Next Article in Journal
Synthesis of High-Crystallinity Mg-Al Hydrotalcite with a Nanoflake Morphology and Its Adsorption Properties for Cu2+ from an Aqueous Solution
Previous Article in Journal
Copper(I)/Triphenylphosphine Complexes Containing Naphthoquinone Ligands as Potential Anticancer Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Coloration of Spinel CoAl2O4 Cobalt Blue Pigments: Composition, Structure, and Cation Distribution

1
School of Packaging and Materials Engineering, Hunan University of Technology, Zhuzhou 412007, China
2
National & Local Joint Engineering Research Center for Advanced Packaging Material and Technology, Hunan University of Technology, Zhuzhou 412007, China
*
Author to whom correspondence should be addressed.
Inorganics 2023, 11(9), 368; https://doi.org/10.3390/inorganics11090368
Submission received: 5 August 2023 / Revised: 6 September 2023 / Accepted: 8 September 2023 / Published: 14 September 2023
(This article belongs to the Section Inorganic Materials)

Abstract

:
Cobalt blue ceramic pigments mainly consisting of CoAl2O4 are subject to the difficulty of color control. Here, a perspective is reported regarding research on the reasons for color change based on the control of the heat treatment and ratio of components. Macroscopically, the composition of pigment powders determines the color. Microscopically, the crystallite characters including size, cation distribution, and structure have an important effect on the color. The ingredient, structural, and color properties of the pigment powders are analyzed using thermo gravimetry–differential scanning calorimetry (TG–DSC), X-ray diffraction (XRD) measurement, Rietveld refinement, energy dispersive spectrometer (EDS), and colorimetry analysis. The color is proven to be associated with cation distribution, such as that of Co2+ and Co3+. It is indicated that high heating temperature, long heating time, and a large proportion of Al3+ can, respectively, induce the Co2+ and Al3+ in tetrahedral and octahedral sites.

1. Introduction

Spinel, a class of cubic crystal structure, is widely researched and applied in many fields related to thermal and chemical stability [1,2,3] such as catalysts [4,5], pigments [6,7], and refractories. The general formula of spinel could be described as AB2O4 (shown in Figure 1), in which ions A and B occupy the tetrahedral and octahedral fully or partly [8,9]. Ions A and B are usually divalent, trivalent, or quadrivalent transitional cations, including zinc, magnesium, titanium, iron, chromium, cobalt, and aluminum, etc.
CoAl2O4 powders acting as the main composition of cobalt blue ceramic pigments are widely used in the ceramics industry [1,10,11,12,13,14]. Cobalt blue pigment is an excellent choice for use in underglaze ceramic products because it possesses high stability in coloration and phase at a high temperature, even above 1360 °C [15]. Thus, it can be utilized in under-glazed ceramic products, which should have physicochemical and colorimetry stability under a high temperature, over 1300 °C. In recent years, many preparations of CoAl2O4 have been introduced, including the co-precipitation method [16,17,18], solution combustion [19,20], grinding and calcining method, molten salt decomposition method, sol–gel method [21], hydrothermal method [22], etc. Among these methods, co-precipitation and sol–gel methods prove helpful for a tiny crystallite size, high purity, and few side reactions.
Pigments mainly composed of CoAl2O4 are reported to exhibit different colors, not only sky-blue or Thenard’s blue. From the research of Gaudon et al., the pigments present blue and cyan colors after doping Zn2+ and Ni2+ in the spinel structure [23]. Even the CoAl2O4 can show the color change from a pink to pale violet color when certain numbers of Co2+ substitute the Al3+ in octahedral sites and the crystallite size is little enough [24]. Apart from the key reason, cation distribution and coloration also change with the crystallite scale from 0.1 to 10 nm [25,26].
With the large scale and sophisticated development of the ceramics industry, pigments with a high stability, fine particle size, and an easily controlled color have lots of advantages, but nowadays many factories are suffering from the color inconsistency of different batches of pigments. The main shortcoming of the current CoAl2O4 pigments is that the coloration [27], including hue and saturation, is hard to control; even the discipline of color based on the crystal structures is indistinct. If this problem is solved effectively, not only color accuracy will be improved, but also the cost of raw materials and thermal treatment [11,13,28,29,30] during aluminum over-stoichiometric spinel synthesis can be cut down.
The reason for color differences consists of two main parts. On the macro-level, related impurities from side reactions are generated during the preparation process. The impurities are associated with the selection of raw materials, the temperature of heat-treatment, pH value, etc. On the micro-level, the free Co2+ and Al3+ cannot stay in the tetrahedral and octahedral sites fully and respectively. In addition, Co2+ can transform into Co3+ and then occupy both the tetrahedral and octahedral sites partly, which will cause changes in the crystal structure including lattice parameter, bond length, ion radii, etc.
In this paper, CoAl2O4 powders are elaborated by the co-precipitation method and sintered under different temperatures (250, 500, 800, 1200, 1400 °C). The macroscopical and microscopical factors are deeply discussed including composition, cation distribution, and cation transfer trend related to the increasing ratio and heating temperatures and time. Based on these, the color effects and changing trends corresponding to the ratio and heating condition are obviously reported. These can be a color prediction of the high-temperature CoAl2O4 pigments and a reference for the doped, cobalt blue pigments through controlling the preparation condition during the fabrication process.

2. Results and Discussion

2.1. Crystal Phases after Heat Treatment at Different Temperatures

The precursors of cobalt blue pigments are analyzed to figure out the thermal influences on pigments using TG–DSC, which can describe the thermodynamic changes during the heat treatment. Figure 2 demonstrates the TG–DSC curve of the precursor sample sintered in an air atmosphere. Two ranges of room temperature—240 °C and 310–430 °C present obvious weight loss, which could represent the loss of free and bound water, respectively [31]. At the same ranges, the DSC curve does not show any endothermic peaks. In the range from 240 to 310 °C, the stair of weight loss indicates the transformation of precursors to the crystal phase [32]. From the relative DSC curve, a specific exothermic peak occurs at 282 °C, probably related to the burning of ethanol [33]. The weight loss in the TG curve turns to be stable when the temperature is above 480 °C, which also indicates the gradually stable phase and the formation of a solid solution with cobalt and aluminum ions. The reaction rate declines as the temperature increases, and the exothermal maximum in the DSC curve occurs at 943 °C, while the reaction does not stop completely. Combined with the TG results, it demonstrates that the sample transforms from amorphous phases to relatively stable phases gradually [34]. The loss of weight during the heat-treatment is 20.75%, and the low weight loss (2.82%) in the range of 1000–1400 °C reveals the high thermal stability of the underglaze pigments.
The XRD spectra of cobalt blue pigment powders after the heat treatment at different temperatures are presented in Figure 3. According to JCPDS-ICDD cards [35], the as-fabricated CoAl2O4 powders are composed with a relatively pure spinel structure (Fd-3m space group) at 1400 °C. Obviously, the higher the temperature the powders sinter at, the sharper the peaks it will present in the patterns, which means the formation of purer spinel structural phases and a larger particle scale. The peaks on the XRD pattern of the material heat-treated at 250 °C indicate the beginning of the transformation to the spinel structure. In the temperature range of 800–1200 °C, substantial spinel phases form but the phase transformation is not obvious, corresponding with the decrease in the reaction rate seen in the DSC curve (Figure 2).
As represented in Figure 4, the EDS spectrum also confirms the composition of the as-prepared powders at 1400 °C. The EDS spectrum and elemental composition (Table 1) present the composition and the ratio of the pigment powders, agreeing with the XRD results. The ratio of Co and Al still remains as about 1:2, which is in accordance with the experimental design. According to JCPDS-ICDD cards [35,36], the peaks (220), (311), (400), (331), (422), (511), and (440) mainly correspond to a cubic spinel structure. Beginning from 1200 °C, the peak (331) occurs and becomes distinct at 1400 °C, which reveals the formation of the main phase, CoAl2O4. The powders heated at 250 °C, 500 °C, and 800 °C are mainly with a spinel phase, with ions occupying the tetrahedral and octahedral positions. However, it seems to be hard to determine the specific phase for diverse cation types. As the heat temperature rises, the main phase transfers to a concrete phase, CoAl2O4, gradually. The ratio of the second strongest to the strongest peak is introduced to distinguish two possible phases, Co3O4 and Co2AlO4, because they have similar XRD spectra, and the same Miller index but different intensities [36]. The former I(220)/I(311) is 0.7 and the latter one is 0.309. Different phases after heat treatment at different temperatures make the pigment powders present different colors. Essentially, the hue is influenced by the cation distribution of Co2+ and Co3+ in the spinel framework with a Fd-3m space group. Both the existence of Co3O4 (one Co2+ ion in the tetrahedral site and two Co3+ ions in the octahedral site) and Co2AlO4 (one Co2+ ion in tetrahedral site and one Co3+ ions in octahedral site) can affect the purity and coloration of cobalt blue pigments. Co3O4 and Co2AlO4 will make the color of pigments darker and greener, respectively. The XRD spectra show that the fabrication temperature of pure CoAl2O4 pigment powders should be 1200 °C or more. Because of the low diffusion rate, many Co2+ and Co3+ ions are in tetrahedral and octahedral positions, respectively, at temperatures below 1200 °C, opposite to the ideal phase structure.

2.2. Influence Factors of Structural Properties

Composite pigments primarily consist of transition metals from the d-block of the periodic table and the molecules or ions (ligands) that bind to these metals. The fundamental source of its color is the absorption of visible light to produce d–d electronic transitions. The coloration of CoAl2O4 is mainly determined by the central ions in the tetrahedral and octahedral coordination fields. Concerning the CoAl2O4 spinel, they are Co2+ and Al3+, but Al3+ is colorless because of the completely empty d electron shell. The main reason for the coloration of CoAl2O4 is Co2+.
The appearance of peaks (331) reveals the formation of CoAl2O4, and the purity of CoAl2O4 rises as the sintering temperature increases. Thus, the main reason for the coloration is related to the crystallite structure.

2.2.1. Crystallite Size

Table 2 presents the crystallite size, Xs, of the pigment powders of different batches. The error of the size estimation is 6%. Over 1200 °C, the crystallite size grows with the temperature and time increasing. As is known from previous research, the crystallite size of pigments can affect the optical performance if the scale is between 0.1 and 10 nm [25,26]. From the analysis data, the scale of the crystallite is not enough to influence the coloration.

2.2.2. Cation Distribution

The formation of a cobalt blue pigment spinel structure for the (CoxAl1−x)[Al1+xCo1−x]O4 (0 ≤ x ≤ 1) system was confirmed by XRD patterns in Figure 5. The fitting profiles and R-factor of the samples are shown as an example in Figure 6. The result is obtained from Retvield refinement using the program Jade. The cation distribution data, including cation position and lattice parameter, are shown in Table 3. Other structural parameters are included in Table 4 for reference. The ratio of cobalt and aluminum is 1:2 in the sample, and the prescription of Sample 2 contains excessive aluminum.
For the normal ratio part, a high heating temperature and long heating time are both related to the increase in the lattice parameter. Compared with the normal ratio samples, the amount of Co3+ transferring from the octahedral site to the tetrahedral site enlarges in the crystallite of sample 2. At the same time, the Co3+ converts into divalent Co, but the quantity of Co3+ in octahedral sites grows correspondingly. Compared with the sample heated at 1400 °C, the stoichiometric ratio, cation position, and lattice parameter of sample 1 at 1200 °C are all close to those of Co2AlO4 because the Co2AlO4 (described as (Co)[CoAl]O4) has one Co2+ in the tetrahedral site and one Co3+ and one Al3+ in the octahedral site, with a smaller lattice parameter than CoAl2O4. Moreover, whether the heating time or temperature increases, the main phase can change into CoAl2O4 in line with cation distribution. The result confirms the analysis of Figure 3 which indicates the formation of CoAl2O4 for the first time at 1200 °C, becoming purer and purer as the temperature rises. Focusing on the cation distribution of sample 1 heated at 1200 °C for 3 h and 1400 °C for 1 h, the data demonstrate the similar phase and structural properties of them. Moreover, the increase in the aluminum ratio can also promote the occupation of Co2+ in the tetrahedral site and Al3+ in the octahedral site. Throughout the increases in the heating time, temperature, and Al element ratio, they all enhance the Al3+ substitution into the Co3+ octahedral site, and “push” a part of the Co ions into tetrahedral sites, with the preference of “pushing” the Co3+ toward the octahedral.
Combined with Table 2, heating sample 1 at 1200 °C for 3 h and 1400 °C for 1 h presents a similar phase. At a low temperature, sample 1, nevertheless, obtains a thinner crystallite size, larger crystalline interfaces, and more curved grain boundaries which are unfavorable for crack propagation and development, resulting in better strength and toughness [37,38]. And, increasing the ratio of aluminum can encourage the cation transfer to the normal position in the crystallite to obtain high-quality cobalt blue pigments. Some researchers also demonstrate that the Co/Al ratio should not exceed 0.3 to obtain pure CoAl2O4 [39]. Gaudon et al. report the same results and explain that the excessive Al3+ ratio prevents the occurrence of Co ions in octahedral sites due to its higher octahedral site preference energy [23].
Cation distribution is significant to the coloration of cobalt blue and many other kinds of pigments. From our research, controlling the heat treatment and doping of excessive ions in the formula prove effective. To fabricate other relational colors, some cations also have effects on the coloration, such as Ni2+ and Mg2+. They can substitute the Co2+ in tetrahedral sites and endow the crystallite with a cyan hue [23]. Hence, cation substitution should be researched in the future to prepare various pigments, and cation distribution may be helpful to control and predict the coloration of pigments. Alongside the factors mentioned above, cation substitution should also be considered in the preparation of the solid solution, not only for the color, but also for other properties like magnetic and catalytic performance.

2.3. Color Analysis

The selected colorimetry analysis system is CIE-L*a*b* value, and the test result is shown in Table 5. L* represents the brightness, with a value from 0 to 100; a* stands for the component from green to red, with a value from −128 to 127; b* stands for the component from blue to yellow, with a value from −128 to 127 [40].
The colorimetric analysis of the samples after heat treatment at different temperatures and time with different ratios are presented. The color of the samples heated at 500 and 800 °C appears close to black, owing to the Co3O4 acting as the main phase. Starting from 1200 °C, the coloration of samples turns from black to blue, which corresponds to the formation of CoAl2O4. This is consistent with XRD patterns in Figure 2. Samples heated at 1200 °C for 3 h and 1400 °C for 1 h present analogous L*a*b* values, which can confirm the results of XRD refinement data, i.e., the similar phases of the two samples. The increase in the aluminum ratio proves effective in color because the value b* shifts obviously toward blue, which is also in agreement with the report by Gaudon [41]. Throughout all colorimetric values of rising temperature, high temperature and an appropriate ratio provide the best conditions for the formation of CoAl2O4.
In essence, the cation distribution should combine with the color. Cobalt, the color element in the spinel, decides on the coloration depending on the valence and site. The distributions of Co2+ and Co3+ in tetrahedral or octahedral sites do not have an obvious connection with the L* value. In another word, the cation distribution in the cobalt spinel could have nothing to do with the brightness. The value of a* has a little bit of connection with the position of Co3+, but this little influence is not specific to the naked eye in our samples. The degree of blue is mainly related to the b* value. It proposed that if too many Co2+ occupy the tetrahedral sites, the b* value can be smaller.
Diffuse absorbance spectra are an auxiliary approach to the colorimetry analysis of spectroscopic data, and also the variation in the cobalt ions. Figure 7 presents the UV–vis absorption spectra of cobalt blue pigments with diverse formulas, heat temperatures, and heat times. All the spectra show the intensive absorption of visible light in the range of 500–700 nm, especially for the ones heated with high temperatures, which corresponds to the blue color region. Moreover, the absorptions at about 380 and 590 nm represent Co3+ in octahedral sites and Co2+ in tetrahedral sites, respectively [42]. This is also characterized as a Co3O4 cluster due to two board bands at around 380 and 700 nm for the samples sintered at 500 °C [43,44]. As the temperature increases, these absorptions turn out to be weak. The maximal absorptions at near 540, 590, and 640 nm turn to become clear with the heat temperature increase, which indicates the formation of CoAl2O4. Additionally, the decline in absorption at about 380 nm shows the decrease in Co3+ in octahedral sites, resulting in a reduction in the green shade.

3. Experimental

3.1. Synthesis Process

Figure 8 illustrates the fabrication process of samples in this study. With a molar ratio of 1:2, cobalt acetate (C4H6CoO4·4H2O, Aladdin Corp., Shanghai, China) and aluminum nitrate (Al(NO3)3·9H2O, Aladdin Corp., Shanghai, China) were dissolved separately in water, and then they were mixed together. Under magnetic stirring at 500 rpm, NH3·H2O was added dropwise to reach a pH of about 9; however, the quantity of the NH3·H2O can be a little more than needed. Then, we continued to stir for 2.5 h to mix them homogeneously. The precursors were separated with a centrifugal machine, and then the precipitates were washed with absolute ethanol to neutralize them in order to obtain cobalt and aluminum hydroxide alcohol mixtures. Then, the jelly-like gel was subject to heat treatment and sintering at 250 °C, 500 °C, 800 °C, 1200 °C, and 1400 °C for 1 h each. The fabrication process was repeated with an excessive proportion of aluminum (ratio of cobalt and aluminum is 1:3) to obtain another batch of precursors. Then, they were heated under different temperatures.

3.2. Characterization

The as-prepared powders were characterized by thermo gravimetry–differential scanning calorimetry (TG–DSC, STA 449 F5, Netzsch, Selb, Germany), X-ray diffraction (XRD, D/MAX2500VL/PC, Rigaku, Tokyo, Japan), energy dispersive spectroscopy (EDS, Hitachi S-800, Hitachi, Tokyo, Japan), and a spectrodensitometer (FD-5, Konica Minolta, Tokyo, Japan). XRD measurements were carried out with a X-celerator detector, using Cu(Kα1/Κα2) radiation. Diffractograms were analyzed via the Rietveld refinement method with a software program named Jade 6.0. Debye Scherrer’s Equation (1) was used with the widths of the peak to evaluate crystallite sizes [30].
D = K γ B cos θ
where D is the average thickness of the crystal grain perpendicular of the crystal plane, K is the Scherrer constant, γ is the X-ray wavelength (generally 1.54056 Å), B is the half-height width or integral width of the measured sample diffraction peak, and θ is the Bragg angle in degrees.

4. Conclusions

In this paper, we have prepared cobalt blue pigments using the coprecipitation method, where the heating temperature, time, and ratio of cobalt and aluminum have been set as the variables. The as-fabricated samples were investigated using TG–DSC, XRD, EDS, and colorimetry analysis. Then, Rietveld refinement was implemented to obtain the crystallite data, including the crystallite size, cation distribution, chemical formula, and lattice parameter.
The color change is related to macro and micro perspectives. On the one hand, the color is corresponding to the composition. When the heating temperature is under 800 °C, Co3+ exists in the octahedral site of spinel, so the color appears to lack a blue hue. On the other hand, when the sintering temperature comes to 1200 °C or more, the phase of CoAl2O4 will become purer. Co2+ in the tetrahedral sites can make the b* value small, which indicates that the color of the samples tends to be blue. The cation distribution could be controlled by the heat treatment and stoichiometric ratio. High heat-treatment temperature, long heat-treatment time, and excessive aluminum content play important roles in controlling the cation positions for pure cobalt blue pigments. The results have the potential for energy saving and color prediction in the fabrication process of cobalt blue and other related Co-based pigments.

Author Contributions

Methodology, W.Z. (Weiran Zhang), Z.L. and G.W.; software, W.Z. (Weiran Zhang), Z.L., W.W. and H.Z.; resources, G.W.; writing—original draft preparation, W.Z. (Weiran Zhang); writing—review and editing, W.Z. (Weiran Zhang), W.W., H.Z. and H.J.; visualization, Q.X.; supervision, H.J., W.Z. (Weili Zhang) and R.W.; project administration, H.J., W.Z. (Weili Zhang) and R.W.; funding acquisition, H.J. and Q.X. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the support from the Projects of Hunan Provincial Department of Education (21C0435), the Science and Technology Innovation Program of Hunan Province (2021RC4065), and the Key Research and Development Program of Hunan Province (2022GK2037).

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationship that could have appeared to influence the work reported in this paper.

References

  1. Bernardi, M.; Feitosa, C.; Paskocimas, C.; Longo, E.; Paiva-Santos, C. Development of metal oxide nanoparticles by soft chemical method. Ceram. Int. 2009, 35, 463–466. [Google Scholar] [CrossRef]
  2. Akdemir, S.; Ozel, E.; Suvaci, E. Solubility of blue CoAl2O4 ceramic pigments in water and diethylene glycol media. Ceram. Int. 2011, 37, 863–870. [Google Scholar] [CrossRef]
  3. Merino, M.C.G.; Estrella, A.L.; Rodriguez, M.E.; Acuña, L.; Lassa, M.S.; Lascalea, G.E.; Vázquez, P. Combustion Syntheses of CoAl2O4 Powders Using Different Fuels. Procedia Mater. Sci. 2015, 8, 519–525. [Google Scholar] [CrossRef]
  4. Wu, T.; Sun, S.; Song, J.; Xi, S.; Du, Y.; Chen, B.; Sasangka, W.A.; Liao, H.; Gan, C.L.; Scherer, G.G.; et al. Iron-facilitated dynamic active-site generation on spinel CoAl2O4 with self-termination of surface reconstruction for water oxidation. Nat. Catal. 2019, 2, 763–772. [Google Scholar] [CrossRef]
  5. Angeletti, C.; Pepe, F.; Porta, P. Structure and catalytic activity of CoχMg1–χAl2O4 spinel solid solutions. Part 1.—Cation distribution of Co2+ ions. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1977, 73, 1972–1982. [Google Scholar] [CrossRef]
  6. Llusar, M.; Forés, A.; Badenes, J.; Calbo, J.; Tena, M.; Monrós, G. Colour analysis of some cobalt-based blue pigments. J. Eur. Ceram. Soc. 2001, 21, 1121–1130. [Google Scholar] [CrossRef]
  7. Monari, G.; Manfredini, T. Coloring Effects of Synthetic Inorganic Cobalt Pigments in Fast-Fired Porcelainized Tiles. In A Collection of Papers Presented at the 97th Annual Meeting and the 1995 Fall Meetings of the Materials & Equipment/Whitewares: Ceramic Engineering and Science Proceedings; Wachtman, J.B., Jr., Ed.; The American Ceramic Society: Columbus, OH, USA, 1996; Volume 17, pp. 167–172. [Google Scholar] [CrossRef]
  8. Tatarchuk, T.; Bououdina, M.; Paliychuk, N.; Yaremiy, I.; Moklyak, V. Structural characterization and antistructure modeling of cobalt-substituted zinc ferrites. J. Alloys Compd. 2017, 694, 777–791. [Google Scholar] [CrossRef]
  9. Agilandeswari, K.; Kumar, A.R. Synthesis, characterisation, optical and luminescence properties of CoAl2O4. AIP Conf. Proc. 2015, 1665, 120022. [Google Scholar] [CrossRef]
  10. Hudson, K.; Winbow, H.; Cowley, J. Colors for Ceramic Bodies. In A Collection of Papers Presented at the 97th Annual Meeting and the 1995 Fall Meetings of the Materials & Equipment/Whitewares: Ceramic Engineering and Science Proceedings; Wachtman, J.B., Jr., Ed.; The American Ceramic Society: Columbus, OH, USA, 1996; Volume 17, pp. 102–110. [Google Scholar] [CrossRef]
  11. Li, W.; Li, J.; Guo, J. Synthesis and characterization of nanocrystalline CoAl2O4 spinel powder by low temperature combustion. J. Eur. Ceram. Soc. 2003, 23, 2289–2295. [Google Scholar] [CrossRef]
  12. Cho, W.-S.; Kakihana, M. Crystallization of ceramic pigment CoAl2O4 nanocrystals from Co–Al metal organic precursor. J. Alloys Compd. 1999, 287, 87–90. [Google Scholar] [CrossRef]
  13. Pozas, R.; Orera, V.; Ocaña, M. Hydrothermal synthesis of Co-doped willemite powders with controlled particle size and shape. J. Eur. Ceram. Soc. 2005, 25, 3165–3172. [Google Scholar] [CrossRef]
  14. Hu, Z.; Xue, M.; Zhang, Q.; Sheng, Q.; Liu, Y. Nanocolorants: A novel class of colorants, the preparation and performance characterization. Dye. Pigment. 2008, 76, 173–178. [Google Scholar] [CrossRef]
  15. Zhang, W.; Li, J.; Zhong, F.; Wu, G.; Jiang, H.; Zhang, W.; Liu, Q. Preparation and characterization of supercritical fluid–fried (CoAl2O4) cobalt blue nano-pigment. J. Asian Ceram. Soc. 2021, 10, 33–39. [Google Scholar] [CrossRef]
  16. Yuan, S.; Yuanlin, Z.; Yufei, T.; Yangfeng, J. Fabrication and Coloration Mechanism of CoAl2O4 Ceramic Pigment by Co-precipitation Method. J. Synth. Cryst. 2017, 46, 79–84. [Google Scholar]
  17. Zhang, A.; Mu, B.; Luo, Z.; Wang, A. Bright blue halloysite/CoAl2O4 hybrid pigments: Preparation, characterization and application in water-based painting. Dye. Pigment. 2017, 139, 473–481. [Google Scholar] [CrossRef]
  18. Ahmed, I.; Shama, S.; Moustafa, M.; Dessouki, H.; Ali, A. Synthesis and spectral characterization of CoxMg1−xAl2O4 as new nano-coloring agent of ceramic pigment. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2009, 74, 665–672. [Google Scholar] [CrossRef] [PubMed]
  19. Xin, L.; Jiake, L.; Kai, C. Preparation of Spinel Type ZnxCo1-xAl2O4 Ultrafine Ceramic Pigment by Solution Combustion Method. J. Synth. Cryst. 2014, 43, 2086–2090. [Google Scholar]
  20. Gilabert, J.; Palacios; Sanz, V.; Mestre, S. Characteristics reproducibility of (Fe, Co)(Cr, Al)2O4 pigments obtained by solution combustion synthesis. Ceram. Int. 2016, 42, 12880–12887. [Google Scholar] [CrossRef]
  21. Chemlal, S.; Larbot, A.; Persin, M.; Sarrazin, J.; Sghyar, M.; Rafiq, M. Cobalt spinel CoAl2O4 via sol-gel process: Elaboration and surface properties. Mater. Res. Bull. 2000, 35, 2515–2523. [Google Scholar] [CrossRef]
  22. Valenzuela, M.; Bosch, P.; Aguilar-Rios, G.; Montoya, A.; Schifter, I. Comparison Between Sol-Gel, Coprecipitation and Wet Mixing Synthesis of ZnAl2O. J. Sol-Gel Sci. Technol. 1997, 8, 107–110. [Google Scholar] [CrossRef]
  23. Gaudon, M.; Robertson, L.; Lataste, E.; Duttine, M.; Ménétrier, M.; Demourgues, A. Cobalt and nickel aluminate spinels: Blue and cyan pigments. Ceram. Int. 2014, 40, 5201–5207. [Google Scholar] [CrossRef]
  24. Fernández-Osorio, A.; Pineda-Villanueva, E.; Chávez-Fernández, J. Synthesis of nanosized (Zn1−Co )Al2O4 spinels: New pink ceramic pigments. Mater. Res. Bull. 2012, 47, 445–452. [Google Scholar] [CrossRef]
  25. Tsurumi, T.; Hirayama, H.; Vacha, M.; Taniyama, T. Nanoscale Physics for Materials Science; CRC Press: Boca Raton, FL, USA, 2009. [Google Scholar] [CrossRef]
  26. Zawadzki, M.; Wrzyszcz, J. Hydrothermal synthesis of nanoporous zinc aluminate with high surface area. Mater. Res. Bull. 2000, 35, 109–114. [Google Scholar] [CrossRef]
  27. Zhang, A.; Mu, B.; Hui, A.; Wang, A. A facile approach to fabricate bright blue heat-resisting paint with self-cleaning ability based on CoAl2O4/kaolin hybrid pigment. Appl. Clay Sci. 2018, 160, 153–161. [Google Scholar] [CrossRef]
  28. Angeletti, C.; Pepe, F.; Porta, P. Structure and catalytic activity of CoxMg1–xAl2O4 spinel solid solutions. Part 2.—Decomposition of N2O. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1978, 74, 1595–1603. [Google Scholar] [CrossRef]
  29. Zayat, M.; Levy, D. Blue CoAl2O4 Particles Prepared by the Sol−Gel and Citrate−Gel Methods. Chem. Mater. 2000, 12, 2763–2769. [Google Scholar] [CrossRef]
  30. Holzwarth, U.; Gibson, N. The Scherrer equation versus the ‘Debye-Scherrer equation’. Nat. Nanotechnol. 2011, 6, 534. [Google Scholar] [CrossRef] [PubMed]
  31. Mindru, I.; Marinescu, G.; Gingasu, D.; Patron, L.; Ghica, C.; Giurginca, M. Blue CoAl2O4 spinel via complexation method. Mater. Chem. Phys. 2010, 122, 491–497. [Google Scholar] [CrossRef]
  32. Zhang, A.; Wang, N.; Mu, B.; Li, H.; Wang, A. Amino-acid-assisted preparation of CoAl2O4/kaolin hybrid pigments. Appl. Clay Sci. 2020, 191, 105611. [Google Scholar] [CrossRef]
  33. Peng, X.; Cheng, J.; Yuan, J.; Jin, N.; Kang, J.; Hou, Y.; Zhang, Q. Environmental blue CoAl2O4 pigment co-doped by Zn2+ and Mg2+: Synthesis, structure and optical properties. Adv. Appl. Ceram. 2017, 117, 303–311. [Google Scholar] [CrossRef]
  34. Lilova, K.I.; Navrotsky, A.; Melot, B.C.; Seshadri, R. Thermodynamics of CoAl2O4–CoGa2O4 solid solutions. J. Solid State Chem. 2010, 183, 1266–1271. [Google Scholar] [CrossRef]
  35. Wong-Ng, W.; McMurdie, H.; Hubbard, C.; Mighell, A. JCPDS-ICDD Research Associateship (cooperative program with NBS/NIST). J. Res. Natl. Inst. Stand. Technol. 2001, 106, 1013–1028. [Google Scholar] [CrossRef]
  36. Visinescu, D.; Paraschiv, C.; Ianculescu, A.; Jurca, B.; Vasile, B.; Carp, O. The environmentally benign synthesis of nanosized CoxZn1−xAl2O4 blue pigments. Dye. Pigment. 2010, 87, 125–131. [Google Scholar] [CrossRef]
  37. Chen, Z.-Z.; Shi, E.-W.; Li, W.-J.; Zheng, Y.-Q.; Zhuang, J.-Y.; Xiao, B.; Tang, L.-A. Preparation of nanosized cobalt aluminate powders by a hydrothermal method. Mater. Sci. Eng. B 2004, 107, 217–223. [Google Scholar] [CrossRef]
  38. Yu, F.; Yang, J.; Ma, J.; Du, J.; Zhou, Y. Preparation of nanosized CoAl2O4 powders by sol–gel and sol–gel-hydrothermal methods. J. Alloys Compd. 2009, 468, 443–446. [Google Scholar] [CrossRef]
  39. Štangar, U.L.; Orel, B.; Krajnc, M. Preparation and Spectroscopic Characterization of Blue CoAl2O4 Coatings. J. Sol-Gel Sci. Technol. 2003, 26, 771–775. [Google Scholar] [CrossRef]
  40. Luo, M.R.; Cui, G.; Rigg, B. The development of the CIE 2000 colour-difference formula: CIEDE2000. Color Res. Appl. 2001, 26, 340–350. [Google Scholar] [CrossRef]
  41. Gaudon, M.; Apheceixborde, A.; Ménétrier, M.; Le Nestour, A.; Demourgues, A. Synthesis Temperature Effect on the Structural Features and Optical Absorption of Zn1−xCoxAl2O4 Oxides. Inorg. Chem. 2009, 48, 9085–9091. [Google Scholar] [CrossRef]
  42. Vandewater, L.; Bezemer, G.L.; Bergwerff, J.A.; Versluijshelder, M.; Weckhuysen, B.M.; Dejong, K. Spatially resolved UV–vis microspectroscopy on the preparation of alumina-supported Co Fischer–Tropsch catalysts: Linking activity to Co distribution and speciation. J. Catal. 2006, 242, 287–298. [Google Scholar] [CrossRef]
  43. Liotta, L.; Pantaleo, G.; Macaluso, A.; Di Carlo, G.; Deganello, G. CoOx catalysts supported on alumina and alumina-baria: Influence of the support on the cobalt species and their activity in NO reduction by C3H6 in lean conditions. Appl. Catal. A Gen. 2003, 245, 167–177. [Google Scholar] [CrossRef]
  44. Dhas, C.R.; Venkatesh, R.; Jothivenkatachalam, K.; Nithya, A.; Benjamin, B.S.; Raj, A.M.E.; Jeyadheepan, K.; Sanjeeviraja, C. Visible light driven photocatalytic degradation of Rhodamine B and Direct Red using cobalt oxide nanoparticles. Ceram. Int. 2015, 41, 9301–9313. [Google Scholar] [CrossRef]
Figure 1. The crystal structure of AB2O4.
Figure 1. The crystal structure of AB2O4.
Inorganics 11 00368 g001
Figure 2. TG–DSC curves of precursor (air atmosphere).
Figure 2. TG–DSC curves of precursor (air atmosphere).
Inorganics 11 00368 g002
Figure 3. XRD patterns of cobalt blue pigments dried via the supercritical method and heated at 1400 °C, 1200 °C, 800 °C, 500 °C, and 250 °C.
Figure 3. XRD patterns of cobalt blue pigments dried via the supercritical method and heated at 1400 °C, 1200 °C, 800 °C, 500 °C, and 250 °C.
Inorganics 11 00368 g003
Figure 4. EDS spectrum of the CoAl2O4 sample heating at 1400 °C.
Figure 4. EDS spectrum of the CoAl2O4 sample heating at 1400 °C.
Inorganics 11 00368 g004
Figure 5. XRD patterns of samples in different conditions. Sample 1 is fabricated by the ratio of Co and Al equal to 1:2; sample 2 is fabricated by excessive Al.
Figure 5. XRD patterns of samples in different conditions. Sample 1 is fabricated by the ratio of Co and Al equal to 1:2; sample 2 is fabricated by excessive Al.
Inorganics 11 00368 g005
Figure 6. Rietveld analysis results of (a) sample 1 after 1200 °C heat treatment for 1 h, (b) sample 2 after 1200 °C heat treatment for 1 h, (c) sample 1 after 1200 °C heat treatment for 3 h, (d) sample 1 after 1400 °C heat treatment for 1 h, and (e) sample 2 after 1400 °C heat treatment for 1 h, including fitting pattern, residual curve, Bragg reflection position, and R−factor.
Figure 6. Rietveld analysis results of (a) sample 1 after 1200 °C heat treatment for 1 h, (b) sample 2 after 1200 °C heat treatment for 1 h, (c) sample 1 after 1200 °C heat treatment for 3 h, (d) sample 1 after 1400 °C heat treatment for 1 h, and (e) sample 2 after 1400 °C heat treatment for 1 h, including fitting pattern, residual curve, Bragg reflection position, and R−factor.
Inorganics 11 00368 g006
Figure 7. UV–vis absorbance spectra of CoAl2O4 pigment powders after heat treatment at different temperatures and times.
Figure 7. UV–vis absorbance spectra of CoAl2O4 pigment powders after heat treatment at different temperatures and times.
Inorganics 11 00368 g007
Figure 8. Synthesis process of cobalt blue pigments.
Figure 8. Synthesis process of cobalt blue pigments.
Inorganics 11 00368 g008
Table 1. Composition parameters of the CoAl2O4 sample heated at 1400 °C.
Table 1. Composition parameters of the CoAl2O4 sample heated at 1400 °C.
ElementWeight%Atomic%
O K37.4658.05
Al K31.3928.85
Co K31.1513.10
Totals100.00
Table 2. Crystallite size of CoAl2O4 pigment powders after heat treatment at different temperatures and times.
Table 2. Crystallite size of CoAl2O4 pigment powders after heat treatment at different temperatures and times.
SampleTemperature (°C)Time (h)Xs (Å)
112001553
212001596
112003580
114001624
214001721
Table 3. Cation distribution data calculated from the Rietveld refinement of (CoxAl1−x)[Al1+xCo1−x]O4 (0 ≤ x ≤ 1) system after heat treatment at different temperatures and times.
Table 3. Cation distribution data calculated from the Rietveld refinement of (CoxAl1−x)[Al1+xCo1−x]O4 (0 ≤ x ≤ 1) system after heat treatment at different temperatures and times.
SampleTemperature (°C)Time (h)Chemical FormulaOccupation of CationsLattice Parameter
Tetrahedral SiteOctahedral Site
112001Co1.91Al1.09O4 ( Co 0.75 2 + Al 0.25 3 + ) [ Co 0.25 2 + Co 0.91 3 + Al 0.84 3 + ] 8.096
212001Co1.46Al1.54O4 ( Co 0.85 2 + Al 0.15 3 + ) [ Co 0.46 2 + Co 0.15 3 + Al 1.39 3 + ] 8.104
112003Co1.39Al1.61O4 ( Co 0.90 2 + Al 0.10 3 + ) [ Co 0.39 2 + Co 0.1 3 + Al 1.51 3 + ] 8.110
114001Co1.39Al1.61O4 ( Co 0.89 2 + Al 0.11 3 + ) [ Co 0.39 2 + Co 0.11 3 + Al 1.50 3 + ] 8.111
214001CoAl2O4 ( Co 0.94 2 + Al 0.06 3 + ) [ Co 0.06 2 + Al 1.94 3 + ] 8.107
Table 4. Structural parameters CoAl2O4 pigment powders after heat treatment at different temperatures and times.
Table 4. Structural parameters CoAl2O4 pigment powders after heat treatment at different temperatures and times.
SampleTemperature (°C)Time (h)Space Groupa (Å)α (°)V (Å3)ZDensity
112001Fd-3m8.09690533.3984.4053
212001Fd-3m8.10490532.3784.4138
112003Fd-3m8.1190532.6784.4113
114001Fd-3m8.11190533.5884.4037
214001Fd-3m8.10790533.1884.5219
Table 5. CIE-L*a*b* colorimetric analysis of CoAl2O4 pigment powders after heat treatment at different temperatures and times.
Table 5. CIE-L*a*b* colorimetric analysis of CoAl2O4 pigment powders after heat treatment at different temperatures and times.
SampleTemperature (°C)Time (h)L*a*b*Color
1500122.52−0.53−11.24Inorganics 11 00368 i001
1800114.54−2.82−9.44Inorganics 11 00368 i002
11200127.70−4.15−27.17Inorganics 11 00368 i003
21200124.60−3.74−27.83Inorganics 11 00368 i004
11200327.87−3.68−30.80Inorganics 11 00368 i005
11400127.82−1.85−31.18Inorganics 11 00368 i006
21400126.417.31−40.54Inorganics 11 00368 i007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, W.; Li, Z.; Wu, G.; Wu, W.; Zeng, H.; Jiang, H.; Zhang, W.; Wu, R.; Xue, Q. Effects of Coloration of Spinel CoAl2O4 Cobalt Blue Pigments: Composition, Structure, and Cation Distribution. Inorganics 2023, 11, 368. https://doi.org/10.3390/inorganics11090368

AMA Style

Zhang W, Li Z, Wu G, Wu W, Zeng H, Jiang H, Zhang W, Wu R, Xue Q. Effects of Coloration of Spinel CoAl2O4 Cobalt Blue Pigments: Composition, Structure, and Cation Distribution. Inorganics. 2023; 11(9):368. https://doi.org/10.3390/inorganics11090368

Chicago/Turabian Style

Zhang, Weiran, Ziyu Li, Guohua Wu, Wei Wu, Hailan Zeng, Haiyun Jiang, Weili Zhang, Ruomei Wu, and Qiong Xue. 2023. "Effects of Coloration of Spinel CoAl2O4 Cobalt Blue Pigments: Composition, Structure, and Cation Distribution" Inorganics 11, no. 9: 368. https://doi.org/10.3390/inorganics11090368

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop