Next Article in Journal
Engineering the Mechanics and Thermodynamics of Ti3AlC2, Hf3AlC2, Hf3GaC2, (ZrHf)3AlC2, and (ZrHf)4AlN3 MAX Phases via the Ab Initio Method
Previous Article in Journal
Impact of Laser Power on Electrochemical Performance of CeO2/Al6061 Alloy Through Selective Laser Melting (SLM)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring Brannerite-Type Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) Oxides: Crystal Structure and Optical Properties

by
Hua-Chien Hsu
1,
Narayanan Lakshminarasimhan
1,2,
Jun Li
1,
Arthur P. Ramirez
3 and
Mas A. Subramanian
1,*
1
Department of Chemistry, Oregon State University, Corvallis, OR 97331, USA
2
CSIR-Central Electrochemical Research Institute, Karaikudi 630003, Tamil Nadu, India
3
Department of Physics, University of California Santa Cruz, Santa Cruz, CA 95064, USA
*
Author to whom correspondence should be addressed.
Crystals 2025, 15(1), 86; https://doi.org/10.3390/cryst15010086
Submission received: 29 December 2024 / Revised: 9 January 2025 / Accepted: 13 January 2025 / Published: 16 January 2025
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
Environmentally benign, highly stable oxides exhibiting desirable optical properties and high near-IR reflectance are being researched for their potential application as pigments. Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) oxides with brannerite-type structures were synthesized by the conventional solid-state reaction method to study their optical properties. These series exhibit structural transitions from brannerite (C2/m) to distorted brannerite (P 1 ¯ ) and NiV2O6-type (P 1 ¯ ) structures. The average color of Mg1−xMxV2O6 compounds varies from reddish-yellow to brown to dark brown. The L*a*b* color coordinates reveal that Mg1−xCuxV2O6 and Mg1−xNixV2O6 show more red hues in color with x = 0.4 and x = 0.5, respectively. The UV–Vis diffuse reflectance spectra indicate a possible origin for these results include the ligand-to-metal charge transfer (O2− 2p-V5+ 3d), metal-to-metal charge transfer (from Mn2+ 3d/Cu2+ 3d/Co2+ 3d/Ni2+ 3d to V5+ 3d), band gap transitions, and d–d transitions. Magnetic property measurements revealed antiferromagnetic behavior for the compounds Mg1−xMxV2O6 (M = Mn, Cu, Co, and Ni), and an oxidation state of +2 for the M ions was deduced from their Curie–Weiss behavior. The system Mg1−xMnxV2O6 has a NIR reflectance in the range between 40% and 70%, indicating its potential to be utilized in the pigment industry.

1. Introduction

Transition metal oxides have attracted significant attention in the field of solid-state chemistry since the discovery of their various captivating colors, such as chrome green (Cr2O3) and cobalt blue (CoAl2O4), which are the currently used commercial inorganic pigments. However, there are concerns about Cr and Co due to their high toxicity to the environment and/or durability among commercial pigments [1,2,3]. The objective of the present study is to explore environmentally friendly, intense, durable, and cost-effective inorganic colorants for the pigment industry. Vanadium, a 3d transition metal, exhibits various oxidation states, +3, +4, and +5, in oxides such as V2O3, VO2, and V2O5. Due to these oxidation states, vanadium oxides exhibit a diverse range of useful properties such as magnetic, catalytic, electrochemical energy storage in lithium-ion batteries, and optical properties [4,5,6]. The optical absorption of vanadium oxides lies in the near-UV-to-visible region, where ligand-to-metal charge transfer (LMCT) electronic transitions from O−2 2p to V+5 3d occur.
The serendipitous discovery of the intense blue pigment Y(In,Mn)O3 highlighted new directions for tuning pigments through structural–substitutional rationales, enabling structures amenable to substituting various transition metal ions to have their optical properties tuned accordingly [1]. For example, brannerite is a brown-colored uranium titanate mineral (UTi2O6), and the composition of this structure can differ extensively [7]. The general formula can be described as AB2O6, and this structure has larger A cations with divalent metal and small-sized B ions [8]. The structure of brannerite (AB2O6) is monoclinic with the space group C2/m in which A and B sites are octahedrally coordinated, forming chains of edge-sharing AO6 connected to edge-sharing BO6 octahedral units (Figure 1). The distance of AA within the chains is 3.53 Å, and the angle of ∠A–O–A is 104.7°; the AA distance between chains is 4.97 Å [9]. The divalent metal vanadium oxides, MV2O6 (M = Mg, Ca, Mn, Co, Ni, Cu, and Zn), have been widely studied for their magnetic, photocatalytic, fuel cell, battery applications, etc. [10,11,12]. Even as an inorganic–organic hybrid, the observed one-dimensional ferromagnetic interactions above the Néel temperature in the antiferromagnetic brannerite-type [{Mn(Bpy)x}(VO3)2] ≈ (H2O)y [x = 0.5 or 1; y = 0.62 or 1.12] reveals the potential of brannerites as quantum materials [13]. Furthermore, solid solution systems can be realized in the brannerite types like MnV2O6–MoO3, MgV2O6–CaV2O6, ZnV2O6–CaV2O6, and MnV2O6–CaV2O6 [9,14,15,16]. Lakshminarasimhan et al. recently reported the structure–optical property relationships of the solid solution Ca1−xMnxV2O6 and Zn1−xCoxV2O6 [17]. We also note that MgV2O6 exhibits a higher reflectance in the near-infrared (NIR) region compared to other divalent metal vanadium oxides. Therefore, MgV2O6, when combined with transition metals, could potentially exhibit strong reflectance in the NIR region, making these materials effective in energy-saving applications on rooftops and vehicles [2,3].
It is not surprising that not all MV2O6 compounds have this ideal brannerite-type structure, as variations in the structure can be expected due to differences in stoichiometry. For example, CuV2O6 is a slight distortion of the brannerite structure with the space group P 1 ¯ and consists of CuO6 and VO6 groups (Figure 2a). The Cu–Cu distance along the b-axis is 3.54 Å, and the ∠Cu–O–Cu bond angle is 103.99° within the chains, whereas the Cu–Cu distance along the a-axis is 4.97 Å between the chains [18]. α-CuV2O6 shows quasi-one-dimensional behavior with short-range magnetic ordering and on cooling towards a long-range antiferromagnetic transition at 24 K [19,20]. On the other hand, MV2O6 (M = Co and Ni) have NiV2O6-type structures with the space group ( P 1 ¯ ) formed by MO6 and corner-shared VO4 and corner- and edge-shared VO6 (Figure 2b). The MM distance along the b-axis is 2.96 Å, and the ∠M–O–M bond angle is 92.67° within the chains, whereas the MM distance along the a-axis is 7.13 Å between the chains in the NiV2O6-type structure [21]. Comparing these three brannerite-related structures, we have the ideal brannerite with a monoclinic angle greater than 110°, the distorted brannerite with a triclinic angle around 110°, and the NiV2O6-type with a triclinic angle approximately 93°. The structures of these compounds show octahedral coordination of the M-site ion by oxygen, often with distortion attributable to the varying magnitudes of the Jahn–Teller effect exhibited by each metal ion [22].
The valence states of transition metal ions and the crystal structure of the series played vital roles in the origin of the color of the samples reported here [2]. To further explore the optical properties of brannerite-type oxides, here, we report a new series Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni; x = 0.0 to 1.0) and their optical properties. The structures in the solid solution are also further investigated through powder X-ray diffraction (XRD), with refinements of the lattice parameters. The structure–optical property relationships are shown for the series Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni).

2. Experimental

2.1. Synthesis

We synthesized Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) compounds, with x ranging from 0.0 to 1.0, using the conventional high-temperature solid-state reaction method. Stoichiometric amounts of MgO (Alfa Aesar, Ward Hill, MA, USA, 99.5%), MnCO3 (Alfa Aesar, Ward Hill, MA, USA, 99.9%), Co(OH)2 (Alfa Aesar, Ward Hill, MA, USA, 97%), CuO (Fluka, Buchs, Switzerland, 99.9%), NiO (Alfa Aesar, Ward Hill, MA, USA, 99.99%), and V2O5 (Johnson Matthey, London, United Kingdom, 99.9%) were ground thoroughly using an agate mortar and pestle. After the mixture was homogeneous and fine, the starting materials were pressed into pellets with an applied pressure of 0.8 psi and transferred into an alumina ceramic crucible and heated in air between 600 and 700 °C for 12 h in a muffle furnace. The solid solutions Mg1−xCuxV2O6 and Mg1−xMnxV2O6 series were sintered at least twice at 600 °C and 700 °C with intermediate grinding, respectively, while the synthesis temperature of Mg1−xCoxV2O6 and Mg1−xNixV2O6 was 650 °C with repeated heating at least twice.

2.2. Characterization

Phase formation was confirmed through powder X-ray diffraction (XRD) using a Rigaku Miniflex II X-ray Diffractometer (Rigaku, Tokyo, Japan) with Cu-Kα radiation. The lattice parameters were obtained by Le Bail fitting using GSAS/EXPGUI software First version [23,24]. An internal standard, NaCl (Fm 3 ¯ m; Sigma Aldrich, St. Louis, MO, USA, 99.99%), was used while recording the XRD. Crystal structures were generated using VESTA 3.5.8 software [25]. Diffuse reflectance UV–Vis (DRUV–Vis) and NIR reflectance spectra (up to 2500 nm) were obtained using a Jasco V-670/V-770 spectrophotometer (JASCO, Tokyo, Japan). The data were converted to absorbance using the Kubelka–Munk equation. The L*a*b* color coordinates were measured using a Konica Minolta CM-700d spectrophotometer (Konica Minolta, Tokyo, Japan). The magnetization measurements were performed using a commercial magnetometer (MPMS3, Quantum Design, San Diego, CA, USA) in the temperature range 5–350 K.

3. Results and Discussion

3.1. Structural Analysis

The phase formation of Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni; x = 0.0–1.0) was analyzed by powder XRD. Figure S1, in the Supporting Information, shows the powder XRD patterns of solid-solution Mg1−xMnxV2O6 prepared at 700 °C in air. The end members, despite having the same space group (C2/m), the solid solution exhibited a mixture of phases, resulting in a miscibility gap (x = 0.4–0.8) due to slight variations in the unit cell edges for the end members. The lattice parameters of the samples were refined in order to further investigate the structures by Le Bail fitting. The lattice parameters and unit cell volume vs. x in the Mg1−xMnxV2O6 are shown in Figure S2. The lattice parameters and unit cell volume increase upon Mn substitution due to the larger ionic radii of Mn2+ (0.83 Å) when compared to Mg2+ (0.72 Å) in six coordination [26].
Similarly, Mg1−xCuxV2O6 (x = 0.0–1.0) samples were synthesized at 600 °C in air, and the powder XRD patterns are shown in Figure 3. The Mg1−xCuxV2O6 series shows a change in structure from monoclinic brannerite (C2/m) to brannerite with a slight triclinic distortion ( P 1 ¯ ) [17]. We can see the sample has a mixed phase starting from x = 0.4 with impurities. Based on the structural analysis (Figure 4a,b), the lattice parameters a, b, and c; angle β; and the unit cell volume are presented for the ranges x = 0.0–0.4 and x = 0.8–1.0. These variations are attributed to a miscibility gap between the C2/m and P 1 ¯ space groups, leading to a mixture of phases with impurities at x = 0.6. The a and c parameters show a decrease in the lattice through the substitution of Mg2+ by Cu2+, while the b parameters exhibit an increase. Since the ionic radii of Mg2+ and Cu2+ are 0.72 Å and 0.73 Å in the 6-fold coordination, respectively, as the copper content increases, the lattice parameters also exhibit a marginal increase [26]. The a and c parameters, however, show a decrease, and this probably can be explained by the structural distortion leading to lattice contraction. We can observe that the lattice parameter c exhibits a larger decrease compared to the other parameters. The change in the angle β and decrease in volume with the increasing amount of Cu2+ in the brannerite-type structure are observed in Figure 4c,d. Therefore, we can conclude that the shrinking of the brannerite structure, accompanied by distortion upon Cu2+ substitution, is influenced by the strong Jahn–Teller effect of Cu2+.
In the series Mg1−xCoxV2O6 (x = 0.0–1.0) and Mg1−xNixV2O6 (x = 0.0–1.0), solid solutions were formed with some miscibility gaps. It is to be noted that CoV2O6 and NiV2O6 have a triclinic structure ( P 1 ¯ ). These compounds were synthesized at 650 °C in air, and a phase transition was clearly observed in the series through XRD analysis, as shown in Figure S3. In the Mg1−xCoxV2O6 series, the phase transition to a triclinic structure ( P 1 ¯ ) occurred when the cobalt content was x = 0.8. For compositions with x ≤ 0.6, the major phase presented was a monoclinic structure (C2/m) with impurities. Similarly, in the series of Mg1−xNixV2O6 samples, the major phase of the brannerite structure was found for compositions x ≤ 0.1, while for x ≥ 0.3, compounds predominantly showed a triclinic structure. Therefore, the presence of two distinct structures in the end members causes the series to exhibit a clear phase transition upon substitution.
The variations in the unit cell edges in the Mg1−xCoxV2O6 series show different scales in two regions: x = 0.0–0.6 and x = 0.8–1.0 (Figure S4). The lattice parameters do not change much except for ‘c’, which shows a decrease with the increasing Co2+ composition from x = 0.0 to x = 0.6. It is possible that the Co2+ substitution leads to a decrease in the average distance between neighboring ions in unit cell c. For the triclinic structure (x = 0.8–1.0), the lattice parameters a and c increase as the cobalt content increases. This can be due to the larger ionic radius of Co2+ (0.735 Å) as compared to that of Mg2+ (0.72 Å) [26]. In the series of Mg1−xNixV2O6, with both end members having different structures (brannerite and NiV2O6 types), a decrease in the a and c lattice parameters was observed as the Ni2+ content increases (Figure S5). This can be attributed to the smaller ionic radius of Ni2+ (0.69 Å) when compared to Mg2+ (0.72 Å) [26]. The lattice parameter ‘b’ also exhibits a decreasing trend upon the Ni2+ substitution in the NiV2O6 ( P 1 ¯ ) phase (x ≥ 0.8). Contrarily, the C2/m phase (x < 0.8) shows an increase in the ‘b’ lattice parameter when Ni increases. To summarize, it is known that brannerite and NiV2O6-type structures have different segments within the morphotropic series and, hence, a miscibility gap between them.

3.2. Optical Properties

The samples in the Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) series exhibit varying colors that can be understood by measuring the optical properties. The color of the Mg1−xMnxV2O6 samples changes from brown to dark brown and eventually to black as the amount of Mn2+ increases (Figure S6a). The colors originate from the combination of band gap, d–d transitions, and charge transfers, transitions which can be confirmed by diffuse reflectance UV–Vis spectra of Mg1−xMnxV2O6 samples, as shown in Figure S6b. The spectra show that, with the increasing concentration of Mn2+, the optical absorption edges are red-shifted. This can be understood by narrowing of the band gap with the increasing Mn2+ concentration, as the end members of the solid solution MgV2O6 (Eg = 2.3 eV) and MnV2O6 (Eg = 1.44 eV) have different band gaps [17]. The band gap in these compounds can be considered as the difference in energy levels between the filled O 2p band and empty V 3d band. The absorption edges not only shift from 580 to 1000 nm but also show a variation in intensity, and these are evidence for the influence of Mn2+ on the ligand-to-metal charge transfer (LMCT: O2−-V5+) in Mg1−xMnxV2O6. In addition to band gap (or LMCT) transition, the introduction of Mn2+ induces absorption bands in the visible region, as observed with humps in the region 450–550 nm for x = 0.1–0.7 and between 600 and 750 nm for x = 0.7. These absorptions are due to the d–d transition of Mn2+ in a distorted six-coordination environment. In general, Mn2+ in a perfect octahedral coordination with an inversion center does not exhibit any d–d transition [2]. However, the forbidden d–d transitions are enhanced in intensity because of the distorted octahedral coordination on Mn2+. Also, the metal-to-metal charge transfer from half-filled Mn 3d orbitals to empty V 3d orbitals influences the color of the samples. The corresponding L*a*b* color coordinates were calculated. The positive values of a* indicate red hues, which decrease in the Mg1−xMnxV2O6 series upon Mn2+ substitution. Similarly, the positive values of b*, which represent yellow, decrease as well in this series. Furthermore, the lightness (L*) decreases with the increase in the Mn2+ substitution, as listed in Table S1. The NIR reflectance spectra of the Mg1−xMnxV2O6 samples are measured (Figure S6c). The percentage reflectance of Mg1−xMnxV2O6 varies between 75% and 40% across the NIR region for x = 0.0 to x = 1.0, revealing a decrease in NIR reflectance with the increasing Mn2+.
In the photographs of the samples, DRUV–Vis absorption and reflectance spectra of the Mg1−xCuxV2O6 samples are shown in Figure 5. The DRUV–Vis reflectance spectrum of TiO2 is also shown in Figure 5c (also in Figures S6–S8) as a comparison since it has a high NIR reflectance due to its high refractive index and application as a white, reflective pigment. In the UV–Vis spectra, it is clear that, as the amount of Cu2+ increases, the optical absorption edge is red-shifted from 600 to 780 nm, which can be understood as due to the decreasing band gap, which is different for MgV2O6 (Eg = 2.3 eV) and CuV2O6 (Eg = 1.69 eV) [16]. The broad optical absorption bands in the visible region (400–600 nm) indicate Cu 3d to V 3d metal–metal charge transfer (MMCT) transitions [27]. In the wavelength range of 650–900 nm, the weak absorption bands arise from partially filled Cu 3d eg orbitals in the octahedral coordination. In addition, the Mg1−xCuxV2O6 series with x = 0.2 to x = 1.0 exhibits a strong absorption in the near-IR region. The images of the samples and L*a*b* parameters showing the color changes in the Mg1−xCuxV2O6 series are presented in Figure 5a and Table 1, respectively. According to L*a*b* color measurement, the solid-solution series Mg1−xCuxV2O6 displays colors ranging from red to yellow hues, particularly the sample with x = 0.4, showing a more pronounced red hue. As the concentration of Cu2+ increases, the color changes from yellowish to brown to black, which makes an agreement with the color measurement.
In the series Mg1−xCoxV2O6, as in the previous cases, the color changes to dark, as shown in the photographs (Figure S7a), and the optical absorption edge shows a red shift with the increasing Co2+ content (Figure S7b). Also, the compositions with x = 0.2, 0.4, and 0.6 show the distinct d–d transitions in the wavelength regions 650–780 nm and 780–1000 nm due to Co2+ present in the octahedral coordination environment. The variation in the optical absorption (x < 0.8 and x ≥ 0.8) is attributed to the differences in the crystal structures (Brannerite vs. NiV2O6-type). This is similar to the observations of optical properties of Co2+ in Zn1−xCoxV2O6 with brannerite-type structures [17]. The NIR reflectance of the samples in the Mg1−xCoxV2O6 series exhibits a decrease from 70 to 20% with the increasing Co2+ concentration (Figure S7c). From the measured L*a*b* color coordinates for the Mg1−xCoxV2O6 samples, it is clear that the color shifts from yellowish to brown to dark brown and black as the cobalt composition increases (Table S2). For the sample with x = 0.4, the color has a more reddish tone as compared to the other compositions. Furthermore, it is evident that the yellow hue diminishes with the increasing cobalt content.
The color of the samples in the series Mg1−xNixV2O6 shows a change from brown to reddish brown to a dark color with the increasing Ni2+ concentration in Figure S8. The DRUV–Vis absorption spectra of these samples exhibit a red shift in the optical absorption edge with the increasing Ni2+, and this is similar to the observed results with the introduction of transition metal ions in MgV2O6, as discussed above. The optical absorption occurs at 350 nm in the near-UV region due to the ligand–metal charge transfer (O2− 2p V5+ 3d). There are two absorption bands in the visible region located at ~ 450 nm and 750 nm, which indicate 3A2g   3T1g (P) and 3A2g   3T1g (F) transitions, respectively, of Ni2+ ions in the octahedral coordination [17,28]. In the NIR reflectance, the spectrum indeed shows an intense absorption from x = 0.2 to x = 1.0 and as the Ni2+ replacement increases, the percentage reflectance decreases. The L*a*b* color coordinates show a decreasing lightness (L*), which means the samples become darker. Meanwhile, the b* decreases, showing the decrease in yellow hue, and the composition with x = 0.5 has a more intense red color of a* that is 20.23 (Table S3).

3.3. Magnetic Properties

To aid in understanding the microscopic origin of the color changes, the valence states of the transition metal ions can be deduced from the magnetic behavior of the compounds discussed above. It is known that MnV2O6 exhibits antiferromagnetic (AFM) behavior with non-frustrated Mn2+ ordering [29]. To understand the influence on the magnetic behavior upon the substitution of Mn2+ with Mg2+, the magnetic properties of Mg1−xMnxV2O6 (x = 0.9 and 1.0) were measured in the temperature range between 5 and 350 K, and the results are shown in Figure S9. It shows antiferromagnetic (AFM) ordering of Mn2+ in octahedral coordination, and the Néel temperatures (TN) are 17.0 and 20.0 K, respectively. These values are in close agreement with the reported TN of 20 K [29]. The results show that the decrease in TN is consistent with the simple mean field dependence of the ordering temperature on the coordination number. Table 2 lists the calculated magnetic moments (μB), Curie constant, C (emu K/mol), and Weiss constant, ΘCW (K). In order to prove the oxidation states of Mn in Mg1−xMnxV2O6, we used the Curie–Weiss law to fit 𝒳 m o l 1 vs. T in the temperature range between 250 and 300 K, and from the linear region, the slope represents 1/C and y-intercept ΘCW/C [29]. The negative values of ΘW (Table 2) confirm the AFM nature of the nearest neighbor interaction, and the downturn in 1/χ on cooling below 50 K indicates the likelihood of next-nearest-neighbor interactions. The calculated effective spin-only magnetic moments of the samples with x = 0.9 and 1.0 are 6.13 and 6.18 μB, respectively, and these values are in good agreement with the theoretical magnetic moment (5.92 μB) of the divalent Mn2+ ion. Thus, the oxidation state of manganese in Mg1−xMnxV2O6 is confirmed as 2+.
For Mg1−xCuxV2O6 (x = 0.8 and 1.0) in Figure 6, as the Cu concentration decreases, we observed a decrease in the magnetic ordering temperature. Additionally, at x = 1.0, antiferromagnetic ordering with TN = 22 K is observed, consistent with previous studies [27]. In the inset in Figure 6, the molar susceptibility of CuV2O6 is displayed, and we observed that the magnitude of the ferromagnetic signal was around 3 emu/mol, which is very small compared to a FM saturation moment. This signal, which accounts for less than 0.1% of the compound, suggests the presence of uncompensated spins. We can conclude that the magnetic coupling is relatively weak in the structures. Mg0.2Cu0.8V2O6 exhibits magnetic interactions along the b-axis, with an edge-sharing octahedra contributing to ferromagnetic ordering. Based on the Curie–Weiss law in the temperature range between 250 and 300 K, a negative Curie temperature (Θ) suggests that the Cu2+ magnetic moments might overall exhibit antiferromagnetic correlations. The calculated effective moments for the samples with x = 0.8 and 1.0 are 2.16 and 2.14 μB, respectively (Table 2). The results show a slightly larger magnetic moment than the theoretical value (μeff = 1.73 μB); however, the values fall within the observed range and are consistent with the literature [27]. The enhanced effective magnetic moment is attributed to the partial orbital contribution resulting from Jahn–Teller distortion. This is further supported by the discussion of lattice parameters in the Structural Analysis (Section 3.1), with the structure of CuV2O6 shown in Figure 2a. The measurements confirm that the oxidation state of copper is indeed +2 in Mg1−xCuxV2O6 (x = 0.8 and 1.0) within the structure.
Similarly, the samples of Mg1−xCoxV2O6 (x = 0.8 and 1.0) and Mg1−xNixV2O6 (x = 0.9 and 1.0) clearly exhibit antiferromagnetic ordering, as shown in Figures S10 and S11. When the concentrations of Co and Ni are decreased, these samples show a corresponding decrease in the ordering temperature. In the Mg1−xCoxV2O6 (x = 0.8 and 1.0) series, the AFM transition with Néel temperatures (TN) was observed at 2.8 and 7.0 K, respectively, while the samples presented TN at 14.5 and 15.0 K, respectively, in Mg1−xNixV2O6 (x = 0.9 and 1.0). It is worth noting that the susceptibility decreases by 10% in Mg0.1Ni0.9V2O6 compared to NiV2O6, which implies the sample was successfully synthesized as a solid solution. The observed magnetic moments in Mg1−xCoxV2O6 (x = 0.8 and 1.0) are 5.27 and 5.24 μB, respectively, which are much larger than the expected spin-only moment (3.87 μB) of high-spin Co2+ due to orbital contribution in the temperature range between 50 and 150 K. For the Mg1−xNixV2O6 (x = 0.9 and 1.0) samples, the observed magnetic moments of the Ni spins with S = 1 are 3.40 and 3.37 μB, respectively, in the temperature range between 50 and 200 K. According to the Curie–Weiss law results, the oxidation states of Co and Ni were confirmed as +2 in Mg1−xCoxV2O6 and Mg1−xNixV2O6. Thus, the magnetic measurements of the studied brannerite samples helped in probing the oxidation states of the transition metal ions, which, in turn, explains the observed optical properties originating from d–d transitions, LMCT, and/or MMCT.

4. Conclusions

New brannerite-type solid solutions Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) were synthesized, and their optical properties were studied. The results showed a miscibility gap between two segments of morphotropic transitions in solid solutions, causing phase transitions within the brannerite structure from C2/m to P 1 ¯ . The color and optical properties of these series originate from a combination of band gap, LMCT, MMCT, and d–d transitions. It was observed that Mg1−xCuxV2O6 and Mg1−xNixV2O6 show more red hues in color with x = 0.4 and x = 0.5, respectively. Few compositions in the series Mg1−xMnxV2O6 exhibit a high reflectance, which shows potential for their application as a ‘cool’ pigment. Furthermore, Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) display antiferromagnetic behavior in the temperature range between 5 and 350 K. Further, the magnetic moments deduced from the measurements confirmed the presence of Mn2+, Cu2+, Co2+, and Ni2+ in these series. It is worth exploring brannerite-type oxides as pigments due to their flexible compositional tuning with less expensive and non-toxic elements and relatively low temperature of synthesis.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/cryst15010086/s1: Figure S1: Powder XRD patterns of Mg1−xMnxV2O6 (x = 0.0–1.0) samples that were prepared at 700 °C in air; Table S1: The measured L*a*b* color coordinates of Mg1−xMnxV2O6; Figure S2: Variations in the (a) lattice parameters ‘a’ and ‘b’, (b) lattice parameter ‘c’, (c) angle β, and (d) volume of the Mg1−xMnxV2O6 series as a function of x; Table S2: The measured L*a*b* color coordinates of Mg1−xCoxV2O6; Figure S3: Powder XRD patterns of (a) Mg1−xCoxV2O6 and (b) Mg1−xNixV2O6 prepared at 650 °C in air; Table S3: The measured L*a*b* color coordinates of Mg1−xNixV2O6; Figure S4: Variations in the (a) lattice parameters ‘a’ and ‘b’ and (b) lattice parameter ‘c’ of the Mg1−xCoxV2O6 series as a function of x; Figure S5: Variations in the (a) lattice parameters ‘a’ and ‘b’ and (b) lattice parameter ‘c’ of the Mg1−xNixV2O6 series as a function of x; Figure S6: (a) Photographs, (b) DRUV–Vis absorbance, and (c) NIR reflectance spectra of the Mg1−xMnxV2O6 solid solution series. The NIR reflectance of TiO2 is also shown for comparison in (c); Figure S7: (a) Photographs, (b) DRUV–Vis absorbance, and (c) NIR reflectance spectra of the Mg1−xCoxV2O6 compounds, along with TiO2; Figure S8: (a) Photographs, (b) DRUV–Vis absorbance, and (c) NIR reflectance spectra of the Mg1−xNixV2O6 compounds, along with TiO2; Figure S9: Magnetic susceptibility and inverse molar susceptibility vs. the temperatures of MnV2O6 and Mg0.1Mn0.9V2O6; Figure S10: Magnetic susceptibility and inverse molar susceptibility vs. the temperatures of CoV2O6 and Mg0.2Co0.8V2O6; Figure S11: Magnetic susceptibility and inverse molar susceptibility vs. the temperatures of NiV2O6 and Mg0.1Ni0.9V2O6.

Author Contributions

N.L. conceived the research interest. H.-C.H. carried out the synthesis and properties characterization. A.P.R. conducted the magnetic measurements. M.A.S., N.L. and J.L. discussed the structural characterization and optical properties. All authors contributed to the drafting and editing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The work done at Oregon State University was supported by NSF Grant No. DMR-2025615 (M.A.S.). The work done at UC Santa Cruz (A.P.R.) was supported by NSF Grant No. DMR-2218130.

Data Availability Statement

The research data will be made available on request.

Acknowledgments

The author N.L. acknowledges the United States–India Educational Foundation (USIEF) for the Fulbright-Nehru Academic and Professional Excellence Fellowship (2020–2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Subramanian, M.A.; Li, J. Challenges in the Rational Design of Intense Inorganic Pigments with Desired Colours. Nat. Rev. Mater. 2022, 8, 71–73. [Google Scholar] [CrossRef]
  2. Subramanian, M.A.; Li, J. YInMn Blue—200 Years in the Making: New Intense Inorganic Pigments Based on Chromophores in Trigonal Bipyramidal Coordination. Mater. Today Adv. 2022, 16, 100323. [Google Scholar] [CrossRef]
  3. Li, J.; Subramanian, M.A. Inorganic Pigments with Transition Metal Chromophores at Trigonal Bipyramidal Coordination: Y(In,Mn)O3 Blues and Beyond. J. Solid State Chem. 2019, 272, 9–20. [Google Scholar] [CrossRef]
  4. von Dreifus, D.; Pereira, R.; Rodrigues, A.D.; Pereira, E.C.; de Oliveira, A.J.A. Sol-Gel Synthesis of Triclinic CoV2O6 Polycrystals. Ceram. Int. 2018, 44, 19397–19401. [Google Scholar] [CrossRef]
  5. Watanabe, M.; Muto, M.; Abe, Y.; Kaneko, T.; Toda, A.; Uematsu, K.; Ishigaki, T.; Sato, M.; Koide, J.; Toda, M.; et al. Synthesis of Red-Emissive CaV2O6 Nanophoshor via a Water Assisted Solid State Reaction Method. ECS J. Solid State Sci. Technol. 2021, 10, 106010. [Google Scholar] [CrossRef]
  6. Wang, F.; Zhang, H.; Liu, L.; Shin, B.; Shan, F. Synthesis, Surface Properties and Optical Characteristics of CuV2O6 Nanofibers. J. Alloys Compd. 2016, 672, 229–237. [Google Scholar] [CrossRef]
  7. Brannerite. Available online: https://www.sciencedirect.com/topics/earth-and-planetary-sciences/brannerite (accessed on 19 August 2023).
  8. Beck, H.P. A Study on AB2O6 Compounds, Part III: Co-Ordination Needs and Patterns of Connectivity. Z. Kristallogr. Cryst. Mater. 2013, 228, 271–288. [Google Scholar] [CrossRef]
  9. Ng, H.N.; Calvo, C. Crystal Structure of and Electron Spin Resonance of Mn2+ in MgV2O6. Can. J. Chem. 1972, 50, 3619–3624. [Google Scholar] [CrossRef]
  10. Mocała, K.; Ziółkowski, J. Polymorphism of the Bivalent Metal Vanadates MeV2O6 (Me = Mg, Ca, Mn, Co, Ni, Cu, Zn, Cd). J. Solid State Chem. 1987, 69, 299–311. [Google Scholar] [CrossRef]
  11. Jin, X.; Ding, X.; Qin, Z.; Li, Y.; Jiao, M.; Wang, R.; Yang, X.; Lv, X. Comprehensive Study of Electronic, Optical, and Thermophysical Properties of Metavanadates CaV2O6 and MgV2O6. Inorg. Chem. 2022, 61, 17623–17633. [Google Scholar] [CrossRef]
  12. Kimber, S.A.; Mutka, H.; Chatterji, T.; Hofmann, T.; Paul, F.H.; Bordallo, H.N.; Argyriou, D.N.; Attfield, J.P. Metamagnetism and Soliton Excitations in the Modulated Ferromagnetic Ising Chain CoV2O6. Phys. Rev. B 2011, 84, 104425. [Google Scholar] [CrossRef]
  13. Fernández de Luis, R.; Urtiaga, M.K.; Mesa, J.L.; Vidal, K.; Lezama, L.; Rojo, T.; Arriortua, M.I. Short-Range and Long-Range Magnetic Ordering, in Third Generation Brannerite Type Inorganic−Organic Vanadates: [{Mn(Bpy)}(VO3)2] ≈ (H2O)1.16 and [{Mn(Bpy)0.5}(VO3)2] ≈ (H2O)0.62. Chem. Mater. 2010, 22, 5543–5553. [Google Scholar] [CrossRef]
  14. Krasnenko, T.I.; Zabara, O.A.; Surat, L.L.; Strepetov, S.V. Investigation of the Ca1-xMnx(VO3)2 Solid Solution. Inorg. Mater. 1987, 23, 1252–1254. [Google Scholar]
  15. Hansen, S.; Nilsson, J.; Teixidor, F.; Viñas, C.; Abad, M.M.; Nielsen, R.I.; Olsen, C.E.; Rosendahl, C.N.; Haugg, M.; Trabesinger-Rüf, N.; et al. Coordination of Vanadium (5+) in Solid Solutions MgV2O6-CaV2O6 and ZnV2O6-CaV2O6. Acta Chem. Scand. 1996, 50, 512–515. [Google Scholar] [CrossRef]
  16. Kozłowski, R.; Ziółkowski, J.; Mocała, K.; Haber, J. Defect Structures in the Brannerite-Type Vanadates. I. Preparation and Study of Mn1−xфxV2−2xMo2xO6 (0 ≤ x ≤ 0.45). J. Solid State Chem. 1980, 35, 1–9. [Google Scholar] [CrossRef]
  17. Lakshminarasimhan, N.; Li, J.; Hsu, H.-C.; Subramanian, M.A. Optical Properties of Brannerite-Type Vanadium Oxides, MV2O6 (M = Ca, Mg, Mn, Co, Ni, Cu, or Zn). J. Solid State Chem. 2022, 312, 123279. [Google Scholar] [CrossRef]
  18. Calvo, C.; Manolescu, D. Refinement of the Structure of CuV2O6. Acta Crystallogra. Sect. B Struct. Sci. 1973, 29, 1743–1745. [Google Scholar] [CrossRef]
  19. Prokofiev, A.V.; Kremer, R.K.; Assmus, W. Crystal Growth and Magnetic Properties of α-CuV2O6. J. Cryst. Growth 2001, 231, 498–505. [Google Scholar] [CrossRef]
  20. Vasil’ev, A.N.; Ponomarenko, L.A.; Antipov, E.V.; Velikodny, Y.A.; Smirnov, A.I.; Isobe, M.; Ueda, Y. Short-Range and Long-Range Magnetic Ordering in α-CuV2O6. Physica B 2000, 284–288, 1615–1616. [Google Scholar] [CrossRef]
  21. Müller-Buschbaum, H.K.; Kobel, M. Ein Neuer Strukturtyp Der Oxovanadate MV2O6: NiV2O6. Z. Anorg. Allg. Chem. 1991, 596, 23–28. [Google Scholar] [CrossRef]
  22. Halcrow, M.A. Jahn–Teller Distortions in Transition Metal Compounds, and Their Importance in Functional Molecular and Inorganic Materials. Chem. Soc. Rev. 2013, 42, 1784–1795. [Google Scholar] [CrossRef]
  23. Larson, A.C.; Dreele, R.B. General Structure Analysis System; Los Alamos National Laboratory Report; Los Alamos National Laboratory: Los Alamos, NM, USA, 2000; p. 86. [Google Scholar]
  24. Toby, B.H. EXPGUI, a Graphical User Interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  25. 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]
  26. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  27. Hossain, M.K.; Sarker, H.P.; Sotelo, P.; Dang, U.; Rodríguez-Gutiérrez, I.; Blawat, J.; Vali, A.; Xie, W.; Oskam, G.; Huda, M.N.; et al. Phase-Pure Copper Vanadate (α-CuV2O6): Solution Combustion Synthesis and Characterization. Chem. Mater. 2020, 32, 6247–6255. [Google Scholar] [CrossRef]
  28. Rossman, G.R.; Shannon, R.D.; Waring, R.K. Origin of the Yellow Color of Complex Nickel Oxides. J. Solid State Chem. 1981, 39, 277. [Google Scholar] [CrossRef]
  29. Kimber, S.A.; Attfield, J.P. Disrupted Antiferromagnetism in the Brannerite MnV2O6. Phys. Rev. B 2007, 75, 064406. [Google Scholar] [CrossRef]
Figure 1. The unit cell of an ideal brannerite structure with the space group, C2/m. The general formula is AB2O6. The A and B sites are octahedrally coordinated, forming chains of edge-sharing AO6 connected to edge-sharing BO6.
Figure 1. The unit cell of an ideal brannerite structure with the space group, C2/m. The general formula is AB2O6. The A and B sites are octahedrally coordinated, forming chains of edge-sharing AO6 connected to edge-sharing BO6.
Crystals 15 00086 g001
Figure 2. (a) The unit cell of a slightly distorted brannerite structure (CuV2O6) with the space group, P 1 ¯ . (b) The unit cell of NiV2O6-type structures with the space group, P 1 ¯ , formed by MO6 (M = Co or Ni) and corner-shared VO4 and corner- and edge-shared VO6.
Figure 2. (a) The unit cell of a slightly distorted brannerite structure (CuV2O6) with the space group, P 1 ¯ . (b) The unit cell of NiV2O6-type structures with the space group, P 1 ¯ , formed by MO6 (M = Co or Ni) and corner-shared VO4 and corner- and edge-shared VO6.
Crystals 15 00086 g002
Figure 3. Powder XRD patterns of Mg1−xCuxV2O6 (x = 0.0–1.0) samples prepared at 600 °C in air.
Figure 3. Powder XRD patterns of Mg1−xCuxV2O6 (x = 0.0–1.0) samples prepared at 600 °C in air.
Crystals 15 00086 g003
Figure 4. Variations in the (a) lattice parameters ‘a’ and ‘b’, (b) lattice parameter ‘c’, (c) angle β, and (d) volume of the Mg1−xCuxV2O6 series as a function of x.
Figure 4. Variations in the (a) lattice parameters ‘a’ and ‘b’, (b) lattice parameter ‘c’, (c) angle β, and (d) volume of the Mg1−xCuxV2O6 series as a function of x.
Crystals 15 00086 g004
Figure 5. (a) Photographs, (b) DRUV–Vis absorbance, and (c) NIR reflectance spectra of Mg1−xCuxV2O6 compounds. The NIR reflectance of TiO2 is also shown for comparison in (c).
Figure 5. (a) Photographs, (b) DRUV–Vis absorbance, and (c) NIR reflectance spectra of Mg1−xCuxV2O6 compounds. The NIR reflectance of TiO2 is also shown for comparison in (c).
Crystals 15 00086 g005
Figure 6. Magnetic susceptibility and inverse molar susceptibility vs. the temperatures of CuV2O6 and Mg0.2Cu0.8V2O6.
Figure 6. Magnetic susceptibility and inverse molar susceptibility vs. the temperatures of CuV2O6 and Mg0.2Cu0.8V2O6.
Crystals 15 00086 g006
Table 1. The measured L*a*b* color coordinates of Mg1−xCuxV2O6.
Table 1. The measured L*a*b* color coordinates of Mg1−xCuxV2O6.
L*a*b*
x = 0.054.5212.2643.08
x = 0.145.2316.2032.47
x = 0.246.2918.2140.08
x = 0.436.1820.7921.89
x = 0.629.5914.085.52
x = 0.828.298.432.42
x = 0.925.414.520.22
x = 1.021.613.880.08
Table 2. Observed magnetic moments μeffB), and Curie and Weiss constants of Mg1−xMxV2O6 (M = Mn, Cu, Co, and Ni). The theoretical spin-only magnetic moments are calculated for Mn2+ (d5), Cu2+ (d9), Co2+ (d7), and Ni2+ (d8).
Table 2. Observed magnetic moments μeffB), and Curie and Weiss constants of Mg1−xMxV2O6 (M = Mn, Cu, Co, and Ni). The theoretical spin-only magnetic moments are calculated for Mn2+ (d5), Cu2+ (d9), Co2+ (d7), and Ni2+ (d8).
CompositionCurie Constant
(emu K/mol)
Weiss Constant (K)Obs. Mag. Moment (μB)Th. Mag. Moment (μB)
Mg0.1Mn0.9V2O64.69−9.366.135.92
MnV2O64.77−8.516.185.92
Mg0.2Cu0.8V2O60.58−52.242.161.73
CuV2O60.57−86.192.141.73
Mg0.2Co0.8V2O63.47−12.225.273.87
CoV2O63.43−11.355.243.87
Mg0.1Ni0.9V2O61.455.433.402.83
NiV2O61.426.803.372.83
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

Hsu, H.-C.; Lakshminarasimhan, N.; Li, J.; Ramirez, A.P.; Subramanian, M.A. Exploring Brannerite-Type Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) Oxides: Crystal Structure and Optical Properties. Crystals 2025, 15, 86. https://doi.org/10.3390/cryst15010086

AMA Style

Hsu H-C, Lakshminarasimhan N, Li J, Ramirez AP, Subramanian MA. Exploring Brannerite-Type Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) Oxides: Crystal Structure and Optical Properties. Crystals. 2025; 15(1):86. https://doi.org/10.3390/cryst15010086

Chicago/Turabian Style

Hsu, Hua-Chien, Narayanan Lakshminarasimhan, Jun Li, Arthur P. Ramirez, and Mas A. Subramanian. 2025. "Exploring Brannerite-Type Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) Oxides: Crystal Structure and Optical Properties" Crystals 15, no. 1: 86. https://doi.org/10.3390/cryst15010086

APA Style

Hsu, H.-C., Lakshminarasimhan, N., Li, J., Ramirez, A. P., & Subramanian, M. A. (2025). Exploring Brannerite-Type Mg1−xMxV2O6 (M = Mn, Cu, Co, or Ni) Oxides: Crystal Structure and Optical Properties. Crystals, 15(1), 86. https://doi.org/10.3390/cryst15010086

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