You are on page 1of 7

Journal of Luminescence 242 (2022) 118554

Contents lists available at ScienceDirect

Journal of Luminescence
journal homepage: www.elsevier.com/locate/jlumin

Structural properties investigation of different alumina polymorphs (η-, γ-,


χ-, θ-, α-Al2O3) using Cr3+ as a luminescent probe
M.G. Baronskiy *, S.V. Tsybulya, A.I. Kostyukov, A.V. Zhuzhgov, V.N. Snytnikov
Boreskov Institute of Catalysis SB RAS, Lavrentiev Ave. 5, 630090, Novosibirsk, Russia

A R T I C L E I N F O A B S T R A C T

Keywords: The structural properties of η-, γB-, γPb-, χ-, θ- and α-Al2O3 polymorphs were studied by X-ray diffraction and
Alumina polymorphs photoluminescence spectroscopy using Cr3+ ion as a luminescent probe in its natural concentration (<10− 3 wt
Luminescence %). The analysis of photoluminescence (PL) and photoluminescence excitation (PLE) data within the crystal field
Cr3+ luminescent probe
theory revealed a decrease in the crystal field strength Dq from 1810 to 1740 cm− 1, the Racah parameter B from
Structure
606 to 432 cm− 1, and the covalence parameter β1 from 1.076 to 1.062 in the series α-Al2O3 → γ-, η-Al2O3. This
indicates that the fraction of covalent bond increases for the low-temperature γ-Al2O3 and η-Al2O3 phases as
compared to α-Al2O3 powder and ruby single crystal. The obtained values of crystal field strength Dq, Racah
parameters B and C, and covalence degree β1 also differ from each other among low-temperature phases η-, γB-,
γPb- and χ-Al2O3. This testifies that each low-temperature Al2O3 polymorph has different ionicity/covalence
ratio. The analysis of PL and PLE data showed that the chromium ion in χ-Al2O3 occupies the local structural
positions that are similar to those in α-Al2O3. This occurs because, according to XRD data, the cubic closest
packing of oxygen anions in χ-Al2O3 partially transforms into the hexagonal closest packing, which is typical of
α-Al2O3.

1. Introduction concepts concerning the structure of low-temperature alumina poly­


morphs indicate that they differ from each other not in the atomic
Alumina (Al2O3) is a multifunctional material having a wide set of structure of the ordered crystal blocks, which have the defect spinel
polymorphs [1–3]. There are high-temperature (δ-, κ-, θ-, α-Al2O3) structure in all cases, but rather in their nanostructure – the morphology
phases and low-temperature (η-, γ-, χ-Al2O3) modifications of Al2O3. The and intergrowth of crystallites [12–15]. It means that differences in the
only thermodynamically stable alumina phase is α-Al2O3. All metastable local structure of low-temperature polymorphs, which are related to the
Al2O3 polymorphs irreversibly transform into α-Al2O3 upon thermal presence of typical point and extended defects (including the boundaries
treatment of the corresponding hydroxides according to the diagram of blocks and the stacking faults), may be much more essential than
displayed in Fig. 1. differences in the averaged crystal structure.
Alumina has various application fields: for example, α-Al2O3 is Photoluminescence (PL) spectroscopy is a powerful method for
widely used in optical industry as a material for the production of optical investigating the local and electronic structures of various materials
elements such as windows, prisms, optical fibers, components of lasing [16–20]. Many publications are devoted to the Cr3+:α-Al2O3 system [6,
medium), thermoluminescent radiation detectors, and others and also 21–23]. In some works the luminescence properties of θ-Al2O3 are
for antireflection and protective coatings [6–9]. Low-temperature η- and studied [24,25]. One of the main advantages of PL spectroscopy in
γ-Al2O3 polymorphs serve as the supports for heterogeneous catalysts exploring the structure of deliberately undoped Al2O3 phases is that the
[10,11]. High- and low-temperature Al2O3 phases are efficiently structure-sensitive probes are represented by impurity ions of 3d ele­
distinguished by X-ray diffraction (XRD) and transmission electron mi­ ments, minor amounts of which (<10− 3 wt%) are always present in
croscopy (TEM) methods. However, the closeness of crystal structures Al2O3 [2,26–28]. This allows studying the local and electronic structures
and the similarity of diffraction patterns of the low-temperature Al2O3 of Al2O3 without the introduction of additional distortions. It should be
polymorphs hinder their identification in a mixture of phases. Modern noted that, in contrast to high-temperature Al2O3 polymorphs, the

* Corresponding author.
E-mail addresses: baronskiy@catalysis.ru, baronskiymg@mail.ru (M.G. Baronskiy).

https://doi.org/10.1016/j.jlumin.2021.118554
Received 27 May 2021; Received in revised form 21 October 2021; Accepted 23 October 2021
Available online 26 October 2021
0022-2313/© 2021 Elsevier B.V. All rights reserved.
M.G. Baronskiy et al. Journal of Luminescence 242 (2022) 118554

literature provides virtually no data about differences in physicochem­ following content of impurities (wt%): Na = 0.11; K = 0.03 and Si =
ical properties of the low-temperature alumina phases revealed by PL 0.01. The initial γ-Al(OH)3 was treated in a DESI-15 double-disk disin­
spectroscopy. In our earlier studies, PL probing of the structure of sub­ tegrator to obtain a powder with the particle size distribution having a
stances by Cr3+, Fe3+ and Mn4+ ions was used to reveal and analyze the maximum at ca. 50 μm. After that, the produced γ-Al(OH)3 powder was
spectroscopic differences between low-temperature polymorphs ground in a ball mill during five days for obtaining a powder with the
γB-Al2O3 and γPb-Al2O3, which were obtained by the sol-gel technology particle size distribution of ca. 1–5 μm. The subsequent calcination of
from different precursors (crystalline boehmite and pseudoboehmite) the γ-Al(OH)3 sample in a muffle furnace upon slow heating for 6 h at
[26,29]. PL method with the Cr3+ probe was employed to elucidate the 550◦ С led to the formation of single-phase χ-Al2O3.
phase composition of Al2O3 nanoparticles, which were obtained by laser η-Al2O3 and θ-Al2O3. Bayerite (β-Al(OH)3) served as the initial
vaporization using a cw CO2 laser. As a result, (γ + δ)-Al2O3, θ-Al2O3 and substance to obtain η-Al2O3 and θ-Al2O3. The bayerite was synthesized
α-Al2O3 phases were identified [27]. by precipitation of Al3+ ions from an aqueous solution of a commercially
The understanding of the spectral features of 3d elements in crys­ available Al(NO3)3 (NevaReaktiv, Russia) salt using a 25% aqueous so­
talline solids is very important for estimating the future applications of lution of ammonia (NevaReaktiv, Russia) as a precipitant. The content of
these materials. Such understanding can be provided by a detailed impurities in β-Al(OH)3 (wt%): Si = 0.03; Cu = 0.02 and S = 0.02. The
analysis of the positions of energy levels, their splitting in the crystal initial aqueous solutions of aluminum nitrate and ammonium hydroxide
field, and the dependence of this splitting on the nature of ligands and with the concentration 1 mol/l were mixed via their simultaneous
chemical bond properties for the particular host materials. For impurity pouring at a rate of 1 and 3 ml/min, respectively, under vigorous stirring
ions of 3d elements located in the matrix of different Al2O3 polymorphs, with a stirrer at 500 rpm. Aging of the mixed solutions was carried out at
such an analysis can be performed within the crystal field theory. pH 10.0 and temperature 90 ◦ C for 12 h. After that, the resulting gel was
Recently, S. Adachi considered the luminescence data for Cr3+ ion in washed using a Buchner funnel with a filter paper to remove impurities
different Al2O3 polymorphs using the Franck-Condon analysis with the and dried to a constant weight at 150 ◦ C for 12 h in order to obtain a
configurational-coordinate model [30]. In particular, he used the spec­ xerogel. Note that washing of the freshly precipitated gel was performed
tral data for η-Al2O3 and γ-Al2O3 that were obtained by the authors of after achieving nearly a neutral pH value. η-Al2O3 and θ-Al2O3 samples
this work. In our study, structural and luminescence data obtained in were synthesized by calcination of bayerite in air for 4 h at 550 and
experiments were systematized and analyzed for a wide set of undoped 950 ◦ C, respectively.
single-phase Al2O3 powders: η-, γB-, γPb-, χ-, θ- and α-Al2O3. Specifically, The steps used for the synthesis of η-, γB-, γPb-, χ-, θ- and α-Al2O3
PL and PLE spectra of the χ-Al2O3 sample were recorded and analyzed powders, which were examined in the work, are summarized in a
for the first time. One of the most important results of the studies re­ scheme displayed in Fig. 2.
ported in this work is the revealed structural closeness of χ- and α-Al2O3.
Thus, the aim of the work was to obtain and analyze within the crystal 2.2. Characterization of samples
field theory the spectral-luminescence characteristics of various Al2O3
polymorphs, particularly the low-temperature η-, γ- and χ-phases, to The chemical composition of the samples under consideration was
compare the results of analysis with the structural data obtained by XRD, controlled by X-ray fluorescence (XRF) analysis on an ARL – Advant’x
and to find differences in the above listed characteristics for the tested analyzer. X-ray diffraction (XRD) patterns of alumina powders were
η-, γB-, γPb-, χ-, θ- and α-Al2O3 single-phase samples. obtained on a Bruker D8 Advance diffractometer using CuKα radiation
(λ = 0.15418 nm). Measurements were performed in the 2θ range of
2. Experimental 10–70◦ with a step 0.05◦ and acquisition time 3 s. Phases were identified
by comparing experimental diffraction patterns with the data of ICDD,
2.1. Preparation of samples PDF 2 database. Photoluminescence (PL), photoluminescence excitation
(PLE) spectra and PL decay curves were measured on a CaryEclipse
γB-Al2O3 and γPb-Al2O3 powders were synthesized according to the (Varian) fluorescence spectrophotometer with a Xe lamp as an excita­
sol-gel procedure described in Ref. [26]. The α-Al2O3 powder was ob­ tion source. The powders under consideration were placed in quartz
tained by thermal treatment of the γB-Al2O3 powder in air at a temper­ cuvettes.
ature of 1250 ◦ C for 4 h.
χ-Al2O3. In the synthesis of χ-Al2O3 powder, the initial substance 3. Results and discussion
was represented by a commercially available aluminum hydroxide –
gibbsite (γ-Al(OH)3) (Achinsk Alumina Refinery, Russia), which had the According to XRF data, the content of impurity ions of 3d elements

Fig. 1. Temperature diagram of transient Al2O3 polymorphs obtained by thermal treatment of the corresponding aluminum hydroxides [4,5].

2
M.G. Baronskiy et al. Journal of Luminescence 242 (2022) 118554

Fig. 2. A scheme of thermal decomposition of aluminum hydroxides in air at atmospheric pressure for the preparation of η-,γB-, γPb-, χ-, θ- and α-Al2O3 powders.

was below the detection limit of the method (<0.01 wt%) virtually in all phase identification, are determined by the nanocrystalline arrangement
the tested Al2O3 powders. A chromium impurity was not detected in any of low-temperature alumina polymorphs [12].
of the tested single-phase Al2O3 samples. Besides, a titanium impurity Fig. 4 a shows the normalized PL spectra of alumina powders
was identified in the γB-Al2O3 sample. An iron impurity was detected in recorded at excitation with λex = 530 nm. Fig. 4 b presents the corre­
η-Al2O3 and θ-Al2O3 samples. Fig. 3 demonstrates XRD data of the tested sponding normalized PLE spectra. As seen in Fig. 4 a, b, differences in
Al2O3 powders. X-ray diffraction patterns displayed in Fig. 3 are typical the luminescence properties between Al2O3 phases manifest themselves
of alumina polymorphs. The acquired XRD data show that all the Al2O3 as the differing positions of band maxima in PL and PLE spectra, and
samples examined in the study are the single-phase systems. their intensity ratios in the series α-Al2O3 → γ-Al2O3. For example, the
α-Al2O3 (hexagonal structure; space group R3c) and θ-Al2O3 luminescence intensity of the α-Al2O3 sample is eight times higher as
(monoclinic structure; space group A2/m) can be identified quite easily compared to γB-Al2O3.
because they are the well-crystallized phases. All low-temperature The PL spectra (Fig. 4 a) correspond to the radiative electronic
alumina species are nanocrystalline and, according to the generally transition 2Eg →4A2g in the Cr3+ ions, which isomorphously substitute
accepted ideas, are considered to have the spinel structure (space group Al3+ and occupy regular octahedral positions in the Al2O3 structure. The
Fd3m). This is why their diffraction patterns look quite similar at the PLE spectra (Fig. 4 b) contain two characteristic broad bands corre­
first glance. However, some important features make it possible to sponding to the electronic transitions 4T1g, 4T2g → 4A2g in Cr3+ ions
identify such species. Thus, χ-Al2O3 is characterized by the presence of incorporated into the structure of different Al2O3 polymorphs. The
an additional diffuse scattering peak at ca. 42.6◦ , which is not typical of presence of such PLE spectra confirms that the PL data were interpreted
the cubic spinel structure. This is the peak that emerges due to the correctly.
presence of planar defects in the closest packing of anions [12]. γ- and The PL decay curves are displayed in Fig. 5. The PL decay curves for
η-Al2O3 can be distinguished using the intensity ratio of peaks 311 and Cr3+ ions in the tested single-phase Al2O3 samples demonstrate that the
222, the shift of peak 311 in opposite directions from the position lifetime (τPL) of excited 2Eg state falls in the millisecond range, which
determined by the lattice parameter, and the presence of a weak but corresponds to the known literature data, particularly for the Cr3+:α-
sharp peak 111 for η-Al2O3. All these features, which are not usual for Al2O3 system [32]. The PL decay curve of Cr3+ ions in α-Al2O3 was

Fig. 3. A comparison of the X-ray diffraction patterns of as-prepared low-temperature χ-, η-, γB-, γPb-Al2O3 polymorphs (a) and high-temperature θ-, α-Al2O3 (b) with
data from ICDD databases for pure χ-, η-, γ-, θ-, α-Al2O3, respectively. Indices of the peaks in the case of low-temperature species γB-, γPb-, χ- and η-Al2O3 correspond
to the cubic spinel structure, in the case of α-Al2O3 – to the hexagonal corundum structure, and in the case of θ-Al2O3 (according to the data reported in Ref. [31]) – to
the monoclinic structure.

3
M.G. Baronskiy et al. Journal of Luminescence 242 (2022) 118554

Fig. 4. Normalized PL (a; λex = 530 nm) and PLE (b; λmax = 695 (γB), 698 (γPb), 687 (θ), 696 (η), 694 (α, χ) nm) spectra for powders of different Al2O3 polymorphs.
The inset in Fig. 4a shows R1,2-lines of the luminescence of Cr3+ ions in α-Al2O3.

Fig. 5. The PL decay curves (log scale) for γB- (a), γPb- (b), θ- (c), η-(d), α- (e) and χ-Al2O3 (f) powders (λmax = 695 (γB), 698 (γPb), 687 (θ), 696 (η), 694 (χ, α) nm). PL
lifetimes of Cr3+ ions in different Al2O3 polymorphs are indicated.

approximated by an exponential function with the formula y = A1 ⋅


( ) Table 1
exp − τx1 + y0 . All other obtained PL decay curves were approximated Lifetimes (τPL)av of the excited 2Eg state of Cr3+ ions calculated for different
( Al2O3 polymorphs.
by a biexponential function according to the formula y = A1 ⋅ exp − Sample (λPL, nm) τ1, ms τ2, ms (τPL)av, ms
) ( ) η-Al2O3 (696) 3.99 0.38 3.23 ± 0.02
x x
τ1 + A2 ⋅exp − τ2 + y0 . The corresponding τ(1,2)PL values are indi­ γB-Al2O3 (695) 2.69 0.25 1.97 ± 0.07
γPb-Al2O3 (698) 3.29 0.31 2.61 ± 0.07
cated in Fig. 5 and in Table 1. χ-Al2O3 (694) 3.69 0.25 3.32 ± 0.07
Particles of the tested η-, γB-, γPb-, χ- and θ-Al2O3 samples are shown θ-Al2O3 (688) 4.78 0.36 4.87 ± 0.05
on a nanometer scale. It is known [17,33] that the PL decay time of α-Al2O3 (694) 4.19 ̶ 4.19 ± 0.03
impurity centers in nanostructured systems is often described by a
biexponential function, which is caused by the presence of nonequiva­
Moreover, our earlier study [20] revealed in η-, γB- and γPb-Al2O3 sam­
lent luminescence centers, some of them corresponding to the bulk
ples not only the luminescence of bulk Cr3+b centers, but also the
centers, while others residing in subsurface layers of nanocrystallites.
luminescence of surface Cr3+s centers.

4
M.G. Baronskiy et al. Journal of Luminescence 242 (2022) 118554

Taking into account that for all the studied samples, except α-Al2O3, Table 3
the PL decay time curves are not monoexponential, the average values of The crystal field strength (Dq), the Racah parameters (B and C) and the cova­
τPL ((τPL)av) were calculated for them by the formula reported in lence degree for Cr3+:Al2O3 polymorphs.
Ref. [34]. Sample Dq B Dq/ С С/B β1
(cm− 1) (cm− 1) B (cm− 1)
A1τ21 + A2τ22
(τPL )av ​ = ​ η-Al2O3 1770 505 3.5 3497 6.9 1.062
A1τ1 + A2 τ 2 γB-Al2O3 1740 432 4.0 3667 8.5 1.062
γPb-Al2O3 1800 504 3.6 3466 6.9 1.054
Results of the calculations for the corresponding single-phase Al2O3
χ-Al2O3 1810 493 3.7 3524 7.1 1.061
samples are listed in Table 1. θ-Al2O3 1740 564 3.1 3401 6.0 1.076
One can see from PL and PLE spectra (Fig. 4) that the spectroscopic α-Al2O3 1810 606 3.0 3273 5.4 1.076
behavior of Cr3+ ions in Al2O3 polymorphs is different. This is evidenced Ruby 1664 640 2.6 3300 5.2 1.105
first of all by different positions of the band maxima in PL and PLE Al2O3:Cr3+ 0,4 at.% 1785 676 2.6 3304 4.9 1.131
(ceramic)
spectra. The acquired luminescence data were analyzed within the
crystal field theory using a set of equations for ions with d3 configuration *B0 = 918 cm− 1 and С0 = 3850 cm− 1 are the parameters of interelectron
in the octahedral field [35,36]: interaction of free Cr3+ ions obtained from the atomic spectra [35,39].

E(4T2g) = 10Dq aluminosilicate glasses [43]. The Dq/B value can be used to estimate the
( )
E 4 T2 g ​ = ​ 10Dq degree of perfection of the local environment of chromium ions with
ligands. Thus, the higher is the Dq/B value, the less distorted is the local
/ [ ]/ octahedral environment of the metal ion.
B Dq = (ΔE/Dq)2 ​ − ​ 10(ΔE / Dq) 15[(ΔE / Dq) ​ − ​ 8]
It is known that the low-temperature η- and γ-Al2O3 polymorphs
( )/ / / have the face-centered cubic (FCC) crystal lattice with oxygen anions
E 2 Eg B ​ = ​ 3.05C B ​ + ​ 7.90 ​ − ​ 1.80B Dq located in its nodes to form the three-layered cubic closest packing.
Aluminum cations reside in the interstitial sites, occupying the octahe­
where. ΔE = E(4 T1 g) − E(4 T2 g) dral and tetrahedral positions and forming the spinel structure [12,44,
E(4T2g), E(4T1g) and E(2Eg) are the energy values for the corre­ 45]. The α-Al2O3 structure has the hexagonal closest packing (HCP) of
sponding energy levels of Cr3+ ions located in the Al2O3 matrix of oxygen anions, in which aluminum cations occupy pairwise two adja­
different polymorphs, which were found from PL and PLE spectra. The cent octahedra [46]. Due to a strong electrostatic interaction of
indicated values for the tested Al2O3 polymorphs are summarized in aluminum cations, which results from connection of the corresponding
Table 2. octahedra by their faces, local positions of the Al3+ cations and,
As a result, values of the crystal field strength (Dq) and Racah pa­ accordingly, the substituting Cr3+ cations, in α-Al2O3 are more distorted
rameters B and C were obtained for all the studied single-phase Al2O3 than in the case of η- and γ-Al2O3. This is consistent with the data ob­
samples. In addition, taking into account the expression β1 = tained by the analysis of luminescence measurements for the tested
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( )2
powders of single-phase Al2O3. Thus, it is seen in Table 3 that for all the
B
B0 + CC0 [37], which quantitatively describes the nephelaux­ tested Al2O3 samples in the series γB-, η-Al2O3 → α-Al2O3 the Dq/B value
etic effect in the spectroscopy of the Cr3+ ion, the β1 values were decreases more than 1.3-fold, and in comparison with the results for
calculated for all the samples under consideration. The calculated values ruby single crystal taken from Ref. [35] – 1.5-fold.
of Dq, B, C and β1 are summarized in Table 3. For γB-, γPb- and α-Al2O3 Noteworthy is the similarity between PL spectra of the χ-Al2O3 and
phases examined in this work, the Dq, B, C and β1 values were obtained α-Al2O3 samples (Fig. 4 a). Thus, in distinction to PL spectra of other
in our earlier studies [26,29]. Here we present the refined values of the low-temperature Al2O3 polymorphs, the spectrum of χ-Al2O3 contains a
indicated spectroscopic parameters for γB-, γPb- and α-Al2O3 samples relatively narrow luminescence band with a maximum at λmax = 694
taking into account the electronic-vibrational nature of the bands cor­ nm, as in the case of PL spectrum of α-Al2O3. In addition, the crystal field
responding to 4T1g, 4T2g → 4A2g and 2Eg → 4A2g transitions in Cr3+ ions of strength Dq for the low-temperature χ-Al2O3 phase is similar to that for
the Al2O3 matrix. For comparison purposes, Table 3 lists the spectro­ α-Al2O3. At the same time, it should be noted that B, С and β1 parameters
scopic data for ruby crystal and Al2O3 ceramics with the Cr3+ concen­ for χ-Al2O3 are close to the corresponding values for low-temperature η-
tration of 0.4 at% that were taken from Refs. [35,38]. and γ-phases (see Table 3). On the other hand, the obtained PLE spec­
The Dq/B value characterizes the strength of the octahedral crystal trum of χ-Al2O3 is more noisy in comparison with the PLE spectra of the
field in which the ion of d element is located. For ions with the d3 other studied Al2O3 polymorphs. Thus, a strong instrumental effect
configuration, which include Cr3+, the case of strong crystal field is could affect the shape of the broad PLE bands of χ-Al2O3, which could
implemented at Dq/B > 2.3 [35,40,41]. Such Dq/B values correspond, lead to slightly different spectroscopic parameters (B, C) from the
for example, to the Cr3+ ions that occupy regular positions in the calculated above. According to XRD data (Fig. 3), the peak at 2θ ~42.6◦
structure of various crystals: corundum, topaz, alexandrite, etc. [42]. in the diffraction pattern of χ-Al2O3 sample indicates that the cubic
Whereas the case of weak crystal field (Dq/B < 2.3) is implemented for closest packing of oxygen atoms in the χ-phase partially transforms into
the Cr3+ ions in strongly disordered systems, for example, in the hexagonal closest packing, which makes it closer to the α-Al2O3
structure. Thus, there is an evident correlation between structural and
luminescence spectroscopy data for the tested χ-Al2O3 and α-Al2O3
Table 2 powders as well as between χ-Al2O3 and other investigated low-
Positions of the band centroids in PL and PLE spectra of Cr3+ ion in the tested temperature η-, γB- and γPb-Al2O3.
Al2O3 samples. The ratio of contributions of ionicity and covalence for the (Me) –
Sample E(4T2g), cm− 1
E(4T1g), cm− 1
E(2Eg), cm− 1 (ligand) chemical bond, in particular, for oxide host materials, is the
η-Al2O3 17700 23100 14400
determining factor in variation of the Racah parameters. If the covalent
γB-Al2O3 17400 22100 14400 bond between 3d cation and O anion grows, this leads to a decrease in
γPb-Al2O3 18000 23400 14300 parameters B and C [38]. As seen in Table 3, the value of parameter B
χ-Al2O3 18100 23400 14400 decreases from 606 cm− 1 for α-Al2O3 to 432 cm− 1 for, e.g., γB-Al2O3.
θ-Al2O3 17400 23300 14500
Such a trend is not observed for parameter C; however, if parameters B
α-Al2O3 18100 24400 14403

5
M.G. Baronskiy et al. Journal of Luminescence 242 (2022) 118554

and C are considered jointly when estimating the covalence parameter Author statement
β1, the β1 value also shows a slight decrease – from 1.08 for α-Al2O3 to
1.06 for γ- and η-phases. Thus, the overall decrease in parameters B and Mark G. Baronskiy: Conceptualization, Investigation, Writing –
β1 in the series of tested polymorphs α-Al2O3 → γ-, η-Al2O3 indicates that original draft. Sergey V. Tsybulya: Conceptualization, Investigation,
the contribution of the covalent bond increases for the low-temperature Writing – original draft. Anton I. Kostyukov: Investigation. Aleksey V.
polymorphs γ- and η-Al2O3 as compared to the α-phase and ruby crystal. Zhuzhgov: Investigation. Valeriy N. Snytnikov: Conceptualization.
The values of parameters Dq, B, C and β1 calculated from PL and PLE
spectra also slightly differ from each other and for different Declaration of competing interest
low-temperature polymorphs γB-, γPb-, η-Al2O3 and χ-Al2O3. This dem­
onstrates that each of the low-temperature Al2O3 polymorphs has The authors declare that they have no known competing financial
different ionicity/covalence ratio. interests or personal relationships that could have appeared to influence
It is important to note that for some of investigated alumina poly­ the work reported in this paper.
morphs the C/B value is very high. This is especially pronounced for
low-temperature modifications η-, γB-, γPb- and χ-Al2O3 (see Table 3). M. Acknowledgements
Back et al. in their studies of Cr3+:CaHfO3 system showed that a more
reliable discussion and comparison of the crystal field effect on Cr3+ in This work was financially supported by the Ministry of Science and
unconventional sites should be based on different Tanabe-Sugano dia­ Higher Education of the Russian Federation within the governmental
grams for each material with different C/B [47]. If we consider, for order for Boreskov Institute of Catalysis SB RAS (projects АААА-А21-
example, the PL spectrum of γB-Al2O3 (Fig. 4 a, black curve) it can be 121011390009-1 and АААА-А21-121011390053-4).
seen that there is a clear emission in longwave spectrum region which
can be assigned to 4T2 → 4A2 electronic transition. Thus, a weaker Dq/B References
with respect to the other polymorphs is expected. However, the value of
Dq/B = 4, is the highest among the investigated Al2O3 polymorphs. This [1] L.D. Hart, Alumina Chemicals: Science and Technology Handbook, The American
Ceramic Society, New York, 1990.
can be explained by considering such a high C/B value. In our case it is [2] L. Trinkler, B. Berzina, D. Jakimovica, J. Grabis, I. Steins, UV-light induced
known that low-temperature alumina polymorphs are spinel structures, luminescence processes in Al2O3 bulk and nanosize powders, Opt. Mater. 32 (2010)
lamellar materials with different ratios of octahedral and tetrahedral 789–795.
[3] J. Gangwar, B.K. Gupta, S.K. Tripathi, A.K. Srivastava, Phase dependent thermal
cation positions. Moreover, as was shown in Ref. [12] low-temperature and spectroscopic responses of Al2O3 nanostructures with different morphogenesis,
aluminas should be classified with respect to their nanostructural fea­ Nanoscale 7 (2015) 13313–13344.
tures. For spinels, the distinctions between Al2O3 phases relate to their [4] P. Euzen, Alumina, Handbook of Porous Solids, 2002, pp. 1591–1677.
[5] G. Busca, Structural, surface, and catalytic properties of aluminas, Adv. Catal. 57
nanostructure, in particular, the morphology and linking of nano­ (2014) 319–404.
particles that have similar crystal structures. The presence of such [6] T.H. Maiman, Stimulated optical radiation in ruby masers, Nature 187 (1960)
structural features is one of the reasons leading to a greater contribution 493–494.
[7] D.H. Jundt, M.M. Fejer, R.L. Byer, Growth and optical properties of single-crystal
of covalence. This fact, in turn, can explain such large C/B values for
sapphire fibers, SPIE Proceedings on Infrared Fiber Optics 1048 (1989) 39–43.
low-temperature Al2O3 polymorphs as compared, for example, with [8] M.S. Akselrod, S.W.S. McKeever, A radiation dosimetry method using pulsed
α-Al2O3. optically stimulated luminescence, Radiat. Protect. Dosim. 81 (1999) 167–176.
[9] N. Yamaguchi, K. Tadanaga, A. Matsuda, T. Minami, M. Tatsumisago, Anti-
reflective coatings of flowerlike alumina on various glass substrates by the sol–gel
4. Conclusion process with the hot water treatment, J. Sol. Gel Sci. Technol. 33 (2005) 117–120.
[10] B.C. Lippens, J.J. Steggarda, in: B.G. Linsen (Ed.), In Physical and Chemical Aspects
It was shown that the PL spectroscopy using Cr3+ ion as a lumines­ of Adsorbents and Catalysts, Academic Press, London, New York, 1970, p. 177.
[11] A.B. Stiles, Catalyst Supports and Supported Catalysts: Theoretical and Applied
cent probe is applicable to investigation of structural differences in a Concepts, Butterworth Publishers, Stoneham, MA, 1987, p. 270.
wide set of Al2O3 polymorphs. For undoped (the content of impurity [12] S.V. Tsybulya, G.N. Kryukova, Nanocrystalline transition aluminas: nanostructure
d elements <10− 3 wt%) η-, γB-, γPb-, χ-, θ- and α-Al2O3, the PL, PLE and features of X-ray powder diffraction patterns of low-temperature Al2O3
polymorphs, Phys. Rev. B 77 (2008), 024112-1-13.
spectra and PL decay curves were recorded for the luminescence of Cr3+ [13] G. Busca, The surface of transitional aluminas: a critical review, Catal. Today 226
ions occupying regular structural positions in the matrix of some Al2O3 (2014) 2–13.
phase. The revealed spectroscopic differences in the behavior of Cr3+ [14] V. Pakharukova, D. Yatsenko, E. Gerasimov, A. Shalygin, O. Martyanov,
S. Tsybulya, Coherent 3D nanostructure of γ-Al2O3: simulation of whole X-ray
ions in the tested samples were analyzed using the crystal field theory. powder diffraction pattern, J. Solid State Chem. 246 (2017) 284–292.
The analysis of PL and PLE spectra showed that for Cr3+ ions in the series [15] M. Rudolph, M. Motylenko, D. Rafaja, Structure model of γ-Al2O3 based on planar
of polymorphs γ-, η-Al2O3 → α-Al2O3 the Dq/B value decreases more defects, IUCrJ 6 (2019) 116–127.
[16] M. Anpo, M. Che, Applications of photoluminescence techniques to the
than 1.3-fold (and more than 1.5-fold as compared to ruby crystal),
characterization solid surfaces in relation to adsorption, Catalysis and
which indicates that structural positions of Cr3+ ions in α-Al2O3 are more photocatalysis, Adv. Catal. 44 (1999) 119–257.
distorted in comparison with γ- and η-Al2O3. [17] A.B. Kulinkin, S.P. Feofilov, R.I. Zakharchenya, Luminescence of impurity 3d and
The PL and PLE spectra and PL decay curves were obtained for the 4f metal ions in different crystalline forms of Al2O3, Phys. Solid State 42 (2000)
857–860.
first time for Cr3+ ions located in the χ-Al2O3 structure. Similar PL [18] M. Anpo, S. Dzwigaj, M. Che, Applications of photoluminescence spectroscopy to
spectra and Dq parameters are observed for χ-Al2O3 and α-Al2O3 sam­ the investigation of oxide-containing catalysts in the working state, Adv. Catal. 52
ples. These observations show that the Me ion, which is represented by (2009) 1–42.
[19] A.A. Rastorguev, M.G. Baronskii, N.A. Zaitseva, L.A. Isupova, A.I. Kostyukov, T.
the Cr3+ ion in this case, occupies close local structural positions in the V. Larina, N.A. Pakhomov, V.N. Sytnikov, Photoluminescence properties of
low-temperature χ-Al2O3 and high-temperature α-Al2O3 phases. microspherical alumina-chromium catalyst, Inorg. Mater. Appl. Res. 5 (5) (2014)
In the series from α-Al2O3 to γB-Al2O3, the observed trend in the 476–481.
[20] M.G. Baronskiy, A.I. Kostyukov, T.V. Larina, V.N. Snytnikov, N.A. Zaitseva, A.
behavior of Racah parameter B (606 → 432 cm− 1) and the covalence V. Zhuzhgov, Photoluminescence of surface chromium centers in the Cr/Al2O3
degree β1 (1.08 → 1.06) indicates an increase in the fraction of covalent system that is active in isobutane dehydrogenation, Mater. Chem. Phys. 234 (2019)
bond for the low-temperature polymorphs γ- and η-Al2O3 as compared to 403–410.
[21] A.L. Schawlow, G.E. Devlin, Simultaneous optical maser action in two ruby
the α-phase and ruby crystal. The low-temperature polymorphs η-, γB-, satellite lines, Phys. Rev. Lett. 6 (1961) 96–98.
γPb- and χ-Al2O3 also have different values of Dq, B and β1 parameters, [22] C.M. MacRae, N.C. Wilson, Luminescence database I—minerals and materials,
which testify to different ionicity/covalence ratio in each of the low- Microsc. Microanal. 14 (2008) 184–204.
temperature Al2O3 polymorphs.

6
M.G. Baronskiy et al. Journal of Luminescence 242 (2022) 118554

[23] V.A. Pustovarov, V.S. Kortov, S.V. Zvonarev, A.I. Medvedev, Luminescent vacuum [35] M.G. Brik, S.J. Camardello, A.M. Srivastava, N. Avram, A. Suchocki, Spin-forbidden
ultraviolet spectroscopy of Cr3+ ions in nanostructured aluminum oxide, J. Lumin. transitions in the spectra of transition metal ions and nephelauxetic effect, ECS
132 (2012) 2868–2873. Journal of Solid State Science and Technology 5 (2016) R3067–R3077.
[24] Q. Wen, D. Lipkin, D. Clarke, Luminescence characterization of chromium- [36] M. Back, J. Ueda, H. Nambu, M. Fujita, A. Yamamoto, H. Yoshida, H. Tanaka, M.
containing θ-alumina, J. Am. Ceram. Soc. 81 (1998) 3345–3348. G. Brik, S. Tanabe, Boltzmann thermometry in Cr3+-doped Ga2O3 polymorphs: the
[25] D. Renusch, M. Grimsditch, J.D. Jorgensen, J.P. Hodges, Pressure dependence of structure matters, Adv. Optical Mater. 9 (2021), 2100033-1-11.
Cr3+ fluorescence in θ-alumina, Oxid. Metals 56 (2001) 299–311. [37] M.G. Brik, S.J. Camardello, A.M. Srivastava, Influence of covalency on the Mn4+
2
[26] A. Rastorguev, M. Baronskiy, A. Zhuzhgov, A. Kostyukov, O. Krivoruchko, Eg → 4A2g emission energy in crystals, ECS Journal of Solid State Science and
V. Snytnikov, Local structure of low-temperature γ-Al2O3 phases as determined by Technology 4 (2015) R39–R43.
the luminescence of Cr3+ and Fe3+, RSC Adv. 5 (2015) 5686–5694. [38] K. Drdlikova, R. Klement, D. Drdlika, D. Galusek, K. Maca, Processing and
[27] A. Kostyukov, M. Baronskiy, A. Rastorguev, V. Snytnikov, Vl Snytnikov, properties of luminescent Cr3+ doped transparent alumina ceramics, J. Eur. Ceram.
A. Zhuzhgov, A. Ishchenko, Photoluminescence of Cr3+ in nanostructured Al2O3 Soc. 40 (2020) 2573–2580.
synthesized by evaporation using a continuous wave CO2-laser, RSC Adv. 6 (2016) [39] C.A. Morrison, Crystal Field for Transition-Metal Ions in Laser Host Materials,
2072–2078. Springer, Berlin, 1992, p. 190.
[28] A.I. Kostyukov, A.V. Zhuzhgov, V.V. Kaichev, A.A. Rastorguev, VlN. Snytnikov, V. [40] R.G. Burns, Mineralogical Applications of Crystal Field Theory, second ed.,
N. Snytnikov, Photoluminescence of oxygen vacancies in nanostructured Al2O3, Cambridge University Press, Cambridge, 1993, p. 551.
Opt. Mater. 75 (2018) 757–763. [41] A.I. Kostyukov, M.G. Baronskiy, T.V. Larina, V.N. Snytnikov, N.A. Zaitseva, A.
[29] M. Baronskiy, A. Rastorguev, A. Zhuzhgov, A. Kostyukov, O. Krivoruchko, A. Pochtar, A.V. Ishchenko, S.V. Cherepanova, V.N. Snytnikov, Laser vaporized
V. Snytnikov, Photoluminescence and Raman spectroscopy studies of low- CrOx/Al2O3 nanopowders as a catalyst for isobutane dehydrogenation, Mater.
temperature γ-Al2O3 phases synthesized from different precursors, Opt. Mater. 53 Char. 169 (2020), 110664-1-14.
(2016) 87–93. [42] M. Gaft, R. Reisfeld, G. Panczer, Modern luminescence spectroscopy of minerals
[30] S. Adachi, Luminescence spectroscopy of Cr3+ in Al2O3, Opt. Mater. 114 (2021), and materials, Springer, 200, 355.
111000-111001-8. [43] D.L. Russell, K. Holliday, M. Grinberg, D.B. Hollis, Broadening of optical transitions
[31] J. Gangwar, B.K. Gupta, P. Kumar, S.K. Tripathi, A.K. Srivastava, Time-resolved in Cr31-doped aluminosilicate glasses, Phys. Rev. B 59 (1999) 712–718.
and photoluminescence spectroscopy of θ-Al2O3 nanowires for promising fast [44] R.-S. Zhou, R.L. Snyder, Structures and transformation mechanisms of the η-, γ- and
optical sensor applications, Dalton Trans. 43 (2014) 17034–17043. θ- transition aluminas, Acta Crystallogr. B47 (1991) 617–630.
[32] F. Auzel, G. Baldacchini, Photon trapping in ruby and lanthanide-doped materials: [45] M. Trueba, S.P. Trasatti, γ-alumina as a support for catalysts: a review of
recollections and revival, J. Lumin. 125 (2007) 25–30. fundamental aspects, Eur. J. Inorg. Chem. (2005) 3393–3403.
[33] Y. Zheng, H. You, G. Jia, K. Liu, Y. Song, M. Yang, H. Zhang, Facile hydrothermal [46] U.L. Bragg, G.F. Claringbull, Crystal Structure of Minerals, 1965, p. 409. London.
synthesis and luminescent properties of large-scale GdVO4:Eu3+ nanowires, Cryst. [47] M. Back, J. Ueda, M.G. Brik, S. Tanabe, Pushing the limit of Boltzmann distribution
Growth Des. 9 (2009) 5101–5107. in Cr3+-doped CaHfO3 for cryogenic thermometry, ACS Appl. Mater. Interfaces 12
[34] I.E. Kolesnikov, A.V. Povolotskiy, D.V. Mamonova, E. Lähderanta, A.A. Manshina, (2020) 38325–38332.
M.D. Mikhailov, Photoluminescence properties of Eu3+ ions in yttrium oxide
nanoparticles: defect vs. normal sites, RSC Adv. 6 (2016) 76533–76541.

You might also like