Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 9

Growth of Passive Films on Valve Metals and Their Alloys

H. Habazaki, in Encyclopedia of Interfacial Chemistry, 2018


Introduction
Thin oxide films play a key role in the corrosion resistance of metal
and alloys. To understand the growth kinetics of the oxide films in
aqueous solutions, we need detailed knowledge of the interfacial
process and the ionic transport process within the oxide films. In
general, the formation of the oxide films on metals and alloys
involves the ionic transport in oxide under the high electric field, the
dissolution of the oxide film at the film/solution interface, and
sometimes the precipitation of dissolving species from the solution
and/or oxygen gas generation at high potentials.
Several models have been proposed for the growth of oxide films
since the 1930s. In the high field model,1–3 the growth kinetics is
described using a following equation,
i=i0expβE
in which i is the oxide growth current, i0 and β are material-
dependent constants, and E is the electric field strength on the
oxide film. The model is based on the hopping mechanism of ions;
ions at regular or interstitial positions jump to neighbor vacancies or
other interstitial positions with an activation energy, which increases
with the jump distance. A point defect model (PDM) was proposed
by Macdonald et al.4–6 The model assumes the transport of anionic
vacancies across the oxide film to the oxide/solution interface, which
is responsible for oxide growth, while the diffusion of cations and/or
cation vacancies results only in the metal dissolution. The interfacial
potential drops are taken into consideration in the PDM, although
they are neglected in the high field model. Recently, a more general
model that can be extended to alloy substrate under the nonsteady
state is reported by Marcus et al.7
Valve metals, such as Al, Bi, Hf, Nb, Ta, W, and Zr, can form
compact, so-called barrier-type oxide films at high current
efficiencies in selected electrolytes. Once relatively thick oxide films
are formed at high potentials, the anodic current is low at potentials
lower than the formation potential, although high cathodic current for
hydrogen evolution flows. The film structure is dependent upon the
metal, formation potential, formation electrolyte, and surface
pretreatment of metals and alloys. The oxide films formed by anodic
polarization of metals are not pure oxide of substrate because of
incorporation of electrolyte anions species in the oxide film. The
incorporated electrolyte-derived species also influence the film
properties. Here, the formation of oxide films on aluminum and its
alloy, titanium and its alloy, zirconium and its alloys, and magnesium
and its alloys is briefly described.

Sol–gel precursor inks and films


Chang Liu, ... Lei Liao, in Solution Processed Metal Oxide Thin
Films for Electronic Applications, 2020
Abstract
Metal oxide films have important applications in various fields,
especially in the semiconductor industry. There are many ways to
prepare metal oxide films, such as atomic layer deposition (ALD),
physical vapor deposition (PVD), chemical vapor deposition (CVD),
and so on. However, the most common method to fabricate metal
oxide film is sol-gel progress because it offers the easy fabrication
of films or coatings of complex oxides and easy control of
composition and microstructure of deposited films. In this chapter,
the basic knowledge and development history of sol-gel progress
will be introduced by some simple diagrams. Moreover, some typical
examples of sol-gel preparation of thin films will be exhibited to help
readers understand, including Al2O3, HfO2, ZnO, and TiO2. In
addition, other precursors-based metal oxides inks and films will
also be briefly introduced in this chapter.
Application and Uses of Graphene Oxide and Reduced
Graphene Oxide
Sekhar C. Ray, in Applications of Graphene and Graphene-Oxide Based Nanomaterials, 2015
2.5.6 GO/rGO as Coating Technology
Multilayer GO films are optically transparent and impermeable under dry conditions. Exposed
to water (or water vapor), they allow passage of molecules smaller than a certain size. The
films consist of millions of randomly stacked flakes, leaving nano-sized capillaries between
them. Closing these nanocapillaries using chemical reduction with hydro-iodic acid creates
rGO films that are completely impermeable to gases, liquids, or strong chemicals >100 nm
thick. Glassware or copper plates covered with such a graphene “paint” can be used as
containers for corrosive acids. Graphene-coated plastic films could be used in medical
packaging to improve shelf life.
View chapterPurchase book

The oxide film and passivity of aluminium


Christian Vargel, in Corrosion of Aluminium (Second Edition), 2020
3.6 Low-temperature oxide film growth
The oxide film formed at high temperature in the liquid state or at
lower temperature in the solid state during metal transformation
operations has a natural tendency to react very slowly with its
environment. The barrier layer, formed at high temperature, is not
affected by the “aging” of the amorphous layer in contact with the
environment.
The oxide film (or rather its amorphous layer) is not inert with
respect to the surrounding environment [55]. It grows by adsorption,
in particular water vapour [56], and is transformed and incorporates
ions present in the environment6 [57].
3.6.1 Growth of the oxide film in air
Vernon [58] was probably the first, in 1927, to study the growth of
oxide film in an industrial atmosphere at room temperature. He
found that the oxide film reached a thickness of 10 nm after 10–
14 days of exposure regardless of the alloy and nature of the
atmosphere.
According to Hart [59], in dry oxygen at 20°C, at a pressure of
0.1 MPa, the growth of the oxide film on polished aluminium is
initially rapid and then slows down to reach a thickness of 3 nm after
several days. The presence of water vapour (80% RH) increases
the growth rate and the final thickness is greater: 4 nm after 7 days.
The growth law is logarithmic initially and then becomes inverse-
logarithmic. In both cases, these were amorphous films.
By simulating the oxidation of three aluminium crystals with low-
index surfaces (100), (110) and (111), Hasnaoui [60] found that the
growth mechanisms of the amorphous oxide film are independent of
crystallographic orientation. The tests on AA3003 sheets exposed
for several years in a natural atmosphere at room temperature
showed that the oxide film grows more because of the relative high
humidity (Fig. B.3.12). It can reach several tens of nanometres after
4 years of exposure.

Oxidation processing of electronic materials


Z. Li, W. Gao, in Developments in High Temperature Corrosion and
Protection of Materials, 2008
Preparation of zinc oxide films
ZnO films can be prepared by many techniques, including chemical
vapour deposition (CVD) [135], electron beam evaporation [136],
molecular beam epitaxy (MBE) [137], pulsed laser deposition (PLD)
[138], sol-gel [139], spray pyrolysis [140], sputtering [141] and
chemical bath deposition [142]. Thermal oxidation of metallic Zn
[143–149], ZnS or ZnSe [150–153] films has been used to prepare
ZnO films. The Zn precursor films were deposited onto the
substrates (glass, sapphire or Si) by magnetron sputtering, thermal/
electron-beam evaporation or filtered cathodic vacuum arc
technique. Oxidation treatment was normally conducted in air or
oxygen atmospheres in a wide temperature range typically from 300
to 1000°C.
Cho et al. [143] prepared ZnO films by oxidation of Zn films in an
oxygen atmosphere and found that the particle size increased with
increasing oxidation temperature, resulting in sharpened and
enhanced X-ray diffraction peaks for ZnO. These samples have
excellent emission properties; and the room-temperature
photoluminescence (PL) spectra only show a strong UV peak
around 383–390 nm. The oxidation temperature had a significant
influence on the PL properties. As the temperature increased, the
PL peak intensity increased and the peaks became sharper. The full
width at half maximum (FWHM) of the PL peak was 107 and
23 meV for the samples grown at 700 and 1000°C, respectively,
indicating a better optical quality when using a higher oxidation
temperature. Similar phenomena had been observed by Chen [144]
and Alivov [147]. In Cho and Chen’s studies, the visible light
emissions were very weak due to the low defect concentrations in
these films.
Wang et al. [145], however, observed a different trend of PL vs.
temperature in the ZnO films thermally grown in ambient air at
temperatures ranging from 320 to 1000°C for 1 h. The UV peak of
the sample oxidized at 410°C has the strongest PL intensity and the
narrowest FWHM of the temperatures investigated, and the intensity
ratios of UV emission to deep-level emission are 13 and 200 for the
samples prepared at 700 and 410°C, respectively. When the
oxidation temperature was higher than 800°C, the PL intensity
increased again. Wang et al. also studied the correlations between
the deep-level emissions and the oxidation temperature and found
that (i) the intensity of the broad visible band increases with
increasing oxidation temperature; (ii) the visible band can be fitted
well by two peaks located in the green and yellow energy range;
and (iii) the position of the green peak depends strongly on the
oxidation temperature. It red-shifts with an increase of temperature:
a shift of 0.3 eV is observed for the sample grown at 1000°C
compared to the 410°C oxidized sample. The yellow peak displays
a weak temperature dependence. The intensity ratio of the green to
yellow peak increases with the oxidation temperature, especially at
temperatures higher than 900°C. Since the melting point of Zn is
419.5°C, the authors suggested that at lower temperatures
(≤419°C) Zn stays in a solid state and can be oxidized at relative
equilibrium conditions. When the temperature is higher than its
melting point, oxidation is carried out at an unstable liquid state.
More defects that are responsible for the non-radiative transition will
be introduced into the films. However, a much higher oxidation
temperature (≥900°C) facilitates the migration of grain boundaries
and promotes the coalescence of small crystals, thus decreasing
the concentration of nonradiative recombination centres. The
complex relations between the formation and annihilation of defects
and oxidation temperature lead to the complicated emission
behaviours observed in thermally grown ZnO films.
In addition to the PL property characterizations, Alivov et al. [147]
also studied the dependence of electrical properties on oxidation
temperature and oxygen partial pressure. From Hall-effect
measurements, they found that the resistivity of ZnO films increases
with oxidation temperature from 381 Ω-cm at 400°C to 6.02 × 104 Ω-
cm at 900°C. At the same time, the carrier (electron) concentration
decreases from 4.56 × 1015 to 7.32 × 1012 cm−3, and mobility
increases from 3.6 to 14.2 cm2/V-s, providing evidence of
improvement of crystal quality of ZnO films with temperature as the
ZnO crystal lattice becomes more stoichiometric with high
temperature annealing in oxygen. Alivov et al. also investigated the
variation of the electrical properties as a function of the oxygen
partial pressure. Resistivity and mobility increase by a factor of 2–4
when oxygen partial pressure is varied from 1.01 × 105 to
1.33 × 10−3 Pa, indicating an improvement of film crystallinity. The
authors believed that decreasing oxygen partial pressure can slow
down the oxidation process of Zn and thus the best crystal quality
can be expected at the lowest crystal formation rates.
As mentioned above, ZnO is a promising material for development
of LEDs due to its high exciton binding energy (~60 meV). For these
applications, however, p-type (hole) conduction must be realized.
Though highly conductive n-type (electron) ZnO films have been
created, progress in fabrication of p-type ZnO with low resistivity is
very slow. Doping is the way to realize p-type conduction, and N is
one of the candidates for p-type doping. Generally, doping of
nitrogen in ZnO could be achieved by sputtering [154,155], CVD
[156,157], PLD [158], MBE [159], implantation [160], and/or solution
approaches [161]. However, the chemical activity of O is higher than
that of N; Zn preferentially combines with O rather than with N. As a
result, it is difficult for N to be incorporated into ZnO films. Recently,
it has been found that oxidation treatment of Zn-N (Zn3N2) can
directly incorporate N atoms into ZnO during the growth process to
occupy oxygen positions, to partly compensate some donors (e.g.
oxygen vacancy) and then to yield p-type conductivity during the
course of thermal oxidation [162–165]. Nakano et al. [165] reported
a significant improvement of p-type doping characteristics by
oxidation at temperatures between 500 and 800°C, where more N
acceptors are activated and the oxidation state is enhanced.
However, other studies indicated that oxidation treatment at high
temperatures (>500 or 600°C) might initialize a conduction-type
transformation from p to n [162,163]. These different observations
can be understood if one can see that the N-containing precursor
films were prepared using different methods under different
conditions.

Preparation of graphene
Kazuyuki Takai, ... Michio Inagaki, in Graphene, 2020
2.3.7 Fabrication of transparent reduced graphene oxide films
Transparent rGO films were prepared by dip-coating a hydrophilic
substrate such as SiO2 into an aqueous GO dispersion at 70°C,
followed by thermal reduction at 550–1100°C under a flow of
Ar [343]. The thickness of the resultant rGO film was tuned by
changing the temperature of the GO dispersion as well as the
dipping repetition. The electrical conductivity was about 550 S/cm
for the 10.1-nm-thick film. It increased to 727 S/cm with an increase
in the film thickness to 29.9 nm. The TEM image
in Fig. 2.121A shows stacking of the rGO nanoflakes; the SAED
pattern (inset) gave a 100 ring together with strong 002 spots, which
suggested the existence of the scrolling or folding of rGO flakes.
Transmittance of the rGO film was 70.7% at a wavelength of
1000 nm, lower than 82.4% for fluorinated tin oxide (FTO) film and
90% for indium tin oxide (ITO) film, although it could be improved to
80.0% by decreasing the film’s thickness. In a high-wavelength
region above 1500 nm, however, the transmittance of rGO films was
kept almost constant, although FTO and ITO showed strong
absorption, as displayed in Fig. 2.121B. Conductive and transparent
films with a thickness of 14–86 nm were prepared from GO by
exfoliation under sonication and reduction with hydrazine in a water
dispersion. Then the films were made on a quartz surface and
annealed up to 1100°C [344], as shown in Fig. 2.122. Transmittance
of the film with 14 nm thickness was well over 80% in the
wavelength range of 1100–3000 nm, and its electrical conductivity
was over 200 S/cm.

Sign in to download full-size image


Figure 2.121. Transparent reduced graphene oxide (rGO) film [343]. (A)

Transmission electron microscopy image with selected area electron diffraction

pattern and (B) transmission spectrum of rGO film, compared with those of

indium tin oxide (ITO) and fluorinated tin indium oxide (FTO).

Sign in to download full-size image


Figure 2.122. Thin sheets prepared at different thicknesses from graphite oxide

through exfoliation, reduction, and annealing at 1100°C [344].

For the anode of solar cells, transparent conductive rGO films were
prepared from GO by different exfoliation and reduction
processes [288,345,346]. Large-area, transparent, and highly
conductive, few-layered graphene films were synthesized via CVD
on an Ni-coated SiO2/Si substrate to apply to the anode of solar
cells [347].
Transparent rGO films were prepared from Nafion-functionalized
rGO nanoflakes [325]. The dispersion was filtered using Millipore
ester membranes until a thin film was formed on the membrane,
followed by drying in air at 90°C, immersion in o-DCB and acetone,
washing with acetone, and drying at 150°C to obtain transparent
rGO films. In Fig. 2.112B, sheet resistance is plotted against
transmittance for Nafion–rGO composite films prepared from the
mixture of Nafion/GO at ratios of 1:1 and 5:1. The latter film gave a
sheet resistance of about 30 × 103 Ω/sq at a transmittance of 80%.
The low resistance of these rGO films was thought to be caused by
the presence of residual electroconductive Nafion, because washing
the film by a large amount of water–ethanol mixture just after the
film formed on the filter made the sheet resistance increase.
Transparent carbon films (graphene films) were successfully
synthesized by the heat treatment of giant polycyclic aromatic
hydrocarbons (hexadodecyl-substituted superphenalene, C96–C12)
(Fig. 2.123A) at 1100°C in Ar on a quartz substrate. Their
photographs and transmission spectra are shown in Fig. 2.123B and
C, respectively [348]. The transmittance of the films with
thicknesses of 4, 12, 22, and 30 nm was 90%, 80%, 66%, and 55%,
respectively, and the electrical conductivity of the film 30 nm thick
was measured as 206 S/cm with a sheet resistance as 1.6 kΩ/sq,
both of which changed slightly with a decrease in film thickness.

You might also like