You are on page 1of 5

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: ACS Omega 2019, 4, 22114−22118 http://pubs.acs.org/journal/acsodf

Hibonite Blue: A New Class of Intense Inorganic Blue Colorants


Brett A. Duell, Jun Li, and M. A. Subramanian*
Department of Chemistry, Oregon State University, Corvallis, Oregon 97331, United States
*
S Supporting Information

ABSTRACT: Commercially available spinel cobalt blue (CoAl2O4) utilizes a significant


amount of carcinogenic Co2+, which makes its synthesis more hazardous and
environmentally harmful. Considerable effort has been put into developing more
environmentally benign and robust blue pigments to replace cobalt blue. A new class of
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

blue pigments with tunable hue were prepared. The solid solution series,
CaAl12−2xCoxTixO19 (0 < x ≤ 1), crystallizes in a hexagonal mineral hibonite
(CaM12O19) structure with five distinct crystallographic sites for M cations (M = Al,
Co, and Ti). The origin of intense blue color is attributed to a synergistic effect of
Downloaded via 101.14.225.108 on April 11, 2020 at 07:30:26 (UTC).

allowed d−d transitions involving the chromophore Co2+ in both tetrahedral and trigonal bipyramidal crystal fields. Compared
with commercial cobalt blue, these tunable hibonite blues possess a reddish hue that intensifies the blue color as observed in
Y(In,Mn)O3 (YInMn) blues, with a significant reduction of Co2+ concentration from 33% to as low as 4% by mass. A significant
advantage of hibonite blues over cobalt blue is the substantial reduction in carcinogenic cobalt content while enhancing the
color properties at a reduced cost for raw materials.

■ INTRODUCTION
For thousands of years, civilizations around the world have
an intense bright blue color because of tetrahedral Co2+ spin-
allowed d−d 4A2(4F)−4T1(4P) transitions between 500 and
sought inorganic materials that could be used to paint things 700 nm. CoAl2O4, which contains 33% Co2+ by mass, is
blue, often with very limited success.1,2 Discoveries of durable expensive and environmentally unfriendly to produce. Finding
new inorganic blue colorants are rare occurrences in materials replacement compounds that have the same durability,
science. The first ancient blue material discovered was lapis intensity, and hue of blue has proven even more difficult.
lazuli, which was about 6000 years ago. It was more expensive The intensity of color generally scales with the chromophore
than gold. Egyptians and Babylonians utilized naturally composition; however, reducing cobalt content is infeasible in
occurring lapis lazuli stones found in Afghanistan to make the CoAl2O4 spinel structure without significant loss in
rudimentary blues. It was not until 1826 a chemical process for coloration. The search for materials with similar local
developing synthetic lapis lazuli (now called ultramarine blue, environments for chromophores has led to the exploration of
Na7Al6Si6O24S3) was identified.2 In 1706, Prussian blue other crystal structures as color-producing hosts. The mineral
(Fe4[Fe(CN)6]3) was accidentally made by a German dye- hibonite, known as “blue angel”, has light blue coloration
maker Johann Jacob Diesbach.2 Although cobalt blue because of a charge transfer between Ti3+ and O2−. Hibonite,
containing mixtures had long been used in Chinese porcelain,3 with the general formula CaM12O19, has five distinct
the pure cobalt blue (CoAl2O4) pigment was independently crystallographic M sites, including three different octahedra,
synthesized by a French chemist in 1802.4 Nearly two a set of tetrahedra, and a set of trigonal bipyramids (TBPs),
centuries later, YInMn blue (YIn1−xMnxO3) was serendip- where the metal center, M, is in pseudotetrahedral
itously discovered when searching for magnetodielectric coordination, offset from the trigonal plane (Figure 1).
materials for electronics and sparked renewed worldwide Hibonites crystallize into a hexagonal crystal structure, with
interest in pigment discoveries.2,5 Most of these compounds spinel blocks identical to CoAl2O4 separated by an “R” block
suffer from stability, cost, color, or toxicity issues. The origin of with TBPs, face-shared octahedral M sites, and Ca2+ sites.7 The
color in ultramarine blue is from a charge transfer within an tetrahedral and TBP sites are of interest because they are
unstable S3− cluster, which decomposes under mild acidic noncentrosymmetric and therefore may produce intense
conditions and with time. Prussian blue is dark and dull transitions as found in YIn1−xMnxO3 and CoAl2O4.5,8 A
because of a charge transfer between Fe2+ and Fe3+ and is previous work has discovered many different colors in hibonite
unsuitable for many coloring applications. The YInMn blue compounds; however, the resulting blue colors are less intense
family of pigments are durable and possess remarkable heat than CoAl2O4.9−14
reflecting properties, which make them excellent candidates for We have investigated the effects of systematic Co2+/Ti4+
energy-saving coatings, nonetheless they are not considered substitution into the hibonite structure and have found that
cost-effective for widespread general coating applications.2,6
The intensity of the color, ease of synthesis, and wide Received: October 2, 2019
applicability have resulted in CoAl2O4 becoming a dominant Accepted: November 19, 2019
commercial blue pigment for the last 200 years. CoAl2O4 gives Published: December 14, 2019

© 2019 American Chemical Society 22114 DOI: 10.1021/acsomega.9b03255


ACS Omega 2019, 4, 22114−22118
ACS Omega Article

Figure 1. Schematic of the unit cell of hibonite. Symmetrically distinct


sites are shown in different colors: light green (M1) and yellow (M5):
edge-shared octahedra; red (M4): face-shared octahedra; light blue
(M2): TBPs; dark blue (M3): tetrahedra.

Co2+ gives rise to an intense, tunable blue coloration with up to


x = 1 Co2+ (CaAl12−2xCoxTixO19, 8% Co2+ by mass). This
coloration at such low Co2+ concentrations is characteristic
because of the unique effect of the cosubstituted Ti4+ within
the hibonite structure. This color is tunable through varying
Co2+ concentration as well as cosubstitution with other
transition metals, such as Ni2+, Mn3+, and Cr3+.

■ RESULTS AND DISCUSSION


Addition of Co2+ and Ti4+ increases the unit cell parameters Figure 2. (a) Lattice parameter evolution with Co2+/Ti4+ substitution
because of the increased size of Co2+ and Ti4+ over Al3+ (Figure in CoAl12−2xCoxTixO19 and (b) neutron Rietveld refinement of
2a).15 The deep blue hues seen in hibonites originate in the CaAl10CoTiO19.
strong site preference of Co2+ and Ti4+ substitution, affirmed
by structural analysis of neutron diffraction data of Table 1. Summary of Rietveld Refinement Occupancy
CaAl10.6Co0.7Ti0.7O19 and CaAl10CoTiO19 (Figure 2b). The Results of CaAl12−2xCoxTixO19
face-shared octahedral sites are the primary occupied sites of CaAl12O19a CaAl10.6Co0.7Ti0.7O19 CaAl10CoTiO19
Ti4+ for both compositions (Table 1). This is consistent with wRp (%) 4.72 4.57 4.60
what was found in the CaAl12−2xNixTixO19 hibonites, with the Χ2 1.57 1.85 1.92
primary difference being the increased site preference of Co2+ a (Å) 5.5592(1) 5.5970(1) 5.6106(2)
for the tetrahedral and TBP sites compared to Ni2+.9 Of the c (Å) 21.902(1) 22.042(1) 22.211(1)
three octahedral coordination environments available in the M1 (Al) 1 1 1
structure (Figure 1), only the face-shared octahedral site is M2 (Al) 0.5 0.369(6) 0.371(9)
capable of a site distortion.11 Because of the larger size and M2 (Co) 0.084(1) 0.087(3)
greater charge of Ti4+ over Al3+, a distortion must occur in M2 (Ti) 0.046(5) 0.042(6)
order to satisfy valency. Therefore, Ti4+ distributes primarily M3 (Al) 1 0.74(2) 0.65(4)
into the face-shared octahedral environment where the local M3 (Co) 0.26(2) 0.35(4)
structure around the site can allow the metal to displace away M4 (Al) 1 0.753(4) 0.605(7)
from the shared face of the other octahedron. Neutron M4 (Ti) 0.247(4) 0.395(7)
structure analysis reveals that 76 and 70% of Co2+ are in the M5 (Al) 1 0.981(2) 0.950(8)
tetrahedral site for CaAl10.6Co0.7Ti0.7O19 and CaAl10CoTiO19, M5 (Co) 0 0.024(4)
respectively. A significant amount of Co2+ also resides in the M5 (Ti) 0.019(2) 0.026(4)
TBP site, 24 and 17% of all Co2+ for CaAl10.6Co0.7Ti0.7O19 and a
From Li, et al.11
CaAl10CoTiO19, respectively, with no Co2+ occupying the M1
edge-shared octahedral or M4 face-shared octahedral sites in
either composition. Co2+ occupies the M5 edge-shared the tetrahedral environments require a Jahn−Teller distortion
octahedral site in small amounts only in the CaAl10CoTiO19 for d7; however, the tetrahedral site has a much higher percent
composition. of Co2+ in both compositions refined. This is attributed to the
The Co2+ ion displays a strong preference for the tetrahedral displacement of the metal center of the TBP away from the
coordination environment, even though the crystal field trigonal plane. This displacement creates a pseudotetrahedral
stabilization energy for octahedral Co2+ is higher. This may environment, which is smaller than the real tetrahedral site and
be because the d7 electron configuration of Co2+ favors a weak thus less likely to be occupied by the larger Co2+ ion. The total
Jahn−Teller distortion in octahedral coordination, which is TBP occupancy in both refined compositions is similar, despite
unavailable in the hibonite structure.16 Neither the TBP nor the difference in the molar percent of cobalt out of all sites.
22115 DOI: 10.1021/acsomega.9b03255
ACS Omega 2019, 4, 22114−22118
ACS Omega Article

This suggests that the TBP site becomes saturated by Co2+ at


low concentrations of Co2+.
This site preference is much more pronounced in
CaAl12−2xCoxTixO19 than in the previously investigated
CaAl12−2xNixTixO19 or Ca1−xLaxAl12−xNixO19.9 The high spin
d8 electron configuration of Ni2+ may contribute to this
reduced preference as octahedral crystal field stabilization
energy is much larger than in Co2+.
The color of these materials has been evaluated by L*a*b*
color measurements. L*a*b* uses three primary numbers to
quantify the brightness (L*), green-red (a*), and blue-yellow
(b*) values. Higher L* values are the result of brighter colors
with higher light reflectivity. Positive a* values correspond to
more red coloration while more negative b* values are more Figure 3. XRD patterns for as-synthesized CaAl10CoTiO19 and after
blue. Hibonite blue pigments give intense negative b* blue 12 h of 50% HNO3 soaking and 12 h of soaking in 1 M NaOH.
values rivaling that of CoAl2O4 and approaching that of YInMn
blue (Table 2). We have found that, while CoAl2O4 gives no stability; therefore, we believe that Ca(Al,Co,Ti) 12 O 19
hibonites are also stable in most pH environments.
Table 2. L*a*b* Color Coordinates of Neutron structure analysis reveals that 70% of Co2+ is in the
Ca(Al,Co,Ni,Ti)12O19 Samples Compared to Other Well- tetrahedral site. The origin of the color of Ca(Al,Co,Ti)12O19
Known Blue Compounds hibonite blue is therefore primarily due to a strong d−d
transition of tetrahedral Co2+ between 500 and 675 nm (Figure
composition % mass Co L* a* b* 4a). The spectroscopy of Co2+ in tetrahedral coordination of
CaAl11Co0.5Ti0.5O19 4.24 45.86 1.12 −38.34
CaAl10.8Co0.6Ti0.6O19 5.05 44.94 3.95 −42.24
CaAl10.6Co0.7Ti0.7O19 5.85 38.26 0.52 −39.47
CaAl10.4Co0.8Ti0.8O19 6.64 38.26 0.52 −39.85
CaAl10.2Co0.9Ti0.9O19 7.41 39.99 0.61 −38.51
CaAl10CoTiO19 8.18 29.91 1.07 −33.57
CaAl10Co0.8Ni0.2TiO19 6.54 42.52 −3.44 −36.87
CaAl10Co0.5Ni0.5TiO19 4.09 43.04 −4.39 −36.21
CaAl10Co0.2Ni0.8TiO19 1.64 50.08 −8.78 −31.37
CaAl10CoTiO19−HNO3c 8.18 30.19 0.92 −32.92
CaAl10CoTiO19−NaOHc 8.18 30.63 0.24 −31.28
CaAl10NiTiO19a 49.59 −14.72 −28.44
CoAl2O4b 33.31 43.51 −4.46 −44.39
lapis lazuli stone 30.06 7.51 −24.21
YIn0.8Mn0.2O3 34.47 9.08 −44.39
YIn0.9Mn0.1O3 40.00 11.90 −47.90
a
From Li, et al.9 bSample from Shepherd Color Company. cData
collected after soaking in 50% HNO3 or 1 M NaOH.

red hue like YInMn blue, Ca(Al,Co,Ti)12O19 substituted


hibonite blues to give more reddish hues, approaching that
of lapis lazuli stones and YInMn blue. These reddish hues are
unobtainable with CoAl2O4 spinel, as reduction in Co2+
content by substitution in the spinel system only reduces the
intensity of blue coloration and has no effect on hue.
Introduction of other chromophores, such as Ni2+, into the
Ca(Al,Co,Ti)12O19 hibonites reduces this redness and adds
certain greenish hues to the hibonite blue. Material brightness
can often be tuned through chromophore concentration;
Figure 4. (a) UV−vis absorbance and (b) NIR reflectance of
however, only recently, Y(In,Mn,Ti,Zn)O3 has been shown to CaAl12−2xCoxTixO19 samples. Commercial Co blue (CoAl2O4) was
have a tunable hue.17,18 This tunability is rare in blue pigments measured for comparison.
and shows that hibonite blue can be used as a versatile pigment
with multiple hues to suit various needs. The CaAl10CoTiO19
composition was chosen for acid and base stability tests. oxides has been well studied, and the broadening of the
4
Specimens of CaAl10CoTiO19 were soaked in either 50% A2(4F)−4T1(4P) Co2+ transition has been explained through
HNO3 or 1 M NaOH for 12 h, filtered, and examined via spin−orbit interactions of the d7 ions.19 Typically, this gives
L*a*b* color measurements and X-ray diffraction (XRD). rise to three noticeable transitions in the visible region, as is the
XRD patterns (Figure 3) show no change in structure and case of CoAl2O4. However, we found that Ca(Al,Co,Ti)12O19
L*a*b* measurements reveal no change in color. This is hibonite has four transitions within the same 500−675 nm
consistent with other hibonites, which have high acid/base region. This difference may be due to the distortion of the
22116 DOI: 10.1021/acsomega.9b03255
ACS Omega 2019, 4, 22114−22118
ACS Omega Article

Figure 5. Images of selected CaAl12−xCoxTixO19 compounds and other commercially available powders.

tetrahedral site in the hibonite that cannot exist in the cubic Discoveries of blue pigments are still scarce as no criteria for
CoAl2O4 spinel.20 This, however, does not explain the nearly investigations of novel systems exist. Despite the abundance of
identical intensity of the color of hibonite blue compared to weakly blue cobalt-containing compounds, we could not have
CoAl2O4. Despite a decrease in Co2+ concentration from 33% foreseen the intensity of color and redness coming from
by mass to as low as 4% by mass, hibonite blue maintains a hibonite blue. Predicting the exact color of new materials
dramatic blue coloration. Both the tetrahedral and TBP Co2+ remains the realm of intuitive guess-work rather than science.23
may give rise to the intense absorption between 500 and 675
nm, as the TBP environment is more accurately described as a
set of disordered face-shared tetrahedra, where the disordered
■ METHODS
Stoichiometric amounts of CaCO3 (Sigma-Aldrich, 99.0%),
site is offset from the face by approximately 0.28 Å. These Al2O3 (Cerac, 99.99%), CoCO3 (Baker Analytical, 99%), and
pseudotetrahedra are smaller in volume than the tetrahedral TiO2 (Cerac, 99%) were ground in an agate mortar, pressed
site, giving the larger Co2+ ion preference for the tetrahedral into pills, and heated between 1623 and 1773 K over 12 h.
site. The combined TBP and tetrahedral Co2+ then gives the Intermittent grinding and reheating were implemented to
spectrum seen in the Ca(Al,Co,Ti)12O19 hibonite. Indeed, ensure sample purity and homogeneity. Compositions for the
substitution of Ni2+ for Co2+ reduces this absorption because formula, CaAl12−2xCoxTixO19, were x = 0.2, 0.4, 0.5, 0.6, 0.7,
of the decrease in site preference of Ni2+ for the tetrahedral and 0.8, 0.9, and 1.0. Phase purity and unit cell parameters were
TBP environments as well as the difference in electron determined using powder XRD on a bench-top Rigaku
configuration.9 The brightness of hibonite blues (Figure 5) is Miniflex II powder diffractometer with Cu Kα radiation and
affected by both the Co2+ and Ti4+ substitutions. The strong a graphite monochromator. Neutron powder diffraction data
absorption in the UV region (250 nm) is the result of a metal− were obtained from the Center for Neutron Research at the
ligand charge transfer between Ti4+ and O2−, which at higher National Institute of Standards and Technology to investigate
Ti4+ concentrations tails into the visible region, darkening the chemical composition and site occupancies. Samples were
color. This effect is also seen in other hibonites. loaded into vanadium canisters, and data were collected at
Lastly, we investigated the usefulness of hibonite blue as a room temperature using a Cu(311) monochromator yielding a
near-infrared (NIR) reflective material. NIR reflectivity is a 1.54 Å wavelength. Initial structures were refined using GSAS
property necessary for compounds to be used as energy-saving EXPGUI software. Diffuse reflectance spectroscopic data were
heat reflection colorants (“cool pigments”) and is the subject collected in the UV−vis region on an in-house spectrometer
of increasing interest.21 Current blue pigments containing with MgO as a reference. The reflectance data were converted
cobalt, such as CoAl2O4, have very low NIR reflectivity. This is to absorbance with the Kubelka−Munk equation. NIR
due to two d−d transitions for tetrahedral Co 2+ , reflectance measurements were performed using a Jasco V-
4
A2(4F)−4T2(4F) at 1400 nm and 4A2(4F)−4T1(4F) at 1600 670 spectrometer up to 2500 nm. Color coordinates in the
nm.22 The high concentration of Co2+ in CoAl2O4 gives a large CIELAB color space were measured using a Konica Minolta
absorption in the NIR region. Thus, most Co2+-containing CM-700d spectrophotometer. Samples of CaAl10CoTiO19
compounds have very low NIR reflectivity, making them were soaked in 50% HNO3 and 1 M NaOH overnight and
unsuitable for energy-saving coatings. The hibonite blues (e.g., dried, and their color properties and phase purity were
determined via XRD in order to determine acid and base
CaAl11Co0.5Ti0.5O19) show an increase in NIR reflectivity by as
resistance.


much as 60% between 1200 and 1600 nm and as high as 125%
increase between 800 and 1000 nm while still maintaining the
intense blue color (Figure 4b). ASSOCIATED CONTENT


*
S Supporting Information

CONCLUSIONS The Supporting Information is available free of charge at


https://pubs.acs.org/doi/10.1021/acsomega.9b03255.
We have found that hibonite blues are stable in both strongly
acidic and basic environments, with no change in structure or Description of L*a*b* color space, magnetic suscepti-
color after treatment. This rivals the durability of both cobalt bility data and results, and detailed Rietveld refinement
blue and YInMn blue. Furthermore, with increasing Co/Ti results of CaAl10.6Co0.7Ti0.7O19 and CaAl10CoTiO19
content, there is increasing ultraviolet absorbance because of (PDF)


the Ti−O charge transfer. UV radiation is a common
degradant of polymer coatings. Additives, such as inorganic AUTHOR INFORMATION
UV-absorbing pigments, are added to reduce this organic
Corresponding Author
degradation. Both cobalt blue and YInMn blue were shown to
be useful as functional colorants in this regard.6 Similarly, the *E-mail: mas.subramanian@oregonstate.edu.
UV absorption of hibonite blue makes it an excellent candidate ORCID
for polymer-based coating applications. M. A. Subramanian: 0000-0001-5487-043X
22117 DOI: 10.1021/acsomega.9b03255
ACS Omega 2019, 4, 22114−22118
ACS Omega Article

Author Contributions Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976,
M.A.S. conceived the research interest. B.A.D. carried out the 32, 751−767.
synthesis. B.A.D. and J.L. performed neutron refinements and (16) Burns, R. G. Mineralogical Applications of Crystal Field Theory,
2nd ed.; Cambridge University Press: New York, 1993; pp 243−249.
performed optical property measurements. M.A.S., B.A.D., and
(17) Li, J.; Lorger, S.; Stalick, J. K.; Sleight, A. W.; Subramanian, M.
J.L. discussed the structure characterization and optical A. From Serendipity to Rational Design: Tuning the Blue Trigonal
properties. All authors contributed to the drafting of the Bipyramidal Mn3+ Chromophore to Violet and Purple through
manuscript. Application of Chemical Pressure. Inorg. Chem. 2016, 55, 9798−9804.
Funding (18) Li, J.; Subramanian, M. A. Inorganic pigments with transition
metal chromophores at trigonal bipyramidal coordination: Y(In,Mn)-
This work was supported by NSF Grant DMR-1508527. O3 blues and beyond. J. Solid State Chem. 2019, 272, 9−20.
Notes (19) Bates, T. Ligand field theory and absorption spectra of
The authors declare no competing financial interest. transition-metal ions in glasses. Mod. Aspects Vitreous State 1961, 2,


195−254.
(20) Bamford, C. R. The application of ligand field theory to colored
ACKNOWLEDGMENTS glasses. Phys. Chem. Glasses 1962, 3, 189−202.
The authors thank Dr. Judith Stalick for neutron data (21) Jose, S.; Joshy, D.; Narendranath, S. B.; Periyat, P. Recent
collection at NIST Center for Neutron Research. The advances in infrared reflective inorganic pigments. Sol. Energy Mater.
Sol. Cells 2019, 194, 7−27.
identification of any commercial product or trade name does
(22) Llusar, M.; Forés, A.; Badenes, J. A.; Calbo, J.; Tena, M. A.;
not imply endorsement or recommendation by the National Monrós, G. Colour analysis of some cobalt-based blue pigments. J.
Institute of Standards and Technology. Eur. Ceram. Soc. 2001, 21, 1121−1130.

■ REFERENCES
(1) Ball, P. Bright Earth: The Art and Invention of Color, The
(23) Orna, M. V. Chemical origins of color. J. Chem. Educ. 1978, 55,
478.

Chicago University Press: Chicago, 2003.


(2) Kupferschmidt, K. In search of blue. Science 2019, 364, 424−
429.
(3) Rose, R.; Wood, N. Science and Civilisation in China; Cambridge
University Press, 2004; Vol. 5, Part 12, Ceramic Technology
(Review), pp 658−692.
(4) Gehlen, A. v. F.; Thenard, V. v. B. Ueber die Bereitung einer
blauen Farbe aus Kobalt, die eben so schön ist wie Ultramarin. Allg. J.
Chem. 1803, 2.
(5) Smith, A. E.; Mizoguchi, H.; Delaney, K.; Spaldin, N. A.; Sleight,
A. W.; Subramanian, M. A. Mn3+ in Trigonal Bipyramidal
Coordination: A New Blue Chromophore. J. Am. Chem. Soc. 2009,
131, 17084−17086.
(6) Smith, A. E.; Comstock, M. C.; Subramanian, M. A. Spectral
properties of the UV absorbing and near-IR reflecting blue pigment,
YIn1-xMnxO3. Dyes Pigm. 2016, 133, 214−221.
(7) Hibonite: Hibonite Mineral Information and Data, Hudson
Institute of Minerology. www.mindat.org (accessed August 2019).
(8) Laporte, O.; Meggers, W. F. Some Rules of Spectral Structure. J.
Opt. Soc. Am. 1925, 11, 459−463.
(9) Li, J.; Medina, E. A.; Stalick, J. K.; Sleight, A. W.; Subramanian,
M. A. Colored oxides with hibonite structure: A potential route to
non-cobalt blue pigments. Prog. Solid State Chem. 2016, 44, 107−122.
(10) Medina, E. A.; Li, J.; Subramanian, M. A. Colored oxides with
hibonite structure II: Structural and optical properties of CaAl12O19-
type pigments with chromophores based on Fe, Mn, Cr and Cu. Prog.
Solid State Chem. 2017, 45-46, 9−29.
(11) Li, J.; Medina, E. A.; Stalick, J. K.; Sleight, A. W.; Subramanian,
M. A. Structural studies of CaAl12O19, SrAl12O19, La2/3+δAl12−δO19,
and CaAl10NiTiO19 with the hibonite structure; indications of an
unusual type of ferroelectricity. Z. Naturforsch., B: J. Chem. Sci. 2016,
71, 475−484.
(12) Leite, A.; Costa, G.; Hajjaji, W.; Ribeiro, M.; Seabra, M.;
Labrincha, J. Blue cobalt doped-hibonite pigments prepared from
industrial sludges: Formulation and characterization. Dyes Pigm. 2009,
81, 211−217.
(13) Hajjaji, W.; Seabra, M. P.; Labrincha, J. A. Recycling of solid
wastes in the synthesis of Co-bearing calcium hexaluminate pigment.
Dyes Pigm. 2009, 83, 385−390.
(14) Subramanian, M. A.; Sleight, A. W.; Li, J. Compounds
Comprising a Hibonite Structure and a Method for Their Use. U.S.
Patent 9,758,385 B2, Sept 12, 2017.
(15) Shannon, R. D. Revised effective ionic radii and systematic
studies of interatomic distances in halides and chalcogenides. Acta

22118 DOI: 10.1021/acsomega.9b03255


ACS Omega 2019, 4, 22114−22118

You might also like