Вы находитесь на странице: 1из 15

Journal of Non-Crystalline Solids 521 (2019) 119493

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


journal homepage: www.elsevier.com/locate/jnoncrysol

Review

Advanced applications of amorphous alumina: From nano to bulk T


a,b,⁎ a,b,⁎⁎ a a
Andraž Mavrič , Matjaz Valant , Chunhua Cui , Zhiming M. Wang
a
Institute of Fundamental and Frontier Sciences, University of Electronic Science and Technology of China, Chengdu 610054, China
b
University of Nova Gorica, Materials Research Laboratory, Vipavska 13, SI-5000 Nova Gorica, Slovenia

ARTICLE INFO ABSTRACT

Keywords: In modern technologies, the researchers are keen on top-down approaches to tailor the materials from bulk into
Amorphous alumina their nanometer forms. For the amorphous alumina, this is the other way around. The amorphous alumina was
Strain engineering receiving attention in the forms of thin films or nanoparticles many years ago, yet any attempts to increase the
Short-range structure dimensions destabilized its amorphous structure and caused crystallization at even relatively modest tempera-
Wear and scratch resistance
tures. This issue has been addressed by smart engineering of its lattice strain, which allows the formation of a
Dielectric
high-temperature stable bulk alumina. In this review, we present the recently developed synthetic methods and
Photoluminescence
Catalysis provide insights in sustaining the amorphous structure of widely used alumina. We highlight its adoptable short-
range structure and morphology, emphasize the corresponding tuning of the mechanical and electronic prop-
erties, and illuminate the advanced applications in wear and scratch resistant coatings, corrosion protection,
insulator layers, capacitor, luminescent material, and catalysis.

1. Introduction transistors [31,32], and hydrophilicity for corrosion protection


[19,33–35]. Highly absorptive nanoparticles have been used for en-
Alumina in all its forms is technologically important material. It can hancement of thermal conductivity of nanofluids [5,18], as a catalyst
exist in many crystallographic modifications with different properties. for biodiesel conversion [36] and for fluoride removal in water treat-
The most frequently occurring form in nature is α-Al2O3 (corundum); it ment processes [37].
is a thermodynamically stable phase at room temperature. Besides,
Al2O3 can be found in many metastable polymorphs. There is a wide 2. Thermodynamic stability of amorphous alumina
choice of reviews dedicated to alumina polymorphs summarizing their
structure [1], catalysis [2–4] and nanoscale properties [5]. In addition, α-Al2O3 starts to crystallize from basic hydroxide gels at tempera-
there are specialized reviews on anodic aluminum oxide (AAO) [6–8] tures as low as 420 °C [38]. If the metastable alumina polymorphs are
that has become its own branch of research. In this review, we focused present, they transform to the thermodynamically stable α-phase
on amorphous alumina (am-Al2O3), emphasizing its synthesis of nano- through different transition sequences at an elevated annealing tem-
and micro-sized structures, thermodynamic stability, functional prop- perature around 1100 °C [1]. Many efforts have been made to lower the
erties, and advanced applications. transformation temperature since the high-temperature annealing leads
In contrast with a variety of alumina crystalline polymorphs, an to strongly agglomerated α-Al2O3 particles that are hard to be pro-
amorphous form possesses some unique properties, originating from the cessed [39–45]. A reduction in the particle size has been identified as
non-crystalline nature. This structural disorder leads to a decrease in the crucial factor for preventing the metastable alumina against trans-
the band gap energy [9,10], a variation in photoluminescence [11–16], formation to α-Al2O3 [46–49]. It was shown that the γ-phase has lower
a change in ferromagnetism [17], and an increase in thermal con- surface energy than the α-phase, which makes it stable at a specific
ductivity [18]. A special attention has been given to the am-Al2O3 for surface area > 125 m2 g−1 at room temperature and > 75 m2 g−1 at
coating applications. For instance, the uniform amorphous matrix was 500 °C [46]. The effect of the specific surface area is so predominant
reported to exhibit excellent mechanical and optical properties that thermodynamically prevents transitions to the α phase forming
[19–30], high dielectric constant for use as a gate insulator in nanocrystals with dimensions < 12 nm [50]. To produce metastable α-

Correspondence to: A. Mavrič, Institute of Fundamental and Frontier Sciences, University of Electronic Science and Technology of China, No.4, Section 2, North

Jianshe Road, Chengdu 610054, China.


⁎⁎
Correspondence to: M. Valant, University of Nova Gorica, Vipavska 13, SI-5000 Nova Gorica, Slovenia.
E-mail addresses: andraz.mavric@uestc.edu.cn (A. Mavrič), matjaz.valant@ung.si (M. Valant).

https://doi.org/10.1016/j.jnoncrysol.2019.119493
Received 19 January 2019; Received in revised form 3 June 2019; Accepted 4 June 2019
Available online 07 August 2019
0022-3093/ © 2019 Elsevier B.V. All rights reserved.
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

Fig. 1. A schematic diagram of the thermodynamic stability of α-, γ- and am-


Al2O3, depending on the specific surface area. Reconstructed from data in Ref.
[46, 50, 51].

Fig. 2. a) TEM images of am-Al2O3, taken at 200 keV, after different e-beam
Al2O3 nanocrystals (sub-10 nm), a high energy input is necessary (e.g. exposure time. The inset at 13 min reveals a crystal lattice embedded in the
by high energy milling [50]). amorphous surrounding. b) Electron diffraction patterns after different ex-
Tavakoli et al. [51] compared the excess enthalpy as a function of posure time. The diffraction rings correspond to 311 and 422 planes of γ- Al2O3.
the specific surface area of the am-, γ- and α-Al2O3 nanoparticles with Reprinted from [55], with permission from Elsevier.
that of the bulk corundum. The excess enthalpy is a linear function of
the specific surface area for the am-, γ and α phase. With the increase of densification was attributed to the loss of oxygen from the peroxy ion.
the specific surface area, the enthalpy increases slowly for the amor- The crystallization temperature of the alumina thin films deposited
phous phase than for the γ and α phases, due to the smaller surface on a mesoporous silica gel was measured with Differential Scanning
energy of am-Al2O3 (Fig. 1). When the specific surface area exceeds Calorimetry (DSC). It was found to be strongly thickness dependent and
370 m2 g−1 (corresponding to a particle diameter of ~6.5 nm), am- shifted from 1020 °C to 998 °C for the 2.7 nm and 13.9 nm films, re-
Al2O3 becomes the most energetically stable phase. The same trends of spectively. When compared to HfO2 this dependence was much weaker
lowering the transition enthalpy with the increase in the specific sur- [52]. Am-Al2O3, a 2–6 nm thick film within a yttrium-stabilized zir-
face area were shown for all the transitions in the system, i.e. the am- to conia multilayer, was reported to be stable up to 1000 °C [57]. In the
α, am- to γ and γ to α transitions. In comparison with the surface en- case of an interface between a metal and aluminum oxide, the interface
thalpy, a contribution of interfacial energy to the excess enthalpy energy of the am-Al2O3 is lower than that of the crystalline state [58].
arising from grain-grain interfaces was found to be negligible. Beside For this reason, the am-Al2O3 thin film that forms during oxidation of
the enthalpy part of the free energy, the configurational entropy, as a the metal aluminum remains amorphous up to 5–10 nm thickness, de-
result of the disorder in am-Al2O3, should also be accounted for. Be- pending on the crystallographic face of the aluminum and growth
cause α-Al2O3 is an ordered structure, it has the configurational entropy temperature [58–61].
equal to zero [51]. On the other hand, both γ- and am-Al2O3 structures The main drawback of amorphous alumina film with thickness >
are to some extent disordered. This results in an increase in the con- 100 nm is its low thermodynamic stability as its unique properties fade
figurational entropy. Being amorphous, am-Al2O3 is the most dis- upon crystallization when using it for coatings or electronic devices.
ordered. As a result, the positive entropy change via the α- to γ- tran- Some thermal treatment of the films is necessary to achieve cohesion
sition is smaller relative to the α- to am-transition. with the substrate or to dehydrate the am-Al2O3 films. This often results
Bloch et al. [52] studied the effect of the am-Al2O3 film thickness on in the full or partial crystallization due to the low crystallization onset
structural order and crystallization induced by an electron-beam on temperature of ~450 °C [62]. Preparing am-Al2O3 that can be thermally
TEM. It was found that for a 4.16 nm thick film there is no change in the treated at higher temperatures without being crystallized requires im-
structure after 20 min of exposure to the 300 keV electron beam. provement of its thermodynamic stability. We have seen before that
Around twice-thicker film (7.39 nm) after exposure showed the electron keeping the dimensions on the nanoscale and the surface area high (for
energy loss spectra resembling γ-Al2O3, without crystallization (e.g. nanoparticles and thin films) the crystallization can be successfully
long-range order). Exposition of a much thicker film (63.1 nm) to the suppressed. In addition, it was shown that the interface formation with
electron beam resulted in crystalline γ-Al2O3 with some α-Al2O3-like metal also suppresses the crystallization onset. In some cases, the
features. The crystallization under the exposure to TEM electron beam structural disorder introduced by doping (e.g. with Y3+ [36,63,64],
was found also by other authors [53–55]. As shown in Fig. 2a, the P5+ [65], Er3+ [64], Eu3+ [16], La3+ [66], Si4+ [67], Fe3+ [68] or
crystallization under the TEM electron beam at 200 keV can be in op- Cr3+ [64,68], Zn2+ [69]) suppressed the crystallization. We demon-
erando observed, starting from local densification to nucleation and strated yet another approach for stabilization of the amorphous alu-
growth of the crystalline phase. This can be confirmed by the electron mina by nanocomposite formation [23,55]. We showed that by dis-
diffraction patterns (Fig. 2b), where we can notice the appearance and persing polysilane molecules into the alumina matrix (Fig. 3), the bulk
intensification of the diffraction rings and dots characteristic for γ- am-Al2O3 can be successfully sustained at 900 °C. During the heat
Al2O3 [55]. In addition, densification of the am-Al2O3 structure was treatment, the polysilane molecules are oxidized, form covalent bonds
also evidenced by a synchrotron X-ray radiation during a near-edge X- with the matrix and induce a homogeneously distributed stress
ray absorption fine structure (NEXAFS) experiment [56]. The throughout the bulk [55].

2
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

Fig. 3. Schematic presentation of forms that induce different stress: a) nanoparticles – surface stress; b) thin films – interfacial stress; and c) nanocomposites – bulk
stress. TEM images of am-Al2O3 in form of d) nanoparticles [Reprinted with permission from [51]. Copyright 2013 American Chemical Society.]; e) thin films
[Reprinted with permission from [52]. Copyright 2014 American Chemical Society.]; f) bulk nanocomposite: and g) a combination of nanocomposite and thin film.

The common strategy of all the reported approaches is to introduce 3. Structure


enough stress into the alumina structural matrix to thermodynamically
favor the amorphous phase over crystalline phases. The origin of the The absence of a distinct structural periodicity in am-Al2O3 does not
stress seems to be not important, as the efficient suppression of the preclude the existence of characteristic atomic configurations, in which
crystallization onset has been demonstrated as a result of interfacial, Al3+ ions retain the type of bonding and coordination polyhedra similar
surface or bulk stress (Fig. 3). to the crystalline forms. Because the interatomic forces in the crystals
Fig. 4 shows the correlation between the crystallization temperature and amorphous material are essentially the same, the amorphous can be
and the thin films thickness or nanoparticles diameters prepared by presented as a three-dimensional network of such polyhedra, which
different methods summarized in Table 1. A decrease in the film lack the periodicity and symmetry [70]. The XRD powder diffraction of
thickness or particle size increases the crystallization temperature. This am-Al2O3 does not show the sharp peaks, which are observed in the
is a direct consequence of the increase in the interface or surface en- polycrystalline powders. The diffractograms of the am-Al2O3 solids
ergy, respectively. In addition, the larger sized particles, stabilized by usually contain one or more broad low-intensity halos at the positions
doping or nanocomposite formation, demonstrated high stability. Their corresponding to the strongest peaks of the related crystalline poly-
stability is comparable with the stability of the < 10 nm sized alumina morphs (Fig. 5). The broad peaks of am-Al2O3 reflect the average dis-
particles owing to the improved strain that is homogeneously in- tances between the atoms in the amorphous structure. Their intensity
troduced into the amorphous structural network. slightly increases after annealing at higher temperatures due to the
densification of the material and the increase of the local order.
Besides 4- and 6-coordinated Al3+, [Al]4 and [Al]6, that are found in
the crystalline polymorphs (with α-Al2O3 containing only [Al]6), the
presence of 5-coordinated Al3+, [Al]5, was found in the amorphous
alumina (Fig. 6a). The presence of [Al]5 is believed to be unique for am-
Al2O3 with some minor exceptions. A small amount of [Al]5 was ex-
perimentally found in highly defective surfaces of nanosized metastable
γ- and κ- alumina [73,74] and simulated by theoretical reconstruction
of their surfaces [75]. The experimental observations of the [Al]5 in am-
Al2O3 have also been supported by theoretical calculations. Some the-
oretical calculations also reported a small amount of 3-coordinated
Al3+, [Al]3, which have not yet confirmed experimentally. The reason
might be in the fact that the calculations have been performed on
models of very low-density alumina clusters that cannot be experi-
mentally realized. The distribution of different Al coordination poly-
hedra found by different studies is presented in Fig. 7.
Different spectroscopy techniques have been used to resolve the
structure of amorphous alumina. By modeling the X-ray radial dis-
tribution function the presence of [Al]5 polyhedra in addition to [Al]4
and [Al]6 was shown for AAO films by Oka et al. [76]. They modeled
the system by mixing the structure of γ-Al2O3 and disordered phase. A
structure very similar to γ-Al2O3 was observed also for AAO by an ex-
Fig. 4. Correlation between the size and crystallization temperature of thin tended electron energy-loss fine structure (EXEELFS) [77,78], EXAFS
films and nanoparticles; the data are taken from references shown in the [79], a near-edge X-ray absorption fine structure (NEXAFS) [56], and
brackets in the plot. The dashed lines show the crystallization temperature of XPS [60,80] and for the RF sputtered films by X-ray radial distribution
bulk particles stabilized by the Eu-doping and nanocomposite formation. function [81]. Lamparter and Kneip [82] combined the X-ray and

3
Table 1
Synthesis methods used for the formation of am-Al2O3. Where applicable the thermal treatment conditions resulting in crystalline form are given.
Method Source, conditions Form, size Crystallization Ref.
A. Mavrič, et al.

PLD α-Al2O3 pallet at RT thin films, ~260 nm γ at deposition temperature > 500 °C [29,105]
PLD polycrystalline Al2O3 film ≥3 μm N/A [104]
PLD Al2O3, at RT thin film 20 nm N/A [31]
PLD polycrystalline Al2O3, at RT ~40 nm N/A [127]
PLD Al2O3 sintered target, RT thin film 0.1–1 μm If deposited at substrate temperature 800 °C. [108]
ALD trimethylaluminum/water, 300 °C thin film 12 nm γ after annealing at 1000 °C, 60 s [106]
ALD trimethylaluminum/water, 300 °C thin film 20 nm > 800 °C [141]
ALD trimethylaluminum/H2O, 150 and 175 °C thin film 80-320 nm N/A [30]
ALD trimethylaluminum in Ar/O2/N2, at 200 °C thin film ~1.4 μm N/A [88]
ion beam during physical vapor Al2O3, the substrate at 200 °C 223-1209 nm N/A [107]
deposition
e-beam evaporation alumina disk, annealing at 400 and 600 °C after deposition thin films, 350 nm 600–800 °C [142]
e-beam evaporation Al2O3, annealed at 700 °C from 1 to 12 h thin film, 116-254 nm N/A [97]
CVD aluminum tri-sec-butoxide nanoparticles, ~10 nm amorphous below 1000 °C [113]
CVD from aluminum acetylacetonate, 500–700 °C a film composed of particles (15–30 nm) N/A [24]
PECVD AlCl3, H2, CO2 at T < 500 °C films, ≥3 μm N/A [109,143]
PECVD trimethylaluminum in an oxygen plasma 0.1-1 μm If deposited at substrate temperature 800 °C. [108]
Reactive magnetron sputter Al2O3 target thin films, ~200 nm Onset at 800 °C. [87]
deposition
Reactive magnetron sputter Al target, Ar-O2 gas mixture, 350 °C thin films, ~1 μm depends on the substrate bias voltage and magnetic [28]
deposition field
Reactive magnetron sputter Al target, Ar-O2 gas mixture, substrate temperature < 200 °C thin films, ~1 μm 600 °C or as deposited when substrate temperature [56,102,144]
deposition during deposition > 200 °C
Pulsed rf magnetron sputtering Al/Al2O3 substrate at RT thin films, ~1 μm annealing at 500 °C [26,145]
High power impulse magnetron Al target, Ar-O2 gas mixture thin films, 0.2-1 μm N/A [146]

4
sputtering
Pulsed rf magnetron sputtering Al target, Ar-O2 gas mixture thin films, 1-2 μm depends on the deposition rate and substrate [146]
temperature, generally > 450 °C
RF sputtering Al target, Ar-O2 gas mixture, annealing at 300 °C thin film, ~50 nm N/A [27]
RF magnetron sputtering Al2O3 target, Ar or Ar-O2 gas mixture, the substrate at 20–500 °C thin film, < 1 μm N/A [116]
MOCVD aluminum tri isopropoxide, 420–650 °C films 0.4–1.5 μm 450–650 °C [19,20,33,110,112]
Sol-gel Al(NO3)3/glacial acetic acid/acetylacetone/polyvinyl alcohol, heat films, ~215 nm > 700 °C [25,147]
treated at 400 °C
Sol-gel boehmite (AlOOH) sol, deposited at 250 °C films, 2-3 μm N/A [34]
Sol-gel Al(NO3)3/water/glycerol/citric acid, heat treated 500–900 °C porous micro particles > 600 °C [13,25]
Sol-gel Al(NO3)3/NH3/pH = 7 heat treated at 550 °C nanoparticles, ~5 nm, agglomerated > 600 °C and when prepared in pH > 7 after first [17]
heat treatment at 550 °C
Sol-gel AlCl3/ethylene glycol monomethyl ether, annealing 600–1000 °C nanoparticles, ~50 nm > 700 °C [23]
Sol-gel AlCl3 and polysilane thin films, ~50 nm amorphous at 1500 °C [23]
Sol-gel AlCl3 and polysilane bulk < 1000 °C due to mullite crystallization [55]
Sol-gel AlCl3/Y(NO3)3/polyvinyl alcohol, calcined at 700 °C nanoparticles, BET specific surface area of N/A [36]
312 m2 g−1 (Y3+ doped)
Sol-gel aluminum isopropylate/2-ethoxyethanol/glycol ether/acetic acid, thin films, ~260 nm N/A [132]
spin-coating, thermal treatment at 450 °C
Sol-gel aluminum isopropoxide/water/HNO3 dip-coating, thermal thin films, ~250 nm N/A [134]
treatment at 550 °C
Sol-gel powder flat-Al13/pluronic-F127/water spin-coating, thermal thin films, 60–350 nm N/A [133]
treatment at 500 °C
Spray pyrolysis AlCl3/water/ethanol 250–550 °C Depending on the amount of deposited material, N/A [131]
nanoparticles or thin films are formed.
Coprecipitation AlCl3/EuCl3/ethanol, fast drying at 350 °C, annealed at 500–900 °C particles, 4-10 μm (10% Eu3+) amorphous at 900 °C [16]
(continued on next page)
Journal of Non-Crystalline Solids 521 (2019) 119493
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

[140]

[148]
[37]
[15]
[11]
[12]
Ref.
Crystallization

> 600 °C

N/A
N/A
N/A
N/A
N/A

Fig. 5. XRD diffractograms of α, γ, and am-Al2O3 in nanocomposite form. The


peak positions (blue bars) for α-Al2O3 and γ-Al2O3 are taken from PDF cards
431,484 [71] and 2,015,530 [72], respectively. Each thick on the y-axis present
nanoparticles < 25 nm dispersed in toluene

200 counts, the bars from PDF card are normalized to the peak with the highest
intensity. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
nanoporous membranes 40-13 nm

thin films on Al, 0.05–0.5 μm


nanostructured microspheres

nanoparticles < 4 nm
Form, size

nanotubes
Al(NO3)3/water/sodium dodecyl sulfonate/NH3 to pH~5, 120 °C,

acetic acid/aluminum isopropoxide


heat treated 500–800 °C

Fig. 6. a) Representation of distorted tetrahedral [Al]4, bipyramidal [Al]5 and


Source, conditions

octahedral [Al]6 polyhedra, Al3+ (blue) and oxygen (red). b) Two-dimensional


27
Al 3QMAS NMR spectra of the rotor background subtracted amorphous ALD
thin films distinguishing between three different Al3+ coordinations. Reprinted
Al anode
Al anode
Al anode
Al anode

with permission from [87,88]. Copyright 2010 American Chemical Society.


(For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

neutron diffraction data of AAO with reverse Monte Carlo simulations


that showed the presence of the 3-, 4-, 5- and 6-coordinated polyhedra,
with [Al]4 being the dominant fragment. A much more direct tool for
Table 1 (continued)

analysis of the Al coordination in amorphous alumina is NMR spec-


Hydrothermal

Solvothremal

troscopy [83,84]. The solid-state NMR (SSNMR) with a Magic Angle


Anodization
Anodization
Anodization
Anodization

Spinning (MAS) of 27Al nuclei allows distinguishing 4-, 5- and 6- co-


Method

ordinated Al3+ at the chemical shifts of ~60, ~35 and ~0 ppm, re-
spectively [85,86]. Due to the local atomic environment variations in

5
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

Fig. 7. Distribution of the Al polyhedron coordina-


tion in the am-Al2O3 obtained with various theore-
tical methods and experimental techniques.
Theoretical results are ordered by increasing the
density of modeled unit cell: low density (here
chosen as < 3.1 g cm−3), middle dense distribution
(3.1 g cm−3 ≤ ρ ≥ 3.25 g cm−3) and high density
(> 3.25 g cm−3). Experimental results are ordered
by the temperature of heat treatment. The references
to the original data are labeled on the X-axis. (For
interpretation of the references to colour in this
figure legend, the reader is referred to the web ver-
sion of this article.)

the amorphous phase, the NMR peak broadening is high [84]. The three showing on the higher density of the structure and large degree of the
different Al coordination environments can be better distinguished corner- to edge-sharing structural transition. This is especially true for
from 2D NMR spectra (Fig. 6b) [87,88]. NMR was used to show the the samples that have been exposed to very high temperatures [55,83].
presence of [Al]5 in amorphous aluminum oxide films prepared by The experimental results on the low-density am-Al2O3 state are in
anodization [89], RF sputtering [87], ALD and PVD [88], sol-gel [90]. fairly good agreement with the theoretical studies while for the high-
Gutiérrez and Johansson [91] performed a molecular dynamic (MD) density state certain deviations exist. The reported concentration of
study of amorphous alumina structure. They revealed the presence of a [Al]4 polyhedra is not much lower than predicted by calculations on the
short-range order dominated by two- to five-fold rings of alternating Al- high-density models while the concentration of [Al]5 is significantly
O bonds, resulting in the structure dominated by [Al]4 tetrahedra. They lower. It seems that [Al]5 is energetically the most unfavorable and at
show that the structures of amorphous and liquid alumina are very si- elevated temperatures tends to change to [Al]4 which can further un-
milar and resemble the surface structure of γ-Al2O3 at room tempera- dergo corner-to-edge-sharing transition [55]. With the increase in the
ture. Similar structural characteristics were later obtained by others by share of [Al]6, the structure is becoming more and more similar to the γ-
MD [92–98] and Density Functional Theory (DFT) [84] calculations. Al2O3 until it finally crystallizes.
Oxygen is preferred on the surface of am-Al2O3 and Al is enriched just
below the surface [99].
It was found both theoretically [92,97,100] and experimentally 4. Synthesis
[55,87,97] that with thermal annealing of am-Al2O3 the distribution of
the different coordination polyhedra changes. The coordination number Dehydration of different aluminum hydroxide species above 300 °C
of the aluminum ion increases. The transformation from the corner- gives the polycrystalline polymorphs. Depending on the hydroxide
sharing [Al]4 and [Al]5 polyhedra to more stable edge-sharing [Al]6 structure the material changes through a series of phase transitions and
octahedra occurs with temperature [87,97]. This transformation results converts to α-Al2O3 on heating between 900 and 1100 °C. These tran-
in the densification of the alumina structural matrix. Pores and nano- sitions are of high technological importance and are well studied.
voids are formed during this process [92,97,100]. During further an- During the transitions, aluminum hydroxide precursor undergoes loss of
nealing, the nucleation in the amorphous alumina matrix occurs by water and densification. At the ambient conditions, am-Al2O3 cannot be
short-range reordering of the structure that resembles the γ phase. In- prepared just by dehydrating aluminum hydroxides. Moreover, most of
itially, the crystallization of the amorphous alumina results in γ-Al2O3 the aluminum hydroxide precursors are by themselves crystalline. The
[10,52,68,101]. Further transformation to more stable α-Al2O3 can be phase transformation to the alumina proceeds without degradation of
direct or indirect (over θ- or κ-phases) [102]. If an initial film already crystalline structure and the polytype of the resulting transition alu-
contains some γ-phase nanocrystals, am-Al2O3 transforms to the α mina is in fact determined by the crystalline structure of the precursor
phase via γ phase only, otherwise the transformation proceeds over a [1]. Nevertheless, am-Al2O3 has been prepared from finely powdered
mixture of polymorphs [102]. aluminum trihydroxide by vacuum drying at 425 °C. It has crystallized
Fig. 7 shows the distribution of the Al polyhedron coordination in at heating to 450 °C [103]. Due to problems with transforming alu-
the amorphous alumina obtained with various theoretical methods and minum hydroxide precursors to amorphous alumina the research has
experimental techniques. The theoretical calculations performed on low moved towards the synthesis from vapor phase such as physical vapor
density am-Al2O3 structural models gave a high share of [Al]4 poly- deposition (PVD) and chemical vapor deposition (CVD). Up-scalable
hedra (> 50%) followed by the [Al]5 polyhedra and the [Al]6 poly- techniques, such as sol-gel synthesis, have also been adopted. However,
hedra with < 5%. As expected, by increasing the structural density the the production was mostly oriented to nanoparticles and thin films due
share of the polyhedral with higher coordination increases. The similar to the low-temperature stability of amorphous bulk alumina. The syn-
trend can be observed from the experimentally reported polyhedron thetic methods used for the synthesis of am-Al2O3 are collected in
distributions. As deposited films that tend to have lower density have a Table 1. Am-Al2O3 is prepared under thermodynamically non-equili-
high amount of the [Al]4 polyhedra, exceeding 50%. The samples ex- brium conditions; therefore, its structure largely depends on the
posed to higher temperatures have a higher amount of [Al]6 polyhedra, synthesis conditions.

6
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

4.1. Vapor phase methods orientation, temperature, heating rate and time of the thermal treat-
ment, the amorphous film starts to crystallize to γ- and α-Al2O3
For the preparation of the thin films and powders, PVD and CVD [59,60,80,123,124]. The crystallization temperature was found to lar-
have been frequently used. PVD is carried out in an ultrahigh vacuum gely depend on an oxide layer thickness [59–61,125].
where pure alumina is vaporized by a laser or a radiofrequency source. Similarly, the electrochemical oxidation gives the amorphous phase
The as-deposited alumina thin films or nanoparticles are amorphous [78,79,82] during the first stages of the growth and later a mixture of
[24,26,29,101,104–108]. The PVD films deposited on a heated sub- am-, γ- and α-Al2O3. The crystallization can be induced by increasing
strate or by heating just above 500 °C have been reported to crystallize the oxide layer thickness or by a thermal treatment [8,126]. However,
to polycrystalline γ-Al2O3 or a mixture of the polymorphs. the crystallization temperature does not only depend on the layer
[26,29,101,105,106,108]. thickness but also on the electrolyte that was used during anodic oxi-
The alumina thin films by CVD at 350 °C include aluminum oxide dation. The AAO is during the growth in contact with the electrolyte,
hydroxide (AlOOH) [109–112]. Thermal treatment is necessary to re- causing the electrolyte anions, water, and hydroxyl groups to be in-
move the hydroxide groups. Since CVD results in hydroxylated alumina, corporated in the structure. The crystallization can occur only when
precise temperature control is needed to dehydrate alumina and to these inclusions are thermally removed from the structure [126].
prevent the crystallization. For example, CVD films (300 nm thick)
prepared from aluminum tri-isopropoxide [111] deposited at tem- 4.3. Solution methods
perature lower than 415 °C resulted in the hydroxylated film, while
crystallization was induced at deposition temperatures over 700 °C. Based on the described mechanism for the stabilization of am-Al2O3
Amorphous nanoparticles ranging from 5 to 100 nm were produced by with the introduction of stress, it is clear why am-Al2O3 was success-
introducing aluminum alkoxide vapors into a high-temperature furnace fully synthesized only in a thin film or nanoparticle form, mainly by
at 1000 °C [113]. The exact temperature of crystallization depends on PVD [127] and CVD [108], or by electrochemical aluminum oxidation
the thickness of the deposit. [11,12,15]. More convenient up-scalable techniques, as for example
Based on the reported studies, we can conclude that CVD technique sol-gel methods, have been considered but have also remained limited
that requires the dehydration at temperatures above 600 °C, cannot to the nanoparticles and thin film production [14,25,128–134]. An-
produce thicker am-Al2O3 films as the PVD technique that does not nealing of the films and nanoparticles, prepared with such methods, is
need the heat treatment. However, after heat treatment above 1000 °C required for dehydration and conversion of the initial product to the
the PVD and CVD films and powders of a high surface area result in a alumina [13,135,136]. The conversion is associated with full or in the
mixture of different polymorphs. Even more, pure α-Al2O3 crystallizes most cases partial crystallization of the oxide [108,137–139]. Recently,
only at 1500 °C [102,114,115]. This demonstrates that the high surface we demonstrated a method for high-temperature stabilization of bulk
area stabilizes not only amorphous but also metastable aluminas. amorphous alumina, prepared from sol-gel [23,55]. The stabilization
The difference between different vapor phase deposition techniques was achieved by dispersing polysilane dendritic molecules in the alu-
is remarkable. For example, the as deposited films prepared by the minum hydroxide gel. During heat treatment of the gel above 600 °C the
atomic layer deposition (ALD) contain much higher fraction of [Al]5 polysilane molecules oxidized, form covalent bonds with the matrix and
with polyhedra more disordered than in the PVD film [88]. The films induce a homogeneously distributed stress throughout the bulk of the
produced by the radio frequency magnetron sputtering showed the am-Al2O3 matrix (Fig. 3f). The stability towards crystallization depends
variation in density from 3.1 to 3.6 g cm−3 with an O/Al ratio of 1.5 or on the amount of the dispersed polysilane molecules and is limited at
higher, depending on the process conditions. Another group showed a 1000 °C. At this temperature, mullite crystallizes on the interface be-
local oxygen deficiency in the films prepared by the direct current tween the dispersed oxidized polysilane and alumina matrix. This high-
magnetron sputtering [93]. Di Fonzo et al. [104] showed three growth temperature stability allows for complete dehydration of the sol-gel
regimes for the pulsed laser deposition of alumina films. At low oxygen derived am-Al2O3 and its use in high-temperature applications.
pressure, the films were homogeneous and dense. An increase in the In addition to the sol-gel synthesis, solvothermal reactions have
deposition pressure led to columnar film growth. At even higher pres- been used to prepare nanostructured am-Al2O3. Nanotubes with a
sure it led to a film assembled from clusters with dendritic and porous diameter from 6 to 8 nm and a length up to 200 nm have been prepared
morphology with significantly lower density (under 3 g cm−3). hydrothermally from aluminum nitrate and sodium dodecyl sulfonate
as a structure-directing template. The amorphous nanotubes have
4.2. Oxidation of metal aluminum formed after a heat treatment at 500 °C and crystallized to γ-Al2O3 at
800 °C [140]. Microspheres consisting of branched nanorods have been
The thin films can also be prepared on metal Al by thermal or anodic prepared solvothermally from aluminum isopropoxide in acetic acid.
oxidation. Immediately after exposure of metal Al to oxygen, an oxide No template was used to prepare these structures. At first 1D nanorods
phase starts to form on its surface [117,118]. Brune et al. [117] studied were formed. By increasing the reaction time the nanorods self-orga-
oxygen interaction with the Al(111) surface by STM at 27 °C. They nized into the microspheres [37].
showed that oxygen starts to stick to the surface already at a 0.005
monolayer (ML) surface coverage. The coverage is random, but oxygen 5. Advanced applications of amorphous alumina
species are mobile and start to form islands. Oxidation starts by nu-
cleation of the first oxide nuclei already at low oxygen surface coverage 5.1. Mechanical properties and related applications
(0.2 ML). With further oxygen intake, the oxidation continues by the
formation of new nucleation spots rather than the grain growth. Before α-Al2O3 was successfully produced in the form of a thin film on a
a continuous oxide layer is formed the surface consists of the small stainless steel substrate with a laser-induced deposition and retained
oxide granules. The oxide formation already occurs at low pressures the hardness of the bulk form [149]. Presence of the γ-phase in the α-
(10−7–10−6 Torr) and temperatures around liquid nitrogen [118–121]. Al2O3 decreases the mechanical properties of the coatings [150]. On the
During the initial stages of the thermal oxidation of Al surfaces the other side, it is reported that the pure γ thin films produced by a pulsed
oxide layer is amorphous [59–61,122,123]. The growth of the oxide reactive magnetron sputtering can exhibit even better mechanical
layer depends on the heat treatment parameters. The limiting factor is properties than the α phase [151]. Molecular dynamics calculations of
the diffusion of the Al ions from the metal to the surface. At higher am-Al2O3 structural and elastic properties showed that the bulk and
temperatures, the diffusion kinetics increase and the oxide layer starts shear moduli are consistently lower in amorphous than the crystalline
to become thicker. At some point, depending on the Al surface Al2O3 phases [96].

7
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

Table 2
Mechanical properties of the am-Al2O3 thin films.
Method Deposition temperature Film thickness (substrate) Hardness [GPa] (method) Young modulus [GPa] Ref
[°C]

PLD RT ~1 μm (Si(100)) 20.8 (nanoindentation) N/A [29]


PLD N/A ≥3 μm (Si(100)) ~11 (nanoindentation) ~190 [104]
PLD RT 1 μm (Si(100)) 11.7 ± 0.3 (Berkovich diamond tip) 182 ± 5 [108]
PECVD RT 1 μm (Si(100) and SiO2/SiO) 6.4 ± 0.2 (Berkovich diamond tip) 116 ± 4 [108]
PECVD 300 ≥3 μm (Corning 7059 glass) 5.3 N/A [143]
350 9.6
500 11.3
(ultra-microhardness)
Reactive magnetron sputter 350 ~1 μm (Si(100)) 10–12 (Berkovich diamond tip) N/A [28]
deposition
e-beam evaporation 400 350 nm (SiO2) 5–8 (nanoindentation) 110–125 [142]
e-beam evaporation 700 116-254 nm (Si(100)) N/A 190–240 (increases with [97]
annealing time)
MOCVD 480 0.4–1.5 μm (Ti alloy) ~11 (nanoindentation) ~155 [19]
MOCVD 480 ~1 μm (Ti alloy and Si) 10.8 ± 0.8 (nanoindentation) 155 ± 6 [20]
Sol-gel 750 ~50 nm (glass) 7.4 ± 0.4 (Berkovich tip nano- 81 ± 5 [23]
indenter)

Table 2 summarizes the mechanical properties of the thin films is applied. As in the case of hardness, the annealing and the consequent
prepared by different deposition techniques. Based on the reported increase in density also cause an increase in the elasticity [21,142]. In
results it can be concluded that the hardness of the thin films containing addition, the coatings containing polycrystalline alumina polymorph
the crystalline forms of Al2O3 is higher than that of the amorphous films embedded in the amorphous matrix were found to benefit from both,
[19–22,28,29,108,142,143]. The hardness of the am-Al2O3 is largely the increased hardness due to the partial crystallinity and the increased
affected by a density of the thin films. The hardness increases with the wear resistance due to more compact morphology, good adhesion,
substrate temperature due to the increase in Al3+ average coordination surface smoothness and change in the coating deformation behavior
number and bond density [21,97,108,142,143]. This is in accordance [21,22,152]. The impurities are mostly affecting the CVD and sol-gel
with the molecular dynamic simulations that showed that the annealing films. Higher temperatures are required to produce the OH-free high-
results in the formation of the 6-coordinated Al and denser amorphous density films with sufficient adhesion to the substrate. In the range of
structure [91]. The calculations confirmed that these structural changes 400–600 °C, most of the am-Al2O3 films crystallize. This prevents fur-
increase Young's modulus and yield stress [98]. In accordance with the ther densification and elimination of the impurities from the film.
density, a decrease in the amount of impurities (chlorine, carbon, hy- We applied the am-Al2O3 coating onto a glass substrate [23]. The
drogen…) also increases the hardness [19,108,143]. Because the alu- coating was prepared in the nanocomposite form (Fig. 3g), by disper-
mina films prepared by PLD are very pure, they have superior char- sing the polycarbosilane molecules into the alumina matrix. The com-
acteristics compared to other methods. In addition, the PLD films can be pletely transparent coating behaved in a plastic manner. The wear re-
thicker because they are deposited at room temperature without further sistance exceeded that of the boro-aluminosilicate glass by a factor of
need for thermal treatment. From the point of morphology, it is im- 35 and the scratch resistance by more than an order of magnitude
portant that the film stays homogeneous and dense to achieve good (Fig. 8b,c). The plastic deformation of the am-Al2O3 coatings in contrast
mechanical properties. Growth regimes that lead to columnar film to fracture deformation of the polycrystalline coatings has also been
growth, a film assembled from clusters or a dendritic and porous observed by others [20–22,28]. The plasticity is attributed to the su-
morphology have a significantly lower density and deteriorated me- perior wear resistance of the amorphous coatings when compared to the
chanical properties [104]. polycrystalline alumina due to the buckling rather than the cracking.
Despite the lower hardness compared to the crystalline polymorphs,
the am-Al2O3 thin films are favorable as wear and starch resistant
coatings [20,21,28]. The continuous amorphous thin films without 5.2. Optical properties and related applications
grain boundaries experience adhesive failure; after a critical load,
buckling of the material occurs on a deformation edge. In contrast, the The am-Al2O3 band gap is in the range of 3.2–4.3 eV [9,55,153]
nanocrystalline coating experience a cohesive failure resulting in a depending on the structural short-range order [9,141]. This value is
plastic deformation already at a minimum load (Fig. 8a). The material much lower in contrast to the bulk crystalline phases, ~8.8 eV [154].
of the amorphous coating is buckling rather than cracking after the load Interestingly, ultrathin crystalline films with abundant defects have
been reported to have the bandgap in the range from 2.5-5 eV similar to

Fig. 8. a) The schematic of cross-section after adhesive and cohesive failure mode. b) Am-Al2O3 coated and d) uncoated glass surface after 1000 cycles with an
applied force 5 N. Reprinted with permission from [23]. Copyright 2016 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

8
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

am-Al2O3 [9,155]. The low band gap of the am-Al2O3 is caused by the 5.3.2. Electronic applications
stress-induced defects that give mid-bandgap states [9,153,156] and a The dense am-Al2O3 thin films have been successfully used as ca-
change in the bottom of the conductive band due to the different Al pacitors [116,132], gate dielectrics or insulators [31,147,148,163] with
coordination symmetries [56,96,141]. Although the early studies hy- high dielectric constant (up to 11), low leakage current, low factor loss
pothesized that the defect states originate from the oxygen vacancies and high breakdown field (300–700 MVm−1). The am-Al2O3 films were
[96,141,153], the NEXAFS measurements showed trapped O-O pairs successfully used as an insulator in thin film electroluminescence de-
inducing mid-band states [56]. Nevertheless, the oxygen vacancies, and vices [134]. The sol-gel and ALD deposited am-Al2O3 films were com-
the O-O pairs vanish upon the densification and so do the mid-band pletely transparent in the visible range and showed no charge transport
states. at the am-Al2O3/ZnS interfaces.
Although the crystalline alumina does not absorb visible light, it is Momida et al. [94] calculated that am-Al2O3 has a lower average
not easy to prepare transparent polycrystalline Al2O3 films. The pre- Al3+ coordination number than the crystalline Al2O3. This causes a
sence of boundaries between crystallites and interfaces with pores and wide Al-O-Al bond angle distribution due to several types of the Al sites
voids causes light scattering that lowers the transparency. To produce ([Al]4, [Al]5 and [Al]6) that are strongly geometrically distorted due to
the transparent polycrystalline films, it is necessary to avoid light the lack of symmetry. As a consequence, a contribution of lattice po-
scattering, which can be avoided by decreasing the number of scat- larization to the dielectric constant is higher in the amorphous than
tering centers or by decreasing the size of the scattering centers much crystalline material. The am-Al2O3 dielectric constant depends on its
below the wavelength of visible light. This is done either by keeping the structure. It can be tuned by increasing the average coordination
low thickness of the films and coatings, or by preparing extremely well around Al ion and the material density. The tunability has never been
sintered dense films with the particles in a range of 10 nm, low surface possible for the crystalline polymorphs [94]. In addition, the similar
and interface roughness, and small pore size [29,127,142,145,157]. On trend was experimentally observed and theoretically evaluated by
the other hand, the am-Al2O3 films have a big advantage to be free from Evangelisti et al. [148]. Combining an electron impedance spectroscopy
the crystallite boundaries that cause light scattering. The continuous and an Auger analysis they showed that the main contribution to the
am-Al2O3 films were successfully prepared and showed transparency in dielectric constant comes from the lattice polarization that pre-
the visible range [23–27,134]. dominantly depends on the atomic arrangement rather than on the
Am-Al2O3 has a lower refractive index than the crystalline phases electronic configuration and corresponding electronic polarization.
[28–30,108,146]. Since the refractive index is related to a density of the By controlling the characteristics of the prepared films such as in-
material it is not surprising that less dense am-Al2O3 has the refractive creasing density and maintaining the stoichiometric ratio, the am-Al2O3
index closer to the amorphous glass substrates. According to Snell's law, dielectric constant can reach that of the crystalline alumina. Cibert
this leads to smaller refraction on the coating-substrate interface, et al. [108] prepared the amorphous thin films with a dielectric con-
making am-Al2O3 more suitable for the glass coatings than the crys- stant of 9–10. After annealing and crystallization to γ-Al2O3, the di-
talline films. Although the refraction index is lower than that of the α electric constant slightly increased to 11–13. The insulating properties
phase; dense amorphous films can have the refraction index similar to get worse with oxygen deficiency [116] and porosity [134]; in contrast,
the γ structure [146]. they improve with the increasing am-Al2O3 density [148]. Thus, the
Another approach for manipulating optical properties of am-Al2O3 insulating properties largely depend on the preparation conditions.
is by doping. Such doped alumina can absorb light in visible range due Vanbeisen et al. [134] compared the leaking current of the sol-gel and
to electron transitions in atomic levels of the dopant ion. Co3+ ions can ALD prepared am-Al2O3 thin films. The sol-gel films have more leakage
be completely dispersed in the alumina matrix up to 7 vol% [158]. The and lower break-down fields due to the higher film porosity. The
tetrahedral coordinated Co3+ gives broad overlapping peaks in the sputtering parameters affected the stoichiometry of the r.f. magnetron
range from 500 to 700 nm. Cu2+ doping gives penta-coordinated Cu2+ sputtered films [116]. An Al/O ratio below the stoichiometric value of
that absorbs visible light above 500 nm [159]. The broad absorption 1.5 gave the films with poorer insulator properties than that of the
peaks are a consequence of a different local environment of dopant oxygen-rich films, demonstrating the effect of the oxygen vacancies on
metal ions, causing different energy level distortions of their energy the conductivity. Interestingly, the cation doping can control the
levels. amount of oxygen vacancies. The presence of the cation vacancies
prevents formation of the oxygen vacancies and increases a breakdown
5.3. Other advanced applications strength [163–165]. The am-Al2O3 co-doped by 2% of Si4+ and 1% of
Mg2+ gives the breakdown strength of 544 MV m−1, which is twice as
5.3.1. Protective coatings high as 276 MV m−1 of the un-doped film [164]. Doping with 10% of
In contrast with the hydrophilic polycrystalline structures (contact Ti4+ gives breakdown strength of 530 MV m−1 [165]. Doping with
angle ~50°) [19] the amorphous form is hydrophobic (contact 10% of Zr4+ increased dielectric constant from 6.2 to 11.8 for sol-gel
angle > 100°) making it great catalyst support for anti-fouling and self- deposited thin films [163]. Zou et al. investigated the effect of La3+
cleaning applications where low wettability is desirable. The amor- doping in the range from 5 to 10% [166] that was shown to increase the
phous alumina coatings were tested for corrosion protection in a saline breakdown strength. Due to the same oxidation state as Al3+, La3+ does
environment. The am-Al2O3 MOCVD films on a titanium alloy, with a not introduce the cation vacancies. Because of a significantly bigger
thickness ranging from 300 to 1500 nm, [19,33] increase the corrosion radius, the La3+ ions make the structure of am-Al2O3 thin films more
resistance for more than two orders of magnitude compared to the compact, which suppresses the ionic mobility. In addition, it was pro-
uncoated alloy. Other forms of alumina provide much more limited posed that La3+ might help to trap and scatter the charge, thus en-
protection. The reason is in the microstructure. A laminar structure of hancing the breakdown strength. The described studies showed that the
aluminum hydroxide and grain boundaries of crystalline coatings allow cation doping is an efficient technique for tuning the am-Al2O3 elec-
for electrolyte penetration towards the metal surface while the con- tronic properties, which is essential for the applications in the field of
tinuous amorphous layer tightly closes down the substrate surface and protective coating and dielectrics.
prevents the contact with the oxidizing environment. Similar observa-
tions have been reported for the 2000–3000 nm thick sol-gel coatings 5.3.3. Energy storage
on magnesium [34] and the mixed amorphous/nanocrystalline coatings Sol-gel derived am-Al2O3 was used in a multilayer Al/am-Al2O3/Pt
on aluminum alloy [152]. Ultrathin (< 5 nm) ALD thin films have been capacitor [132]. The leakage current density was suppressed to
used to protect silicon photocathodes for hydrogen evolution reaction ~5.4 × 10−5 A cm−2, the operating voltage was enhanced to 591 MV
[160–162], stability exceeded 100 h [161]. m−1 and the energy density was ~14 J cm−3. The increased

9
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

performance of the capacitor was attributed to the am-Al2O3 layer that


was grown by the anodic oxidation of Al electrode on the sol-gel de-
posited Al/am-Al2O3 interface (Fig. 10a-NAO) and provided the ultra-
high resistivity (9.4 × 1012 Ω cm−1). With this, they demonstrated the
application of the interface formed by two am-Al2O3 films that are
structurally different. Although the disorder and O-vacancies are fa-
vorable for electron trapping, they also lower the dielectric constant,
resulting in a high leaking current. Dense anodic film (Fig. 10a - NAO)
provided high resistivity, whereas more disordered sol/gel film pro-
vided a volume for energy storage (Fig. 10a-AmAO).
The supercapacitor in which the am-Al2O3 is not structurally
homogenous was also prepared by Fukuhara et al. [167]. They prepared
the am-Al2O3 supercapacitor by anodic oxidation of the metallic AlY3
amorphous alloy. The growth of am-Al2O3 films by anodic oxidations
results in different density and structure near the metal/oxide interface
and the oxide/electrolyte interface that results from the diffusion and
inclusion of the ions through the oxide layer. The [Al]6 clusters were
accumulated on the surface, the surface resistance was high and the
leaking of charge was prevented. The structure in the core of the film
was more disordered. This resulted in enhanced electron trapping on
the positive charged O-vacancy and improved energy storage. Inter-
estingly, with an increase of the film thickness, the film partially
crystalized into the mixture of γ-Al2O3 and α-Al2O3. Regardless of the
higher amount of the material, the amount of the stored charge was
decreased due to lower population of O-vacancies in crystalline do-
mains.
Another approach for decreasing the leaking of the am-Al2O3 su-
percapacitor was introduced by Feng et al. [164]. With co-doping by
Si4+ and Mg2+, the cation vacancies were generated. This prevented
the formation of oxygen vacancies and increased the breakdown
strength. Compared to the un-doped am-Al2O3 thin film, the energy
density of (Al.97Si.02Mg.01) was enhanced from 6.2 J cm−3 to
9.2 J cm−3. However, with a further increase in dopant concentration,
the stored energy density decreased significantly, revealing the im-
portance of the oxygen vacancies for the charge storage.

5.3.4. Photoluminescent applications


The oxygen vacancies have been found to be responsible for in- Fig. 9. a) Fluorescence image of the dendritic cells after the uptake of am-Al2O3
tensive blue photoluminescence in the am-Al2O3 particles [12–14] and nanoparticles excited at 370 nm and examined with a fluorescence microscope.
porous membranes [15]. The oxygen vacancies cause F+ center defects The inset shows the microscopic image of a contrast cell without am-Al2O3
nanoparticles uptake. Reprinted with permission from [12]. Copyright 2012
(singly ionized oxygen vacancies) in the band gap. These defects act as
American Chemical Society. b) Room temperature photoluminescence emission
charge traps and cause the luminescence [12,15]. For the 3-nm am-
from am-Al2O3 doped with 10% Eu3+. The light emission shows the char-
Al2O3 nanoparticles, higher photoluminescence intensity was observed acteristic peaks associated with radiative transitions between the electron en-
in contrast to the same sized γ- and α- phase nanoparticles [12]. These ergy levels of Eu3+ ions. Reprinted with permission from [16]. Copyright 2002
am-Al2O3 nanoparticles are small enough (smaller than the spread of by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
wave function) for the quantum confinement effect to occur [11,12].
Although crystalline nanoparticles show weak blue photoluminescence
Eu3+ doped am-Al2O3 after excitation in the UV range, due to the most
[12], its intensity is low due to the low concentration of oxygen va-
intensive 5D0-7F2 transition at approximately 615 nm [16,171,172].
cancies. It was shown, by combining EELS and photoluminescence
The luminescence intensity depends on the concentration of Eu3+
measurements with quantum simulations, that the oxygen vacancies
[16,171]. It increases with increasing the Eu3+ concentration and
have similar energy positions in the crystalline and am-Al2O3, however,
reaches a peak at 10% of Eu3+ [16]. Similar luminescence, due to the
the oxygen deficiency in am-Al2O3 is higher [153]. In addition to the F
transitions between the energy levels of the dopant ion, has also been
+ centers, weak photoluminescence is also produced by the F center
observed for other rear-earth metal ions (Gd3+, Ho3+, Pr3+, Sm3+,
defects (doubly ionized oxygen vacancies) [15,153] and carbonyl im-
Dy3+ and Tm3+) [173].
purities [13,168]. Ultrathin am-Al2O3 nanoparticles (< 4 nm), pro-
Interestingly, in addition to the luminescence, an upconversion
duced by ultrasonication of anodic am-Al2O3 membranes, were used as
emission has been observed for Er3+ and Nd3+ doped am-Al2O3 and
fluorescent biological labels [12]. After the exposure of the dendritic
excited with 800 nm laser light [174]. The upconversion from Er3+
cells to the am-Al2O3 nanoparticles for 24 h, the labeled cells excited at
emits at 548 nm (4S3/2–4I15/2) and 640 nm (4F9/2–4I11/2), and Nd3+
370 nm showed a clear blue fluorescence (Fig. 9a). Most likely, the am-
emits at 640 nm (4G7/2–4I11/2). The emission is weak and the origin of
Al2O3 nanoparticles penetrate the cell membrane by endocytosis. Nei-
the upconversion is unknown. It is speculated that it may be related to a
ther detectable toxicity nor inhibitory effect on cell proliferation was
multi-phonon relaxation resulting from inhomogeneity of the alumina
observed after continuous exposure for two days.
matrix [174,175].
In contrast to wide photoluminescence peaks of pure am-Al2O3, the
doping with lanthanide ions (e.g. Eu3+ [16,169–172]) gives sharp
emission lines, due to transitions between atomic levels of the dopant 5.3.5. Ferromagnetism
ion (Fig. 9b). In general, a red light luminescence has been reported for The oxygen vacancies, in particular F+ centers, are also responsible

10
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

and pH, these OH sites can act as acid or base active sites [2,4].
However, am-Al2O3 has rarely been considered for the purpose of the
catalysis or adsorption. The nanoparticles doped with La have been
used as the catalyst for the biodiesel conversion [36] and the nanofibrils
for the fluoride removal from water [37].
The molecular dynamics studies have shown that the am-Al2O3
surface is oxygen terminated, causing Al3+ enrichment just below the
surface [99]. The surface relaxes by decreasing the average coordina-
tion of Al ions, i.e. higher amount of three coordinated Al3+ and edge-
sharing tetrahedra near the surface in contrast to the bulk. Although the
surfaces of the crystalline alumina are also terminated with oxygen
atoms, the surface is reconstructed by cation vacancies rather than
lower Al coordination, due to significant inward relaxation of Al atoms.
Due to the higher distribution of low coordinated Al3+ on the surface,
the amorphous surface is superior catalyst to the crystalline surfaces.
The importance of exposed Al3+ sites was shown on the (100) facets of
nanosized γ-Al2O3, where the thermal annealing and dehydration at
300 °C result in some amount of under-coordinated [Al]5 sites [74,177].
When preparing a Pt/γ-Al2O3 catalyst, firstly Pt is bound to those [Al]5
Fig. 10. Advanced properties of am-Al2O3. a) TEM image of a cross-section of sites resulting in uniform atomic dispersion. It is suggested that the
the Al/am-Al2O3 capacitor layer for energy storage with newly formed alu- presence of the [Al]5 sites could potentially lead to favorable dispersion
minum oxide (NAO) providing high resistivity. Republished with permission of of deposited guest metals [74]. The presence of the [Al]5 sites has been
Royal Society of Chemistry (Great Britain), from [132]; permission conveyed experimentally confirmed in amorphous silica-alumina (ASA) and on
through Copyright Clearance Center, Inc. b) Room temperature M-H curves of
the surface of some zeolites [178,179]. It was shown that [Al]5 and
9–12 nm nanoparticles heat treated at different temperatures (S600 – am-Al2O3;
[Al]4 coexist on such surfaces, serving as Brønsted acid sites (proton
S700 – am−/γ-Al2O3; S800 and S900 – γ-Al2O3; S1000 – α-Al2O3. The M-H
curves of bulk Al2O3 shown in the inset. Reprinted with permission from [17]. donating sites). By increasing the Al concentration in the ASA, the
Copyright 2011 American Chemical Society. c) Temperature dependence of number of [Al]5 increases. Consequently, the amount of acid sites in-
thermal conductivity (TC) enhancement of am-Al2O3/ethylene glycol nano- creases. As an example, at the alumina concentration of 70%, phe-
fluids for am-Al2O3 loadings of 0.25, 0.50, and 1.0 vol%. Reprinted from [18], nylglyoxal conversion to ethyl mandelate in ethanol proceeds with a
with the permission of AIP Publishing. d) Schematic representation of an in- highest yield of 99.8% [179]. Thus, we conclude that am-Al2O3 with its
teraction of the amorphous and crystalline nanoparticles in ethylene glycol (EG) high and adaptable amount of [Al]5 sites offers a good perspective for
fluid, showing stronger solvation of amorphous nanoparticles. Reprinted from alumina supported catalysis.
[18], with the permission of AIP Publishing.
6. Conclusions
for intrinsic ferromagnetism that has been reported for the am-Al2O3
nanoparticles at room temperature [17]. As the magnetization is related Aluminum is an earth-abundant, inexpensive element; therefore its
to the concentration of vacancies and is the highest for am-Al2O3, it crystalline oxide polymorphs have anchored in many fields of tech-
decreases with improved crystallinity (formation of γ) and higher nology. Yet, a little attention has been given to applications of the
average coordination number of the Al ions (α) (Fig. 10b). In bulk amorphous alumina, both in the fields where crystalline polymorphs
alumina, no ferromagnetism was detected. dominate (catalysis, catalysis support, coatings, adsorption, etc.) and
for more advanced applications. Certainly, the reason lies in its ther-
modynamic stability issue that makes amorphous alumina hard to be
5.3.6. Thermal conductivity enhancement
processed in a bulk form. Recently, our group and other authors de-
Due to the high thermal conductivity of Al2O3, its nanoparticles
veloped new preparation methods to sustain the amorphous structure
were used for thermal conductivity enhancement of heat transfer fluids.
under hostile conditions, characterized their unique structures and
The thermal conductivity of ethylene glycol increased for up to 75%
successfully applied in many fields. This has triggered new interests in
after the addition of 1% Al2O3 [18] (Fig. 10c). The enhancement is
the amorphous alumina and new attempts to stabilize it to resist the
higher by the addition of am-Al2O3 particles relative to the γ-Al2O3 and
high-temperatures, for instance by the smart engineering of the lattice
α-Al2O3. The amorphous nanoparticles contain a larger amount of
strain. This high temperature crystallization issue has been addressed so
randomly distributed tetra-, penta- and octahedra that serve as active
nicely that the potential applications of this unique material will have
sites for glycol adsorption [5,18]. These results in stronger solvation
received great attention.
and better dispersion of the amorphous nanoparticles compared to
The main advantages of am-Al2O3 originate from its amorphous
crystalline nanoparticles (Fig. 10d). As a result, a higher volume frac-
structure consisting of randomly distributed polyhedra that results in
tion of the dispersed nanoparticles can be achieved, which improves the
lower coordination number and density in contrast to the crystalline
thermal conductivity.
polymorphs. Due to the absence of crystallite boundaries, thin films are
usually uniform and transparent. By controlling the density, thin films
5.3.7. Catalysis and adsorption with various hardness, elasticity, resistivity, and dielectric constant can
The metastable alumina polymorphs are generally poorly crystal- be obtained. This further extends into better wear and scratch re-
lized [1]. This results in highly defected structures with many active sistance, hydrophobicity, corrosion protection properties, light ab-
sites for catalysis and adsorption [4,176]. In the fully dehydrated alu- sorption, photoluminescence, ferromagnetism, and catalysis. Moreover,
mina polymorphs, the active site is [Al]4, that acts as a strong Lewis tailoring the coordination of Al ions and lattice density of am-Al2O3 can
acid [2,3]. Because the completely dehydrated alumina is hard to easily control all these properties.
produce at mild temperatures, required to avoid structural changes, the Up-to-date different structures of am-Al2O3 have exploited in many
catalytic reactions mainly take place in a water medium. The alumina advanced applications. For instance, a combination of two am-Al2O3
catalysts normally contain OH groups or water molecules bonded on layers with different densities has been used as a supercapacitor for
their surface. Depending on the chemical surrounding of the OH groups energy storage. Controlling over nanoparticle size was demonstrated to

11
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

allow tuning of the photoluminescence wavelength. Nanoparticles were Induced room temperature ferromagnetism in amorphous and crystalline Al2O3
employed as photoluminescent labels and heat exchange fluids. The nanoparticles, J. Phys. Chem. C 115 (2011) 16814–16818.
[18] J. Gangwar, A.K. Srivastava, S.K. Tripathi, M. Wan, R.R. Yadav, Strong enhance-
am-Al2O3 thin film dielectrics were used in electronic devices and ment in thermal conductivity of ethylene glycol-based nanofluids by amorphous
super-capacitors. The surface of am-Al2O3 has higher amount of ex- and crystalline Al2O3 nanoparticles, Appl. Phys. Lett. 105 (2014).
posed Al3+ sites in contrast to the crystalline alumina, making it a [19] D. Samelor, A.-M. Lazar, M. Aufray, C. Tendero, L. Lacroix, J.-D. Beguin,
B. Caussat, H. Vergnes, J. Alexis, D. Poquillon, N. Pebere, A. Gleizes, C. Vahlas,
promising candidate for catalysis and adsorption. Evidently, there is a Amorphous Alumina coatings: processing, structure and remarkable barrier
lot of room for further investigations of am-Al2O3. By now, little at- properties, J. Nanosci. Nanotechnol. 11 (2011) 8387–8391.
tention has been given to applications in catalysis, extremely aggressive [20] Y. Balcaen, N. Radutoiu, J. Alexis, J.D. Beguin, L. Lacroix, D. Samelor, C. Vahlas,
Mechanical and barrier properties of MOCVD processed alumina coatings on
oxidizing environment, corrosion coatings, superhydrophobicity, oleo- Ti6Al4V titanium alloy, Surf. Coat. Technol. 206 (2011) 1684–1690.
phobicity, icephobicity and optomagnetics. It is almost certain that in [21] F. Garcia Ferre, E. Bertarelli, A. Chiodoni, D. Carnelli, D. Gastaldi, P. Vena,
the next decades the scientists' and engineers' ingenuity will come up M.G. Beghi, F. Di Fonzo, The mechanical properties of a nanocrystalline Al2O3/a-
Al2O3 composite coating measured by nanoindentation and Brillouin spectro-
with a number of very exciting new applications that will take ad-
scopy, Acta Mater. 61 (2013) 2662–2670.
vantage of the uniqueness of amorphous alumina. [22] M. Sieber, T. Mehner, D. Dietrich, G. Alisch, D. Nickel, D. Meyer, I. Scharf,
T. Lampke, Wear-resistant coatings on aluminium produced by plasma anodising
Declaration of interest statement A correlation of wear properties, microstructure, phase composition and dis-
tribution, Surf. Coat. Technol. 240 (2014) 96–102.
[23] M. Valant, U. Luin, M. Fanetti, A. Mavric, K. Vyshniakova, Z. Siketic, M. Kalin,
That there are no known conflicts of interest associated with this Fully Transparent Nanocomposite Coating with an Amorphous Alumina Matrix
publication and there has been no significant financial support for this and Exceptional Wear and Scratch Resistance, Adv. Funct. Mater. 26 (2016)
4362–4369.
work that could have influenced its outcome. [24] B.P. Dhonge, T. Mathews, S.T. Sundari, C. Thinaharan, M. Kamruddin, S. Dash,
A.K. Tyagi, Spray pyrolytic deposition of transparent aluminum oxide (Al2O3)
Acknowledgment films, Appl. Surf. Sci. 258 (2011) 1091–1096.
[25] B.F. Hu, M.W. Yao, R.H. Xiao, J.W. Chen, X. Yao, Optical properties of amorphous
Al2O3 thin films prepared by a sol-gel process, Ceram. Int. 40 (2014)
AM and MV acknowledge the financial support from the Slovenian 14133–14139.
Research Agency (research core funding No. P2-0412). MV acknowl- [26] I.N. Reddy, V.R. Reddy, N. Sridhara, V.S. Rao, M. Bhattacharya,
P. Bandyopadhyay, S. Basavaraja, A.K. Mukhopadhyay, A.K. Sharma, A. Dey,
edges the distinguished international expert program appointed by the Pulsed rf magnetron sputtered alumina thin films, Ceram. Int. 40 (2014)
Chinese Ministry of Education. ZW acknowledge National Program on 9571–9582.
Key Basic Research Project (973 Program) 2013CB933301 and National [27] F.O. Ogundare, I.O. Olarinoye, He+ induced changes in the surface structure and
optical properties of RF-sputtered amorphous alumina thin films, J. Non-Cryst.
Natural Science Foundation of China 51272038.
Solids 432 (2016) 292–299.
[28] J.L. Wang, Y.H. Yu, S.C. Lee, Y.W. Chung, Tribological and optical properties of
References crystalline and amorphous alumina thin films grown by low-temperature reactive
magnetron sputter-deposition, Surf. Coat. Technol. 146 (2001) 189–194.
[29] G. Balakrishnan, S.T. Sundari, R. Ramaseshan, R. Thirumurugesan, E. Mohandas,
[1] I. Levin, D. Brandon, Metastable alumina polymorphs: crystal structures and D. Sastikumar, P. Kuppusami, T.G. Kim, J.I. Song, Effect of substrate temperature
transition sequences, J. Am. Ceram. Soc. 81 (1998) 1995–2012. on microstructure and optical properties of nanocrystalline alumina thin films,
[2] M. Trueba, S.P. Trasatti, gamma-Alumina as a support for catalysts: a review of Ceram. Int. 39 (2013) 9017–9023.
fundamental aspects, Eur. J. Inorg. Chem. (2005) 3393–3403. [30] M. Tulio Aguilar-Gama, E. Ramirez-Morales, Z. Montiel-Gonzalez, A. Mendoza-
[3] G. Busca, Structural, surface, and catalytic properties of aluminas, Adv. Catal. 57 Galvan, M. Sotelo-Lerma, P.K. Nair, H. Hu, Structure and refractive index of thin
(2014) 319–404. alumina films grown by atomic layer deposition, J. Mater. Sci. Mater. El. 26
[4] G. Busca, The surface of transitional aluminas: A critical review, Catal. Today 226 (2015) 5546–5552.
(2014) 2–13. [31] P. Katiyar, C. Jin, R.J. Narayan, Electrical properties of amorphous aluminum
[5] J. Gangwar, B.K. Gupta, S.K. Tripathi, A.K. Srivastava, Phase dependent thermal oxide thin films, Acta Mater. 53 (2005) 2617–2622.
and spectroscopic responses of Al2O3 nanostructures with different morphogen- [32] C. Avis, J. Jang, High-performance solution processed oxide TFT with aluminum
esis, Nanoscale 7 (2015) 13313–13344. oxide gate dielectric fabricated by a sol-gel method, J. Mater. Chem. 21 (2011)
[6] H.O. Ali, Review of porous anodic aluminium oxide (AAO) applications for sen- 10649–10652.
sors, MEMS and biomedical devices, Trans. Inst. Met. Finish. 95 (2017) 290–296. [33] G. Boisier, M. Raciulete, D. Samelor, N. Pebere, A.N. Gleizes, C. Vahlas,
[7] A.M.M. Jani, D. Losic, N.H. Voelcker, Nanoporous anodic aluminium oxide: Electrochemical behavior of chemical vapor deposited protective aluminum oxide
Advances in surface engineering and emerging applications, Prog. Mater. Sci. 58 coatings on Ti6242 titanium alloy, Electrochem. Solid-State Lett. 11 (2008)
(2013) 636–704. C55–C57.
[8] C. Lu, Z. Chen, Anodic Aluminium Oxide-Based Nanostructures and Devices, in: [34] I.B. Singh, P. Gupta, A. Maheshwari, N. Agrawal, Corrosion resistance of sol-gel
H.S. Nalwa (Ed.), Encyclopedia of Nanoscience and Nanotechnology, American alumina coated Mg metal in 3.5 % NaCl solution, J. Sol-Gel Sci. Techn. 73 (2015)
Scientific Publishers, 2011, pp. 235–259. 127–132.
[9] I. Costina, R. Franchy, Band gap of amorphous and well-ordered Al2O3 on Ni3Al [35] J.S. Daubert, G.T. Hill, H.N. Gotsch, A.P. Gremaud, J.S. Ovental, P.S. Williams,
(100), Appl. Phys. Lett. 78 (2001) 4139–4141. C.J. Oldham, G.N. Parsons, Corrosion protection of copper using Al2O3, TiO2,
[10] A.K.N. Kumar, S. Prasanna, B. Subramanian, S. Jayakumar, G.M. Rao, A trans- ZnO, HfO2, and ZrO2 Atomic layer deposition, ACS Appl. Mater. Interfaces 9
mission electron microscopy and X-ray photoelectron spectroscopy study of an- (2017) 4192–4201.
nealing induced gamma-phase nucleation, clustering, and interfacial dynamics in [36] G. Amini, G.D. Najafpour, S.M. Rabiee, A.A. Ghoreyshi, Synthesis and character-
reactively sputtered amorphous alumina thin films, J. Appl. Phys. 117 (2015). ization of amorphous Nano-Alumina powders with high surface area for biodiesel
[11] W.J. Zhang, X.L. Wu, J.Y. Fan, G.S. Huang, T. Qiu, P.K. Chu, Luminescent amor- production, Chem. Eng. Technol. 36 (2013) 1708–1712.
phous alumina nanoparticles in toluene solution, J. Phys-Condens. Mat. 18 (2006) [37] D.J. Kang, S.R. Tong, X.L. Yu, M.F. Ge, Template-free synthesis of 3D hierarchical
9937–9942. amorphous aluminum oxide microspheres with broccoli-like structure and their
[12] X.L. Wu, S.J. Xiong, J.H. Guo, L.L. Wang, C.Y. Hua, Y.Y. Hou, P.K. Chu, Ultrathin application in fluoride removal, RSC Adv. 5 (2015) 19159–19165.
Amorphous Alumina Nanoparticles with Quantum-Confined Oxygen-Vacancy- [38] P. Brand, R. Troschke, H. Weigelt, Formation of alpha-Al2O3 by thermal-decom-
Induced Blue Photoluminescence as Fluorescent Biological Labels, J. Phys. Chem. position of basic aluminum chlorides at low-temperatures, Cryst. Res. Technol. 24
C 116 (2012) 2356–2362. (1989) 671–675.
[13] C.K. Lin, M. Yu, Z.Y. Cheng, C.M. Zhang, O.G. Meng, J. Lin, Bluish-white emission [39] S. Rajendran, Production of ultrafine alpha-alumina powders and fabrication of
from radical carbonyl impurities in amorphous Al2O3 prepared via the Pechini- fine-grained strong ceramics, J. Mater. Sci. 29 (1994) 5664–5672.
type sol-gel process, Inorg. Chem. 47 (2008) 49–55. [40] C.S. Nordahl, G.L. Messing, Sintering of alpha-Al2O3-seeded nanocrystalline
[14] A. Amirsalari, S.F. Shayesteh, Effects of pH and calcination temperature on gamma-Al2O3 powders, J. Eur. Ceram. Soc. 22 (2002) 415–422.
structural and optical properties of alumina nanoparticles, Superlattice. Microst. [41] A.R. Boccaccini, C. Kaya, Alumina ceramics based on seeded boehmite and elec-
82 (2015) 507–524. trophoretic deposition, Ceram. Int. 28 (2002) 893–897.
[15] G.G. Khan, A.K. Singh, K. Mandal, Structure dependent photoluminescence of [42] H.S. Kim, M. Kang, Rapid crystal phase transformation into hexagonally shaped
nanoporous amorphous anodic aluminium oxide membranes: role of F+ center alpha-alumina using AlF3 seeds, J. Sol-Gel Sci. Technol. 68 (2013) 110–120.
defects, J. Lumin. 134 (2013) 772–777. [43] Z. Zivkovic, N. Strbac, J. Sestak, Influence of fluorides on polymorphous trans-
[16] A.E. Esparza-Garcia, M. Garcia-Hipolito, M.A. Aguilar-Frutis, C. Falcony, formation of alpha-Al2O3 formation, Thermochim. Acta 266 (1995) 293–300.
Cathodoluminescent and photoluminescent properties of Al2O3 powders doped [44] Z. Li, Z. Li, A. Zhang, Y. Zhu, Synergistic effect of alpha-Al2O3 and (NH4)3AlF6 co-
with Eu, Phys. Stat. Sol. A 193 (2002) 117–124. doped seed on phase transformation, microstructure, and mechanical properties of
[17] G.J. Yang, D.Q. Gao, J.L. Zhang, J. Zhang, Z.H. Shi, D.S. Xue, Evidence of Vacancy- nanocrystalline alumina abrasive, J. Alloys Compd. 476 (2009) 276–281.

12
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

[45] Q.B. Tian, Y.Y. Zhang, X.J. Yang, J.S. Dai, Z.J. Lv, Influences of NH4F on trans- [75] H.P. Pinto, R.M. Nieminen, S.D. Elliott, Ab initio study of gamma-Al2O3 surfaces,
formation and morphology of high-pure alpha-alumina, Mater. Res. Express 4 Phys. Rev. B 70 (2004).
(2017). [76] Y. Oka, T. Takahashi, K. Okada, S.I. Iwai, Structural-analysis of anodic alumina
[46] J.M. McHale, A. Auroux, A.J. Perrotta, A. Navrotsky, Surface energies and ther- films, J. Non-Cryst. Solids 30 (1979) 349–357.
modynamic phase stability in nanocrystalline aluminas, Science 277 (1997) [77] P.S. Sklad, P. Angelini, J. Sevely, Eextended electron-energy loss fine-structure
788–791. analysis of amorphous Al2O3, Philos. Mag. A 65 (1992) 1445–1461.
[47] X. Bokhimi, J.A. Toledo-Antonio, M.L. Guzman-Castillo, B. Mar-Mar, [78] A.J. Bourdillon, S.M. Elmashri, A.J. Forty, Application of TEM extended electron-
F. Hernandez-Beltran, J. Navarrete, Dependence of boehmite thermal evolution on energy loss fine-structure to the study of aluminium-oxide films, Philos. Mag. A 49
its atom bond lengths and crystallite size, J. Solid State Chem. 161 (2001) (1984) 341–352.
319–326. [79] S.M. Elmashri, R.G. Jones, A.J. Forty, An electron-yield EXAFS study of anodic-
[48] B.K. Gan, I.C. Madsen, J.G. Hockridge, In situ X-ray diffraction of the transfor- oxide and hydrated-oxide films on pure aluminum, Philos. Mag. A 48 (1983)
mation of gibbsite to alpha-alumina through calcination: effect of particle size and 665–683.
heating rate, J. Appl. Crystallogr. 42 (2009) 697–705. [80] L.P.H. Jeurgens, F. Reichel, S. Frank, G. Richter, E.J. Mittemeijer, On the devel-
[49] H.N. Kim, S.K. Lee, Effect of particle size on phase transitions in metastable alu- opment of long-range order in ultra-thin amorphous Al2O3 films upon their
mina nanoparticles: A view from high-resolution solid-state Al-27 NMR study, Am. transformation into crystalline y-Al2O3, Surf. Interface Anal. 40 (2008) 259–263.
Mineral. 98 (2013) 1198–1210. [81] R. Manaila, A. Devenyi, E. Candet, Structural order in amorphous aluminas, Thin
[50] J.W. Drazin, D.A. Kazerooni, E.P. Gorzkowski, C.R. Feng, S.B. Qadri, R. Goswami, Solid Films 116 (1984) 289–299.
B.N. Feigelson, J.A. Wollmershauser, Reducing the Size of Nanocrystals below the [82] P. Lamparter, R. Kniep, Structure of amorphous Al2O3, Phys. B 234 (1997)
Thermodynamic Size Limit, Cryst. Growth Des. 17 (2017) 1752–1758. 405–406.
[51] A.H. Tavakoli, P.S. Maram, S.J. Widgeon, J. Rufner, K. van Benthem, S. Ushakov, [83] G. Kunathfandrei, T.J. Bastow, J.S. Hall, C. Jager, M.E. Smith, Quantification of
S. Sen, A. Navrotsky, Amorphous Alumina Nanoparticles: Structure, Surface aluminum coordinations in amorphous aluminas by combined central and satellite
Energy, and Thermodynamic Phase Stability, J. Phys. Chem. C 117 (2013) transition magic-angle-spinning NMR-spectroscopy, J. Phys. Chem. 99 (1995)
17123–17130. 15138–15141.
[52] L. Bloch, Y. Kauffmann, B. Pokroy, Size effect on the short range order and the [84] R. Lizarraga, E. Holmstroem, S.C. Parker, C. Arrouvel, Structural characterization
crystallization of nanosized amorphous alumina, Cryst. Growth Des. 14 (2014) of amorphous alumina and its polymorphs from first-principles XPS and NMR
3983–3989. calculations, Phys. Rev. B 83 (2011).
[53] J. Murray, K. Song, W. Huebner, M. O'Keefe, Electron beam induced crystallization [85] J.F. Stebbins, S. Kroeker, S.K. Lee, T.J. Kiczenski, Quantification of five- and six-
of sputter deposited amorphous alumina thin films, Mater. Lett. 74 (2012) 12–15. coordinated aluminum ions in aluminosilicate and fluoride-containing glasses by
[54] R. Nakamura, M. Ishimaru, H. Yasuda, H. Nakajima, Atomic rearrangements in high-field, high-resolution Al-27 NMR, J. Non-Cryst. Solids 275 (2000) 1–6.
amorphous Al2O3 under electron-beam irradiation, J. Appl. Phys. 113 (2013). [86] J.L. Yarger, K.H. Smith, R.A. Nieman, J. Diefenbacher, G.H. Wolf, B.T. Poe,
[55] A. Mavrič, M. Fanetti, G. Mali, M. Valant, High-temperature stabilization of bulk P.F. McMillan, Al coordination changes in high-pressure aluminosilicate liquids,
amorphous Al2O3, J. Non-Cryst. Solids 499 (2018) 363–370. Science 270 (1995) 1964–1967.
[56] C. Arhammar, A. Pietzsch, N. Bock, E. Holmstroem, C.M. Araujo, J. Grasjo, [87] S.K. Lee, S.B. Lee, S.Y. Park, Y.S. Yi, C.W. Ahn, Structure of Amorphous Aluminum
S. Zhao, S. Green, T. Peery, F. Hennies, S. Amerioun, A. Foehlisch, J. Schlappa, Oxide, Phys. Rev. Lett. 103 (2009).
T. Schmitt, V.N. Strocov, G.A. Niklasson, D.C. Wallace, J.-E. Rubensson, [88] S.K. Lee, S.Y. Park, Y.S. Yi, J. Moon, Structure and disorder in amorphous alumina
B. Johansson, R. Ahuja, Unveiling the complex electronic structure of amorphous thin films: insights from high-resolution solid-state NMR, J. Phys. Chem. C 114
metal oxides, Proc. Natl. Acad. Sci. U. S. A. 108 (2011) 6355–6360. (2010) 13890–13894.
[57] N. Kemik, S.V. Ushakov, M. Gu, N. Schichtel, C. Korte, N.D. Browning, [89] I. Farnan, R. Dupree, Y. Jeong, G.E. Thompson, G.C. Wood, A.J. Forty, Structural
Y. Takamura, A. Navrotsky, Yttria-stabilized zirconia crystallization in Al2O3/YSZ chemistry of anodic alumina, Thin Solid Films 173 (1989) 209–215.
multilayers, J. Mater. Res. 27 (2012) 939–943. [90] L.A. O'Dell, S.L.P. Savin, A.V. Chadwick, M.E. Smith, A Al-27 MAS NMR study of a
[58] L.P.H. Jeurgens, W.G. Sloof, F.D. Tichelaar, E.J. Mittemeijer, Thermodynamic sol-gel produced alumina: identification of the NMR parameters of the theta-
stability of amorphous oxide films on metals: application to aluminum oxide films Al2O3 transition alumina phase, Solid State Nucl. Magn. Reson. 31 (2007)
on aluminum substrates, Phys. Rev. B 62 (2000) 4707–4719. 169–173.
[59] M.A. Trunov, M. Schoenitz, X.Y. Zhu, E.L. Dreizin, Effect of polymorphic phase [91] G. Gutierrez, B. Johansson, Molecular dynamics study of structural properties of
transformations in Al2O3 film on oxidation kinetics of aluminum powders, amorphous Al2O3, Phys. Rev. B 65 (2002).
Combust. Flame 140 (2005) 310–318. [92] V.V. Hoang, Molecular dynamics study on structure and properties of liquid and
[60] L.P.H. Jeurgens, W.G. Sloof, F.D. Tichelaar, E.J. Mittemeijer, Structure and mor- amorphous Al2O3, Phys. Rev. B 70 (2004).
phology of aluminium-oxide films formed by thermal oxidation of aluminium, [93] H.J. Chang, Y.M. Choi, K.J. Kong, B.H. Ryu, Atomic and electronic structures of
Thin Solid Films 418 (2002) 89–101. amorphous Al2O3, Chem. Phys. Lett. 391 (2004) 293–296.
[61] F. Reichel, L.P.H. Jeurgens, G. Richter, J. Mittemeijer, Amorphous versus crys- [94] H. Momida, T. Hamada, Y. Takagi, T. Yamamoto, T. Uda, T. Ohno, Theoretical
talline state for ultrathin Al2O3 overgrowths on Al substrates, J. Appl. Phys. 103 study on dielectric response of amorphous alumina, Phys. Rev. B 73 (2006).
(2008). [95] E.A. Chagarov, A.C. Kummel, Generation of realistic amorphous Al2O3 And ZrO2
[62] P. Brand, P. Dietzmann, Amorphous precursors of alpha-Al2O3 - An electron-op- samples by hybrid classical and first-principle molecular dynamics simulations,
tical study, Cryst. Res. Technol. 27 (1992) 529–534. ECS Trans. 16 (2008) 773–785.
[63] J. Alkebro, S. Begin-Colin, A. Mocellin, R. Warren, Mechanical alloying of alu- [96] S. Davis, G. Gutierrez, Structural, elastic, vibrational and electronic properties of
mina-yttria powder mixtures, J. Eur. Ceram. Soc. 20 (2000) 2169–2174. amorphous Al2O3 from ab initio calculations, J. Phys.-Condens. Mat. 23 (2011).
[64] D.D. Ragan, T. Mates, D.R. Clarke, Effect of yttrium and erbium ions on epitaxial [97] M. Tane, S. Nakano, R. Nakamura, H. Ogi, M. Ishimaru, H. Kimizuka, H. Nakajima,
phase transformations in alumina, J. Am. Ceram. Soc. 86 (2003) 541–545. Nanovoid formation by change in amorphous structure through the annealing of
[65] F. Abbattista, A. Delmastro, G. Gozzelino, D. Mazza, M. Vallino, G. Busca, amorphous Al2O3 thin films, Acta Mater. 59 (2011) 4631–4640.
V. Lorenzelli, Effect of phosphate ions on the surface-chemistry and microstructure [98] V.V. Le, V.H. Nguyen, K.H. Pham, The structure and mechanical properties in
of amorphous alumina, J. Chem. Soc. Faraday T. 86 (1990) 3653–3658. amorphous alumina under pressure, Comput. Mater. Sci. 79 (2013) 110–117.
[66] Z.F. Wang, W.Y. Wu, X. Bian, Y.F. Wu, Synthesis and characterization of amor- [99] S.P. Adiga, P. Zapol, L.A. Curtiss, Atomistic simulations of amorphous alumina
phous Al2O3 and gamma-Al2O3 by spray pyrolysis, Green Process. Synth. 5 surfaces, Phys. Rev. B 74 (2006).
(2016) 305–310. [100] M.E. Prokhorskii, A.D. Fofanov, L.A. Aleshina, E.A. Nikitina, Structure of amor-
[67] H. Bolvardi, M.T. Baben, F. Nahif, D. Music, V. Schnabel, K.P. Shaha, S. Mraz, phous oxide Al2O3: results of a molecular-dynamics experiment, Crystallogr. Rep.
J. Bednarcik, J. Michalikova, J.M. Schneider, Effect of Si additions on thermal 49 (2004) 631–634.
stability and the phase transition sequence of sputtered amorphous alumina thin [101] T.W. Simpson, Q.Z. Wen, N. Yu, D.R. Clarke, Kinetics of the amorphous
films, J. Appl. Phys. 117 (2015). - > gamma - > alpha transformations in aluminum oxide: effect of crystal-
[68] N. Yu, T.W. Simpson, P.C. McIntyre, M. Nastasi, I.V. Mitchell, Doping effects on lographic orientation, J. Am. Ceram. Soc. 81 (1998) 61–66.
the kinetics of solid-phase epitaxial-growth of amorphous alumina thin-films on [102] P. Eklund, M. Sridharan, G. Singh, J. Bottiger, Thermal stability and phase
sapphire, Appl. Phys. Lett. 67 (1995) 924–926. transformations of gamma-/Amorphous-Al2O3 thin films, Plasma Process. Polym.
[69] N. Kim, R. Bassiri, M.M. Fejer, J.F. Stebbins, The structure of ion beam sputtered 6 (2009) S907–S911.
amorphous alumina films and effects of Zn doping: High-resolution Al-27 NMR, J. [103] J. Beretka, M.J. Ridge Dehydration products of aluminium hydroxides, J. Chem.
Non-Cryst. Solids 405 (2014) 1–6. Soc. A 0 (1967) 2016–2019.
[70] W.H. Zachariasen, The atomic arrangement in glass, J. Am. Chem. Soc. 54 (1932) [104] F. Di Fonzo, D. Tonini, A.L. Bassi, C.S. Casari, M.G. Beghi, C.E. Bottani, D. Gastaldi,
3841–3851. P. Vena, R. Contro, Growth regimes in pulsed laser deposition of aluminum oxide
[71] A. Kirfel, K. Eichhorn, Accurate structure analysis with synchrotron radiation. The films, Appl. Phys. A Mater. Sci. Process. 93 (2008) 765–769.
electron density in Al2O3 and Cu2O, Acta Cryst A46 (1990) 271–284. [105] G. Balakrishnan, P. Kuppusami, S.T. Sundari, R. Thirumurugesan, V. Ganesan,
[72] L. Smrcok, V. Langer, J. Krestan, gamma-Alumina: A single-crystal X-ray diffrac- E. Mohandas, D. Sastikumar, Structural and optical properties of gamma-alumina
tion study, Acta Crystallogr. C 62 (2006) I83–I84. thin films prepared by pulsed laser deposition, Thin Solid Films 518 (2010)
[73] J.H. Kwak, J.Z. Hu, A. Lukaski, D.H. Kim, J. Szanyi, C.H.F. Peden, Role of pen- 3898–3902.
tacoordinated Al3+ ions in the high temperature phase transformation of gamma- [106] E. Beche, F. Fournel, V. Larrey, F. Rieutord, C. Morales, A.M. Charvet, F. Madeira,
Al2O3, J. Phys. Chem. C 112 (2008) 9486–9492. G. Audoit, J.M. Fabbri, Direct bonding mechanism of ALD-Al2O3 thin films, ECS J.
[74] J.H. Kwak, J. Hu, D. Mei, C.-W. Yi, D.H. Kim, C.H.F. Peden, L.F. Allard, J. Szanyi, Solid State Sci. Technol. 4 (2015) P171–P175.
Coordinatively unsaturated Al3+ centers as binding sites for active catalyst [107] L. Parfitt, M. Goldiner, J.W. Jones, G.S. Was, Residual stresses in amorphous
phases of platinum on gamma-Al2O3, Science 325 (2009) 1670–1673. alumina films synthesized by ion-beam-assisted deposition, J. Appl. Phys. 77

13
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

(1995) 3029–3036. [139] J.H. Roque-Ruiz, E.A. Cabrera-Ontiveros, G. Gonzalez-Garcia, S.Y. Reyes-Lopez,
[108] C. Cibert, H. Hidalgo, C. Champeaux, P. Tristant, C. Tixier, J. Desmaison, Thermal degradation of aluminum formate sol-gel; synthesis of alpha-alumina and
A. Catherinot, Properties of aluminum oxide thin films deposited by pulsed laser characterization by H-1, C-13 and Al-27 MAS NMR and XRD spectroscopy, Results
deposition and plasma enhanced chemical vapor deposition, Thin Solid Films 516 Phys. 6 (2016) 1096–1102.
(2008) 1290–1296. [140] L.H. Qu, C.Q. He, Y. Yang, Y.L. He, Z.M. Liu, Hydrothermal synthesis of alumina
[109] C.H. Lin, H.L. Wang, M.H. Hon, Preparation and characterization of aluminum nanotubes templated by anionic surfactant, Mater. Lett. 59 (2005) 4034–4037.
oxide films by plasma enhanced chemical vapor deposition, Surf. Coat. Technol. [141] E.O. Filatova, A.S. Konashuk, Interpretation of the changing the band gap of
90 (1997) 102–106. Al2O3 depending on its crystalline form: connection with different local symme-
[110] A. Gleizes, M. Sovar, D. Samélor, C. Vahlas, Low temperature MOCVD-processed tries, J. Phys. Chem. C 119 (2015) 20755–20761.
alumina coatings, Adv. Sci. Technol. 45 (2006) 1184–1193. [142] P. Nayar, A. Khanna, D. Kabiraj, S.R. Abhilash, B.D. Beake, Y. Losset, B. Chen,
[111] A.N. Gleizes, C. Vahlas, M.-M. Sovar, D. Samelor, M.-C. Lafont, CVD-Fabricated Structural, optical and mechanical properties of amorphous and crystalline alu-
aluminum oxide coatings from aluminum tri-iso-propoxide: Correlation between mina thin films, Thin Solid Films 568 (2014) 19–24.
processing conditions and composition, Chem. Vap. Depos. 13 (2007) 23–29. [143] H.L. Wang, C.H. Lin, M.H. Hon, The dependence of hardness on the density of
[112] H. Vergnes, D. Samelor, A.N. Gleizes, C. Vahlas, B. Caussat, Local kinetic modeling amorphous alumina thin films by PECVD, Thin Solid Films 310 (1997) 260–264.
of aluminum oxide metal-organic CVD from aluminum Tri-isopropoxide, Chem. [144] M. Sridharan, M. Sillassen, J. Bottiger, J. Chevallier, H. Birkedal, Pulsed DC
Vap. Depos. 17 (2011) 181. magnetron sputtered Al2O3 films and their hardness, Surf. Coat. Technol. 202
[113] K. Okuyama, Y. Kousaka, N. Tohge, S. Yamamoto, J.J. Wu, R.C. Flagan, (2007) 920–924.
J.H. Seinfeld, Production of ultrafine metal-oxide aerosol particles by thermal [145] I.N. Reddy, V.R. Reddy, N. Sridhara, S. Basavaraja, A.K. Sharma, A. Dey, Optical
decomposition of meral alkoxide vapors, AICHE J. 32 (1986) 2010–2019. and microstructural characterisations of pulsed rf magnetron sputtered alumina
[114] P. Wong, M. Robinson, Chemical vapor deposition of polycrystalline Al2O3, J. Am. thin film, J. Mater. Sci. Technol. 29 (2013) 929–936.
Ceram. Soc. 53 (1970) 617–621. [146] J. Houska, J. Blazek, J. Rezek, S. Proksova, Overview of optical properties of
[115] T.C. Chou, D. Adamson, J. Mardinly, T.G. Nieh, Microstructural evolution and Al2O3 films prepared by various techniques, Thin Solid Films 520 (2012)
properties of nanocrystalline alumina made by reactive sputtering deposition, 5405–5408.
Thin Solid Films 205 (1991) 131–139. [147] B.F. Hu, M.W. Yao, P.F. Yang, W. Shan, X. Yao, Preparation and dielectric prop-
[116] B.G. Segda, M. Jacquet, J.P. Besse, Elaboration, characterization and dielectric erties of dense and amorphous alumina film by sol-gel technology, Ceram. Int. 39
properties study of amorphous alumina thin films deposited by r.f. magnetron (2013) 7613–7618.
sputtering, Vacuum 62 (2001) 27–38. [148] F. Evangelisti, M. Stiefel, O. Guseva, R.P. Nia, R. Hauert, E. Hack, L.P.H. Jeurgens,
[117] H. Brune, J. Wintterlin, J. Trost, G. Ertl, J. Wiechers, R.J. Behm, Interaction of F. Ambrosio, A. Pasquarello, P. Schmutz, C. Cancellieri, Electronic and structural
oxygen with Al(111) studied by scanning-tunneling microscopy, J. Chem. Phys. 99 characterization of barrier-type amorphous aluminium oxide, Electrochim. Acta
(1993) 2128–2148. 224 (2017) 503–516.
[118] W.H. Krueger, S.R. Pollack, The initial oxidation of aluminum thin films at room [149] Y. Adraider, Y.X. Pang, F. Nabhani, S.N. Hodgson, M.C. Sharp, A. Al-Waidh, Laser-
temperature, Surf. Sci. 30 (1972) 263–279. induced deposition of alumina ceramic coating on stainless steel from dry thin
[119] E.E. Hueber, C.T. Kirk, Work function changes due to the chemisorption of water films for surface modification, Ceram. Int. 40 (2014) 6151–6156.
and oxygen on aluminum, Surf. Sci. 5 (1966) 447–465. [150] A. Aryasomayajula, N.X. Randall, M.H. Gordon, D. Bhat, Tribological and me-
[120] F. Jona, Preparation and properties of clean surfaces of aluminum, J. Phys. Chem. chanical properties of physical vapor deposited alpha alumina thin film coating,
Solids 28 (1967) 2155–2160. Thin Solid Films 517 (2008) 819–823.
[121] M.W. Roberts, B.R. Wells, Chemisorption of oxygen by aluminium, Surf. Sci. 15 [151] R. Cremer, M. Witthaut, D. Neuschutz, G. Erkens, T. Leyendecker, M. Feldhege,
(1969) 325–332. Comparative characterization of alumina coatings deposited by RF, DC and pulsed
[122] G. Litrico, P. Proulx, J.B. Gouriet, P. Rambaud, Controlled oxidation of aluminum reactive magnetron sputtering, Surf. Coat. Technol. 120 (1999) 213–218.
nanoparticles, Adv. Powder Technol. 26 (2015) 1–7. [152] X. Nie, E.I. Meletis, J.C. Jiang, A. Leyland, A.L. Yerokhin, A. Matthews, Abrasive
[123] C.D. Kong, D. Yu, S.Q. Li, Q. Yao, Mechanism and modelling of aluminium na- wear/corrosion properties and TEM analysis of Al2O3 coatings fabricated using
noparticle oxidation coupled with crystallisation of amorphous Al2O3 shell, plasma electrolysis, Surf. Coat. Technol. 149 (2002) 245–251.
Combust. Theor. Model. 20 (2016) 296–312. [153] T.V. Perevalov, O.E. Tereshenko, V.A. Gritsenko, V.A. Pustovarov, A.P. Yelisseyev,
[124] M. Schoenitz, B. Patel, O. Agboh, E.L. Dreizin, Oxidation of aluminum powders at C. Park, J.H. Han, C. Lee, Oxygen deficiency defects in amorphous Al2O3, J. Appl.
high heating rates, Thermochim. Acta 507-08 (2010) 115–122. Phys. 108 (2010).
[125] P.C. Snijders, L.P.H. Jeurgens, W.G. Sloof, Structure of thin aluminium-oxide films [154] R.H. French, Electronic band-structure of Al2O3, with comparison to AlON and
determined from valence band spectra measured using XPS, Surf. Sci. 496 (2002) AlN, J. Am. Ceram. Soc. 73 (1990) 477–489.
97–109. [155] W.J.W. Gignac, R.S. Kowalczyk, S. P, Valence- and conduction-band structure of
[126] E. Rocca, D. Vantelon, S. Reguer, F. Mirambet, Structural evolution in nanoporous the sapphire (1102) surface, Phys. Rev. B 32 (1985) 1237–1247.
anodic aluminium oxide, Mater. Chem. Phys. 134 (2012) 905–911. [156] B. Ealet, M.H. Elyakhloufi, E. Gillet, M. Ricci, Electronic and crystallographic
[127] R. Boidin, T. Halenkovic, V. Nazabal, L. Benes, P. Nemec, Pulsed laser deposited structure of gamma-alumina thin-films, Thin Solid Films 250 (1994) 92–100.
alumina thin films, Ceram. Int. 42 (2016) 1177–1182. [157] J.-H. Park, D.S. Park, B.-D. Hahn, J.-J. Choi, J. Ryu, S.-Y. Choi, J. Kim, W.-H. Yoon,
[128] Y. Berkovich, A. Aserin, E. Wachtel, N. Garti, Preparation of amorphous aluminum C. Park, Effect of raw powder particle size on microstructure and light transmit-
oxide-hydroxide nanoparticles in amphiphilic silicone-based copolymer micro- tance of alpha-alumina films deposited by granule spray in vacuum, Ceram. Int. 42
emulsions, J. Colloid Interface Sci. 245 (2002) 58–67. (2016) 3584–3590.
[129] Z.F. Zhu, S. Cheng, H. Liu, Size-controlled synthesis of alumina nanoparticles [158] G.A. Niklasson, Optical properties of cobalt-dopedamorphous aluminum oxide, J.
through an additive-free reverse cation-anion double hydrolysis method, Mater. Appl. Phys. 57 (1985) 157–158.
Lett. 161 (2015) 720–723. [159] L. Fu, X. Li, M. Liu, H. Yang, Insights into the nature of Cu doping in amorphous
[130] J.T. Wang, L. Ge, Z.Q. Li, L. Li, Q. Guo, J.G. Li, Facile size-controlled synthesis of mesoporous alumina, J. Mater. Chem. A 1 (2013) 14592–14605.
well-dispersed spherical amorphous alumina nanoparticles via homogeneous [160] M.J. Choi, J.Y. Jung, M.-J. Park, J.-W. Song, J.-H. Lee, J.H. Bang, Long-term
precipitation, Ceram. Int. 42 (2016) 8545–8551. durable silicon photocathode protected by a thin Al2O3/SiOx layer for photo-
[131] A.B. Khatibani, S.M. Rozati, Synthesis and characterization of amorphous alu- electrochemical hydrogen evolution, J. Mater. Chem. A 2 (2014) 2928–2933.
minum oxide thin films prepared by spray pyrolysis: effects of substrate tem- [161] R. Fan, W. Dong, L. Fang, F. Zheng, X. Su, S. Zou, J. Huang, X. Wang, M. Shen,
perature, J. Non-Cryst. Solids 363 (2013) 121–133. Stable and efficient multi-crystalline n(+)p silicon photocathode for H-2 pro-
[132] Z. Su, M.W. Yao, M.X. Li, W.B. Gao, Q.X. Li, Q. Feng, X. Yao, A novel and simple duction with pyramid-like surface nanostructure and thin Al2O3 protective layer,
aluminium/sol-gel-derived amorphous aluminium oxide multilayer film with high Appl. Phys. Lett. 106 (2015).
energy density, J. Mater. Chem. C 6 (2018) 5616–5623. [162] I.A. Digdaya, G.W.P. Adhyaksa, B.J. Trzesniewski, E.C. Garnett, W.A. Smith,
[133] C.K. Perkins, R.H. Mansergh, J.C. Ramos, C.E. Nanayakkara, D.-H. Park, Interfacial engineering of metal-insulator-semiconductor junctions for efficient
S. Goberna-Ferron, L.B. Fullmer, J.T. Arens, M.T. Gutierrez-Higgins, Y.R. Jones, and stable photoelectrochemical water oxidation, Nat. Commun. 8 (2017).
J.I. Lopez, T.M. Rowe, D.M. Whitehurst, M. Nyman, Y.J. Chabal, D.A. Keszler, [163] W. Yang, K. Song, Y. Jung, S. Jeong, J. Moon, Solution-deposited Zr-doped AlOx
Low-index, smooth Al2O3 films by aqueous solution process, Opt. Mater. Express 7 gate dielectrics enabling high-performance flexible transparent thin film transis-
(2017) 273–280. tors, J. Mater. Chem. C 1 (2013) 4275–4282.
[134] K. Vanbesien, P. De Visschere, P.F. Smet, D. Poelman, Electrical properties of [164] Q. Feng, M. Yao, Z. Su, X. Yao, Significantly enhanced energy density of amor-
Al2O3 films for TFEL-devices made with sol-gel technology, Thin Solid Films 514 phous alumina thin films via silicon and magnesium co-doping, Ceram. Int. 44
(2006) 323–328. (2018) 11160–11165.
[135] A.C. Dillon, A.W. Ott, J.D. Way, S.M. George, Surface chemistry of Al2O3 de- [165] M. Yao, R. Xiao, Y. Peng, J. Chen, B. Hu, X. Yao, The influence of titanium doping
position using Al(CH3)3 and H2O in a binary reaction sequence, Surf. Sci. 322 on the electric properties of amorphous alumina films prepared by sol-gel tech-
(1995) 230–242. nology, J. Sol-Gel Sci. Technol. 74 (2015) 39–44.
[136] A.J. Fanelli, J.V. Burlew, Preparation of fine alumina powder in alcohol, J. Am. [166] P. Zou, M. Yao, J. Chen, Y. Peng, X. Yao, Leakage current and dielectric break-
Ceram. Soc. 69 (1986) C174–C175. down in lanthanum doped amorphous aluminum oxide films prepared by sol-gel,
[137] K.R. Murali, P. Thirumoorthy, Characteristics of sol-gel deposited alumina films, J. Ceram. Int. 42 (2016) 4120–4125.
Alloys Compd. 500 (2010) 93–95. [167] M. Fukuhara, T. Kuroda, F. Hasegawa, T. Hashida, E. Kwon, K. Konno, Amorphous
[138] Y. Adraider, S.N.B. Hodgson, M.C. Sharp, Z.Y. Zhang, F. Nabhani, A. Al-Waidh, aluminum-oxide supercapacitors, EPL 123 (2018).
Y.X. Pang, Structure characterisation and mechanical properties of crystalline [168] D.V. Kysil, A.V. Vasin, S.V. Sevostianov, V.Y. Degoda, V.V. Strelchuk, V.M. Naseka,
alumina coatings on stainless steel fabricated via sol-gel technology and fibre laser Y.P. Piryatinski, V.A. Tertykh, A.N. Nazarov, V.S. Lysenko, Formation and lumi-
processing, J. Eur. Ceram. Soc. 32 (2012) 4229–4240. nescent properties of Al2O3:SiOC nanocomposites on the Base of alumina

14
A. Mavrič, et al. Journal of Non-Crystalline Solids 521 (2019) 119493

nanoparticles modified by phenyltrimethoxysilane, Nanoscale Res. Lett. 12 Matjaz Valant is Head of Materials Research Laboratory
(2017). and Dean of School for Environmental Sciences at the
[169] H. Nasu, D. Hirota, K. Inoue, T. Hashimoto, A. Ishihara, Luminescent properties of University of Nova Gorica in Slovenia. He is Adjunct
amorphous Al2O3 prepared by sol-gel method, J. Ceram. Soc. Jap. 116 (2008) Professor at Institute of Fundamental and Frontier Sciences,
835–836. University of Electronic Science and Technology of China
[170] J. Reyes, D.Y. Medina, M. Aguilar, M.A. Barron, E. Garfias, A.J. Morales, Red, and a visiting professor at London South Bank University.
white and blue light emission fromEuropium doped Al2O3 confined into a silica His main research interest is oxide semiconductors, di-
matrix, Open J. Appl. Sci. 8 (2018) 338–345. electrics, magnetoelectric, topological insulators, and elec-
[171] J.C.V. Brito, N.V. Gaponenko, K.S. Sukalin, T.F. Raichenok, S.A. Tikhomorov, trocaloric materials as well as the synthesis of new nanos-
X. Wang, Z. Cheng, N.I. Kargin, Europium luminescence from amorphous yttrium tructured functional materials.
alumina films on fused silica substrates, IOP Conf. Series: Mater. Sci. Eng. 475
(2019) 012019.
[172] K. Smits, D. Millers, A. Zolotarjovs, R. Drunka, M. Vanks, Luminescence of Eu ion
in alumina prepared by plasma electrolytic oxidation, Appl. Surf. Sci. 337 (2015)
166–171.
[173] T. Ishizaka, Y. Kurokawa, Optical properties of rare-earth ion (Gd3+, Ho3+, Chunhua Cui has been engaged as Professor of Molecular
Pr3+, Sm3+, Dy3+ and Tm3+)-doped alumina films prepared by the sol-gel Electrochemistry since 2017 in the Institute of Fundamental
method, J. Lumin. 92 (2000) 57–63. and Frontier Sciences at the University of Electronic Science
[174] T. Ishizaka, Y. Kurokawa, T. Makino, Y. Segawa, Optical properties of rare earth and Technology of China. Prior to his appointment, he was
ion (Nd3+, Er3+ and Tb3+)-doped alumina films prepared by the sol-gel an independent researcher within URPP LightChEC of the
method, Opt. Mater. 15 (2001) 293–299. University of Zurich. His group concentrates on molecular
[175] S. Tanabe, T. Ohyagi, T. Hanada, N. Soga, Upconversion and Local Structure of electrochemistry, semiconductor electrochemistry, in-situ
Er3+ Doped Aluminate Glasses, J. Ceram. Soc. Jp. 101 (1993) 78–83. spectroelectrochemistry, and electrochemical micro-reac-
[176] T.K. Phung, A. Lagazzo, M.A.R. Crespo, V.S. Escribano, G. Busca, A study of tions.
commercial transition aluminas and of their catalytic activity in the dehydration of
ethanol, J. Catal. 311 (2014) 102–113.
[177] J.H. Kang, L.D. Menard, R.G. Nuzzo, A.I. Frenkel, Unusual non-bulk properties in
nanoscale materials: Thermal metal-metal bond contraction of gamma-alumina-
supported Pt catalysts, J. Am. Chem. Soc. 128 (2006) 12068–12069.
[178] Z. Wang, Y. Jiang, O. Lafon, J. Trebosc, K.D. Kim, C. Stampfl, A. Baiker, J.-
P. Amoureux, J. Huang, Bronsted acid sites based on penta-coordinated aluminum Zhiming M. Wang received his PhD degree in Condensed
species, Nat. Commun. 7 (2016). Matter Physics from the Institute of Semiconductors at the
[179] Z. Wang, Y. Jiang, F. Jin, C. Stampfl, M. Hunger, A. Baiker, J. Huang, Strongly Chinese Academy of Sciences in Beijing, China in 1998. He
enhanced acidity and activity of amorphous silica–alumina by formation of pen- is now a Professor of the National 1000-Talent Program,
tacoordinated AlV species, J. Catal. 372 (2019) 1–7. working in the University of Electronic Science and
Technology of China. His research on the optoelectronic
properties of low-dimensional semiconductor nanos-
Andraž Mavrič received his PhD at the University of Nova tructures and corresponding applications in photovoltaic
Gorica, Slovenia where he developed a method for high- devices.
temperature stabilization of bulk amorphous alumina. He is
currently a postdoctoral researcher at Molecular
Electrochemistry Laboratory, Institute of Fundamental and
Frontier Sciences, University of Electronic Science and
Technology of China. He works in the area of electro-
catalysis and is actively involved in the development of
catalysts and materials for corrosion protection.

15

Вам также может понравиться