Next Article in Journal
One-Step Synthesis of Polyethyleneimine-Grafted Styrene-Maleic Anhydride Copolymer Adsorbents for Effective Adsorption of Anionic Dyes
Previous Article in Journal
Design of Na3MnZr(PO4)3/Carbon Nanofiber Free-Standing Cathodes for Sodium-Ion Batteries with Enhanced Electrochemical Performances through Different Electrospinning Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Control of Optical Reflection in Ca2MgWO6 by Co and Mo Doping

1
Department of Chemistry and Biotechnology, Faculty of Engineering, Tottori University, 4-101, Koyama-cho Minami, Tottori 680-8552, Japan
2
Center for Research on Green Sustainable Chemistry, Tottori University, 4-101, Koyama-cho Minami, Tottori 680-8552, Japan
3
Field of Advanced Ceramics, Department of Life Science and Applied Chemistry, Nagoya Institute of Technology, Gokiso, Showa, Nagoya 466-8555, Japan
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(8), 1886; https://doi.org/10.3390/molecules29081886
Submission received: 14 March 2024 / Revised: 16 April 2024 / Accepted: 18 April 2024 / Published: 21 April 2024
(This article belongs to the Section Colorants)

Abstract

:
To develop novel inorganic red pigments without harmful elements, we focused on the band structure of Ca2(Mg, Co)WO6 and attempted to narrow its bandgap by replacing the W6+ sites in the host structure of Mo6+. Ca2Mg1−xCoxW1−yMoyO6 (0.10 ≤ x ≤ 0.30; 0.45 ≤ y ≤ 0.60) samples were synthesized by a sol-gel method using citric acids, and the crystal structure, optical properties, and color of the samples were characterized. The Ca2Mg1−xCoxW1−yMoyO6 solid solution was successfully formed, which absorbed visible light at wavelengths below 600 nm. In addition, the absorption wavelength shifted to longer wavelengths with increasing Mo6+ content. This is because a new conduction band composed of a Co3d-W5d-Mo4d hybrid orbital was formed by Mo6+ doping to reduce the bandgap energy. Thus, the color of the samples gradually changed from pale orange to dark red, with a hue angle (h°) of less than 35°. Based on the above results, the optical absorption wavelength of the Ca2Mg1−xCoxW1−yMoyO6 system can be controlled to change the color by adjusting the bandgap energy.

Graphical Abstract

1. Introduction

Inorganic pigments have been utilized for the coloration of ceramics, plastics, glasses, etc., because of their high thermal stability, light resistance, and hiding power. Red inorganic pigments, which indicate warning colors, have been in large demand for applications such as traffic paints. Minium (Pb3O4·2PbO·PbO2), vermilion (HgS), and cadmium red (CdS·CdSe) were used as the red inorganic pigments. Although they exhibit a vivid red color and excellent durability, they contain highly toxic elements, such as Pb, Hg, Cd, and Se, which have negative effects on the human body and environment. Therefore, they are either regulated or banned worldwide. This regulation affects various inorganic materials and is not limited to pigments. For example, toxic HgCl2 catalysts have been made to be replaced by mercury-free catalysts [1]. In this context, the development of environmentally friendly inorganic pigments containing less or no toxic elements has been strongly desired to replace existing harmful inorganic pigments.
Recently, sulfides and oxynitride-based pigments such as Ce2S3 and Ca1−xLaxTaO2−xN1+x have attracted attention as very brilliant color materials [2,3,4]. However, the chemical stability of sulfide pigments is poor and may cause discoloration when mixed with other pigments. The synthesis of oxynitride pigments requires a prolonged flow of toxic ammonia gas. In addition, sulfur oxide or nitrogen oxide gases are generated during calcination. Thus, oxide pigments are preferred for practical use because they are chemically stable. Although several studies have been reported on oxide pigments [5,6,7,8,9,10,11,12,13], environmentally benign reddish inorganic pigments are not yet comparable to conventionally harmful ones.
In this study, we focus on divalent Co ions (Co2+) to develop a novel red inorganic pigment. Octahedrally coordinated Co2+ ions are expected to exhibit a red or reddish violet color because they absorb green visible light between 490 and 560 nm based on the d–d transition attributed to 4T1(F) → 4T1(P) [14,15,16]. LiCoPO4 and (Zn1−xCox)Al2O4 with Co2+ ions have been studied for the development of reddish pigments [15,16], and cobalt violet (Co3(PO4)2) is well known as commercially available. Although there are many materials containing Co for pigments, alloys, catalysis, etc. [17,18,19,20], the Co content tends to be reduced to avoid price escalation because of its high cost. Against this background, there have been many studies on the development of inorganic pigments with low Co contents [21,22,23,24,25,26].
Here, we adopted double-perovskite-type Ca2MgWO6 as a host material for novel inorganic red pigments. Double-perovskite-type oxides are generally referred to as A2BB′O6, where A is a cation with a large ionic radius, B and B’ are cations smaller than A, and the total valences of A, B, and B’ are +12. Ca2MgWO6 has a space group of P21/n and a monoclinic structure, which consists of 8-coordinate Ca2+ at the A site and 6-coordinate octahedral [MgO6] and [WO6] alternating at the B and B’ sites [27,28]. In addition, Ca2MgWO6 is a promising phosphor candidate because of its excellent fluorescent properties [27,28,29]; however, few studies have been conducted on this topic.
From the above, we synthesized Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples that were partially substituted with Co2+ at the Mg2+ site of the Ca2MgWO6 host structure and evaluated their color. Upon substitution with Co2+, the samples became light orange, and the color of the samples changed to brown as the Co2+ content increased, enhancing the light absorption attributed to the d–d transition of Co2+. Among these samples, Ca2Mg1−xCoxWO6 (0.10 ≤ x ≤ 0.30) can be a reddish pigment by strengthening the absorbance of the green-light region between 490 and 560 nm while maintaining the reflectance of the red-light region between 605 and 750 nm.
For the double-perovskite structure of Ca2NiWO6, the bandgap energy can be reduced by partially replacing the W6+ site with Mo6+ [30]. The band structure of Ca2NiWO6 as shown in Figure 1 has a valence band with an orbital hybridized by the t2g orbital of the Ni3d and O2p orbitals and two conduction bands: the low-energy eg orbital of Ni3d and the high-energy W5d orbital [30,31]. By substituting Mo6+ for the W6+ site in this compound, a conduction band composed of the Ni3d-W5d-Mo4d hybrid orbital was newly formed, and the bandgap energy was reduced [30]. In this study, we attempted to replace the W6+ site in the host structure with Mo6+ to narrow the bandgap and exert a red color by expanding the light absorption wavelength to longer wavelengths. In other words, the Ca2Mg1−xCoxW1−yMoyO6 (0.10 ≤ x ≤ 0.30; 0.45 ≤ y ≤ 0.60) samples were synthesized, and their color was evaluated.

2. Results and Discussion

2.1. Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50)

2.1.1. X-ray Powder Diffraction

Figure 2a shows the X-ray powder diffraction (XRD) patterns of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples. The Ca2MgWO6 phase was obtained for all samples as the main phase, but a few peaks of CaWO4 were observed as an impurity phase.
To investigate the lattice strain of these samples, a Williamson-Hall (W-H) analysis was conducted following the equation βcosθ/λ = 2εsinθ/λ + K/D, where K = 0.94 and β, θ, λ, D, and ε represent the peak width at half-maximum intensity, diffraction position, wavelength of radiation, crystallite size, and strain component, respectively [32,33]. The W-H plot of Ca2MgWO6 is shown in Figure 2b, where the slope and intercept represent the strain component and crystallite size, respectively. The lattice strain of all the samples was also estimated using the W-H plot and is summarized in Figure 2c. The lattice strain of the samples increased with increasing Co content, indicating that Co2+ was partially introduced into the host lattice. Therefore, the probability of the d–d transition of Co2+ should increase with the Co concentration in these systems.
The Ca2MgWO6 double-perovskite structure was illustrated by the VESTA program [34], as shown in Figure 3. It has octahedral Mg2+ sites that can be partially replaced by Co2+ ions. The composition dependence of the lattice volume of the samples, calculated from each XRD pattern, is shown in Figure 4. The lattice volume of the samples increased monotonically with good linearity (determination coefficient: R2 > 0.99) by increasing the Co2+ content in the 0 ≤ x ≤ 0.50 range. This phenomenon means that larger Co2+ (ionic radius: 0.0745 nm, 6-coordination site [35]) ions were partially introduced into the Mg2+ (ionic radius: 0.072 nm, 6-coordination site [35]) sites of the host structure. Thus, Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) solid solutions were successfully synthesized.

2.1.2. Ultraviolet–Visible Reflectance Spectra

The optical properties of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples were evaluated by ultraviolet–visible (UV–vis) reflectance spectroscopy. The UV–vis reflectance spectra of the samples are shown in Figure 5. Ca2MgWO6 (x = 0), as a host compound, strongly reflected all visible light regions. In contrast, the samples with Co2+ absorbed visible light at wavelengths of 350 nm and between 500 and 600 nm. The former light absorption at approximately 350 nm was attributed to the ligand-to-metal charge transfer (LMCT) transition of O2— to Co2+ [36], and the latter between 500 and 600 nm was assigned to the d–d transition (4T1(F) → 4T1(P)) of Co2+ [14,15,16]. The absorption wavelength for the LMCT transition of Co2+ shifted to longer wavelengths with the Co2+ content owing to lattice volume expansion, leading to longer Co–O bond lengths. In addition, the optical absorption based on the d–d transition was enhanced with increasing Co2+ concentration, as shown in Figure 2c. This d–d transition (4T1(F) → 4T1(P)) is a spin-allowed transition, and the transition probability increases with not only the Co2+ content but also the lattice strain.

2.1.3. Color Property

The L*a*b*Ch° color coordinate data for the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) powder samples are listed in Table 1. Photographs of the samples are shown in Figure 6. In the case of x = 0, the sample strongly reflected visible light and exhibited a white color. In contrast, for x ranging from 0.10 to 0.50, light absorption in the green light region around 490–560 nm increased with increasing Co content. Since the brightness (L*) decreased and the redness (a*) as well as the yellowness (b*) increased, the sample color showed a pale orange for x ranging from 0.10 to 0.30 and brown for x ranging over 0.40. Thus, it was found that a reddish pigment could not be realized by doping the host lattice with only Co.

2.2. Ca2Mg1−xCoxW1−yMoyO6 (0.10 ≤ x ≤ 0.30; 0.45 ≤ y ≤ 0.60)

2.2.1. X-ray Powder Diffraction

The Co2+-doped samples, Ca2Mg1−xCoxWO6 (0.10 ≤ x ≤ 0.50), exhibited a pale orange or brown color, and reddish pigments were not obtained. To improve their color, Ca2Mg1−xCoxW1−yMoyO6 (0.10 ≤ x ≤ 0.30; 0.45 ≤ y ≤ 0.60), in which the W6+ site of pale orange Ca2Mg1−xCoxWO6 was partially replaced by Mo6+, was synthesized and characterized. These samples are hereafter referred to as CoaMob, where a and b are the Co and Mo content, respectively. (For example, when Co content is 10% and Mo content is 50%, the sample is described as Co10Mo50.)
The XRD patterns of the CoaMob powder samples are shown in Figure 7a–e. For comparison, the XRD patterns of the CoaMo0 samples are also shown. Although the target Ca2MgWO6 phase was obtained as the main phase for all samples, a few peaks of CaMoO4 were detected as impurities, resulting in a mixed phase. The lattice strain of all the samples was estimated using a W-H plot, as summarized in Figure 7f. The lattice strain of the samples decreased with increasing Mo content, indicating that Mo6+ was partially introduced into the host lattice.
The dependence of the lattice volume on the composition of all the samples calculated from each XRD pattern is shown in Figure 8. The introduction of smaller Mo6+ ions (ionic radius: 0.0590 nm at the 6-coordination site [35]) into the W6+ sites (ionic radius: 0.060 nm at the 6-coordination site [35]) should result in a decrease in lattice volume with increasing Mo6+ content. However, the lattice volume of the CoaMob samples increased linearly. This is because W6+, which has a higher electronegativity (6-coordination: 2.175 [37]) in the Ca2(Mg, Co)WO6 structure, was partially replaced by Mo6+ with less electronegativity (6-coordination: 2.101 [37]), extending the bond length for W/Mo–O to expand the lattice volume. Thus, the synthesized samples successfully formed solid solutions.

2.2.2. UV–Vis Reflectance Spectra

The UV–Vis reflectance spectra of CoaMob are shown in Figure 9. The results for the CoaMo0 (10 ≤ a ≤ 30) samples are shown for comparison. The Mo-doped samples absorbed intensely visible light at wavelengths below 600 nm, and the absorption wavelength shifted to longer wavelengths with increasing Mo concentrations. This is due to the formation of a new conduction band of the Co3d-W5d-Mo4d hybrid orbital by Mo6+ doping [30], which results in a reduction in the bandgap energy, as shown in Figure 10.
The bandgap energy (Eg) of these samples was investigated using a Tauc plot [38], which was obtained by converting the corresponding UV–vis reflectance spectra, as shown in Figure 11. The bandgap energy was determined from the Tauc plot by extrapolating the linear area across the axis of the graph. The intersection with the axis is an estimation of the corresponding Eg. The estimated Eg values are listed in Table 2. The Eg values of the Co- and Mo-doped samples were smaller than those of the Co-doped samples, which means that the Co3d-W5d-Mo4d hybrid orbital as a conduction band was newly constructed by Mo6+ doping, as shown in Figure 10. As expected, the optical bandgap energy decreased with increasing Mo content when the Co content was fixed. Therefore, the bandgap energy of the sample could be finely controlled by co-doping with various concentrations of Co and Mo. In addition, the reduction in bandgap energy leads to a higher electrical conductivity of the materials. The samples synthesized using this bandgap-controlling strategy have potential applications in other fields, such as catalysts for water splitting [39].

2.2.3. Color Property

The color properties of the CoaMob samples were evaluated using a colorimeter. The L*a*b*Ch° color coordinates and photographs are presented in Table 3 and Figure 12, respectively. Brightness (L*), redness (a*), and yellowness (b*) decreased with increasing Mo6+ content. This is because the samples absorbed not only the green light region (490–560 nm) of the complementary color against the red but also the yellow to red light regions (580–750 nm), as shown in Figure 9. However, the hue angle (h°) decreased with increasing Co and Mo content. When the Co concentration was greater than 15%, ranged from 0 to 35, indicating a red color. In fact, the sample color changed from reddish-brown to dark red upon Mo doping. In addition, the Co15Mo50 samples exhibited the highest a* value among the samples with a red hue angle (0 ≤ h° ≤ 35) to be the most nearly red color. Consequently, the color of the sample could be gradually controlled by Co and Mo doping.
The chromatic parameters of the CoaMob sample were compared with those of commercially available red pigments such as Bengal red (Fe2O3), vermillion (HgS), and cadmium red (CdS·CdSe), as listed in Table 4. The photographs are shown in Figure 13. Although the values of a* and b* for Co15Mo50 were smaller than those for commercial red pigments, the pigment synthesized in this study showed the lowest hue angle ( = 33.9), which means that its color is close to the purest red color among these red pigments. However, further improvements are necessary to make them comparable to the red color of conventionally harmful pigments.

2.2.4. Chemical Stability Test

If pigments are practically used for various applications, such as tableware, their acid-base resistance is an important property. The chemical stability of the Co15Mo50 powder samples was evaluated. The powder samples were soaked for 7 h at room temperature in 4% CH3COOH and 4% NH4HCO3 aqueous solutions, assuming vinegar and baking soda, which are possibly the acids and bases most likely to come into contact with tableware. The samples were washed with deionized water and ethanol and dried at ambient temperature for 24 h. Table 5 summarizes the chromatic parameters of the samples after acid and base resistance tests, and the corresponding photographs of the samples are shown in Figure 14. Unfortunately, the Co15Mo50 pigment was less chemically stable because the color tone changed after the leaching tests in the acid and base solutions. To suppress color degradation, it is necessary to protect the surface with inert substances such as silica.

3. Materials and Methods

3.1. Synthesis

Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples were synthesized using a citrate sol-gel method. The starting materials were Ca(NO3)2·4H2O (FUJIFILM Wako Pure Chemical Industries Ltd., Osaka, Japan, 98.5%), Mg(NO3)2·6H2O (FUJIFILM Wako Pure Chemical Industries Ltd., 99.5%), Co(NO3)2·6H2O (FUJIFILM Wako Pure Chemical Industries Ltd., 99.5%), and WO3 (Kishida Chemical Co. Ltd., Osaka, Japan, 98.0%). The materials were weighed stoichiometrically to obtain the desired compositions, as listed in Table 6. WO3 was dissolved in 30 cm3 of five-fold diluted aqueous ammonia (FUJIFILM Wako Pure Chemical Industries Ltd., 28.0 wt. %), and the metal nitrates were dissolved in 50 cm3 of deionized water. These solutions were mixed and stirred uniformly, and citric acid (CA; FUJIFILM Wako Pure Chemical Industries Ltd., 98.0%) was added as a chelator to complex the cations in the solution. The molar ratio of the total cations (Ca, Mg, Co, and W) to CA was 2:1. The mixed solution was stirred by heating at 80 °C until a gel was obtained, which was then oven-dried at 120 °C for 24 h. The dried gel was pulverized using an agate mortar and calcined in an alumina crucible at 500 °C for 8 h in air. After calcination, the sample was again heated in an alumina boat at 1250 °C for 5 h in air. The samples were ground in an agate mortar and pestle before characterization.
CoaMob (10 ≤ a ≤ 30; 45 ≤ b≤ 60), namely Ca2Mg1−xCoxW1−yMoyO6 (0.10 ≤ x ≤ 0.30; 0.45 ≤ y ≤ 0.60), samples were also prepared using a procedure similar to that described above. The starting materials, Ca(NO3)2·4H2O, Mg(NO3)2·6H2O, Co(NO3)2·6H2O, MoO3 (FUJIFILM Wako Pure Chemical Industries Ltd., 99.9%), and WO3, were stoichiometrically weighed, as shown in Table 7. MoO3 and WO3 were dissolved in 30 cm3 of five-fold diluted aqueous ammonia.

3.2. Characterization

Powder X-ray diffraction (XRD) analysis was performed using an Ultima IV X-ray diffractometer (Rigaku Corporation, Tokyo, Japan) to identify the crystal phases and structures of the samples. XRD patterns were obtained using Cu-Kα radiation, which was operated at a tube voltage of 40 kV and a tube current of 40 mA. The data were collected by scanning over a 2θ range of 20°–80°. The sampling width was 0.02°, and the scan speed was 6° min−1. The lattice volumes were calculated from the XRD peak angles refined by α-Al2O3 as a standard and using CellCalc Ver. 2.20 software. An ultraviolet–visible (UV–Vis) spectrometer (JASCO Corporation, Tokyo, Japan, V-770 with an integrating sphere attachment) was used to record the optical reflectance spectra of the as-prepared samples using a standard white plate as a reference. The step width was 1 nm, and the scan rate was 1000 nm min−1. The bandgap energies of the samples were calculated from the absorption edge of the absorbance spectrum, represented by the Kubelka-Munk function f(R) = (1 − R)2/2R, where f(R) is the theoretical absorbance and R is the measured reflectance [36]. The chromatic parameters of the powder samples were evaluated based on the Commission Internationale de l’Éclairage (CIE) L*a*b*Ch° system using a colorimeter (Konica-Minolta, Inc., Tokyo, Japan, CR-400). A standard C illuminant was used for colorimetric measurements. The L* parameter shows the brightness or darkness in neutral grayscale. Positive and negative a* values represent reddish and greenish colors, respectively. Positive and negative b* values indicate yellowish and bluish colors, respectively. The chroma parameter (C) is the color saturation, which is expressed by the formula C = [(a*)2 + (b*)2]1/2. The hue angle (h°) ranged from 0° to 360° and was calculated using the equation h° = arctan(b*/a*). For the L*a*b*Ch° color coordinate data, all values showed standard deviations of less than 0.1.

4. Conclusions

To develop novel inorganic red pigments with fewer harmful elements, we focused on double-perovskite-type Ca2MgWO6 and Co2+ ions as the host material and chromophores, respectively. Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples were synthesized using the citrate sol-gel method. The sample color turned from white to brown with an increase in Co2+ content due to enhanced light absorption based on the d–d transition, but a reddish color was not obtained. Thus, we attempted to lower the bandgap energy of Ca2Mg1−xCoxWO6 to achieve a red color, and the CoaMob (10 ≤ a ≤ 30; 45 ≤ b ≤ 60) samples were prepared and characterized. When Mo6+ was introduced into the W6+ site of Ca2Mg1−xCoxWO6, a new conduction band corresponding to the Co3d-W5d-Mo4d hybridized orbital was formed, and their bandgap energies decreased, as expected. The bandgap energy decreased with increasing Mo6+ concentrations because of the widening conduction band. As a result, the samples absorbed longer wavelengths of light and exhibited a reddish-brown or dark-red color. Among the samples whose hue angle ranged in red (0 ≤ h° ≤ 35), Co15Mo50 showed the highest a* value of 22.3. Although the vividness of the sample synthesized in this study was less than that of commercially available inorganic red pigments, its chromatic purity was the highest among them. It is noteworthy that the bandgap energy can be controlled by introducing the Mo4d orbital between the Co3d and W4d orbitals to form a new wide conduction band; thus, the color of the sample is also controllable.

Author Contributions

The following are the author contributions to this study: conceptualization, R.O. and T.M.; methodology, K.Y., K.M., R.O. and T.M.; validation, K.Y., R.O. and T.M.; investigation, K.Y., K.M. and R.O.; data curation, K.Y., K.M., R.O. and T.M.; writing—original draft preparation, K.Y.; writing—review and editing, R.O. and T.M.; supervision, T.M.; funding acquisition, T.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS KAKENHI, grant numbers JP22K04698 and JP20H02439.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, B.; Jin, C.; Shao, S.; Yue, Y.; Zhang, Y.; Wang, S.; Chang, R.; Zhang, H.; Zhao, J.; Li, X. Electron-Deficient Cu Site Catalyzed Acetylene Hydrochlorination. Green. Energy Environ. 2023, 8, 1128–1140. [Google Scholar] [CrossRef]
  2. Zhang, S.; Ye, M.; Chen, S.; Han, A.; Zang, Y. Synthesis and Characterization of Mica/γ-Ce2−xYxS3 Composite Red Pigments with UV Absorption and High NIR Reflectance. Ceram. Int. 2016, 42, 16023–16030. [Google Scholar] [CrossRef]
  3. Jansen, M.; Letschert, H.P. Inorganic Yellow-Red Pigments without Toxic Metals. Nature 2000, 404, 980–982. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, S.-G.; Li, Y.-M.; Wang, Z.-M.; Shen, Z.-Y.; Xie, Z.-X. Enhanced High Temperature Oxidization Resistance of Silica Coated γ-Ce2S3 Red Pigments. Appl. Surf. Sci. 2016, 387, 1147–1153. [Google Scholar] [CrossRef]
  5. Shiraishi, A.; Takeuchi, N.; Masui, T.; Imanaka, N. Novel Environment Friendly Inorganic Red Pigments Based on Bi4V2O11. RSC Adv. 2015, 5, 44886–44894. [Google Scholar] [CrossRef]
  6. Yoshida, T.; Masui, T.; Imanaka, N. Novel Environmentally Friendly Inorganic Red Pigments Based on Calcium Bismuth Oxides. J. Adv. Ceram. 2015, 4, 39–45. [Google Scholar] [CrossRef]
  7. Masui, T.; Imanaka, N. Novel Environment-Friendly Inorganic Red Pigments Based on (Bi, Er, Y, Fe)2O3 Solid Solutions. J. Asian Ceram. Soc. 2014, 2, 195–198. [Google Scholar] [CrossRef]
  8. Shirpour, M.; Sani, M.A.F.; Mirhabibi, A. Synthesis and Study of a New Class of Red Pigments Based on Perovskite YAlO3 Structure. Ceram. Int. 2007, 33, 1427–1433. [Google Scholar] [CrossRef]
  9. Kim, S.W.; Hasegawa, T.; Watanabe, M.; Sugimoto, K.; Saito, Y.; Uematsu, K.; Toda, K.; Sato, M. Environmentally Friendly Rb3V5O14 Fluorescent Red Pigment. Dye. Pigm. 2017, 136, 219–223. [Google Scholar] [CrossRef]
  10. Jiang, P.; Li, J.; Sleight, A.W.; Subramanian, M.A. New Oxides Showing an Intense Orange Color Based on Fe3+ in Trigonal-Bipyramidal Coordination. Inorg. Chem. 2011, 50, 5858–5860. [Google Scholar] [CrossRef]
  11. Kumari, L.S.; Rao, P.P.; Reddy, M.L. Environment-Friendly Red Pigments from CeO2–Fe2O3–Pr6O11 Solid Solutions. J. Alloys Compd. 2008, 461, 509–515. [Google Scholar] [CrossRef]
  12. Kumari, L.S.; Rao, P.P.; Sameera, S.; Koshy, P. Synthesis and Optical Properties of Ce0.95Pr0.05−xMxO2 (M = Mn, Si) as Potential Ecological Red Pigments for Coloration of Plastics. Ceram. Int. 2012, 38, 4009–4016. [Google Scholar] [CrossRef]
  13. Jovaní, M.; Sanz, A.; Beltrán-Mir, H.; Cordoncillo, E. New Red-Shade Environmental-Friendly Multifunctional Pigment Based on Tb and Fe Doped Y2Zr2O7 for Ceramic Applications and Cool Roof Coatings. Dye. Pigm. 2016, 133, 33–40. [Google Scholar] [CrossRef]
  14. Meseguer, S.; Tena, M.A.; Gargori, C.; Badenes, J.A.; Llusar, M.; Monrós, G. Structure and Colour of Cobalt Ceramic Pigments from Phosphates. Ceram. Int. 2007, 33, 843–849. [Google Scholar] [CrossRef]
  15. Serment, B.; Corucho, L.; Demourgues, A.; Hadziioannou, G.; Brochon, C.; Cloutet, E.; Gaudon, M. Tailoring the Chemical Composition of LiMPO4 (M = Mg, Co, Ni) Orthophosphates To Design New Inorganic Pigments from Magenta to Yellow Hue. Inorg. Chem. 2019, 58, 7499–7510. [Google Scholar] [CrossRef] [PubMed]
  16. Fernández-Osorio, A.; Pineda-Villanueva, E.; Chávez-Fernández, J. Synthesis of Nanosized (Zn1−xCox)Al2O4 Spinels: New Pink Ceramic Pigments. Mater. Res. Bull. 2012, 47, 445–452. [Google Scholar] [CrossRef]
  17. Miccichè, F.; Oostveen, E.; Haveren, J.V.; Linde, R.V.D. The Combination of Reducing Agents/Iron as Environmentally Friendlier Alternatives for Co-Based Driers in the Drying of Alkyd Paints. Prog. Org. Coat. 2005, 53, 99–105. [Google Scholar] [CrossRef]
  18. Uner, D.; Demirkol, M.K.; Dernaika, B. A Novel Catalyst for Diesel Soot Oxidation. Appl. Catal. B-Environ. 2005, 61, 334–345. [Google Scholar] [CrossRef]
  19. Frías, M.; Sánchez De Rojas, M.I. Total and Soluble Chromium, Nickel and Cobalt Content in the Main Materials Used in the Manufacturing of Spanish Commercial Cements. Cem. Concr. Res. 2002, 32, 435–440. [Google Scholar] [CrossRef]
  20. Russell, N.V.; Wigley, F.; Williamson, J. Microstructural Changes to Metal Bond Coatings on Gas Turbine Alloys with Time at High Temperature. J. Mater. Sci. 2000, 35, 2131–2138. [Google Scholar] [CrossRef]
  21. Llusar, M.; Vitásková, L.; Šulcová, P.; Tena, M.A.; Badenes, J.A.; Monrós, G. Red Ceramic Pigments of Terbium-Doped Ceria Prepared through Classical and Non-Conventional Coprecipitation Routes. J. Eur. Ceram. Soc. 2010, 30, 37–52. [Google Scholar] [CrossRef]
  22. Llusar, M.; Forés, A.; Badenes, J.A.; Calbo, J.; Tena, M.A.; Monrós, G. Colour Analysis of Some Cobalt-Based Blue Pigments. J. Eur. Ceram. Soc. 2001, 21, 1121–1130. [Google Scholar] [CrossRef]
  23. Ahmed, I.S.; Shama, S.A.; Moustafa, M.M.; Dessouki, H.A.; Ali, A.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]
  24. Robertson, L.; Duttine, M.; Gaudon, M.; Demourgues, A. Cobalt–Zinc Molybdates as New Blue Pigments Involving Co2+ in Distorted Trigonal Bipyramids and Octahedra. Chem. Mater. 2011, 23, 2419–2427. [Google Scholar] [CrossRef]
  25. Gorodylova, N.; Kosinová, V.; Dohnalová, Ž.; Bělina, P.; Šulcová, P. New Purple-Blue Ceramic Pigments Based on CoZr4(PO4)6. Dye. Pigm. 2013, 98, 393–404. [Google Scholar] [CrossRef]
  26. Erič, S.; Kostić-Gvozdenović, L.; Miladinović, M.; Pavlović, L. Synthesis of cobalt-based ceramic pigments from industrial waste materials. Ceramics-Silikáty 1990, 34, 61–67. [Google Scholar]
  27. Cao, R.; Quan, G.; Shi, Z.; Luo, Z.; Hu, Q.; Guo, S. A Double Perovskite Ca2MgWO6:Bi3+ Yellow-Emitting Phosphor: Synthesis and Luminescence Properties. J. Lumin. 2017, 181, 332–336. [Google Scholar] [CrossRef]
  28. Xu, D.; Wu, X.; Zhang, Q.; Li, W.; Wang, T.; Cao, L.; Meng, J. Fluorescence Property of Novel Near-Infrared Phosphor Ca2MgWO6:Cr3+. J. Alloys Compd. 2018, 731, 156–161. [Google Scholar] [CrossRef]
  29. Jiang, Y.; Tong, Y.; Chen, S.; Zhang, W.; Hu, F.; Wei, R.; Guo, H. A Three-Mode Self-Referenced Optical Thermometry Based on up-Conversion Luminescence of Ca2MgWO6:Er3+, Yb3+ Phosphors. Chem. Eng. J. 2021, 413, 127470. [Google Scholar] [CrossRef]
  30. Luo, Y.; Xue, J.; Zhu, X.; Daniel, J.; Gao, X.; Sun, S.; Gao, C.; Bao, J. Enhanced Photocatalytic Oxygen Evolution over Mo-Doped Ca2NiWO6 Perovskite Photocatalyst under Visible Light Irradiation. RSC Adv. 2017, 7, 5821–5826. [Google Scholar] [CrossRef]
  31. Li, D.; Zheng, J.; Zou, Z. Band Structure and Photocatalytic Properties of Perovskite-Type Compound Ca2NiWO6 for Water Splitting. J. Phys. Chem. Solids 2006, 67, 801–806. [Google Scholar] [CrossRef]
  32. Williamson, G.K.; Hall, W.H. X-Ray Line Broadening from Filed Aluminium and Wolfram. Acta Metall. 1953, 1, 22–31. [Google Scholar] [CrossRef]
  33. Guan, D.; Zhong, J.; Xu, H.; Huang, Y.-C.; Hu, Z.; Chen, B.; Zhang, Y.; Ni, M.; Xu, X.; Zhou, W.; et al. A Universal Chemical-Induced Tensile Strain Tuning Strategy to Boost Oxygen-Evolving Electrocatalysis on Perovskite Oxides. Appl. Phys. Rev. 2022, 9, 011422. [Google Scholar] [CrossRef]
  34. Momma, K.; Izumi, F. VESTA 3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  35. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  36. Barreca, D.; Massignan, C.; Daolio, S.; Fabrizio, M.; Piccirillo, C.; Armelao, L.; Tondello, E. Composition and Microstructure of Cobalt Oxide Thin Films Obtained from a Novel Cobalt(II) Precursor by Chemical Vapor Deposition. Chem. Mater. 2001, 13, 588–593. [Google Scholar] [CrossRef]
  37. Li, K.; Xue, D. Estimation of Electronegativity Values of Elements in Different Valence States. J. Phys. Chem. A 2006, 110, 11332–11337. [Google Scholar] [CrossRef] [PubMed]
  38. Tauc, J.; Grigorovici, R.; Vancu, A. Optical Properties and Electronic Structure of Amorphous Germanium. Phys. Stat. Sol. 1966, 15, 627–637. [Google Scholar] [CrossRef]
  39. Guan, D.; Wang, B.; Zhang, J.; Shi, R.; Jiao, K.; Li, L.; Wang, Y.; Xie, B.; Zhang, Q.; Yu, J.; et al. Hydrogen Society: From Present to Future. Energy Environ. Sci. 2023, 16, 4926–4943. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of band structure for (a) Ca2NiWO6 and (b) Ca2Ni(W, Mo)O6. The black arrows indicate the band gap energy of the host compound, whereas the red arrows indicate the band gap energy of the Mo-doped compound.
Figure 1. Schematic illustration of band structure for (a) Ca2NiWO6 and (b) Ca2Ni(W, Mo)O6. The black arrows indicate the band gap energy of the host compound, whereas the red arrows indicate the band gap energy of the Mo-doped compound.
Molecules 29 01886 g001
Figure 2. (a) XRD pattern of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples; (b) Williamson-Hall (W-H) plot of Ca2MgWO6; and (c) composition dependence of lattice strain (ε) calculated by W-H analysis.
Figure 2. (a) XRD pattern of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples; (b) Williamson-Hall (W-H) plot of Ca2MgWO6; and (c) composition dependence of lattice strain (ε) calculated by W-H analysis.
Molecules 29 01886 g002
Figure 3. Crystal structure of double-perovskite Ca2MgWO6; Ca: blue, Mg: orange, W: gray, O: red.
Figure 3. Crystal structure of double-perovskite Ca2MgWO6; Ca: blue, Mg: orange, W: gray, O: red.
Molecules 29 01886 g003
Figure 4. Composition dependence of the lattice volume of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples.
Figure 4. Composition dependence of the lattice volume of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples.
Molecules 29 01886 g004
Figure 5. UV–vis reflectance spectra of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples.
Figure 5. UV–vis reflectance spectra of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples.
Molecules 29 01886 g005
Figure 6. Photographs of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) powder samples.
Figure 6. Photographs of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) powder samples.
Molecules 29 01886 g006
Figure 7. XRD patterns of the CoaMob samples: (a) a = 10; (b) a = 15; (c) a = 20; (d) a = 25; and (e) a = 30. (f) Composition dependence of lattice strain (ε) for CoaMob calculated by WH analysis.
Figure 7. XRD patterns of the CoaMob samples: (a) a = 10; (b) a = 15; (c) a = 20; (d) a = 25; and (e) a = 30. (f) Composition dependence of lattice strain (ε) for CoaMob calculated by WH analysis.
Molecules 29 01886 g007
Figure 8. Composition dependence of the lattice volume of the CoaMob samples: (a) a = 10; (b) a = 15; (c) a = 20; (d) a = 25; and (e) a = 30.
Figure 8. Composition dependence of the lattice volume of the CoaMob samples: (a) a = 10; (b) a = 15; (c) a = 20; (d) a = 25; and (e) a = 30.
Molecules 29 01886 g008
Figure 9. UV–vis reflectance spectra of the CoaMob samples: (a) a = 10; (b) a = 15; (c) a = 20; (d) a = 25; and (e) a = 30.
Figure 9. UV–vis reflectance spectra of the CoaMob samples: (a) a = 10; (b) a = 15; (c) a = 20; (d) a = 25; and (e) a = 30.
Molecules 29 01886 g009
Figure 10. Schematic illustration of band structure for (a) Ca2(Mg, Co)WO6 and (b) Ca2(Mg, Co(W, Mo)O6. The blue arrows indicate the band gap energy of the host compound, whereas the red arrows indicate the band gap energy of the Mo-doped compound.
Figure 10. Schematic illustration of band structure for (a) Ca2(Mg, Co)WO6 and (b) Ca2(Mg, Co(W, Mo)O6. The blue arrows indicate the band gap energy of the host compound, whereas the red arrows indicate the band gap energy of the Mo-doped compound.
Molecules 29 01886 g010
Figure 11. Tauc plot of the CoaMob samples: (a) a = 10; (b) a = 15; (c) a = 20; (d) a = 25; and (e) a = 30.
Figure 11. Tauc plot of the CoaMob samples: (a) a = 10; (b) a = 15; (c) a = 20; (d) a = 25; and (e) a = 30.
Molecules 29 01886 g011
Figure 12. Photographs of the CoaMob samples.
Figure 12. Photographs of the CoaMob samples.
Molecules 29 01886 g012
Figure 13. Photographs of Ca2Mg0.85Co0.15W0.50Mo0.50O6, Bengal red, vermillion, and cadmium red pellets made from powder samples.
Figure 13. Photographs of Ca2Mg0.85Co0.15W0.50Mo0.50O6, Bengal red, vermillion, and cadmium red pellets made from powder samples.
Molecules 29 01886 g013
Figure 14. Photographs of Ca2Mg0.85Co0.15W0.50Mo0.50O6 samples before and after the chemical stability test.
Figure 14. Photographs of Ca2Mg0.85Co0.15W0.50Mo0.50O6 samples before and after the chemical stability test.
Molecules 29 01886 g014
Table 1. Color coordinates of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples.
Table 1. Color coordinates of the Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50) samples.
xL*a*b*Ch°
096.4+0.70−0.080.70353
0.1086.5+5.05+12.313.367.7
0.1583.1+6.12+14.916.167.7
0.2078.8+8.24+18.220.065.6
0.2575.0+9.57+22.324.366.8
0.3071.0+12.3+25.628.464.3
0.4067.0+11.4+25.027.765.5
0.5062.7+12.3+23.026.061.9
Table 2. Bandgap energy (Eg) for the CoaMob samples estimated from the UV–vis reflectance spectra.
Table 2. Bandgap energy (Eg) for the CoaMob samples estimated from the UV–vis reflectance spectra.
SamplesEg/eV
Co10Mo01.99
Co10Mo451.86
Co10Mo501.84
Co10Mo551.83
Co10Mo601.82
Co15Mo01.99
Co15Mo451.81
Co15Mo501.80
Co15Mo551.79
Co15Mo601.78
Co20Mo01.97
Co20Mo451.80
Co20Mo501.78
Co20Mo551.77
Co20Mo601.75
Co25Mo01.97
Co25Mo451.79
Co25Mo501.78
Co25Mo551.75
Co25Mo601.74
Co30Mo01.97
Co30Mo451.77
Co30Mo501.75
Co30Mo551.74
Co30Mo601.73
Table 3. Color coordinates for the CoaMob samples.
Table 3. Color coordinates for the CoaMob samples.
SampleL*a*b*Ch°
Co10Mo086.5+5.05+12.313.367.7
Co10Mo4540.4+26.3+31.340.950.0
Co10Mo5036.8+24.8+23.634.243.6
Co10Mo5536.0+25.8+24.035.242.9
Co10Mo6037.2+24.5+24.234.444.6
Co15Mo083.1+6.12+14.916.167.7
Co15Mo4535.0+24.8+20.732.339.9
Co15Mo5031.2+22.3+15.026.933.9
Co15Mo5531.1+22.1+14.626.533.5
Co15Mo6029.8+21.6+13.425.431.8
Co20Mo078.8+8.24+18.220.065.6
Co20Mo4531.0+22.9+17.128.636.7
Co20Mo5030.0+20.2+11.723.330.1
Co20Mo5527.9+19.2+10.621.928.9
Co20Mo6028.2+19.0+10.321.628.5
Co25Mo075.0+9.57+22.324.366.8
Co25Mo4529.7+21.9+13.725.832.0
Co25Mo5026.3+20.2+11.823.430.3
Co25Mo5525.4+15.9+7.5317.625.3
Co25Mo6025.7+16.3+7.6418.025.1
Co30Mo071.0+12.3+25.628.464.3
Co30Mo4527.0+16.0+6.8617.423.2
Co30Mo5025.4+14.4+5.4615.420.8
Co30Mo5525.5+14.3+5.5415.321.2
Co30Mo6025.4+10.8+3.2411.316.7
Table 4. Color coordinates of various red pigments.
Table 4. Color coordinates of various red pigments.
SamplesL*a*b*h°
Co15Mo5031.2+22.3+15.033.9
Bengal red (Fe2O3)36.7+33.1+25.037.0
Vermillion (HgS)53.5+55.6+42.937.7
Cadmium red (CdS·CdSe)54.0+61.8+55.341.8
Table 5. Color coordinate data of the Co15Mo50 samples before and after the chemical stability test.
Table 5. Color coordinate data of the Co15Mo50 samples before and after the chemical stability test.
TreatmentL*a*b*Ch°
As synthesized31.2+22.3+15.026.933.9
4% CH3COOH39.8+22.0+21.630.844.5
4% NH4HCO340.2+20.2+22.028.444.7
Table 6. Amounts of reagents used to synthesize Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50).
Table 6. Amounts of reagents used to synthesize Ca2Mg1−xCoxWO6 (0 ≤ x ≤ 0.50).
SampleCa(NO3)2·4H2OMg(NO3)2·6H2OCo(NO3)2·6H2OWO3CA
Ca2MgWO61.2290 g0.6672 g-0.6033 g3.9994 g
Co101.2180 g0.5951 g0.0751 g0.5979 g3.9637 g
Co151.2126 g0.5296 g0.1121 g0.5952 g3.9460 g
Co201.2072 g0.5243 g0.1488 g0.5926 g3.9286 g
Co251.2019 g0.4894 g0.1852 g0.5900 g3.9113 g
Co301.1966 g0.4548 g0.2212 g0.5874 g3.8941 g
Co401.1862 g0.3864 g0.2924 g0.5823 g3.8602 g
Co501.1760 g0.3192 g0.3623 g0.5773 g3.8269 g
Table 7. Amounts of reagents used to synthesize CoaMob (10 ≤ a ≤ 30; 45 ≤ b≤ 60).
Table 7. Amounts of reagents used to synthesize CoaMob (10 ≤ a ≤ 30; 45 ≤ b≤ 60).
SampleCa(NO3)2·4H2OMg(NO3)2·6H2OCo(NO3)2·6H2OWO3MoO3CA
Co10Mo451.3564 g0.6627 g0.0836 g0.3662 g0.1860 g4.4139 g
Co10Mo501.3737 g0.6712 g0.0846 g0.3372 g0.2093 g4.4703 g
Co10Mo551.3915 g0.6799 g0.0857 g0.3074 g0.2333 g4.5281 g
Co10Mo601.4097 g0.6888 g0.0869 g0.2768 g0.2578 g4.5875 g
Co15Mo451.3496 g0.6228 g0.1247 g0.3644 g0.1851 g4.3920 g
Co15Mo501.3668 g0.6338 g0.1263 g0.3355 g0.2083 g4.4479 g
Co15Mo551.3844 g0.6388 g0.1280 g0.3058 g0.2321 g4.5051 g
Co15Mo601.4025 g0.6472 g0.1296 g0.2754 g0.2565 g4.5640 g
Co20Mo451.3430 g0.5833 g0.1655 g0.3626 g0.1842 g4.3704 g
Co20Mo501.3600 g0.5907 g0.1676 g0.3338 g0.2073 g4.4257 g
Co20Mo551.3774 g0.5982 g0.1698 g0.3043 g0.2309 g4.4824 g
Co20Mo601.3953 g0.6060 g0.1720 g0.2740 g0.2552 g4.5406 g
Co25Mo451.3364 g0.5442 g0.2059 g0.3607 g0.1833 g4.3490 g
Co25Mo501.3532 g0.5510 g0.2085 g0.3321 g0.2062 g4.4037 g
Co25Mo551.3705 g0.5580 g0.2111 g0.3027 g0.2297 g4.4599 g
Co25Mo601.3882 g0.5652 g0.2139 g0.2726 g0.2539 g4.5175 g
Co30Mo451.3299 g0.5054 g0.2458 g0.3590 g0.1824 g4.3278 g
Co30Mo501.3466 g0.5117 g0.2489 g0.3305 g0.2052 g4.3820 g
Co30Mo551.3637 g0.5182 g0.2521 g0.3012 g0.2286 g4.4376 g
Co30Mo601.3812 g0.5249 g0.2553 g0.2712 g0.2526 g4.4946 g
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

Yamaguchi, K.; Minagawa, K.; Oka, R.; Masui, T. Control of Optical Reflection in Ca2MgWO6 by Co and Mo Doping. Molecules 2024, 29, 1886. https://doi.org/10.3390/molecules29081886

AMA Style

Yamaguchi K, Minagawa K, Oka R, Masui T. Control of Optical Reflection in Ca2MgWO6 by Co and Mo Doping. Molecules. 2024; 29(8):1886. https://doi.org/10.3390/molecules29081886

Chicago/Turabian Style

Yamaguchi, Kazuki, Kohei Minagawa, Ryohei Oka, and Toshiyuki Masui. 2024. "Control of Optical Reflection in Ca2MgWO6 by Co and Mo Doping" Molecules 29, no. 8: 1886. https://doi.org/10.3390/molecules29081886

Article Metrics

Back to TopTop