You are on page 1of 11

Phase Transitions

A Multinational Journal

ISSN: 0141-1594 (Print) 1029-0338 (Online) Journal homepage: https://www.tandfonline.com/loi/gpht20

Synthesis and structural characterization of


alumina nanoparticles

Puneet Kaur, Atul Khanna, Nirmal Kaur, Priyanka Nayar & Banghao Chen

To cite this article: Puneet Kaur, Atul Khanna, Nirmal Kaur, Priyanka Nayar & Banghao Chen
(2020): Synthesis and structural characterization of alumina nanoparticles, Phase Transitions, DOI:
10.1080/01411594.2020.1765245

To link to this article: https://doi.org/10.1080/01411594.2020.1765245

Published online: 12 May 2020.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=gpht20
PHASE TRANSITIONS
https://doi.org/10.1080/01411594.2020.1765245

Synthesis and structural characterization of alumina nanoparticles


Puneet Kaura, Atul Khannaa, Nirmal Kaura, Priyanka Nayara and Banghao Chenb
a
Department of Physics, Guru Nanak Dev University, Amritsar, India; bDepartment of Chemistry & Biochemistry,
Florida State University, Tallahassee, FL, USA

ABSTRACT ARTICLE HISTORY


Alumina nanoparticles were synthesized by thermal dehydration of Al Received 29 November 2019
(OH)3 (gibbsite), which shows the sequential formation of AlOOH Accepted 27 April 2020
(bohemite), γ, δ, θ, and α-alumina phases upon heat treatment in the
KEYWORDS
temperature range: 150–1150°C. Alumina nanoparticles were Alumina nanoparticles;
characterized by X-ray diffraction and 27Al MAS-NMR studies to annealing; 27Al-MAS NMR; Al
investigate the structural transitions, in particular, the changes in Al—O —O speciation; transition
speciation as a function of heat treatment. γ-alumina is produced by alumina phases
annealing bohemite at 400°C, and transforms into α-alumina via the
formation of δ and θ-alumina as the intermediate phases. The formation
of θ and α-alumina starts at 1020°C and the concentration of these
phases increases upon annealing at 1100°C, while the concentrations of
γ and δ-alumina decreases steadily. The heat treatment at 1150°C
produces α-phase (84%) and the remnant θ-phase (16%). 27Al MAS-NMR
revealed the existence of two types of Al—O speciation i.e. AlO4 and
AlO6 structural units whose relative concentration depends upon the
transition alumina phases.

1. Introduction
Aluminum oxide is one of the most important ceramic materials with several outstanding properties
such as high electrical insulation, high hot-hardness, strength, wear-resistance, catalytic activity, stiff-
ness, thermal conductivity, good dielectric properties, chemical inertness, high melting temperature
low friction and high optical band gap [1–7]. Aluminum oxide exists in several transitional phases
such as η, γ, δ, θ, β, κ, χ and α-alumina [5, 8–10]. The crystal structure of transitional aluminas has
been associated with different ordering and occupancies of Al3+ cations at octahedral and tetrahedral
sites in different proportions of occupation [11]. Cai et al. theoretically investigated the migration of
aluminum ions between the adjacent unoccupied tetrahedral or octahedral interstitial sites with oxy-
gen anions remaining fixed during γ to θ-alumina transformation through detailed atomistic
description and also described the way to enhance the thermal stability of γ-alumina at high temp-
eratures for its use in catalytic applications [12]. The structure of aluminum oxide based on fcc pack-
ing of oxygen includes γ (cubic or tetragonal), η (cubic), θ (monoclinic) and δ (either tetragonal or
orthorhombic) phases, whereas hcp packing includes α (trigonal), κ (orthorhombic), and χ (hexago-
nal) phases [9, 10, 13, 14]. γ-alumina is stable up to 900°C but gets irreversibly transformed into the
thermodynamically stable α-alumina phase upon annealing at higher temperatures through a series
of metastable phase transformations: γ → δ → θ→ α [5, 9, 11, 12, 15].
Petlinger et al. reported the structural description of these alumina phases [14]. Al(OH)3, com-
monly known as gibbsite, consists of layers of AlO6 octahedra that share one edge along the plane
whereby each oxygen atom is bonded to a hydrogen atom. Bohemite or aluminum oxy-hydroxide

CONTACT Atul Khanna atul.phy@gndu.ac.in


© 2020 Informa UK Limited, trading as Taylor & Francis Group
2 P. KAUR ET AL.

(AlOOH) contains oxygen and aluminum ions in a double layer of octahedra with hydrogen
atoms located in a zig-zag fashion [14]. The structure of γ-alumina is generally described as hav-
ing defective cubic spinel-like structure, whose oxygen lattice is built by cubic closed packing of
oxygen layers with vacancies existing on cationic sites with AlO6 and AlO4 structural units
(denoted as Al[6] and Al[4] respectively) [16, 17]. θ-alumina consists of equally distributed alumi-
num ions at the octahedral and tetrahedral sites i.e. Al[4] = Al[6] = 50% [14, 18]. α-alumina (cor-
undum) contains only Al[6], and can be described as having hexagonal close packing of
spheres in which aluminum ions occupy two-thirds of octahedral interstitial sites [10, 17].
Further, the different alumina phases have different applications based on their physical proper-
ties. Ultrafine particle size, large surface area and high porosity of γ-alumina serves applications as
catalysts, adsorbents, coatings, soft abrasives, catalysts in industrial activities like petroleum refin-
ing, alcohol dehydration, oxidation of CO into CO2 and reduction of automotive pollutants such
as NOx [9, 12, 19]. Corundum is a widely used ceramic due to its high resistance to electricity,
temperature and corrosion [14].
Nuclear Magnetic Resonance (NMR) employing Magic Angle Spinning (MAS) is a commonly
used structural probe to accurately determine the local environments of the atoms. The distribution
of aluminum ions at the tetrahedral and octahedral sites can be determined precisely by 27Al MAS-
NMR and this technique has provided the knowledge of structural environments of aluminum ions
around oxygen lattice [11, 15, 18, 20–23]. In the ideal spinel structure (γ-alumina), octahedrally
coordinated cations are twice as many as the tetrahedral ones [18]. Lee et al. found by 27Al MAS-
NMR and computational studies that ∼70% of Al ions are octahedrally coordinated in γ-alumina
[22]. Ansell et al. studied the changes in Al—O coordination from octahedral to tetrahedral upon
melting of Al2O3 [24].
Amorphous aluminum oxide has been found to contain penta-coordinated aluminum ions i.e.
AlO5 (denoted as Al[5]) along with Al[4] and Al[6] units. Three peaks are centered at ∼70, 40 and
7 ppm in NMR spectra due to tetra, penta and hexa-coordinations of Al ions respectively [6, 25,
26]. Nayar et al. reported that upon annealing of amorphous alumina thin films, the concentration
of Al[5] units decreases, while the concentration of Al[6] increases, resulting in the crystallization of
amorphous films. Further amorphous alumina films are found to contain Al[5] and Al[4] units in the
ratio 1:2 [26]. Shi et al. determined the average Al—O coordination number by high energy X-ray
and neutron diffraction and reported that amorphous alumina thin films contain 51% Al[4], 41%
Al[5] and 8% Al[6] whereas supercooled alumina at 1817 K contains 0.14% AlO3 (Al[3]) in addition
to 55.9% Al[4], 38.7% Al[5] and 5.23% Al[6] while the molten alumina at 2700 K contains 63% Al[4]
and 32% Al[5] units with small quantities of Al[3] (2%) and Al[6] (3%). Thus, the concentration of
Al[5] in amorphous alumina films is found to be higher than that in molten and supercooled alumina
[27].
The synthesis of crystalline Al2O3 nanoparticles has been reported by various techniques includ-
ing sol-gel [2], precipitation method [3], hydrothermal method [4], spray pyrolysis [28], laser abla-
tion [29], microwave-assisted solvothermal [30] and the combustion method [31]. Tavakoli et al.
prepared amorphous alumina nanoparticles using gas phase condensation setup designed for laser
evaporation of Al2O3 target [6]. In the present study, ultra-fine grained aluminum oxide nanopar-
ticles were synthesized using aluminum metal powder as a precursor. The evolution of different
phases of aluminum oxide under heat treatment has been studied by powder X-Ray diffraction
(XRD) and 27Al-Magic Angle Spinning-Nuclear Magnetic Resonance (27Al-MAS NMR) spec-
troscopy. NMR results are found to be consistent with the XRD findings.

2. Experimental methods and characterization


Aluminum metal powder (99.99%) was mixed with dilute hydrochloric acid to produce a clear sol-
ution of AlCl3. Hydrogen gas evolved out with fizzing sound, accompanied by the liberation of a lot
PHASE TRANSITIONS 3

of heat, by the following chemical reaction:


2 Al + 6 HCl  2 AlCl3 + 3H2  (1)
Ammonium hydroxide solution was then added slowly to the solution of AlCl3, and white gela-
tinous precipitates of aluminum hydroxide were obtained by the following reaction:
3NH4 OH + AlCl3  Al(OH)3  + 3 NH4 Cl (2)
The precipitates of Al(OH)3 were then washed with distilled water several times, and filtered out
on a filter paper. These precipitates were dried at 100°C for 30 min and then heated at 150°C for
90 min. Further, heat treatment was given to the nanoparticles sequentially at 250°C, 400°C, 600°
C, 800°C, 950°C, 1020°C, 1100°C, 1150°C for 6 h each in air, to study the effects of heating on
the structural transformations of aluminum oxide.
X-ray diffraction studies were carried out with Bruker D8 Focus X-ray diffractometer with Cu-
Kα1,2 radiations of the wavlength: λ = 1.5406 and 1.5444 Å in the 2θ range of 10°–70°. 27Al-MAS
NMR spectra were measured on a Bruker AVIII HD NMR spectrometer operating at a magnetic
field of 11.74 T with a 4 mm Bruker MAS probe at a Larmor frequency of 130.38 MHz for 27Al
nuclei. Sample spinning rate was 14 kHz. Short RF pulses (<15°) with recycle delay of 1s
were used. Spectra were collected after 2048-4096 scans and referenced to 1M Al(NO3)3 (aq.) at
δ = 0 ppm.

3. Results and discussion


3.1. Phase formation
The XRD patterns show the formation of different phases of alumina under the effect of heat
treatment. Figure 1(a) shows the evolution of aluminum oxide phases from the initial
aluminum hydroxide after heat treatment in the temperature range of 150°C–600°C. Upon
heating the Al(OH)3 precipitates at 150°C for 90 min, sharp XRD peaks at 22.8°, 32.7°, 40.2°,
46.8°, 52.8° are observed (denoted as $ in Figure 1(a)) which match with the Joint Committee on
Powder Diffraction Standards (JCPDS) #85-1049 of gibbsite i.e. Al(OH)3. The peaks centered at
58.3°, 68.4° can be due to both AlOOH (JCPDS #83-2384) and Al(OH)3. With the increase in temp-
erature to 250°C, besides the above-mentioned peaks, broad peaks due to AlOOH are also detected at
28.4°, 38.3°, 49.1°, 64.9° (denoted as * in Figure 1(a)). The decrease in the intensity of peaks due to Al
(OH)3 confirmed the reduction in the concentration of hydroxyl ions in the sample with heat treat-
ment at 250°C.
The progressive dehydration and removal of hydroxyl groups lead to the formation of a mixture
of γ-Al2O3 (JCPDS #10-0425) and δ-Al2O3 phases (JCPDS #46-1131) at 400°C, whose XRD patterns
exhibit broad peaks centered at 37.0°, 45.7° and 66.3°. On further increasing the annealing tempera-
ture to 600°C, the broad peaks are detected at 37.4°, 39.6°, 45.8°, 60.7° and 66.7° and are indexed to
both γ and δ-alumina phases (Figure 1(a)).
Figure 1(b) shows the XRD patterns of the samples heat treated in the higher temperature range:
800°C to 1150°C. The sample heated at 800°C shows peaks at 31.9° (γ/δ), 36.9° (δ), 45.8° (γ/δ), 61.0°
(δ) and 66.9° (γ/δ). In the sample annealed at 950°C, XRD peaks are observed at 32.6° (δ), 37.2° (δ),
39.7° (γ/δ), 45.7° (γ/δ), 60.7° (γ) and 67.0° (γ/δ). The enhancement in the intensity and sharpening of
peaks is observed with each heat treatment step from 400°C to 950°C and confirms the growth of the
size of nanoparticles.
After annealing treatment at 1020°C, the sample is found to be a mixture of θ, α, γ and δ alumina
phases. The growth of peaks at 25.6°, 35.1°, 43.2°, 52.5° and 57.5° mark the beginning of the for-
mation of α-alumina phase (JCPDS #10-0173). Small peaks at 32.8°, 36.7°, 46.6°, 50.7°, 60.1° and
67.2° in this sample are due to the mixture of δ and θ-phases of alumina. In addition, other peaks
at 31.7°, 39.4° (γ, δ and θ-phases) and 45.4°, 66.7° (δ-phase) are also present.
4 P. KAUR ET AL.

Figure 1. XRD patterns of heat treated aluminum hydroxide in the temperature range: (a) 150°C–600°C and (b) 800°C–1150°C.
Peaks labeled as $ and * are due to Al(OH)3 and AlOOH respectively.

Further increase in annealing temperature to 1100°C, leads to the enhancement in intensity of


XRD peaks at 25.6°, 35.1°, 37.8°, 41.6°, 43.3°, 52.5°, 57.5°, 66.5° and 68.2° due to the growth of the
α-phase. This sample also shows peaks at 19.4°, 31.4°, 32.7°, 36.6°, 38.8°, 39.9°, 44.8°, 46.4°, 47.6°,
50.7°, 62.5°, 63.9° and 67.5° due to θ-alumina. Other small peaks are observed at 45.6° (δ) and
59.8°, 61.2° (θ/α-phase). Thus, the sample annealed at 1100°C is a mixture of θ and α-alumina
phases, possibly with very small concentration of the remnant δ-phase.
The XRD pattern of the sample annealed at 1150°C, shows peaks at 25.5°, 35.1°, 37.8°, 41.2°, 43.3°,
52.6°, 57.5°, 66.4°, 68.2° due to α-phase with enhanced intensity, which confirms the growth of α-
alumina at the expense of θ-alumina. Small peaks at 32.6°, 36.7°, 38.9°, 54.3°, 64.0° and 67.3° corre-
sponding to the remnant θ-phase are also detected, along with peaks at 36.0° (δ), and 59.7°, 61.1° (α/
θ-phases).
Thus, XRD studies indicate the thermodynamically driven phase transformation from Al(OH)3 to
AlOOH, γ, δ, θ and α-phases of alumina under the heat treatment from 150°C to 1150°C. The sample
PHASE TRANSITIONS 5

annealed at 1150°C show the presence of majority and stable α-phase, along with remnant θ and δ
phases.
Therefore, under heat treatment of aluminum hydroxide, the following structural transform-
ations take place:
◦ ◦ ◦
Al(OH)3 +AlOOH −− − Al(OH)3 +AlOOH−−
250 C
− (g + d)−Al2 O3 −−
400 C
− (g + d)−Al2 O3
600 C

◦ ◦ ◦ ◦
−− − (g + d)−Al2 O3 −−
800 C
− (g + d)−−
950 C
−− Al2 O3 (u + a + d + g)−−
1020 C
−− Al2 O3 (u + a)
1100 C


−− −− Al2 O3 (a + u).
1150 C

The above structural transformation sequence is in good agreement with the earlier findings [7,
10, 32–35].
The average crystallite size (D) of thealumina nanoparticles
 was estimated from the width of the
0.9 l
XRD peak by Debye-Scherrer formula D = and the values are given in Table 1.
b Cosu
Ultra-fine grained γ-alumina nanoparticles having crystallite size in the range: 3–5 nm are formed
upon annealing in the temperature range: 400°C–950°C. The crystallite size of θ-alumina nanopar-
ticles (that first formed at 1020°C) increases from 15 nm to 27 nm after annealing at 1150°C. The α-
alumina nanoparticles are comparatively larger in size and their crystallite size increases from 25 nm
in the sample annealed at 1020°C to 57 nm in the sample annealed at 1150°C.
Cava et al. had reported the structural evolution of different phases of alumina nanoparticles
under the effect of heat treatment from 700°C to 1200°C and reported the values of crystallite size
from XRD and transmission electron microscopy (TEM). These authors reported that in the temp-
erature range: 800–950°C, only γ-Al2O3 exists and the crystallite size was found to be in the range: 7–
9 nm from XRD and ∼10 nm from TEM studies. Upon increasing the annealing temperature to
1025°C, γ-Al2O3 converted into α-Al2O3 whose crystallite size is found to be 59.5 nm from XRD
and 55 nm from TEM [32]. Thus, the values of crystallite size of different phases of alumina nano-
particles as determined from XRD was found to be in good agreement with its value from TEM
studies [32].
Johnston et al. prepared γ-alumina nanopowders by pulsed laser ablation and heat treated the
nanopowders in the temperature range: 200°C–1200°C [36]. At lower temperature 200°C–800°C,
only γ-Al2O3 was detected which transformed to α-phase at 1100°C. γ-alumina had particle size
of 5–10 nm, which increased to 30 nm upon the formation of α-alumina, as calculated by Bru-
nauer–Emmett–Teller (BET) technique. Further Johnson et al. reported that the value of crystallite
size in the sample annealed at 800°C was in agreement with the values determined by TEM and BET
studies [36].

3.2. 27Al MAS-NMR


NMR spectroscopy is a very effective technique to elucidate the short-range order around a specific
atom of interest. Figure 2(a,b) shows the 27Al MAS-NMR spectra of alumina samples heat treated at

Table 1. Crystallite size of alumina phases as a function of the annealing temperature.


Annealing temperature (°C) Phases Crystallite size (D) (nm)
400 γ/δ 3 (γ/δ)
600 γ/δ 4 (γ/δ)
800 γ/δ 5 (γ/δ)
950 γ/δ 5 (γ/δ)
1020 θ+δ+α+γ 15 (θ + δ) + 25 (α)
1100 θ+α 20 (θ) + 30 (α)
1150 α+θ 27 (θ) + 57 (α)
6 P. KAUR ET AL.

Figure 2. 27Al-MAS NMR spectra of alumina nanoparticles annealed (a) 150°C–600°C (b) 800°C–1150°C.

different temperatures, which show two major peaks around 8 and 68 ppm, due to two types of Al—
O speciation: hexa-coordinated structural units (denoted as Al[6]) and tetra-coordinated units
(denoted as Al[4]), respectively [6]. The quantitative fraction of Al[4] and Al[6] units were determined
from the ratio of areas under the respective NMR resonance peaks. Table 2 gives the values of the
chemical shifts (δ), fraction of areas (A4 and A6) under the NMR resonant peaks due to Al[4] and
Al[6] structural units respectively.
Figure 2(a) shows that the samples that were heat-treated at 150°C and 250°C exhibit a single
NMR resonance peak at ∼8.2 ppm showing that both the bohemite and gibbsite phases are made
up of only Al[6] structural units. This is because the structure of both these phases consist of only
hexa-coordinated Al—O units and no tetrahedral Al—O units are present, as discussed by Peitinger
et al. [14]. The samples annealed in the range of 400°C to 950°C show two NMR peaks centered at ∼8
and 67 ppm, showing the presence of both the Al[6] and Al[4] units respectively [21], due to the for-
mation of γ and δ-phases. Due to the similar atomic environment of Al ions in γ and δ-Al2O3, it was
not possible to resolve their NMR peaks [20]. The concentration of Al[4] in the samples annealed at
400°C and 600°C is about 24% in both the samples with remaining 76% that of Al[6] units. The
sample annealed at 800°C contains 31% Al[4] and 69% Al[6] units while the sample annealed at
PHASE TRANSITIONS 7

Table 2. Chemical shifts (δ) and the fraction of areas (A4 and A6) under the NMR resonance peaks in samples annealed at different
temperatures.
Annealing temperature (°C) Phase δ4 (ppm) δ6 (ppm) A4 A6 NAl—O = 4A4 + 6A6
150 Al(OH)3 + AlOOH – 8.2 – 1 6
250 Al(OH)3 + AlOOH – 8.2 – 1 6
400 γ, δ 67.9 8.6 0.24 0.76 5.52
600 γ, δ 66.7 8.4 0.24 0.76 5.52
800 γ, δ 67.4 8.4 0.31 0.69 5.38
950 γ, δ 65.7 8.3 0.32 0.68 5.36
1020 θ, δ, γ 66.7 7.9 0.40 0.60 5.20
α – 13.6
1100 θ, δ 65.3 8.6 0.22 0.78 5.56
α – 14.0
1150 θ, δ 65.1 8.3 0.08 0.92 5.84
α – 14.2

950°C contains 32% Al[4] and 68% Al[6]. Lee et al. also reported the presence of 70% Al[6] units in γ-
Al2O3. [22].
After annealing the sample at 1020°C, the concentration of Al[6] units decreases from 68% to 60%,
while the concentration of Al[4] increases from 32% to 40%. This is due to the formation of θ-Al2O3,
as confirmed by XRD patterns (Figure 1(b)), which is reported to contain equal concentrations of
Al[4] and Al[6] [14]. The deviation from an exact equal concentration of Al[4] and Al[6] is due to
the fact that the sample annealed at 1020°C contains significant concentrations of α, γ and δ-
alumina, all of which do not contain equal concentrations of Al[6] and Al[4]. The sample annealed

Figure 3. Deconvoluted 27Al-MAS NMR spectra of alumina nanoparticles heat treated at (a) 1020°C (b) 1100°C and (b) 1150°C.
8 P. KAUR ET AL.

at 1100°C shows an increase in Al[6] concentration to 78% at the expense of Al[4], which decreases to
22%. With further annealing at 1150°C, the sample is found to contain 8% Al[4] and 92% Al[6] which
confirms the formation of α-alumina, since α-phase is known to contain only Al[6] [15]. The small
quantity of Al[4] in this sample confirms the presence of the residual θ-alumina in the sample
annealed at 1150°C. Thus, both XRD and NMR results are consistent with each other.
The average or overall Al—O coordination number (NAl—O) in samples annealed at different
temperatures is given as:
NAl−O = 4A4 + 6A6 (3)
The data in Table 2 shows that the average aluminum-oxygen coordination first decreases from 6
to ∼5 with the formation of mixture of γ and δ phase from the starting Al(OH)3 and AlOOH com-
pounds and then increases and approaches to the value of 6 again with the growth of the α-phase.
The deconvolution of 27Al-MAS NMR spectra provides quantitative fractions of each phase in the
samples with the multiple phases. NMR peaks were deconvoluted by Peakfit software using Lorent-
zian profile for the peaks. The deconvoluted NMR spectra of alumina samples heat treated at 1020°C,
1100°C and 1150°C is shown in Figure 3(a-c). It is clear from this figure that the peak due to Al[6]
ions in α-alumina shifts to ∼14 ppm [11].
Since the samples heat-treated at 1100°C and 1150°C contain very small amount of δ-phase, we
can neglect its contribution in order to calculate the percentage of α and θ-phases in them. Knowing
that θ-alumina contains equally amounts Al[4] and Al[6] units, whereas α-alumina contains only Al[6]
and using the formula of the total average coordination number, we can calculate the concentration
of α and θ-Al2O3 phases in the samples using the relation:
1
NAl−O = 6x + [6(1 − x) + 4(1 − x)] (4)
2
where NAl—O is the average Al—O coordination, and x and (1-x) is the fraction of α-Al2O3 and θ-
Al2O3 phases respectively in the sample.
Using this formula, the percentage of the α-phase in the sample annealed at 1100°C was calculated
to be 56% which increased to 84% with the increase in temperature to 1150°C at the expense of θ-
alumina whose concentration decreased from 44% to 16%.

4. Conclusions
Aluminum oxide was prepared by the chemical routes using aluminum metal powder and ammonia
solution as precursors. The heat treatment of aluminum hydroxide in the temperature range: 150°C–
1150°C produced structural phase transformations from Al(OH)3 to AlOOH, γ, δ, θ and α phases of
aluminum oxide. XRD and 27Al-MAS NMR techniques were used to elucidate the alumina phase
formation and the local environment of the aluminum ions in the different phases. NMR studies
found the formation of 84% of α-alumina and 16% θ-alumina in the sample annealed upto 1150°
C. The heat treatment of aluminum oxide nanoparticles produced variation in the average Al—O
coordination number. The XRD and 27Al MAS-NMR results provide a consistent picture of the
structural transformations in alumina nanoparticles with heat treatment.

Acknowledgements
Atul Khanna acknowledges Council of Scientific and Industrial Research, New Delhi for research grant. Puneet Kaur
acknowledges the Department of Science and Technology, New Delhi, India for the award of DST-Inspire fellowship.

Disclosure statement
No potential conflict of interest was reported by the author(s).
PHASE TRANSITIONS 9

Funding
This work was supported by Council of Scientific and Industrial Research, New Delhi, India.

References
[1] González W, Álvarez E, Martínez-Martínez R, et al. Cold white light generation through the simultaneous
emission from Ce3+, Dy3+ and Mn2+ in 90Al2O3· 2CeCl3·3DyCl3· 5MnCl2 thin film. J Lumin. 2012;132
(8):2130–2134.
[2] Li J, Pan Y, Xiang C, et al. Low temperature synthesis of ultrafine α-Al2O3 powder by a simple aqueous sol–gel
process. Ceram Int. 2006;32(5):587–591.
[3] Song X-L, Qu P, Yang H-P, et al. Synthesis of γ-Al2O3 nanoparticles by chemical precipitation method. J Centr
South Univ Technol. 2005;12(5):536–541.
[4] Qu L, He C, Yang Y, et al. Hydrothermal synthesis of alumina nanotubes templated by anionic surfactant. Mater
Lett. 2005;59(29–30):4034–4037.
[5] Issa I, Joly-Pottuz L, Réthoré J, et al. Room temperature plasticity and phase transformation of nanometer-sized
transition alumina nanoparticles under pressure. Acta Mater. 2018;150:308–316.
[6] Tavakoli AH, Maram PS, Widgeon SJ, et al. Amorphous alumina nanoparticles: structure, surface energy, and
thermodynamic phase stability. J Phys Chem C. 2013;117(33):17123–17130.
[7] Pati RK, Ray JC, Pramanik P. A novel chemical route for the synthesis of nanocrystalline α-Al2O3 powder. Mater
Lett. 2000;44(5):299–303.
[8] Wefers K, Misra C. Oxides and hydroxides of aluminum, ALCOA Technical paper No. 19, ALCOA Laboratories.
USA, 1987.
[9] Levin I, Brandon D. Metastable alumina polymorphs: crystal structures and transition sequences. J Am Ceram
Soc. 1998;81(8):1995–2012.
[10] Boumaza A, Favaro L, Lédion J, et al. Transition alumina phases induced by heat treatment of boehmite: an X-
ray diffraction and infrared spectroscopy study. J Solid State Chem. 2009;182(5):1171–1176.
[11] O’Dell LA, Savin SL, Chadwick AV, et al. A 27Al MAS NMR study of a sol–gel produced alumina: Identification
of the NMR parameters of the θ-Al2O3 transition alumina phase. Solid State Nucl Magn Reson. 2007;31(4):169–
173.
[12] Cai S-H, Rashkeev SN, Pantelides ST, et al. Phase transformation mechanism between γ-and θ-alumina. Phys
Rev B. 2003;67(22):224104.
[13] Levin I, Gemming T, Brandon D. Some metastable polymorphs and transient stages of transformation in
alumina. Phys Status Solidi. 1998;166(1):197–218.
[14] Peintinger MF, Kratz MJ, Bredow T. Quantum-chemical study of stable, meta-stable and high-pressure alumina
polymorphs and aluminum hydroxides. J Mater Chem A. 2014;2(32):13143–13158.
[15] John C, Alma N, Hays G. Characterization of transitional alumina by solid-state magic angle spinning alu-
minium NMR. Appl Catal. 1983;6(3):341–346.
[16] Ionescu A, Allouche A, Aycard J-P, et al. Study of γ-alumina surface reactivity: Adsorption of water and hydro-
gen sulfide on octahedral aluminum sites. J Phys Chem B. 2002;106(36):9359–9366.
[17] Mo S-D, Ching W. Electronic and optical properties of θ−Al2O3 and comparison to α−Al2O3. Phys Rev B.
1998;57(24):15219.
[18] Pecharroman C, Sobrados I, Iglesias J, et al. Thermal evolution of transitional aluminas followed by NMR and IR
spectroscopies. J Phys Chem B. 1999;103(30):6160–6170.
[19] Minnermann M, Neumann B, Zielasek V, et al. Alumina-promoted cobalt and iron xerogels as catalyst for the
Fischer–Tropsch synthesis. Catal Sci Technol. 2013;3(12):3256–3267.
[20] Kim HN, Lee SK. Effect of particle size on phase transitions in metastable alumina nanoparticles: a view from
high-resolution solid-state 27Al NMR study. Am Mineral. 2013;98(7):1198–1210.
[21] Samain L, Jaworski A, Edén M, et al. Structural analysis of highly porous γ-Al2O3. J Solid State Chem.
2014;217:1–8.
[22] Lee M-H, Cheng C-F, Heine V, et al. Distribution of tetrahedral and octahedral Al sites in gamma alumina.
Chem Phys Lett. 1997;265(6):673–676.
[23] O’Reilly D. Quadrupolar broadened nuclear magnetic resonance of polycrystalline solids. J Chem Phys. 1958;28
(6):1262–1264.
[24] Ansell S, Krishnan S, Weber JR, et al. Structure of liquid aluminum oxide. Phys Rev Lett. 1997;78(3):464.
[25] Lee SK, Lee SB, Park SY, et al. Structure of amorphous aluminum oxide. Phys Rev Lett. 2009;103(9):095501.
[26] Nayar P, Khanna A, Kabiraj D, et al. Structural, optical and mechanical properties of amorphous and crystalline
alumina thin films. Thin Solid Films. 2014;568:19–24.
[27] Shi C, Alderman OL, Berman D, et al. The structure of amorphous and deeply supercooled liquid alumina. Front
Mater. 2019;6:38.
10 P. KAUR ET AL.

[28] Varatharajan K, Dash S, Arunkumar A, et al. Synthesis of nanocrystalline α-Al2O3 by ultrasonic flame pyrolysis.
Mater Res Bull. 2003;38(4):577–583.
[29] Ismail RA, Zaidan SA, Kadhim RM. Preparation and characterization of aluminum oxide nanoparticles by laser
ablation in liquid as passivating and anti-reflection coating for silicon photodiodes. Appl Nanosci. 2017;7
(7):477–487.
[30] Padilla-Rosales I, Cabañas-Moreno J, Jiménez G, et al. Near UV excitable Eu-doped alumina nanophosphors
synthesized by the microwave assisted solvothermal technique. Mater Res Express. 2017;4(12):125007.
[31] Monteiro MAF, Brito HFD, Felinto M, et al. Photoluminescence behavior of Eu3+ ion doped into γ-and α-
alumina systems prepared by combustion, ceramic and Pechini methods. Microporous Mesoporous Mater.
2008;108(1–3):237–246.
[32] Cava S, Tebcherani S, Souza I, et al. Structural characterization of phase transition of Al2O3 nanopowders
obtained by polymeric precursor method. Mater Chem Phys. 2007;103(2–3):394–399.
[33] Chang P-L, Yen F-S, Cheng K-C, et al. Examinations on the critical and primary crystallite sizes during θ-to α-
phase transformation of ultrafine alumina powders. Nano Lett. 2001;1(5):253–261.
[34] Li JG, Sun X. Synthesis and sintering behavior of a nanocrystalline α-alumina powder. Acta Mater. 2000;48
(12):3103–3112.
[35] Musil J, Blažek J, Zeman P, et al. Thermal stability of alumina thin films containing γ-Al2O3 phase prepared by
reactive magnetron sputtering. Appl Surf Sci. 2010;257(3):1058–1062.
[36] Johnston GP, Muenchausen R, Smith DM, et al. Reactive laser ablation synthesis of nanosize alumina powder. J
Am Ceram Soc. 1992;75(12):3293–3298.

You might also like