You are on page 1of 170

Nano Porous Alumina based Composite Coating for

Tribological Applications

A Thesis
Submitted for the Degree of

Doctor of Philosophy

In the Faculty of Engineering

By

Arti Yadav

Department of Mechanical Engineering


Indian Institute of Science
BANGALORE-560012

DECEMBER 2014
Dedicated to my parents, brother and family members…..
Acknowledgements
I would like to thank my advisor, Prof M.S. Bobji for his support, help and guidance
extended to me during the course of my research. It is from him, that I have imbibed the
logical thinking process which is fundamental to pursuing any research. The numerous
talks that I had with him regarding the topics related to my research, and also of general
and personal nature have transformed me not only as a researcher but also into a better
human being.

I would like to thank the Chairman, Prof. R. Narasimhan and all other faculty
members of the Mechanical Engineering Department for their willingness to help and for
being approachable. I also specially thank Prof. Jaywant Arakeri for his valuable
suggestions during my presentations. I also thank the office staff of the Department,
especially Mrs. Somavathi, Mrs. Banu, Mr. Nagaraja and Mrs Devaki for all the assistance
rendered to me.

Special thanks to my all lab mates Prashant, Jaydeep, Dilip, Muthukumar, Vikram,
shrikant, Lavanya, Sudipta and Sulochana. I thank my senior Anantheshwara for his
guidance in the initial stages. I thank all the present and past project assistants and
summer interns.

I am thankful to Shama Sunder H S for all the help and encouragement received
from him.

I am thankful to CENSE IISc for allowing me to use equipments.

I thank Texas Instruments (TI) for providing me free samples for my experiments. I
would like to extend my thanks to National Instrument (NI) for their software packages.

I bring to close with my appreciation to my friends especially Meenakshi, Sweeti,


Aswin, Arti Tripathi, Guganaeswram, Priyanka, Gauri, Anju Babu, Eswaramoorthy, Kuldeep,

i
Ajay, Shyam, Nibedita, Shayali and Dipti for their moral support. They were always with me
at each and every moment in all offices.

I am also really thankful to my younger brother Vishnu and my cousins Mithun,


Vimal , Saloni , Aditi , Vivek who have constantly spoken of encouragements and their love
over my life throughout the years.

Last but not least, I am thankful to my parents who have always encouraged me to
pursue my dreams even from a very young age. I dearly and deeply thank to them for all
the sacrifices they have made into making what I am today.

- Arti Yadav

ii
Abstract
Anodisation is a surface treatment process, commonly used to form a protective
oxide coating on the surface of metals like aluminium. Anodised coatings, being grown out
of the base metal have excellent interface strength but are porous and brittle. Porosity of
the coating reduces the hardness and the brittle nature of the oxide induces cracking. In
practice, the pores are typically filled with organic dye and sealed. Under certain controlled
electrochemical conditions, anodisation results in a highly ordered hexagonal porous
structure in pure aluminium. In this work, we explore the possibility of using this ordered
porous alumina to form a novel metal nanocomposite as a tribological coating. By
optimizing the nonporous structure and tuning the electrodeposition process, we
uniformly filled the ordered pores with copper. We have measured the hardness of the
resulting ordered and aligned nanocomposite. We explore the possibility of using this
composite coating for tribological applications by carrying out some preliminary
reciprocating wear test.

Ordered porous alumina layer is formed by a two-step anodisation process. By


optimizing the anodisation conditions, we control the thickness of the coating and the pore
size. The interface of the porous structure and aluminium substrate is defined by a non-
conducting dense barrier oxide layer. However, to deposit metal into the pores, a
conducting path should be established through the barrier layer. One possibility is to etch
out the bottom of the pores at the cost of the interface strength and losing out on the main
advantage of anodised coatings. To be able to fill metal without this sacrifice, we utilised
the dendritic structure in the barrier layer formed by a step-wise reduction of voltage
towards the end of anodisation process. Optimisation of this dendritic structure led to
uniform deposition of metal into pores, achieved by pulsed electrodeposition. In pulse
electrodeposition, a positive pulse is applied to remove accumulated charge near to the

iii
bottom of pores, followed by a negative pulse to deposit metal and a delay to allow
diffusion of ions. By optimising the pulse shape and duration, we have achieved uniform
growth of metal into pores. Further, monitoring the deposition current helped us to
identify and control different phases of growth of the nanowire.

The properties of the porous alumina and the nanocomposite were measured by
nanoindentation. The deformation characteristics were obtained by observing the indents
in a FE-SEM. We find that dendritic modification of interface has very little effect on the
hardness of the porous alumina layer. We also found that the porous alumina deformed
either by compaction or by forming circumferential and radial cracks. When copper is filled
in the nano pores, the hardness increased by 50% and no circumferential cracks were
found up to the load of 10 mN for a film thickness of about 1 µm. Coefficient of friction of
the coating reciprocated against steel in dry condition is found to be around 0.4. Minimal
wear was observed from the SEM images of wear track.

In summary, a novel nanocomposite coating with ordered porous alumina as matrix


embedded with aligned metal nano rods has been developed. This was achieved by
optimally modifying the barrier layer without sacrificing the interfacial strength. Uniform
coating has been achieved over an area of 10 mm x 10 mm. The coating is found to have
high hardness and high wear resistance.

iv
Contents

1. Introduction ................................................................................................................................................... 1
1.1 Anodisation of Aluminium .................................................................................................................. 1
1.2 Formation of Porous Alumina............................................................................................................ 3
1.3 Controlling Geometry of the Pores .................................................................................................. 8
1.3.1 Anodizing Voltage ....................................................................................................................... 9
1.3.2 Concentration and type of the Electrolyte ..................................................................... 11
1.3.3 Anodisation Temperature .................................................................................................... 12
1.4 Two-step Anodisation ........................................................................................................................ 13
1.5 Ordered Porous Structure as a Template ................................................................................... 14
1.6 Characterization ................................................................................................................................... 17
1.6.1 Morphology ................................................................................................................................ 17
1.6.2 Electrochemical Impedance Spectroscopy .................................................................... 19
1.6.3 Mechanical Properties ........................................................................................................... 21
1.7 Scope of Work ....................................................................................................................................... 25
1.8 Aim and Objectives.............................................................................................................................. 25
1.9 Organization of Thesis ....................................................................................................................... 26
2. Formation of Ordered Pores ................................................................................................................. 28
2.1 Introduction ........................................................................................................................................... 28
2.2 Electro Polishing of Aluminium ..................................................................................................... 29
2.3 First Anodisation.................................................................................................................................. 32
2.4 Chemical Etching for Patterning .................................................................................................... 35
2.5 Second Anodisation ............................................................................................................................ 37
2.6 Pore-widening....................................................................................................................................... 41
2.7 Choosing the Anodising Parameters ............................................................................................ 42
2.7.1 Anodisation Voltage ................................................................................................................ 43

v
2.7.2 Type of Electrolyte .................................................................................................................. 45
2.7.3 Effect of Temperature Control ............................................................................................ 47
2.8 Summary ................................................................................................................................................. 50
3. Controlling Barrier Oxide Layer .......................................................................................................... 51
3.1 Introduction ............................................................................................................................................. 51
3.2 Experimental Details ............................................................................................................................. 54
3.3 Different Methods of Barrier Thinning .......................................................................................... 56
3.3.1 Chemical Etching ............................................................................................................................ 56
3.3.2 Cathodic Polarization .................................................................................................................... 58
3.3.3 Step-wise Voltage Reduction Method..................................................................................... 60
3.4 Controlling Dendrite Geometry ........................................................................................................ 62
3.4.1 Influence of Voltage Reduction Rate on Dendrite Structure ......................................... 63
3.4.2 In situ Capacitance Measurements .......................................................................................... 69
3.4.3 Effect of Pore Widening ............................................................................................................... 71
3.5 Impedance Spectroscopy of Porous Alumina .............................................................................. 75
3.5.1 Equivalent Circuits Model Analysis ......................................................................................... 75
3.5.2 Impedance Analysis ....................................................................................................................... 77
3.6 Summary .................................................................................................................................................... 80
4. Electrochemical Deposition of Metal into Porous Alumina Film ............................................ 81
4.1 Introduction ............................................................................................................................................. 81
4.2 Experimental Details ............................................................................................................................. 82
4.3 Electrodeposition of Copper .............................................................................................................. 84
4.3.1 DC Deposition .................................................................................................................................. 84
4.3.2 Voltage Controlled AC Electrodeposition.............................................................................. 85
4.3.3 Voltage Controlled Pulse Electrodeposition ........................................................................ 88
4.4 Effect of Dendrite Geometry on Deposition ................................................................................. 91
4.5 Electrodeposition of Silver and Nickel ......................................................................................... 102
4.6 Discussion ............................................................................................................................................... 103
4.7 Summary .................................................................................................................................................. 106
5. Measurement of Hardness by using Nanoindentation ............................................................. 107

vi
5.1 Introduction ........................................................................................................................................... 107
5.2 Experimental Details ........................................................................................................................... 108
5.3 Load-Displacement Characteristics .............................................................................................. 111
5.3.1 Porous Alumina ............................................................................................................................. 111
5.3.2 Porous Alumina with Modified Interface ............................................................................ 115
5.4 Effect of Film Thickness ..................................................................................................................... 118
5.5 Hardness of Film ................................................................................................................................... 121
5.6 Indentation on Nanocomposite ...................................................................................................... 124
5.7 Discussion ............................................................................................................................................... 125
5.8 Conclusions ............................................................................................................................................. 127
6. Tribological Characterization of Porous Alumina based Nanocomposite ........................ 128
6.1 Introduction ........................................................................................................................................... 128
6.2 Experimental Details ........................................................................................................................... 129
6.3 Friction ..................................................................................................................................................... 132
6.4 Wear Mechanisms ................................................................................................................................ 134
6.4.1 Wear Track on Pure Aluminium ............................................................................................. 135
6.4.2 Wear Track on Porous Alumina .............................................................................................. 136
6.4.3 Wear Track on Porous Alumina/Copper Nanocomposite ............................................ 137
6.5 Discussion ............................................................................................................................................... 141
6.6 Summary .................................................................................................................................................. 143
7. Conclusions ................................................................................................................................................ 144
Appendix……………………………………………………………………………………………………………..……..146
Appendix A Electrochemical Impedance spectroscopy (EIS) ………………………………..….146
Bibliography……………………………………………………………………………….……………………………...149

vii
Chapter 1

1. Introduction
1.
1.1 Anodisation of Aluminium
Anodisation is a surface treatment process, which is generally used to form a
protective oxide coating to improve the surface properties [1]. Anodisation, of aluminium
and its alloys, is commonly used in the industries to protect the surface from corrosion [2].
Anodisation refers to electrolytic oxidation of metal in which chemical conversation to
oxide takes place on the surface of metal. Anodisation is performed in water based
electrochemical solution, and is controlled by various parameters such as pH, temperature,
electrolyte and applied voltage [3]. The chemical reaction taking place at the anode is

(1)

Two types of oxide coatings are formed during anodisation of aluminium. One is
compact oxide layer and the other is porous layer. The formation of both these type of
structures depends upon the forming acids and anodisation conditions [4]. Compact oxide
layer on aluminium substrate is highly uniform in thickness and relatively dense. This type
of film is formed in nearly neutral electrolytes and can protect the surface from corrosion
[5,6]. Film growth at constant current density develops a constant electric field of the order
of 106 -107 V/cm across the layer [4]. In order to maintain this electric field, the voltage
across the layer must increase as the thickness of this oxide layer increases [7]. Beyond a
critical thickness of the film voltage exceeds the breakdown limit and the uniform film
1
formation is terminated [4]. Although compact oxide layer have excellent corrosion
resistance, the maximum thickness of compact oxide layer that can be obtained is thus
limited.

Porous oxide layer formed on aluminium substrate is not limited in thickness as the
compact layer. This type of layer is readily formed in certain electrolytes in which oxide
layer is slightly soluble [2]. The thickness of this layer is time dependent and can grow
many times the upper limit of compact oxide layer [2]. Thickness depends upon the current
density, time of electrolyte and temperature. In industry, anodisation is extensively used to
protect aluminium alloy components against corrosion and abrasion. The porosity of the
coating is dealt by sealing the pores by exposing the oxide films to hot water or steam.
Typically some dyes are used to fill the pores before sealing them [2,8].

Under certain controlled conditions, a highly ordered porous alumina can be formed
on pure aluminium. Keller et al [9] were first to use Transmission Electron Microscopy
(TEM) to observe a hexagonal closed packed structure. In 1970, O’Sullivan and Wood [10]
explained the growth mechanism of porous alumina and reported that the pores structure
primarily depends on anodizing voltage. In 1995, Masuda and Fakuda [11] introduced a
two-step process, to get highly ordered porous alumina. Since then, porous alumina has
attracted significant interest in nanotechnology as it is seen as ideal scaffold for the
synthesis of low numerous, low dimensional nano-scale materials [12–18].

In this work, we explore the possibility of using porous alumina film as a nano-
composite coating. For this we have fabricated ordered porous alumina on pure aluminium
substrates using two-step anodisation process. The pores in the attached oxide films are
then filled with metal uniformly by electrodeposition. For this the structure of the bottom
of the pores has to be carefully tuned to make it electrically conducting enough without
mechanically weakening the interface of the resulting coating. We evaluate the nano-
composite coating by measuring hardness using nanoindentation and measuring friction
and wear in a reciprocating wear test.

2
1.2 Formation of Porous Alumina
Porous alumina formation can be carried out in either galvanostatic or
potentiostatic mode depending on whether applied current density or the applied potential
is kept constant respectively. The variation of voltage and current in galvanostatic and
potentiostatic mode while anodising in 20% H2SO4 at 1 0C is shown in figure 1.1 [3]. In the
galvanostatic mode, initially voltage increases with time (stage a) representing gradual
growth of dense oxide layer. As the thickness of this compact layer increases, nucleation of
pores happens on its surface (stage b). The potential reached a maximum and as the pores
propagates into the oxide layer (stage c) starts decreasing. Eventually a steady state voltage
is reached and the thickness of the generated ordered porous alumina keeps growing. In
potentiostatic mode (figure 1.1B) the initial high current density decreases to a minimum
as the pores are formed in the compact oxide layer and reaches a steady state value during
the steady growth of ordered film.

Figure 1.1 Schematic of the voltage and current transient in (A) Galvanostatic and (B)
Potentiostatic modes [3,19].

The two basic chemical processes that take place during anodisation of aluminium
are oxidation of aluminium and the dissolution of the formed alumina [12]. The aluminium
ions form and gets uniformly distributed at the aluminium – oxide interface.

3
(2)

These aluminium ions gets oxidised either by oxygen ions or hydroxyl ions.

(3)

(4)

Either the aluminium cation has to diffuse through interstitial in the oxide layer to
the electrolyte [20] or the anions (oxygen or hydroxide) have to diffuse through the layer
to the aluminium metal (figure 1.2). The necessary oxygen and or hydroxyl ions are
supplied from the electrolyte by water splitting reaction.

(5)

(6)

Negative O2- ions can also form from absorption of OH- ions at the electrolyte
interface [20–22]. Under steady state, the rate of these reactions deter mines the thickness
and morphology of the resulting oxide film.

Figure 1.2: Schematic of O2- and OH- ions formation at the oxide/electrolyte interface from
water interacting with SO42- anions [10].
4
In near neutral solutions of boric acid, ammonium borate, ammonium tartrate or
ammonium tetra borate, the rate of water splitting reaction (equation 5) is slow. As the
compact oxide layer grows, the aluminium, oxygen and hydroxyl ions have to transit
through the existing oxide layer. The currently held view is that all the three ions move
through the oxide [23,24] . The movement of these ions is governed by high field
conduction equation,

(7)

where A and B are the constants for a given temperature and j is the current density
contributed by any one of the ion’s transport across the oxide film of thickness t under an
external applied field E. The electric field can be approximated to

⁄ (8)

where U is the potential across the film. In galvanostatic mode, the voltage will keep
on increasing as the oxide layer thickness (t) increases, such that a constant electric field of
the order 106 to 107 V/cm [4,23]. This will ultimately lead to dielectric breakdown. In
potentiostatic mode, since the voltage applied is constant the electric field across the oxide
layer will gradually drop as the oxide thickness increases. The corresponding current
density and hence the transport of the ions through oxide drops (figure 1.1b). The
oxidation will eventually stop when the oxide film thickness reaches a critical value (tc)
[2,24].

In acidic solutions, similar process takes place but with increased dissolution of
alumina at the oxide electrolyte interface (equation 4). This result in an equilibrium oxide
film thickness that is greater than the stable oxide film formed in the neutral solutions.
However in the acidic solutions any local variation in the dissolution rates due to
inhomogeneities present in the system will result in pores being formed. Increased oxide
dissolution could occur at the locations of defects such as impurities dislocation, grain
boundary, non-metallic inclusions or roughness [4,24,25]. It is also possible that electro-

5
dynamic convection resulting from local current and density fluctuations could also lead to
formation of a distinct patterns of local dissolution of the oxide [26–28]. Whatever may be
the mechanism of the pore nucleation, once formed the pores will grow because the local
electric field is enhanced due to the decrease in the oxide layer thickness.

Higher electric field near the pores lead to higher field assisted dissolution [29] .
That is both the rate of oxide formation at the metal-oxide interface and the rate of
dissolution at the oxide electrolyte interface are enhanced. Many other mechanisms [3]
have been proposed for the enhanced dissolution ranging from increased temperature due
to joule heating [2,9,10,30,31]. Evaporation of electrolyte leading to local melting of
aluminium [32] formation of gel like structure consisting of hydroxide and hydrate
compounds, proton assisted dissolution [33].

Under potentiostatic mode, transient current density (figure 1.3) can be thought of
as a resultant of two distinct mechanisms [33]. Dense barrier film formation current
decreases exponentially with time (equation 7) while the pore formation current increases
initially and reaches a steady state of constant pore growth.

Figure 1.3: Schematic of overlapping processes occurring during the porous alumina the
porous alumina growth under the potentiostatic mode [33].

Once the pores are formed in the oxide layer, the local electric field is enhanced due
to the reduced thickness of the insulating oxide layer. Enhanced electric field at the bottom

6
of the pores results in oxide-metal interface assuming the shape of the pore [24]. The
dissolution and formation of oxide then proceeds to keep the electric field constant. This
results in a cylindrical hole with a hemispherical bottom (figure 1.4). If the pores are far
apart they are free to grow both in diameter and depth and hence ultimately approach each
other. Once the aluminium between the side walls of the pores is exhausted the diametrical
growth ceases and pores increase only in depth. If the pores are nucleated closer than the
reduced field at the section AB’ in figure 1.4 H will move the pores apart. This model of
uniform pore formation is known as equifield strength model [24].

Figure 1.4: Schematic of pores growth [24].

It has also been proposed that along with a short-range order decided by the
equifield strength model, there has to be a long range ordering mechanism to explain the
ordered pores over a large area. As aluminium gets oxidised to alumina the volume can
almost double. This increased volume can be accommodated only by expansion of oxide in
vertical direction [34], since the oxidation takes place only at the bottom of the pore. This
could result in mechanical stresses in the film. Since the formation of porous alumina is
accompanied by dissolution of oxide as well, the effective volume expansion will be less

7
than 2. It has been found that the ordering of pores occur for the volume expansion ration
of about 1.2. [35].

1.3 Controlling Geometry of the Pores


Typical geometrical structure of the ordered porous alumina is characterised by
cylindrical pores arranged in a closed packed hexagonal pattern (figure 1.5). The diameter
(dp) of these pores can range from few tens of nanometre to hundreds of nanometre. The
bottom of the pores is closed by a spherical cap of barrier layer. The thickness of this
barrier layer (t) is dependent on the maximum electric field developed across the barrier
layer and hence is related to applied anodisation voltage.

Figure 1.5: (a) Schematic structure of porous alumina and (b) schematic cross-sectional view
of the anodized oxide layer [3].

The depth of the vertically aligned pore channel is determined mainly by the
anodising time. It can even exceed 100 micrometre [36,37]. However, it has been observed

8
by some researchers that the ordered domain size will decrease for very long (>30hr)
times of anodisation [35].

1.3.1 Anodizing Voltage

Anodisation voltage is the main parameter which affects the pore diameter.
O’Sullivan and Wood [10] have shown that diameter, is linearly proportional to the
anodisation voltage.

(9)

Where U is the applied anodic voltage, Dp is the pore diameter andp is the
proportionality constant. The interpore distance (Dint) and the barrier layer thickness (t)
also follow the linear trend. While anodising in phosphoric acid they obtained the
proportionality constants for pore diameter, inter pore distance and barrier layer thickness
as 1.29 nm/V, 2.77 nm/V and 1.04 nm/V respectively (figure 1.6).

Figure 1.6: The variation of (a) the pore diameter (b) the interpore distance on the anodising
voltage [10].

9
Nielsch et al [35] reported that the pore diameter formed by anodisation under the
optimized self-ordering conditions [21,35], leading to a quasi-perfect hexagonal
arrangement of pores can be calculated using the following equation:

( 10 )
√ √

where U is the applied anodic voltage, Dp is the pore diameter and k is the proportionality
constant (~2.5) and P is the porosity.

Figure 1.7: Effect of anodizing voltage on barrier thickness for porous alumina formed at
various electrolytes [38].

Chu et al [38] evaluated the barrier layer thickness after anodisation in sulphuric
acid, glycolic acid, tartaric acid, malic acid and phosphoric acid solutions. They deduced a
general and constant relationship between the barrier layer thickness and anodizing
voltage as given in figure 1.7. Lee et al [36] found that for common conditions of porous
alumina formation in oxalic acid the barrier layer thickness is about 1.3 nm/V.

10
1.3.2 Concentration and type of the Electrolyte

The rate of dissolution of oxide in the electrolyte is dependent on electrolyte. Higher


dissolution rate will reduce the barrier layer thickness. The pore diameter increases with
increase in pH of the solution and the aggressiveness of the electrolyte [2,10]. It has been
reported that the interpore distance decreases using high concentration of sulphuric acid
in a mixture sulphuric acid-oxalic electrolyte at constant pH [39]. They found that interpore
distance is almost proportional to 2.6 nm/V and 2.5 nm/V for the electrolyte of
oxalic-sulphuric mixture with low (0.206) and high (0.214) sulphuric concentrations at
constant pH of 0.57 (figure 1.8). They also reported that the concentration of the electrolyte
affects the ordering of porous alumina. For low sulphuric acid concentration, the ordering
was low because of lower growth rate and for high sulphuric concentration higher ordering
was observed.

Figure 1.8: Effect of pH on interpore distance at various anodizing voltage [39].

Generally, non-aggressive electrolytes produce thicker barrier layers, large pores,


large interpore distance than aggressive electrolyte [2,10]. Sulphuric acid is more
aggressive than phosphoric acid and hence the pores grow much faster in sulphuric acid
than in phosphoric acid at a given voltage [10].

11
Nielsch et al [35] reported that, the self-organized arrangement and shape of
alumina pores are determined by the anodizing parameter and requires a porosity of 10%
independent of the anodizing parameters. It means that the best self-ordering arrangement
of the pores for optimum anodisation parametric conditions always results in a porosity of
10% as shown in table 1.

Table 1.1: Results of structural property of self-ordered Porous alumina [35]

Electrolyte Interpore Inner wall Pore Porosity


distance thickness diameter

0.3 M H2SO4 66.3 nm 7.2 nm 24 nm 12%

25V,

0.3 M C2H2O4 105 nm 9.1 nm 31 nm 8%

40V,

0.1 M H3PO4 501 nm 54 nm 158.4 nm 9%

195 V

1.3.3 Anodisation Temperature

The electrolyte temperature affect the pore structure [19,35,40,41–45]. The


anodising ratio (the ratio of the thickness of barrier layer to the anodic voltage) of porous
alumina formed in 0.4 M phosphoric acid decreases from 1.14 to 1.04 when the electrolyte
temperature varies from 20 0C to 30 0C at the constant voltage [10]. With increase in the
temperature, both the dissolution of the oxide layer and growth rate increases [42].
Ebihara et al [45] showed that there was no effect of the electrolyte temperature on
diameter of pores in porous alumina formed by anodisation in oxalic acid and sulphuric
acid.

12
Figure 1.9: (a) Effect of anodizing temperature at various anodizing voltage on (a) pore
diameter (b) interpore distance [41].

1.4 Two-step Anodisation


To obtain a highly ordered porous alumina on an aluminium surface by anodisation
Masuda et al used two step anodisation process [11,46,47]. They used ordered pattern
produced by first anodisation as nucleus for a second anodisation step. This eliminated the
random nucleation process and enables ordered structure from the top of the surface. A
schematic representation of the stages used by Masuda et al [46] is shown in figure 1.10. It
is a three step process; during first step of anodisation, the non-uniform oxide layer with
the growth nucleated randomly is obtained. The degree of pore configuration on the
surface is low. Anodizing parameters such as pH of the electrolyte, anodizing voltage, and
temperature of the solution also effect pore distribution. But, after some period of time,
first step of anodisation process, the alumina growth consist of an order structure at the
bottom of the pores.

13
Figure 1.10: (a) Schematic representation of evolution of porous alumina formed by two step
anodisation method [11].

Therefore, in the second process etching is done to dissolve this oxide layer and
obtain uniform and ordered structure. This oxide porous structure formed during first
anodisation step should be protectively etched such that the surface of porous structure
after etching drops an ordered hexagonal dimple on the surface. These dimples will act as
nuclei sites for the second step of anodisation. In the final step, second anodisation is
carried out, where an array of dimples are leftover from etching. In this process of
anodisation, ordered arrays of uniformly arranged pores are generated throughout the
surface [3].

1.5 Ordered Porous Structure as a Template


To fabricate the nanostructure material for the emerging applications such as nano-
devices, nano-electronics, nano-optics, sensing there is requirement of a nano-patterning

14
techniques. Nano structure materials are conventionally produced by various physical and
chemical methods such as chemical vapour deposition (CVD), thermal evaporation,
conventional metal organic chemical vapour deposition (MOCVD), sol-gel approach,
lithography and template-based electrodeposition method [48–53]. In order to control and
improve the nanostructure devices it is necessary to fabricate order nanostructure with
tuneable shape, size and geometry. Lithography is the most common method for
controlling the size and geometry. But it has inherent limitations, such as area of the nano-
patterning, high efforts, high cost. To overcome these restrictions; researchers need to
explore new approaches like porous alumina nano-pattern template-assisted
electrodeposition.

Electro-deposition technique for nano growth fabrication is field assisted and


happens selectively upon exposing to conductive part of cathode that executes it possible
to avoid pore closure and filling efficiency. The difficulty of achieving a homogeneous filling
in porous alumina is non-conducting barrier layer at the bottom. In general, there are two
type of techniques being used so far for filling the metals into nano-pores of alumina
template by electrodeposition.

One type of the method is to remove the ultrathin porous alumina membrane from
the aluminium substrate and then chemically dissolving the barrier layer from the porous
alumina nano structure. The barrier layer was removed after detached aluminium
substrate from porous alumina. As a final step before deposition, a metallic layer (mostly
gold) is sputtered on one side of the free standing alumina membrane. The drawback of
this technique is that the ultrathin porous alumina template is difficult to handle without
an external support of the substrate and hence it is applicable only to thicker
templates (>20 µm). This method mainly employs DC or pulsed DC potential for the
deposition process [54].

The other type of method is reducing or removal of barrier layer through different
processes without removing the porous alumina membrane from the aluminium substrate.
The barrier oxide layer removal processes such as chemical etching, cathodic polarization
and limiting current step wise or reducing anodizing voltage [55–57]. Chemical etching

15
process is always accompanied by pore widening. The bottom part of porous membrane
was chemically etched. The drawback of this process is that the removal of the barrier layer
may cause large pores with ultra-thin pore wall thickness.

Figure 1.11: Schematic representing of the electrochemical process during cathodic


polarization [58].

In cathodic polarization, the polarity of the applied voltage between the sample and
the platinum electrode is reversed, making the anodised sample negative. H+ ions from the
electrolyte pass rapidly through the barrier layer and get reduced at sample-oxide interface
[58]. This results increase in the concentration of OH- ions at the bottom of pores. Before
these ions get transferred to anode, they react chemically with the Al2O3 barrier layer, due
to the following reaction:

( 11 )

( 12 )

Thus alumina layer gets dissolved with the reaction produce Al(OH)3 removed to
the electrolyte solution. This results increase in the pore diameter. Lee et al [59] have
observed increase in pore diameter and corrosion pits on the Al substrate.

16
Nielsch et al [14] reported and described the pore filling mechanism by pulse
electrodeposition. They have suggested that when metal is deposited on the pore tips, there
is depletion in concentration of metal ions in its vicinity. Along each pore the concentration
builds up, and it takes time for the metal ions to migrate from the pore opening to pore tip
by diffusion. Consequently, it is better to use pulsed mode deposition with a relative delay
time between sequential deposition square pulses, sufficient for the metal ion
concentration to provoke ions near the pore tips before the next deposition pulse begins
[14,60]. There is a drawback on using continuous pulses source, that as explained
previously, there is an exhaustion of metal ions near the pore tip, hydrogen evolution will
become dominant inhibiting homogeneous deposition [14]. This will reduce both pore
filling and current efficiency. By limiting current density process or reducing voltage, the
pores at the bottom split up and create dendrite like structure. This dendrite like structure
from the pore tips will be the nucleation site for the filling metal into porous alumina by
electrodeposition. AC pulse electrodeposition has also been carried out by several
investigators after thinning barrier layer [61–64].

1.6 Characterization
1.6.1 Morphology

Transmission electron microscope is generally used to analyse micro-structure


through high-resolution imaging and it has the ability to evaluate the internal structure of
material. Characterization of porous alumina by using TEM was carried out in 1953 by
Keller et al. [9] for the first time. They observed the structured of porous alumina and
illustrated the structure of porous alumina as a hexagonal closed packed structure
consisting pores and barrier layer. In order to characterize the surface, the porous alumina
thickness should be maintained under less than 100 nm since electrons need to go through
the samples. O’ sullivan and Wood [10] have been studied the morphology and growth of
porous alumina in various electrolyte at constant current density or voltage quantitatively
revealing porous alumina morphology using TEM [10]. They found that the most of the

17
walls of the pores are amorphous and at intersection of the walls they found thin
crystalline alumina.

Thompson et al [23] using combination of transmission electron microscopy along


with secondary ion mass spectrometry studied film morphology and composition in great
detail. They have obtained TEM micrographs of porous alumina samples anodised for
various time durations, from 20s- 120s, in phosphoric acid. Using these along with the in-
depth profiles of species related to phosphate anion, obtained by mass spectrometry, they
have suggested a model for initiation and growth of porous alumina by ionic migration
through the oxide film. Similar studies on morphology of porous alumina was conducted by
Le Coz et al [65], where they have obtained high resolution TEM images of highly ordered
hexagonal porous alumina structure and mapped various regions based on the crystallinity
and distribution of chemical species from the electrolyte solution.

Figure 1.12: SEM micrographs of hole-arrays of anodic alumina (a) surface and (b) cross-
sectional views [66].
18
Scanning Electron Microscope (SEM) is generally used to deter mine the surface
morphology, topography and the chemical composition of the specimen. In SEM a focused
electron beam is scanned over a surface and the interaction of this electron beam is used to
form high resolution image (figure1.12). The resolution of the image depends on many
factors. Field emission SEM has been used studying porous alumina morphology. SEM is
not only used to observe the surface and the cross-section of the porous alumina while
tilting [11,66], but also to evaluate the elemental chemical composition using energy
dispersive X-ray analysis [67].

The main disadvantage of SEM imaging is that the sample must have good electrical
conductivity. Since porous alumina is non-conducting material, it needs to be coated with a
very thin layer of gold. Good electrical conductivity can be achieved by coating gold at a
very small sputtering current and appropriate sputtering time with a direct current
sputtering equipment. The time of sputtering should not be too long and the current should
not be too high. From middle of the last century, the characterization of surface
morphology and analysis of pore distributions structure of ordered porous alumina such as
pore diameter, wall thickness, interpore distance, porosity and thickness of the film has
been widely obtained by SEM [3].

1.6.2 Electrochemical Impedance Spectroscopy

Electrochemical impedance spectroscopy (EIS) is an efficient technique for the


characterization of electrochemical systems. The fundamental approach of the AC
impedance spectroscopy is to apply a very small amplitude sinusoidal current and measure
the voltage response at a wide range of frequencies. In recent years, impedance
spectroscopy has found widespread applications in the field of characterization of
materials. An appropriate equivalent circuit (EC) model can fit to the impedance spectra
which allows to obtain the time dependence of important properties of the porous alumina

19
such as the impedance, the capacitance and the resistance of the barrier and the porous
layers [68,69]. EIS can deter mine the changes of the barrier layer and porous layers as a
function of time by analysing the EC.

Sulka et al [70] studied the electrochemical behaviour of porous alumina obtained


by two step anodisation method at the different anodizing voltage and different thickness
(figure1.15). They used electro polished aluminium as reference for impedance
measurement of porous alumina and also studied the response of electrochemical
behaviour of sample before and after thinning of the barrier layer [70].

Figure 1.13: Nyquist plot for a two-steps anodized aluminium produced at 15 and 23 V of (a)
at various thickness (b) before and after barrier layer thinning [70].

Hitzig et al [71] proposed an equivalent circuit model to characterize the porous


alumina. In this model they have separately considered the thick porous alumina layer and
barrier layer and explained the electrochemical properties. Zhao et al [58] studied the
electrochemical properties of porous anodic films on aluminium in different solutions
using EIS. They have simulated experimental results using equivalent electrical circuits and
studied the capacitive behaviour of the porous alumina layer and barrier layer separately
[58].

20
1.6.3 Mechanical Properties

Hardness of the material is often defined as resistance of the material to penetration


depth [72,73]. A hard indenter of know geometry is pressed on to the sample and the
resultant permanent impression is used to define hardness. Though there are many
hardness testing methods and each one use slightly different principle to obtain
quantitative information, all can be reduced to one identical value if Meyer’s hardness of
mean pressure is used [73–76]. It is obtained as the ratio of the applied load to the project
area of the impression on the sample surface. Nanoindentation is a technique of hardness
measurement in which the penetration of the indenter is of the order tens and hundreds of
nanometres. Since the resultant impression cannot be seen optically, it measures the
penetration depth of the indenter and hardness is obtained from a calibrated area function
that relates the penetration depth to the area of the impression. Nano indentation is used
mainly to deter mine the hardness and the elastic modulus. Nanoidentation, along with
proper models, has been used to measure elastic, creep, cracking, phase transformation,
strain hardening behaviour of the material.

Since the penetration depths are less it was realized that the load and depth sensing
nanoindentation could be useful for measuring mechanical properties of a very thin film
and surface layer[74–76]. Oliver, Hatching and Pethica suggested a method to deter mine
the hardness by using load-displacement (P-h) curve and the cross-sectional area of the
indenter as a function of distance from the tip. In this method at peak load, depth can be
obtained from the P-h data and projected area of the contact can be taken as a shape
function [77]. A typical load-displacement curve obtained from nanoindentation is shown
in figure 1.14, the various key parameters are the maximum load (Pmax), maximum
displacement (hmax) and the initial unloading contact stiffness (S=dP/dh).

21
Figure 1.14: Schematic representing of load versus displacement [77].

As the indenter is pressed against the material, both elastic and plastic deformation
occur and leave an impression that conforms the shape of indenter used while only elastic
deformation is recovered during unloading. An estimate of the contact area of indenter
with known geometry, at maximum load Pmax is obtained from P-h curve. Hardness can be
calculated as

( 13 )

Although porous alumina is very attractive for fabricating Nanomaterials, very few
studies have been carried out for evaluating mechanical properties. Alchala et al [78] have
measured the hardness and elastic modulus of porous alumina by using Nanoindentation.
They found that hardness and modulus decreased linearly as the pore size is increased
[78,79]. Xia et al developed a FEM based model for the indentation behaviour of porous
alumina and found that the porous structure collapsed in a shear band deformation mode
than crack formation. They suggested that the material can have multi-axial damage
tolerance where the pores greatly improve the toughness of materials [80,81].

22
Figure 1.15: SEM image of a Berkovich indent revealing two cracks at maximum load 120 mN
[82].

Ng et al performed the nanoindentation test on porous alumina. They have observed


discontinuous with periodic displacement extrusion in P-h. SEM image (figure 1.15) of the
indent showed median cracks and bilinear cracks [82,83]. Propagation of the cracks
involved in the deformation of porous structure is shown to be reason for discontinuous
behaviour in P-h curve. They have developed a phenomenological model for porous
alumina [82]. Bobji et al [84] studied the mechanical properties of porous alumina while
controlling the area fraction of pores. They have found that the hardness is proportional to
the solid area fraction [84].

Tribological properties such as friction and wear of surfaces, depends on surface


roughness and hardness. By controlling the hardness of coating, the tribological properties
can be modified. Hard coating on a softer substrate is desirable to prevent the wear and to
protect the surface [85,86]. When two surfaces come close to each other and interact
mechanically and is contact established over a wide range of scales ranging including
atomic scales. One possible way to control the tribological properties is by modifying the
geometry of these contacts by controlling texture or roughness of the surface.

Texture on the surfaces has been produced by different methods such as creating
nanometre sized holes, by anodizing the surface or by monolayer coating of nanoparticles.
Hamilton et al [87] studied the surface texturing in the form of micro asperities that act as
23
micro hydrodynamic bearing [87,88]. Achanta et al [89] performed reciprocating sliding
test on hard TiN and DLC coating against SiN ball using micro tribometer at normal forces
in millinewton (mN). After just a very few reciprocating cycles, the wear track were
observed [89]. H Yu et al [90] have studied the triblogical performance on textured surface.
The results indicate that there is a better friction reduction effect on textured surfaces
compared with the un-textured ones [90]. Wang et al performed the tribological properties
on thick anodic oxide film and coated impregnated with PTFE using pin on plate [91,92]. It
is found that coated PTFE is having less coefficient of friction and less wear rate. Recently
Ngan et al [93]performed the scratch test on ordered porous alumina and disordered
porous alumina at the same normal load (figure1.16). They found that the coefficient of
friction is higher in disordered porous alumina. For ordered porous alumina they
conducted the scratch test in both dry and wet condition and found coefficient of friction is
more in dry condition.

Figure 1.16: SEM images of the scratch tracks under the load of 5 mN normal load on (a)
ordered porous alumina (b) disordered porous alumina, (c) in air and (d) in water [93].

Ning-ning et al [94] studied tribological properties of nanoporous alumina by using


pin on disc tester and found high hardness and good wear resistance. They have also
studied C60 nanoparticle embedded into the porous oxide film and found that there was a

24
decrease in friction and reduced wear rate when C60 nanoparticle were embedded into
porous alumina. The test results concluded that porous alumina could be used as a
reservoir for C60 nanoparticle [94].

1.7 Scope of Work


Wear resistance of a substance is proportional to hardness [95]. Thus to protect a
soft material from wear a hard coating is being used. From Case carburising to diamond
like carbon films, this principle has been applied in many engineering applications. Hard
coatings fail mainly due to brittle cracks and delamination when subjected to frictional
forces. The hard particles generated by these get entrapped in the contact and increases the
wear rate by acting as an abrasive. Hard coatings are good to protect the surface of the
material only if the coatings do not fail by cracking and delamination.

Composite materials are made up of two materials one material distributed in the
matrix of the other. The resulting material has properties that are superior to the both in
certain aspects. Nanocomposites have been tried by dispersing nanometre sized materials.
The main problem in such nanocomposites is agglomeration of the small particles into big
lumps thus offsetting the advantages of small size.

In this work, we distribute metallic nanorods in a matrix of alumina by filling the


ordered pores resulting from anodisation through electrodeposition. We exploit the
excellent adhesion alumina with the aluminium substrate to form a hard composite
coating. The barrier layer forming the mechanical interface between porous alumina and
the aluminium substrate is optimally modified such that electrodeposition can be carried
out without compromising the strength.

1.8 Aim and Objectives


The overall aim of the present work is “To form ordered composite coatings based
on nanoporous alumina for tribological applications”

25
The objectives of present work are

 To optimize the anodisation conditions to obtain ordered porous anodic alumina


coating in pure aluminium
 To modify aluminium – porous alumina interface for metal deposition
 To deposit metals electrochemically, in the pores such that the nanorods are formed
in all the pores.
 To characterize the mechanical properties of the resultant composite coating.

1.9 Organization of Thesis


In the chapter 2, we study the optimization of porous alumina during anodisation on
pure aluminium. It is found that the geometry of porous alumina can be controlled by
varying the anodizing parameters. We have studied the effect of various electrolysis
parameters like concentration of the electrolyte, anodizing voltage, temperature and time
of anodisation on the surface. We have found that the pore diameter and interpore distance
can be controlled by varying the anodizing voltage. We have also found that the thickness
of the porous alumina film is highly dependent on anodizing time. Further, by controlling
the current density in potentiostatic mode, the pore channel can be tuned.

Chapter 3 deals with the interface modification process used to reduce the barrier
layer thickness, as needed for metal electrodeposition. Various processes that can be used
for this are discussed along with their advantages and disadvantages. In particular voltage
reduction method that can be used to reduce the barrier layer thickness without
weakening the interface is discussed. The dendritic structures generated, by the voltage
reduction technique, at the interface are described

Chapter 4 deals with the electrodeposition of metal into porous alumina structure.
Three methods of electrodeposition that we have explored for filling of the metal in to the
pores are discussed. In particular pulse electrodeposition process that we found to be
effective for uniform deposition is discussed. Optimization of dendrites structures,

26
generated due to the interface modification process described in Chapter 3, to obtain
uniform metal filling is also discussed.

Chapter 5 deals with the indentation behaviour of porous alumina by tuning the
thickness of the porous alumina. The film thickness of the porous alumina was accurately
tuned by controlling the time of anodisation. It has been observed that there is a direct
influence of the film thickness on deformation behaviour. Two different deformation
mechanisms have been observed. For high film thickness and low penetration depths, the
porous alumina deforms by compaction. On the other had for thin films and penetration
depths of the order of the thickness and above, circumferential cracks have been observed
on the coating. Hardness is also found to be dependent on the deformation mechanism.
When the NPA film deforms due to cracks measured hardness is less and when the porous
film deforms by compaction the hardness is high. Hardness of the Cu filled porous alumina
films is found to be higher.

Chapter 6 discusses the evaluation of the friction and wear characteristics of porous
alumina and porous alumina/copper composite coating using reciprocating friction test.
Magnitudes of coefficient of friction for porous alumina and porous alumina/copper
composites are presented. Analysis of wear track generated using scanning electron
microscopy is discussed.

27
Chapter 2

2. Formation of Ordered Pores

2.1 Introduction
Alumina with ordered pores can be formed through electrochemical oxidation of
pure aluminium in an acidic electrolyte for a narrow band of anodising parameters. The
rates of alumina formation at the metal – oxide interface has to be finely balanced with the
rate of dissolution of oxide in the acidic electrolyte. At steady state, either potentiostatic
mode or galvanostatic mode, the porous alumina grows at a constant rate with the
thickness of the bottom barrier layer remaining constant [25]. Pores are nucleated on the
surface at random sites on the surface and get ordered with anodising time. Therefore, a
conventional one-step anodisation process will result in porous alumina with disordered
pores on the top and regular pores at the bottom of the porous alumina film. As we have
discussed in chapter 1, a two-step anodisation process that was proposed by Masuda et al
[11] that provides an ordered nuclei for pore formation is used in this study.

The pore formation during anodisation process depends on many parameters such
as anodizing voltage, pH of the electrolyte and temperature of the solution. The geometrical
parameters of porous alumina such as pore diameter, interpore distance, wall thickness
and the barrier film thickness can be controlled by controlling the anodisation parameters.

In this chapter, we optimize the anodisation parameters to obtain desired geometry


of porous alumina on pure aluminium substrate. Though many studies on this have been

28
reported in the literature, we found that forming repeatable pore structure of desired
dimension still needs careful optimisation of parameters in our laboratory. In this chapter,
we detail the experimental conditions under which the ordered porous structures have
been obtained.

2.2 Electro Polishing of Aluminium


Pure aluminium sheet (99.999%, thickness 2 mm and 0.2 mm) was obtained from
M/s Advent Research Materials. The sheet was cut into required dimension (10 mm x 10
mm) by wire cut electrical discharge machining (EDM). EDM is used to minimise the effect
of work hardening due to sample preparation. After EDM cut, the samples were cleaned in
a series of steps. The samples are first rinsed with de-ionised water and then cleaned with
ethanol for 10 min in an ultrasonicator. After rinsing again with the de-ionized water the
samples were dried with nitrogen jet.

Mechanical polishing was performed using silicon carbide grit emery paper with
size varying sequentially from 400 to 2000. The samples were polished against emery
paper in one direction and then turned 90o before polishing again. The samples were
polished till all the groves left over from previous step are removed. After cleaning with DI
water, the samples were polished with diamond paste on selvet cloth in a polishing
machine. The grit size of diamond paste used was 0.5 µm and the rotation of plate was kept
constant at 800 rpm. The polishing was performed till mirror finish is obtained. After the
polishing, samples were cleaned in DI water before electro-polishing.

Electropolishing is a standard pre-treatment process used to remove roughness and


make the surface homogeneous before generating nanostructures. In this work, all the
aluminium samples were electropolished before the anodisation. Electropolishing was
performed in an electrochemical cell with a solution of 20 wt% perchloric acid in ethanol at
a constant voltage of 10 V. Graphite was used as counter electrode and aluminium sample
as working electrode. Distance between both the electrodes was maintained at around
20 mm for all the experiments. Electro polishing of each sample was carried out for 3 min.

29
Power was supplied through the voltage/current source meter (Keithley 2611B) and the
current was simultaneously recorded during the electro polishing. The temperature of the
electrolyte was maintained at 10 °C by a Peltier device attached with a magnetic stirrer.
After electro polishing samples were cleaned with DI water and purged using a nitrogen jet.

Figure 2.1: I-V curve of aluminium sample in perchloric acid solution.

Figure 2.2: Current vs time during electro polishing at constant voltage of 10 V.


30
To determine the ideal electropolishing voltage we measured the current for
different applied voltages. From the figure 2.1 it can be seen that the current is more of less
constant at about 0.15 A for voltages between 6 V and 18 V. After 20 V the current
increases with the applied voltage and increases drastically beyond 40 V. It has been shown
in the literature [96] that the electropolishing conditions are observed when the I-V
current characteristics are in the plateau region. For our experiments we chose a voltage of
10 V that lies in plateau region figure 2.1.

Variation of current during electropolishing at 10 V is shown in figure 2.2. Current


density as soon as the sample is inserted into the solution is around 0.210 A/cm2. This
drops over time to about 0.07 A/cm2 by 1 min and then a steady state current density of
about 0.065 A/cm2. We find the best surfaces are obtained for 2-3 minutes of
electropolishing. Longer period of electropolishing results in waviness of the surface.

Figure 2.3: AFM images of (a) mechanically polished surface (b) electropolished surface.

Electropolishing produced reflective and smooth surface [97]. Average surface


roughness of these samples was measured using Bruker AFM (Atomic Force Microscopy).
Figure 2.3 shows AFM images of the aluminium samples after mechanical polishing and
after electropolishing. Mechanically polished surfaces revealed linear grooves in the
direction of polishing with the width ranging from 500 nm to 2 micron. Mean roughness
(Ra) measured for 10 µm sampling length was around 14 nm. Electropolishing of the

31
samples for 3 min has removed all the grooves and surface looks flat. The mean roughness
(Ra) for 10 µm sampling length was found to be less than 1 nm. Due to the uniform
polishing effect, the number of in-homogeneities on the electropolished surface may have
been greatly reduced compared to mechanically polished surface [98].

2.3 First Anodisation

Figure 2.4: Schematic of the experimental setup.

After electro polishing, the samples were anodised in an electrolyte cell, with
platinum as counter electrode (figure 2.4) in potentiostatic mode. Two types of sample
holders were used. One holder is made of polycarbonate substrate of thickness 2 mm with
a conductive track for connecting the sample with electrode the power source. The
conductive track was screen printed with silver ink on by using semi-automatic screen
printing machine. The ink was cured at 80 °C for 20 minutes. A dielectric ink is printed on
top of the conducting silver layer (figure 2.5) to protect silver from external electrolyte.
Other sample holder was made of a copper plate with the aluminium sample was using
conducting silver paint and wrapped with teflon tape in such a way that only the front
portion of the sample is exposed to the solution.
32
Anodisation is performed at a constant voltage of 40 V in 0.3 M oxalic acid solution.
The temperature was maintained at 18 °C with the help of a Peltier device (figure 2.4).
Electrolyte was continuously stirred during the process to effectively remove hydrogen
bubbles that evolved on the platinum electrode and to minimise local overheating of the
sample. The rotation of the magnetic stirrer is kept constant at about 200 rpm. Current was
recorded on using Keithly equipment.

Figure 2.5: Schematic diagram of screen printed sample holder.

Figure 2.6 shows a typical variation of current during first anodisation. Initially, the
current decreases rapidly with the time. This is associated with a growth of dense oxide
layer [19,23]. As the oxide layer grows in thickness, at a constant applied potential, the
electric field across the oxide film starts decreasing. After reaching a minimum of about
4 mA/cm2, the current starts rising. At this point the current is low and hence the
dissolution rate is higher than the oxide formation rate. The resultant local dissolution on
the oxide layer forms the nucleus for pore growth. The current reached a maximum of
about 5.5 mA. Once all the pores are formed, the current stabilises with contribution only
from pore growth. The steady state current was found to depend on the anodizing
conditions such as voltage, concentration of electrolyte and temperature during
anodisation.
33
Figure 2.6: Current vs. anodizing time at a constant voltage of 40 V in 0.3 M oxalic acid
electrolyte at 18 °C.

Figure 2.7: SEM top view of porous anodic alumina after first anodisation in 0.3 M oxalic
solution at 18 °C after 30 min of anodisation.

34
The morphology of samples was characterized using Field Emission Scanning
Electron Microscopy (FESEM). The SEM image of the top surface porous alumina after the
first anodisation is as shown in figure 2.7. It can be observed that the pores are randomly
distributed over the surface. It can be seen that the diameter and pitch of pores have huge
variation. The randomly distributed pores get ordered after a certain time of anodisation
below the surface. Figure 2.8 shows the cross-section image of five min anodisation surface.
It can be seen the thickness of about 300 nm from top is not uniform, after that pores starts
getting ordered. Therefore, the pores on the surface are not found to be uniform but after a
certain thickness pores are ordered.

Figure 2.8: SEM cross-section image after 5 min first anodisation.

2.4 Chemical Etching for Patterning


The oxide layer formed after first anodisation has to be removed by chemical
etching to leave patterns on the surface that will act as the nuclei for next anodisation step.
Porous alumina has been selectively dissolved without affecting the aluminium substrate
using two methods. In the first method, a mixture of phosphoric and chromic acid etchant

35
was used. The dissolution rate is high in this method. Hence, it was used to completely etch
the porous alumina structure. In the second method, anodised sample containing the
porous alumina structure was treated with the oxalic acid. The rate of dissolution in this
process is slow compare with previous method. Hence, this procedure was used to increase
the pore diameter.

Figure 2.9: SEM images of alumina after chemical etching for (a) 5 min, (b) 15 min, (c) 30 min
and (d) 1hr.

For complete dissolution of porous alumina a mixture of chromic acid and


phosphoric acid (6 wt% H3PO4 + 1.8 wt% H2Cr2O4) was used. The temperature of the
solvent was maintained at 50 °C. The samples were placed vertically in the solution in such
a way that the porous alumina was facing the solvent. The alumina dissolves isotropically
with time. The morphology of dissolved alumina layer after exposing it to etchant for
various time periods was characterized in SEM to obtain optimum time of etching.

Figure 2.9 shows the SEM images of chemically dissolved surface after various time
periods in chromic acid and phosphoric acid. Figure 2.9 (a) is the SEM image of the surface,
when the oxide layer has been etched for five minutes. It can be observed that the pore
36
have become widened and start merging with neighbouring pores. Some pores dissolve
faster which might be because of the inhomogeneous and amorphous nature of oxide layer.
It was observed that pits are formed on the surface due to highly concentrated strong acid.
Figure 2.9 (b) shows the SEM image of top view of the surface after 15 min of anodisation.
It can be seen that the amorphous oxide layer has dissolved leaving behind the forming
hair like crystalline alumina [10]. At the bottom, it can also been seen that some of these
structures have started to detach from the surface. Figure 2.9 (c) shows the surface after 30
minutes of dissolution. It can be seen that almost all the alumina has been etched. The dark
contrast is the aluminium while the alumina appears bright. It was observed that the
optimal dissolution time is about 30 minutes to leave desired pattern on the surface if the
first anodisation was carried out for 30 min. Longer anodisation will result in thicker
alumina layer and hence would need more etching time.

It has been further observed in SEM etching for 30 min removes the oxide layer
almost uniformly over the whole surface. The resulting uniform pattern consists of closely
packed dimples. The space between dimples is decorated with tiny left-overs from the
crystalline alumina rods forming hexagonal vertices. Figure 2.9(d) shows the SEM image of
surface after 1 hr of chemical etching time. It can be seen that the oxide layer is completely
removed from the aluminium surface and the pattern is destroyed. There are tiny
randomly distributed pits resulting from etching aluminium. For most of our studies, first
anodisation for 30 min followed by etching for 30 min was used.

2.5 Second Anodisation


Second anodisation was carried out after the chemical etching that left the surface
with an array of dimples. It was carried out in the same electrolytic cell like the first
anodisation and under identical conditions. Dimples act the nucleus for the growth of
alumina and hence ordered array of pores was generated throughout the surface.

The anodisation current recorded during second anodisation at 40 V in 0.3 M oxalic


acid solution maintained at 18 °C shown in figure 2.10. Initially the current decreases

37
quickly with the time and reached a minimum value of current around 2.7 mA. After
reaching a maximum of about 3 mA the current slightly decreases and reached a steady –
state value of about 2.6 mA. Current during second anodisation follows the same phases as
the first anodisation (figure 2.10). However the magnitudes of the extremes are reduced.
Also the steady state current is reached early as the pore merging is absent.

Figure 2.10: Variation of current during second anodisation at 40 V.

Figure 2.11 shows the top view of porous alumina surface after second anodisation
for 10 minutes. It can be observed that the porous alumina structure is regular with
hexagonal arrangement. Sui et al [98] has reported that the regular array of 100% defect
free hexagonal arrangement can be obtained in oxalic acid solution over the surface area of
0.25 µm2. The regularity of the pores arrangement occurs over the large surface area. Using
image processing techniques, the diameter (Dp) and the interpore distance (Dint) were
obtained. The average pore diameter was found to be 56 nm and the pitch was around
114±5 nm.

The thickness of the porous alumina layer formed ranged from around 0.3 μm to
200 μm [99] depending upon the anodisation conditions and as well the rate of growth and
dissolution in an acidic electrolyte [100–102]. The film thickness primarily depends on

38
anodisation time [37,103]. In order to measure the thickness of porous alumina film, the
samples were cut using a diamond saw cutter. Figure 2.12 shows the cross section of the
alumina layer obtained after anodisation for different duration of 2 min, 5 min 10 min and
1 hr. The corresponding position, for first three duration, in the anodisation current
variation graph (figure 2.10) is marked as A, B and C respectively.

Figure 2.11: SEM image of the surface after second anodisation.

At the stage A in (figure 2.10), when the time of anodisation was 1 min, the
thickness of the film was measured to be around 0.14-0.17μm. The growth of 2 min
anodisation film was measured around 0. 31 (figure 2.12a). It can be seen that the bottom
of pores has U shape like structure. At the stage B, the anodisation time of 5 min, the
thickness of the film was around 0.55 µm (figure 2.12b). It should be noted that the
magnification of these images are not constant. At the stage C, for the anodisation time of
10 min(figure 2.12c), the thickness was measured to be around 1.1 µm and it was around
6.5 µm for 1hr time of anodisation (figure 2.12 d).

39
Figure 2.12: Cross-section of porous alumina after different duration of second anodisation
(a) 2 min, (b) 5 min, (c) 10 min and (d) 1 hr.

Figure 2.13: Film thickness vs anodisation time curves of porous.

40
Variation of the film thickness with anodisation duration is shown in figure 2.13. It
can be seen that the thickness is linearly varying with the anodisation time with a slope
about 110 nm/min. It is interesting to note that at zero anodisation time, the best fit line
gives a thickness of about 100 nm. This could be the thickness of the barrier layer left over
from the chemical etching process.

2.6 Pore-widening
As briefly mentioned in section 2.4, mild etching conditions can increase the pore
diameter without affecting the other parameters. This pore-widening process was
conducted in 0.63 M oxalic acid solution at 25 °C. The rate of pore-widening is dependent
on the pH value of the electrolyte. When the pH of the solution is low then the pore size
increases faster in comparison with high pH. Therefore, 0.63 M oxalic acid is used for small
pore size and 5 wt% H3PO4 was used for large pore size.

The pore widening is a process after the formation of porous alumina. In this
process the isotropic dissolution of porous oxide layer occurs in an acidic electrolyte. Many
studies have been performed for optimizing the affecting parameters during the pore-
widening. As we have discussed in chapter 1, during pore widening process the wall
thickness decreases and the pore diameter increases. In literature, pore widening is widely
examined. Pore widening in phosphoric acid was observed to result in inhomogeneous
dissolution of porous alumina [104]. We have used two different solutions oxalic acid and
phosphoric acid during pore-widening of porous alumina.

The effect of pore-widening of porous alumina using oxalic acid and phosphoric acid
is as shown in figure 2.14. The dissolution was carried out at the same temperature of 25 °C
and the amount of stirring is also kept constant. When the pore widening is performed for
2hr at 25 °C in an oxalic acid solution (figure 2.14 a), the pore diameter is measured to be
around 57 nm. The diameter has been increased to around 20 nm in pore widening. The
rate of increase after the chemical dissolution is slow (0.17 nm/min) and it can be observed
that, after the pore widening, the pores are regularly arranged. The dissolution of oxide

41
layer is uniform over the surface. In contrast, the pore widening in phosphoric acid solution
is performed for 1 hr (figure 2.14 b). The pore diameter is widened to around 40 nm and
the dissolution rate of 0.67 nm/ min. It is noticed in the pore widening with the phosphoric
acid, that some pores have started merging with the neighbouring pores. Therefore, the
phosphoric acid etching rate is faster than oxalic acid. It is illustrated that the pore
diameter can be controlled by chemical dissolution using various solutions. It is found that
the rate is slow and more isotropic in oxalic acid solution compared with phosphoric acid
solution. Hence oxalic acid dissolution was chosen as it gives better control over the whole
process.

Figure 2.14: SEM image of pore widening in (a) 0.63 M oxalic acid solution for 2hr and (b) 5
wt% phosphoric acid solutions for 1hr.

One of the main reasons for carrying out pore widening is to control the dendrite
layer formed in the barrier layer [105]. Controlling this dendrite layer is very important (as
can be seen later in chapter 3) for uniform filling of metallic nanowires in the pores.

2.7 Choosing the Anodising Parameters


The geometry of porous alumina depends on the anodisation parameters as
discussed in chapter 1. Some of the parameters that affected the final choice of parameters
were studied in detail and reported below. Our aim was to make porous alumina that can
be filled with metals for use in tribological applications. Hence the main concern was the
42
ease with which we can control the geometry without compromising on the total time
needed for making repeatable samples. Some of the choices were driven by the limitation
on the experimental facilities like power supplies; temperature controllers while others
have been driven by the initial hypothesis that for tribological applications we may need
thin alumina films rather than thick ones. Here we discuss some of the studies that helped
in deciding the final anodising parameters.

2.7.1 Anodisation Voltage

Pore diameter of the anodised alumina depends on the equilibrium achieved


between the rates of dissolution and formation of the oxide. Increasing anodising voltage
can increase the formation rate while the dissolution rate is kept constant more or less by
maintaining the pH of the electrolyte. Probably, the actual process is more complex and
depends on the interplay between the various anodising parameters.

Figure 2.15 shows the SEM images of porous alumina samples obtained for different
anodizing voltages in oxalic acid following. The rest of the anodising conditions are kept
constant as detailed before for achieving ordered pores following two-step anodisation
process. The surface morphology of anodised sample at different voltage contains more or
less circular pores, uniformly distributed over the surface. It can be observed that the
diameter of pores is increasing as the anodizing voltage increase. When anodised at the
voltage of 10 V, it is seen that the pores are formed (figure 2.15a). It noticed that some
pores are not appropriately opened up but all the pores are regularly arranged. The size of
pore diameter measured from image analysis is around ~13 nm. Figure 2.15 (b) shows the
SEM image of porous alumina when the anodizing voltage was 30 V. The average diameter
of porous alumina was measured to be around 39 nm. In the case of applied anodizing
voltage of 50 V, the pore average diameter was around 63 nm. When the anodised voltage
at 60 V is as shown in figure 2.15 (d), the size of pores becomes larger, to around 79 nm.

43
Figure 2.15: SEM images of perfectly ordered porous alumina formed at different anodisation
potential of (a) 10 V, (b) 30 V, (c) 50 V and (d) 60 V.

Figure 2.16: Variation of pore diameter for different anodizing voltages.

The pore diameter was measured from the SEM images using an image processing
code written in Matlab. First the image converted to the gray scale and then a gray scale

44
threshold value was chosen to distinguish the pores. The threshold value is selected by
taking the average in between the two peaks of the intensity distribution along a line in the
image. Using the threshold value the image was converted to binary image. The size and the
centroid of all the pores in the image were then obtained. The pore diameter measured
over an area of about 1.2 µm x 0.9 µm. The variation of the pore diameter as a function of
the anodising voltage is given in figure 2.16. The pore diameter can be seen increasing
linearly with the anodizing voltage. The experimental results are in a close agreement with
the theoretical prediction calculated from equation 9.

2.7.2 Type of Electrolyte

The types of anodisation forming electrolyte is influenced both the formation and
the resultant geometrical parameters. Typical formation rate of porous alumina increase in
order to: sulphuric acid > oxalic acid > phosphoric acid [23]. The typical formation ratio of
porous alumina is around 2.7 and 1.2 nm/V. Results show that in sulphuric acid, the pore
diameter is smaller than the oxalic acid and phosphoric acid.

Three different electrolytes were tried out sulphuric acid, oxalic acid and
phosphoric acid. Ordered pores had been reported in all these electrolytes [3]. SEM image
of porous alumina obtained by the two step anodisation in the solution of sulphuric acid,
oxalic acid and phosphoric acid under the potential of 25 V, 40 V and 80 V respectively
shown in figure 2.17. The time of anodisation was 10 min for all the cases. In
figure 2.17 (a), it can be seen that the pores are visible and ordered in the solution of
0.3 M sulphuric acid under the anodizing voltage of 25 V at 18 °C. Some pores look larger
which might be because of rise in temperature during anodisation. The average pore
diameter was around 48 nm and the interpore distance was measured to be 60 nm. The
diameter was measured around 56 nm and pitch was around 114 nm when porous alumina
is formed in an electrolyte of 0.3 M oxalic acid at anodizing voltage of 40 V at 18 °C
(figure 2.17b). In the case of phosphoric acid (0.4 M H3PO4) solution, the applied anodizing
voltage was 80 V at the temperature of 18 °C. The average pore diameter was measured to
be 70 nm and the pitch was around 120 nm.

45
Figure 2.17: SEM images of porous alumina formed in different electrolytes of (a) 0.3 M H2S04
at 25 V, (b) 0.3 M H2C2O4 at 40 V and (c) 0.4 M H3PO4 at 80 V.

The final choice of electrolyte depended on the ability to control the temperature of
the electrolyte with the Peltier device of capacity 62 W. For aggressive sulphuric acid, the
drawback in our experimental setup was maintaining the temperature constant during
anodisation. After a certain period, the electrolyte temperature starts increasing as the
amount of heat generated during anodisation is higher than that can be removed by the
Peltier device. Similar temperature rise was also observed in phosphoric acid at the

46
anodizing voltage of 80 V. Since there is an increase in temperature, the rate of dissolution
increases resulting in bigger sized pores compared to the empirical prediction (like in
figure 2.16). Since there is a temperature increase in both the electrolytes, sulphuric acid
and phosphoric acid, they are not suitable for an anodisation temperature of 18 °C, which
can be achieved in using oxalic acid. Further it was found that for phosphoric acid, the
current density was higher and often the limitation was the maximum current limit of the
DC power supply used.

To summarize, all the three acids provide good arrangement of circular and
uniformly arranged pores. After considering all the above facts, 0.3 M oxalic acid at 40 V
seems to be the most appropriate electrolyte since there is no notable increase in
temperature during the process.

2.7.3 Effect of Temperature Control

The temperature plays an important role in the formation of porous alumina during
anodisation. Temperature should be lower than the room temperature to avoid local
heating at the bottom of the pores during anodisation. Without controlling the
temperature, the growth of alumina is not uniform [41]. Local heating caused by
uncontrolled temperature could result in an inhomogeneous electric field distribution.
Moreover, without controlling the temperature, the porous film could generate cracks.
Here, we have performed the effect of porous film in uncontrolled temperature while
keeping all anodisation conditions constant.

Figure 2.18 shows the SEM image of porous alumina obtained using 0.3 M oxalic
acid at an anodizing voltage of 40 V. The time of anodisation was 10 min. The anodisation
was carried out at room temperature. From the figure 2.18 (a) and figure 2.18(b), it is
observed that the surface has cracks. These cracks are generated because of increase in the
temperature of electrolyte during anodisation process started at room temperature. The
temperature of the electrolyte kept on increasing during the process. When the electrolyte
temperature increased, it was noticed that the fluctuation in the flowing current was very
high as shown in figure 2.19. It can be seen that current was very high within a minute and
47
was increasing up to two minutes. The fluctuation in the current might be because of
inhomogeneous electric field distribution occurring at the bottom of the pores due to local
heating. Hence without constant temperature the porous oxide layer was found to have
cracks.

Figure 2.18: (a) SEM image of porous alumina at 0.3 M oxalic acid solution at room
temperature (b) Enlarged SEM image at different location of the same sample.

48
Figure 2.19: Current density –time graph of porous alumina formed in oxalic acid solution
without controlling the temperature.

49
2.8 Summary
We describe the synthesis of porous anodic alumina by two step anodisation
method, an ordered hexagonal pore arrays formed by anodisation during growth with
oxalic acid as an electrolyte. We have systematically optimized the geometrical parameters
and obtained the precise control over the geometry of porous alumina.

 We have found out that the pore diameter and interpore distance is highly
dependent on the anodizing voltage. The pore diameter is linearly proportional to
the anodizing voltage.
 Among the three electrolyte solutions used, we found that the oxalic acid has better
control compared to sulphuric acid and phosphoric acid.
 We have studied the effect of temperature on the porous film while keeping all the
other anodisation conditions constant. It was found that the uncontrolled
temperature generate cracks on the oxide layer.
 For precise control of the film thickness, we have varied the anodizing time from 1
min to 60 min and found that the thickness of the film increases linearly as the time
of anodisation increases.

50
Chapter 3

3. Controlling Barrier Oxide Layer

3.1 Introduction
Porous alumina has been widely used as a template to make uniform long aspect
ratio of nanorods/wires [12–18]. Most often the alumina is removed from the aluminium
substrate and used as a free standing membrane in which metal is deposited
electrochemically. Our aim is to use the porous alumina as a matrix in which ordered and
aligned nanorods are embedded to create a composite coating. We would like to exploit the
good adhesive property of the alumina with the substrate along with the high hardness of
alumina.

Figure 3.1: Cross section image of typical porous alumina.


51
In order to fill the pores with metal the good conducting aluminium substrate has to
be exposed to the electrolyte through the non-conducting alumina layer. Figure 3.1 shows a
typical cross section of the coating we have obtained. It can be seen that the bottom of the
pores are closed with a spherical cap of alumina barrier layer. This insulated barrier layer
has to be removed for in situ synthesis of nanowires by electrodeposition. The thickness of
the barrier layer is proportional to the applied anodizing voltage and it varies from few nm
to a few hundreds of nm. It should be noted that the alumina layer shown in figure 3.1 is
subjected to 3 hr of pore widening and hence the barrier layer has already been thinned to
some extent. The thickness of the barrier layer was around 52 nm before pore widening
and around 30 nm after pore widening.

Barrier oxide layer can be removed by chemical etching, cathodic polarization and
reducing anodizing voltage in steps [56,58,104,106,107]. Chemical etching is isotropic and
as seen in the previous chapter results in complete removal of the alumina hence it is not
suitable for composite coatings. In cathodic polarisation, the sample is converted to
cathode after anodisation and as the current is passed, oxide layer is dissolved selectively
along the conducting path and hence pores are selectively etched. Stepwise voltage
reduction process uses the fact that the pore diameter and the thickness of the barrier layer
formed depends on the anodising voltage applied. As the voltage is reduced in steps the
pore diameter and the thickness reduces forming a dendrite like structure [108–110].

Dendrites like structures (figure 3.2) are thought to be formed under non-steady
state conditions of anodisation [110]. These dendrite like nanoporous structures can be
used as a new type of template towards the fabrication of nanomaterial to improve the
shape dependent optical, magnetic, electrical and mechanical properties [111–114].

52
Figure 3.2: Fabrication of dendrite like structure [110].

Figure 3.3: (A) Potential-time and (B) Current-time in non-steady state anodisation [110]

Cheng et al [110] reported that the dendrite like structures obtained under the non-
steady state anodisation condition is by decreasing the potential exponentially, which is
53
shown in figure 3.3. The dendrite like structure might be formed because of the electrical
treeing and the mechanical stress during the growth of structure [34,115]. The negative O2-
and OH- ions will build up in the oxide layer due to the unbalanced formation rate of oxide
and dissolution rate of oxide. At some point of time, partial discharge takes place leading to
non-uniform current pathway during the growth. This electric field-induced dissolution
results in the generation of branched like structure. Mechanical stress may also play an
important role in the formation of branch pore morphology. When the new pore tries to
develop from its parent pores, it has to first overcome repulsion forces from its
neighbouring pores. If it succeeds, the parent pore will branch into two other pores,
otherwise it will end up as a node [110].

In this work we have studied and optimized appropriate conditions to reduce the
barrier layer thickness. We have created varying thickness of dendrite structure by
controlling the rate of reduction of the anodising voltage. We have measured thickness of
the barrier layer obtained during this process by monitoring the capacitance in-situ using
LCR meter.

3.2 Experimental Details


At the end of the second anodisation process, we tried to reduce the barrier layer
thickness by all the three methods, by chemical etching, cathodic polarization and reducing
anodizing voltage in steps. In chemical etching, barrier layer is thinned by exposing it to an
acid. After the second anodisation sample is taken out from the electrolyte and then the
chemical etching is carried out in a solution of 0.63 M oxalic acid at a temperature of 25 0C.

In cathodic polarization [58] method the polarity of the sample was reversed, to
dissolve the barrier layer. The applied DC anodisation voltage was reduced from 40 to 15 V,
and then is maintained at 15 V for 15 min before reversing the polarity. This is done to
form the dendritic layer at the bottom of the pores. Cathodic polarization is then carried
out in 0.5 M KCl solution at 14 0C. A constant reverse potential of -5 V is applied for 20 min.

54
After that pore widening is carried out in a solution of 6 wt% of H3PO4 solution for 1hr at a
temperature of 25 0C.

Third method of barrier thinning was performed by reducing voltage gradually in


steps In this voltage reduction method, after the second anodisation, the anodisation
voltage is dropped from 40 to 0 V in steps of 1V in 0.3 M oxalic acid. The reduction voltage
rates of 1V/4s, 1V/8s, 1V/10s, 1V/12s, 1V/15s, 1V/20s, 1V/25s, 1V/30s, 1V/40s and
1V/100s are considered at 18 0C. 1V/12s indicates that after reducing the current
anodising voltage by 1V voltage was maintained constant for 12 sec before the next step of
voltage reduction. After voltage reduction, the pore widening is performed in a solution of
0.63 M oxalic acid solution. The temperature for pore-widening was maintained at 25 0C.

During the barrier thinning process, capacitance was measured in situ in the same
electrolyte. For this a LCR meter (E4980 from M/s Agilent Technologies) with a wide
capacitance measurement range with accuracy of ±0.05% was used. The operating
frequency was maintained at 100 Hz. The LCR meter and Source meter (Source Measure
Unit 2611B from Keithly) were connected to the computer through the GPIB (General
Purpose Interface Bus). The connection to the sample of the source meter or the LCR meter
was controlled through a relay as shown in figure 3.4.

Figure 3.4: Schematic of in situ capacitance measuring setup

To characterise the interface between aluminium and the alumina, AC impedance


analysis was performed by using the same LCR meter with the two electrode system
control. The absolute impedance and phase angle were recorded over a frequency range of

55
20 Hz to 2 MHz on a logarithmic scale. A voltage of 100 mV was applied with the zero offset
potential. Impedance data was recorded through GPIB using labview.

3.3 Different Methods of Barrier Thinning


The compact dielectric layer formed at the pore bottom is dependent on anodizing
voltage with the proportionality constant of about 1.3-1.4 nm/V [3]. For filling metals by
electrodeposition, barrier layer should be less than 10 nm [14]. Thus the barrier layer has
to be thinned for electrodeposition. The three methods that has been tried are Chemical
etching, cathodic polarisation and stepped voltage reduction.

3.3.1 Chemical Etching

Chemical etching is the process where the dissolution of porous oxide layer occurs
in a suitable electrolyte acid. Reducing barrier layer by chemical etching, so far many
studies have been performed to fill metal into the pores[3]. During etching process the
barrier layer thickness reduces. We have used oxalic acid solution for chemical etching as
we have optimized the process and found that the homogeneous dissolution takes place in
oxalic acid (section 2.6).

Porous alumina is obtained by using 0.3 M oxalic acid at constant voltage of 40 V for
10 min of anodisation at 18 0C. SEM cross-section images of porous alumina before
thinning barrier layer shows the barrier layer thickness was around 50 nm (figure 3.5a &
3.5b). The variation in the barrier thickness was negligible. Afterwards, the chemical
etching performed in a solution of 0.63 M oxalic acid for 3 hr is shown in figure 3.5(c) & (d).
It was observed that the thickness of barrier layer is reduced and the thickness has been
reduced from about 50 nm to about 25 nm. After chemical etching, the diameter of porous
alumina increases from 56 nm to 80 nm and wall thickness reduces from 77 nm to 53 nm. If
the barrier layer is reduced further, the wall thickness decreases and pores will start
merging with each other. The hardness of the coating will decrease as it depends on the
solid area fraction of the alumina [84]. Therefore, chemical etching or pore-widening can
56
be used to reduce the barrier layer thickness to some extent (~ 30-35 nm) but not
completely. Since the metal filling process takes place only if the barrier layer thickness is
below 10 nm, if it is more than 10 nm, the efficiency of the deposition will be low and a
higher deposition voltage is required.

Figure 3.5: (a) SEM cross-sectional view of porous alumina before etching and (c) after pore-
widening, insets (b) and (d) show the enlarged image of barrier layer.

57
3.3.2 Cathodic Polarization

Figure 3.6: Voltage-Current –time plots before cathodic polarization.

Before cathodic polarization, the anodizing voltage was reduced step wise from 40 V
to 15 V with a time gap of 9 sec between two successive reduction steps in the same
electrolyte cell as second anodisation. The graph (figure 3.6) shows the change in current
as a function of time before the process of cathodic polarization. From the graph, it is
observed that after 10 min of second anodisation, steady state current density was close to
5mA/cm2, and as the voltage reduction process occurs, the current flow decreases
gradually. Then as the samples are held for 15 min at 15 V, the current reaches to a value
of about 1.8 mA.

58
Figure 3.7: Variation in current during cathodic polarization.

The cathodic polarization was performed at -5 V at an electrolyte temperature 14 0C.


The graph in figure 3.7 shows the change in current as a function of time during cathodic
polarization. In cathodic polarization, the current flows due to reduction of hydrogen ions
into hydrogen gas [59]. Initially, the current remains constant, as the ions are transported
the pores. After 12 min, the current starts to increase rapidly which indicates that more
numbers of OH- ions are penetrating into the surface. Moreover, these ions can corrode the
interface of porous alumina. Figure 3.8 shows the corresponding SEM image of porous
alumina after cathodic polarization. From the SEM images it is seen that film is completely
removed at the interface and lots of pits are observed at the bottom of pores.

59
Figure 3.8: (a) SEM cross-sectional view of porous alumina after cathodic polarization, (b)
and (c) show the enlarged image of (a).

3.3.3 Step-wise Voltage Reduction Method

Both the chemical etching and cathodic polarization method have some problem. In
the ordinary chemical etching process, because of etching the barrier layer thinning rate
and the wall thickness etching rate are found to be the same. In the cathodic polarization,
due to the chemical attack of OH- ions, the barrier etching rate was much faster than the
wall thickness etching rate. However, there was no control over the temperature and it is
observed that film had damaged at the interface as discussed in section 3.3.2 (figure 3.8).

60
Figure 3.9: Variation of current during step-wise reduction process.

In the step-wise voltage reduction process, after second anodisation the voltage was
reduced from 40 V to 0 V in steps of 1 V at a various reduction rate. Figure 3.9 shows the
applied voltage and recorded current as a function of time during this process. From the
figure 3.9, it can be seen that at each step as the voltage is reduced, the anodizing current
decreases immediately and then increases more or less exponentially. It takes finite time
for the current to reach a steady state value. Next reduction step typically follows before a
steady state current is reached. The presence of exponential current variation indicates
some sort of capacitive behaviour and the interface has in fact been represented by an
equivalent electrical circuit consisting of resistors and capacitors.

The morphology of porous alumina which was obtained after voltage reduction with
the rate of 1V/10s is shown in figure 3.10. It is seen that, at the bottom, U shape barrier
layer is thinned without affecting the porous alumina wall thickness. Therefore, by using
voltage reduction method bottom oxide layer can be removed with minimal damage to the
porous structure. The thickness of dendrite like structure depends upon the reduction rate
of anodizing voltage.

61
Figure 3.10: SEM cross-section image of (a) porous alumina after voltage reduction with the
reduction rate of 1V/10s, (b) after pore widening for 2 hr (c) enlarged image of (b).

3.4 Controlling Dendrite Geometry


One of the important requirements for the composite coating is that all the pores
have to be filled with the metals. Once electrodeposition starts in one of the pores it
becomes the path of least resistance and hence the nanowire will grow preferentially in
that pore. If all the pores have to be filled they have to be filled simultaneously. This would
mean that all the pore bottoms should have identical resistance to flow of ions with
assistance from applied potential or by diffusion. This translates into having identical
electrical and physical structure at the bottom of the pores. Hence, it is important to control
the dendrite geometry such that uniform filling is made possible.

62
3.4.1 Influence of Voltage Reduction Rate on Dendrite Structure

Under steady state anodisation condition, the barrier layer thickness is constant and
depends on the equilibrium between the rate of formation of oxide and the rate of
dissolution of oxide. When the anodisation voltage is reduced both the formation and
dissolution rates will decrease and a new equilibrium will be attained. The barrier layer
thickness will reduced in tune with the reduced voltage. However the presence of pore
walls will prevent uniform pore formation. If the current is allowed to reach a steady state
then pores of reduced diameter will grow. At the bottom of porous alumina, the dendrite
structure geometry depends upon the reduction rate of the anodizing voltage.

Figure 3.11 shows how the current varies during stepwise reduction of voltage as
the anodisation voltage is decreased from 40 V to 0 V at different voltage reduction rates.
The general trend is similar to the one described above (figure 3.9). Overall the current
decreases exponentially. It should be noted that all the ions and electrons contributing to
the current would involve in formation of oxide. Hence the current only gives a broad
notion of oxide layer thickness. Whenever the voltage is suddenly dropped the
environment at the bottom of the pores varies and finite time is required to attain the new
equilibrium state. The time constant for this exponential process depends on the
capacitance of the existing oxide layer and the double layer in the solution along with the
resistance to the current flow probably determined by oxide layer. We find that this time
constant is of the order of 15 sec and increases as the voltage is reduced.

As the anodising voltage is reduced, the sharp decrease in current may depend on
the magnitude of this change followed by an exponential return to a new steady state value.
Probably most of the current during this initial stage is contributed from the readjustment
of the ion environment in the solution as well as at the metal-oxide interface. Hence very
little oxide layer is formed. If the voltage is further reduced before considerable oxide is
formed then the dendrite structure is not formed or not very prominent.

63
Figure 3.11: Variation of current during the stepwise voltage reduction for reduction rate
of (a) 1V/4s, (b) 1V/8s, (c)1V/10s, (d) 1V/12s, (e) 1V/15s, (f) 1V/20s, (g) 1V/30s and (h)
1V/40s.
64
Cross section images of the oxide coating with dendrite formed by various reduction
rates are shown in figure 3.12. In order to make the dendrite layer visible we had to do
pore widening for 2 hr before taking the cross section. The pore widening as can be seen
later helped us to ensure uniformity in metal growth during deposition. Cross section was
obtained by mechanically breaking the oxide layer and hence the surface is not uniform.
Sectioning of porous alumina using Focused ion beam (FIB) milling was tried, but it was
unsuccessful because of charge accumulation of the alumina as the layer is non-conducting.

It can be seen from the figure 3.12 that no dendrite structure can be observed for
rates of 1V/4s and 1V/8s. The thickness of the dendrite layer increases with the reduction
in rate. Dendrite structures formed for the rates around 1V/15s were with minimal
branching and for 1V/40s produced very complicated with multitude of branching. As the
dendrite branches grow and the pores start merging at the bottom and hence it starts
appearing as if two different layers of structure exist (figure 3.12h).

Figure 3.13 shows the total film thickness and dendrite layer thickness variation at
different voltage reduction rates. The voltage reduction rate of 1V/4s is marked as 0.25 V/s
in the figure. When the voltage reduction rate is slow, the dendrite film thickness is more
and the dendrite structure formation depends on the total time of oxide formation. The
time of second anodisation was constant for each of the samples, so the thickness of the
pores above the dendrite structure will be almost constant. Hence, the overall total
thickness (after the second anodisation film thickness plus dendrite thickness) increases
with the dendrite thickness. For the anodisation voltage rate of 1V/4s, the dendrite
thickness was around 0.032 µm. For anodisation voltage reduction rates of 1V/8s, 1V/10s,
1V/12s, 1V/15s, 1V/20s, 1V/25s, 1V/30s and 1V/40s, the dendrite thickness was found to
be 0.18 µm, 0.30 µm, 0.31 µm, 0.52 µm, 0.79 µm, 1.00 µm, 1.05 µm respectively. From figure
3.13, it is clear that the dendrite thickness is inversely proportional to the voltage
reduction rate. As a result dendrite structure can be tuned by controlling the anodisation
voltage reduction rate.

65
Figure 3.12: SEM cross-section image of porous alumina with the various reduction rates of
(a) 1V/4s, (b) 1V/8s, (c) 1V/10s, (d) 1V/12s, (e) 1V/15s, (f) 1V/20s, (g) 1V/30s and (h)
1V/40s.

66
Figure 3.13: Dendrite layer thickness resulting from various voltage reduction rates

If the current transient following the sudden drop in voltage is allowed to reach
equilibrium then the pore structure is expected to stabilise at a new pore diameter.
However the new pore is allowed to nucleate only at the bottom of the existing pores. To
check this out, after second anodisation the voltage was reduced in steps to 25 V, 12.5 V,
6.25 V and 2 minutes of time was allowed for the current to equilibrate. The time elapsed
between successive step is about 8 times the typical time constant. Due to capacitive nature
at interface, a step change in voltage produces an exponential change in the current with a
time constant. The final voltage of 6.5 V was held constant for 10 min. The cross section of
the resulting pore structure is shown in channel can be seen in the figure 3.14. It can be
seen that for the initial voltage drop from 40 V to 25 V only one pore has formed inside the
existing pore. The resulting inter pore distance is dictated by the pores formed at 40 V but
the pore diameter is being decided by the reduced voltage of 25 V. However for the next
drop in voltages the pores starts merging and finally moving towards the new equilibrium
structure dictated by the anodising voltage. Hence at the bottom many branches of
dendrite like structure were found at the bottom.

67
Figure 3. 14: (a) SEM cross-section image of porous alumina formed in 0.3 M oxalic acid
solution at 18 0C by tuning the voltage (b) Enlarged image of bottom part

Figure 3.15 compares the structure formed by two extreme rates of voltage
reduction of 1V/10s and 1V/100s. From the figure 3.15 it can be observed that the dendrite
thickness is very less when the anodizing voltage reduction rate was 1V/10s (figure 3.15a
& 3.15b). Some pores are open at the bottom. When the reduction rate was 1V/100s, the
dendrite thickness kept on increasing due to long anodisation voltage reduction time. It is
noted that in each step of voltage reduction, the parent pore has two branches and these
branches are generating two more branches in the second voltage reduction. This process
is repeated until the end of anodisation. Therefore, the dendrite thickness keeps on
increasing with time at the slow reduction rate. The dendrite like structure will be useful
for structure dependent properties.

68
Figure 3.15 (a) and (c) SEM cross-sectional images of porous alumina with the voltage
reduction rate of 1V/10s and 1V/100s (b) and (d) are the enlarged image of bottom of figure
(a) & (c) respectively.

3.4.2 In situ Capacitance Measurements

Capacitance is inversely proportional to the thickness of the dielectric medium


separating two conductors. The barrier oxide layer is the dielectric separating aluminium
and the conducting solution. By measuring the capacitance, it was hoped that a qualitative
measure of variation in barrier layer structure can be obtained. The motivation is to get a
quick and non-destructive measure for the barrier layer structure without having to take a
cross section and use FE-SEM to obtain the structure. Hence we built a relay in the circuit
to alternatively connect the power source and a LCR meter during anodising (figure 3.3)

69
Figure 3.16: Variation of Capacitance with time at various voltage reduction rates

Figure 3.16 shows how the capacitance varied with time for different voltage
reduction rates. After second anodisation at 40V, the capacitance was found to be around
0.27 µF. This capacitance value gradually increases and reached a plateau towards the end
for almost all the rates except the fastest one of 1V/4s. The same data is plotted as a
function of anodising voltage in figure 3.17. It is seen that, when the voltage reduction rate
is high (1V/8s), the capacitance increases to around 0.52 µF. On the other hand, when the
voltage reduction rate is low (1V/40s), the final capacitance was found to be around 2.3 µF.
It can also be seen that the capacitance plateaus at around 8-10 V.

70
Figure 3.17: Variation in Capacitance as a function of applied anodising voltage

From the capacitance point of view, the dielectric oxide layer thickness varies over
the whole surface of aluminium electrode. Since the potential between the far away
solution and the aluminium is constant, the interface can be thought of as a number of
capacitors connected in parallel. The effective capacitance is sum of all the individual
capacitance and will be dominated by the capacitance of the thinnest section as the
capacitance is inversely proportional to the thickness. Thus the final capacitance indicates
the thickness of the barrier layer that the electrodeposition has to overcome. Low rate of
voltage reduction gives the thinner barrier layer and hence higher capacitance.

3.4.3 Effect of Pore Widening

Pore diameter increases during the pore widening process. The pore-widening on
dendrite structure was performed after voltage reduction process, to smoothen fine
dendritic structures. Figure 3.18 compares cross section of the oxide layer subjected to
different periods of chemical etching of 30 min and 2 hr at 25°C. The voltage reduction rate
was 1V/15s after second anodisation. It can be seen that the diameter after 2 hr of pore
71
widening increases, compared to that of 30 min of pore widening. From the figure 3.18, it is
noticed that the chemical dissolution of the film was uniform and isotropic. Therefore at
the bottom of pore channel, dendritic structure is etched to ensure uniform structure for
filling of metal.

Figure 3.18: SEM cross-section image of pore widening for (a) 30 min and (b) 2hr

Figure 3.19 shows the capacitance variation versus anodising voltage reduction rate
from 1V/4s to 1V/40s before and after pore widening for 2 hr. It is seen that the
capacitance increases after pore widening. Since, after pore widening, the bottom thickness
of porous alumina reduces due to etching, thus the capacitance increases. The percentage
of increase in capacitance due to pore widening appears to be constant.

The capacitance measurement for different time of pore widening after the barrier
layer thinning at a rate of 1V/15s is shown in figure 3.20. In the curve, it is observed that
capacitance increases as the duration of pore-widening increases. When the pore-widening
time was 30 min, the capacitance was around 0.71 µF. Further, when the pore was widened
for 60 min, 90 min, 120 min, 150 min and 180 min, the capacitance were 0.743 µF, 0.792
µF, 0.842 µF, 0.914 µF and 1.02 µF respectively. It is evident that the capacitance increases
more or less quadratically with pore-widening time. In addition, capacitance is
proportional to the area and inversely proportional to the thickness. So, in case of pore
widening, the bottom thickness of oxide layer decreases and its area may increase as well
72
with increasing pore widening duration. Since the capacitance is proportional to the total
area of the pores and inversely proportional to the oxide layer thickness thereby the
capacitance increases proportionally as pore widening time increases.

Figure 3.19: Capacitance vs voltage reduction rate after the barrier thinning and the pore
widening

Figure 3.20: Capacitance at different pore widening time of porous alumina after voltage
reduction from 40V to 0V at the rate of 1V/15s

73
Figure 3.21 shows the comparison chart of the capacitance variation during the
overall process, starting from porous alumina to dendrite structure formation. Usually, the
oxide layer forms when the pure aluminium is exposed to air(less than 5 nm) [116] and
capacitance of this very thin oxide layer will be high. Hence, in the figure 3.21, after both
electro polishing and etching of oxide layer on pure aluminium, the capacitance is found to
be very high. The capacitance after second anodisation is very less (around 250 nF).
Capacitance is inversely proportional to the insulating separation thickness between two
conducting electrode. After second anodisation, the insulating oxide layer is thick
(~50 nm), as a result the capacitance is less. The capacitance during barrier reduction rate
of 1V/15s and pore widening is more compared to second anodisation. In case of barrier
reduction rate of 1V/15s, the capacitance is higher than second anodisation because the
oxide layer is thinned. In pore widening, which is performed for 2 hr after the barrier
reduction, the barrier layer decreases further and hence capacitance increases. Capacitance
increases because of dissolution of oxide, thickness of the barrier oxide layer further
decreases and area increases. Hence, by measuring the capacitance, the dendrite oxide
layer thickness can be controlled.

74
Figure 3.21: Capacitance variation after electro polishing (EP), etching (ET), second
anodisation (SA), barrier reduction with the reduction at 1V/15s (BR) and pore widening
(PW).

3.5 Impedance Spectroscopy of Porous Alumina


In situ capacitance measurements were carried out at a single signal frequency of
100 Hz. Analysis of wide frequency range impedance can be carried out after the process is
completed. This impedance spectroscopic measurement is generally analysed with the aid
of Nyquist plot or Bode plot. Nyquist plot shows the variation of the reactive component of
impedance with respect to the resistive components at various probing frequencies. Bode
plot shows the logarithmic modulus of impedance and the phase angle as a function of
frequency (explained in appendix A).

3.5.1 Equivalent Circuits Model Analysis

Usually, porous electrodes have large surface areas, resulting in high capacitances and
compensating slow electrochemical reactions. A metal electrode with a thin oxide layer in

75
an electrolytic solution has been modelled typically by Randles circuit [117] consisting of
an electrolytic solution resistance (Rs) in series with a parallel combination of a double
layer and oxide layer capacitance (Cdl) and a charge transfer resistance. Introducing a pore
resistance Rp indicating the resistance for the ions to diffuse through the narrow path in
pores and through dendritic structure, each pore can be modelled by an individual Randles
circuit. Since all the pores are connected electrically in parallel, the equivalent circuit may
be simplified to a single Randles circuit as shown in figure 3.22.

The impedance Z, of equivalent circuit in figure 3.22 is given by

( 14 )

where, ω is the measuring frequency. The transient current through this equivalent
circuit can be obtained as follows

[ ]
( 15 )

It can be seen that for short times the current is dominated by the exponential term, and as the
time tends towards a large value the exponential term tends towards zero and hence the current is
given by the term outside the square bracket.

76
Figure 3.22: Electrical equivalent circuit model for porous alumina

3.5.2 Impedance Analysis

The equivalent circuit and simulation results with the circuit show that the electrical
equivalent circuit can be successfully applied to the experimental data to explain the
porous alumina electrochemical behaviour. Usually, at low frequency, the resistance of the
sample dominates. At high frequency, the capacitance of the sample dominates. We have
studied the effect of these characteristics quantitatively by plotting the frequency
dependent impedance response of this circuit by a Nyquist plot.

77
Figure 3.23: Electrochemical Impedance spectroscopy: Nyquist plot for porous alumina; first
anodisation, second anodisation, barrier thinning at 1V/15s and pore-widening

Figure 3.24: Electrochemical Impedance spectroscopy: Nyquist plot for different pore
widening times for a voltage reduction rate of 1V/15s.

78
Figure 3.25: Electrochemical Impedance spectroscopy: Bode-phase plot for porous alumina;
first anodisation (FA), second anodisation (SA), barrier thinning (BR) and pore-widening
(PW).

Figure 3.26: Electrochemical Impedance spectroscopy: Bode-amplitude plot for porous


alumina; first anodisation (FA), second anodisation (SA), barrier thinning (BR) and pore-
widening (PW)

79
Impedance spectra over a frequency range of 2 Hz to 2 MHz after the first
anodisation (FA), second anodisation (SA), barrier thinning (BR) and pore widening (PW)
are shown in figure 3.23. After FA, the thickness of film was around 1.4 µm and after SA, BR
and PW the film thickness was around 1.7 µm. Nyquist plot response shows that, the
resistance (z’) after FA is less than the corresponding value after SA, because the film
thickness is less after FA compared to SA. After barrier layer thinning, the resistance
decreases from 6 KΩ to 3KΩ which reveals that the barrier layer is thinned. After pore
widening, resistance is observed to be even lesser than post barrier layer thinning.
Furthermore, we have studied the Nyquist response after pore widening at various times
(figure 3.24). As the pore-widening time increases, the resistance decreases since the oxide
layer reduces in thickness and area increases. Bode-magnitude plots are presented in
figures 3.25 and 3.26. At low frequency, phase angle is in between 20 degree to 80 degree
which means the process is under diffusion control at that corresponding frequency.

3.6 Summary
To deposit metal into the pores methods for barrier layer thinning were explored in
this chapter. We have studied chemical etching, cathodic polarization and voltage reduction
processes for this purpose. We have found that using step-wise voltage reduction process,
we were able to create a dendritic structure in the barrier layer that can enable the metal
filling into pores without weakening the interface. The thickness of this dendrite like
structure depended on the time over which this modification was carried out. A systematic
study on optimization of this voltage reduction process for reducing barrier layer was
carried out.

80
Chapter 4

4. Electrochemical Deposition of Metal into Porous


Alumina Film

4.1 Introduction
For depositing metal into the pores electrochemically, there has to be a conducting
path through the oxide layer. One way is to remove the oxide layer completely from the
bottom of the pores. This will weaken the interface and will compromise on the main
advantage of anodised films. The other way is to thin the barrier layer sufficient enough for
electrons to conduct through without much compromise on the integrity of the anodised
film. The dendrite like structure formed by the step-wise voltage reduction method comes
handy as thinning happens only over a small area. However, the main challenge would now
be to ensure uniform and simultaneous deposition of the metal in the pores.

Probably the conduction of the electron through the residual thin barrier layer
happens by tunnelling [14,118]. It has been proposed that the nucleation of the metal into a
pore happen at more than one location but under certain conditions of electro-deposition
single crystal metallic wires can be grown into the pores [16]. In this chapter we explain in
detail all out attempts at growing nanowires uniformly in the ordered nanoporous alumina.
We show that even though the process was optimised with respect to copper deposition,
the same methodology can be used to deposit nickel and silver as well.

81
4.2 Experimental Details
Depending on the type of the voltage signal used, electrodeposition of metals into
the pores can be carried out by three different methods.
 D.C deposition
 A.C. deposition
 Pulsed deposition

For D.C. electrodeposition, constant voltage signal was applied between the sample
and a platinum gauge electrode. Anodised aluminium was made as cathode. For A.C.
deposition A.C. signal was generated using a National Instruments data acquisition card
(DAQ-PCI6221). An external voltage amplifier used to amplify the voltage signal supplied
by the data acquisition card. The block diagram of the amplifier along with a current
measurement setup is shown in figure 4.1. For the pulsed deposition also the same circuit
was used. Desired voltage shape and amplitude was defined in a computer through a
Labview program. For A.C deposition a pure sinusoidal signal was generated while pulsed
deposition a positive square pulse followed by a negative pulse and a time gap was used.
The amplifier circuit amplifies the voltage generated by the DAC card by 10 times to the
range 0-50 V.

Figure 4. 1: Block diagram of the A.C deposition circuit.

82
Rf

Vcc

Ri
Q3

U3

R3 Q4
VIN
R4

Sample
Vee
R1
Vsense
R2

Isense_V U1

Risense R5

Figure 4. 2 Circuit diagram for pulsed deposition amplifier.

The diagram of the actual circuit designed and built is given in figure 4.2. The
voltage signal from the DAC card is given to the voltage amplifier as input Vin. This voltage
signal is amplified by op-amp “U3” and applied to the sample. The two transistors Q3 and
Q4 are used as a buffer which helps to deliver current of upto 500 mA to the sample. The 1
 resistor is used in series with the sample for current measurement. Current through the
sample creates a small voltage drop across the 1 Ω resistor. This voltage is amplified by the
opamp “U1” and given to DAQ card. Thus the applied voltage and the current could be
simultaneously are logged into PC.

Copper deposition into the porous alumina was carried out in a mixture of
0.5M copper sulphate (CuSO4) and 0.57 M boric acid (H3BO3) solution. The DC deposition
voltage was kept at 10V. Current data was recorded through Labview by using DAQ card.
The electrodeposition using AC signal was carried out at 17 Vac at the frequencies of
250 Hz, 500 Hz and 750 Hz at 250C. Pulse electrodeposition was carried out by applying a
positive pulse of 7 V and negative pulse of 17V for different pulse time with a delay time of
50 ms. The filled porous alumina samples with copper was examined using SEM to deter

83
mine the degree of filling the sample. The percentage of filling metal was obtained after
removing the few nm of thickness from top surface using mechanical polishing.

Electrodeposition of Ag and Ni into porous alumina was carried out by applying


pulse electrodeposition. Ag deposition was carried out in a solution of silver sulphate
(8.5 g/l) di-ammonium hydrogen citrate (200 g/l) and potassium thio-cyanate (105 g/l).
The pH was maintained between 4-4.5. The applied voltage was + 6V. The duration of pulse
of positive and negative voltage was 6ms and the delay time was 50 ms.

Electrodeposition of Ni was carried out in a mixture of nickel sulphate (300 g/l),


nickel chloride (45 g/l) and boric acid (45 g/l). For Ni deposition, the temperature was
maintained at 35 0C. The pH of the solution was 4.5. The applied voltage for positive pulse
was 3 V for 2 ms and negative pulse was 10V for 8ms. The delay time was 50 ms.

4.3 Electrodeposition of Copper


Metal filling by DC electrodeposition into porous alumina is usually performed by
two processes. One is removing the aluminium completely by etching and deposition of
conducting layer on the reverse side, but it is difficult to handle ultrathin oxide layer as we
have discussed in pervious chapter [106]. Second process involves removing the barrier
layer from the bottom of the pores completely, but it is difficult to obtain uniformity over
the entire surface [56,106]. We follow the second process and as outlined in previous
chapter, we try to get the electrochemical conditions at the bottom of the pores as uniform
as possible. Even after the barrier layer thinning, the main challenge was to achieve
uniform filling of all the pores simultaneously.

4.3.1 DC Deposition

Figure 4.3 shows the top view SEM micrograph of porous alumina after filling
copper by DC electrodeposition method. The surface of the alumina was seen scattered
with tiny pyramids of copper. The size of the pyramids varied from few nanometres to few
micrometres. To deter mine the amount of pore filling of copper into the samples were
84
mechanically polished gently on a selvyt cloth. It can we clearly seen from the SEM image
(figure 4.3b) most of the pore are empty. In DC deposition method it was found that we had
very little control over the process.

Figure 4. 3: DC deposited surface showing a) tiny pyramids and b) pores filled with copper
revealed after polishing.

4.3.2 Voltage Controlled AC Electrodeposition

Without separation or removing barrier layer electrodeposition of metal in porous


alumina is possible under AC electrodeposition. When an AC voltage or current is applied
to the barrier layer, it acts as a rectifier allowing current preferentially in one direction
[61]. This is possibly because of the fact that electrons can move through the barrier layer
with much ease compared to bigger ions like Al3+ or O- and OH- ions. Deposition of metal
like nickel, cobalt, cadmium, bismuth, iron, silver and gold in porous alumina using

85
AC electrodeposition has been reported [119]. However, AC electrodeposition through the
barrier layer is complicated process because the deposition depends on frequency of the
signal applied [18,120]. To optimize the metal deposition and effect of deposition
frequency for AC electrodeposition of copper in porous alumina, a systematically study has
been performed.

Figure 4. 4: Applied voltage during AC electrodeposition.

Figure 4. 5: Current trace during AC electrodeposition.

86
The uniform growth of metal by AC electrodeposition at the voltage controlled mode
with a continuous 250 Hz, 500 Hz and 750 Hz frequency sine wave was used under the
17 V. Figure 4.4 shows the sinusoidal voltage applied and the corresponding current
(figure 4.5) flowing through the electrochemical cell. It can be seen that the current is more
or less symmetric about the x-axis indicating that the total effective charge transferred per
cycle is very small. The effect of AC electrodeposition on porous alumina during copper
filling is characterized using SEM.

Figure 4. 6: SEM images of porous alumina after voltage reduction with the reduction rate of
1V/15s followed by 2 hr pore widening, filled with copper using AC electrodeposition at
various frequencies of (a) 250 Hz, (c) 500 Hz and (e) 750 Hz.

87
Figure 4.6 shows the SEM image of top surface of porous alumina after filling copper
into porous alumina. When a continuous sine wave of 17 V AC at 250 Hz frequency was
applied (figure 4.6 a), oval shape damaged regions were observed on the surface. The
operating frequency of applied voltage was increased to 500 Hz, numerous of pyramids
were observed on the top of surface. The size of pyramids varied from ~200 nm
to ~800 nm over the all surface (figure 4.6b). When the frequency was maintained at
700 Hz, it was seen that bunches of pyramids were accumulated together and forming an
island of pyramids. Surface was found damaged at the edges of these pyramids and theses
damaged were observed oval shape. It might be because of aluminium oxide/ hydroxide
dissolution producing hydrogen gas as we have discussed in chapter 1 [58]. Therefore the
evolution of hydrogen gas is becoming more dominant resulting in damage of the porous
alumina layer.

4.3.3 Voltage Controlled Pulse Electrodeposition

Figure 4. 7: (a) Applied voltage transient of two square pulses during pulse electrodeposition.

88
It is identified in the above section that damage occurs on porous alumina layer
under continuous AC electrodeposition conditions. To avoid this effect, instead of
continuous AC signal, the pulse electrodeposition based on step square wave with relaxing
voltage is used for filling. The square wave signal gives a better pore filling compared to
sine waveform [14,60,120]. The alternating voltage pulse signal in millisecond range with a
delay is used for copper filling into porous alumina.

Figure 4.7 shows the applied voltage pulse with the delay time. The polarization
voltage of 7 V making the sample positive was applied for 3.2 ms followed by a negative
voltage of 17 V applied for 3.2 ms. Switch in voltage values occurred over few
microseconds. Deposition of copper happened during this negative pulse. A delay time of
50 ms was given before the next cycle. This delay helps in restoration of copper ion
concentration at the pore bottom before the next pulse and prevents excessive hydrogen
evolution [60]. This improves the homogeneity of deposition. The delay should not be more
than 2 sec, because the dissolution reaction between the oxide and electrolyte can become
dominant [14,60].

Figure 4. 8: Current trace of two square pulses during pulse electrodeposition.

89
Figure 4.8 shows the current traces measured at the beginning of the deposition. As
the positive pulse is applied, the current increase to 40 mA and as the voltage is held
constant at 7 V for 3.2 ms, the current decreases to a value of about 10 mA. The positive
pulse is applied to discharge the capacitance of the barrier layer. When this positive cycle
switch to the negative cycle, a sudden jump in current (250 to -230 mA) is occurs. During
each pulse of negative current copper is deposited at the pore bottom. At the beginning of
the delay time a sharp increment of current around 70mA occurs for a few micro seconds
and afterwards current decreases immediately and reaches zero. During delay between
pulses as no voltage is applied, the ions move from higher concentration region to lower
concentration region due to diffusion.

Figure 4. 9: SEM images of porous alumina after voltage reduction with the reduction rate of
1V/15s followed by 2hr pore widening filled with Cu using pulse electrodeposition (a) and (b)
are the top view, (c) and (e) are the cross section image of porous alumina after copper
deposition and (d) Top view of deposited copper on porous alumina after polishing.

90
The copper filled into porous alumina samples were examined by SEM as shown in
figure 4.9. From the cross-sectional images (figure 4.9c and 4.9e) it is shown that the
copper is filled into the pores. The top surface is the image after removing the overgrowth
shown in the figure 4.9d, it can be seen that almost the entire surface is filled with copper
nanostructures. Cross section and the surface image after mechanical polishing to remove
overgrowth further confirm this.

4.4 Effect of Dendrite Geometry on Deposition


Figure 4.10 and figure 4.11 show the SEM images of porous alumina after filling
copper into porous alumina samples modified at various voltage rates after second
anodisation. It is observed that when the voltage reduction rate was 1V/4s, almost all pores
are empty on the top surface and the cross-sectional image confirms that pores are not
filled at bottom (figure 4.10 a and b). A thick barrier layer still exists at the pore bottom
which might be preventing the copper deposition. The pores are observed to be empty on
the top surface when the samples were modified with a reduction rate of 1V/8s, but the
bottom of pores are filled with copper, the length of the copper nanowire is found to be less
than 30% of the porous alumina thickness (figure 4.10 c and d). For the samples obtained
with the voltage reduction rate of 1V/10s, few pyramids were detected on the top surface
(figure 4.10 e and f) of porous alumina. As soon as a copper nucleus is formed it provides
the path of least resistance resulting in its rapid growth. Hence the pores wherever the
nucleus is formed gets filled fast. Once the copper nanowires reach the surface it is freed
from the restrictions to ion flow imposed by the insulating alumina. The wire grows
uniformly in all the directions. The single crystal nature of the nanowire and the
preferential dissolution of any secondary nucleus formed during the positive cycle results
in pyramidal structure. Once pyramids start forming on the top surface of the porous
alumina film, the filling of the pores comes to a halt. It has been observed however long the
deposition was continued the pores never filled. The variation in the length of copper
nanowire in pores of this sample was found to be around 80 ± 20% of the porous alumina
thickness.

91
For the 1V/12s modified sample, a bunch of pyramids are observed (figure
4.10 g and h), it can be seen in the cross-section of the image that the variation in the length
of copper wire was around 5% of the porous alumina thickness and almost all the
nanowires reach the top surface of porous alumina. The size of the pyramids were found to
be almost double to that of the pyramids formed at 1V/10s. During deposition, it is
observed that the growth of nanowire is faster at 1V/12s.

When the dendrite length was around 0.5 µm (voltage reduction rate 1V/15s),
numerous pyramids are observed on the surface and it is seen from the cross-sectional
image that every pore is filled with copper till the top surface and found the growth of
filling is extremely uniform in almost every pore channel (figure 4.11a and b). At the
reduction rate of 1V/20s with the dendrite thickness around 0.8 µm, a few pyramids are
observed at the top surface (figure 4.11c), and cross-sectional image showed that the
bottom of pores channel, all the dendrites are filled and highly uniform (figure 4.11d). The
length of copper nanowire is measured to be around 50±10% of the porous alumina
thickness. At the voltage reduction rate of 1V/25s, it is observed from cross-sectional
image that only dendrite thickness is filled (figure 4.11e and f). As the dendrite thickness is
increasing (voltage reduction rate 1V/30s), it is seen that at the bottom of pores, a very few
roots of dendrite is filled (figure 4.10g and 4.10h) .This implies that, when the length of
dendrite is more (1 µm), the dendrite is preventing the copper filling in pores. That might
be because of the duration of negative pulses during deposition are very short.

From these observations, it is clear for the given voltage pulse the best copper filling
is obtained only if the dendrites are formed with voltage reduction range in the range of
1V/10s-1V/20s. Noticeably, for the rate of 1V/15s, a highly uniform nanowire growth is
observed.

92
Figure 4. 10: Top SEM view of porous alumina after copper deposition at the barrier layer
thinning with voltage reduction rate of (a) 1V/4s , (c) 1V/8s , (e) 1V/10s , (g) 1V/12s;
(b),(d),(f) and (h) are the corresponding Cross-sectional image.

93
Figure 4. 11: SEM Top surface image of porous alumina after copper deposition at the barrier
layer thinning with voltage reduction rate of (a) 1V/15s, (c) 1V/20s, (e) 1V/25s, and (g)
1V/30s; (b),(d),(f) and (h) are the corresponding Cross-sectional image.

94
Figure 4. 12: Effect of pore-widening time on deposition.

After fixing the voltage reduction at the rate of 1V/15s, we have studied the
influence of pore widening on growth of copper. Pore widening was carried out for 2 hr
and 3 hr. The duration of deposition was 8 min and was kept constant in both the cases
(figure 4.12). It is found that for 3 hr of the pore widening time, the overgrowth is in
pyramid shape as shown in figure 4.12a and b and when pore widening time is 2 hr, a
cauliflower shape is observed (figure 4.12 c and d). From the current trace during
deposition it is noticed that the overgrowth in 2 hr pore widening started much early
(~ 4 min) than in 3 hr pore widening (~8 min). It is noticed that the size of pyramids were
bigger than the size to cauliflower like structures.

95
Figure 4. 13: Current vs deposition time curves of 1V/15s barrier reduction rate at three
stages Electrodeposition time.

Figure 4.13 shows the current pulses at three different stage of deposition: initial
stage, mid stage, and last stage. The stages are chosen based on the pulse shape. Initially
the magnitude of the current at the start of the pulse is more than that at the end of the
pulse. In the mid phase the initial current and end current of the pulse are almost equal.
The final phase is marked by a very high current and the initial point of current is more
than end point of current.

96
Figure 4. 14: SEM images of copper nanowires at various deposition time of (a)10s, (b)28s, (c)
30s, (d)90s, (e)120s and (f)180s.

To study the copper growth during these three different phases, the porous alumina
was removed by chemical etching in sodium hydroxide solution after deposition. This
exposes the copper nano wires which can then easily imaged. Figure 4.14 shows the copper
nanowires on aluminium surface obtained at various deposition times. Nanowires forms
bundle like structure due to adhesive interaction between them. The bundles must have
been formed by the surface tension as the solution dries. The stability of bundle depends
97
on the attractive adhesive force between the individual wires and the stiffness that oppose
bending. Thus the size of the bundle will depend on the stiffness of the copper wires which
in this case depends on the length of the wires and the adhesive interaction. Since adhesive
interaction remains constant being dependent on the material, small bundles are formed by
low stiff wires.

It can be observed that there is very little variation in the length of the wires
confirming simultaneous deposition. It is seen that the copper nanowire length is
increasing with deposition time. At a less deposition time of 10 sec, the copper wire length
was found to be in 0.23 nm. The copper length increases as the electrodeposition time
increases, (figure 4.14b). After deposition time of 30 sec, the dendrite is filled completely
and the copper nanowire starts growing into the pores towards the surface. When the
deposition time was around 180 sec, it is seen that the copper nanowire have reached the
top surface.

Figure 4. 15: Evolution of the current pulse characterised by the initial current (I1) and final
current (I2).

98
Comparing the measured current pulse line, the ones in figure 4.13 and the SEM
images of figure 4.14, the different phases of the deposition can be easily identified. The
initial phase corresponds to the dendrite filling and the mid phase corresponds to uniform
growth of nanowires into the pores and the final phase marks the projection of nanowire
outside the pores and formation of overgrowths resulting in pyramid or cauliflower like
structures.

Figure 4.15 shows how the magnitude of the current at the start of the negative
voltage pulse and the current at the end of the pulse varies with the deposition time. The
dendrite structure starts filling up first. This phase is characterised by nucleation at
multiple sites within a single pore. The current pulse can be characterised by Randle circuit
as discussed earlier. The initial current (I1) as soon as the voltage reversal occurs is high
and the current exponentially decreases to a steady state value (I2). As the deposition
progresses, while the I1 remains same the I2 starts increasing as it can be seen from figure
4.13. When I1 and I2 are equal, the dendrites seem to be fully filled. Afterwards, the current
pulse more or less remains flat with I2 slightly greater (more negative in figure 4.15) than
I1. As soon as the voltage pulse switch to reverse direction, the current response shows
typical capacitive behaviour with current suddenly changing to I1 followed by potential
decay to I2. This indicate uniform filling of dendritic structures. As the pores get filled up
the behaviour changes with no exponential decay which is shown in figure 4.13. It may be
that once all the dendritic structures filled the current becomes flat. Thus longer it take for
I1 and I2 to be same, more dendritic are getting filled. This results in uniform growth of
nanowires. Finally, when the nanowires start to overgrow from the pores, a sudden
increase in current is observed.

Figure 4.16 shows the time taken for filling of the dendrite structure generated at
various voltage reduction rates. It is found that for the given experimental conditions the
time period where I1 crosses I2 (as indicated by circle in figure 4.15) is greater than 15 sec
then uniform filling occurs. If the time is less than 5s, then the uniform growth of metal is
not achieved. That is uniform filling of the metal in the porous structure is achieved when
the dendrite structures at the interface are generated using the voltage reduction rates in

99
between 1V/12s and 1V/20s (figure 4.16). We have observed almost 100% filling when
dendrite structures are produced at voltage reduction rate of 1V/15s.

As soon as the voltage pulse reverses direction, the current response shows typical
capacitive behaviour with current suddenly changing to a value of I1 followed by potential
decay to a value of I2. This indicate uniform filling of dendritic structures. As the pores get
filled up, this behaviour changes as shown in figure 4.13. It may be that once all the
dendritic structures filled the current pulse becomes flat instead of exponential decay from
I1 to I2. Thus longer it take for I1 and I2 to be same, more dendritic are getting filled. This
results in uniform growth of nanowires.

Figure 4. 16: Voltage reduction rate when the initial point and end point of negative pulse are
equal during deposition

100
Figure 4. 17: SEM cross-section images of copper nanowires with porous alumina after partial
dissolution of porous alumina in 0.1 M NaOH at room temperatures.

Figure 4. 18: Deposition time of 1V/15s barrier reduction rate samples as a function of
nanowire length.

As the bunched up nanowires prevent accurate length measurement, the porous


alumina was partially etched to get better estimate as shown in figure 4.17. It can be seen
that the copper nanowire at the bottom has two roots. The length of roots of the nanowire
was almost same. Figure 4.18 shows the length of the copper wire deposited as a function
of time. It can be seen that the rate of growth remains linear till the wire reaches the
101
surface. The total electrodeposition time to fill the porous alumina with 1.7 µm pore
thickness was around 180 sec.

4.5 Electrodeposition of Silver and Nickel


It is found that deposition of porous alumina has occurred when the optimal range
of voltage reduction of barrier layer is 1V/10s-1V/25s and the best copper filling is
obtained when the voltage reduction rate was 1V/15s. We have studied the
electrodeposition for Ag and Ni under same conditions of modified barrier layer.

Figure 4. 19: Top view SEM images of porous alumina after Ag deposition at the barrier layer
thinning with voltage reduction rate of 1V/15s (b) after removing overgrowth.

The pores are filled with Silver under the pulse electrodeposition based on square
wave with relaxing voltage. The pH of the solution was maintained at 4-4.5. The filling in
porous alumina was examined to deter mine the degree of pore filling which is as shown in
figure 4.19. Top image exhibits some over growth on the surface. However, overgrowth
was removed by mechanical polishing which deter mines the percentage of pore filling the
alumina filled with Ag. The magnified SEM image (figure 4.19b) shown that almost all the
pores is filled with Ag. About 95% of pores were found to be filled.

102
Figure 4. 20: Top view SEM images of porous alumina after Ni deposition at the barrier layer
thinning with voltage reduction rate of 1V/15s, (b) after removing overgrowth.

After Ag filling, we tried to fill Ni into porous alumina under the specific condition of
pulse electrodeposition. The Watts bath contains boric acid which is usually added as a
buffering agent. The pH of the solution was maintained at 4.5.

Figure 4.20 shows an SEM image of porous alumina filled with Ni before and after
polishing. Mechanical polishing ensured that the Ni nanowire is filled into porous alumina.
The filling of Ni is found homogeneous and uniform into the pores. Hence it is possible to
fill metals into pores uniformly by pulse electrodeposition method. The thinning of the
barrier layer with the voltage reduction rate of 1V/15s, is found to be highly efficient with
almost 90% filling of metal into porous alumina.

4.6 Discussion
Porous alumina formed by anodisation of aluminium has a non-conducting barrier
oxide layer. This layer does not allow the metal to be deposited into the pores using
electro-chemical deposition process as a conducting path is required. One of the methods,
[106] that is being used, is removing aluminium substrate completely by etching and
depositing a conducting metal layer like a thin layer of gold on the reverse side of the free

103
standing alumina film. However, it is difficult to handle ultrathin oxide films. Another
method is by thinning the barrier layer at the bottom of the pores. The barrier layer is
thinned to nanometer thickness such that electrons can tunnel through the oxide
layer [14,118]. Uniform thinning of barrier layer will result in reduction in te interface
strength. Hence a gradual reduction in the diameter of the pores resulting in a dentritic
structure at the bottom of the pores is used. However the dendritic roots formed at the
bottom of the pores by this method are non-uniform resulting in non-uniform filling of the
pores [56,106].

Dendritic structure has been obtained by Cheng et al [110], at non-steady state


anodisation conditions by exponentially decreasing the anodisation voltage. In the
experiments presented in this study the thickness of the dendrite like structure was
controlled by thinning the barrier layer at various voltage reduction rates (discussed in
chapter 3). The thickness of barrier layer was determined by measuring capacitance of this
insulating dendritic layer. To remove the intricate paths in the dentritic structure that
could introduce variability in conduction, we carried out a chemical thinning process. This
process will also cause an increase in pore diameter and we refer it as pore widening
process in this chapter. It was found that pore widening after barrier thinning resulted in
dendritic roots that facilitate uniform filling of metal in the pores. This method can
therefore be successfully used to create uniform tribological coatings. By varying the
voltage reduction rate and the subsequent pore widening time, it is possible to obtain an
optimal dendritic layer so that conduction of electrons through the thin barrier layer
occurs by tunnelling without compromising much on the adhesion of the alumina layer.

Once the barrier layer is thinned sufficiently the nanowires can be grown
electrochemically using Direct Current (DC), Alternating Current (AC) or pulsed
electrodeposition. Metal filling by DC electrodeposition [106] into pores of the porous
alumina results in pyramids of copper on the free surface of the alumina. Most of the pores

104
were found empty. It was found that the main variable in the process was the voltage
applied and it has very little effect on the uniformity in filling.

AC electrodeposition without separation or removing barrier layer has also been


used for filling of pores in alumina. Metals like nickel, cobalt, cadmium, bismuth, iron, silver
and gold has been filled into porous alumina [119]. When an AC voltage or current is
applied to sample, the barrier layer acts as a rectifier allowing current preferentially in one
direction [61]. The deposition rate and uniformity depends on frequency of the signal
applied [18,120]. We have studied AC electrodeposition at three operating frequency
250 Hz, 500 Hz and 750 Hz. At 250 Hz frequency, it was seen that porous alumina surface
was damaged in few places (figure 4.6 a). When the frequency was increase (500 Hz and
700 Hz), it was seen that bunches of pyramids were accumulated together and forming an
island of pyramids. Surface was found damaged around the edges of these pyramids. To
avoid this effect, instead of continuous applying the ac signal, a pulse electro deposition
based on step square wave with relaxing voltage can be used for filling. A rectangular pulse
is found to give a better pore filling compared to sine waveform [14,60,120].

Pulsed electrodeposition is reliable for deposition into nanoscale multifunctional


materials and can compensate for the slow diffusion-driven transport of ions in the
pores [14]. A negative pulse is applied to fill the metal and this is followed by a small
positive pulse discharge the charge accumulation in the barrier layer which acts like a
capacitor. A delay, following this positive pulse, helps in restoration of copper ion
concentration at the pore bottom before the next pulse by diffusion. Delay also prevents
excessive hydrogen evolution and thus improves the homogeneity of deposition [60]. A
longer duration of delay would result in the dissolution of the oxide in the
electrolyte [14,60].

The uniformity in filling is also dependent on the pores structure at which the
tunnelling and charge accumulation are taking place. To optimise, we have filled metal into
porous alumina with different dendritic structure. Different structures were obtained by
varying the voltage reduction rate at which the barrier layer was thinned. It is observed the
uniform copper filling is obtained only if the dendrites are formed with voltage reduction

105
range in the range of 1V/10s-1V/20s. We found that, for the rate of 1V/15s, a highly
uniform nanowire growth is observed. Even though the process was optimised with
respect to copper deposition, we have found that the same methodology can be used to
deposit nickel and silver as well. Hence, uniform filling of pores can easily be achieved by
carrying out the barrier layer thinning process followed by the pore winding process.

4.7 Summary
Electrodeposition processes for filling of metal into the porous structure are
explored. DC, AC and pulse electrodeposition were explored and we have found that the
use of pulse electrodeposition is highly efficient and well suited method for metal filling
into the pores. Using this process along with optimization of barrier layer thinning process
we have achieved uniform metal filling into porous alumina. We have identified the metal
filling characteristics of porous alumina for different voltage reduction rates. We have
found that barrier layer thinning using the voltage reduction process at the rate of 1V/15s
was highly efficient for almost 100% filling of metal into pores.

106
Chapter 5

5. Measurement of Hardness by using


Nanoindentation

5.1 Introduction
Wear resistance is proportional to the hardness of the material (Archard law) [121].
Hence it is a well-established practice to coat a soft material with a hard coating to improve
wear resistance [122–124]. However, hard coatings fail due to cracks [122] and delaminate
[125]. The wear particles generated due to delamination can get entrapped between the
contacting surfaces and may act as abrasive, thereby increasing wear rate[126]. Hard
coatings are good to protect the surface of the material only if the coatings do not crack
[122]. In this chapter, we report on our exploration to use porous alumina and porous
alumina nanocomposite as tribological coating. We have studied the mechanical properties
of porous alumina and porous alumina filled with metal by measuring the hardness. We
have also studied the deformation characteristics of porous alumina coating and the effect
of modifying the interface barrier layer on the deformation characteristics.

Indentation response of porous alumina has been studied as a function of film


thickness using nanoindentation. The film thickness of the porous alumina was controlled
by varying the time of anodisation and indentation was carried out on all these films at the
same maximum load. We find that the film thickness has a direct influence on the
deformation behaviour. Two different deformation mechanisms, cracking and compaction
were observed and correspondingly the hardness also varied as function of film thickness.
This observation is confirmed by indentation at different loads such that the penetration to
107
film thickness ratio is kept constant. Indentation on film with modified interface revealed
that the effect of the modification has minimal effect on hardness and no delamination of
alumina film was observed. Copper filled alumina nano composite coatings showed higher
hardness and lesser propensities to crack compared to porous alumina of same thickness.

5.2 Experimental Details


We have carried out nanoindentation studies on three types of samples: Porous
alumina samples, Porous alumina samples with modified interface, and metal filled porous
alumina samples. Porous alumina samples with varying thickness (as shown in Table 5.1)
were produced as described in chapter 2. Interface modified porous alumina samples for
the nanoindentation studies were generated using voltage reduction method as described
below. The interface modification was carried out on the samples after anodisation with
the step-wise voltage reduction method at a reduction rate of 1V/9s from 40 V to 10 V.
After this, the voltage was maintained at 10 V for 15 min. No pore widening was carried
out. The total time for interface modification for all the samples was around 19-20 min.
These experiments were carried out before the copper deposition process was optimized.
Hence the barrier layer modification process was slightly different.

Table 5.1: Parameters for all the experiment

Sample Anodising Film Thickness Modified film Total Film


Time ( min) ( µm) thickness Thickness (
( µm) µm)
Porous alu 1 0.17 - 0.17
mina 2 0.31 - 0.31
3 0.55 - 0.55
5 0.69 - 0.69
10 1.18 - 1.18

Modified 2 0.31 0.18 0.49


Porous 4 0.55 0.30 0.85
alumina 5 0.69 0.23 0.92
10 1.18 0.22 1.4

108
SEM images of cross-section samples used for nanoindentation experiments are
shown in figure 5.1. Anodisation time for the samples shown in figure 5.1a and figure 5.1b
was 5 min and 10 min respectively and with interface modification process carried out for
20 min as described earlier. Though the total thickness of the porous alumina film is
different for samples with different anodizing time, the thickness of the dendritic structure
generated at the interface due to the modification procedure is found to be varying from
0.18 µm to 0.30 µm. It was difficult to find out the boundary between the unmodified and
modified layers. The porous alumina film thickness for normal and interface modified
samples as a function of anodizing time is plotted in the figure 5.2. From the figure, it can be
seen that the film thickness increases linearly with time of anodisation. For both porous
alumina and interface modified porous alumina samples, the thickness of the film increased
at the rate of ~0.108 µm/ min.

Figure 5. 1: SEM cross-section image of porous alumina modified interface in 0.3 M oxalic acid
solution at 18 0C with the anodizing time of (a) 5 min and (b) 10 min

109
Metal filled porous alumina samples were produced with film thickness of 1.2 µm.
The optimised barrier layer modification was used to obtain uniform filling according to
the procedure described in chapter 4. The thickness of all the samples was measured by
obtaining cross-section images of the samples in SEM. Indentation was carried out after
removing the overgrowth of copper by slight mechanical polishing.

Figure 5. 2: Film thickness-total time curves of porous alumina (PAA) and porous alumina
with modified interface (MPAA)

Indentations were carried out in a Nanoindenter (M/s Hysitron Inc) with a diamond
Berkovich tip. The tip radius was around 70 nm. The indentions were performed in load
controlled mode. For all the samples, the maximum load of 10 mN was applied with a
constant loading rate of 0.1mN/sec. At the end of loading, load was held constant for 2 sec.
unloading was carried out at the same rate as that of the loading. The indentations were
carried out in an array of 3x3 such that they can be easily located for imaging. SEM images
of all the indents were obtained to study the deformation characteristics. The deformation
in and around the indent observed were correlated with the load-displacement curves
obtained to understand the deformation characteristics. As there was differing amount of

110
elastic recovery obtaining hardness using area function method [77] was found not to be
appropriate. Instead areas of the indents were obtained directly from the images and were
used to obtain hardness.

5.3 Load-Displacement Characteristics


5.3.1 Porous Alumina

Porous alumina samples anodized for various time durations, with varying anodised
layer thickness (Table1) were indented. Figure 5.3 shows the load (P) –displacement (h)
curves of porous alumina sample on various film thicknesses at a maximum applied force
(Pmax) of 10 mN. Samples with low porous alumina thickness show greater depth of
penetration compared to higher thickness samples. Initial part of loading portion of all the
curves overlapped with each other. After that, the curves start to diverge. The low
thickness samples start to diverge first and pop-in behaviour follows. For high thickness
samples, the pop-in behaviour is observed at a displacement of about 100-200 nm.

Figure 5. 3: P-h curves of porous alumina at various film thicknesses

111
At the low thickness (0.17 µm and 0.31 µm), there is no pop-in like behaviour
observed, it might be because, when the thickness of the coating is very small pop in occurs
immediately after the start of the indentation and hence may not be prominently visible in
the P-h curve. At low film thickness the deformation is more as compared to high thickness.
At the thickness of 0.55 µm, 0.69 µm and 1.18 µm pop-in is observed. For the thickness of
0.55 µm pop-in is observed at around 650 nm 830 nm, it can be due to the sudden failure of
film and interface. For the thicknesses of 0.69 µm and 1.18 µm, the pop-in is observed
around 200-220 nm. Some of the loading curves show more than one ‘pop-in’ behaviour.
The unloading portions of all the P-h curves show plastic deformation with unloading
slopes almost identical.

Figure 5. 4: SEM images of indent of porous alumina at the thickness of(a) 0.17 µm,(b) 0.31
µm,(c) 0.52 µm,(d) 0.55 µm,(e) 0.69 µm and (f) 1.18 µm

112
SEM images of the residual indents for the samples whose P-h curves (shown in
figure 5.3) are shown in figure 5.4. All the indent show median and circumferential cracks
except for the 1.18 micron thick sample. The median cracks extend radially from the centre
of the indent along the line where the edge of the indenter came in contact. The
circumferential cracks are found on the face of the impression and almost parallel to a line
joining the vertices. For the indent on the sample with highest thickness only median
cracks are seen. Even the width of this median crack is minimal. It can also be seen that the
area of the residual impression is also very small.

As the spherical tip of the indenter comes into contact with the sample, the hard
alumina film initially deforms elastically. As the penetration increases, the subsurface
stresses increase. Eventually the soft aluminium will start deforming plastically. Since the
amount of deformation in aluminium is higher than that in the alumina, the film will start
bending eventually leading to cracks. The cracks are probably responsible for the pop-in
behaviour observed in the P-h curves. Once the film cracks the behaviour of P-h curve is
dominated by the deformation behaviour of the soft aluminium substrate.

Figure 5. 5: P-h curves of pure aluminium and Porous alumina with thicknesses of 0.55 µm
and modified porous alumina with the thickness of 1.4 µm.

113
Load-displacement curves from the nanoindentation experiments can be used to
obtain both qualitative and quantitative information. Comparing the P-h curves taken with
and without coating and with varying thickness of the coating can give information about
the deformation characteristics of the sample. Figure 5.5 shows the P-h curves for pure
aluminium and hard porous alumina films, on aluminium substrate, with 0.5 µm and 1.4
µm thickness. The penetration depth of pure aluminium is more, compared to porous
alumina. For pure aluminium the maximum penetration depth was around 1.3 µm, while
for the porous alumina thickness of 0.5 µm and 1.4 µm, the maximum penetration depth
was around 0.9 µm and 0.8 µm respectively. Unloading portion of P-h curve of pure
aluminium shows perfectly plastic behaviour. For sample with 0.5 µm thick film, the
unloading portion shows similar plastic behaviour. However, 1.4 µm thick film shows
considerable elastic recovery.

The SEM images of indents at constant load of 10 mN are shown in figure 5.6.
Figure 5.6(a) is the SEM image of indent on aluminium. There is no crack found in the
indent image but a little pileup is observed around the indent. Figure 5.6 (b) shows the SEM
image of indent of porous alumina with ta (=0.55 µm). This exhibits the edge cracks and the
circumferential cracks. At ta (=1.18 µm), it has edge crack (figure 5.6c) and the pop-in is
observed. Figure 5.6 d shows the SEM image of indent of porous alumina with the high film
thickness (=1.4 µm). There are no cracks found although the pop-in is observed, which
might be due to effect of substrate.

114
Figure 5. 6: SEM images of indent of (a) pure aluminium, (b) Porous alumina with thicknesses
of 0.55 µm, (c) Porous alumina with thicknesses of 1.1 µm and (d) Porous alumina with
thicknesses of 1.4 µm

Depth of penetration recorded for 0.17 µm thick porous alumina film for a
maximum load of 10 mN is around 1 µm. That is the contribution of deformation of
aluminium is high to the total depth of penetration. At the high thickness (1.4 µm)
maximum penetration depth in the P-h curve is less (0.8 µm) compare to the film thickness.
So in this case, the film is getting indented, not the substrate. That is the porous alumina is
harder than the aluminium substrate and can resist the indentation better.

5.3.2 Porous Alumina with Modified Interface

Interface of porous alumina was modified to fill metal into pores as we have
discussed in chapter 3. However, it is important to understand if there is any change in the
hardness and deformation characteristics of the sample due to the modified interface.
Figure 5.7 shows the P-h curve of porous alumina samples with the modified interface after
second anodisation for various film thicknesses for a maximum applied force (Pmax) of 10
mN. The P-h curves of the samples were similar for first few nanometres of depth of

115
penetration. After about 50 nm, the curves start to diverge. It is observed that the
maximum penetration depth decreases with increasing film thickness.

At low film thickness to high film thickness (0.49 µm to 1.4 µm), the deformation is
more as compared to high thickness. For the thickness of 0.49 µm pop-in is observed at
around 640 nm to 790 nm, it can be due to the sudden failure of interface. At the thickness
of 0.85 µm, pop-in is observed around 270 nm. At the thickness of 0.92 µm and 1.4 µm, pop-
in like behaviour is observed at around 510-520 nm, it might be because of high thickness.
The loading portion of P-h curve of porous alumina and modified porous alumina film is
almost same. Only at the thickness of (0.92 µm and 1.4 µm), the pop-in is repeatable but
does not occur at the same penetration. This might be because the porosity films are
different at the bottom compared to the top.

P-h curves for almost all the indents show ‘pop-in’ kind of behaviour. However, P-h
curves for low thickness samples exhibit multiple ‘pop-in’ at various indenter
displacements. Unloading portions of the curves are not same unlike normal porous
alumina samples. High thickness samples show significant elastic recovery during
unloading. While that of low thickness samples show almost perfect plastic behaviour.

Figure 5. 7: P-h curves of modified porous alumina with the film thicknesses of 0.49 µm, 0.85
µm, 0.92 µm and 1.4 µm

116
Figure 5. 8: SEM images of indent of modified porous alumina at the thickness of (a) 0.49 µm,
(b) 0.85 µm, (c) 0.92 µm, and (d) 1.4 µm

The corresponding SEM images of indents are shown in figure 5.8 a-d on the
samples with thicknesses of 0.49 µm, 0.85 µm, 0.92 µm and 1.4 µm respectively. Very low
film thickness exhibit multiple cracks which include circumferential cracks, and median
cracks as in the sample without modified. As the film thickness increases, it shows edge
cracks and circumferential cracks. Samples of thickness more than 0.85 µm show median
cracks but no circumferential cracks (figure 5.8 b and c). The sample with highest film
thickness of 1.4 m shows no cracks with the indent formed by compacted alumina. The
area of this residual impression is also small while the elastic recovery as observed from
P-h curve is highest.

117
5.4 Effect of Film Thickness
For hard coating on the softer substrate, in the initial stages of the indentation, as
the indenter comes into the contact with the film, support comes from the film than the
substrate. As the depth of penetration increases, depending on the thickness of the hard
coating, at a certain depth of penetration, substrate starts to influence the indentation
process. To minimize the influence of the substrate on the indentation process, the
maximum depth of the penetration is usually maintained to be less than 10% of the
thickness of the film[127].

Figure 5. 9: P-h curves of porous alumina with a thickness of 0.55 µm at different loads of 10
mN, 5 mN, 2.5mN and 1 mN

For porous alumina on aluminium substrate, when the indentation begins, at a very
low load, the film starts deforming. At a certain depth of penetration, substrate starts
deforming. Since the aluminium substrate is softer, a very low stress generates high strain
and the amount of deformation or strain will vary in both the porous alumina film and the
substrate. Since the strain is different in film and substrate, the substrate will deform more
plastically and hence the hard film has to bend to avoid delamination. This can result in
cracks either on the surface or at the interface. The presence of discontinuities in the
118
P-h curves can reveal information about any such deformation phenomenon such as
cracking and delamination.

We have carried out nanoindentation on porous alumina substrate for the loads
ranging from 1 mN to 10 mN to understand the deformation characteristics as a function of
applied load. P-h curve obtained for porous alumina samples for different load is shown in
figure 5.9. It shows that the loading portion of all the load displacement curves overlap
with very minimal scatter. The maximum depth of penetration at maximum load of 10 mN,
5mN, 2.5 mN and 1 mN are about 0.91 µm, 0.59 µm, 0.35 µm and 0.11 µm respectively. The
loading portion of the P-h curves show change in curvature at around 150 – 200 nm
probably due to the influence of the substrate which is 1.2 µm thick. At this point, P-h curve
with maximum load of 10 mN shows a pop-in at around 200 nm. In the same P-h curve, a
second pop-in is observed at about 800 nm. Unloading portions of the P-h curve with
maximum load of 1 mN showed significant elastic recovery. For the P-h curve at higher
loads unloading portion shows slight elastic recovery which is small compared to that
observed for smaller loads.

Figure 5. 10: SEM images of indent of porous alumina the thickness of 0.55 µm at various load
of (a) 1mN,(b) 2.5 mN,(c) 5mN and (d) 10 mN

119
Figure 5.10 shows the SEM image of indents at various maximum loads of 1mN, 2.5
mN, 5mN and 10 mN. When the maximum load applied is 1mN on a sample, there is no
crack seen, but in the indent at a load of 2.5 mN, it is seen that within the indent
median/edge cracks are visible, running parallel to the indenter edges. For the maximum
load of 10 mN, indent exhibits both circumferential and median/edge cracks.

Correlating the P-h curves with the SEM images of the indents can bring out the
indentation behaviour of the coating. For indentations at lower load (for example with
maximum load 1mN and 2.5 mN), as seen in figure 5.9, when the depth of penetration is
less compared to the film thickness, the indenter penetrates only into the film but not to
the substrate. Sample deforms due to the compaction on aluminium substrate (figure 5.6d).
However, the sharp edges of the indenter would generate a crack in the coating (figure
5.10b, c). These cracks can show a pop-in in the P-h curve. Since only single phase material
is indented there would not be any change in the curvature of the P-h curve (figure 5.9).

For indentations at higher load (for example with maximum load at 10mN in
figure 5.9), when depth of penetration is comparable to the film thickness of porous
alumina, the deformation due to the penetration of the indenter not only influences the film
but it is extended into the substrate. Since aluminium is softer than the porous alumina, the
strain in aluminium is more compared to the strain in porous alumina layer. This would
make the porous alumina layer bend and the area circumscribing the indent would come
under tension. This can lead to the circumferential cracking as seen in SEM image
(figure 5.10d). Generation of these types of cracks would give a pop-in in the P-h curve.

From the analysis, it can be inferred that the edge cracks occur when the maximum
depth of penetration is nearly equal to the film thickness. The film and substrate deform
plastically but the film does not seem to be bending as indent just touches to the substrate.
On other hand circumferential cracks on the surface occur when the maximum penetration
depth is more than the film thickness. In that case both the film and substrate deform
plastically and forms cracks due to bending.

120
5.5 Hardness of Film
Oliver-Pharr method which is generally used to compute the hardness and elastic
modulus from the P-h curve is accurate for the indentation of a single phase material
[128,129]. However, for the indentation of non-homogeneous material such as nano-
porous alumina this method cannot be applied [77]. The hardness of the film may not be
obtained precisely using the flat punch assumption made in the model [130]. Therefore, in
this study the hardness of the porous alumina coating on the aluminium substrate is
calculated by obtaining the projected area of the each indent from the corresponding SEM
images. Using projected area calculated from SEM images of indents, the hardness of the
porous alumina samples is calculated as

( 16 )

Figure 5. 11: Variation of hardness as a function of film thickness

121
Projected area of the indents measured by SEM is observed to be lower than the
area obtained by using the area function of the indenter. Hence the hardness calculated
from the projected area of the indent from SEM image is relatively higher as compared to
the one calculated by calibrating area function of the indenter. Figure 5.11 shows the
variation in the hardness as a function of film thickness. It is seen that the hardness is
increasing as the film thickness increases. The hardness is more for the film thickness of
0.9 µm compare with the thickness of 1.18 µm. That might be because of modified interface,
the porosity of porous alumina will be different, as hardness is proportional to the solid
area fraction [84,131].

Figure 5. 12 Maximum penetration depth vs film thickness

Figure 5.12 shows variation in maximum penetration depth (hmax ) as a function of


film thickness. hmax is decreasing as the film thickness increases. At the low film thickness
of 0.17 µm and 0.31 µm hmax is high (around 1.03 µm and 1 µm respectively). At the high
film thickness of 1.18 µm and 1.4 µm hmax is high around 0.90 µm and 0.819 µm
respectively.

122
Figure 5. 13:Hardness (H) vs maximum penetration depth to film thickness ratio (hmax/Ft) of
(a) porous alumina, (b) porous alumina with modified interface and (c) combined both
porous alumina and modified porous alumina

Figure 5.13 shows the hardness vs. maximum penetration (hmax) to film thickness
ratio (ft) of porous alumina with modified interface and combined both porous alumina and
modified porous alumina. The deformation characteristics of porous alumina are examined
in figure 5.13a, when the hmax/ft is more than 1, the circumferential and edge cracks are
observed. When the hmax/ft is more or less equal to 1, the edge cracks are observed. When

123
the hmax/ft is less than 1, there are no cracks observed and deformation is only due to
compaction. Similar trend is found in modified porous alumina (figure 5.13b).

Both porous alumina and modified porous alumina films exhibits similar
deformation characteristics. The combined hardness (H) vs maximum penetration depth to
film thickness ratio (hmax/Ft) is shown in figure 5.13c. The hmax/ft is almost equal to 1 the
corresponding hardness is around 2GPa. When the hmax/ft is less than 1 the corresponding
hardness is about 5.5GPa and when it is more than 1 the corresponding hardness is about
1 GPa. In order to prevent cracks, the porous alumina film, hmax/ft should be less than or
equal to 1 otherwise film will fail.

“Hence it is seen that porous alumina exhibit two different deformation mechanisms
as a function of hmax/Ft. Hardness varies from 1GPa to 5GPa. When the hardness is about
1GPa, oxide, the film deforms and generates circumferential cracks on the surface. When
the hardness is about 5GPa, the film deforms by compaction”.

5.6 Indentation on Nanocomposite


Nanoporous alumina coating is brittle. Empty pores of the coating crumble under
load and undergo compaction under compressive stresses. The film under tension can
easily crack. Here we explore the possibility to improve the mechanical properties of the
coating by understanding the deformation characteristics of metal filled porous alumina
samples.

Ordered porous alumina generated through two step anodisation processes and
filled with metal (as described in chapter 4) is indented. P-h curve for metal filled porous
alumina sample with thickness of the film measured to be around 1.2 µm along with that of
empty porous alumina sample is shown in figure 5.14b. It can be seen from the loading
portion that two curves diverge within few nanometers of depth of penetration. The curve
for metal filled porous alumina is steeper compared to the normal porous alumina sample.
At around 200 nm- 250 nm of depth of penetration the curvature of both the samples
decreases sharply indicating that samples are now deforming more easily. This might due

124
to the effect of aluminium substrate. At this point, P-h curve for the empty sample showed a
pop-in while for the metal filled sample no such pop-in is observed at any point in the
P-h curve.

Figure 5. 14: (a)SEM images of indent of Cu filled porous alumina and (b) P-h curves of porous
alumina and Cu filled porous alumina. Porous alumina samples are used in both the cases
after voltage reduction with the reduction rate of 1V/15s followed by 2hr pore widening.

The maximum depth of penetration for metal filled sample is significantly smaller
than for empty sample. Hardness calculated for the metal filled sample is around 4.2 GPa
which is around 50% increase from that of the empty sample. SEM image (figure 5.14a) of
the corresponding indent from the metal filled sample showed no cracks.

5.7 Discussion
As we have discussed in chapter 4, porous alumina has wide application in
fabricating nanomaterials like nanowires. Many studies have been carried out in
characterising optical and electrical properties of porous alumina [104,132–135]. However,
very few studies, for evaluating mechanical properties like hardness and elastic modulus of
porous alumina, has been reported [78,82,84,136]. Nanoporous alumina coating is brittle,

125
and under the concentrated load the porous alumina coating fails by compaction and
cracking.

Ng et al [82,83] have studied the deformation behaviour of porous alumina of very


large thickness compared to the penetration depth. They observed the cracks (median and
bilinear) on the indent area. We have carried out indentation for different thickness to
penetration depth ratio. It was found that for a given indenter penetration, the film
thickness has a direct influence on the deformation behaviour and hence the hardness.
Hardness is found to increase with the film thickness for a given maximum load of 10 mN
(figure 5.11) upto a film thickness of about 1.5 m. Two different deformation mechanisms
cracking and compaction are observed, at low thicknesses, the film deformed by cracking
and by compaction at higher thicknesses. The behaviour of PAA film with modified
interface is found to be similar to that of PAA. The interface modification has very minimal
effect on the measured hardness.

Mechanical properties of soft film/hard substrate and hard film/soft substrate


material has been widely studied [137–142]. Generally it is recommended that the
penetration should be less than 10% of the film thickness [140,143] to avoid substrate
effect. Assimina et al [140] reported that hard film/soft substrate material system, the
calculated hardness of the thin film decrease when maximum penetration depth to film
thickness hmax/Ft ratio is equal to 0.40. We have carried out indentation with this ration
greater than 0.4. Deformation behaviour of the PAA coating was observed to be dependent
on maximum penetration depth to film thickness (hmax/ft) ratio. Compaction occurs when
this ratio is less than 1 (hmax/ft < 1) and Cracking occurs at higher ratio (hmax/ft > 1). In the
case of hmax/ft > 1, the substrate deforms plastically and the film has to bend to
accommodate the substrate deformation. This results in tensile forces along the radial
direction resulting in circumferential cracks. For a tribological coating, cracking is not
desirable [122]. So, in order to avoid cracks, the contact loads should be such that the
penetration of the asperity or any hard indenter should be less than the film thickness..

To study the effect of filled pores on the hardness indetentation of modified porous
alumina and metal filled porous alumina nanocomposite samples were carried out for the

126
same overall film thickness. Copper filled alumina nanocomposite coatings showed lesser
propensity to crack compared to porous alumina. No cracks were seen on the metal filled
porous alumina up to a maximum load of 10 mN (figure 5.14a). P-h curve for the porous
alumina showed a pop-in [8] while for the metal filled porous alumina no such pop-in is
observed at any point in the P-h curve (figure 5.14b). This further proves that the copper
filled alumina has higher resistance to cracking. P-h curve for metal filled porous alumina
is found to be much steeper compared to the porous alumina sample. This means that the
metal filled alumina structure higher hardness. Hardness is found to be increased by about
50%.

5.8 Conclusions
In this chapter, we studied the hardness and deformation characteristics of porous
alumina coatings in order to understand its effectiveness as a tribological coating. Hardness
of the porous alumina coating was observed to be dependent on hmax/ft ratio. The
magnitude of hardness varied from 1 GPa for hmax/ft ratio of around 6 and 5 GPa for hmax/ft
of around 1. The coatings exhibited two deformation behaviours at the two extreme hmax/ft
ratios. Porous alumina coating deformed by compaction when the ratio of hmax/ft < 1.
However, when the ratio was hmax/ft > 1, the coating was deformed by forming median and
circumferential cracks inside and around the indent respectively. For a tribological coating,
cracking is not desirable. Hence, porous alumina can be used as a tribological coating when
the ratio of hmax/ft < 1.

Metal filled in pores of the alumina structure increased the hardness of the coating
by around 50%. No cracks were observed in the SEM image of the indent for a maximum
load of 10 mN. Hence it is interesting to explore the tribological properties of metal filled
porous alumina coating.

127
Chapter 6

6. Tribological Characterization of Porous Alumina


based Nanocomposite

6.1 Introduction
Nanoindentation of the porous alumina and the copper filled composite showed that
they have higher hardness than the aluminium substrate. The hardness of the filled
nanocomposite is 50% higher than the corresponding unfilled porous alumina. Further the
nanocomposite had no circumferential cracks. In this chapter we present the details of our
study on friction and wear characteristics.

This study is preliminary study. More detailed and controlled experiments would be
needed to explore all the tribological advantages of the nanocomposite we have developed.
Hence we report on the flat on flat reciprocating experiments (section 6.2). Friction
measurements for different sample are shown in section 6.3. It is shown that copper
nanocomposite has less coefficient of friction compared to pure aluminium and porous
alumina. In section 6.4, the wear mechanism is discussed for different loads. For low
coating thickness, nanocomposite develop has cracks whereas for high thickness no cracks
were observed. It is also seen that for high thickness nanocomposite undergoes negligible
wear compared to aluminium and porous alumina.

128
6.2 Experimental Details
A ball on flat reciprocating tribometer (Ducom TR 282-M111) was modified to a flat
on flat configuration. Since we had developed the nanocomposite with a film thickness of
about 1 micron, the tribological experiments had to be carried out at lower loads. For
experiments with higher load we would need a thick nanocomposite coating. We have
observed uniform filling of copper can be carried out in porous alumina with much higher
thickness. However since most of our study had concentrated on 1 µm coating, we
restricted tribological study also to the sample with the same thickness. We have used flat
on flat, as contact pressure can be controlled easily without damaging the film.

Most of the reciprocating contact experiments were performed under dry condition.
The schematic diagram of modified tribometer is shown in figure 6.1. A indenter made of
steel ball of radius 6 mm was reciprocated with the help of a DC motor. The rotary motion
of the motor was converted to linear reciprocation motion by a Scotch-Yoke mechanism
and a linear bearing. This alumina composite coating sample was attached to the ball
through a spherical bearing as shown in the inset of the figure 6.1. Enough precautions
were taken to make sure that conformal contact was established between the sample and
the stationary steel counter surface. The steel counter surface was mechanically polished
stainless steel fixed in the sample holder. The sample was fixed on a reciprocating arm
which moves linearly. Normal load of 100 g, 200 g and 300 g were applied using dead
weights. All the experiments were performed at room temperature of 25°C in air of 48%
relative humidity. The experimental conditions are given in table 1. All experiments were
done at constant rate of 5 cycles/sec (5 Hz), and stroke length of 2mm. Before reciprocating
tests, all samples and stainless steel counter surfaces were cleaned with an ultrasonic
cleaner in deionized water for 5 min and then in ethanol for 10 min.

129
Figure 6. 1: Schematic diagram of the experimental setup

Table 6.1: Experimental Conditions

Experimental Parameters Experimental data

Normal Load 100 g, 200 g, 300g


Reciprocating distance 2 mm/cycle
Sliding frequency 5 Hz
Average sliding velocity 10 mm/sec
Relative humidity 44 ±5%
Temperature Room temperature
Lubrication condition Dry
Samples Pure aluminium, porous alumina,
porous alumina/copper
nanocomposite
Substrate
Stainless steel -Flat

130
Figure 6. 2: The raw friction force signal in volts. The mean values of difference in upper and
lower values give the friction force.

A piezoelectric force sensor (PCB Model 208 C01) was used to measure the
frictional force continuously using DAQ card and Labview. A constant voltage of about 2 V
was supplied to the DC (Agni 0-24 V) motor to get 5 cycles/sec reciprocation motion. The
friction force (in Newton) data was collected at an acquisition rate of 1000 per sec. Figure
6.2 show the raw friction force data in volts against the sliding time obtained from the
piezoelectric sensor. The raw data was processed using MATLAB and the mean values of
upper and lower cycle were obtained (shown by dotted lines in figure 6.2). The difference
in mean value of upper and lower cycles gives two times the mean frictional force in each
cycle. The friction force is obtained in newton by using the following equation:

( 17 )

where LM, UM and S are the mean of upper cycle voltage output from the force
sensor, mean of lower cycle voltage and sensitivity respectively.
131
The coefficient of friction (COF), the ratio of frictional force and the normal load was
obtained as,

( 18 )

where f is the frictional force and N is the Normal load.

6.3 Friction
Tribological measurement was performed on pure aluminium, porous alumina and
porous alumina/copper nanocomposite to deter mine the coefficient of friction against
stainless steel counter surface. Figure 6.3 shows the frictional force traces measure for
pure aluminium, porous alumina and porous alumina/copper nanocomposite at the normal
load of 100 g. It can be seen that, for porous alumina/copper friction force is less compared
to porous alumina and aluminium. The frictional force of pure aluminium was around 0.5
and remains almost constant over the time. The frictional force of porous alumina, porous
alumina/copper was around 0.6 and 0.4 respectively. It is observed that for both low and
high thickness of porous alumina/copper, the frictional force remains almost same.

132
Figure 6. 3: Friction force against the sliding cycles for pure aluminium, porous alumina and
porous alumina/copper nanocomposite. For all the experiments the substrate was
mechanically polished stainless steel.

Figure 6.4 shows the comparison chart of the coefficient of friction variation for
pure aluminium, porous alumina and porous alumina/copper. The coefficient of friction of
pure aluminium was around 0.5. For aluminium, the coefficient of friction was high which
might be because of the ductility and probably because aluminium suffered a bulk plastic
deformation. Under sliding, a very heavy transfer and retransfer of particle which are
firmly bonded occurs between aluminium and stainless steel. This might be due to high
adhesive force between pure aluminum and steel. The coefficient of friction of pure
aluminum was measured to be around 0.62.

For porous alumina, the coefficient of friction was very high, which might be
because of the substrate and peeling of a layer of porous alumina when porous alumina and
stainless steel come in contact. As a result, a high tangential force is required to carry out
the reciprocating sliding motion resulting in high coefficient of friction in porous alumina.
Both, abrasion resulting from the hard alumina wear particles and adhesion between
exposed aluminum or transfer layer and steel may be contributing towards the friction
simultaneously.
133
When porous alumina/copper nanocomposite was tested, the coefficient of friction
was around 0.39 and 0.4 for both low and high thickness sample respectively. The
coefficient of friction was reduced from roughly 0.62 to 0.39 compared to porous alumina.
The reduction in the coefficient of friction at both low and high thickness of porous
alumina/copper nanocomposite might be because of Cu particles coming in contact
between the porous alumina/copper composite and stainless steel and acting as a self-
lubricant, thus resulting in the decreased coefficient of friction.

Figure 6. 4: Bar chart of coefficient of friction with varying load

6.4 Wear Mechanisms


Material properties such as hardness, elastic modulus, thermal conductivity and
thermal diffusivity influences wear. Yet at the same time, the parameters that control wear
are not clear [122]. However under the dry sliding conditions, material properties like
hardness are important in deter mining the wear resistance. As we have demonstrated in
chapter 5, the porous alumina has very high hardness. In this section, results from dry
sliding wear test performed on pure aluminium, porous alumina and porous
alumina/copper nanocomposite are discussed.
134
6.4.1 Wear Track on Pure Aluminium

Figure 6. 5: SEM image of wear track of Aluminium sliding against stainless steel substrate

Figure (6.5) shows the SEM images of wear track on aluminum after normal load of
300 g was applied against the stainless steel flat substrate under sliding. Sliding direction
is as indicated by the double arrow in the figure. Even though the mean contact pressure
applied (of the order of 10 MPa) was much smaller than the hardness of the aluminium,
localized wear occurred because of surface roughness. It should be noted that the
roughness of the surface reciprocated were very smooth. The steel was diamond polished
while the alumina sample was electropolished before anodisation. A highly layered
structure is observed at the top surface which might be because of the high ductility of
pure aluminium which might cause it to easily undergo high plastic deformation under
sliding. When aluminium slides against stainless steel, a transfer film is formed on
stainless steel during the forward stroke, which gets transferred back to aluminium
during the reverse stroke [144]. The transfer film formed may be a mixed layer which is a
combination of both stainless steel and aluminium. The origin of the mixed transfer layer
might be the beginning of wear. Aluminium is highly ductile and hence a highly thick
135
mixed layer structure would exist which causes gouging of the aluminium with adhesion
thereby leading to a high adhesion wear rate. Wear debris observed on the surface might
be the loose adhered particles.

6.4.2 Wear Track on Porous Alumina

The process of wear results in geometrical and structural changes when two
surfaces come into contact. These changes range from nano-scale to macro-scale [122]. At
the nano-scale, asperities get reduced under sliding whereas at the micro-scale cracks gets
generated and debris is released. At the macro-scale, wear debris is agglomerated and a
surface layer is formed and deformed. However, the nano textured structure of porous
alumina can provide good wear resistance due to the high hardness porous alumina. In this
section, we study the wear characteristics of porous alumina.

Figure 6. 6: SEM images of wear track of (a) porous alumina and (b),(c) and (d) are the
enlarged image of porous alumina at three different locations

136
Figure 6.6 shows the SEM image of wear tracks of porous alumina after a normal
load of 300 g were applied under sliding against the stainless steel flat substrate. Numerous
cracks are observed on the surface (figure 6.6a), which may have been caused due to the
brittle nature of porous alumina. The porosity and brittle nature of porous alumina may
give rise to cracks when subjected to tangential loads caused by sliding. The enlarged
image near the crack reveals that pores still exist at the bottom (figure 6.6b) and the darker
contrast can be dimpled structure of aluminium. This shows that the cracks are forming
only on top layer while the inner subsurface of the porous film still exist without
delamination. The transfer layer can be the combination of porous alumina and stainless
steel substrate. The crack formation and propagation that happens only at the top layer
might be due to the small penetration depth of the porous alumina against the steel
substrate.

The enlarged partial portion of the wear track area of porous alumina in figure
(6.6c) shows a compacted layer kind of structure with sparsely distributed wear debris.
The wear debris found adhered inside the wear track area, may have been formed due to a
combination of abrasive and adhesive wear and may result in high values of friction
coefficients. The sliding wear response in figure 6.6d shows interesting features with few
pores found to be filled with wear debris. The filling of the pores may have occurred due to
the tiny abrasive particles formed just a few cycles after the commencement of sliding.

The above discussion shows that when porous alumina is subjected to sliding under
the applied normal load, even though crack formation and propagation occurs, the cracks
are restricted only to the top layers. This shows that porous alumina is strong enough to
withstand the contact pressure applied under sliding.

6.4.3 Wear Track on Porous Alumina/Copper Nanocomposite

With a very hard porous alumina film on a softer aluminium substrate, it has been
observed in the above section that at the given load cracks formation occurs on the top
layer. The coefficient of friction was also found to be high. A combination of brittle and
ductile coating may prove to be more beneficial. For example in the combination of ductile
137
and brittle DLC/Carbide multilayered coating, the elastic layer allows the brittle layer to
slide over each other in a manner of multi leaf book when bent [122].The combination of
ceramic/metal nanocomposites can improve wear characteristics as it provide one hard
phase and one soft phase [145,146]. Here we are studying the wear mechanism of porous
alumina/copper nanocomposite coating which is textured within the matrix. In the
following section, we have studied the wear mechanism of this coating at both low and high
coating thickness.

6.4.3.1 At Low Thickness

Figure 6.7 shows the SEM image of wear track of porous alumina/copper
nanocomposite of low thickness (~1 µm) sliding against stainless steel after applying a
normal load of 300 g. The wear mechanism observed is different from that of porous
alumina. Inside the wear tracks, very less wear debris and fewer cracks are observed. In
figure (6.7a), a transfer layer is observed, which could be the combination of porous
alumina/copper nanocomposite and stainless steel (figure 6.7a). At the top, few scratches
are observed which might have been caused due to a hard abrasive particle entrapped in
between the contact during sliding. However these scratches confirm that a transfer layer
exists. However, the surface beneath the top layer seem to be unaffected by the sliding.

138
Figure 6. 7: SEM images of wear track of (a) porous alumina/copper nanocomposite at low
thickness (~1 µm) and (b),(c) and (d) are the enlarged image of porous alumina/copper
nanocomposite at three different locations

A wavy surface morphology with peak and valley is observed in figure (6.7b) and
(6.7d). This might be because of the fact that, under dry conditions abrasive wear occurs
during sliding. The hard abrasive wear particles formed could be reason for the multitude
of scratches observed on the transfer layer. In Figure (6.7c) some cracks are observed. The
small wear particles that form during the initial cycles after the commencement of sliding
would agglomerate to form larger particles in the subsequent cycles. These large sharp and
hard particles rub against the steel substrate during sliding which eventually results in the
generation the crack.

6.4.3.2 At High Thickness

Usually, geometry related parameters like surface roughness, film thickness, hardness and
wear debris influences on the wear mechanism of the coating. In this section we use high

139
thickness (~3 µm) of porous alumina/copper nanocomposite coating for sliding. The wear
track image of porous alumina/copper nanocomposite at high thickness sliding against
steel is shown in figure (6.8).

Figure 6. 8: SEM images of wear track of (a) porous alumina/copper nanocomposite at high
thickness (~3 µm)

From the figure, the surface is completely free from cracks. The wear debris
formation is almost negligible. As sliding ensued some Cu debris forms on the surface. This
Cu debris acts as a lubricating medium between the coating and steel and prevents not only
the wear of the coating but also the crack formation. The white spots on the surface seen in
the image may be the Cu debris particles. Hence we find that porous alumina/copper based
nanocomposite coating proves to be extremely effective in improving the abrasive wear
resistive coating under dry conditions.

140
6.5 Discussion
According to Archard law, wear resistance is proportional to hardness of the
material [86]. Any hard coating will thus improve wear resistance [122–124]. However one
of the main issues with hard coatings is that they are brittle and hence fail by cracking and
delamination[122, 125] . This problem is higher for thin hard coatings on softer substrate
as the substrate deform plastically subject the coating to bending stresses. However, softer
substrate is desirable for higher toughness of the component. One possible solution for this
problem is to provide with alternate layers of hard and soft coatings [145,146]. The
combination of soft and hard DLC/Carbide multilayered coating, the soft material allows
the hard material to slide over each other in a manner of multi leaf book when bent [122].
Other possible solution is to develop a composite coating like Cu filled porous alumina in
which the layers are aligned in vertical direction with soft copped nano-rods blunting the
crack tips.

For pure aluminium, the coefficient of friction (COF) was found to be about 0.6. This
high value might be because of the fact that as aluminium is ductile it can undergo large
amount of plastic deformation. Also, of formation of a transfer layer that has high adhesion
between aluminium and stainless steel could result in high COF. PAA being ceramic with
covalent bonds has lower adhesion with steel and hence has lower COF.

The COF of porous alumina/copper nanocomposite was found to be lowest


compared to PAA and aluminium. This was found for both low (~1 µm) and high (~3 µm)
thickness films. Cu particles from the wear debries may be forming a low shear strength
transfer layer between hard alumina and the steel counterface. This could be the reason for
the decreased coefficient of friction. Though SEM images give some clues to this, further
detailed analysis may be needed for conclusive proof.

From SEM images it is clear that pure aluminium has highest wear. The worn
surface exhibits show a highly mixed layer. It has been reported [144] that when
aluminium slides against stainless steel, a transfer film is formed on stainless steel during
the forward stroke, which gets transferred back to aluminium during the reverse stroke.
The formation of the mixed transfer layer starts from the very beginning of the wear
141
process [122–124] . This transfer film might be a mixed layer consisting both stainless
steel and aluminium. Any break in the film causes gouging of the aluminium due to its
high adhesion with steel. The worn aluminium particle then gets entrapped forming new
transfer layer. Few wear debris observed on the surface might be such worn particles
loosely adhering to the surface.

Ning-ning et al [94] found that the porous alumina have high hardness and wear
resistance. But as we have discussed that the porosity and brittle nature of porous
aluminium may give rise to cracks when subjected to tangential loads caused by sliding as
shown in figure 6.6. These cracks are formed only on top layer without affecting the inner
subsurface of the porous film. This might be due to the small penetration depth of the
PAA against the steel substrate at low loads. Wear debris were found adhered within the
wear track area and this may indicate that a combination of abrasive and adhesive wear
mechanisms are operating. Few pores are found to be filled with wear debris. This filling
of the pores may have occurred due to the tiny abrasive particles formed just a few cycles
after the commencement of sliding.

Wang et al [91–93] has reported that the thick anodic oxide film filled with PTFE
having low coefficient of friction and less wear. In our experiments, the porous
alumina/copper nanocomposite of low thickness (~1 µm) and high thickness (~3 µm) is
found to have negligible wear. At low thickness within the wear track, transfer layer and
few scratches are observed. Those scratches might be due to a hard abrasive particle
entrapped in between the contact during sliding (figure 6.7a). However, the surface
beneath the top layer was unaffected. Small wear particles would be formed during the
initial cycles of sliding and these sharp and hard particles may get embedded on the soft
steel substrate. This could result in the generation of cracks on the top layer of the
composite coating. For high thickness sample, the surface was found completely free from
cracks. The formation of wear debris was almost negligible. In this case under the sliding,
some Cu debris might have formed on the surface and acts as a lubricating medium
between the coating and steel and prevents wear. As a result, porous alumina/copper
based nanocomposite coating may prove to be effective in improving the wear resistance
under dry conditions.
142
6.6 Summary
 Coefficient of friction for porous alumina and aluminium samples were high while
it was slightly less for porous alumina/copper nanocomposite.
 In porous Alumina crack formation occurs at the top layer but the sub-layer seems
to be remaining unaffected, it could be the dimpled structure of aluminium.
 Wear and crack were found to be less in porous alumina/copper than in porous
alumina and aluminium.
 A thin transfer layer is observed on the surface of porous alumina/copper
nanocomposite which could be the combination of copper/steel gets coated on
alumina
 At low thickness of porous alumina/copper nanocomposite coating few cracks are
found on wear track area.
 At high thickness of porous alumina/copper nanocomposite coating no cracks are
found.

143
Chapter 7

7. Conclusions

The main focus of this work was to develop a novel porous alumina based metal
nanocomposite coating for tribological applications. Using a two-step anodisation process,
in 0.3 M oxalic acid, and by controlling anodisation parameters, we have obtained a highly
ordered hexagonal porous structure on pure aluminium. It was observed that the pore
diameter and inter-pore distance are highly dependent on the anodizing voltage. The pore
diameter was linearly proportional to the anodizing voltage. The thickness of the film
increased linearly with the time of anodisation. By performing anodisation at constant
voltage and varying the anodisation time, we achieved precise control of the film thickness
with minimal change in the porous structure.

The interface of the porous structure and aluminium is characterised by an


insulating barrier oxide layer. To deposit metal into the pores using electrodeposition
process, we were able to create a dendritic structure in the barrier layer by step-wise
voltage reduction towards the end of anodisation process. This dendritic structure helped
to form a conducting path of electrolyte to aluminium substrate thereby enabling
electrodeposition. The thickness of this dendrite like structure depended on the time over
which this modification was carried out.

To fill metal into the porous structure we have tried 3 different electrodeposition
processes: DC, AC and pulse electrodeposition. We have found that the use of pulse
electrodeposition is highly efficient and well suited method for metal filling into the pores.

144
Using this process along with optimization of barrier layer thinning process we have
achieved uniform metal filling into porous alumina. We have found that barrier layer
thinning using the voltage reduction process at the rate of 1V/15s was highly efficient for
almost 100% filling of metal into pores without weakening the interface.

We have carried out indentation studies of hard porous alumina coating on the soft
aluminium substrate to understand the hardness and deformation characteristics. The
nano porous alumina coating thickness depends on anodisation time and increases with
anodisation time at the rate of ~0.108 µm min-1. It is observed that the coating shows two
deformation behaviours with the change in depth of penetration to film thickness (hmax/Ft)
ratio. At high hmax/Ft ratio coating exhibits cracks and at low hmax/Ft ratio the deformation
was mainly due to compaction. We have found that dendritic modification of interface has
very little effect on the deformation characteristics compared to that of samples with no
interface modification. When copper is filled in the nano pores, the hardness increased by
50%.

We have evaluated the tribological properties of the nanocomposite coating by


measuring friction and wear in a reciprocating wear test. Coefficient of friction for porous
alumina and Al samples were high while it was 30% less for porous alumina/copper. In
porous alumina, crack formation occurs at the top layer but the sub layers remain
unaffected. From the SEM images wear and crack were found to be less in
porous alumina/copper than in porous alumina and aluminium. A thin transfer layer of
copper/steel gets coated on alumina. At low thickness of porous alumina/copper
nanocomposite coating few cracks are found on wear track area. At high thickness of
porous alumina/copper nanocomposite coating no cracks are found.

To conclude, we have developed a novel nanocomposite coating with nanoporous


alumina as matrix with aligned metal nano rods. This was achieved by optimally modifying
the barrier layer without sacrificing the interfacial strength. Uniform coating was achieved
over an area of 10 mm x 10 mm. The coating is found to have higher hardness and higher
wear resistance.

145
Appendix

Appendix A Electrochemical Impedance spectroscopy (EIS)


Electrochemical Impedance spectroscopy (EIS) is used extensively to characterize
the frequency dependent properties of electrochemical interfaces. EIS is obtained by
applying an electrical stimulus (voltage) between sample and the solution and observing
the response (current) at the different frequencies.

Most widely used equivalent circuit to model the electrochemical interfaces is


Randles circuit. Randles circuit assumes that electrolyte resistance, RS is in series with a
parallel combination of the double-layer capacitance, Cdl and a charge transfer resistance,
RCT [116]. The equivalent circuit for electrochemical interfaces can be simplified as shown
in figure A.1. The impedance Z, of the Randles circuit can be expressed as

( A1 )

where, Rp representing the resistance for the ions to diffuse through the narrow path in
pores and ω is the measuring frequency. The impedance Z is the combination of both real
value (resistance) and complex value (reactance).

146
Figure A.1: Simplified version of electrical equivalent randles circuit for porous alumina

More widely, the EIS is expressed as a Nyquist plot which provides visible and
characterizing information about electrode process. Nyquist plot shows the real versus
imaginary components of impedance at different frequency of the applied voltage signal
figure A.2 shows Nyquist plot obtained from EIS measurements, carried out on porous
alumina after 2 hr pore widening at the voltage reduction rate of 1V/15s. A 100mV (rms)
amplitude signal has been applied to the porous alumina/solution interface, with a
frequency sweep from 20 Hz to 2MHz, with a logarithmic increment. The Nyquist plot
obtained is closed to semicircle and matches with that of Randles circuit except for the low
frequency diffusion dominated section.

Figure A.2: Nyquist plot for different pore widening times for a voltage reduction rate of
1V/15s.

147
In general, the most reliable parameter an EIS can determine is solution resistance
(Rs), when the AC frequency is high (>1MHz). Since when frequency is very large, the
imaginary component of impedance tends to zero, and real component of impedance will
be equal to the solution resistance. Thus the Nyquist plot intercept on the real component
of impedance-axis at the high frequency end (figure A.2) will be the value Rs. The value
of Rs, obtained from the curve, for porous alumina is 68 ± 2Ω. The diameter of the semi-
circle in the Nyquist plot is called the charge transfer resistance RCT. Thus the Nyquist plot
intercept on the x-axis at the low frequency end gives the sum of the charge transfer
resistance and the solution resistance (Rs+RCT =2200 ± 10Ω). At the centre on the real axis,
and the angular frequency corresponding to highest point in the imaginary axis is called the
characteristic frequency ω, which is equal to 1/RCTCdl.

148
Bibliography

[1] Wernick S, Pinner R and Sheasby P G 1992 The surface treatment and finishing of
aluminum and its alloys Wernick R Pinner P G Sheasby 1320 Pages 5 Th Ed. 2 Vol. 492
Illus. 230 Tables 2534 Ref. Price 170US 340 00 Book.

[2] Diggle J W, Downie T C and Goulding C 1969 Anodic oxide films on aluminum Chem.
Rev. 69 365–405.

[3] Alkire R C, Gogotsi Y, Simon P and Eftekhari A 2008 Nanostructured materials in


electrochemistry (John Wiley & Sons).

[4] Thompson G 1997 Porous anodic alumina: fabrication, characterization and


applications Thin Solid Films 297 192–201.

[5] Pringle J 1972 A very precise sectioning method for measuring concentration
profiles in anodic tantalum oxide J. Electrochem. Soc. 119 482–91.

[6] Brown F and Mackintosh W 1973 The Use of Rutherford Backscattering to Study the
Behavior of Ion-Implanted Atoms During Anodic Oxidation of Aluminum: Ar, Kr, Xe,
K, Rb, Cs, Cl, Br, and l J. Electrochem. Soc. 120 1096–102.

[7] Cabrera N and Mott N 1949 Theory of the oxidation of metals Rep. Prog. Phys. 12
163.

[8] Hoar T and Wood G 1962 The sealing of porous anodic oxide films on aluminium
Electrochimica Acta 7 333–53.

[9] Keller F, Hunter M S and Robinson D L 1953 Structural Features of Oxide Coatings on
Aluminum J. Electrochem. Soc. 100 411–9.

149
[10] O’sullivan J and Wood G 1970 The morphology and mechanism of formation of
porous anodic films on aluminium Proc. R. Soc. Lond. Ser. Math. Phys. Sci. 511–43.

[11] Masuda H and Fukuda K 1995 Ordered metal nanohole arrays made by a two-step
Science 268 1466–8.

[12] Lei Y, Cai W and Wilde G 2007 Highly ordered nanostructures with tunable size,
shape and properties: a new way to surface nano-patterning using ultra-thin alumina
masks Prog. Mater. Sci. 52 465–539.

[13] Sander M s., Prieto A l., Gronsky R, Sands T and Stacy A m. 2002 Fabrication of High-
Density, High Aspect Ratio, Large-Area Bismuth Telluride Nanowire Arrays by
Electrodeposition into Porous Anodic Alumina Templates Adv. Mater. 14 665–7.

[14] Nielsch K, Müller F, Li A-P and Gosele U 2000 Uniform nickel deposition into ordered
alumina pores by pulsed electrodeposition Adv. Mater. 12 582–6.

[15] Gao T, Meng G, Zhang J, Wang Y, Liang C, Fan J and Zhang L 2001 Template synthesis
of single-crystal Cu nanowire arrays by electrodeposition Appl. Phys. A 73 251–4.

[16] Yin A, Li J, Jian W, Bennett A and Xu J 2001 Fabrication of highly ordered metallic
nanowire arrays by electrodeposition Appl. Phys. Lett. 79 1039–41.

[17] Kline T R, Tian M, Wang J, Sen A, Chan M W and Mallouk T E 2006 Template-grown
metal nanowires Inorg. Chem. 45 7555–65.

[18] Sun M, Zangari G and Metzger R M 2000 Cobalt island arrays with in-plane
anisotropy electrodeposited in highly ordered alumite Magn. IEEE Trans. On 36
3005–8.

[19] Parkhutik V and Shershulsky V 1992 Theoretical modelling of porous oxide growth
on aluminium J. Phys. Appl. Phys. 25 1258.

[20] Hoar T P and Mott N F 1959 A mechanism for the formation of porous anodic oxide
films on aluminium J. Phys. Chem. Solids 9 97–9.

[21] Li F, Zhang L and Metzger R M 1998 On the growth of highly ordered pores in
anodized aluminum oxide Chem. Mater. 10 2470–80.

150
[22] Siejka J and Ortega C 1977 An O18 Study of Field-Assisted Pore Formation in
Compact Anodic Oxide Films on Aluminum J. Electrochem. Soc. 124 883–91.

[23] Thompson G E and Wood G C 1981 Porous anodic film formation on aluminium
Nature 290 230–2.

[24] Su Z and Zhou W 2008 Formation mechanism of porous anodic aluminium and
titanium oxides Adv. Mater. 20 3663–7.

[25] Thompson G, Furneaux R, Wood G, Richardson J and Goode J 1978 Nucleation and
growth of porous anodic films on aluminium.

[26] Orlik M 1999 On the onset of self-organizing electrohydrodynamic convection in the


thin-layer electrolytic cells J. Phys. Chem. B 103 6629–42.

[27] Su Z and Zhou W 2009 Formation, microstructures and crystallization of anodic


titanium oxide tubular arrays J. Mater. Chem. 19 2301–9.

[28] Lu S, Su Z, Sha J and Zhou W 2009 Ionic nano-convection in anodisation of


aluminium plate Chem. Commun. 5639–41.

[29] Thompson G 1997 Porous anodic alumina: fabrication, characterization and


applications Thin Solid Films 297 192–201.

[30] Cherki C and Siejka J 1973 Study by nuclear microanalysis and O18 tracer techniques
of the oxygen transport processes and the growth laws for porous anodic oxide
layers on aluminum J. Electrochem. Soc. 120 784–91.

[31] Parkhutik V, Belov V and Chernyckh M 1990 Study of aluminium anodization in


sulphuric and chromic acid solutionsII. Oxide morphology and structure
Electrochimica Acta 35 961–6.

[32] Akahori H 1961 Electron Microscopic Study of Growing Mechanism of Aluminium


Anodic Oxide Film J. Electron Microsc. (Tokyo) 10 175–85.

[33] Hoar T and Yahalom J 1963 The initiation of pores in anodic oxide films formed on
aluminum in acid solutions J. Electrochem. Soc. 110 614–21.

151
[34] Jessensky O, Müller F and Gosele U 1998 Self-organized formation of hexagonal pore
arrays in anodic alumina Appl. Phys. Lett. 72 1173–5.

[35] Nielsch K, Choi J, Schwirn K, Wehrspohn R B and Gosele U 2002 Self-ordering


Regimes of Porous Alumina:  The 10 Porosity Rule Nano Lett. 2 677–80.

[36] Lee W, Ji R, Gosele U and Nielsch K 2006 Fast fabrication of long-range ordered
porous alumina membranes by hard anodization Nat. Mater. 5 741–7.

[37] Hwang S-K, Jeong S-H, Hwang H-Y, Lee O-J and Lee K-H 2002 Fabrication of highly
ordered pore array in anodic aluminum oxide Korean J. Chem. Eng. 19 467–73.

[38] Chu S, Wada K, Inoue S, Isogai M, Katsuta Y and Yasumori A 2006 Large-scale
fabrication of ordered nanoporous alumina films with arbitrary pore intervals by
critical-potential anodization J. Electrochem. Soc. 153 B384–91.

[39] Kashi M A, Ramazani A, Rahmandoust M and Noormohammadi M 2007 The effect of


pH and composition of sulfuric–oxalic acid mixture on the self-ordering configuration
of high porosity alumina nanohole arrays J. Phys. Appl. Phys. 40 4625.

[40] Ono S, Saito M, Ishiguro M and Asoh H 2004 Controlling factor of self-ordering of
anodic porous alumina J. Electrochem. Soc. 151 B473–8.

[41] Sulka G and Parkola K 2007 Temperature influence on well-ordered nanopore


structures grown by anodization of aluminium in sulphuric acid Electrochimica Acta
52 1880–8.

[42] Sulka G D and Parkola K G 2006 Anodising potential influence on well-ordered


nanostructures formed by anodisation of aluminium in sulphuric acid Thin Solid
Films 515 338–45.

[43] Paolini G, Masoero M, Sacchi F and Paganelli M 1965 An investigation of porous


anodic oxide films on aluminum by comparative adsorption, gravimetric and
electronoptical measurements J. Electrochem. Soc. 112 32–8.

152
[44] Li A P, Muller F, Birner A, Nielsch K and Gosele U 1998 Hexagonal pore arrays with a
50–420 nm interpore distance formed by self-organization in anodic alumina J. Appl.
Phys. 84 6023–6.

[45] Ebihara K, Takahashi H and Nagayama M 1983 Structure and density of anodic oxide
films formed on aluminium in oxalic acid solutions J Met Finish Soc Jpn 34 548–53.

[46] Masuda H, Nishio K and Baba N 1993 Fabrication of a one-dimensional microhole


array by anodic oxidation of aluminum Appl. Phys. Lett. 63 3155–7.

[47] Masuda H and Fukuda K 1995 Ordered metal nanohole arrays made by a two-step
replication of honeycomb structures of anodic alumina Science 268 1466–8.

[48] Dai Z, Gole J, Stout J and Wang Z 2002 Tin oxide nanowires, nanoribbons, and
nanotubes J. Phys. Chem. B 106 1274–9.

[49] Pan Z W, Dai Z R, Ma C, Wang Z L, Dai Z, Pan Z and Wang Z 2002 Gallium oxide
nanoribbons and nanosheets J. Am. Chem. Soc. 106 902–4.

[50] Liu H, Zhao Q, Li Y, Liu Y, Lu F, Zhuang J, Wang S, Jiang L, Zhu D, Yu D and others
2005 Field emission properties of large-area nanowires of organic charge-transfer
complexes J. Am. Chem. Soc. 127 1120–1.

[51] Pan Z W, Dai Z R, Ma C and Wang Z L 2002 Molten gallium as a catalyst for the large-
scale growth of highly aligned silica nanowires J. Am. Chem. Soc. 124 1817–22.

[52] Dai Z R, Pan Z W and Wang Z L 2002 Growth and structure evolution of novel tin
oxide diskettes J. Am. Chem. Soc. 124 8673–80.

[53] Mathur S, Barth S, Shen H, Pyun J-C and Werner U 2005 Size-Dependent
Photoconductance in SnO2 Nanowires Small 1 713–7.

[54] Masuda H, Yotsuya M and Ishida M 1998 Spatially selective metal deposition into a
hole-array structure of anodic porous alumina using a microelectrode Jpn. J. Appl.
Phys. 37 L1090.

[55] Martin C R 1996 Membrane-based synthesis of nanomaterials Chem. Mater. 8 1739–


46.

153
[56] Rabin O, Herz P R, Lin Y-M, Akinwande A I, Cronin S B and Dresselhaus M S 2003
Formation of thick porous anodic alumina films and nanowire arrays on silicon
wafers and glass Adv. Funct. Mater. 13 631–8.

[57] Tian M, Xu S, Wang J, Kumar N, Wertz E, Li Q, Campbell P M, Chan M H and Mallouk T


E 2005 Penetrating the oxide barrier in situ and separating freestanding porous
anodic alumina films in one step Nano Lett. 5 697–703.

[58] Zhao X, Seo S-K, Lee UJ and Lee KH 2007 Controlled electrochemical dissolution of
anodic aluminum oxide for preparation of open-through pore structures J.
Electrochem. Soc. 154 C553–7.

[59] Lee W and Park S-J 2014 Porous anodic aluminum oxide: anodization and templated
synthesis of functional nanostructures Chem. Rev. 114 7487–556.

[60] Sauer G, Brehm G, Schneider S, Nielsch K, Wehrspohn R B, Choi J, Hofmeister H and


Gosele U 2002 Highly ordered monocrystalline silver nanowire arrays J. Appl. Phys.
91 3243–7.

[61] Al Mawlawi D, Coombs N and Moskovits M 1991 Magnetic properties of Fe deposited


into anodic aluminum oxide pores as a function of particle size J. Appl. Phys. 70 4421–
5.

[62] Li F, Metzger R M and Doyle W 1997 Influence of particle size on the magnetic
viscosity and activation volume of α-Fe nanowires in alumite films Magn. IEEE Trans.
On 33 3715–7.

[63] Routkevitch D, Bigioni T, Moskovits M and Xu J M 1996 Electrochemical Fabrication


of CdS Nanowire Arrays in Porous Anodic Aluminum Oxide Templates J. Phys. Chem.
100 14037–47.

[64] Routkevitch D, Chan J, Xu J and Moskovits M 1997 Porous anodic alumina templates
for advanced nanofabrication Proceedings of the International Symposium on Pits and
Pores: Formation, Properties, and Significance for Advanced Luminescent Materials vol
97 pp 350–7.

154
[65] Le Coz F, Arurault L and Datas L 2010 Chemical analysis of a single basic cell of
porous anodic aluminium oxide templates Mater. Charact. 61 283–8.

[66] Masuda H, Hasegwa F and Ono S 1997 Self‐Ordering of Cell Arrangement of Anodic
Porous Alumina Formed in Sulfuric Acid Solution J. Electrochem. Soc. 144 L127–30.

[67] Du G, Chen Q, Che R, Yuan Z and Peng L-M 2001 Preparation and structure analysis
of titanium oxide nanotubes Appl. Phys. Lett. 79 3702–4.

[68] Wang M-H and Hebert K R 1996 An electrical model for the cathodically charged
Aluminum electrode J. Electrochem. Soc. 143 2827–34.

[69] Huang Y, Shih H, Huang H, Daugherty J, Wu S, Ramanathan S, Chang C and Mansfeld F


2008 Evaluation of the corrosion resistance of anodized aluminum 6061 using
electrochemical impedance spectroscopy (EIS) Corros. Sci. 50 3569–75.

[70] Sulka G, Moshchalkov V, Borghs G and Celis JP 2007 Electrochemical impedance


spectroscopic study of barrier layer thinning in nanostructured aluminium J. Appl.
Electrochem. 37 789–97.

[71] Hitzig J, Juttner K, Lorenz W and Paatsch W 1986 AC-Impedance Measurements on


Corroded Porous Aluminum Oxide Films J. Electrochem. Soc. 133 887–92.

[72] O’Neill H 1934 The hardness of metals and its measurement (Chapman & Hall, ltd.).

[73] Tabor D 1951 The hardness of metals vol 10 (ClarendonP).

[74] Pethicai J, Hutchings R and Oliver W 1983 Hardness measurement at penetration


depths as small as 20 nm Philos. Mag. A 48 593–606.

[75] Loubet J, Georges J, Marchesini O and Meille G 1984 Vickers indentation curves of
magnesium oxide (MgO) J. Tribol. 106 43–8.

[76] Stone D, LaFontaine W, Alexopoulos P, Wu T-W and Li C-Y 1988 An investigation of


hardness and adhesion of sputter-deposited aluminum on silicon by utilizing a
continuous indentation test J. Mater. Res. 3 141–7.

155
[77] Oliver W c. and Pharr G m. 1992 An improved technique for determining hardness
and elastic modulus using load and displacement sensing indentation experiments J.
Mater. Res. 7 1564–83.

[78] Alcala G, Mato S, Skeldon P, Thompson G E, Mann A B, Habazaki H and Shimizu K


2003 Mechanical properties of barrier-type anodic alumina films using
nanoindentation Surf. Coat. Technol. 173 293–8.

[79] Alcala G, Skeldon P, Thompson G E, Mann A B, Habazaki H and Shimizu K 2002


Mechanical properties of amorphous anodic alumina and tantala films using
nanoindentation Nanotechnology 13 451.

[80] Xia Z, Riester L, Sheldon B W, Curtin W A, Liang J, Yin A and Xu J M 2004 Mechanical
properties of highly ordered nanoporous anodic alumina membranes Rev Adv Mater
Sci 6 131–9.

[81] Xia Z, Riester L, Curtin W A, Li H, Sheldon B W, Liang J, Chang B and Xu J M 2004


Direct observation of toughening mechanisms in carbon nanotube ceramic matrix
composites Acta Mater. 52 931–44.

[82] Ng K Y, Lin Y and Ngan A H W 2009 Deformation of anodic aluminum oxide nano-
honeycombs during nanoindentation Acta Mater. 57 2710–20.

[83] Ng K Y and Ngan A H W 2012 Effects of pore-channel ordering on the mechanical


properties of anodic aluminum oxide nano-honeycombs Scr. Mater. 66 439–42.

[84] Bobji M S, Pendyala P, Gupta P and Kalode P 2011 Effect of porosity on the
indentation behaviour of nanoporous alumina films Int. J. Surf. Sci. Eng. 5 51–62.

[85] Holmberg K and Mathews A 1994 Coatings tribology: a concept, critical aspects and
future directions Thin Solid Films 253 173–8.

[86] Holmberg K, Ronkainen H and Matthews A 2000 Tribology of thin coatings Ceram.
Int. 26 787–95.

[87] Hamilton D, Walowit J and Allen C 1966 A theory of lubrication by


microirregularities J. Fluids Eng. 88 177–85.

156
[88] Etsion I 2005 State of the art in laser surface texturing J. Tribol. 127 248–53.

[89] Achanta S, Drees D and Celis JP 2005 Friction and nanowear of hard coatings in
reciprocating sliding at milli-Newton loads Wear 259 719–29.

[90] Yu H, Deng H, Huang W and Wang X 2011 The effect of dimple shapes on friction of
parallel surfaces Proc. Inst. Mech. Eng. Part J J. Eng. Tribol. 225 693–703.

[91] Wang H, Yi H and Wang H 2005 Analysis and self-lubricating treatment of porous
anodic alumina film formed in a compound solution Appl. Surf. Sci. 252 1662–7.

[92] Wang H and Wang H 2007 Fabrication of self-lubricating coating on aluminum and
its frictional behaviour Appl. Surf. Sci. 253 4386–9.

[93] Wang S, Ngan A H W and Ng K Y 2012 Anomalous load dependence in single-asperity


tribological behaviour of anodic aluminium oxide nanohoneycombs Scr. Mater. 67
360–3.

[94] Hu N, Ge S and Fang L 2011 Tribological properties of nano-porous anodic aluminum


oxide template J. Cent. South Univ. Technol. 18 1004.

[95] Williams S R 1942 Hardness and hardness measurements (American Society for
Metals Cleveland, Ohio, USA).

[96] Landolt D 1987 Fundamental aspects of electropolishing Electrochimica Acta 32 1–


11.

[97] Pendyala P, Bobji M and Madras G 2014 Evolution of Surface Roughness During
Electropolishing Tribol. Lett. 1–9.

[98] Raicheva S 1984 The effect of the surface state on the electrochemical behaviour of
copper electrodes Electrochimica Acta 29 1067–73.

[99] Li A-P, Müller F, Birner A, Nielsch K and Gösele U 1999 Fabrication and
Microstructuring of Hexagonally Ordered Two-Dimensional Nanopore Arrays in
Anodic Alumina Adv. Mater. 11 483–7.

[100] Li F, Zhang L and Metzger R M 1998 On the growth of highly ordered pores in
anodized aluminum oxide Chem. Mater. 10 2470–80.
157
[101] Garcia-Vergara S, Skeldon P, Thompson G and Habazaki H 2006 A flow model of
porous anodic film growth on aluminium Electrochimica Acta 52 681–7.

[102] Garcia-Vergara S, Skeldon P, Thompson G and Habazaki H 2007 A tracer


investigation of chromic acid anodizing of aluminium Surf. Interface Anal. 39 860–4.

[103] Wu Z, Richter C and Menon L 2007 A study of anodization process during pore
formation in nanoporous alumina templates J. Electrochem. Soc. 154 E8–12.

[104] Xu W, Chen H, Zheng M, Ding G and Shen W 2006 Optical transmission spectra of
ordered porous alumina membranes with different thicknesses and porosities Opt.
Mater. 28 1160–5.

[105] Sui Y and Saniger J M 2001 Characterization of anodic porous alumina by AFM Mater.
Lett. 48 127–36.

[106] Shaban M, Hamdy H, Shahin F, Park J and Ryu S-W 2010 Uniform and reproducible
barrier layer removal of porous anodic alumina membrane J. Nanosci. Nanotechnol.
10 3380–4.

[107] Choi J, Sauer G, Nielsch K, Wehrspohn R B and Gosele U 2003 Hexagonally arranged
monodisperse silver nanowires with adjustable diameter and high aspect ratio Chem.
Mater. 15 776–9.

[108] El-Sayed M A 2001 Some interesting properties of metals confined in time and
nanometer space of different shapes Acc. Chem. Res. 34 257–64.

[109] El-Sayed M A 2004 Small is different: shape-, size-, and composition-dependent


properties of some colloidal semiconductor nanocrystals Acc. Chem. Res. 37 326–33.

[110] Cheng W, Steinhart M, Gösele U and Wehrspohn R B 2007 Tree-like alumina


nanopores generated in a non-steady-state anodization J. Mater. Chem. 17 3493.

[111] Martin C R 1994 Nanomaterials–a membrane-based synthetic approach (DTIC


Document).

158
[112] Wang J-G, Tian M-L, Kumar N and Mallouk T E 2005 Controllable template synthesis
of superconducting Zn nanowires with different microstructures by electrochemical
deposition Nano Lett. 5 1247–53.

[113] Pena D J, Mbindyo J K, Carado A J, Mallouk T E, Keating C D, Razavi B and Mayer T S


2002 Template growth of photoconductive metal-CdSe-metal nanowires J. Phys.
Chem. B 106 7458–62.

[114] Park S, Lim J-H, Chung S-W and Mirkin C A 2004 Self-assembly of mesoscopic metal-
polymer amphiphiles Science 303 348–51.

[115] Lee W, Ji R, Gosele U and Nielsch K 2006 Fast fabrication of long-range ordered
porous alumina membranes by hard anodization Nat. Mater. 5 741–7.

[116] Zhang S 2010 Organic nanostructured thin film devices and coatings for clean energy
(CRC Press).

[117] Bard A J and Faulkner L R 1980 Electrochemical methods: fundamentals and


applications vol 2 (Wiley New York).

[118] Simmons J G 1963 Electric tunnel effect between dissimilar electrodes separated by
a thin insulating film J. Appl. Phys. 34 2581–90.

[119] Trahey L, Becker C R and Stacy A M 2007 Electrodeposited bismuth telluride


nanowire arrays with uniform growth fronts Nano Lett. 7 2535–9.

[120] Gerein N J and Haber J A 2005 Effect of ac electrodeposition conditions on the


growth of high aspect ratio copper nanowires in porous aluminum oxide templates J.
Phys. Chem. B 109 17372–85.

[121] Archard J F 1953 Contact and Rubbing of Flat Surfaces J. Appl. Phys. 24 981–8.

[122] Stachowiak G and Batchelor A W 2013 Engineering tribology (Butterworth-


Heinemann).

[123] Bhushan B 1998 Handbook of micro/nano tribology (CRC press).

[124] Bhushan B 2010 Modern Tribology Handbook, Two Volume Set (CRC Press).

[125] P. Suh N 1973 The delamination theory of wear Wear 25 111–24.


159
[126] Anon Measuring mechanical properties of coatings : a methodology applied to nano-
particle-filled sol-gel coatings on glass.

[127] Gamonpilas C and Busso E P 2004 On the effect of substrate properties on the
indentation behaviour of coated systems Mater. Sci. Eng. A 380 52–61.

[128] Oliver W C and Pharr G M 2004 Measurement of hardness and elastic modulus by
instrumented indentation: Advances in understanding and refinements to
methodology J. Mater. Res. 19 3–20.

[129] Kan Q, Yan W, Kang G and Sun Q 2013 Oliver–Pharr indentation method in
determining elastic moduli of shape memory alloys—A phase transformable material
J. Mech. Phys. Solids 61 2015–33.

[130] Yan W, Pun C L and Simon G P 2012 Conditions of applying Oliver–Pharr method to
the nanoindentation of particles in composites Compos. Sci. Technol. 72 1147–52.

[131] Choi D, Lee S, Lee C, Lee P, Lee J, Lee K, Park H and Hwang W 2007 Dependence of
the mechanical properties of nanohoneycomb structures on porosity J.
Micromechanics Microengineering 17 501.

[132] Masuda H, Ohya M, Nishio K, Asoh H, Nakao M, Nohtomi M, Yokoo A and Tamamura
T 2000 Photonic band gap in anodic porous alumina with extremely high aspect ratio
formed in phosphoric acid solution Jpn. J. Appl. Phys. 39 L1039.

[133] Masuda H, Ohya M, Asoh H and Nishio K 2001 Photonic band gap in naturally
occurring ordered anodic porous alumina Jpn. J. Appl. Phys. 40 L1217.

[134] Xu W L, Zheng M J, Wu S and Shen W Z 2004 Effects of high-temperature annealing


on structural and optical properties of highly ordered porous alumina membranes
Appl. Phys. Lett. 85 4364–6.

[135] Nahar R K, Khanna V K and Khokle W S 1984 On the origin of the humidity-sensitive
electrical properties of porous aluminium oxide (sensor application) J. Phys. Appl.
Phys. 17 2087.

160
[136] Alcala G, Skeldon P, Thompson G E, Mann A B, Habazaki H and Shimizu K 2002
Mechanical properties of amorphous anodic alumina and tantala films using
nanoindentation Nanotechnology 13 451.

[137] Chen S, Liu L and Wang T 2004 Size dependent nanoindentation of a soft film on a
hard substrate Acta Mater. 52 1089–95.

[138] Chen S H, Liu L and Wang T C 2007 Small scale, grain size and substrate effects in
nano-indentation experiment of film–substrate systems Int. J. Solids Struct. 44 4492–
504.

[139] Bahr D F, Woodcock C L, Pang M, Weaver K D and Moody N R 2003 Indentation


induced film fracture in hard film – soft substrate systems Int. J. Fract. 119 339–49.

[140] Pelegri A A and Huang X 2008 Nanoindentation on soft film/hard substrate and hard
film/soft substrate material systems with finite element analysis Compos. Sci.
Technol. 68 147–55.

[141] Saha R and Nix W D 2002 Effects of the substrate on the determination of thin film
mechanical properties by nanoindentation Acta Mater. 50 23–38.

[142] D. Nix W 1997 Elastic and plastic properties of thin films on substrates:
nanoindentation techniques Mater. Sci. Eng. A 234–236 37–44.

[143] Gelli N V R V, Bobji M S and Mohan S Effect of contact stresses on shape recovery of
NiTiCu thin films Thin Solid Films.

[144] Schey J A and Nautiyal P C 1991 Effects of surface roughness on friction and metal
transfer in lubricated sliding of aluminium alloys against steel surfaces Wear 146
37–51.

[145] Musil J, Zeman P, Hruby H and Mayrhofer P 1999 ZrN/Cu nanocomposite film—a
novel superhard material Surf. Coat. Technol. 120 179–83.

[146] Musil J 2000 Hard and superhard nanocomposite coatings Surf. Coat. Technol. 125
322–30.

161

You might also like