameter is approximately 100 nm) and closed at the bottom by a solid TiO2 layer.
The side-view of the tube layer (Fig. 4.3a) reflects an uneven morphology. Obviously, the
wall thickness and the inner tube diameter significantly vary along the tube wall. The
thickness of the tube walls ranges from a few nm in the uppermost part and may reach 20 nm
at the bottom part. The inner tube diameter adopts the opposite situation. At the bottom, the
tubes are closely packed together (Fig. 4.3c). These features are common for all the tube
layers presented in this work and will be discussed later in Chapter 5. However, the thinner
the tube layers, the larger the differences in wall thickness and tube diameter.
As explained in Chapter 2 and as given by an example of polarization curves and SEM
images of anodized Ti samples, the oxide dissolution occurs in the presence of fluoride anions
via the formation of a soluble hexafluorotitanium complex - [TiF6]2- (see Eqs. 2.14 and 2.15).
In fact, during the anodization in acidic solution (such as 1 M H2SO4 / 0.16 M HF), apart from
the field-aided TiO2 dissolution reactions that contribute to the current (see Fig. 2.13, in
addition to the field-aided oxidation in Eq. 2.3), also the chemical dissolution of TiO2 is to a
large extent responsible for the fact that only a small amount of TiO2 is left on the surface
after anodization (having the nanotube shape). Additional experiments using extensive
anodization period (several dozens of hours) have revealed that the thickness of the nanotube
layer does not increase further after a few hours, as shown in Fig. 4.3d. There is also no
significant thickness variation, when comparing structures formed either in a large volume of
electrolytes (1 liter) to maintain a constant amount of HF for the whole anodization time), or
small volumes (e.g. 50 ml). In the first case, anodization can be carried out as long as there is
available Ti substrate. In the latter case, the growth of the nanotube layer is simply terminated,
or significantly slowed down as soon as there is not a high enough HF concentration.
A measure of the speed of the etching front moving into the Ti substrate can be
obtained by comparing the original thickness of the Ti substrates and the thickness of the
anodized area from the cross-sectioned samples after anodization. A set of these experiments
has been carried out for samples that were anodized in 0.2 M HF at 20 V for different times.
Figure 4.4 shows a loss of the Ti substrate thickness as a function of anodization time. As
evident, considerable loss of the material takes place, e.g. after 3 days of anodization, nearly
0.5 mm of the substrate is etched away. Thus, when anodizing very thin Ti foil (100 m), the
whole thickness will be etched (the Ti volume on the anodized area will be consumed) within
less than 17 hours of anodization. That is, together with the limiting thickness of the nanotube
54

layers, a factor that significantly affects the application of these nanostructures, particularly
when considering that for some applications, only a thin Ti layer sputtered on a conductive
substrate is provided. During the anodization process used for construction of Fig. 4.4, the
anodic current density was approx. 5 mA / cm2, which corresponds to approximately 6 m /
hour of the Ti metal removal according to Eq. 2.8 (i.e. 462 m in 72 hours - a value that
correlates very well with the thickness loss measurement).

2.00
Ti thickness / mm

1.75

1.50
Ti substrate

1.25

nanotube layer
(anodized area)
1.00
0

12 18 24
48
72
Anodization time / hours

Figure 4.4 A loss of Ti thickness underneath the growing nanotube layer during anodization
in 0.2 M HF electrolyte at 20V.

4.3 Growth of the self-organized nanotube layers in neutral


aqueous electrolytes
It has been described in Chapter 2 that the morphology and dimensions of selforganized porous alumina are strongly influenced by a range of parameters, such as
electrolyte composition and temperature, anodization potential, anodization duration and
surface pretreatment [1-13]. Furthermore, within a particular set of conditions, there is a
degree of freedom that allows, to a certain extent, to change for example the anodic potential,
but to still keep the system in the self-organization regime [13]. It can be assumed that these

55

freedoms will be allowed also in the case of TiO2 nanotube formation. Therefore the
following section deals with tailoring the morphology of self-organized TiO2 nanotubular
layers in aqueous electrolytes and seeks the parameters that lead to self-organization and
different morphologies simultaneously.
Within this work, the anodization experiments were carried out by holding the
potential constant after a potential ramp from the open-circuit potential (typically around 0
mV vs. Ag / AgCl electrode). Preliminary experiments showed that best results in view of the
resolution, reproducibility and homogeneity of the nanotube layers (such as shown in Fig. 4.3)
were obtained within the frame of the conditions used here at the potential of 20 V.

4.3.1 Anodization in Na2SO4 / NaF electrolytes


In order to quantify the pH effect on the chemical dissolution of TiO2, a set of etching
experiments was performed with solutions of different pH and for different times. XPS depth
profiles were acquired afterwards to determine the remaining thickness. Mirror-like polished
Ti specimens were anodically oxidized in 1 M H2SO4 at 20 V to grow ~ 50 nm thick compact
TiO2 layer. Afterwards, these samples were immersed in the etching solution. The
composition of the etchant was chosen to simulate anodizing conditions, as much as possible.
Therefore 1 M Na2SO4 / 0.14 M NaF / x H2SO4 mixtures were used (that lead to selforganized TiO2 nanotube layer growth, as shown later) and the solution pH could be adjusted
by changing the amount of concentrated H2SO4. For the sputter profiles, the initial oxide
thickness (50 nm) was taken as 100% and all other values were compared with this value. Fig.
4.5 gives the dissolution rate of anodically grown TiO2 versus the pH. A strong influence of
the pH on the chemical dissolution rate of TiO2 can be seen. For example, the highest
dissolution rate is observed in case of the solution with pH = 2 with a rate of approximately
38 nm.min-1, whereas at pH = 5 the dissolution rate is drastically smaller - less than 1.5
nm.min-1. Therefore it is evident that for higher pH values the chemical etching of TiO2 is
drastically slower. This finding was therefore considered to be a possible key reason for
limited nanotube layer thickness grown in acidic electrolytes and a key element, when
considering the growth of thicker self-organized TiO2 nanotube layer.

56

Figure 4.5 Thickness of the TiO2 anodic layer acquired from XPS sputter profiles as a
function of dissolution time in 1 M Na2SO4 / 0.14 M NaF / X H2SO4 solutions with different
pH. Small additions of concentrated H2SO4 were used to tune the pH (0 ml for ph 7; 20 l for
pH 5.4; 300l for pH 4; 3ml for pH 2.2). Example of an XPS profile is shown as an inset for
the non-etched TiO2.

Based on these findings, a new idea was born considering:


(i)

acidity is needed to grow nanotubes (at the tube bottom),

(ii)

neutral conditions would reduce the etching of the tube walls,

i.e. a desired pH profile along the tube would look like this:

Neutrality

Acidity
Figure 4.6 Outline of the pH distribution
along the nanotube that could diminish
chemical dissolution of TiO2 in the walls.

TiO2
Ti

57

The key idea to achieve this is the fact that during anodization permanent acidity is
produced according to Eq. 2.3 (Ti + 2 H2O TiO2 + 4 H+ + 4 e-). Therefore by tailoring the
anodization, a situation in Fig. 4.6 could be created that would lead to both i) tube growth and
ii) protected walls.
Therefore, we started using electrolytes that contained fluoride anions, but with higher
pH (slightly acidic, or neutral). A set of experiments was performed at 20 V in 1 M Na2SO4
with different NaF concentrations for 2 hours, after sweeping from the open-circuit potential
(OCP) to 20 V (vs. Ag / AgCl) with a sweep rate of 0.1 V / s. SEM characterization was
performed after the experiments, as shown in Fig. 4.7.

b)

a)

450

280

500 nm

500 nm
d)

c)
1950

500 nm

2400

500 nm

Figure 4.7 SEM images of the TiO2 layers with porous/nanotubular features. Ti samples were
anodized in 1 M Na2SO4 with different NaF concentrations: a) 0.025 M, b) 0.08 M, c) 0.14 M
and d) 0.27 M for 2 hours after sweeping from 0 to 20 V (vs. Ag / AgCl) with a sweep rate of
0.1 V / s. The insets show cross-sectional views.

A type of nanotubular layer is grown in these electrolytes, with a relatively strong


variation in the surface morphology. This is even more pronounced, when looking from the
side. Anodization in 1 M Na2SO4 electrolyte containing 0.025 M and 0.08 M NaF (Fig. 4.7a
and b, respectively) results in a relatively heterogenous hazy structure with apparent porous

58

features. The self-organization proceeds neither uniformly, nor completely across the whole
surface. Approximately half of the surface area is covered with islands that are preferentially
etched to a greater depth. Additionally there are cloudy precipitates on the rest of the anodized
area (XPS measurements on the precipitated layer composition showed that the insoluble
layer consist of Ti-, O- and F-species). Anodization in 1 M Na2SO4 electrolytes containing
0.14 M NaF results in more homogenous structures with distinguishable nanotubular
appearance from the side (thickness of the layer ~ 2 m). Anodization in 1 M Na2SO4
electrolytes with almost twice as high concentration (0.27 M NaF) results in even thicker
nanotube layers (~ 2.4 m), but the surface morphology is not as good as in the 0.14 M NaF
case, probably due to the accelerated dissolution / growth kinetics that can be ascribed to a
higher fluoride concentration. In view of these first experiments, the concentration of 0.14 M
NaF was taken as an optimum and further experiments were performed only with this
concentration of electrolyte [190].
Figure 4.8 shows SEM images of the self-organized nanotube TiO2 layer formed in 1
M Na2SO4 / 0.14 M NaF electrolyte after 6 hours of anodization. As can be seen, compared to
the situation after two hours of anodization (as in Fig. 4.7c), the thickness and the surface
morphology of the layer further evolved into a thicker (Fig. 4.8a) and a more ordered state
(Fig. 4.8b). In other words, the self-organization of the nanotubular layer evolved further into
a more organized level. Additionally, the nanotubes in the layer have approximately a 5 times
higher aspect ratio 24 compared to 5. Aspect ratio is defined as a ratio between the tube
length and the diameter, e.g. 2400 nm length / 100 nm diameter 24. The amount of
precipitates on the top of the layer is negligible. Interestingly, from the bottom morphology
shown in Fig. 4.8c and from the Ti substrate separated by a lift-off of the tube layer (Fig. 8d),
it can be seen that the diameter deviation is now much higher as compared to the H2SO4 / HF
results (Fig. 4.3). That means that some very small diameter nanotubes (diameter ~ 30 nm)
can be seen amongst the larger ones. However, the morphology of these thicker layers reflects
the same morphology of samples anodized in H2SO4 / HF i.e. the nanotubes are open on the
top of the layer and closed on the bottom.
By comparing Fig. 4.8b and Fig. 4.8c it can be seen that the average tube diameter of
100 nm is reflected also in the bottom morphology, where hillock spacing is also about 100
nm. This suggests that the lateral spacing of the ordering phenomenon is kept constant during
etching. In these images, the tube walls have a thickness between 10 to 20 nm.

59

a)

top

b)

2.4 m

bottom

200 nm

c)

d)

200 nm

200 nm

Figure 4.8 SEM images of the self-organized nanotube TiO2 layer formed in 1 M Na2SO4 /
0.14 M NaF electrolyte after 6 hours of anodization at 20 V (sweep rate 0.1 V / s).

In order to gain information on thickness gain during the nanotube layer growth, SEM
cross-sectional images were taken from the mechanically bent samples, exactly in the same
fashion as the tube layers formed in the acidic HF containing electrolyte (Fig. 4.3d). Figure
4.9 shows the thickness of the nanotube layer as a function of the anodization time for the 1 M
Na2SO4 / 0.14 M NaF electrolyte. The thickness initially increases steeply from about 1 m at
30 min to about 2 m at 2 hours. This is followed by a plateau of approximately 2.4 m
thickness after 6 hours of anodization [190].
Another very interesting finding is that during potentiostatic polarization in the 1 M
Na2SO4 / NaF electrolytes, periodic current oscillations are observed, as previously shown
also for the H2SO4 / HF case [119] in Fig. 2.13. Figure 4.10a shows current transients
recorded while holding the potential at 20 V constant after a potential ramp (0.1 V / s) in 1 M
Na2SO4 electrolyte with and without addition of NaF (0.14 M). During anodization in a NaF free Na2SO4 electrolyte, the current drops rapidly as a consequence of the classical high-field
oxide layer formation, discussed in Chapter 2.

60

Figure 4.9 Thickness of the self-organized TiO2 nanotubular layer as a function of


anodization time (the same conditions as in Fig. 4.8).

If NaF is added to the Na2SO4 electrolyte, the current transient deviates from the NaF free curve (Fig. 4.10a). This phenomenon can be assigned to an increase of the surface area
and conversion from an originally flat layer into a self-organized nanotube layer. More
importantly, regular current oscillations are evident in the presence of NaF as shown in the
magnified view in Fig. 4.10a. These events remain active during the entire polarization time.
Fig. 4.10b shows magnified SEM side-view on the tube walls of the sample, whose current
transient is shown in Fig. 4.10a. It can be seen that the tubes have rings (ripples) on their walls.
The presence of the rings on each single tube and observation of the current oscillations
occurring during the experiment are most likely related. The correlation between rings and
oscillations is supported by the fact that the tube wall dissolution rate of approximately 20-30
nm / min is comparable to the ring spacing of approximately 25-30 nm observed in Fig. 10b.
The dissolution rate in terms of the tube length is based on an assumption that the pH along
the walls (in the pH range of 3-5) is lower compared with the bulk pH of the electrolyte (pH
7) due to the hydrolysis (Eq. 2.3) and dissolution reactions (Eq. 2.15) and absence of the fieldaided TiO2 dissolution along the tube walls. According to Fig. 4.5, the TiO2 dissolution rate at
these lower pH values is in the range of 20-30 nm / min, thus the current oscillation frequency
(~1 / min) can be converted into a tube length dissolution scale of approximately 20-30 nm /
min.

61

Figure 4.10 a) current transients recorded at 20 V in 1 M Na2SO4 electrolyte with and without
NaF after a potential sweep (sweep rate 0.1 V / s). Magnifications show detail of current
behavior; b) SEM image of the tube walls showing ripples (marked by an arrow) periodically
grown on the tube walls.

It can be concluded that nearly a steady-state tube growth has been reached after 6
hours from the fact that the current-time curve (Fig. 4.10a) has reached nearly a constant
current density and there is almost a constant thickness of the nanotube layer (Fig. 4.9). In
other words, on the longer time scale, the thickness of the layer does not significantly increase
anymore and the current does not decrease any longer, because the chemical and field-aided
TiO2 dissolution and field-aided TiO2 formation proceed at the same equilibrated rate [190].
In order to gain additional information about the anodization in the neutral sulfate /
fluoride electrolytes, pH - measurements of small electrolyte portions (each time 2 ml taken
from the vicinity of the Pt counter and the Ti working electrode) were carried out. In the first
type of experiment, one cell filled with a large electrolyte volume (to minimize the
concentration drift) was used for the anodization with both Pt and Ti electrodes, but
sufficiently separated from each other. In another experiment, the anodic and the cathodic
reactions were separated via a salt bridge between two cells with Ti in one and Pt in the other.
Both experiments led to the same result. This is shown in Fig. 4.11 for the Ti working
electrode area. The pH of the as-prepared 1 M Na2SO4 / 0.14 M NaF electrolyte was 7. In the
course of anodization, the pH - values were changing due to electrode processes, namely due
to the field-aided Ti oxidation, as in Eq. 2.3, and the counter reaction at the Pt electrode, as in
Eq. 2.2 (due to the expected alkalization H2O OH-, the pH value reaches 10). In other
words, the reason for lowering the pH at the working electrode is a permanent generation of
acidity (H+) as a product of the hydrolysis reaction at the Ti / TiO2 interface (Eq. 2.3) and

62

permanent mixing with the fresh electrolyte species entering the tubes, or present on the TiO2
tube layer.

.
Figure 4.11 Measurement of the pH in the vicinity of Ti (working) electrode during
anodization in 1 M Na2SO4 / 0.14 M NaF electrolyte at 20 V.

In other words, the concept outlined in Fig. 4.6 seems to be applicable and the findings
are in line with a presumption that under slow oxide etching conditions, the self-organized
TiO2 nanotube layer may be grown to a higher thickness (e.g. 2.4 m). Extended discussion
about the pH profile within the tubes and its influence on the morphology will be given in
Chapter 5.
At present there are no experimental straight forward techniques to measure pH inside
a single nanotube with 100 nm diameter, or to remove the electrolyte from different parts of
the tube for chemical analysis. However, dissolution experiments conducted in Fig. 4.5 show
in every case that pH of the electrolyte has a great impact on the TiO2 dissolution rate. It
should be noted, that adding HF to Na2SO4 (or eventually NaF to H2SO4) results still in too
high acidity. As a consequence, the dimensions of these layers still resemble dimensions of
those formed by anodization of Ti in H2SO4 / HF mixtures.
Furthermore, cross-sections of samples produced in the neutral electrolytes showed
less loss of the Ti metal compared to the acidic electrolytes (Fig. 4.4). This is also evident
from the lower current densities (compare Fig. 2.13 for HF electrolyte and Fig. 4.10a for NaF

63

electrolyte). On the other hand, the top views (Figs. 4.7 and 4.8b) show that there is still some
irregularity in the tube formation process. In fact some of the tubes do not grow absolutely
straight and some etch traces on the surface are apparent. It is clear that within the present
conditions that the optimum parameters have not been found.
Other electrolytes were tested, apart from Na2SO4 / NaF, composed of the
``successful SO42- / F- pair coupled with Li+, K+, Cs+ and NH4+ cations [191]. In all these
cases, more or less organized structures could be grown, but some specific features of these
electrolytes came into play. Anodization in the Li+ and Cs+ containing solutions led to the
least ordered and fairly heterogenous structures. Another drawback of Cs-containing
electrolytes was their high price. Therefore, they were not considered further for the
experiments. In the K-containing electrolytes, structures similar to those of Na-electrolytes
could be grown, but in all cases the nanotubular layers were covered with precipitates
consisting of potassium titanate [117,191]. Finally, experiments with NH4F (in 1 M
(NH4)2SO4) resulted in the most ordered and homogenous self-organized TiO2 nanotubular
layers in the dimensions very similar to the NaF case [191,192].
All findings in Fig. 4.3 - 4.11 clearly indicate that the morphology of the nanotube
layers depend on a range of electrochemical parameters. High aspect ratio ordered nanotubes
were obtained in Na2SO4 / NaF and (NH4)2SO4 / NH4F electrolytes (AR > 20).

4.3.2 Anodization in (NH4)2SO4 / NH4F electrolytes


The following section aims at more detailed exploration of anodization parameters in
(NH4)2SO4 / NH4F electrolytes. Additionally, apart from the high resolution of the layers
formed in the latter electrolyte, a very good solubility of NH4F ( 100g / 100 ml of water)
enabled anodization in the electrolytes with a wide variety of concentrations, including very
high NH4F concentrations compared to the NaF case ( 4g / 100 ml of water). The results of
this particular investigation are given in Fig. 4.12, which shows a sequence of SEM images
taken from samples anodized in 1 M (NH4)2SO4 concentrations, including very high
concentration ( > 1 M NH4F), at 20 V for 2 hours.
Strong differences in the morphology can be observed. In general, using a very low
NH4F concentration (Fig. 4.12a, 0.03 M) is unsatisfactrory because no nanotube layers are
obtained. Use of a higher NH4F concentration (Fig. 12 b and c) leads to a growth of somewhat

64

regular self-organized tubular structures, while in very concentrated electrolytes (Fig. 12 d, e)


the structure is rough and cross-linked.

a)

b)
1 m

1 m

200 nm

c)

200 nm

d)
1 m

1 m

200 nm

e)

200 nm

f)
0.6

2 m

200 nm
Figure 4.12 SEM images of the self-organized nanotube TiO2 layers formed by anodization
for 2 hours at 20 V (sweep rate 0.1 V / s) in 1 M (NH4)2SO4 with different NH4F
concentrations: a) 0.03 M; b) 0.14 M; c) 0.27 M; d) 0.6 M, e) 1.5 M. The insets show crosssectional views of the samples. Corresponding current transients recorded during anodization
(f).

These findings clearly demonstrate that under "suitable" anodization conditions (such
as potential, time and fluoride content) the self-organization can and likely will take place.
The opposite situation will occur, even if only one parameter, such as demonstrated here on
the example of high NH4F content, does not match the ideal set of all conditions. This can be
deduced also from the current transients recorded during anodization of these samples.

65

Typically, the current densities accompanying the self-organized tube growth are in the range
of 1 to 3 mA / cm2. If the values are higher, anodization never leads to the growth of selforganized structures, but only porous and cross-linked structures, such as shown in Fig. 4.12d,
e. Based on these findings, the NH4F concentration of 0.14 M was selected as the optimal
concentration for the nanotube growth and further experiments were performed mainly in this
solution [192].
The electrochemical behavior of Ti in 1 M (NH4)2SO4 / 0.14 M NH4F was
characterized by aid of standard electrochemical techniques [193]. Fig. 4.13 shows
polarization curves (a) recorded from the OCP to different potential values and the current
transients (b) recorded at the final constant potential. In the polarization curves, a current
plateau is observed from 7 V until approximately 40 V. At higher potentials, the current starts
to increase due to possible passivity breakdown events and accelerated field-aided Ti
dissolution.

Figure 4.13 Polarization curves (a) and current transients (b) recorded during anodization in 1
M (NH4)2SO4 / 0.14 M NH4F electrolyte at different potentials. The sweep rate is 1 V / s.

From Fig. 4.13b it is clear that the steady state currents gradually increase with a
higher applied potential from 1.1 mA / cm2 at 16 V to 2.3 mA / cm2 at 25 V and up to
approximately 50 mA / cm2 at 60 V. Remarkable is a high number of current spikes
(breakdown events) at anodization potentials of 50, or 60 V. SEM characterizations of these
samples revealed that the most regular structures were obtained by anodization at 20 V. The
layers formed at potentials significantly below 20 V exhibit a relatively disordered and thin
nanotubular structure. The layers do not thicken significantly with time, because there is
insufficient electric field to promote the field-aided TiO2 dissolution to be high enough to
66

facilitate the formation of more ordered and thicker layers. The films formed between 22 V
and 30 V have irregular, partially tubular structure located randomly in the porous layer
covering the rest of the area. Above 30 V, large cracked areas in the porous oxide film are
observed, as well as craters at the Ti substrate / TiO2 layer interface with dimensions on the
m-scale. These can be attributed to a series of breakdown events and stresses in the layers
established at these higher potentials. Therefore, a potential of 20 V was selected to be the
best for further investigations (as in previous cases with HF- or NaF-containing electrolytes).
Figure 4.14 shows SEM images of the self-organized TiO2 nanotube layer formed in 1
M (NH4)2SO4 / 0.14 M NH4F electrolyte after 6 hours of anodization at 20 V [192].

a)
top
700 nm

2.5 m

4 m

bottom
c)

b)

200 nm

200 nm

Figure 4.14 SEM images of the self-organized TiO2 nanotube layers formed at 20 V for 6
hours (sweep rate 0.1 V / s) in 1 M (NH4)2SO4 / 0.14 M NH4F showing the cross-sectional (a),
the top (b) and the bottom (c) view of a 2.5 m thick layer.

From the cross-sectional view (Fig. 4.14a) it can be seen that a thickness of
approximately 2.5 m has been attained [192]. Based also on the top-view (Fig. 4.14b) and
the bottom-view (Fig. 4.14c) it is evident that the tubular structure consists of regular
nanotube arrays with a tube diameter of approximately 100 nm with an average spacing of
150 nm. As in previous cases, the tube mouths are open on the top and closed on the bottom

67

with a barrier layer of TiO2 on the inner interface between the Ti substrate and the nanotube
layer. By comparing SEM images of samples anodized in 1 M Na2SO4 / 0.14 M NaF (Fig. 4.8)
with SEM images of samples anodized in 1 M (NH4)2SO4 / 0.14 M NH4F (Fig. 4.14) it is
evident that the thickness and the level of ordering has slightly improved. More specifically,
the diameter distribution is narrower and the nanotubes seem to grow not only longer, but also
straighter. Remarkably, there is no precipitation on the surface.
Another crucial factor that was found to affect the pore morphology is the scan rate of
the potential during polarization. Fig. 4.15a shows polarization curves recorded while
scanning the potential with different sweep rates from the OCP to 20 V. The experiments
were carried out in the 1 M (NH4)2SO4 / 0.14 NH4F electrolytes and for comparison also in a
NH4F-free electrolyte. Clearly, the faster are the sweep rates, the higher are the current
densities during the potential sweep.
Fig. 15b shows the current transients recorded while holding the potential constant at
20V. From these current transients it can be seen that in the NH4F-free (NH4)2SO4 electrolyte
the current drops following a classical high-field passivation behavior.

mV

Figure 4.15 Polarization curves (a) and current transients (b) recorded during anodization in 1
M (NH4)2SO4 / 0.14 M NH4F at 20V with different sweeping rates (in mV).

More importantly, in the NH4F electrolytes, the ramping speed has a lasting impact on
the steady state current density observed at the end potential of 20 V. Even after 2 hours at 20
V, the current density of the sample anodized with a high sweep rate did not drop to values
obtained with a lower sweep rate.
To demonstrate, how influential the sweep rate was on the initiation and growth

68

process of the tube layer, Fig. 4.16 is included here with intentionally high NH4F
concentrations used (0.6 M and 1.5 M). For comparison see Fig. 4.12, where anodization in
these electrolytes did not result in nanotubular structures because the conditions for selforganization were not fulfilled. However, Fig. 4.16 shows that it is possible to attain a selforganization typical for lower fluoride concentrations, by using a very slow potential sweep.
Anodization in 1 M (NH4)2SO4 electrolyte containing 1.5 M and 0.6 M NH4F results in thin
tubular layers as shown in Fig. 4.16a and 4.16c, respectively, when using very low sweep rate
(10 mV / s). Using a fast sweep rate (5 V / s) leads to the growth of several micrometers thick
porous layers that are shown in Fig. 4.16b and d, respectively. In this case, the conditions for
self-organization are not fulfilled due to very high current flowing through the samples. The
structures are very rough and have irregular surfaces that from the side view appear loosely
cross-linked and not tubular.

1.5 M

10 mV / sec

5 V / sec

0.6 M

Figure 4.16 Dependence of anodization parameters on the morphology of nanotubular TiO2


layers. Samples were anodized in 1 M (NH4)2SO4 with 1.5 M NH4F (a, b) and 0.6 M NH4F (c,
d) addition at 20 V with a lower ramp speed (10 mV / s, left columns) and with a high ramp
speed (5 V / s, right columns).

It is noteworthy that for the 1 M (NH4)2SO4 / 0.14 M NH4F electrolyte, the diameter is
only slightly influenced by the sweep rate. The tube diameter increases from 90 to 110 nm (on
average), as the sweep rate increases.

69

4.3.3 Stages of the tube growth in aqueous electrolytes


In order to extract key factors about the tube growth and particularly about the
initiation phase, SEM images were acquired from the samples removed from the 1 M
(NH4)2SO4 / 0.14 M NH4F electrolyte after different times [193]. Fig. 4.17 shows an evolution
of the morphology of these samples that were anodized at 20 V (after a potential ramp to 20 V
with a rate of 1 V / s). Immediately after the sweep from the OCP to 20 V, a thin rough layer
( 20 - 50 nm) covers the entire Ti surface of the titanium (Fig. 4.17a). On the sample
removed after 3 min a compact layer is observed, with a thickness of about 50 nm (Fig. 4.17b).
The layer thickening in this anodization phase is in agreement with the drop of current density
after a constant potential has been established (Fig. 4.15b). From 3-6 minutes, the current
density becomes more stabilized and the first signs of localized attack become apparent on the
surface. The breakdown sites that are randomly distributed over the film surface and the round
shaped holes in the substrate can be observed in Fig. 4.17b and Fig. 4.17c. By 6 minutes, the
SEM images reveal that underneath the relatively compact top layer some pore structures are
becoming visible most likely due to the localized penetration and dissolution of TiO2 by the
F- species in the electrolyte. This porous layer has grown to a thickness of 200 to 300 nm and
a relatively regular worm-like structure (Fig. 4.17c). By 15 - 20 minutes at some areas, the
structure has transformed from the porous structure into a nanotubular structure due to TiO2
dissolution and self-organization. However, there are still areas of the rough top layer, which
are found across the entire surface after 15 min (Fig. 4.17d).
By 20 minutes the current drops further (Fig. 4.15b), while the thickness of the
nanotubular layers increases up to 800 nm (Fig. 4.17e). In this period, the amount of the
unwanted remnant top layer is strongly decreasing, as the etching front moves towards the
substrate and this remnant layer is partially dissolved and partially released from the surface.
Anodization for longer times (more than 1 hour) leads to a gradual reshaping of the initially
rough irregular structure into a homogenous self-organized nanotubular structure (with a
thickness of approximately 2 m). This covers the entire anodized Ti surface with no
significant presence of a remnant layer after 2 hours, as shown in Fig. 4.17f [191].

70

Figure 4.17 An evolution of the morphology of Ti samples anodized in 1 M (NH4)2SO4 / 0.14


M NH4F at 20 V after a potential ramp (rate 1 V / s) for 0 min (a), 3 min (b), 6 min (c), 15
min (d), 20 min (e), 2 hrs (f). The upper parts shown the top-views, the lower parts show the
cross-sectional views.

It is necessary to mention that for slower sweeping rates (less than 100 mV / s), the
transition from one state to another, as shown in Fig. 4.17, can not be so clearly distinguished,
particularly in the early stages of anodization. The reason for this is that for samples anodized
e.g. with a sweep rate of 20 mV / s, it takes nearly 17 minutes to reach the final potential of
20V and during this period significant morphological changes (initiation of the nanotube
growth) of the growing layer are occuring. This is not so sharply pronounced for faster
sweeps. This fact is supported by the behavior and the magnitude of the current, as shown in
the polarization curves and current transients in Fig. 4.15. For the slower sweeping rates, the
current transients do not show significant fluctuations and are not as high as those for faster
sweeping rates. In other words, for slow sweep rates, the oxide growth and also the
dissolution proceed before the potentiostatic hold is reached.

71

4.4 Tailoring the dimensions of the self-organized nanotube layers


in non-aqueous electrolytes
This section deals with studies of the growth of self-organized TiO2 nanotubular layers
by anodization in organic electrolytes containing fluorides, namely glycerol and ethylene
glycol (purchased as anhydrous compounds). The main reason for selecting these electrolytes
(apart from their low water content) is that their high viscosity may influence the diffusion of
the electrolytes species according to the Stokes-Einstein equation:
D = kBT 6a ,

(4.1)

where D is diffusion constant, kB is Boltzmanns constant, T is absolute temperature,


is the dynamic viscosity and a is the radius of a spherical body.
According to the inversely proportional dependence of the diffusion constant D on the
viscosity , it can be expected that the high viscosity of these electrolyte will have a strong
influence on the kinetics of the nanotube layer growth and will strongly influence the
resulting morphology. This presumption goes in hand with the fact that the field-aided oxide
formation (Eq. 2.3) and dissolution at the tube botttom (Eqs. 2.14 and 2.15) are accompanied
by a significant generation of acidity. In other words, there must be a locally acidic pH at the
tube bottoms, no matter whether acidic or neutral electrolytes are used. Therefore, another
idea is put forward, that is, by using a highly viscous electrolyte, better confinement of pH can
be achieved and the dissolution should therefore occur only at the very bottom of the tube
(and also at lower current densities). This situation is depicted in Fig. 4.18.
Aqueous electrolyte

TiO2
Ti

Viscous electrolyte

low pH

TiO2
Ti

Figure 4.18 Difference in the pH confinement at the tube bottom in aqueous and viscous
(organic) electrolytes.

72

Based on these assumptions, it can be speculated that diffusion of species is the main
factor balancing local acidification at the tube bottom. Thus, a straightforward remedy to alter
the pH confinement at the tube bottom is to dampen the fluctuations between the oxide
growth and dissolution by changing the diffusion of the fluoride species. This can be achieved,
as already mentioned, by using highly viscous electrolytes, such as glycerol solutions [194].

4.4.1 Anodization in glycerol electrolytes


In order to quantify an effect of the electrolyte viscosity, anodization experiments were
performed in both the aqueous and non-aqueous electrolytes under otherwise the same
anodization conditions. It can be clearly seen in Fig. 4.19 that the current-time transient
recorded with very dense data acquisition (100 data points / s) after a potential sweep from 0
to 20 V (1 V / s) in a fluoride containing glycerol electrolyte (0.14 M NH4F) shows much
lower current densities than in the purely aqueous electrolyte (1 M (NH4)2SO4) / 0.14 M
NH4F). This demonstrates that diffusion could possibly be the rate-determining step for the
tube growth. However, more important than the lowering of the current by glycerol is the fact
that the curve does not exhibit any detectable current fluctuations [194].
Due to the decrease in the diffusion of the electrolyte species (and slowing immediate
TiO2 dissolution vs. formation kinetics), a better confinement of the acidic area at the tube
bottom takes place and the tubes grow now to a length of 7 m as shown in Fig. 4.20 in line
with the presumption outlined in Fig. 4.18. The inner nanotube diameter is approximately 40
nm, although they range between 30 and 60 nm. This means that tubes with higher aspect
ratios AR 175 are obtained (with some scatter due to the diameter variation), compared with
the sulfate-based electrolytes with ratios 25.
Fig. 4.19b plots the nanotubular layer thickness as a function of anodization time (at
20 V) for both electrolytes in Fig. 19a (thickness was measured directly from the crosssectional images of the samples). A sweep rate of 0.1 V / s was selected in order to have
identical conditions with the experiments in purely aqueous electrolytes. The thickness time
evolution is exactly in the same style as in the case of H2SO4 / HF (Fig. 4.3d) and Na2SO4 /
NaF (Fig. 4.9). For the aqueous electrolyte the thickness levels off after few hours. However,
for the glycerol case, it takes a much longer time to establish a steady-state thickness, the tube
length increases almost linearly and even after 18 hours (~ 6 m) it is still not steady state
[195].

73

a)

b)

Figure 4.19 (a) Current transients recorded at 20 V in 1 M (NH4)2SO4) / 0.14 M NH4F and in
glycerol / 0.14 M NH4F after a potential sweep (sweep rate 0.1 V / s). Insets in (a) show
magnifications of the current density; (b) thickness of the nanotube layer as a function of
anodization time.

Figure 4.20 SEM images of the self-organized TiO2 nanotube layers formed in glycerol /
0.14 M NH4F at 20 V for 13 hours (sweep rate 1 V / s) showing the cross-sectional (a), the top
(b) and the bottom (c) view of a 7 m thick layer.

74

It is noteworthy to mention that for a sweep rate of 1 V / s, the thickness of 7 m has


been attained after 13 hours of anodization, as shown in Fig. 4.20, whereas with 0.1 V / s it is
only around 5 m. This means, that the sweep rate (and the corresponding current density)
has a lasting impact on the structure, as seen in the previous case for 1 M (NH4)2SO4) / 0.14
M NH4F electrolyte (Fig. 4.15).
Even more importantly, the tubes formed under these high viscosity conditions show a
smooth appearance over their entire length as shown in Fig. 4.21. The tubes are grouped in
very tight bundles and do not seem to be connected together. This is in contrast to all previous
growth attempts, where there was always a connection of the side-walls (as in Fig. 4.3, 4.10,
4.14 etc.). Images of the interface between the TiO2 nanotubes and the underlying Ti substrate
reveal that the cap-like shape of the closed ends of the nanotubes is imprinted in the Ti
substrate (Fig. 4.21c and d). This feature is identical to nanotube layers grown by anodization
in aqueous fluoride containing electrolytes (Fig. 4.8d, 4.14c), but in this case, the tubes are
more unifom and spatially distributed. I believe that this cap-like (nanotubes) vs. dimple-like
(Ti substrate) construction plays a significant role for the good adhesion of the nanotubes to
the substrate.

Figure 4.21 SEM images of the side walls (a,b) of the layer shown in Fig. 4.20, lower part of
the nanotubes (of the same layer) attached to the Ti substrate (c) and the Ti substrate with
typical dimples after anodization (d).

75

4.4.1.1 Influence of temperature during anodization in glycerol electrolytes

In this section, the growth of TiO2 nanotube layers is further explored in glycerol
electrolytes at different temperatures. It will be shown that the temperature during anodization
has a great impact on the dimensions of the nanotubular layer. In the case of porous alumina,
the temperature during the anodization also plays an important role [1-13]. From the previous
findings it is evident that the resulting morphology and dimensions of the TiO2 nanotube
layers are strongly influenced by the electrolyte composition, anodization potential and
duration. Nanotubular layers grown in the glycerol electrolyte clearly exhibit a higher degree
of homogeneity and self-ordering, than those grown in aqueous electrolytes. Investigation of
the influence of temperature is particularly interesting for glycerol electrolyte, as its viscosity
strongly changes with temperature [196].
Preliminary experiments with various viscous electrolytes (glycerol, acetic acid,
ethylene glycol, dimethyl sulfoxide) showed that tube formation is possible in glycerol /
NH4F mixtures in a temperature range from 0 80C. These electrolytes were selected in
order to explore the influence of electrolyte viscosity on the tube growth, while keeping all
other experimental conditions, such as the amount of F- species and the applied potential
constant. The viscosities of electrolytes used for this series are given in Table 4.1 for different
temperatures.
Table 4.1 Viscosities of selected electrolytes at various temperatures
Electrolyte
Glycerol / 0.14 M NH4F
Glycerol / 0.14 M NH4F
Glycerol / 0.14 M NH4F
Glycerol / 0.14 M (NH4)2SO4
Glycerol : H20 (50:50 vol. %) / 0.14 M H4F
1 M (NH4)2SO4 / 0.14 M NH4F
Ethylene Glycol / 0.14

Temperature / C
0
20
40
20
20
20
20

Dynamic viscosity / Pa.s


~ 12
~ 1.5
~ 0.3
~ 1.5
~ 0.004
~ 0.001
~ 0.022

A set of anodization experiments in glycerol / 0.14 M NH4F electrolyte at different


electrolyte temperatures (0, 20, 40C) was performed at 20 V for 6 hours. Figure 4.22 shows
the corresponding polarization curves (a) and the current transients (b). For comparison,
curves recorded with fluoride-free glycerol with addition of 0.14 M (NH4)2SO4 at 20C are
also included. In this case, only a compact oxide layer is observed after anodization (as in
previous cases, e.g. Fig. 4.2). For the samples anodized in fluoride containing electrolytes,

76

evidently the current densities increased with increasing temperature and kept these
proportions during the whole anodization period [195].

Figure 4.22 Polarization curves (a) and current transients (b) recorded during 6 hours of Ti
anodization at 20 V in glycerol / 0.14 M NH4F electrolytes at different temperatures. For
comparison flouride-free glycerol electrolyte is included; (sweep rate 50 mV / s).

Figure 4.23 shows corresponding SEM images of the nanotubular layers achieved by
anodization process shown in Fig. 4.22. From these cross-sectional views it is apparent that at
all temperatures, 0C (Fig. 4.23a), 20C (Fig. 4.23b) and 40C (Fig. 4.23c), self-organized
nanotubes are formed. The insets show the lowest parts of the tubes in detail. From these
images it can be seen that all the nanotubes have very smooth walls, i.e. they are ripple-free,
as in Figs. 4.20 and 4.21. Fig. 4.23d plots the tube diameter and the length as a function of the
temperature during anodization. It is evident that the higher the temperature, the larger the
diameter and the longer the tubes. The average tube diameters are 405 nm, 506 nm and
607 nm and lengths are 80050 nm, 2.20.1 m and 3.40.1 m for 0, 20 and 40C,
respectively. Some experiments were performed at 60 and 80C, however, only mechanically
unstable bundles of tubes could be formed on the Ti surface (partially cracked and
disconnected from the substrate).
The strong dependence of the results on the electrolyte temperature (and viscosity)
indicates that the dominant limiting factor is diffusion of the reactants to the pore tip (i.e. tube
bottom), or of reaction products away from the tip. It should be noted, however, that extended
anodization (longer than 6 hours) at 40C does not lead to longer tubes than at 20C, because

77

a lift-off of the tubular layers from the substrate occurred once the layers reached thicknesses
of 4-5 m.

Figure 4.23 SEM cross-sectional images of TiO2 nanotube layers grown by anodization in
glycerol / 0.14 M NH4F at 0C (a), at 20C (b) and at 40C (c). The insets show detailed
bottoms of the smooth nanotubes at higher magnification. (d) an evaluation of the tube
diameter and length as a function of the glycerol electrolyte temperature.
4.4.1.2 Influence of water content in the glycerol electrolyte

In this section, the influence of the water content in the organic electrolytes is
demonstrated. It will be shown that the addition of even a small amount of water has an
extraordinary effect on the morphology of the nanotubular layers [197].
Figure 4.24 shows polarizaton curves (a) and current transients (b) recorded during
anodization of Ti at 20 V for 3 hours in five different electrolytes consisting of glycerol, 0.27
M NH4F and different amounts of water, as marked in the figure (in volume %). The higher
the water content, the higher the current density is during the potential ramping. However,
during the constant potential polarization the differences become smaller.

78

Figure 4.24 Polarization curves (a) and current transients (b) recorded during 3 hours of Ti
anodization at 20 V in glycerol / water / 0.27 M NH4F electrolytes with different volume
ratios of glycerol : water; (sweep rate 250 mV / s).

As it was shown in Fig. 4.17, in the case of (NH4)2SO4) / NH4F electrolytes with slow
sweeping rates, there is no sharp transition between the different stages of the nanotubular
growth, because the nanotube growth initiated during the potential ramping. There is a
similarity between the growth speed in aqueous electrolytes (sweep rate related) and in
organic electrolytes with the same sweep rate, but different water content. For instance, for
the glycerol / water curve (50:50) in Fig. 4.24, the value of the current density is similar to the
values obtained during anodization in aqueous electrolytes. The other curves are significantly
lower, particularly in the later stages of the anodization. This is because the field-aided
processes proceed at comparably slower rates, as evident from the fact that the tube layer
contained remnants of the initial oxide (due to a lower etching ability) as observed by SEM.
The explanation is likely that the viscosity of the glycerol electrolyte (a function of water
content [196]), has a huge impact on the diffusion of all the species involved in the reactions
and thus on the magnitude of the field-assisted TiO2 formation and dissolution.
In order to grow an oxide layer, water needs to be present in the electrolyte (Eg. 2.3).
Naturally, a question arises, where does the water in anhydrous glycerol or ethylene glycol
electrolyte come from? The answer is very simple. Although they are reported to be
anhydrous (or non-aqueous), they are not completely water-free, because they still contain
some traces of residual water from their production (less than 0.4 wt. %). Additionally, it is
very likely that they absorb water upon exposure to the humid atmosphere.

79

Figure 4.25 shows SEM images of the nanotubular layers obtained in glycerol / water /
0.27 M NH4F mixtures with 16.7 % (a), 6.7 % (b), 0.67 % (c) and 0 % (d) of added water
(volume %), respectively. The left columns show top-view images, the middle columns show
cross-sectional images and the right columns show magnified images of the tube walls. It can
clearly be seen that anodization in all investigated electrolytes leads to the formation of selforganized TiO2 nanotubes, but the resulting structures are significantly different, namely, the
length, the diameter and the shape of the tubes differ strongly. The dimensions of the resulting
nanotubes are summarized in Table 4.2.

Figure 4.25 SEM top-views (left columns), cross-sectional images (middle) and magnified
images of the tube walls (right) of samples anodized as in Fig. 4.24 with different water
contents (volume % given in the images).

80

For the pure glycerol electrolyte, the tube walls are completely smooth over the entire
length and the tubes have the smallest diameter of all. When only 1 ml of H2O is added into
149 ml of glycerol (corresponding to 0.67 vol% of H2O), the nanotubes lose their smooth
appearance and some small ripples, or rings, can be found on the tubes walls (Fig. 4.25c).
Additionally, the tube length and diameter have slightly increased. This could be a result of
the added H2O that helps form more TiO2 ( increased oxidation length) and to consequently
dissolve more TiO2 ( larger diameter), as compared with the nanotubes grown in ``pure
glycerol.
Table 4.2 Dimensions of the nanotubes formed in glycerol / H2O / 0.27 M NH4F electrolyte
with different amount of H2O (images shown in Fig. 4.24).
Water content in glycerol / H20 / 0.27
M NH4F mixture /vol. %
50
16.7
6.7
0.67
0

Tube diameter /
nm
10510
10510
7017
505
405

Tube length / nm
145060
86050
90050
175050
170050

Important is the appearance from the top. In "pure" glycerol, patches of the original
oxidized Ti surface are present, whereas in the water-containing electrolyte, it is dissolved. As
soon as water is added, the smooth wall appearance completely disappears, the diameter of
the tubes starts to increase and resemble to the 100 nm diameter of nanotubes grown in
aqueous solutions at 20 V [197].
A different situation is observed with the tube length. It seems that glycerol / water
mixtures with between a few % and 50 % of H2O are not suitable for growing long tubes. In
other words, these conditions are not ideal for the growth of long tubes because they lack the
``right ratio between the diffusion of the species (influenced by the viscosity) and the tube
growth kinetics. Addition of more than 50 vol.% of H2O leads to the growth of nanotubes,
whose morphology and dimensions reflect the nanotubes grown in e.g. (NH4)2SO4 / NH4F.
However, for the conditions given in Fig. 4.24, the longest tubes are obtained for pure
glycerol / 0.27 M NH4F electrolytes (containing almost no water). On the other hand, as can
be seen from the top views in Fig. 4.25, the initial layer from the early growth stages is still
partially present on the nanotubular surface. After 3 hours of anodization, it has not dissolved
yet - in line with previous finding that the dissolution of the upper parts of the nanotubes
grown in glycerol / NH4F electrolyte is very mild in contrast to for instance (NH4)2SO4 /
NH4F electrolyte, as shown in Fig. 4.17.

81

4.4.1.3 Variation of the tube diameter influence of applied potential

In the previous results it was reported that for self-organized nanotube growth, the best
results were achieved, when a potential of 20 V was used. Interesting work by Bauer et al.
[198] however showed the possibility to achieve the nanotube growth in H3PO4 / HF
electrolytes over a range of anodic potentials (1 - 25V) and with a range of different tube
diameters (10 - 120 nm). In this section, I would like to demonstrate the possibility to grow
nanotube layers with even larger tube diameters by changing the applied potential, while
keeping all other conditions the same [197].
Figure 4.26 shows a sequence of SEM images taken from Ti samples anodized at
different potentials in mixture of glycerol and deionized water (50:50 vol. %) with 0.27 M
NH4F electrolyte. All samples have been anodized for 3 hours with the sweep rate of 250 mV
/ s. As can be seen, the formation of a self-organized and uniform nanotube layers with
different tube diameters is possible. To the best of my knowledge, this glycerol / water / NH4F
electrolyte is the only electrolyte up to now that allows growth of such a wide variety of
nanotube diameters. The higher the applied potential the larger is the tube diameter. The
smallest diameter was approximately 20 nm for the sample anodized at 2 V, the largest
diameter of ~ 280 nm was achieved at 40 V. Fig. 4.27 plots the tube diameter and also the
nanotube layer thickness as a function of anodization potential [197].
As it can be seen from Fig. 4.27, the tube diameter increases linearly with the applied
potential, however, this is not completely true for the layer thickness. After 20V, the thickness
starts to deviate from the linear behavior. This type of nanotubes with different diameters
could be profitably used for instance as a template for making nanowires with different
diameter, based on electrodeposition and selective removal of the TiO2 that builds the
nanotubes, as demonstrated in Fig. 2.4. Additionally, they could be suitable in biology to tune
the interactions between the tissue and the Ti implants (covered with a layer of highly
biocompatible TiO2 nanotubes). This has been demonstrated for the tubes grown by Bauer et
al. [198] that showed strongly different interactions with stem cells [199].

82

Figure 4.26 SEM images of TiO2 nanotube layers grown in a mixture of glycerol and
deionized water (50:50 vol. %) containing 0.27 M NH4F at different applied potentials during
3 hours (sweep rate 250 mV / s).

As these mixed glycerol/water electrolytes offer the possibility to grow nanotubes with
a wide range of diameters, attention should be paid also to the bottom tube layer (or, say, a
barrier layer at the tube bottom) of the nanotubes, a tube part that is connected with the Ti
substrate. A remarkable feature of the nanotubes shown in Fig. 4.26 is that upon scratching
they can crack along the tube wall (vertically to the substrate), whereas the tubes grown in
purely aqeuous solutions do not crack and disconnect from each other. This allows one to
look inside the broken tube and to measure the the bottom layer thickness. Examples of
several nanotube bottoms are shown in Fig. 4.28 that plots the tube bottom layer thickness
versus the applied potential. It is very interesting, that with increasing potential the thickness
83

also increases this is in line with growth of anodic compact oxides, as explained in Chapter
2. It is not only the tube diameter that increases with potential, but also the tube bottom layer
thickness [197].

Figure 4.27 Length and diameter of the TiO2 nanotube layers shown in Fig. 4.26.

Figure 4.28 Dependence of the tube bottom layer thickness on the applied potential. Insets
show SEM images of the tube bottoms.

84

Another particularly interesting finding is that for anodization in these mixed


glycerol/water electrolytes, a relatively broad range of NH4F concentrations can be used to
grow nanotubes [197]. This is in contrast to the tube growth in e.g. (NH4)2SO4 / NH4F under
similar conditions (i.e. 20 V, sweep rate between 0.1 - 0.5 V / s) see Fig. 4.12. Several
examples of nanotube layers grown by anodization in glycerol/water electrolytes (50:50 vol.
%, 20 V, 3 hours) with four different concentrations (0, 0.135, 0.27, 0.54 M NH4F) are shown
in Fig. 4.29. In the absence of fluorides, no nanotube layer is grown (the roughness originates
from the non-polished thin foil used here). In all other cases, with fluoride additions, growth
of nanotube layers is achieved. The higher the concentration of fluoride, the thicker is the
layer. However, when viewing from the top, the tube layers formed with very high F- content
(0.54 M) are partially collapsed and do not have a very homogenous structure over the entire
Ti surface.

Figure 4.29 SEM images of TiO2 nanotube layers grown by anodization during 3 hours at
20V (sweep rate 250 mV / s) in the glycerol / water mixture (50:50 vol. %) containing
different amounts of NH4F (given as insets).

It is important to compare the electrochemical data recorded during anodization in


these electrolytes since the nanotube growth was achieved over comparably wider Fconcentration range. Figure 4.30 shows polarization curves (a) and current transients (b)
recorded during anodic growth of corresponding layers shown in Fig. 4.29. The magnitude of

85

the current is clearly affected by fluorides in the electrolyte in the same trend as for the
thickness.

Figure 4.30 (a) Polarization curves and current transient (b) recorded during anodization
during 3 hours at 20V (sweep rate 250 mV / s) in the glycerol / water mixture (50:50 vol. %)
containing different amounts of NH4F.

This relation is not surprising, because with decreasing F- concentration, the


dissolution rate of TiO2 becomes slower, therefore the anodic current is smaller and the tubes
are shorter (in contrast to e.g. NaF / Na2SO4 electrolytes, where they do not grow at all).
To summarize: it is evident that self-organized TiO2 nanotubular layers grown in
glycerol electrolytes show a higher degree of regularity and homogenity in comparison with
the layers grown in the aqueous electrolytes, where considerable thinning of the tube wall
towards the mouth and ripples on the tubes are observed. The smoothness of the nanotubes
could play an important role for their behaviour, e.g. affect the light reflection and thus pave a
way to applications such as photonic crystals or waveguides. Additionally, the glycerol /
water / NH4F electrolytes provide a high degree of freedom for the self-organization of the
tubes and offer very spectacular possibilites for tailoring the nanotube geometry. The
morphology of the nanotubes can be strongly altered by using different anodization
temperatures, additions of water and by using a wide range of anodic potentials.

4.4.2 Anodization in ethylene glycol electrolytes


In adition to highly viscous glycerol-based electrolytes, viscous ethylene glycol-based
electrolytes were also used for anodization experiments (for viscosities see Table 4.1).
Strikingly, larger differences in the nanotubular morphology and dimension compared with

86

glycerol electrolytes could be observed between the nanotubes grown in anhydrous ethylene
glycol / NH4F electrolytes and those grown in ethylene glycol / water / NH4F.
Figure 4.31 shows polarization curves (a) and current transients (b) recorded during
anodization in anhydrous ethylene glycol / 0.27 M NH4F and in the same electrolyte with
intentionally added water (5 vol. %) at 60 V (sweep rate 0.1 V / s). Both the polarization
curves and current transients have similar shapes, but quite different current densities are
obtained. It is somewhat suprising that anodization in electrolytes with added water is
accompanied by lower currents, compared to the almost water-free containing electrolyte.
Fig. 4.32 shows SEM images of the corresponding samples. Fig. 4.32a shows that in
the non-aqueous electrolyte only relatively shaky and disorded nanotube layers are grown and
they are significantly thinner than the nanotube layers grown in the same electrolyte
containing 5 vol. % of water (Fig 4.32b). In fact, the first type of sample after passing ~ 40
V shows signs of anodic breakdown. The breakdown situation in not completely reached as
evident from rapid current decrease after reaching 60 V (Fig. 4.31b). When the same
experiment is done only up to 20 V (instead of 60 V), approximately 1 m thick nanotube
layer can be grown [195]. In other words, in this non-aqueous electrolyte at higher potentials,
the system lacks suitable conditions for self-organization, possibly due to a lack of water
needed for the oxidation reaction. In contrast, in the water-containing electrolyte at a high
potential (e.g. at 60 V), a growth of high aspect ratio nanotubes can be attained, as shown in
Fig. 4.32b (AR 200).

Figure 4.31 Polarization curves (a) and current transients (b) recorded during anodization in
different ethylene glycol electrolytes at 60 V during 6 hours (sweep rate 100 mV / s).

87

For comparison, Fig. 4.31 shows also data recorded during anodization in an ``aged
ethylene glycole containing HF instead of NH4F. Aged electrolyte means an electrolyte that
had been pre-anodized before the current growth attempt has been made in order to improve
the electrolytes performance. In other words, if a freshly prepared electrolyte was used, only
layers similar to those in Fig. 4.2 would be obtained. This aged electrolyte contains water
(since the HF comes from the 48 wt.% aqueous solution) and also some [TiF6]2- species as a
result of aging (Eq. 2.15). The current values are significantly lower for this electrolyte, but
still an ordered array of nanotubes can be grown, as shown in Fig. 4.32c.

Figure 4.32 SEM images of TiO2 nanotube layers grown as in Fig. 4.29 in a water-free /
NH4F (a), water-containing / NH4F (b) and HF-containing ethylene glycol (c).

A special class of nanotubes can be synthesized by anodization in these aged ethylene


glycol / HF electrolytes, as shown in Fig. 4.32, but at high anodic potentials (high electric
fields) [200,201]. A specific synergism between the etching ability of HF and its solvatization
(or, say, encapsulation) by ethylene glycol in these aged electrolytes has been revealed. In an
absence of applied potential, HF seems to be very passive, since Ti samples immersed in this

88

solution (purely chemical term ``solution is intentionally used here, instead of


electrochemical term ``electrolyte) did not visibly undergo any significant attack after
several minutes (this would not be the case in e.g. H2SO4 / HF solutions). Under anodic bias,
this synergism leading to a limited chemical etching of the tube walls and comparably strong
field-enhanced dissolution (drilling) of TiO2 at the tube bottoms can be profitably used to
grow exteremely thick nanotube layers (hundreds of micrometers) when ``strong anodizing
conditions are used (high anodic potentials applied in either one step, or few steps). An
example of a 145 m thick TiO2 nanotube layer is shown in Fig. 4.33. As can be seen, the top
morphology is not the same as in the glycerol electrolytes. In other words, the nanotubes are
covered by a hazy TiO2 ``grass, as a result of disintegration of the nanotube shape on the
very top of the nanotube layer. It should be noted, however, that the diameters of the tubes
grown in these electrolytes (~ 150 nm at 100V) do not correspond to diameters obtained in
glycerol / water / NH4F electrolytes (Fig. 4.27). According to conductivity measurements, this
could be due to the lower potential at the anode, due to the high resistivity (IR drop) of these
electrolytes (see Table 5.1 in Chapter 5).

Figure 4.33 SEM images of a TiO2 nanotube layer grown during 12 hours of anodization at
100V (reached with a sweep rate of 10 V / s) in aged ethylene glycol / 0.16 M HF electrolyte;
(a) side-view and (b) top-view of the layer; (c) tube walls in detail.

89

4.4.2.1 Hexagonal ordering of the nanotubes

In this section, key factors that influence the degree of ordering of the TiO2 nanotube
layers will be shown. As described in Chapter 2, highly ordered self-organized porous
alumina oxides structures were for the first time reported by Masuda and Fukuda [6]. In their
work, the main factors that were crucial for the perfectness of the arrangement and the ideality
of self-ordering were the anodization voltage, the purity of the material and self-induced or
external imprint molds that serve as initiation sites for a secondary pore growth [6,7]. The first
two factors were considered for the nanotube layers. Throughout all experiments performed in
this thesis, the highest degree of ordering of the nanotubes in terms of hexagonal shapes and
close-packing was achieved with ethylene glycol / NH4F electrolytes [202].
Similar to porous alumina, it was found that a repeated anodization of the Ti substrate
strongly increased the degree of order. This degree can be further improved by using a high
purity Ti. Figure 4.34 demonstrates the improvement of ordering of the nanotube layers
grown on the 99.99% pure Ti after the first (a) and the second (b) anodization step.
The nanotubes grown during the first anodization are not perfectly ordered and have
irregular shapes; during the second anodization step closed-packed nanotubes with hexagonal
shape are grown, as a result of a suitably pre-textured surface from the first anodization.
Furthermore, repetition of anodization process and using a high purity Ti significantly
reduced the variation in the average pore diameter and strongly reduced the areal density of
polygon ordering / packing errors, as shown in Fig. 4.34c. It shows additional data acquired
for the 99.6% pure Ti substrates. After the second anodization, number of pentagons and
heptagons has decreased, as well as the standart deviation of the hexagonal cell size [202].
It should be mentioned that a potential of 50 V and the use of high purity Ti substrates
were found to optimize the degree of ordering, minimize the defects in the nanotube layers
and reduce the standard deviation of the hexagonal cell size. At lower potentials (< 50 V),
these features were clearly less pronounced, and at higher voltages, breakdown events and
lift-off of the layers occurred.
An interesting finding is that for the nanotube layer grown under the best conditions
(99.99% purity, 50V, two anodization steps), the hexagonal order as well as the hexagonal
shape of the nanotubes was seen throughout the entire layer thickness. Figure 4.35 shows
SEM images that were taken at different depths in the tube layer (as marked by the windows).
It can be seen that each nanotube keeps its hexagonal shape from the bottom to the top.
Additionally, a strong variation of the wall thickness can be observed ranging from approx. 65
90

nm at the bottom to approx. 12 nm at the tube top. The inner tube diameter increases from
approx. 50 nm to 110 nm. While at the bottom the hexagonal tubes are closed-packed (Fig.
4.35c), there is clearly a gap of few nm at the upper part (Fig. 4.35a). This feature can be
ascribed to the chemical dissolution of the TiO2 nanotubes due to fluoride anions present in
the bulk electrolyte.

Figure 4.34 SEM bottom views of ordered domains of TiO2 nanotubes grown on 99.99%
pure Ti after 1st (a) and 2nd anodization (b) in aged ethylene glycol / 0.27 M NH4F electrolyte
at 50 V; (c) evaluation of cell defects (based on the SEM images) in terms of pentagons and
heptagons (present in otherwise hexagonally packed nanotube layers, area of 100 m2) and
standard deviation of the unit cell size for the nanotube layers after 1st and 2nd anodization of
99.6% and 99.99% Ti substrates. The first anodization steps were done with aged electrolytes
for 12 hours (sweep rate 1 V / s). The second anodization steps were done with the
electrolytes used in the 1st step for 6 hours (sweep rate 5 V / s). For removal of the first
nanotube layer, samples were polarized with a cathodic pulse in 1 M H2SO4 at -3 V leading to
a lift-off and breaking the layers (due to strong H2 evolution).

For the conditions described here, the degree of ordering is very high, but it is still not
perfect. However, performing more than two anodization steps does not significantly improve
the ordering. Considering the sister material - porous alumina - it is very likely that the degree
of ordering could be further extended by several approaches, including electropolishing,
91

uniform nanoimprinting of the Ti surface and possibly also annealing the Ti substrates prior to
the experiments in order to release stresses in the surface [4-7].

TOP

BOTTOM

Figure 4.35 SEM images of the close-packed, hexagonal TiO2 nanotubes grown as in Fig.
4.33b. Windows in the left image indicate roughly depths, where from images on the right
side were taken (i.e. upper part, middle and bottom of the layer).

4.5 Structure of the nanotube layers


It has been noted in Chapter 2 that one of the crucial factors for a solid performance of
TiO2 in various applications is its structure. Hence, it is necessary to determine the structure
of the TiO2 nanotube layer. This section deals with investigations of the structure of the selforganized nanotube layers, carried out by modern instrumental analytical techniques
transmission electron microscopy (TEM) and an X-ray diffraction (XRD).

92

Figure 4.36a shows a TEM image of the self-organized TiO2 nanotubes shown in Fig.
4.20 that were grown by anodization in glycerol / 0.14 M NH4F. This TEM image confirms
that the walls of these nanotubes are very smooth and straight. In order to gain more
information about the structure of the nanotubes, high-resolution TEM investigations have
been carried out. Fig. 4.36b shows an image of the bottom part of the nanotube layer. An
amorphous (disordered) structure is revealed that is confirmed by the diffraction pattern
shown in Fig. 4.36c, taken from the same area. The broad and diffuse ring confirms the
amorphous nature of the nanotube [194]. According to extensive TEM investigation, the
thickness of the walls is 12 nm, and the thickness of the cap-like barrier oxide is 16 nm. By
comparing TEM pictures of the walls from the bottom part and from the upper part (tube
mouth) essentially no variation in thickness could be determined. This is different to the
nanotubes grown in aqueous electrolytes (Fig. 4.3) and ethylene glycol electrolytes (Fig. 4.35),
where considerable thinning of the walls at the pore mouth is observed.
In order to convert the as-formed amorphous nanotubes into crystalline ones, the
sample was thermally annealed at 450 C for 3 h in air. This temperature has been selected
based on previous experience [144], and to have the temperature sufficiently above the
threshold crystallization temperature of 290 C, but below the temperature, where the
sintering of the TiO2 would occur, typically above 600 C [144]. Figure 4.37 shows HRTEM
images of the annealed nanotubes and a detail of the tube bottom and the tube wall. As it can
be seen, the annealed nanotubes have a crystalline structure. By comparing the bottom part of
the as-formed nanotubes (Fig. 4.36b) and the annealed nanotubes (Fig. 4.37b) it is evident that
after crystallization the shape of the nanotube bottom remains smooth and nearly round. The
electron diffraction pattern (inset in Fig. 4.36b) has several continuous rings which indicate
that the nanotubes are composed mainly of randomly oriented TiO2 crystallites (size ~ 10 to
20 nm). There is a slight deviation in the size distribution towards rather big crystallites. A
few large crystallites (> 20 nm) diffract much stronger and make the ring pattern in some
parts ``spotty. The first five diffracted rings correspond to approximate lattice plane
spacings d1=3.51, d2=2.37, d3=1.90, d1=1.71 and d1=1.67. These spacings can be
attributed to the d-spacings d{101}, d{004}, d

{200},

{105},

{211}of

anatase TiO2. Interesting is

the situation in the tubes wall (Fig. 4.37c): here the structure consists of very large crystallites
( 50 nm). The crystal lattice fringes can clearly be seen that have distances of 3.51
corresponding to the spacing of (101) lattice plane of TiO2 anatase phase. The Fourier
transformation of this image (inset in Fig. 4.37c) identifies anatase with intensive reflections
from the (101) plane pointing out the single crystalline nature of the tube walls [203].
93

Figure 4.36 TEM image of several nanotubes (a), a high-resolution TEM image of the single
nanotube bottom (b) and the TEM diffraction pattern of the tube bottom (c).

Figure 4.37 HRTEM images of two annealed TiO2 nanotubes separated from the nanotube
layer (a) and individual nanotube bottom (b) and the tube wall (c). Selected area diffraction
patterns (SAED; shown as insets) were taken from the corresponding regions shown in the
images.

94

The samples investigated by the HRTEM were studied also by XRD, as shown in Fig.
4.38. Crystalline anatase peaks (planes are annotated in the figure) are detected in the
annealed sample, while they are absent prior to the annealing. An evaluation of the peaks
reveals anatase planes that are consistent with the HRTEM findings (101, etc.).

Figure 4.38 XRD spectra of the as-formed (1) and the annealed (2) nanotubular layer
consiting of smooth and high aspect ratio nanotubes. Different anatase planes are annotated at
the corresponding peaks.

As noted in Chapter 2, the most common crystalline TiO2 phases are anatase and rutile.
In order to evaluate the crystallinity of the tubes at different temperatures and to find the
transition points from one phase to the other, a series of annealing experiments at different
temperatures was performed. For these experiments, 2 m thick nanotube layers grown in 1
M (NH4)2SO4 / 0.14 M NH4F electrolyte (at 20 V for 3 hours) were selected, since these were
the most common layers used in the investigations. After annealing, XRD measurements were
conducted, as shown in Fig. 4.39 [203].
The following observation can be made from Fig. 4.39: i) the as-formed nanotube
layer is amorphous; ii) the initiation of the crystallization from an amorphous phase into an
anatase phase begins at 250 C, when first anatase peak appears in the spectra. At 280 C, a
higher anatase peak can be seen (at 2 ~ 25). However, based on the TEM investigations, in
this temperature range there are only small crystallites embedded in an otherwise amorphous
structure. iii) Between 300 and 450C there is a strong X-ray response from the anatase
95

crystals (mainly 101 plane). Based on the TEM studies of these samples, the entire amorphous
structure has been transformed to crystalline anatase phase. iv) At comparably higher
temperatures (more than 500C), a mixture of anatase and rutile (110) is obtained. The higher
the temperature (> 500C), the more rutile is in the structure.

Figure 4.39 XRD spectra of the nanotube layer formed in 1 M (NH4)2SO4 / 0.14 M NH4F
electrolyte at 20 V and annealed at different temperatures for 3 hours with heating and cooling
rate of 30C / min (A represents anatase peaks, R rutile peaks, Ti represents Ti peaks).

The upper limit for the stability of the layers is the temperature of 600 C, as
confirmed by SEM investigations. Sintering and cracking of these layers occurs above 600 C
as a result of stuctural disintegrity. On the samples annealed below 500 C, neither
morphological changes, nor any cracks have been observed in the nanotube layers.
It is noteworthy that the temperature gradient heating and cooling rates was kept
routinely the same (30C per minute) during the entire crystallographic investigation. This
rate is very likely to have an influence on the size of the crystals. A more detailed
investigation of this rate would be desired and would give more insight into the annealing
process.

96

4.6 Chemical composition of the nanotube layers


This section deals with the chemical composition of the self-organized nanotube layers
as investigated by modern spectroscopic techniques, including X-ray photoelectron
spectroscopy (XPS) and energy dispersive X-ray analysis (EDX).
The XPS spectra revealed the chemical composition and also the oxidation states of
the relevant species. Figure 4.40 shows a typical XPS spectra taken at 45 of the nanotubular
layer grown by anodization in 1 M (NH4)2SO4 / 0.14 M NH4F electrolyte at 20 V. The Ti peak
and the O peak reveal the composition of the layer to be TiO2. Some residual amount of C is
also detected, however, that is very difficult to eliminate, because it comes from the
contamination by organic molecules adsorbed on the TiO2 surface.

Figure 4.40 XPS spectra of an amorphous nanotubular layer grown in 1 M (NH4)2SO4) / 0.14
M NH4F.

Within surface analytical techniques, a very useful tool for determining the
composition throughout the layer (in depth) is sputtering. This is based on the periodical
removal of a thin layer of the investigated material (typically up to 2 nm per step) and the
analysis step. This cycle is repeated, until the desired sputter depth is reached. However, this
method has two drawbacks. As reported in literature for both single and polycrystalline TiO2
surfaces [204,205] sputtering of TiO2 in vacuum by e.g. Ar+ ions with high energy may lead

97

to reduction of Ti4+ to lower oxidation states, most likely to Ti3+, and under extereme
conditions to Ti2+. Another sputtering artifact for TiO2 is that as soon as the metallic Ti
beneath the oxide is reached, it may reoxidize spontaneously even in the vacuum due to traces
of oxygen present in the chamber. As a consequence, the proportions of the O and Ti amounts
are misleading.
Another drawback of the TiO2 nanotubular layer for the XPS depth profiling might be
its microscopic roughness. This is because the interface between the Ti substrate and the
nanotube layer is not completely flat (at the m-scale) because the Ti substrate is
polycrystalline. Fig. 4.41 shows an optical image with clearly visible grains and their
boundaries for a flat polished Ti surface that was anodized in 1 M H2SO4 / 0.16 M NH4F at 20
V to form a 500 thick nanotubular layer. Since the orientation of the grains is different and the
TiO2 growth proceeds at different rates on these grains, the nanotube layer does have slightly
different thickness on the different grains, as observed from the SEM cross-sections of the
tube layers grown on the grain boundaries and in their vicinity. Even if only a thin tubular
layer would be examined by sputtering, one location could be sputtered bare from the
nanotube layer, whereas the other might not. Additionally, for extensive studies of the
chemical composition throughout the tube layers, XPS sputtering is not very suitable, also
because of the extremely long times that are needed to sputter a several micrometers thick
layer.

Figure 4.41 Optical image (taken under polarized light) of a 500 thick TiO2 nanotubular layer
grown on originally polished, mirror-like and flat Ti substrate

98

Therefore, in addition to XPS, another method, Energy dispersive X-ray spectroscopy


(EDX), was used to determine the chemical composition througout the nanotubular layers.
Figure 4.42 shows an example of the EDX spectra taken from the 2.5 m thick TiO2 nanotube
layer grown in 1 M (NH4)2SO4) / 0.14 M NH4F electrolyte (Fig. 4.14). The analysis revealed
that the as-formed nanotubular structure consists of Ti, O and a certain amount of F
(percentual composition is given in the upper right corner). An inset of this figure (white
frame in the bottom part) shows a short calculation provided to express the weight proportion
of Ti and O in the TiO2.

M (TiO2) = 80 g / mol,
Elemental: M (Ti) = 48 g / mol, M (O2) = 32 g / mol
Weight ratio of both elements for 1 mol TiO2:
x(Ti) = 48 / 80 = 0.6 x 100 % = 60 %
x(O2) = 32 / 80 = 0.4 x 100 % = 40 %

Figure 4.42 EDX spectra of the as-formed TiO2 nanotube layer.

According to the TiO2 stoichometry, 1 mol of TiO2 consists of 1 mol of Ti and 1 mol
of O2. Therefore, Ti contributes to 60% of the TiO2 molecule weight, and O the remaining 40
%. These calculated weight ratios correspond well to the values obtained by EDX

99

measurements. Thus, it can be again concluded that the nanotube layers consist of TiO2.
However, there is a significant amount of F species present in the layer, coming from the
fluorides (from the electrolyte) that remained inside the nanotubes after the anodization.
These traces are slightly influencing the EDX weight proportions.
Neverthless, based on the results of all three techniques (XPS, EDX, XRD), it can be
concluded that the nanotube layers are composed of TiO2. More attention to fluorides is given
in the next section.

4.6.1 Fluoride removal


The fluoride anion uptake during anodization is quite common for oxide layers grown
electrochemically in the presence of HF, or fluoride salts [116]. The effect is usually ascribed
to competition of fluoride ions with mobile oxygen ions during oxide growth, or to a specific
adsorption of these ions and encapsulation by overgrowth, or to precipitation. The presence of
fluoride ions can be detrimental to several applications: fluoride ions are considered to be
toxic (bio-applications) and corrosive [107,108], they may also introduce undesired electronic
states into an optoelectronic material [114,115]. In this work, two different approaches have
been performed in order to minimize, or completely remove the unwanted F content.
The first approach was long-term soaking of the samples in deionized water (and
changing the water many times), but this led to only a partial removal of the fluoride, as
confirmed by XPS and conductivity (of water) measurements before and after the soaking.
Fig. 4.43 shows an example of spectra taken from the freshly prepared sample and 24hrssoaked sample (grown in 1 M (NH4)2SO4 / 0.14 M NH4F electrolyte at 20 V).
A faster and more efficient approach was achieved by the thermal annealing of the
layers. During the crystallization of the nanotube layer, the fluoride concent in the layer was
strongly reduced as a side effect [203]. Fig. 4.44 shows (a) EDX and (b) XPS high-resolution
spectra acquired on the same sample (obtained also by anodization in 1 M (NH4)2SO4 / 0.14
M NH4F electrolyte) before and after annealing. In the as-formed layer, there is a considerable
amount of fluorine species present (up to about 5 wt. %), as shown in Fig. 4.42. The annealing
treatment (at 450C for 3 hours) reduced the fluorine species to less than 1 wt. %. It can only
be speculated, as to how and why this happens (evaporation as HF or F2 ). Nevertheless, an
important factor for the fluoride incorporation into the nanotubes could be the flexibility of
the amorphous state. Incorporation of additional impurities into an amorphous solid is allowed

100

by reduced structural constraints in an amorphous network. During the crystallization, foreign


atoms are driven out of the growing highly-ordered crystallites.

Figue 4.43 XPS spectra acquired before and after soaking of TiO2 nanotubular layers in
deionized water for 24 hours.

Figue 4.44 EDX (a) and XPS (b) spectra of TiO2 nanotubular layer, acquired before and after
annealing at 450C for 3 hours.

It should be noted, however, that whenever the use of the as-prepared (amorphous)
nanotube layers was desired for further investigations, the latter approach leading to the tube
crystallization could not be used, thus only a long-term soaking in water was performed.

101

4.7 Applications of self-organized TiO2 nanotube layers


As already pointed out in Chapter 2, a whole range of TiO2 applications is based on its
semiconducting nature. Therefore, it is desirable to investigate the photoelectrochemical
response of the tube layers [206]. Fig. 4.45 shows a photocurrent spectra recorded across the
UV and near VIS light and plots the photocurrent as so-called ``incident photon to electron
conversion efficiency (IPCE, often called also quantum efficiency), calculated according to
the following equation:
IPCE (%) =

i ph hc

Pq

100

(4.2)

where iph is the photocurrent density, h is Plancks constant, c velocity of light, P the light
power density, is the irradiation wavelength, and q is the elemental charge.
In Fig. 4.45 are shown data recorded for the as-formed (amorphous) layer, the annealed
anatase tube layer and for comparison also the annealed flat anodic TiO2 (grown in HF-free
H2SO4 electrolyte). The nanotube layers were grown by anodization in 1 M (NH4)2SO4 / 0.14
M NH4F electrolyte at 20 V for 2 hours (sweep rate 0.1 V / s).

Figure 4.45 Photocurrent action spectra comparing the tubular and flat TiO2 layers with
different structures.

In all three cases, as soon as the wavelength exceeded 390 nm, the IPCE values
(respectively the generated photocurrents) approached zero. The reason for this lies in the
band gap energy of anatase TiO2 (Eg 3.2 eV [36]), that corresponds to a wavelenght of

102

approx. 390 nm. Therefore, any light with wavelength longer than that can not excite
electrons from the valence to the conduction band thus, no photocurrents are generated. The
band gap energy of the TiO2 nanotube layer can be evaluated using the Grtner model [164]
that allows one to estimate the band gap of the semiconductor by plotting (Iphhv)1/2 versus hv.
Values of Eg of approximately 3.2 eV for the annealed layers, and a slightly lower value of
3.1 eV for the amorphous layer can be revealed from the intersections of the hv axis [206]
in very good accord with the literature [159]. Evidently, the fact that TiO2 has nanotubular
morphology does not change its classical semiconducting behaviour.
Nevertheless, as evident from Fig. 4.45, the magnitude of the generated photocurrent
(resp. IPCE) is about 4 times higher for the nanotube film compared to the flat TiO2 layer.
Moreover, the as-formed nanotube layer shows only a poor photoresponse. This is in line with
expectations that (i) the amorphous structure provides a high number of defects that lead to a
high carrier recombination rate and (ii) that for amorphous nanotubes essentially only their
bottoms are active, whereas the tube walls are completely inactive due to a large number of
traps. The dramatic increase of the photocurrent after annealing indicates that by conversion
to anatase the tube walls are activated and contribute to the photocurrent [206].
Filling of the self-ordered porous alumina by a secondary material, shown in Chapter 2,
provides one of the major platforms for applications profiting from the highly ordered
structure and a high-aspect ratio. One of the most successful filling approaches for porous
alumina is based on the electrodeposition by current, or voltage pulses, sufficiently high to
trigger a high and stable nucleation rate at the pore bottom, but sufficiently short to avoid
diffusion control of the deposition reaction [48]. However, adopting this strikingly simple
approach to TiO2 nanotube layers fails due to the semiconducting nature of TiO2 and an
electric leakage across these layers. Figure 4.46 demonstrates a novel strategy for successful
filling of TiO2 nanotubes: by optimized electrochemical self-doping. The tube bottoms can be
selectively switched to a higher conductivity, while the tube walls remain highly resistive.
This allows uniform tube filling by electrodeposition from the tube bottom to the top [207].
As-grown nanotubes are amorphous and therefore significantly less conductive than
forward-biased crystalline TiO2. Selectively enhanced conductivity at the tube bottom,
(electrochemical self-doping of TiO2 by Ti3+) can be achieved by applying sufficiently
reducing conditions to the nanotube material in aqueous electrolytes [208,209]. In other words,
by applying a cathodic voltage, part of the Ti4+ from TiO2 can be reduced to Ti3+ according to
the following reaction:

103

Ti4+ + e- + H+ Ti3+H+

(4.3)

This electrochemical reduction step is accompanied by the charge compensation via


proton intercalation and thus it is strongly dependent on the pH of the electrolyte. The onset of
this reduction occurs at around -900 mV (vs. Ag / AgCl). However, at more negative
potentials hydrogen evolution will occur (below -1.2 V, the exact value depends on whether a
potential step, or scan was used to reach the potential):
2 H+ + 2 e- H2

(4.4)

Figure 4.46 Principle of the electrodeposition into TiO2 nanotubes demonstrated on a Cu


filling. After tube formation (a), selective doping of the tube bottom is performed by
electrochemical reduction of Ti in the oxide from Ti4+ to Ti3+ (b); (c) at the conductive
bottoms Cu deposition can be initiated and continuously fills the tubes.

In neutral electrolytes, the intercalation reaction (Eq. 4.3) is sufficiently slow to allow
a precise control of the reaction sequence. Additionally, the amount of H2 evolution is
significantly lower due to a higher pH of the electrolyte. By using a short cathodic step (e.g.
104

from 0 to -1.5 V) in neutral solutions (e.g. (NH4)2SO4), the reduction of the bottoms of the
tubes according to Eq. 4.3 can occur without any observable H2 evolution in the time range of
3 5 seconds. Should this step last longer, noticable H2 evolution will occur in the tubes that
will hamper subsequent tube filling. The reduction of the tube bottom results in the reduction
of the measured mobility gap to approx. 2.4 eV [210,211] making these locations behave
more metallic. This can be profitably used for an electrodeposition of the desired material at
these reduced regions.
Figure 4.47 shows an example of successful Cu electrodeposition into the tube layer
an almost complete filling of the tubes indeed can be achieved. The layers were first
cathodically polarized in 1 M (NH4)2SO4 electrolyte (potential step to -1.5V for 3 seconds),
afterwards the electrolyte was exchanged by 1.5 M CuSO4 and a current pulsing approach was
employed to electroplate copper inside the tubes [207]. XRD investigations showed the Cu
deposit was purely metallic crystalline copper (cubic) with an average grain size of 10 nm.

Figure 4.47 SEM images showing the top (a) and the cross-section (b) of a 500 nm thick
nanotube layer filled with electrodeposited Cu inside the tubes using an approach shown in
Fig. 4.46.

This approach may be used for the deposition not only of Cu, but also of various other
metals and semiconductors. These would have a potential in magnetic applications (filling of
Ni, Co etc., [43,44,48,212]), in piezoelectric devices (filling of Pb, thermally convertable to
PbTiO3, e.g. [213]) and in the solid-state sollar cells (CdS, CdTe etc. in p-n junction with
TiO2, e.g. [214]). In parallel with porous anodic alumina, this approach might be used for
making arrays of nanowires or nanorods of the deposited material after selective dissolution
of TiO2, although in this case it is more complicated to selectively dissolve TiO2 compared
with Al2O3.

105

5. DISCUSSION
In this chapter, a more detailed discussion about the results presented in Chapter 4 will
be given, supported by the existing theoretical knowledge about the flat TiO2 anodic films
given in Chapter 2 and further by comparison to aspects of porous alumina formation, both
presented in Chapter 2. In addition to the traditional electrochemical parameters that are of
crucial importance for the tube growth, such as anodization potential and presence of
fluorides, additional parameters of various nature have been found (or thought) to have
importance for the tube growth. These features will be discussed as well.

5.1 Self-organized tube growth


5.1.1 Overview
In this section, details about the tube formation mechanism and aspects directly related
with it are discussed. Particular attention will be given to different stages of the tube growth.
Considering the porous alumina formation, as shown in Fig. 1.2, a pore nucleation
starts at the defect sites of the oxide, where there is a non-uniform thickness. The pore growth
takes place via an inward dissolution, assisted by the electric field. The formation of the
porous layer begins with a worm-like structure, which develops into an ordered one by a selfequilibration of the different competing processes. The stabilization of the growing pores
depends on the penetration ability of the different anions and the local oxide solubility and the
field strength in the oxide [36].
In the case of TiO2 nanotube formation, the process is similar to porous alumina. It
starts with a formation of the compact TiO2 layer and it ends up with a self-organized TiO2
structure in this case nanotubes. In the most simplest situation, this can be schematically
presented by two distinctly different regimes shown in Fig. 5.1. At first, the high-field
mechanism described in Chapter 2 is operative for the TiO2 formation, as shown in Fig. 5.1a.
The oxide layer is grown by hydrolysis of Ti via the electric field assisted migration of Ti
cations from the Ti / TiO2 interface towards the TiO2 / electrolyte interface, according to Eq.
2.3. According to the available literature [94,160,161], the transport number of number of Ti4+

106

is approximately 0.4, thus 0.6 for O2-. This, however, takes place only at the very beginning of
the anodization (the first seconds up to a few minutes, depending on the potential sweep rate).
At a later stage, due to a presence of fluorides in the electrolyte, field-aided TiO2
dissolution can occur that is accompanied by more, or less significant chemical TiO2
dissolution (depending on the electrolyte pH), as in Eqs. 2.14 and 2.15. In the fluoridecontaining electrolytes not only oxygen anions are moving towards the Ti / TiO2 substrate, but
also flouride anions, as recently stressed in a paper by Habazaki and coworkers [215]. They
used a Si marker during Ti anodization in fluoride containing electrolytes (although
unfortunately performed at comparably high anodic voltages).

Electrolyte
H2 O H+

Electrolyte

Ti(OH)xOy

TiO2

[TiF6]2U
F=
d

O2-

F-

Ti4+
TiO2

O2-

F-

etching
Ti4+
oxidation

Ti

Ti

Figure 5.1 Schematic representations of two basic morphological states in the nanotube
formation. At the very beginning, a flat TiO2 layer is formed (a); due to presence of fluorides
in the electrolyte and consequent TiO2 dissolution (via complexation of Ti4+ with F- species),
nanotube morphology is obtained.

However, Fig. 5.1 represents only the initial and the final stage of the tube formation.
Now lets consider the different intermediate stages. In Chapter 4, a sequence of SEM images
was made from the samples that were removed from 1 M (NH4)2SO4) / 0.14 M NH4F
electrolyte at different anodization times (Fig. 4.17). On the basis of these images (and dozens
of similar ones), one can deduce that the initiation and growth of the self-organized TiO2
nanotubular layers are essentially following the scheme shown in Fig. 5.2 and are
accompanied by significant changes in the current density, as shown in Fig. 4.15.
For simplicity, let us assume first anodization with a high sweep rate (say, 1 V / s). As
shown in Fig. 5.1, during and just after the potential sweep, a TiO2 (and also hydroxide) layer
is grown to a thickness of approximately 50 nm that corresponds to a compact high-field
oxide expected at 20 V (using a growth factor of 2.5 nm / V, see e.g. Ref. [19]). It is known

107

from the literature that the OH- / O ratio - depends strongly on the ramping speed [154,155].
The potential sweep is accompanied by a rapid current increase, followed by rapid current
decrease, after it has reached the final potential of 20 V, in line with the kinetics of the oxide
layer growth described in Chapter 2.

Figure 5.2 Schematic representation of TiO2 nanotube layer formation. The black color
represents a compact TiO2 from the initial stage, the gray color represent an oxide used to
build up the pores and the nanotubes.

In parallel, the influence of the F- species comes into play in the form of localized
dissolution and breakdown / repassivation events in the oxide layer (Fig. 5.2a) accompanied
by the remarkable current increase (see Fig. 4.15). It is very likely that these events do not
occur randomly on the surface, but at defect sites of the oxide, related to the development of a
non-uniform thickness and presence of stresses. The breakdown sites act as seeds for the
growth of a disordered worm-like porous structure underneath the remaining compact layer
(Fig. 5.2b). In the course of time, due to an earlier onset of dissolution at more defect sites, the
108

pores grow deeper and bigger, compared with the rest of the area (less defective), where the
pores are smaller and shallower (Fig. 5.2c). With an increasing number of pores, the effective
surface area increases and so does the current.
At this particular stage of the growth, two important effects start to play a significant
role: (i) the pore depth and (ii) the pH profile in the pores.
First, as explained above, deeper and bigger pores grow initially on the account of
smaller and shallower pores that are formed later during the anodization [216]. The deeper
pores have a larger area available for the electrochemical oxidation of substrate than the
shallower pores. Considering this two-fold morphology and presence an applied potential, the
oxidation front (as well as dissolution front due to fluoride etching) is at the tip of the deeper
pore bottom. This is the area that is the closest to the metal substrate and therefore these
reactions are going to occur preferentially there.
Secondly, there is a pH-gradient established in the pores, as a result of the dissolution
front. In other words, due to field-aided oxidation reactions occuring in the pores, a localized
acidification of the pore bottoms takes place due to the presence of H+ protons. As a
consequence of the lower pH at these locations, an accelerated TiO2 dissolution and further
pore penetration into the Ti substrate takes place. As a result, the deeper pores become even
deeper. On the other hand, the shallower pore generates a smaller amount of H+ ions, and its
growth becomes slower and slower. Finally, the weak pore stops to grow when all of the
neighboring metal ions are already oxidized by the near strong pores. This pore selection
process is summarized by the autocatalysis sequence shown in Fig. 5.3. The deeper the pore,
the larger reaction area it has. A deeper pore induces a formation of more H+ ions. The large
amount of H+ ions accelerates the dissolution of oxide, and results in an accelerated
advancement of the deeper pore [216]. The combination of these steps results in the selforganization of the pores. A similar autocatalytic system was already reported in the porous
alumina formation [36].
Thus, there is a natural selection process: only the deep and ``strong pores can
compete with one another for the available oxidizable area and the current. For the other pores,
it is very diffucult to survive and under critical conditions, the growth of these pores is ``shut
down (see the different pore proportions and the depth evolution between Fig. 5.2c and
5.2d).

109

Deeper pore

Larger reaction area

Faster growth

Larger H+ production

Figure 5.3 Autocatalytic process during the initial stages of the tube growth (according to
Refs. 36 and 216).

After this natural selection process, the formation of the worm-like structure finishes
and all of the remaining pores grow straight (Fig. 5.2d). In this stage, a self-organization of
the pores starts to be established approximately 20 minutes from the beginning of the
anodization in all electrolytes that contain a significant amount of water. In sulfate / fluoride
electrolytes (such as electrolytes consisting of corresponding ammonium salts used for
construction of Fig. 5.2) the remnant initial oxide layer is still present on the layer that is
completely dissolved after approximately 1 hour (or detached from the surface). In other
electrolytes, such as in the glycerol / water / 0.27 M NH4F mixture (as in Fig. 4.25) this oxide
layer disappears earlier during the growth (max. 20 minutes), most likely due to a higher
fluoride content that might be responsible for faster removal of these parts.
Approximately 30 minutes from the beginning of the anodization, a completely selforganized nanotubular layer is obtained (Fig. 5.2e). The thickness of the layer increases
within the next few hours of anodization until a certain limiting thickness value is reached.
The exact time to reach this point varies with the electrolyte. For instance, in the 1 M Na2SO4
/ 0.14 M NaF, or in 1 M (NH4)2SO4 / 0.14 M NH4F, where the nanotube formation proceeds
at 20 V (with a sweep rate of 0.1 V / s), the nanotube layer thickness almost linearly increases
during the first three hours of anodization and then the growth becomes slower. The steadystate thickness of approximately 2.5 m is reached after 10 hours (max.) of anodization, as
shown in Fig. 4.19b. For organic electrolytes, it takes longer until the steady-state nanotubular
layer thickness is attained. However, upon extensive anodization time (more than 12 hours)
the nanotube thickness increases only of few %. Nevertheless, for all electrolytes used in this
work, the growth sequence shown in Fig. 5.2 is valid. The only remarkable difference is the
speed of transition from one stage to another stage, a parameter that is strongly dependent on
the fluoride concentration in the electrolyte and the sweep rate (this will be discussed later).
110

The increase of the nanotube layer thickness is always accompanied by a gradual


current decrease, as shown for example in Fig. 4.15 and 4.19. This would strongly suggest
that the current is diffusion limited (Eq. 2.1). More details about the diffusion are discussed
later in the section Diffusion.
The fact that the layer thickness and the current density reach a limiting value after a
certain polarization time can be explained by a steady-state situation depicted in Fig. 5.4.
During anodization, permanent growth of the oxide layer takes place at the inner interface (Ti
/ TiO2), and dissolution of the oxide layer occurs simultaneously at the outer interface (TiO2 /
electrolyte) and outside the tubes. The steady-state situation is established when the tube
growth rate at the metal / oxide interface at the bottom v1 (or, say, the rate of inward
movement of the metal / oxide interface) is equal to the dissolution rate of the oxide (v2) at the
tube mouth of the film. In this situation, the nanotube oxide layer just permanently penetrates
deeper and deeper into the titanium substrate, without further thickening. As the steady state
current densities are typically considerably high (~ 1 mA / cm2), this occurs with an almost
constant velocity.

Figure 5.4 Steady-state nanotube growth situation characterized by equal rates of TiO2
formation (v1) and dissolution (v2).

111

It should be noted that the chemical dissolution of TiO2 occurs to a certain extent over
the entire tube length, particularly in the HF contaning electrolytes. Thus the tubes with time
become increasingly V-shaped in morphology [217,218], as seen for instance in Figs. 4.3a
and 4.35. In other words, at the top, the tubes possess a significantly thinner wall than at their
bottom. Thus the tubes in Fig. 5.4 also have a V-shape.
One should mention that the mechanism shown here does not provide an exact
reason, why the tubes are formed instead of pores. The reason for separation into tubes, as
opposed to a nanoporous structure, is not yet entirely clear. Attention to this point is provided
in the section Stress effects.
It is generally accepted that the thickness of a compact TiO2 layer is determined
mainly by the applied potential, as discussed in Chapter 2. For porous alumina, it is clear that
its growth is strongly related to the formation potential, selection of the right electrolyte and
the thickness of the layer is time-dependent. From all results given in Chapter 4 it is evident
that all these parameters are also of crucial importance for the TiO2 tube growth and
morphology. However, another important parameter which comes into play to grow
nanotubes is the presence of the right amount of fluorides in the electrolyte. In the next
sections, the influence of anodization potential and fluoride content are discussed in detail.

5.1.2 Influence of potential


For the discussion about the potential influence on the tube growth, let me first
summarize some facts. An applied (anodic) potential influences the electric field strength (F)
across the growing oxide (Eq. 2.5). The arisen current density is exponentially proportional to
the field strength across the oxide layer (Eq. 2.4). The electric field across the oxide layer has
the vital importance for the transport of ionic species through the oxide and thus it is
responsible for the nanotube growth that also requires permanent oxide dissolution. In order
to maintain a continuous, non-disturbed growth of a nanotube layer, the electric field should
be maintained as stable, as possible. Common to all anodic treatments shown here, or reported
elsewhere, is that the nanotube layers are achieved by potentiostatic polarization, typically by
a defined ramping of the potential from the open-circuit potential (OCP) to the constant
potential value, or, less frequently, by a potential step to a desired anodization voltage.
Galvanostatic anodization (under constant current conditions) appears up to now not suitable

112

for the nanotube growth, as it may lead to significant voltage oscillations and destruction of
the nanotube layer [219].
According to the previous reports on the nanotube growth before this work was started
[20,21,118,119], all nanotube growth attempts shown in Chapter 4 of this work and reports of
other groups [220-223], the most suitable anodization potential for obtaining a homogenous
and well-ordered nanotube layer in aqueous electrolytes is 20 V. There are exceptions, where
other researchers have succeeded to synthesize nanotube layers also at lower potentials, e.g. at
10 V in purely HF-based electrolytes [118], or 5 V in chromic acid / HF - based electrolytes
[20], or in HF / H3PO4 mixtures even at as low potential as 1 V [198]. A remarkable feature of
these nanotubes is that they have a smaller tube diameter than those prepared at 20 V (on
average 100 nm). In anhydrous, or partly aqueous electrolytes, the potential range, in which
the nanotube growth can be achieved, was recently extended, as shown in this work (Chapter
4) by use of mixed glycerol / water / NH4F electrolytes [197], ethylene glycol / HF [201] and
mixed ethylene glycol / water / NH4F electrolytes [224].
In order to get a better understanding about the influence of anodic potential on the
tube diameter, one may consider first an anodization of Ti in the absence of fluoride anions in
the electrolytes resulting in a compact oxide formation. As presented in Chapter 2, for anodic
TiO2 films usually values of 2.5 nm of thickness gained per applied anodic volt are reported
(see Ref. [19] and refs. therein). Let us assume first that the nanotube is grown in one step
without any initial compact oxide layer and dissolution. The maximum oxidation length
should follow the growth rate of compact oxide formation [216]. As indicated in Fig. 5.5a, a
schematic drawing of the tube bottom, this value appears as the radius of a hemisphere for the
oxidizable area. The frontier spot for anodization is the tip of the tube bottom. As evident
from Fig. 5.5a, the diameter of the hemisphere is two times as large as the oxidation length,
i.e., ~ 5 nm / V. Thus, the final diameter of the nanotube is closely related with the growth
factor [197,216]. It is noteworthy that the formation of porous alumina arrays follows
basically the same oxidation behavior, as in Fig. 5.5a, i.e. the pore distance has the same
meaning as the nanotube diameter since it follows the two times relationship (see e.g. Ref. [13]
and references in there). The relationship for porous alumina strongly supports that the final
diameter of nanotubes basically becomes two times as large as the oxidation length.

113

a)

b)

Figure 5.5 (a) schematical representation of the tube diameter determination [taken from Ref.
216]; (b) experimentally determined dependence of the tube diameter, the bottom layer
thickness and the tube layer thickness on the applied potential (evaluated from SEM images,
all distance values in nm, potential (U) in V, taken from Ref. [197]).

Indeed, exactly this diameter vs. potential relation is revealed in the real nanotube
growth situation. Fig. 5.5b shows the experimentally obtained dependences of the tube
diameter, bottom layer thickness and the tube length with respect to applied potential [197].
All values are determined from SEM images of the nanotubes grown in glycerol / water
(50:50 vol.%) / 0.27 M NH4F electrolytes (Figs. 4.26 4.28). The constants (factors) for these
relationships fit very well to those of nanotubes obtained in other aqueous electrolytes. Some
deviations are found for i) the tube diameter of nanotubes grown in organic electrolyte
(discussed later) and ii) for the tube length of nanotubes grown at potentials higher than 20 V.
Considering the tube bottom layer thickness (Fig. 4.28), the growth rate of 2.2 nm per
applied volt has been revealed (Fig. 5.5b). This is a bit lower than the growth rate of a
compact film ( 2.5 nm / V) and this difference is not entirely negligible, particularly when
considering the difference at quite high potentials. The lower oxide growth rate might be
explained in terms of increased dissolution rate of the oxide layer at these locations. As
explained earlier in the text, the nanotube growth occurs essentially via permanent field-aided
oxide formation (at the interface between the TiO2 and Ti substrate) and a combination of
field-aided and chemical TiO2 dissolution (at the oxide / electrolyte interface). Thinner tube
bottoms (relative to theory) might have origin in the fact that a part of the current goes to
permanent oxide layer dissolution, thus the bottom layer cannot reach its full expected
thickness.
114

According to the diameter vs. potential relationship, nanotubes grown at 20 V should


have an average diameter of 100 nm. This is indeed the case for nanotubes grown in aqueous
electrolytes or mixed glycerol / water electrolytes (50:50 vol. %). In organic electrolytes,
typically higher voltages are reported to grow tubes of a given diameter. For instance, in
glycerol / NH4F electrolyte a diameter of ~ 40 nm is obtained at 20 V (Fig. 4.21). At the
present, these lower values (compared to the aqueous electrolytes) must be attributed to the
very significant IR drop that occurs in these electrolytes due to their low conductivity (high
resistivity). Table 5.1 shows the conductivities and respective IR drops measured for various
glycerol-based electrolytes and also others used in Chapter 2 (taking 1 mA / cm2 as a standard
current density value accompanying the tube growth, if not denoted otherwise). As can be
seen, the conductivity drastically drops with decreasing water amount. This means that with
the same electrode architecture, there is a much larger IR drop (potential loss due to the
electrical resistance of the electrolyte) between the reference and working electrode. As a
natural consequence, the tube diameter becomes smaller. This can be demonstrated with the
water-free glycerol / NH4F example. The IR drop is 13.3 V, that means, when anodizing at 20
V, the real potential on the electrode is 6.7 V. By multiplication with a factor of 5 (Fig. 5.5),
the diameter results in ~ 34 nm. This is well in accord with the experimentally obtained value
of ~ 40 nm.
Table 5.1 pH values and calculated IR drops for different electrolytes
Electrolyte
Water
content
(vol.%)

c [F-]
(mol/l)
0.0
0.0*
0.135
0.27
0.54
0.27
0.27
0.27
0.27
0.27
0.14
0.14
0.16
<0.16

pH
3.0
5.0
5.5
5.7
6.2
6.6
6.5
6.1
6.0
5.8
5.6
7
1.6
2.7

Conductivity
(mS/cm)
0.02
6.1
1.7
3.4
6.9
0.44
0.21
0.10
0.095
0.09
117
72
0.07
0.3

50
16,7
6.7
3.7
0.67
0
1 M (NH4)2SO4 / NH4F
(100)
1 M Na2SO4 / NaF
(100)
Ethylene glycol / HF
~ 0.35
Ethylene glycol / HF aged***
~ 0.35
* instead of H2O, 0.5 M (NH4)2SO4 was used.
** IR drop calculated using standard current denisty of 1 mA / cm2, distance of
electrodes 1.2 cm and electrode area 1 cm2
*** IR drop calculated using standard current density of 5 mA / cm2
glycerol / NH4F electrolyte

IR drop**
(V)
60
0.2
0.7
0.35
0.17
2.7
5.7
12
12.6
13.3
0.01
0.02
86
20

115

5.1.3 Influence of Fluorides


As mentioned in Chapter 2, fluorides belong to the few species able to dissolve TiO2.
It is evident that they have a vital role in the tube formation (together with the applied
potential), the right amount of fluorides is needed for the growth of ordered nanotube layers.
According to Figs. 4.7, 4.12, the fluoride concentration for the tube growth in purely aqueous
electrolytes has a relatively narrow range between 0.1- 0.3 M F-. Outside of this range, no
ordered nanotubes can be formed. It should be noted that nanotube layer can be grown at
higher concentrations (e.g. 1.5 M in Fig. 4.16), but it requires first the use of a very slow
potential ramping and secondly, the nanotube layers are significantly covered with a kind of
precipitation (the AES measurements show it is a complex of the electrolyte species with
hydrated initial oxide layer).
For purely organic electrolytes, the concentration range can be extended towards
higher concentrations (up to 0.5 M F-) while still growing ordered nanotubes, but is difficult
to prepare these electrolytes, due to the limited solubility of fluorides in them. Typically, the
more fluorides, the longer these solutions must be heated prior to anodization and they can
undergo decomposition (as evident from the change of color and smell) and become useless
for anodization.
At present it seems that the electrolyte that enables growing ordered and homogenous
layers in the widest possible F- concentration range is composed of glycerol, water and NH4F.
As shown in Fig. 4.29, in these electrolytes with different F- concentration, the nanotube
layers can be grown under otherwise identical conditions, that is, without the necessity to
tailor other anodization parameters. As further shown in Fig. 4.30 the current density
increases with increasing F- concentration. This is completely in agreement with the theory
described in Chapter 2, (increasing F- content in the electrolyte causes the current density to
increase). In the light of the results presented here, this is not surprising, since the field-aided
dissolution (driven by F-) is accelerated by increased fluoride content and thus leads to the
growth of thicker nanotube layers. This is only true in the neutral fluoride-containing
electrolytes. As demonstrated in Fig. 2.12, in acidic HF electrolytes with too high Fconcentration (0.4 wt. % corresponds to approximately 0.2 M HF), no nanotube growth can
be achieved.

116

It is noteworthy to mention that XPS and SEM analysis of layers anodized from 0 V to
5 V in F--free and F--containing (0.54 M NH4F) electrolytes show that in both cases a compact
oxide layer of approx. 10 nm in thickness has been formed. However, for the layer built in F- containing electrolytes F species can be detected throughout the sample. Therefore, the higher
currents observed in the initial stage in F- -containing electrolytes are either related to the
faster field-aided F- transport through the growing oxide layer (in comparison with competing
with O2- transport), or are due to the nano-scale porosity (beyond the resolution of FE-SEM)
in the oxide grown in fluoride electrolytes. The first theory would be in line with the recent
study of Habazaki et al. [215] that claims that the fluorides are moving twice as fast as the
oxygen anions (in other words, they have a higher transport number) through the anodically
grown TiO2.

5.2 Other factors influencing growth of the nanotube layers


As evident from the results shown in Chapter 4, the growth of self-organized TiO2
nanotubes is a very complex process and except for the primary factors potential and
fluorides it is influenced by a whole range of other factors. These are discussed here,
namely: the influence of pH, stress effects, temperature and diffusional effects. Surprisingly,
the influence of the initial compact oxide layer is not as important as some may expect.

5.2.1 pH

The following part deals with a more detailed description of the pH variations that
occur during the nanotube growth and are believed to have a significant influence on the
nanotube dimensions.
Based on the growth mechanism presented in Fig. 5.2 (deduced from the theoretical
background and experimental findings presented in Chapters 2 and 4, respectively) it can be
concluded that the field-aided TiO2 oxidation (on the interface between the Ti substrate and
growing TiO2) and dissolution (on the interface of TiO2 and electrolyte) are operative
simultaneously. In addition to these reactions, pH-dependent chemical TiO2 dissolution along
the tube walls and in the tube surrounding plays a significant role in the nanotube formation,
as evident from a difference in the thickness between the nanotube layers grown in acidic HFbased electrolytes and neutral aqueous, or organic fluoride-based electrolytes. In other words,

117

during anodization in an acidic environment and in the presence of F- species, the TiO2
etching is simply too fast to enable growth of high aspect ratio nanotubes. In contrast, in
neutral aqueous or organic electrolytes, where HF is replaced with fluoride salts, the TiO2
etching rate is significantly smaller. This in turn enables the nanotubes to grow substantially
longer.
The field-induced hydrolysis of Ti (resulting in an oxide formation) is accompanied by
a generation of acidity (Eq. 2.3). Under an externally applied potential (field) and in the
presence of water, this reaction leads to a permanent acidification of the reaction spot.
Considering the tube growth (under conditions described earlier in Fig. 5.2), this reaction,
together with simultaneously occuring TiO2 field-aided dissolution, leads to a gradual
reshaping of the initially flat oxide to a porous oxide at early stages and finally to a
nanotubular structure. When considering the steady-state tube growth (a phase when the
nanotube layer is already formed and does not change its thickness significantly), this reaction
goes on permanently at the tube bottom and therefore a permanent localized acidification of
these locations takes place. As explained in Fig. 5.4, the autocatalytic nature of this reaction
maintains this self-induced acidity situation during the whole anodization process.
Therefore, in the case of non-HF-containing electrolytes (that have pH in the range of
5 to 7), it can be expected that there is a difference between the pH at the tube bottom and the
pH in the bulk electrolyte (a sink for reaction products and a source of fresh reactants). In
other words, there might be a pH gradient established inside and outside the nanotube layer,
as a result of mixing these opposing species. It is unfortunately not possible to measure the pH
gradient inside a single nanotube with 100 nm diameter. However, it is possible to introduce a
simple model that takes in account the current (that flows through the nanotube layer during
the anodization) and considers the pH of the bulk electrolyte.
In order to estimate the pH at the tube bottom, a concentration profile of H+ species is
outlined in Fig. 5.6. For the construction of this profile, essentially several assumptions are
put forward that include one-dimensional flux of the species (transport only in the direction of
the tube axis), linear diffusion and no influence of other species from the electrolyte during
the tube growth. Then the flux of the H+ species from the tube bottom to the bulk electrolyte
can be expressed using a difference in the equilibrium concentrations of H+ at these distinct
locations. For that it is necessary to consider the Ficks first law of diffusion that is expressed
as [225]:
J = D

dc
dx

(5.1)

118

where J is the flux of the species (number of particles passing through an imaginary window
in a given time interval), D is diffusion coefficient, dc is the concentration gradient over a
distance dx.

Tube
CT
H+ concentration

Bulk

CEL
Distance x from the tube
Figure 5.6 Simplified profile of the H+ concentration within the tube (CT) to the bulk
electrolyte (CEL).

The concentration of H+ species at the tube bottom can be deduced from the current
density (or charge density, respectively) that flows through the sample. Since the field-aided
hydrolysis is a four-electron reaction, it leads to a generation of four protons (H+, according to
the stoichometry of Eq. 2.3). Now, lets consider for instance anodization in 1 M (NH4)2SO4 /
0.14 M NH4F electrolyte that has a pH value of 5.6. By using the steady-state current density
of 1x10-3 A / cm2, charge of an electron q 1.602x10-19 C, the number of transported
electrons (and protons H+) can be obtained (6.25x1015 cm-2). By using Avogadros number of
NA 6.022x1023 mol-1 (expressing the number of particles in 1 mole of a substance), the H+
flux of ~ 1.038 x10-8 mol / cm2 can be obtained. Furthermore, by using Ficks first law (Eg.
5.1), calculated H+ flux and previously published diffusion coefficient of H+ (DH+ 7.8 x10-5
cm2 / s, taken from Ref. [226]), the pH at the tube bottom can be estimated via the bulk pH
across a certain distance (pH is equal to the negative decadic logarithm of the H+
concentration [225]). In fact, this distance represents a sum of the tube layer thickness and a
width of the mixing zone, where acidic species generated at the tube bottom (and leaving the
tube) are buffered by a more neutral species from the electrolyte driven by the natural
convection, or a flux arisen from the electric field. By using the nanotube layer thickness of
2.5 m and different widths of the mixing zone (10, 50, 10, 150 and 200 m), the pH at the

119

tube can be calculated that is as low as ~ 2.5. The fact that the tube bottom gets acidic during
anodization and that there is a relatively wide mixing zone was further confirmed by
experimentally measured pH values of small electrolyte portions taken form the vicinity of the
samples (as shown for example for the 1 M Na2SO4 / 0.14 M NaF electrolyte in Fig. 4.11) that
were always lower than the bulk pH of the electrolyte. For the 1 M (NH4)2SO4 / 0.14 M NH4F
electrolytes the difference was not so big (pH 0.5). This is likely either due to a higher
buffering capacity of the NH4+ species, or because the electrolytes bulk pH value of ~ 5.6 is
lower compared to pH ~ 7 of the Na-based electrolyte, or it is a result of both effects.
As evident from the results shown in Chapter 4 (nanotubes can grow comparably
longer in neutral electrolytes compared to the acidic ones and also considerably less Ti
material is consumed), it is very likely that the pH profile calculated above is established
during the anodization. This is due to a gradual mixing of the low pH electrolyte (attempting
to leave the tubes) and the higher pH from the bulk electrolyte (attempting to enter the tubes),
there is also considerable chemical TiO2 dissolution (highly dependent on the pH, see Fig. 4.5)
of the tube walls along the whole tube length, not only the tube bottom. Otherwise,
theoretically unlimited thickness of the nanotube layer could be achieved. Since the upper
parts of the nanotubes (tube mouths) have been exposed to the more dissolving environment
(established by a flux of H+) for the longest time, they show also the biggest TiO2 loss. In
other words, the tube walls are thinner at the top than at the tube bottom, as shown on several
examples in Chapter 4.
Considering all aspects mentioned above (a numerical simulation of the relevant ion
fluxes and the pH measurements), the pH profile of the tube can be constructed, as shown in
Fig. 5.7 [192]. It is again made for 1 M (NH4)2SO4 / 0.14 M NH4F with a staring pH of ~ 5.6.
By using this as a buffer and in the same time fluoride-containing solution as an electrolyte
and adjusting the anodic current flow to an ideal value, the acidity (low pH environment) can
be created where it is needed, that is, at the pore tip. The pH there can be as low as ~ 2.
Higher pH-values (in the range of 4-5) are established at the pore mouth, due to migration,
diffusion and gradual mixing of the pH buffering species [NH4F, (NH4)2SO4)] with the acidic
species. Assuming equilibrium, the flux of dissolving species (leading to acidification at the
pore tip) and the flux of buffering species are equal. It should be noted here that the bulk
electrolyte pH at the tube mouth corresponds to a drop in the chemical etch rate of about 20
times.

120

Figure 5.7 Tuning the electrochemical conditions to achieve high aspect ratio nanotube layers:
the field-aided reactions occur at the tube bottom (left), the pH profile within a tube (middle)
and dissolution rates (vdiss) along the tube wall [192].

This pH profile is the key to achieve growth of the high aspect ratio nanotubes. By
tailoring the localized acidification at the tube bottom and thus promoting the TiO2 dissolution
there, while having a more protective environment against the dissolution along the tube
mouth, longer tubes are obtained. The fact that TiO2 is soluble in HF forming [TiF6]2- is on
the one hand essential for the tube formation, on the other hand, it is responsible for the fact
that previous attempts to form the nanotube layers in HF electrolytes always resulted in layer
thicknesses in the range of only a few 100 nm.
At this point, it is necessary to note that the complex TiO2 dissolution rate during the
tube formation can be tuned by the dissolution current. As shown in Fig. 5.7, to establish the
desired pH-profile, a steady state current situation needs to be achieved. In the present work a
voltage sweep technique was used. As shown in Fig. 4.15, the ramping speed (potential sweep
rate) has a huge impact on the current density not only during the polarization (consistent with
121

different electric fields driving ion transport during sweeping), but more importantly, it has an
impact on the current also in the potentiostatic period of the experiment. This implies that
there must be some difference between the as-grown nanotube layers. Figure 5.8 shows the
nanotube layer thickness evaluation, based on SEM images taken from samples anodized in
glycerol / water (50:50 vol. %) / 0.27 M NH4F at 20 V with different sweep rates. As can be
seen, the thickness of the nanotube layers indeed varies significantly with the sweep rate. It
should be noted that differences in the tube layer thickness with different sweep rates are
observed for all electrolytes.

Figure 5.8 Nanotube layer thickness of samples grown by anodization in glycerol / water
(50:50 vol. %) / 0.27 M NH4F at 20 V for 3 hours with different sweeping rates.

In the light of the simulations presented in Figs. 5.6 and 5.7, the dimensional differences,
influenced by the ramping speed, can be associated to a degree of the acidification of the tube
during anodization leading to an enhanced or retarded etching of the side walls of the tubes.
These results thus imply that the mechanism of tube formation is somewhat different from the
relatively well investigated case of the porous alumina [1-5,36]. While in the alumina case,
the field aided ion transport through the pore bottom is the dominant mechanism (Fig. 2.2)
and chemical oxide dissolution plays a minor role, it is evident that in the case of TiO2
nanotube growth the chemical effects can become highly significant.

122

5.2.2 Stress effects


Self-ordering of the nanotubes must also be related to stresses induced by the tube
growth. Presumably, the forming tube bottom TiO2 layer at the interface between the
nanotubes and Ti substrate is permanently exposed to a compressive stress, due to at least two
reasons. Firstly, the lowest part of each tube tends to maintain its diameter (shown in Fig. 5.5),
therefore the oxide volume that builts up this part of the tube is permanently under stress.
Once the nanotubes are formed, this stress is more or less equal in all nanotubes, since each
nanotube is fighting for available space with all other neighboring nanotubes. Secondly, the
stress could be assigned to a significant TiO2 expansion during the oxidation of Ti metal
parameter presented in Chapter 2 as the Pilling-Bedworth ratio (Eq. 2.13, for amorphous TiO2
~ 2.4). Thus an internal stress is present in the growing oxide also in the case of oxide
nanotubes. It can be assumed that the contribution of electrostriction (Eq. 2.14-2.16) to the
total generated stress is only minor, based on the work of Seo et al. [100].
As mentioned earlier in this Chapter, the tube growth mechanism does not explain,
why there are gaps in between the tubes. The intention of the following text is to provide a
possible reason for the gap formation [216]. It is likely that the main factor for the generation
of gaps is the mechanical stress in the nanotubes. In order to discuss the stress at the nanotube
layer / Ti substrate interface, the volume expansion is schematically drawn in Fig. 5.9, based
on the marker experiments by Garcia-Vergara et al. for the porous alumina formation [227].
Since the nanotubes do not have a chance to expand their volume horizontally and to release
the stress of each tube into the other tubes (stress due to a volume expansion), an assumption
of upward one-dimensional expansion in the vertical direction to the substrate can be made. In
Fig. 5.9, five marker lines from (1) to (5) are drawn in the Ti substrates before anodization
with a constant interval [216]. The broken and the solid lines show the positions of the Ti
atoms before and after oxidation, respectively. According to the Pilling-Bedworth ratio (Eq.
2.9), significant volume expansion is generated during the oxide growth, thus the atoms move
to the positive direction of the Z-axis. For the line (2) that represents an initial stage of
oxidation, the oxidation starts at the bottoms of the tubes, and the solid line slightly bends to
Z-axis. With a progress of oxidation, lines (3) and (4), the bending becomes larger due to the
increase of oxidized length in the Z-direction, and the bend location shifts to the center
between the two tubes. As the curvature of the solid lines show, the largest stress is generated
at the tube / oxide interface. The stress reaches the maximum at the center between the tubes

123

as shown with line (5). The primary oxidation proceeds mainly on the side of the tube bottom.
In other words, the Ti atoms located on the side from the gap are oxidized from the same sides,
as indicated with the arrows in Fig. 5.9. Therefore, the center between the two tubes can be
oxidized by either of the neighboring tubes, i.e. this location has two different oxidation
directions, and considerable stress is generated at this center. The generated stress very likely
leads to a generation of gaps between the pores, resulting in a transformation of the
morphology into a tubular one.

(5)
(4)

Direction of
oxidation

(3)
(2)
(1)
Figure 5.9 Schematic drawing of the volume expansion at the metal/oxide interface during
the oxide tube formation (taken from Ref. [216]).

Apart from the stresses present between the nanotubes, the tube bottoms, as such, are
essentially centers with electric fields present across the bottom oxide layer. The field strength
can be as high as 107 V / m [67]. As proposed by Nelson and Oriani [94], the transport of the
species across the growing oxide film may also have influence on the stress generation. Based
on the already mentioned paper of Habazaki et al. [215], the fluoride anions move
approximately twice as fast as oxide anions into the growing oxide. This means that the
higher the applied potential, the deeper (and faster) the fluoride anions move towards the Ti
substrate. As a consequence, accumulation of the fluoride species may occur on the tube
bottom to a certain extent. This would lead to a formation of TiF4 at the Ti substrate /

124

nanotube interface. The TiF4 would be responsible for the poor adhesion of the nanotube layer
to the Ti substrate, as in the case of TaF5 [161]. This would be well in line with the finding
that some nanotube layers formed at the highest potential shown here (40 V) occasionally
undergo detachment from the underlying substrate. The detachment does not occur for the
nanotube layers grown at lower voltages.
Apart from the poor adhesion, the fluoride accumulation at the tube / Ti interface
could lead to an establishment of a more soluble TiO2 structure, that is, as soon as it becomes
the part of the bottom accessible to the electrolyte, it is relatively easier to be dissolved than
pure TiO2. As shown in Fig. 4.44, the fluoride content after the tube formation can be
significantly lowered upon annealing. However, the remaining F traces could also be part of
the TiF4 at the tube layer / Ti interface.

5.2.3 Diffusion
From the results presented in Chapter 4, in particular for the viscous glycerol, or
ethylene glycol electrolytes, it is evident that diffusion has a huge impact on the morphology
and dimensions of the nanotubes, because these electrolytes form thicker and a more ordered
nanotube layers than the aqueous electrolytes. This must be ascribed to a comparably slow
diffusion of the chemical species during anodization in viscous electrolytes, accompanied by a
lower current density, resulting in a lower rate of the local tube bottom acidification and a
lower rate of competing chemical and field-aided TiO2 dissolution.
A straight-forward concept to prove the diffusion limitation of the nanotube growth is
to introduce electrolyte stirring [52,63,228,229]. In this limiting case, the relationship between
the limiting current density (jL) and the width of the diffusion layer (so-called Nernstian
diffusion layer) can be expressed by following equation:
jL =

zFDc

(5.2)

where F is Faraday`s constant and z in number of exchanged electrons and D is diffusion


coefficient and c is concentration. If the diffusion limitation of the reaction takes place, the
surface concentration of the reacting species is zero as all arriving ions react immediately.
This means that the current density becomes voltage independent and depends only on the
width of the diffusion layer (), diffusion coefficients (D) of the species involved and the bulk
concentration of anions (c). The width of the diffusion layer varies significantly with the

125

degree of the agitation of the electrolyte. In many electrochemical kinetic studies, the limiting
current density can be varied by changing of the hydrodynamics of the solution, for instance
by using rotating disc or ring electrodes with defined hydrodynamic conditions [52,228].
Fig. 5.10 plots a comparison of polarization curves and current transients recorded
during anodization of Ti in a F-free and F-containing electrolyte. It can be seen that if stirring
is switched on and off at different stages, the current always sharply increases and rather
slower decreases, respectively, in the case of the F-containing electrolyte (leading to the
nanotube layer growth). In the case of the fluoride-free electrolyte (compact oxide forming)
stirring has no effect and only an exponential decay of the current density can be observed, in
line with the theory described in Chapter 2.

Figure 5.10 Polarization curves (left) and current transients (right) recorded during
anodization of Ti in water and glycerol (50:50 vol.%) with 0.27 M NH4F (leading to a
nanotube TiO2 layer) and in 1 M H2SO4 (leading to a compact TiO2 layer), stirring of the
electrolyte (300 rpm for 2 min) has been employed several times, as marked in the figure.

In line with the current transients recorded in various electrolytes shown in Chapter 4,
the anodic current gradually decreases with time as the tube morphology is being built and the
tube length is prolongated. In other words, considering Eq. 5.2, the current decreases as a
result of increasing diffusion length for the ionic species in the electrolytes, as the nanotube
layers grow in thickness. It would be good to gain more insight into this issue.
According to the nanotube growth mechanism presented in this thesis, there are only a
few processes that have the potential to be the rate-determining step in the nanotube formation.
These include: i) field-aided transport of the ions through the growing oxide; ii) field-aided

126

dissolution of the oxide, iii) transport of fresh reactants into the nanotubes; iv) transport of the
reaction products out of the tubes.
In order to evaluate, which of these process(-es) is the rate-determining step in the
steady-state stage of the nanotube growth, it is very helpful to introduce again the electrolyte
stirring during the anodization process and to add another new parameter additional
chemical species in the electrolyte.
Figure 5.11 shows current transients recorded during anodization of 4 samples at 20V
in electrolyte consisting of water / glycerol (50:50 vol.%) mixture with 0.27 M NH4F. In all
cases, no stirring was used during the first 2 hours of anodization. After 2 hours, several
different operations involving stirring (10 min at 300 rpm) were done: (a) stirring was
switched on and off (dotted line), (b) at the very moment as the stirring was switched on,
NH4F was added to the electrolyte to increase the concentration from 0.27 to 0.54 M (dashed
line); (c) for reference no stirring was used (solid line). In the last experiment (d), 0.27 M
(NH4)2[TiF6] was added in the basic electrolyte from the beginning and stirring was switched,
as in previous cases, after 2 hours of anodization (dashed-dotted line). I should mention that
this particular electrolyte had to be prepared before the anodization, due to comparably lower
solubility of (NH4)2[TiF6] to NH4F. The slow dissolution of the (NH4)2[TiF6] after adding it in
the basic electrolyte under stirring (as in the NH4F case) would lead to an unfair comparison.
In the case (a), the current density increased and decreased, after the stirring has been
switched on and off, respectively, in line with the previous experiment shown in Fig. 5.10. In
time, the current density returned to the value that would be expected in the case of no stirring.
This indicates that the steady state current density is controlled by the diffusion. This is in a
way not surprising, as during stirring the fresh fluoride species are more rapidly delivered to
the tube and reaction products are more rapidly removed out from the vicinity of the nanotube
layer, resulting in an enhanced field-aided TiO2 dissolution rate and therefore increased
current. In the case (b) the current density increased even higher than in the case of stirring
only (a) and after the stirring has been switched off, the current density decreased to a steadystate value that one would expect for electrolytes with higher NH4F concentration (in line
with the difference in the steady-state current density values in Fig. 4.30 for the corresponding
NH4F concentrations). In addition to the current density, the tube layer thickness also
increased from 1.7 to 2.3 m (measured after 3 hours of anodization). In the last case (d), the
current density was lower during the entire anodization, which would indicate a lower
dissolution TiO2 rate (as there are already [TiF6]2- species in the electrolyte). After stirring has

127

been switched on and off, the current increased and decreased, respectively, as in other cases,
and a value was established that would be expected without stirring.
The findings of Figs. 5.10 and 5.11 clearly confirm that the growth of the tubes is
determined by the diffusion (or, say, is diffusion limited), because upon the stirring the width
of the diffusion layer is shortened (Eq. 5.2) leading to an increased current density.
Additionally, by adding more fluorides (reactants) the current density increases, whereas by
adding more hexafluorotitanate species (reactions products), the current density decreases.

Figure 5.11 Current transients recorded anodization in modified electrolytes and under the
same hydrodynamic conditions.

During the anodization process, the concentration of [TiF6]2- in the bulk electrolyte is
increasing, whereas the concentration of fluorides is decreasing. As evident from Fig. 5.11,
the equilibrium between the fluxes of fresh fluorides species and formed [TiF6]2- species is the
rate determining factor. In the non-stirred case, this equilibrium is driven by natural
convection of the species from / into the bulk electrolyte (based on random movement of the
species induced by local thermal differences) and transport through the diffusion layer that is
established in the vicinity and also inside the nanotube layer (i.e. along the nanotube walls). In
the case of stirring, the effect of natural convection is diminished and the diffusion of the
species occurs via a thinner diffusion layer.

128

It can be assumed that the diffusion process can be separated into two parts, diffusion
within the tubes and diffusion across the tube-adjacent layer of the electrolyte to the bulk
electrolyte. To determine the diffusion behavior of the F- and [TiF6]2- species, one may
consider the model presented in Fig. 5.12 (according to Ref. [218]), which is derived from the
diffusion model of porous alumina [9].

Figure 5.12 Concentration profiles of F- and [TiF6]2 species in and outside the nanotubular
layer and the equations used for the calculations. L represents the tube length, p is the tube
porosity (estimated to be 50 %), z number of exchanged electrons (4), F is the Faradays
constant, D is diffusion coefficient, x0 is the diffusion layer width.

Based on the findings of Fig. 5.11, we can assume that the diffusion of both species
play an important role for the reaction rate inside the tubes (although their flux is slightly
fluctuating with time, as deduced from the current oscillations, Fig. 4.19). At the beginning of
anodization, the [TiF6]2- concentration is equal to zero. Lets assume that: (i) the nanotubular

129

structure has been formed, (ii) the dissolution of the tube top is negligible, (iii) F- are
consumed and [TiF6]2- are produced only at the tube bottom, respectively, and (iv) there is a
linear diffusion between the tube bottom and the bulk electrolyte. Then the concentration of
[TiF6]2- ions along the walls is dependent on the [TiF6]2- production rate at the pore bottom.
This means that in the limiting case, when all fluorides at the tube bottom are consumed by
the reaction to [TiF6]2- (Eqs. 2.14 and 2.15), the [TiF6]2- molar concentration is equal to 6 x
the F- bulk concentration, according to the stoichometry of 6 F- [TiF6]2-.
Once the stirrer is switched off, the [TiF6]2-containing electrolyte is no longer swept
from the surface, as shown in Fig. 5.12. Therefore, a significant [TiF6]2- concentration is built
up in the bulk electrolyte near the nanotube surface so that the concentration gradient inside
the tubes decreases. This results in a lower diffusion rate of [TiF6]2- and theoretically should
lead to a higher equilibrium concentration at the tube bottom. Likewise, a similar relationship
holds for the fluoride species being consumed by the reaction.
The rate of the change of the concentration profiles after the stirrer is switched off is
determined by the diffusion and accumulation in the electrolyte outside the tubes in the
diffusion layer. In other words, by stirring mainly the part within the electrolyte can be
affected. For this reason, it takes several minutes before the new steady-state conditions and
the approximately constant anodization current have been reached (Fig. 5.11 and 5.12). Due
to the change in the diffusion layer thickness, the concentration profile at the tube mouth will
also be affected. Based on the present results it is not entirely clear, if F- inward diffusion, or
[TiF6]2- outward diffusion is the key process as adding either species significantly affects the
resulting current density. However, by using a model described by Yasuda and Schmuki [218],
it is possible to estimate the diffusion term zDc0 to be 2.4x10-8 mol / cm.s. This corresponds
to a diffusion constant of F- 2.2x10-5 cm2 / s and the diffusion layer thickness 1.5 m. The
fact that [TiF6]2- and F- additions influence the resulting current density may be ascribed to
inter-related equilibration (e.g. dissociation of [TiF6]2-) that affects the results. Nevertheless,
earlier results showed [218] that for Ti and Zr under similar conditions, very different limiting
current densities are obtained, which would suggest that [TiF6]2- is the diffusion rate
controlling species.
In order to explain the impact of stirring on the formation of ordered pore arrays, two
different considerations can be put forward. First, the electrolyte composition at the tube
bottoms with and without stirring is different, so that the currents differ by more than 60%.
Therefore, it cannot be expected that identical nanotube structures would grow both with and
130

without stirring, even if the other experimental parameters remain unchanged. Secondly, by
SEM observations of these samples, it has been revealed that the tube walls are thinner in the
case of stirring and the uppermost parts of the tubes are partially collapsed, presumably due to
agitation of the electrolyte. Thirdly, the stirring may contribute to spatially homogenous
etching conditions and decrease the standard deviation of the tube thickness.

5.2.4 Temperature
Based on the facts that the nanotubes are grown under competition of oxide formation
and dissolution reactions that are diffusion-limited, it can be easily deduced that temperature
must have a significant effects on the rates of all reactions taking place during the nanotube
growth. For glycerol electrolytes (as shown in Figs. 4.22 and 4.23), temperature has a huge
influence on the viscosity and the resulting morphology. For other electrolytes, such as for the
aqueous electrolytes (consisting of e.g. (NH4)2SO4 / NH4F), ambient laboratory temperatures
were the most suitable to obtain ordered nanotube layers.
However, one particular case has been found, in which the temperature tremendously
decreased the complex TiO2 dissolution [230]. In Fig. 4.4, there is a considerable loss of Ti
substrate upon anodization in HF-based electrolytes. Due to the fact that the starting material
is being ``eaten-up, it hampers using either very thin Ti substrates, or thin Ti layers
deposited on other material (e.g. conductive glass). However, for some applications, it would
be highly desirable to facilitate growth of nanotube layers on different substrates than Ti, such
as on silicon or conducting glass. This would enable the nanotube layers to be used for
various purposes including fabrication of electronic devices and solar cells. In order to grow
nanotubes, the parameters such as the applied potential and the fluoride concentration must be
maintained in the self-organization regime. Thus, the most straight forward factor that could
be considered for reducing the substrate removal rate is the temperature during anodization
(decreasing diffusion of the species, taking the glycerol case as an example). Results of two
anodization experiments performed on Ti sputtered on a Si wafer at two different
temperatures (2 and 20C) in 1 M H2SO4 / 0.16 M HF are given in Fig. 5.13 [230]. The
thickness of the Ti layer on the Si wafer was originally 500 nm. As it can be seen from Fig.
5.13a, after 1 hour of anodization at 20 V at 20C, the Ti layer was nearly dissolved and only
patches of nanotube formation can be observed on the Si surface.

131

Figure 5.13 SEM images of Ti (sputtered on the Si wafers) anodized at 20C (a) and at 2C
(b) in 1 M H2SO4 / 0.16 M HF during 1 hour; (c) current transients recorded during anodizatin
of both samples. The insets show the side-views on the layers.

A completely different situation was observed on the sample anodized at 2C (Fig.


5.13b), which resulted in an ordered nanotube layer (there is still a Ti left in between the tube
layer and the Si wafer). The diameter and the thickness of the tubes is similar to anodized Ti
sheet that is approximately 100 nm and 400 nm, respectively. Using the lower temperature
with the otherwise ``usual conditions has no influence on the tube dimensions. Additionally,
the tube growth kinetics remain similar, i.e. after 1 hour the nanotube layer has the same
thickness as at ambient laboratory conditions [230]. Therefore, it can be concluded that
lowering the temperature had a vital effect on the rates of TiO2 chemical and field-aided
dissolution resulting in much lower removal of Ti. As shown in Fig. 5.13c, the current
densities recorded are also distinctly different the lower temperature has the lower the
current density.

132

5.2.5 Role of the compact oxide layer


Some may assume that the compact film grown in the early stages of the nanotube
formation (as shown in Figs. 4.17 and 5.3) plays an important role in the nanotube formation.
It is at the defect sites (cracks, non-uniform thickness), that the first localized attack by
fluoride anions takes place.
In this regard, a very practical comparison can be shown that is based on the nanotube
growth attempt using a bare Ti foil and already pre-anodized Ti foil (in 1 M (NH4)2SO4 at 20
V for 5 min), thus covered with a compact TiO2 layer [216]. Fig. 5.14 presents SEM images
taken from both types of samples after 3 minutes and after 1 hour of anodization in 1 M
(NH4)2SO4 / 0.14 M NH4F at 20 V (sweep rate 1 V / s). Careful analysis of the images taken
from samples anodized for 3 minutes revealed significant morphological differences. The asanodized sample (Fig. 5.14a) has clearly patches of the porous structure underneath a
perforated top layer (former compact layer grown at the very beginning of anodization in a
fluoride-containing electrolyte). However, for the pre-anodized sample (Fig. 5.14b), this is not
the case. It appears that the layer consists of the initial compact film that became fluffy, likely
as a result of the fluoride penetration and local dissolution.

Figure 5.14 SEM side-view images of the layer grown by anodization in 1 M (NH4)2SO4 /
0.14 M NH4F at 20 V for 3 min from on a Ti foil (a) and on a pre-anodized Ti foil (b) with
anodic TiO2 layer formed in 1 M (NH4)2SO4 at 20 V. Images (c) and (d) represent samples
prepared in the same way, but anodized for 1 h, respectively.

133

Nevertheless, after 1 hour of anodization, a very different picture is obtained for both
samples - the different origins of the samples have disappeared. Obviously the presence of the
pre-anodized compact layer only slows down the self-organization process, but does not affect
the final structure of the nanotube layer. After 1 h in both cases approximately a 1 m thick
self-organized TiO2 nanotube layer was grown (Fig. 5.14c, d). In light of these findings, one
can conclude that the compact TiO2 layer that is formed in the early stages of anodization, no
matter at what potential ramping rate, does not have an important role for the subsequent pore
initiation and nanotube formation. A presence of the compact oxide layer at the beginning of
anodization is just a consequence of the applied potential. Even anodization of a pre-anodized
sample will lead to the nanotube formation, with only a small delay in the initial stage
compared to the other sample. However, the potential sweep rate does have an influence on
the dimensions of the resulting nanotube layer, as shown earlier in the section on pH.

5.3 Other features related to the nanotubes


5.3.1 Growth efficiency
Another aspect of the oxide film growth is its growth efficiency in terms of the electric
current that is consumed by side-reactions and current that is used to build the oxide layer. For
evaluation of the growth efficiency, it is necessary to consider the Faradays law given by Eq.
2.8.
As an example, a set of experiments was done for nanotube layers grown in glycerol /
water (50:50 vol. %) / 0.27 M NH4F electrolytes by anodization at 20 and 30 V for 6 and 15
hours, respectively. The charge density that was going through the samples during the
anodization could be extracted from the recorded electrochemical data (polarization curves
and current transients). The total charge passed during the polarization period and the
potentiostatic period of the experiment was considered. All values used for this calculation
were taken for 1 cm2 of the sample that was also the anodized area. Therefore the current was
equal to the current density and the charge to the charge density. By converting the charge
into mass via Eq. 2.8, the values that are shown in Table 5.2 were obtained. In order to
calculate the growth efficiency, the real mass of TiO2 in the nanotubes is required. This can be
done either by calculation (based on the tube dimensions from SEM images), but this leads,
however, to a large scatter in the values, or, more accurately, by weighing the samples (after

134

removing the Ti substrate by selective dissolution in Br2 / MeOH solution). The values from
weighing are also shown in Table 5.2. By comparing the theoretical and the real TiO2 mass,
the growth efficiencies were obtained.
At 6 hours of anodization, the efficiencies were quite high (~ 89 %). Upon longer
anodization times, there was a decrease in the efficiencies in all cases for 15 hours of
anodization. This is not surprising, because, in fact even after extensive anodization, a high
current still flows through the oxide layer without having a large impact on the thickness of
the layers. In other words, the nanotube growth front moves continually towards the Ti
substrate, but the thickness of the nanotube layers remains nearly the same (max. 5 % of
change) for times longer than 6 hours.
Table 5.2 Growth efficiencies of the nanotube layers
U (V) / sample

Time
m calculated
m real
Growth
(hours)
from Q (mg)
(mg)
efficiency (%)
20 / TiO2 nanotubes
6
0.36
0.32
89
15
0.70
0.32a
46
30 / TiO2 nanotubes
6
0.35
0.31
89
15
1.44
0.31a
22
- result from weighing the samples
a
- dimensions of nanotubes do not significantly change between 6 and 15 hours of
anodization

It is noteworthy also to mention that for neutral aqueous electrolytes, such as 1 M


(NH4)2SO4 / 0.14 M NH4F, the growth efficiency (calculated for 6 hours of anodization i.e.
nearly constant tube layer thickness) is in the range of 40 - 50 %. For acidic HF electrolytes,
the growth efficiencies are comparably small (3 - 10 %), because a large part of the current
goes to the TiO2 dissolution as a consequence of the low pH. Interestingly, for the acidic
electrolyte, the growth efficiencies showed an increase during anodization at low
temperatures, as shown in Fig. 5.13. For purely organic electrolytes, the tube growth proceeds
at nearly the 100 % rate. This is confirmed by SEM images (as in Fig. 4.25) that show very
minor chemical dissolution of the substrates in these electrolytes [195, 197].

5.3.2 Structure
As shown in Chapter 4, all as-formed nanotube layers have an amorphous structure
that can be converted into anatase, or a mixture of anatase and rutile, depending on the
temperature during the thermal annealing. In the early report of Beranek et al. [119], a small

135

amount of anatase was detected in the nanotube layers that were anodized for a very long time
(24 hours). There have been two reports showing an annealing temperature series that
practically contradict each other [144,231]. The first report by Grimes and coworkers [144] on
short tubes grown in 0.5 wt. % HF aqueous electrolyte shows that after annealing at 430C a
minor rutile peak is detected. This report also claims that nanotubes with single phase
crystallinity can not be achieved. The second paper [231] points out the opposite, that single
crystalline anatase nanotubes (grown in H3PO4 / 0.5 wt. % HF electrolyte) can be obtained by
annealing at temperatures up to 450C and that at higher temperatures rutile traces are always
detected. In the present work, there is a hint (Fig. 4.37c) that annealing of the high aspect ratio
and smooth nanotubes at 450C causes a formation of the single phase anatase structure,
inclining to the latter report [231]. This experimental finding would be in line with the
thermodynamical stability of the anatase phase of titania in small crystallites [232]. Large
surface tension exerted by large surfaces present in the tube walls may be responsible for the
fact that crystallites prefer to form into an anatase phase (and suppress rutile formation).
There is very strong support from ab-initio molecular dynamics investigation which shows
that the average surface energy of a crystal in the anatase phase is lower than that of a rutile
crystal [232]. Why this effect may be predominant in a high aspect ratio nanotube remains
open to further work, but may possibly be due to typically lower wall thicknesses of these
nanotubes.
The mechanical stability of the nanotube layers diminished, when annealed at
temperatures above 600C, as mentioned in Fig. 4.37. The nanotubes have collapsed and
cracked apart in line with the literature [144]. Interestingly, it has been also shown that tube
shrinkage and cracking was even more pronounced in humid argon environments, when
compared to oxygen containing ambient atmosphere [144]. It appears that enhanced mass
transport in the humid atmosphere played a major role in the shrinkage and sintering
processes. At elevated temperatures reducing atmospheres create oxygen vacancies in the
nanotubes, which can enhance the mass transport in nanotube walls. Humid environments
also increase the surface diffusion, which influences the crystallization and sintering behavior
of titania. The density of anatase is less than of rutile (3.81 vs. 4.25 g / cm3), hence elastic
strain (or, say, shrinking) is caused when the rutile phase is formed in the anatase phase [143],
which is likely the reason for the nanotube cracking.

136

6. CONCLUSIONS
In the present thesis, efforts to achieve growth of self-organized TiO2 nanotube layers
by anodization of Titanium have been carried out. In comparison with the available early
work, the results of this thesis clearly show that significant improvement of the dimensions,
morphology and structure of the nanotube layers has been achieved by tailoring a wide range
of experimental conditions.
In accord with the growth of other self-organized nanostructures (e.g. porous
alumina), the presence of an electric field is required that drives the titania formation and
contributes significantly to its dissolution. In contrast to porous alumina, this dissolution does
not occur only due to acidity (H+) that is generated inside the pores, but also due to a presence
of fluoride anions that dissolve titania and form soluble complexes.
Several other important parameters for the nanotube growth have been identified
within the present work. Most importantly, an optimized pH profile in the tube is crucial for
the growth of high aspect ratio nanotubes. By using neutral or slightly acidic electrolytes, a
low pH is generated only via the field-aided reactions at the tube bottom and the tube walls
are protected from the dissolution simply due to a higher pH along them that significantly
slows down the TiO2 dissolution. Another significant influence has been shown to be the
electrolyte viscosity that is responsible for slow diffusion of the species within the tube
(influencing the kinetics of the oxide formation vs. dissolution equilibria), while enabling
growth of long and smooth nanotubes. These effects can be significantly altered by the
electrolyte temperature during anodization. By tailoring the amount of water and the fluoride
content in the glycerol / NH4F electrolytes, a large range of diameters and lengths has been
obtained. The diameter of the nanotubes was shown to be closely related to the growth rate of
compact oxide versus the applied potential, similar to porous alumina. In agreement with the
conditions that are required for the growth of perfectly ordered porous alumina, the two-step
anodization approach and optimized experimental conditions (high purity Ti substrate, applied
potential, ethylene glycol based electrolytes) have been shown to be critical for the growth of
the ordered and hexagonally packed nanotube layers.
Based on the available theoretical background of anodic porous alumina, anodic TiO2
films and the detailed analysis of the nanotube growth shown herein, a growth mechanism for
the nanotube layers has been proposed. This growth mechanism consists of several stages and
relies on the field-aided TiO2 formation and the field-aided TiO2 dissolution accompanied by
more or less significant chemical TiO2 dissolution. At the beginning of the anodization, a
137

compact TiO2 layer is formed. After a while, the presence of fluoride anions comes into play,
as the fluorides start to penetrate the freshly forming oxide leaving behind some nanochannels.
A worm-like structure is grown within a few minutes, as a result of a competition for the
available space between non-equal neighboring pores. In the course of time, deeper pores
have an advantage over shorter pores, as they have a larger oxidation area available. An
autocatalytic process accompanies this stage, which leads to accelerated production of H+ at
these larger pores, which further enhances their drilling. As a consequence, growth of the
shallower pores is terminated in time. Due to the equilibration of the oxide formation and
dissolution kinetics and likely also due to stresses present at the Ti metal/ TiO2 interface,
continuous anodization leads to a conversion of the originally smooth Ti surface into a selforganized TiO2 nanotube layer. This process requires some anodization time typically at
least 1 hour. With further anodization, the thickness of the layer increases and diffusion
strongly influences the tube growth in this later stage.
In general, for all electrolytes developed within this thesis, there was always a set of
anodization conditions that were suitable for the nanotube growth. However, if one parameter
of this set becomes unsuitable for the self-organization regime, no ordered layers can be
obtained.
All nanotube layers synthesized in this work have an amorphous structure and consist
of TiO2. Some traces of fluorides are present in the nanotube layers after their growth. For
different functional applications (some of them shown here), this amorphous structure can be
converted to a crystalline structure by annealing.

138

7. OUTLOOK
7.1 TiO2 nanotube layers as a promising and prospective material
Considering the results presented in this thesis, the key points on the TiO2 nanotube
growth have been identified. However, more investigations are needed to obtain a complete
picture on the nanotube growth. One may consider the role example of porous anodic alumina,
where it took decades before the growth mechanisms became clear. Even at the present time,
surprizing morphologies can be obtained, for instance by using a hard anodization approach,
as shown in recent reports [12,13]. It will be particularly interesting to find out, what exactly
is the reason behind the nanotubular, instead of nanoporous morphology, as for some valve
metals (Ti, Zr and Hf) nanotube layers are grown upon anodization in fluoride containing
electrolytes, whereas others (Ta, Nb, W) form nanopores.
Noteworthy is that recently a new research direction was established by alloying these
elements and anodizing them to a mixed valve metal oxide nanostructures (for an overview
see Ref. [233]). These novel materials could be used for very interesting applications, where
the intrinsic specific properties of the oxides are mixed, giving for instance new mechanical or
semiconducting characteristics.
It is not only the growth mechanism that remains to be solved and understood, but
other aspects that are connected with the nanotube structure and morphology. For instance,
based on recent experiments in our laboratory, it seems that the thermal annealing process has
strong implications for the nanotube layers. By changing the heating and cooling rates, and
also annealing durations, the mechanical and thermal stability of the nanotube layers can be
significantly altered. Apart from the crystallites with different sizes grown in the nanotube
material, the nanotubes may even be transformed by annealing to fused nanoporous layers, in
which the gaps between the nanotubes have been diminished. These layers seem to be very
mechanically and thermally stable, even more than ``traditional nanotube layers.
A fundamental drawback of TiO2 can be considered its in-activity in the visible light
region, as discussed in Chapter 2. Nevertheless, there are two basic approaches, how to extend

139

the photoelectrochemical response of TiO2 to the visible part of the solar spectrum in order to
harvest it. The first approach is closely related to the dye-sensitized solar cell shown and
explained in Chapter 2. There have been a large number of reports using various
photosensitizers anchored to TiO2 of different forms. In contrast, there have been only a few
reports [234-238], including ours [234,238] that showed that the nanotube layers are
promising candidates for high efficiency dye-sensitized solar cells. The second approach is
based on doping the TiO2 with suitable species, usually by N [165,166], or C [167] that create
a sub-band gap in the material with lower energies, therefore electrons are excitable by the
visible light illumination. Various techniques have been used to introduce different species
into TiO2 compact layers or powders. Ion implantation is the most straight forward approach
for doping but, most efforts carried out on TiO2 (by transition metal implantation) were
accompanied by structural damage. In the case of TiO2 nanotubes, we have shown successful
nitrogen doping of the tube layers achieved by ion implantation [239], or by thermal treatment
in NH3 [240]. In contrast, wet chemical approaches have shown only limited success up to
now [241].
Features that may be highly significant for biomedical applications is that the tubular
layers apparently stimulate hydroxyapatite formation on their surface [242,243] and also cell
interactions that have been found to be strongly dependent on the tube diameter [199].

7.2 Specific features of the nanotube layers


Particular advantages of the self-organized nanotube layers presented in this thesis are
their large surface area and their defined geometry. For all the functional TiO2 applications
described in Chapter 2, it is desirable to exploit the new nanotube features and arrangements
and their large surface area and replace often unspecified nanoparticles. In other words,
nanotubular oxide films combine geometrical advantages, given by their array structure, with
specific TiO2 properties. For example, ordered TiO2 nanotubes with a tube length in the mrange are promising candidates for a higher light-to-electricity conversion efficiency, as losses
that take place due to the grain boundaries in the sintered nanoparticles can be avoided. The
defined geometry results in a narrow distribution for the diffusion path for entering the tubular
depth (e.g. reactants to be transported to the tube bottom, light to be absorbed). In other words,
the tubes provide a more optimized geometry with significantly shorter carrier diffusion path
for electrons and holes in the tube walls (10 - 20 nm thick) to the active surface area and
provide lower trapping and recombination kinetics, in comparison with the TiO2 nanoparticles
140

(e.g. P25), where the electron-hole pairs have to travel between single nanoparticles (that are
often sintered and convoluted). Therefore, the response and the performance of ordered tube
arrays in applications, such as sensing, or photocatalysis, is expected to be much more defined
than using classical high surface area layers, where nanoparticles are compacted, or sintered
to produce an open porous network.
Figure 7.1 shows a sketch of some effects that can expected to be especially
pronounced in the nanotubular geometry.

Diffusion of the
reaction species

Reactants
Reactant diffusion path

Light absorption
= f (absorption coefficient,
wavelength, depth)

Products
Recombination of the charge
carriers = f (structure,doping,
length, wall thickness)

Figure 7.1 Advantages the nanotubular geometry represented in terms of physico-chemical


aspects.

It can be expected that in at least some of these TiO2 applications (if not all), the
nanotubular structures presented in this work can provide a significant advancement. For
example, it has been shown recently that the nanotube layers are highly efficient for the
photocatalytical decompositon of organic compounds [244].
A key advantage of the nanotube layers over the nanoparticles is that essentially any
form of a Ti surface (sheets, foils, sputtered layers) can be treated by a quick and low cost
anodization approach leading to a TiO2 nanotube layer coating. In other applications, a
separation of the tube array into isolated tubes or selective removal of the nanotubes (e.g. for
templating in parallel to porous alumina) may be desirable.

141

8. REFERENCES
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]

F. Keller, M. S. Hunter, and D. L. Robinson, J. Electrochem. Soc., 100 (1953) 411.


J. W. Diggle, T. C. Downie, C. W. Coulding, Chem. Rev., 69 (1969) 365.
J.P. O' Sullivan, G. C. Wood, Proc. of the Royal Society of London, Series A Mathematical
and Physical Sciences, 317 (1970) 511.
G. E.Thompson, G.C. Wood, Nature 290 (1981) 230
G. Thompson, G. C. Wood, Treatise on materials science and technology, New York
(Academic Press) 1983, 23.
H. Masuda, K. Fukuda, Science, 268 (1995)1466.
H. Masuda, H. Yamada, M. Satoh, H. Asoh, M. Nakao, and T. Tamamura, Appl. Phys. Lett., 71
(1997) 2770.
G. E. Thompson, Thin Solid Films 297 (1997) 192.
O. Jessensky, F. Mller and U. Gsele, J.Electrochem.Soc. 145 (1998) 3735.
S. Ono, M. Saito, M. Ishiguro, H. Asoh, J. Electrochem. Soc. 151 (2004) B473.
S. Ono, M. Saito, H. Asoh, Electrochim. Acta 51 (2005) 827.
S.Z. Chu, K. Wada, S. Inoue, M. Isogai, A. Yasumori, Adv. Mater. 17 (2005) 2115.
W. Lee, R. Ji, U. Gsele, K. Nielsch, Nat. Mater. 5 (2006) 741.
L.T.Canham, App.Phys.Lett. 57 (1990) 1046.
V. Lehmann and H. Fll, J. Electrochem. Soc., 137 (1990) 653.
V. Lehmann, U. Gsele, Adv. Mater. 4 (1992) 116.
J.-N. Chazalviel, R.B. Wehrspohn, F. Ozanam, Mat. Sci. Eng., B6970 (2000) 1.
U. Grning, V. Lehmann, S. Ottow, K. Busch, Appl. Phys. Lett. 68 (1996) 747.
J.W. Schultze, M.M. Lohrengen, Electrochim. Acta 45 (2000) 2499.
V. Zwilling, E. Darque-Ceretti, A. Boutry-Forveille, D. David, M.Y.Perrin and M. Aucouturier,
Surf. Interface Anal., 27 (1999) 629.
V. Zwilling, M. Aucouturier and E. Darque-Ceretti, Electrochim. Acta, 45 (1999) 921.
L. Young, Anodic Oxide Films, New York (Plenum), 1961.
D. A. Vermilyea, Anodic Films, in Advances in Electrochemistry and Electrochemical
Engineering, London (Wiley), 1963, 248.
A.K. Sharma, Thin Solid Films, 208 (1992) 48.
G. P. Burns, J. Appl. Phys., 65 (1989) 2095.
G.He, Q. Fang, L. Zhu, M. Liu, L.Zhang. Chem. Phys. Lett., 395 (2004) 259.
S. Takeda, S. Suzuki, H. Odaka, H. Hosono, Thin Solid Films, 392 (2001) 338.
J. M. Bennet, E. Pelletier, G. Albrand, J.-P. Borgogno, B. Lazarides, C.K. Carniglia, R.A.
Schmell, T.H. Allen, T. Tuttle-Hart, K.H. Guenther, A Saxer, Applied Optics, 28 (1989) 3303.
W. Li, S. Ismat Shah, C.-P. Huang, O. Jung, C. Ni, Mater. Sci. Eng. B, 96 (2002) 247.
M. Ritala, M. Leskel, E. Nyknen, P. Soininen, L. Niinist, Thin Solid Films, 225 (1993) 288.
H.K. Ardakani, Thin Solid Films, 248 (1994) 234.
Y. Ohya, H. Saiki, T. Tanaka, and Y. Takahashi, J. Am. Ceram. Soc., 79, 825 (1996).
I. Zhitomirsky, NanoStructured Materials, 8 (1997) 521.
Information Bulletin 14 (1948), ADS (The Aluminium Development Association) London.
H. Asoh, K. Nishio, M. Nakao, T. Tamamura, H. Masuda, J. Electrochem.Soc., 148 (2001)
B152.
V. P. Parkhutik, V. I. Shershulsky, J. Physics D Appl. Phys., 25 (1992) 1258.
J. Choi, Y. Luo, R.B. Wehrspohn, R. Hillebrand, J. Schilling, U. Gsele, J. Phys.Chem. 94
(2003) 4757.
G. C. Wood, P. Skeldon, G. E. Thompson, and K. Shimizu, J. Electrochem. Soc., 143 (1996) 74.
F. Matsumoto, K. Nishio, H. Masuda, Adv. Mater. 16 (2004) 2105.
G.E. Possin, Rev. Sci. Instrum. 41 (1970) 772.
S. Kawai, R. Ueda, J. Electrochem. Soc. 122 (1975) 32.
L.S. Van Dyke, C.R. Martin, Langmuir 6 (1990) 1123.
D. Al Mawlawi, N. Coombs, M. Moskovits, J. App. Phys., 70 (1991) 4421.
C. R. Martin, Science, 266 (1994) 1961.

142

[45] R M. Metzger, V.V. Konovalov, M. Sun, T. Xu, G. Zangari, B. Xu, M. Benakli, W. D. Doyle,
IEEE Transactions on Magnetics, 36 (2000) 30.
[46] H. Masuda, M. Ohya, H. Asoh, M. Nakao, M. Nohtomi, T. Tamamura, Jap. J. Phys. Chem., 39
(1999) L1403.
[47] N. Itoh, K. Kato, T. Tsuji, M. Hongo, J. Membrane Sci 117 (1996) 189
[48] K. Nielsch, F. Mller, A.-P. Li, U. Gsele, Adv. Mater., 12 (2000) 582.
[49] W. Lee, R. Scholz, K. Nielsch, U. Gsele, Angew. Chem.Int.Ed., 44 (2005) 6050.
[50] Y. Piao, H. Lim, J.Y. Chang, W.Y. Lee, H. Kim, Electrochim. Acta, 50 (2005) 2997.
[51] M. Pourbaix, Atlas d` Equilibres lectrochimiques, Paris (Gautier-Villars), 1963, p. 213
[52] E. Horwood, Instrumental Methods in Electrochemistry, Chichester (Wiley), 1985, 178.
[53] J. Tafel, Z. Physik. Chem., 50 (1905) 641.
[54] P. Schmuki, From Bacon to Barriers, J. Solid State Electrochem., 6 (2002) 145.
[55] V. E. Carter, Metallic Coatings for Corrosion Control, Hants (Butterworth), 1977,27
[56] R. Memming, Comprehensive Treatise of Electrochemistry, Plenum Press (New York), 1983,
Vol. 7.
[57] A. Gnterschulze, H. Betz, Z. Phys., 92 (1934) 367.
[58] E.J.W. Verwey, Physica 2, (1935) 1059.
[59] N.F. Mott, Trans. Faraday Soc., 36 (1940) 472.
[60] H. Cabrera, N.F. Mott, Rep. Prog. Phys., 12 (1948-49) 163.
[61] D. A. Vermilyea, Acta Met. 1, (1953) 282.
[62] J.F. Dewald, J. Electrochem. Soc., 102 (1955) 1.
[63] K. J. Vetter, Electrochemical Kinetics, Berlin (Springer-Verlag), 1967
[64] J. W. Schultze, M.M.Lohrengel, D. Ross, Electrochim. Acta, 28 (1983) 973.
[65] S. Kudelka, A. Michaelis, J.W. Schultze, Electrochim. Acta, 41 (1996) 863.
[66] C.K. Dyer, J.S.L. Leach, J. Electrochem. Soc., 125 (1978) 23.
[67] J.-L. Delplancke, R. Winard, Electrochim. Acta, 33 (1988) 1539 + 1551.
[68] N. Khalil, J.S.L. Leach, Electrochim. Acta, 31 (1986) 1279.
[69] Y.T. Sul, C.B. Johansson, Y. Jeong, T. Albrektsson, Medical Eng. & Phys., 23 (2001) 329.
[70] R. M. A. Azzam and N. M. Bashara, Ellipsometry and Polarized Light, Netherlands (Elsevier),
1987.
[71] G. Blondeau, M. Froelicher, M. Froment, A. Hugot-Le Goff, Thin Solid Films, 42 (1977) 147.
[72] H. Pelouchova, P. Janda, J. Weber, L. Kavan, J. Electroanal. Chem., 566 (2004) 73.
[73] J.A. Harrison, D.E. Williams, Electrochim. Acta, 27 (1982) 891.
[74] K. Asoaka, L. Yokoyama, H. Nagumo, Metallurg. Mater. Trans., 33A (2002) 495.
[75] Z. Tun, J.J. Noel, D. W. Shoesmith, J. Electrochem. Soc., 146 (1999) 988.
[76] J. Yahalom, J. Zahavi, Electrochim. Acta, 15 (1970) 1429.
[77] G.C. Wood, C. Pearson, Corros. Sci., 7 (1967) 119.
[78] J.C. Marchenoir, J. P. Loup and J. Masson, Thin Solid Films 66 (1980) 357.
[79] Y. Mueller, S. Virtanen, in Pits and Pores II, P. Schmuki, D. J. Lockwood, Y. H. Ogata and H. S.
Isaacs, Editors, The Electrochemical Society Proceedings Series, Pennington, NJ (2000), PV
2000-25, p. 294.
[80] N.Sato, Electrochim. Acta, 16 (1971) 1683.
[81] C. K. Dyer, J. S. L. Leach, J. Electrochem. Soc., 125 (1978) 1032.
[82] S. Ikonopisov: Electrochim. Acta, 22 (1977) 1077.
[83] D. A. Vermiyela, J. Electrochem. Soc., 110 (1963) 345.
[84] Y.T. Sul, C.B. Johansson, S. Petronis, A. Krozer, Y. Jeong, A. Wennerberg, T. Albrektsson,
Biomaterials, 23 (2002) 491.
[85] B. Yang, M. Uchida, H.-M. Kim, X. Zhang, T. Kokubo, Biomaterials, 25 (2004) 1003.
[86] P. Schmuki, M.J. Graham, in Encyclopedia of Chemical Physics and Physical Chemistry (vol. 3,
Ed. J.H. Moore, N.D. Spencer), 2001, p. 2423.
[87] Corrosion Resistance of Ti, Bulletin of TIMET (Titanium Metals Corporation).
[88] N. Cassilas, S.J. Charlebois, W.H. Smyrl, H.S. White, J. Electrochem. Soc., 141 (1994) 636.
[89] A.Takamura, Corrosion, 23 (1967) 306.
[90] I. Dudgale, J.B. Cotton, Corros. Sci. 4 (1964) 397.
[91] S. Virtanen, C. Curty, Corrosion 60 (2004) 643.

143

[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]
[101]
[102]
[103]
[104]
[105]
[106]
[107]
[108]
[109]
[110]
[111]
[112]
[113]
[114]
[115]
[116]
[117]
[118]
[119]
[120]
[121]
[122]
[123]
[124]
[125]
[126]
[127]
[128]
[129]
[130]
[131]
[132]
[133]
[134]
[135]
[136]
[137]
[138]

N. Cassilas, S.J. Carlebois, W.H. Smyrl, H.S. White, ibid., 140 (1993) L142.
M. Faraday, Phil. Trans., 124 (1834) 77.
J.C. Nelson, R.A. Oriani, Corros. Sci., 34 (1993) 307.
J.-D.Kim, S.-I. Pyun, R.A. Oriani, Electrochim. Acta, 40 (1995) 1171.
D. H. Bradhurst, J.S.L. Leach, J.Electrochem. Soc. 113 (1966) 1245.
N. Wthrich, Electrochim. Acta, 25 (1980) 819.
J. C. Nelson, R. A. Oriani, Electrochim. Acta 37 (1992) 2051.
T. Shibata, Y. C. Zhu, Corros. Sci., 37 (1995) 253.
K. Ueno, S. -I. Pyun and M. Seo, J. Electrochem. Soc., 147 (2000) 4519.
N.B. Pilling, R.E. Bedworth, J. Inst. Metals, 29 (1923) 529.
N. Wtrich., Electrochim. Acta, 26 (1981) 1617.
T. Hepel, M. Hepel, R. A. Osteryoung, J. Electrochem. Soc., 129 (1982) 2132.
W. Wilhelmsen, A. P. Grande, Electrochim. Acta, 32 (1987) 1469.
E. Bright, D.W. Readey, J. Am. Ceram. Soc.70 (1987) 907
M. Nakagava, S. Matsuya, T. Shiraischi, M. Ohta, J. Dent. Res. 78 (1999) 1568
Y.J.Kwon, H.-J. Seol, H.-I. Kim. K.-J.Hwang, S.-G. Lee, K.-H. Kim, J. Biomed. Mater. Res. B,
73 (2005) 285.
N. Schiff, B. Grosgogeat, M. Lissac, F. Dalard, Biomaterials, 23 (2002) 1995.
E.M.M. Sutter, G.J. Goetz-Grandmont, Corros. Sci. 30 (1990) 461.
N.T. Thomas, K. Nobe, Ibid., 116 (1969) 1748.
M. J. Mandry, G. Rosenblatt, J. Electrochem. Soc., 119 (1972) 29.
A. Caprani, I. Epelboin, P. Morel, J. Electroanal. Chem., 43 (1973) App. 2.
J. J. Kelly, Electrochim. Acta, 24 (1979) 1273.
I. Frateur, S. Catarrin, M. Musiani, B. Tribollet, J.Electroanal. Chem., 482 (2000) 202.
S. Catarrin, M. Musiani, J.Electroanal. Chem., 517 (2001) 101.
K. Shimizu, K. Kobayashi, G.E. Thompson, P. Skeldon, G.C. Wood, Phil. Mag. B, 73 (1996)
461.
N. Bao, X. Feng, X. Lu, Z. Yang, J. Mater.Sci., 37 (2002) 3035.
D. Gong, C. A. Grimes, O. K. Varghese, Z. Chen and E. C. Dickey, J. Mater. Res., 16 (2001)
3331.
R. Beranek, H. Hildebrand and P. Schmuki, Electrochem. Solid-State Lett., 6 (2003) B12.
S. Frey, B. Grsillon, F. Ozanam, J.-N. Chazalviel, J. Carstensen, H. Fll, R. B. Wehrspohn,
Electrochem. Solid-State Lett. 8 (2005) B25.
H. Asoh, S. Ono, T. Hirose, I. Takatori, H. Masuda, Jpn. J. Appl. Phys. 43 (2004) 6343.
H. Tsuchiya, J. M. Macak, I. Sieber, P. Schmuki, Small, 1(2005) 722.
W.J. Lee, W.H. Smyrl, Electrochem. Solid-State Lett., 8 (2005) B7.
H. Tsuchiya, P. Schmuki, Electrochem. Commun., 7 (2005) 49.
I. Sieber, H. Hildebrand, A. Friedrich and P. Schmuki, Electrochem. Commun., 7 (2005) 97.
S. Ono, T. Nagasaka, H. Shimazaki, H. Asoh, in Pits and Pores II: Formation Properties and
Significance for Advanced Materials, Proc. Electrochem. Soc, Pennington, PV 2004-19, p. 123.
H. Tsuchiya, J. M. Macak, I. Sieber, L. Taveira, A. Ghicov, K. Sirotna, P. Schmuki,
Electrochem. Commun., 7 (2005) 295.
S. Berger, H. Tsuchiya. A. Ghicov, P. Schmuki, Appl. Phys. Lett. 88 (2006) 203119
I. Sieber, B. Kannan, P. Schmuki, Electrochem. Solid-State Lett., 8 (2005) J10.
W.Wei, J. M. Macak, P. Schmuki, Electrochem. Commun. 10 (2008), in press.
H. Tsuchiya, S. Berger, J. M. Macak, A. G. Munoz, P. Schmuki, Electrochem. Commun. 9
(2007) 545.
J. M. Macak, H. Tsuchiya, L. Taveira, A. Ghicov, P. Schmuki, J.Biomed.Mat.Res. 75A (2005)
928.
H. Tsuchiya, J. M. Macak, A. Ghicov, P. Schmuki, Small 2 (2006) 888.
A. Ghicov, S. Aldabergerova, H. Tsuchiya, P. Schmuki, Angew. Chem. Int.Ed. 45 (2006) 6993.
X. Feng, J.M. Macak, S. P. Albu, P. Schmuki, Electrochem. Commun.9 (2007) 2403.
K. Yasuda, P. Schmuki, Adv. Mater. 19 (2007) 1757.
H. Tsuchiya, S. Berger, J. M. Macak, P. Schmuki, Electrochem. Commun. 9 (2007) 2397.
U. Diebold, Surface Science Reports, 48 (2003) 53.

144

[139]
[140]
[141]
[142]
[143]
[144]
[145]
[146]
[147]
[148]
[149]
[150]
[151]
[152]
[153]
[154]
[155]
[156]
[157]
[158]
[159]
[160]
[161]
[162]
[163]
[164]
[165]
[166]
[167]
[168]
[169]
[170]
[171]
[172]
[173]
[174]
[175]
[176]
[177]
[178]
[179]
[180]
[181]
[182]
[183]
[184]
[185]
[186]

M.R.Kozlowski, P. S. Tyler, W.H. Smyrl, R. R. Atanasoki, Surf. Sci., 194 (1988) 442.
T. Ohtsuka, M. Masuda, N. Sato, J. Electrochem. Soc., 132 (1985) 787.
J. S.L. Leach, B.R. Pearson, Corros. Sci., 28 (1988) 43.
W.D. Kingery, H.K. Bowen, D.R. Uhlmann, Introduction to Ceramics, New York (Wiley) 1976.
T. Nakamura, T. Ichitsubo, E. Matsubara, A. Muramutsu, N. Sato, H. Takahashi, Scripta
Materialia, 53 (2005) 1019.
O.K. Varghese, D. Gong, M. Paulose, C.A. Grimes, E. C. Dickey, J. Mater. Res., 18 (2003) 156.
B.ORegan and M.Grtzel, Nature 353 (1991) 737.
N. Serpone, E. Pelizzetti, Photocatalysis - Fundamentals and Applications, New York (Wiley),
1989.
A. Fujishima, K. Honda, Nature, 238 (1972) 37.
G. Sberveglieri Ed., Gas Sensors, Dordrecht (Kluwer Academic Publishing), 1992.
A.Aladjem, D.G .Brandon, J. Yahalom, J. Zahavi, Electrochim. Acta, 15 (1970) 663.
T. Hartmann et al., Nucl. Instr. and Meth. B, 141 (1998) 398.
J.W. Schultze, L. Elfenthal, G. Hansen, T. Patzelt, B. Siemensmeyer, J. Thietke, Corros. Sci., 31
(1990) 213.
D.J. Blackwood, L.M.Peter, Electrochim. Acta, 34 (1989) 1505.
D.J. Blackwood, G. Greef, L.M. Peter, Electrochim. Acta, 34 (1989) 875.
T. Ohtsuka, T. Otsuki, Corros. Sci., 39 (1997) 1253.
T. Ohtsuka, T. Otsuki, Corros. Sci., 40 (1998) 951.
K. Azumi, M. Seo, Corros. Sci., 43 (2001) 533.
T. Ohtsuka, T. Otsuki, Corros. Sci., 45 (2003) 1793.
J.P.S Pringle, Electrochim. Acta, 25 (1980) 398.
R. Memming, Semiconductor Electrochemistry, Weinheim (Wiley), 2001.
J.A. Davies, B. Domeiji, J. Electrochem. Soc., 110 (1963) 849.
H. Habazaki, K. Shimizu, S. Nagata, P. Skeldon, G.E. Thompson, G.C. Wood, J. Electrochem.
Soc.. 149 (2002) B70.
M.-H. Wang, K.R. Hebert, J. Electrochem. Soc. 146 (1999) 3741
L. Kavan, M.Grtzel, J. Rathousky, A. Zukal, J. Electrochem. Soc., 143 (1996) 394.
W.W. Grtner, Phys. Rev., 116, 84 (1959).
R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science, 293 (2001) 269.
T. Lindgren, J.M. Mwabora, E. Avendano, J. Jonsson, A. Hoel. C.G Granvist, S.E. Lindquist,
J.Phys. Chem. B 107 (2003) 5709.
S. Sakthivel, H. Kisch, Angew. Chem. Int. Ed., 42 (2003) 4908.
D.M.Brunette, P.Tengvall, M.Textor and P.Thomsen, Titanium in Medicine, Berlin (Springer),
2001.
Z.Z. Gu, A. Fujishima, O. Sato, Appl. Phys. Lett., 85 (2004) 5037.
A. L. Linsebigler, G. Lu, J. T. Yates, Chem. Rev., 95 (1995) 735.
R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima,A. Kitamura, M. Shimohigoshi,
T. Watanabe, Nature, 388 (1997) 431.
Y. Takahashi, Y. Matsuoka, J. Mater. Sci., 23 (1988) 2259.
W. Choi, A. Termin, M.R. Hoffman, J. Phys. Chem., 98 (1994) 13669.
M.K. Akhtar, Y. Xiong, S.E. Pratsinis, AIChE J., 3 (1991) 1561.
H. Selhofer, Vacuum Thin Film (August 1999), 15.
G. V. Samsonov, The Oxide Handbook, New York (IFI/ Plenum Press), 1982.
S. Iijima, Nature, 354 (1991) 56.
T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, K. Niihara, Langmuir, 14 (1998) 3160.
M.H. Huang, Y. Wu, H. Feick, N. Tran, E. Weber, P. Yang, Adv. Mater., 13 (2001) 113.
M. Adachi. Y. Murata, M. Harada. S. Yoshikawa, Chem. Lett., 29 (2000) 942.
S. Z. Chu, S. Inoue, K. Wada. D. Li, H. Haneda, S. Awatsu, J. Phys. Chem. B, 107 (2003) 6586.
S. Uchida, R. Chiba, M. Tomiha, N. Masaki, M. Shirai, Electrochemistry, 70 (2002) 418.
M. Adachi. Y. Murata T.Okada. S. Yoshikawa, J. Elelectrochem. Soc., 150 (2003) G488.
Q. Chen. W. Z. Zhou. G. H. Du. L. H. Peng, Adv. Mater., 14 (2002) 1208.
P.Hoyer, Langmuir, 12 (1996) 1411.
B. B. Lakshmi, P. K. Dorhout. C. R. Martin. Chem. Mater., 9 (1997) 2971.

145

[187] A. Michalowski, D. Al Mawalawi, G. Cheng, M. Moskovits, Chem. Phys. Lett., 349 (2001) 1.
[188] Z. R. R. Tian, 1. A. Voigt. L. Liu. B. McKenzie. H. F. Xu, J. Am. Chem. Soc., 125 (2003)
12384.
[189] A.V. Chichagov, A.B. Belonozhko, A.L. Lopatin, Crystal Structure Data of Minerals,
Kristallographiya, 35 (1990) 610.
[190] J. M. Macak, K. Sirotna, P. Schmuki, Electrochim. Acta, 50 (2005) 3679.
[191] J. M. Macak, L. V. Taveira, H. Tsuchiya, K.Sirotna, J. Macak, P. Schmuki, J. Electroceram., 16
(2006) 29.
[192] J. M. Macak, H. Tsuchiya, P. Schmuki, Angew. Chem. Int. Ed., 44 (2005) 2100.
[193] L. V. Taveira, J. M. Macak, H. Tsuchiya, L. F. P. Dick, P. Schmuki, J.Electrochem. Soc., 152
(2005) B405.
[194] J. M. Macak, H. Tsuchiya, L. Taveira, S. Aldabergerova, P. Schmuki, Angew. Chem. Int. Ed. 44
(2005) 7463.
[195] J. M. Macak, P. Schmuki, Electrochimica Acta 52 (2006) 1258.
[196] R.Gilmont, CEP Magazine 10 (2002) 36.
[197] J.M. Macak, H. Hildebrand, U. Marten-Jahns, P. Schmuki, J. Electroanal. Chem., in press.
[198] S. Bauer, S. Kleber, P. Schmuki, Electrochem. Commun., 8 (2006) 1321.
[199] J. Park, S. Bauer, K. Von der Mark, P. Schmuki, Nano Letters, 7 (2007) 1686.
[200] M. Paulose, K. Shankar,S. Yoriya,H. E. Prakasam,O. K. Varghese, G. K. Mor, T.A. Latempa,A.
Fitzgerald, C. A. Grimes, J. Phys. Chem B, 110 (2006) 16179.
[201] S. Albu, A. Ghicov, J M. Macak, P. Schmuki, Phys. Stat. Sol. (RRL), 1 (2007) R65.
[202] J.M. Macak, S. Albu, P. Schmuki, Phys. Stat. Sol. (RRL), 1 (2007) 181.
[203] J.M. Macak, S. Aldabergerova, A. Ghicov, P. Schmuki, Phys. Stat. Sol. A, 203 (2006) R67.
[204] R. Hengerer, B. Bolliger, M. Erbudak, M. Grtzel, Surf. Sci., 460 (2000) 162.
[205] J. Kunze, A. Ghicov, H. Hildebrandt, J.M. Macak, L. Taveira, P. Schmuki, Z. Phys. Chem., 129
(2006) 1561.
[206] R. Beranek, H. Tsuchiya, T. Sugishima, J. M. Macak, L. Taveira, S. Fujimoto, H. Kisch, P.
Schmuki, Appl. Phys. Lett., 87 (2005) 243114.
[207] J. M. Macak, B. G. Gong, M. Hueppe, P. Schmuki, Adv. Mater. 19 (2008) 3027.
[208] H. Tokudome, M. Miyauchi, Angew. Chem. Int. Ed., 44 ( 2005) 1974.
[209] L. A. Lyon, J.T. Hupp, J. Phys. Chem. B, 103 (1999) 4623.
[210] D.D. Cronemeyer, Phys. Rev., 113 (1959) 1222.
[211] P.J.D. Lindan, N.M. Harrison, M.J. Gillan, J.A. White, Phys. Rev. B, 55 (1997) 15919.
[212] N. V. Mung, D. Y. Park, M. Schwartz, K. Nobe, in Pits and Pores: Formation, Properties and
Significance for Advanced Materials, Proc. Electrochem. Soc, Pennington, 2000, PV 25, 333.
[213] W.D. Kingery, H.K. Bowen, and D.R. Uhlmann, Introduction to Ceramics, New York (Wiley)
1976.
[214] B. P. Rai, Solar Cells, 25 (1998) 265.
[215] H. Habazaki, K. Fushimi, K. Shimizu, P. Skeldon, G.E. Thompson, Electrochem. Commun., 9
(2007) 1222.
[216] K. Yasuda, J. M. Macak, S. Berger, A. Ghicov, P. Schmuki, J. Electrochem. Soc., 154 (2007)
C472.
[217] G.K. Mor, O.K. Varghese, M. Paulose, N. Mukherjee, C.A. Grimes, J. Mater. Res. 18 (2003)
2588.
[218] K.Yasuda, P. Schmuki, Electrochim. Acta 52 (2007) 4053.
[219] L. Taveira, J.M. Macak, K. Sirotna, L.F.P. Dick, P. Schmuki, J. Electrochem. Soc., 153 (2006)
B137.
[220] K.S. Raja, M. Misra, K. Paramguru, Electrochim. Acta, 51 (2005) 154
[221] W.-J. Lee, M. Alhoshan, W.H. Smyrl, J. Electrochem. Soc., 153 (2006) B499.
[222] J. Zhao, X. Wang, R. Chen, L. Li, Solid State Commun., 134 (2005) 705.
[223] F. M. Bayoumi, B.G. Ateya, Electrochem. Commun., 8 (2005) 38.
[224] J.M.Macak, S. Albu, P. Schmuki, MRS Proceeding, 963 (2007) 963-Q11-06.
[225] P. Atkins, J. de Paula, Elements of Physical Chemistry, Oxford, 2005, p. 270
[226] J.V. Macpherson, P.R. Unwin, Anal. Chem., 69 (1997) 2063.

146

[227] S. J. Garcia-Vergara, P. Skeldon, G. E. Thompson, and H. Habazaki, Electrochim.Acta, 52 (200)


681.
[228] V. G. Levich, Physicochemical Hydrodynamics, London (Prentice Hall), 1962.
[229] N. Sato, Electrochemistry at Metal and Semiconductor Electrodes, Amsterdam (Elsevier), 1998.
[230] J. M.Macak, H. Tsuchiya, S. Berger, S. Bauer, S. Fujimoto, P. Schmuki, Chem. Phys. Lett., 482
(2006) 421.
[231] J. Zhao, X. Wang, T. Sun, L. Li, Nanotechnology 16 (2005) 2450.
[232] H. Zhang, J. Banfield, J. Mater. Chem. 8 (1998) 2073.
[233] J.M. Macak, H.Tsuchiya, A. Ghicov, K. Yasuda, R. Hahn, S. Bauer, Curr. Opin. Solid State
Mater. Sci. (2007), in press
[234] J. M. Macak, H. Tsuchiya, A. Ghicov, P. Schmuki, Electrochem. Commun. 7 (2005) 1138
[235] M. Paulose, K. Shankar, O.K. Varghese, G.K. Mor, B. Hardin, C.A. Grimes, Nanotechnology
17 (2006) 1446.
[236] H. Wang, C.T. Yip, K.Y. Cheng, A.B. Djurisic, M. H. Xie, Y. H. Leung, W.K. Chan, Appl.
Phys. Lett. 89 (2006) 023508.
[237] K. Zhu, N. R. Neale, A. Miedaner, A. J. Frank, Nano Letters 7 (2007) 69
[238] R. Hahn, T. Stergiopoulus, J. M. Macak, D. Tsoukleris, A. G. Kontos, S. P. Albu, D. Kim , A.
Ghicov, J. Kunze, P. Falaras, P. Schmuki, Phys. Stat. Sol. (RRL) 1 (2007) R135.
[239] A. Ghicov, J.M. Macak, H. Tsuchiya, J. Kunze, V. Heaublein. L. Frey, P.Schmuki, Nano Letters
6 (2006) 1080.
[240] J. M. Macak, A. Ghicov, R. Hahn, H. Tsuchiya, P. Schmuki, J. Mater. Res. 21 (2006) 2824.
[241] K. Shankar, K.Ch. Tep, G. K. Mor, C. A. Grimes, J. Phys. D: Appl. Phys. 39 (2006) 2361.
[242] H. Tsuchiya, J. M. Macak, L. Muller, J. Kunze, F. Muller, S. P. Greil, S. Virtanen, P. Schmuki,
J.Biomed.Mat.Res., 77A (2006) 534.
[243] S. Oh, R. Finones, Ch. Daraio, L. Chen, S. Jin, Biomaterials 26 (2005) 4938.
[244] J.M. Macak, M. Zlamal, J. Krysa, P. Schmuki, Small, 3 (2007) 300.

147

9. LIST OF SYMBOLS
Acronyms
AES
AFM
AR
AFM
CB
EDX
EIS
FB
IPCE
OCP
PAA
R
RS
SEM
TEM
UV
VB
VIS
XPS
XRD

Auger Electron Spectroscopy


Atomic Force Miscroscopy
Aspect ratio
Atomic Force Microscope
Conduction Band
Energy Dispersive X-ray Spectroscopy
Electrochemical Impedance Spectroscopy
Flat band
Incident photon-to-electron efficiency
Open-circuit potential
Porous anodic alumina
Pilling-Bedworth ratio
Raman Spectroscopy
Scanning Electron Microscope
Transmission Electron Miscroscope
Ultraviolet light
Valence band
Visible light
X-ray Photoelectron Spectroscopy
X-ray Diffractometry

Roman letters
c
C
d
D
E
Eg
eV
F
G
h
j
J
k
kb
l
m
M
Me
NA
Nd
q
S
SHE
t
T
U
V
W
z

Molar concentration
Capacitance
Diameter
Diffusion constant
Standard electrode potential
Band gap energy
Electron-volt
Faradays constant or field strength
Gibbs energy
Plancks constant
Current density
Flux of species
Growth constant
Boltzmann constant
Thickness
Mass
Molecular weight
Metal
Avogadros number
Doping concentration
Electric charge or charge of an electron
Surface area
Standard hydrogen electrode
Transport number or thickness
Thermodynamic temperature
Potential
Volume
Space charge layer width
Number of exchanged electron

148

Greek letters

Absorption coefficient
Diffusion layer
Dielectric constant
Viscosity
Ionic conductivity
Wavelenght
Frequency or Poissons ratio
Density
Mechanical stress
Potential difference
Time

149

Acknowledgements

My deepest gratitude goes to my Doktorvater, Prof. Dr. Patrik Schmuki, who has
made this thesis possible. Having the advantage of being his PhD student, I spent several
fortunate past years of my life in his laboratory, full of scientific enthusiasm, and we enjoyed
together many bright moments.
For the first three-year period financial support I would like to thank to the Bavarian
Ministry of Education, for my overall sponsorship I am grateful to the University of ErlangenNuremberg.
I would like to thank also to my former colleagues and friends at the chair, namely Dr.
Thierry Djenizian and Dr. Luciano Taveira for their extensive help, particularly in the early
stages of my thesis and to Dr. Hiroaki Tsuchiya for his great contribution to my scientific
outcome, for his kindness and all-time friendship and support.
Additionally, I would like to thank numerous other present and former members from
the chair. These include Prof. Dr. Sannakaisa Virtanen (for her various support and
particularly proofreading of all manuscripts over the years) and my other colleagues Andrei
Ghicov, Michael Hueppe, Robert Hahn, Johannes Brunner, Sebastian Kleber, Florian Kellner,
Yan Zhang, Dr. Udo Schlierf, Dr. Stefan Maupai and Dr. Herman Kaiser.
I am thankful to a number of colleagues technicians who helped me with various
measurements and different other problems. Namely, Mrs. Ulrike Marten-Jahns, Mrs. Helga
Hildebrand and Mrs. Anja Friedrich are acknowledged for XRD, XPS and SEM
investigations. Technical help of Mr. Hans Rollig, Mr. Martin Kolacyak, Mrs. Ingrid Tontsch
and Mr. Karl Werner is highly appreciated.
Furthermore, I would like to thank to my friend and former class-mate Dr. Radim
Bernek for his open mind and bright attitude to my work. I am also grateful to my former
boss at ITC Prague, Dr. Jan Mack, who kindly released me from his lab, when I got the
chance to study in Germany.
The English proof-reading by Prof.Dr.David Burleigh is greatly appreciated.
Last, but not least, I would like to thank my parents, my sister and my wife Marta for
their sustained support, patience and love over the years. Without them I would have never
gotten where I am.

Curriculum Vitae
Name:

Jan Mack

Address:
Date of birth:
Place of birth:
Nationality:

Rotkappenweg 2, 91058 Erlangen, Germany


15.08.1979
Pardubice (Czech Republic)
Czech

Tel:
E-mail:

+420777268358
janmacak@seznam.cz

Education:
2004-2008

PhD study
Chair for Surface Science and Corrosion,
Department of Materials Science,
University of Erlangen Nuremberg (Germany)

1998-2003

Diploma Engineer in power engineering


Institute of Chemical Technology in Prague (Czech Republic

1994-1998

Maturity certificate in analytic chemistry


College of Chemistry in Pardubice (Czech Republic)

Main research interest:


Material science, electrochemistry, nanotechnology, thin film characterization,
corrosion, semiconductor chemistry
International experience:

Stay in Steamboat Springs (Colorado, US) 5 months during summer 2001


Stay at FAU Erlangen (Germany) from February till June 2002 (Diploma thesis)
Stay at UFRGS Porto Alegre (Rio Grande do Sul, Brazil) from October till December
2004 (visiting scientist)

Miscelaneous:

Participation Award of the 56th Meeting of Nobel Prize Winners in Chemistry, Lindau
2006
Georg Kurlbaum Prize 2007

Languages:

Czech (mother tongue), English

Publication list
Status of February 2008
1.
A. Ghicov, H. Tsuchiya, J.M.Macak, P.Schmuki
Titanium oxide nanotubes prepared in phosphate electrolytes
Electrochemistry Communications, 7 (2005) 505-509
2.
H.Tsuchiya, J.M. Macak, L.Taveira, E. Balaur, A. Ghicov, K.Sirotna and P.Schmuki
Self-organized TiO2 nanotubes prepared in ammonium fluoride containing acetic acid electrolytes
Electrochemistry Communications 7 (2005) 576-580
3.
J.M.Macak, H.Tsuchyia, A Ghicov and P.Schmuki
Dye-sensitized Anodic TiO2 nanotubes
Electrochemistry Communications 7 (2005) 1138-1142
4.
E. Balaur, J.M. Macak, L. Taveira, P. Schuki
Tailoring the wettability of TiO2 nanotube layers
Electrochemistry Communications 7 (2005) 1066-1070
5.
A. Ghicov, H. Tsuchiya, R. Hahn, J.M. Macak, A.G. Munoz, P. Schmuki
TiO2 nanotubes: H+ insertion and strong electrochromic effects
Electrochemistry Communications, 8 (2006) 528-532

6.
R. Vitiello, J.M. Macak. H. Tsuchiya, A. Ghicov, L.F.P. Dick, P. Schmuki
N-doping of anodic TiO2 nanotubes using heat treatment in ammonia
Electrochem. Communications, 8 (2006) 544-548
7.
J.M. Macak, F. Schmidt-Stein, P. Schmuki
Efficient Oxygen Reduction on Layers of Ordered TiO2 Nanotubes Loaded with Au nanoparticles
Electrochemistry Communications, 9 (2007) 1783-1787
8.
M. Zlamal, J.M.Macak, P. Schmuki, J. Krysa
Electrochemically Assisted Photocatalysis on Self-Organized TiO2 nanotubes
Electrochemistry Communications, 9 (2007) 2822-2826
9.
H. Tsuchiya, S. Berger, J. M. Macak, A. Ghicov, P. Schmuki
Self-organized porous and tubular oxide layers on TiAl alloys
Electrochemistry Communications, 9 (2007) 2397-2402
10.
X. Feng, J.M. Macak, P. Schmuki,
Flexible self-organization of two-size scales oxide nanotubes on Ti45Nb alloy
Electrochemistry Communications, 9 (2007) 2403-2407

11.
H. Tsuchiya, S. Berger, J. M. Macak, A. Ghicov, A.G. Munoz, P. Schmuki
A new route for the formation of self-organized anodic porous
alumina in neutral electrolytes
Electrochemistry Communications, 9 (2007) 545-550
12.
R. Hahn, J. M. Macak, P. Schmuki
Rapid anodic growth of TiO2 and WO3 nanotubes in fluorid-free electrolytes
Electrochem. Communications, 9 (2007) 947-952
13.
I. Paramasivam, J. M. Macak, P. Schmuki,
Photocatalytic activity of TiO2 nanotube layers loaded with Ag and Au nanoparticles
Electrochemistry Communications, 10 (2008) 71-75
14.
A.Ghicov, J. M. Macak, H. Tsuchiya, J. Kunze, V. Haeublein, S. Kleber and P. Schmuki,
TiO2 nanotube layers: Dose effects during nitrogen doping by ion implantation
Chem. Phys. Lett. 419 (2005) 426-429
15.
I. Paramasivam, J.M. Macak, A. Ghicov, P. Schmuki
Enhanced Photochromism of Ag Loaded Self-Organized TiO2 Nanotube
Layers
Chemical Physics Letters, 445 (2007) 233-237
16.
H. Tsuchiya, J. M. Macak, A. Ghicov, Y. Ch. Tang, S. Fujimoto,
M. Niinomi, T. Noda and P. Schmuki,
Nanotube oxide coating on Ti29Nb13Ta4.6Zr alloy prepared by
self-organizing anodization
Electrochimica Acta, 52 (2006) 94-101
17.
J.M. Macak, H. Tsuchiya, A. Ghicov, K. Yasuda, R. Hahn, S. Bauer, P.Schmuki,
Review: TiO2 nanotubes: Self-organized electrochemical formation,
properties and applications
Current Opinion in Solid State and Materials Science, in press.
18.
H.Tsuchiya, J.M. Macak, A. Ghicov, L. Taveira, P.Schmuki
Self-organized porous TiO2 and ZrO2 produced by anodization
Corrosion Science 47 (2005) 3324-3335
19.
H. Tsuchiya, J.M. Macak, A. Ghicov, A. S. Rder, L. V. Taveira, P. Schmuki
Characterization of electronic properties of TiO2 nanotube films
Corrosion Science, 49 (2007) 203-210
20.
R. Hahn, A. Ghicov, H. Tsuchiya, J. M. Macak, A.G. Munoz, P. Schmuki
Lithium-ion insertion in anodic TiO2 nanotubes resulting in high electrochromic contras
Physica Status Solidi (a), 207 (2007) 1281-1285.
21.
A.Ghicov, H. Tsuchiya, J. M. Macak, P. Schmuki
Annealing effects on the photoresponse of TiO2 nanotubes
Physica Status Solidi A 203 (2006) R28-R30.

22.
R. Hahn, T. Stergiopoulus, J. M. Macak, D. Tsoukleris, A. G. Kontos, S. P. Albu, D. Kim , A. Ghicov, J.
Kunze, P. Falaras, P. Schmuki
Dye-sensitization of perchlorate tubes efficient solar energy conversion
using TiO2 nanotubes produced by rapid breakdown anodization a comparison
Physica Status Solidi (RRL), 1 (2007) 135-137.
23.
H.Tsuchiya, J.M. Macak, I.Sieber and P.Schmuki,
Electrochemical Formation of Self -Organized High Aspect Ratio Porous Zirconium Oxide
Small 1 (2005) 722-725
24.
H. Tsuchiya, J. M. Macak, A. Ghicov and P. Schmuki,
Self-organization of anodic oxide nanotubes on two size-scales
Small, 2 (2006) 888-891
25.
J.M.Macak, M. Zlamal, J. Krysa, P. Schmuki
Self-Organized TiO2 Nanotube Layers as Highly Efficient Photocatalyst
Small, 3 (2007) 300-303
26.
H. Tsuchiya J. M. Macak, L. Mller, J. Kunze, F. Mller, P. Greil, S. Virtanen, P. Schmuki
Hydroxyapatite growth of anodic TiO2 nanotubes
Journal of Biomedical Research 77A (2006) 534-541.
27.
J. M.Macak, Hiroaki Tsuchiya, Luciano Taveira, Andrei Ghicov and Patrik Schmuki
Self-Organized Nanotubular Oxide Layers on Ti-6Al-7Nb and Ti-6Al-4V Formed by Anodization in
NH4F Solutions
Journal of Biomedical Research 75A (2005) 929-933.
28.
L.Taveira, J.M.Macak, K.Sirotna, L.F.P.Dick, P.Schmuki,
Voltage oscillations and morphology during the galvanostatic formation of self-organized TiO2
nanotubes
Journal of Electrochemical Society, 153 (2006) B137-143
29.
J.M.Macak, S.Albu, D.G. Kim, I. Paramasivam, S. Aldabergerova, P. Schmuki
Multilayer TiO2 - Nanotube Formation by Two Step Anodization
Electrochemical Solid-State Letters, 10 (2007) K28-K31
30.
P. J. Barczuk, H. Tsuchiya, J. M. Macak, P. Schmuki, D. Szymanska, O. Makowski, K.Miecznikowski
and P. J. Kulesza
Enhancement of the Electrocatalytic Oxidation of Methanol at Pt/Ru Nanoparticles Immobilized in
Different WO3 Matrices
Electrochemical and Solid-State Letters, 9 (2006) E13-E16
31.
X. Feng, J.M. Macak, P. Schmuki
Robust Self-Organization of Oxide Nanotubes in Wide pH Range
Chemistry of Materials, 19 (2007) 1757-1759

32.
J.M. Macak, A. Ghicov, H. Tsuchiya, R. Hahn, P. Schmuki
Photoelectrochemical Properties of N-doped Self-Organized Titania Nanotube Layers with Different
Thicknesses
Journal of Materials Research, 21 (2006) 2624-2828.
33.
E. Balaur, J.M. Macak, H. Tsuchiya, P. Schmuki,
Wetting behaviour of layers of TiO2 nanotubes with different diameters
Journal of Materials Chemistry, 15 (2005) 4488-4491
34.
S. P. Albu, A. Ghicov, J.M. Macak, P. Schmuki
Self-Organized, Free-Standing TiO2 Nanotube Membrane for Flow-through Photocatalytic Applications
Nano Letters, 7 (2007) 1286-1289
35.
A. Ghicov, J.M.Macak, H. Tsuchiya, J. Kunze, V. Haeublein, L.Frey and P.Schmuki
Nitrogen-doped TiO2 nanotube arrays with strongly enhanced photoresponse
in the visible range
Nano Letters, 6 (2006) 1080-1082.
36.
J.M. Macak, H. Tsuchiya. S. Bauer, A. Ghicov, P. Schmuki, P. J. Barczuk, M.Z.
Nowakowska, M. Chojak, P.J. Kulesza
Self-organized Nanotubular TiO2 Matrix as Support for Dispersed Pt/Ru
Nanoparticles; Enhancement of the Electrocatalytic Oxidation of Methanol
Electrochemistry Communications 7 (2005) 1417-1422.
37.
J.M.Macak, K.Sirotn and P.Schmuki
Self-organized porous titanium oxide in Na2SO4 / NaF Electrolytes,
Electrochimica acta, 50 (2005) 3679-3684
38.
J.M.Macak, H.Tsuchyia and P.Schmuki
High Aspect Ratio TiO2 Nanotubes by Anodization of Ti
Angewandte Chemie Int. Edit. 44 (2005) 2100-2102
Angewandte Chemie 117 (2005) 2163-2166
39.
J.M.Macak, H.Tsuchyia, L.Taveira, S. Aldabergerova and P.Schmuki
Smooth Anodic TiO2 Nanotubes,
Angewandte Chemie Int. Edt. 44 (2005) 7463-7465
Angewandte Chemie, 117 (2005) 7629-7632
40.
J.M.Macak, H. Tsuchiya, L. Taveira, K.Sirotn, J.Macak and P.Schmuki
Influence of fluoride containg electrolytes on the formation of self-organized nanotubes by anodization.
Journal of Electroceramics, 16 (2006) 29-34
41.
L.Taveira, J.M.Macak, L.F.P.Dick, P.Schmuki,
Initiation and growth of self-organized TiO2 nanotubes anodically formed in NH4F/(NH4)2SO4
electrolytes
Journal of The Electrochemical Society 152 (2005) B405-B409

42.
J.M.Macak and P.Schmuki
Anodic growth of self-organized anodic TiO2 nanotubes in viscous electrolytes
Electrochimica Acta, 52 (2006) 1258-1264
43.
J.M. Macak, S. Aldabergerova, A. Ghicov, P. Schmuki
Smooth anodic TiO2 nanotubes: Annealing and Structure
Physica Status Solidi A 203 (2006) R67-R69.
44.
J. M. Macak, H. Tsuchiya, S. Berger, S. Bauer, S. Fujimoto and P. Schmuki
On Wafer TiO2 Nanotube-layer Formation by Anodization of Ti-films on Si
Chemical Physics Letters, 428 (2006) 421-425
45.
J.M. Macak, S. P. Albu, P. Schmuki
Towards Ideal Hexagonal Ordering of TiO2 Nanotubes
Physica Status Solidi (RRL) 1 (2007) 181-183.
46.
S. P. Albu, A. Ghicov, J.M. Macak, P. Schmuki
250 m long anodic TiO2 nanotubes with hexagonal self-ordering
Physica Status Solidi (RRL),1 (2007) R65-R67
47.
K. Yasuda, J.M. Macak, A. Ghicov, S. Berger, P. Schmuki
Mechanistic Aspects of the Self-Organization Process for Oxide Nanotube Formation on Valve Metals
Journal of Electrochemical Society, 154 (2007) C472-C478
48.
J.M. Macak, H. Hildebrand, U. Marten-Jahns, P. Schmuki
Mechanistic Aspects and Growth of Large Diameter Self-Organized TiO2 Nanotubes
Journal of the Electroanalytical Chemistry, in press.
49.
J.M. Macak, B.G. Gong, M. Hueppe, P. Schmuki
Filling of TiO2 nanotubes by self-doping and electrodeposition
Advanced Materials 19 (2007) 3027-3031
50.
R. Beranek, H. Tsuchiya, T. Sugishima, J.M.Macak, L. Taveira, S. Fujimoto, H. Kisch and P. Schmuki
Enhancement and limits of the photoelectrochemical response from anodic TiO2 nanotubes
Applied Physics Letters 87 (2005) 243114
51.
H.Tsuchiya, J.M. Macak, L.Taviera and P.Schmuki
Fabrication and characterization of smooth high aspect ratio zirconia nanotubes
Chemical Physics Letters 410 (2005) 188-191
52.
H.Tsuchiya, J.M. Macak, I.Sieber, L. Taveira, A. Ghicov, K.Sirotna, P.Schmuki,
Self - organized porous WO3 formed in NaF electrolytes,
Electrochemistry Communications, 7 (2005) 295-298
53.
J. Kunze, A. Ghicov, H. Hildebrand, J. M. Macak, L. Taveira, P. Schmuki,
Challenges in the surface analytical characterization of anodic TiO2 films
Zeitschrift fr Physikalische Chemie 219 (2005) 1561-1582

54.
H. Tsuchiya, J. M. Macak, I. Sieber and P. Schmuki,
Anodic Porous Zirconium Oxide Prepared in Sulfuric Acid Electrolytes
Materials Science Forum, 512 (2006) 205-208
55.
X. Feng, J. M. Macak, S. P. Albu, P. Schmuki
Electrochemical formation of self-organized anodic nanotube coating on Ti28Zr8Nb biomedical alloy
surface
Acta Biomaterialia, 4 (2008) 318-323
56.
W. Wei, J.M. Macak, P. Schmuki,
High Aspect Ratio Ordered Nanoporous Ta2O5 Films by Anodization of Ta
Electrochemistry Communications, 10 (2008) 428-432

You might also like