Novel environment friendly inorganic red pigments based on Bi4V2O11

Wendusua, Atsunori Shiraishib, Naoki Takeuchia, Toshiyuki Masuia and Nobuhito Imanaka*a
aDepartment of Applied Chemistry, Faculty of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan. E-mail: imanaka@chem.eng.osaka-u.ac.jp; Fax: +81-6-6879-7354; Tel: +81-6-6879-7352
bSaga Ceramics Research Laboratory, 3037-7 Kuromuta-hei, Arita-cho, Nishimatsuura-gun, Saga 844-0022, Japan

Received 18th March 2015 , Accepted 11th May 2015

First published on 11th May 2015


Abstract

Novel environment friendly inorganic red pigments, (Bi1−xyZrxAly)4V2O11+δ (0 ≤ x ≤ 0.15; 0 ≤ y ≤ 0.10), were successfully synthesized by a conventional solid state reaction method, and their colour properties were characterized. The colour of the obtained pigments was dependent on the calcination condition and the sample composition, and the highest red hue was obtained for the (Bi0.92Zr0.07Al0.01)4V2O11.34 pigment calcined at 800 °C for 10 h in a flow of pure O2. The a* value corresponding to the red chromaticity of this pigment was +41.9, which was significantly greater than that of the commercial iron oxide pigment (a* = +28.9). Bi4V2O11 is considered to be a nontoxic compound, and the other components (Zr and Al) are also safe elements. Therefore, the present pigment could be an attractive candidate as a novel environment friendly inorganic red pigment.


Introduction

Inorganic pigments are used for a wide variety of products such as paints, inks, plastics, ceramic tiles, and porcelains.1 Among the inorganic pigments with various colours, red ones are especially in demand due to their high visibility. Several red pigments such as cadmium red (CdS·CdSe), mercuric sulfide red (HgS), and lead oxide red (Pb3O4 or PbO·Pb2O3) have been popularly used as industrial inorganic red materials for a long period of time, because these compounds can exhibit a brilliant red colour. However, the use of these materials has been restricted because they contain toxic elements (Cd, Se, Hg, and Pb), which are harmful to human health and the environment. On the other hand, iron oxide red (Fe2O3) are still widely used for many applications, since this material is environmental friendly. Unfortunately, its colour is insufficient compared with those of the conventional toxic ones. Although a number of inorganic compounds have been reported as environmental friendly inorganic red pigments, the redness of these pigments still cannot exceed that of the iron oxide red.2–10 Perovskite-type oxynitrides, Ca(1−x)LaxTaO(2−x)N(1+x), have also been reported as nontoxic red pigments, but toxic and flammable ammonia gas is required for the synthesis process.11 In addition, these oxynitride compounds are decomposed to oxides at 420 °C and higher temperatures,11 accompanying with generation of toxic nitrogen oxides.

In our previous studies, bismuth oxide (Bi2O3) was chosen as a base material because this compound has already been reported to be nontoxic and insoluble.12 Although Bi2O3 is a pale yellow compound, its colour was altered to be reddish by the control of the lattice size and the band structure by doping Er3+, Y3+, and Fe3+ ions into the Bi2O3 lattice.13,14 Among them, a ((Bi0.72Er0.04Y0.24)0.80Fe0.20)2O3 pigment exhibited the most brilliant red hue (a* = +33.1 in the CIE L*a*b* system).14 The redness value of this pigment surpassed that of a commercially available Fe2O3 pigment (a* = +28.9), but the difference was small. Accordingly, it is highly required to synthesize new red pigments with more reddish hue.

Because of this situation, we focused on bismuth vanadate (Bi4V2O11) with orthorhombic structure (space group: Aba2)15,16 as a new base material. The structure of the orthorhombic Bi4V2O11 consists of Bi2O2 layers interleaved with vanadium oxide sheets, which are composed of edge-shared VO4 and VO6 units. However, half of the VO6 units contain an oxide anion deficiency, and accordingly, the VO4 layer acts as VO3.5.16 The colouring mechanism of the Bi4V2O11 is based on a charge transfer transition from a hybrid Bi6s–O2p orbital to V3d orbital in the band structure. Therefore, it is expected that the colour of the Bi4V2O11 can be tuned by introducing other elements into the lattice to control the lattice size, because the extent of the orbital hybridization in the Bi6s–O2p valence band should depend on the interionic distance between Bi3+ and O2−.

In our previous works, it has been elucidated that lattice shrinkage positively affects the colour properties by modulating the width of valence band, whereas, an excess amount of oxide anion vacancies gives a negative effect on the colour properties.17–19 In accordance with these facts, in the present study, oxide anion deficiencies in the Bi4V2O11 lattice were filled by the calcination in a flow of pure oxygen, accompanying with the oxidation state change of a part of bismuth ions from Bi3+ to Bi5+. Furthermore, smaller Zr4+ (ionic radius: 0.072 nm)20 and/or Al3+ (ionic radius: 0.0535 nm)20 ions were doped into the Bi3+ (ionic radius: 0.103 nm)20 and Bi5+ (ionic radius: 0.076 nm)20 sites of the lattice. In this study, therefore, novel (Bi1−xyZrxAly)4V2O11+δ (0 ≤ x ≤ 0.15; 0 ≤ y ≤ 0.10) red pigments were synthesized as advanced environmental friendly red materials. The colour properties of the samples were characterized, and the doping effects of Zr4+ and Al3+ were revealed.

Experimental

The (Bi1−xyZrxAly)4V2O11+δ (0 ≤ x ≤ 0.15; 0 ≤ y ≤ 0.10) pigments were synthesized using a conventional solid state reaction method. Bi2O3, ZrO2, Al2O3, and V2O5 powders were mixed in a stoichiometric amount in an agate mortar. After mechanically mixed with a planetary-type ball-milling apparatus (PULVERISETTE 7, Fritsch GmbH), the homogenous mixture was calcined at 800 °C for 10 h in a flow of air or pure oxygen. Before characterization, all samples were gently ground in an agate mortar.

The crystal structures of the samples were identified by X-ray powder diffraction (XRD; Rigaku SmartLab) using Cu-Kα radiation (40 kV, 30 mA). The compositions of the samples were analyzed by X-ray fluorescence spectrometer (XRF, Rigaku Supermini200), and it was confirmed that they were in good agreement with the theoretical values. For some representative samples, quantitative analysis of oxygen was performed with an Oxygen/Nitrogen/Hydrogen Analyser (EMGA-930, HORIBA, Japan). X-ray photoelectron spectroscopy (XPS; ULVAC PHI5000 Versa Probe II) was performed at room temperature with Al-Kα radiation (1486.6 eV) without Ar+ ion etching. The effect of charging on the binding energies was corrected with respect to the C 1s peak at 284.6 eV. Particle morphology and size distribution were examined using scanning electron microscopy (SEM; Shimadzu, SS-550). The size distribution and average particle size were estimated by measuring the diameters of 200 particles from the SEM micrographs.

Bi LIII XANES (X-ray absorption near-edge structure) spectra of the samples were measured in the transmission mode using the BL-11 beam line at SAGA Light Source (SAGA-LS). The synchrotron radiation was monochromized with a Si (111) double-crystal monochromator. The I0 and I ionization chambers were filled with N2 85%–Ar 15% mixture gas and pure Ar gas, respectively. XANES spectrum of α-Bi2O3 was also measured as a standard material.

Optical reflectance spectra of the samples were measured using a UV-vis spectrometer (Shimadzu UV-2600) with barium sulfate as a reference. The bandgap energies of the samples were determined from the absorption edge of the absorbance spectra represented by the Kubelka–Munk function, f(R) = (1 − R)2/2R, where R is reflectance.21,22 The colour properties of the samples were evaluated in terms of CIE L*a*b*CH° system with a chromometer (Konica-Minolta CR-300). In this system, the parameter L* represents the brightness or darkness of a colour relative to a neutral grey scale, while the parameters a* (the red-green axis) and b* (the yellow-blue axis) qualitatively express the colour. Chroma parameter (C) represents the colour saturation of the pigments and is calculated according to the following formula: C = [(a*)2 + (b*)2]1/2. The parameter H° ranges from 0 to 360°, and is calculated with the formula, H° = tan−1(b*/a*) (H° = 0, for pure red). To evaluate the humidity resistance of the samples, a representative sample was heated at 100 °C for 48 h in a flow of water vapour saturated air atmosphere at 50 °C, and the colour property was evaluated after this treatment using the chromometer.

Results and discussion

Bi4V2O11

Fig. 1 shows the XRD patterns of the Bi4V2O11 samples, which were synthesized in a flow of air and pure O2. Hereafter, the former and the latter samples are denoted as Bi4V2O11 (air) and Bi4V2O11 (O2), respectively. The orthorhombic Bi4V2O11 structure was obtained in a single-phase form for both samples. A weak peak at around 2θ = 24° was assigned to the superlattice structure of the orthorhombic Bi4V2O11.16 The lattice parameters (a, b, c, and V) calculated from the diffraction angle in the XRD patterns are listed in Table 1. The lattice parameters (a, b, c, and V) of Bi4V2O11 (O2) were smaller than those of Bi4V2O11 (air). These results suggest that pentavalent Bi5+ ions were generated in the lattice of Bi4V2O11 (O2) by oxidization of trivalent Bi3+ ions, because the ionic radius of Bi5+ (0.076 nm)20 is smaller than that of Bi3+ (0.103 nm).20
image file: c5ra04734j-f1.tif
Fig. 1 XRD patterns of the Bi4V2O11 samples.
Table 1 The orthorhombic lattice parameters (a, b, c, and V) of the samples
Samples a/nm b/nm c/nm V/nm3
Bi4V2O11 (O2) 0.5604 1.5257 0.5519 0.4719
Bi4V2O11 (air) 0.5606 1.5263 0.5528 0.4729


The valence states of the chemical species on near the surface of the two samples were identified by XPS analysis. Fig. 2 illustrates the XPS of the Bi 4f, V 2p, and O 1s core-levels for the samples. The peak shape and binding energies were obviously depended on the calcination atmosphere. In the case of the Bi 4f spectrum of Bi4V2O11 (air), the 4f5/2 and 4f7/2 peaks located at 164.1 eV and 158.8 eV can be assigned to those of trivalent Bi3+.23,24 For Bi4V2O11 (O2), however, the additional shoulder peaks were observed at 165.8 eV and 160.5 eV corresponding to those of pentavalent Bi5+.24


image file: c5ra04734j-f2.tif
Fig. 2 XPS of the Bi 4f (a), V 2p and O 1s (b) core-levels for the samples.

On the other hand, in the cases of V 2p and O 1s for Bi4V2O11 (air), the 2p1/2 and 2p3/2 peaks located at 524.1 eV and 516.4 eV can be attributed to those of pentavalent V5+,25 and the 1s peak at 529.7 eV can be assigned to the lattice oxygen.25 In the spectrum of Bi4V2O11 (O2), however, some additional peaks appeared: a shoulder peak located at 531.3 eV (O 1s′) can be attributed to an excess amount of oxygen species introduced into the lattice,25 and a shoulder peak located at 517.9 eV (V5+ (2p3/2)′) is considered to be attributable to the excessive oxygen species.

From the above results, it is revealed that partial oxidization of Bi3+ to Bi5+ occurred, due to the calcination of Bi4V2O11 in a flow of pure O2. As a result, oxide anions were introduced by the charge compensation mechanism into the oxide anion deficiencies, which originally exist in the Bi4V2O11 lattice. Accordingly, it is precise to express Bi4V2O11 (air) and Bi4V2O11 (O2) as Bi4V2O11 and Bi4V2O11+δ, respectively.

Fig. 3 depicts the UV-vis diffuse reflectance spectra for Bi4V2O11 and Bi4V2O11+δ. Between the two samples, Bi4V2O11+δ reflected the red light (605–780 nm) more effectively. Table 2 summarizes the CIE L*a*b* colour coordinate data and bandgap energies (Eg) for the samples. The a* value corresponding to red chromaticity for Bi4V2O11+δ (a* = +37.1) was by about 10 points higher than that of Bi4V2O11 (a* = +27.3).


image file: c5ra04734j-f3.tif
Fig. 3 UV-vis diffuse reflectance spectra for Bi4V2O11 and Bi4V2O11+δ.
Table 2 CIE L*a*b* colour coordinate data and bandgap energies (Eg) for the samples
Samples L* a* b* Eg/eV
Bi4V2O11+δ 47.3 +37.1 +36.6 2.20
Bi4V2O11 43.5 +27.3 +29.3 2.15


As already mentioned in the XPS analysis, pentavalent Bi5+ ions were partially generated in the Bi4V2O11+δ sample, and the amount of the oxide anion deficiencies was decreased. Consequently, the impurity energy levels originated in these oxide anion deficiencies in the band structure were partially eliminated from the Bi4V2O11+δ sample. Therefore, the reflection in the red light increased for Bi4V2O11+δ, leading to the significant increase in the redness (a*), although the bandgap energy of this sample was slightly larger than that of Bi4V2O11, as seen in Table 2. Based on these results, all samples were synthesized in a flow of pure O2 in the following sections.

(Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.15)

XRD patterns of the (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.15) samples are shown in Fig. 4. For the (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.08) samples, the orthorhombic Bi4V2O11 structure was obtained in a single phase form, and there were no diffraction peaks of impurity phases. On the other hand, a monoclinic BiVO4 structure was observed as a secondary phase for the (Bi1−xZrx)4V2O11+δ1 (0.10 ≤ x ≤ 0.15) samples. Table 3 lists the orthorhombic Bi4V2O11 lattice parameters (a, b, c, and V) for the (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.15) samples. The lattice parameters (a, b, c, and V) linearly decreased with increasing Zr4+ content, in the single-phase region (0 ≤ x ≤ 0.08), because the ionic radius of Zr4+ (0.072 nm)20 is smaller than those of Bi5+ (0.076 nm)20 and Bi3+ (0.103 nm).20 However, the lattice parameters (a, b, c, and V) became approximately constant in the two-phase region (0.10 ≤ x ≤ 0.15). These results indicated the formation of solid solutions with a single-phase orthorhombic Bi4V2O11 structure, and the solubility limit of Zr4+ ions was about 8 mol% in the (Bi1−xZrx)4V2O11+δ1 system.
image file: c5ra04734j-f4.tif
Fig. 4 XRD patterns of the (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.15) samples.
Table 3 The orthorhombic lattice parameters (a, b, c, and V) of the samples
Samples a/nm b/nm c/nm V/nm3
Bi4V2O11+δ 0.5604 1.5257 0.5519 0.4719
(Bi0.95Zr0.05)4V2O11+δ1a 0.5585 1.5244 0.5510 0.4690
(Bi0.92Zr0.08)4V2O11+δ1b 0.5580 1.5240 0.5508 0.4686
(Bi0.90Zr0.10)4V2O11+δ1c 0.5581 1.5238 0.5506 0.4684
(Bi0.85Zr0.15)4V2O11+δ1d 0.5582 1.5239 0.5507 0.4685


Fig. 5 shows the UV-vis diffuse reflectance spectra for the (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.15) samples. To realize a red pigment, it is important to satisfy both high reflection in the red light region (605–780 nm) and effective absorption (i.e., low reflection) in the green light region (490–560 nm), because green and red are the complementary colours. For all samples, strong reflection in the red light region (605–780 nm) and effective absorption in the green light region (490–560 nm) were observed, and, therefore, these samples are red. Among the single-phase samples of (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.08), (Bi0.92Zr0.08)4V2O11+δ1b reflected the red light most effectively. In the two-phase region (0.10 ≤ x ≤ 0.15), the green light absorption decreased, although further increase in the red light reflection was observed.


image file: c5ra04734j-f5.tif
Fig. 5 UV-vis diffuse reflectance spectra of the (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.15) samples.

The CIE L*a*b* colour coordinate data and bandgap energies (Eg) for the (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.15) samples are listed in Table 4. The a* value, which corresponds to red chromaticity in the positive direction, increased with increasing the content of Zr4+ ions in the single phase region (0 ≤ x ≤ 0.08), and, the (Bi0.92Zr0.08)4V2O11+δ1b sample exhibited the highest a* value at +40.5. However, the redness (a*) decreased in the two-phase region (0.10 ≤ x ≤ 0.15), due to the formation of the impurity monoclinic BiVO4 phase, which exhibits yellow hue.17–19

Table 4 The CIE L*a*b* colour coordinate data and bandgap energies (Eg) of the (Bi1−xZrx)4V2O11+δ1 (0 ≤ x ≤ 0.15) samples
Samples L* a* b* Eg/eV
Bi4V2O11+δ 47.3 +37.1 +36.6 2.20
(Bi0.95Zr0.05)4V2O11+δ1a 49.4 +39.7 +40.6 2.19
(Bi0.92Zr0.08)4V2O11+δ1b 49.6 +40.5 +41.1 2.18
(Bi0.90Zr0.10)4V2O11+δ1c 50.1 +40.4 +42.5 2.19
(Bi0.85Zr0.15)4V2O11+δ1d 52.1 +38.9 +44.5 2.33


(Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.10)

Fig. 6 depicts the XRD patterns of the (Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.10) samples. For the (Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.05) samples, the orthorhombic Bi4V2O11 phase was observed and there were no extra lines due to other compounds or impurities. In contrast, for the samples in which the Al3+ content was more than 5 mol%, the monoclinic BiVO4 structure was observed as an impurity phase, in addition to the main orthorhombic Bi4V2O11 phase. The orthorhombic Bi4V2O11 lattice parameters (a, b, c, and V) for the (Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.10) samples are tabulated in Table 5. The lattice parameters (a, b, c, and V) proportionally decreased with increasing the amount of Al3+ in the single-phase region (0 ≤ y ≤ 0.05), because the ionic radius of Al3+ (0.0535 nm)20 is smaller than those of Bi5+ (0.076 nm)20 and Bi3+ (0.103 nm).20 However, the lattice parameters (a, b, c, and V) became nearly constant in the two-phase region (0.07 ≤ y ≤ 0.10). From these results, it is elucidated that solid solutions of the orthorhombic Bi4V2O11 structure were successfully formed for the (Bi1−yAly)4V2O11+δ3 (0.03 ≤ y ≤ 0.05) samples, and the solubility limit of Al3+ ions was around 5 mol% in (Bi1−yAly)4V2O11+δ3. Furthermore, it is considered that the Al3+ ions were doped not into the trivalent Bi3+ site, but dominantly into the pentavalent Bi5+ site, because the ionic radius of Al3+ (0.0535 nm)20 is closer to that of Bi5+ (0.076 nm)20 than that of Bi3+ (0.103 nm).20
image file: c5ra04734j-f6.tif
Fig. 6 XRD patterns of the (Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.10) samples.
Table 5 The orthorhombic lattice parameters (a, b, c, and V) of the samples
Samples a/nm b/nm c/nm V/nm3
Bi4V2O11+δ 0.5604 1.5257 0.5519 0.4719
(Bi0.97Al0.03)4V2O11+δ3a 0.5586 1.5229 0.5507 0.4681
(Bi0.95Al0.05)4V2O11+δ3b 0.5582 1.5217 0.5505 0.4678
(Bi0.93Al0.07)4V2O11+δ3c 0.5583 1.5219 0.5503 0.4679
(Bi0.90Al0.10)4V2O11+δ3d 0.5584 1.5214 0.5504 0.4677


Fig. 7 illustrates the UV-vis diffuse reflectance spectra of the (Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.10) samples. For (Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.05), which were obtained in the single orthorhombic Bi4V2O11 phase, the absorption in the green light region (490–560 nm) increased with increasing the content of Al3+. Among them, (Bi0.97Al0.03)4V2O11+δ3a reflected the red light (605–780 nm) most effectively, although the absorption in the green light region (490–560 nm) of this sample was slightly weaker than that of (Bi0.95Al0.05)4V2O11+δ3b.


image file: c5ra04734j-f7.tif
Fig. 7 UV-vis diffuse reflectance spectra of the (Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.10) samples.

The CIE L*a*b* colour coordinate data and bandgap energies (Eg) for the (Bi1−yAly)4V2O11+δ3 (0 ≤ y ≤ 0.10) samples are listed in Table 6. The a* value corresponding to the red chromaticity of these pigments depended on their compositions, and the highest redness was obtained for (Bi0.97Al0.03)4V2O11+δ3a (a* = +41.6), as in the UV-vis diffuse reflectance measurement.

Table 6 The CIE L*a*b* colour coordinate data and bandgap energies (Eg) of the samples
Samples L* a* b* Eg/eV
Bi4V2O11+δ 47.3 +37.1 +36.6 2.20
(Bi0.97Al0.03)4V2O11+δ3a 47.7 +41.6 +38.0 2.17
(Bi0.95Al0.05)4V2O11+δ3b 47.8 +40.6 +36.8 2.19
(Bi0.93Al0.07)4V2O11+δ3c 52.5 +40.2 +38.8 2.20
(Bi0.90Al0.10)4V2O11+δ3d 53.0 +36.9 +43.4 2.21


(Bi1−xyZrxAly)4V2O11+δ2 (0.01 ≤ x ≤ 0.07; 0.01 ≤ y ≤ 0.07; x + y = 0.08)

As mentioned in the previous sections, dissolution of Zr4+ or Al3+ into the Bi4V2O11+δ lattice is effective to increase the redness of the resulting pigments. In this section, therefore, Zr4+ and Al3+ were co-doped into the Bi4V2O11+δ lattice and the composition was optimized. In this regard, the total amount of Zr4+ and Al3+ was kept constant at 8 mol%, because an excess amount of the dopants had negative effect on the red hue of the pigments by the formation of yellowish monoclinic BiVO4 phase as an impurity.

XRD patterns of the (Bi1−xyZrxAly)4V2O11+δ2 (0.01 ≤ x ≤ 0.07; 0.01 ≤ y ≤ 0.07; x + y = 0.08) samples are shown in Fig. 8. The orthorhombic Bi4V2O11 structure was obtained in a single-phase form for all samples. As shown in Table 7, the lattice parameters (a, b, c, and V) of the (Bi1−xyZrxAly)4V2O11+δ2 (0.01 ≤ x ≤ 0.07; 0.01 ≤ y ≤ 0.07; x + y = 0.08) samples became smaller than those of the non-doped Bi4V2O11+δ sample. These results indicate the formation of solid solutions with the orthorhombic Bi4V2O11 structure for all samples.


image file: c5ra04734j-f8.tif
Fig. 8 XRD patterns of the (Bi1−xyZrxAly)4V2O11+δ2 (0.01 ≤ x ≤ 0.07; 0.01 ≤ y ≤ 0.07; x + y = 0.08) samples.
Table 7 The orthorhombic lattice parameters (a, b, c, and V) of the samples
Samples a/nm b/nm c/nm V/nm3
Bi4V2O11+δ 0.5604 1.5257 0.5519 0.4719
(Bi0.92Zr0.07Al0.01)4V2O11+δ2a 0.5587 1.5227 0.5506 0.4684
(Bi0.92Zr0.04Al0.04)4V2O11+δ2b 0.5586 1.5222 0.5505 0.4681
(Bi0.92Zr0.01Al0.07)4V2O11+δ2c 0.5585 1.5220 0.5503 0.4678


Fig. 9 depicts the UV-vis diffuse reflectance spectra of the (Bi1−xyZrxAly)4V2O11+δ2 (0.01 ≤ x ≤ 0.07; 0.01 ≤ y ≤ 0.07; x + y = 0.08) samples. Both the reflection in the red light (605–780 nm) and the absorption in the green light (490–560 nm) were dependent on their compositions. Among these pigments, the (Bi0.92Zr0.07Al0.01)4V2O11+δ2a pigment reflected the red light (605–780 nm) most effectively.


image file: c5ra04734j-f9.tif
Fig. 9 UV-vis diffuse reflectance spectra of the (Bi1−xyZrxAly)4V2O11+δ2 (0.01 ≤ x ≤ 0.07; 0.01 ≤ y ≤ 0.07; x + y = 0.08) samples.

Table 8 summarizes the CIE L*a*b*CH° colour coordinate data and bandgap energies (Eg) for the (Bi1−xyZrxAly)4V2O11+δ2 (0.01 ≤ x ≤ 0.07; 0.01 ≤ y ≤ 0.07; x + y = 0.08) samples. The a* value (red chromaticity) depended on their compositions, and the highest red hue was obtained for (Bi0.92Zr0.07Al0.01)4V2O11+δ2a and (Bi0.92Zr0.04Al0.04)4V2O11+δ2b, where a* = +41.9 for both samples. Furthermore, the hue angle (H°) corresponding to colour purity, for (Bi0.92Zr0.07Al0.01)4V2O11+δ2a (H° = 39.0) was smaller than that for (Bi0.92Zr0.04Al0.04)4V2O11+δ2b (H° = 40.0). Therefore, (Bi0.92Zr0.07Al0.01)4V2O11+δ2a produced the deepest red hue among these pigments.

Table 8 The CIE L*a*b*CH° colour coordinate data and bandgap energies (Eg) of the samples
Samples L* a* b* C H° Eg/eV
Bi4V2O11+δ 47.3 +37.1 +36.6 52.1 44.6 2.20
(Bi0.92Zr0.07Al0.01)4V2O11+δ2a 49.6 +41.9 +34.0 53.9 39.0 2.18
(Bi0.92Zr0.04Al0.04)4V2O11+δ2b 49.7 +41.9 +35.2 54.7 40.0 2.18
(Bi0.92Zr0.01Al0.07)4V2O11+δ2c 51.2 +38.9 +43.2 58.1 48.0 2.19


To investigate chemical states of bismuth species in the bulk for the samples, Bi LIII XANES spectra of non-doped Bi4V2O11+δ, (Bi0.92Zr0.08)4V2O11+δ1b, and (Bi0.97Al0.03)4V2O11+δ3a were measured as shown in Fig. 10. Bi LIII XANES spectra of α-Bi2O3 were also collected as the standard of trivalent Bi3+. In the spectra of non-doped Bi4V2O11+δ, (Bi0.92Zr0.08)4V2O11+δ1b, and (Bi0.97Al0.03)4V2O11+δ3a, a peak assigned to pentavalent Bi5+ ions was observed at 13[thin space (1/6-em)]440 eV.26 On the contrary, this peak was absent in the spectrum of α-Bi2O3, in which bismuth ions exist only in trivalent state. In addition, the absorption edge shifted to the higher energy side for these pigments, compared to that of α-Bi2O3. These results elucidate the existence of pentavalent Bi5+ ions in the bulk of non-doped Bi4V2O11+δ, (Bi0.92Zr0.08)4V2O11+δ1b, and (Bi0.97Al0.03)4V2O11+δ3a.


image file: c5ra04734j-f10.tif
Fig. 10 Bi LIII XANES spectra of non-doped Bi4V2O11+δ, (Bi0.92Zr0.08)4V2O11+δ1b, and (Bi0.97Al0.03)4V2O11+δ3a.

Fig. 11 illustrates XPS of (Bi0.92Zr0.08)4V2O11+δ1b, (Bi0.97Al0.03)4V2O11+δ3a, and (Bi0.92Zr0.07Al0.01)4V2O11+δ2a. The spectrum of non-doped Bi4V2O11+δ is also depicted as a reference. The proportion of pentavalent Bi5+ in total bismuth ions (Bi5+/(Bi5+ + Bi3+)) for these pigments was estimated from the area of corresponding peak in the spectra, and the results are tabulated in Table 9. The proportion of Bi5+ was 35% for (Bi0.92Zr0.08)4V2O11+δ1b, while it was 26% for non-doped Bi4V2O11+δ. This result suggests that the amount of Bi5+ was increased by the Zr4+ doping into the bismuth sites of the Bi4V2O11+δ lattice, and thereby, the oxide anion deficiencies were decreased in (Bi0.92Zr0.08)4V2O11+δ1b, because of the reason already mentioned above. In the case of (Bi0.97Al0.03)4V2O11+δ3a, the proportion of Bi5+ was 21%, and this value was smaller than that for non-doped Bi4V2O11+δ (26%). The binding energies of Bi 4f7/2 (159.6 eV) and Bi 4f5/2 (164.9 eV) attributable to those of pentavalent Bi5+ were observed at lower energies than those of non-doped Bi4V2O11+δ: Bi 4f7/2 (160.5 eV) and Bi 4f5/2 (165.8 eV). These results indicate that significant lattice shrinkage was occurred in the (Bi0.97Al0.03)4V2O11+δ3a lattice to induce significant electronic repulsion in the 4f orbital, even the 4f electrons are well shielded from the surroundings by the outer 5d and 6s orbitals. As a result, the conduction band composed of the V3d orbital was also enlarged by the lattice shrinking to increase the crystal field splitting. In fact, the bandgap energy for this pigment was the smallest among the three samples. Furthermore, in the case of the (Bi0.92Zr0.07Al0.01)4V2O11+δ2a pigment, both of the increase in the proportion of Bi5+ and the decrease in the binding energies (Bi 4f7/2 and Bi 4f5/2) were confirmed.


image file: c5ra04734j-f11.tif
Fig. 11 XPS of (Bi0.92Zr0.08)4V2O11+δ1b, (Bi0.97Al0.03)4V2O11+δ3a, and (Bi0.92Zr0.07Al0.01)4V2O11+δ2a.
Table 9 The proportion of pentavalent Bi5+ in total bismuth ions (Bi5+/(Bi5+ + Bi3+)) for the samples
Samples Proportion of Bi5+/%
Bi4V2O11+δ 26
(Bi0.92Zr0.08)4V2O11+δ1b 35
(Bi0.97Al0.03)4V2O11+δ3a 21
(Bi0.92Zr0.07Al0.01)4V2O11+δ2a 30


To determine the oxygen content in the representative samples, quantitative analysis of oxygen was performed for Bi4V2O11, Bi4V2O11+δ, (Bi0.92Zr0.08)4V2O11+δ1b, (Bi0.97Al0.03)4V2O11+δ3a, and (Bi0.92Zr0.07Al0.01)4V2O11+δ2a samples, and the results are listed in Table 10. The oxygen amounts in the samples calcined in a flow of pure O2 were higher than that of Bi4V2O11 calcined in a flow of air. Taking account of the results in Fig. 11 and Table 9 into consideration, it was found that the oxide ion content monotonically increased with the proportion of Bi5+.

Table 10 The analysed compositions of the representative samples
Samples Analysed compositions
Bi4V2O11 Bi4V2O11.00
Bi4V2O11+δ Bi4V2O11.17
(Bi0.92Zr0.08)4V2O11+δ1b (Bi0.92Zr0.08)4V2O11.41
(Bi0.97Al0.03)4V2O11+δ3a (Bi0.97Al0.03)4V2O11.14
(Bi0.92Zr0.07Al0.01)4V2O11+δ2a (Bi0.92Zr0.07Al0.01)4V2O11.34


Fig. 12 shows SEM images and size distributions of the Bi4V2O11.17, (Bi0.92Zr0.08)4V2O11.41, (Bi0.97Al0.03)4V2O11.14, and (Bi0.92Zr0.07Al0.01)4V2O11.34 samples. Particles with indefinite shapes were observed for all samples, and the mean particle sizes of the Bi4V2O11.17, (Bi0.92Zr0.08)4V2O11.41, (Bi0.97Al0.03)4V2O11.14, and (Bi0.92Zr0.07Al0.01)4V2O11.34 samples were 7.52, 7.20, 7.58, and 7.29 µm, respectively. There is no evident change in the mean particle sizes, size distributions, and morphologies of these samples. These results indicate that the change in the colour properties of these samples was caused by the doping elements, and the contribution from the particle size and morphology can be eliminated.


image file: c5ra04734j-f12.tif
Fig. 12 SEM images and size distributions of the (a) Bi4V2O11.17, (b) (Bi0.92Zr0.08)4V2O11.41, (c) (Bi0.97Al0.03)4V2O11.14, and (d) (Bi0.92Zr0.07Al0.01)4V2O11.34 pigments.

From the results mentioned above, the doping effect of Zr4+ and Al3+ on the red hue of the pigments can be concluded as follows.

Doping effect of Zr4+. Zr4+ doping into the bismuth sites in the Bi4V2O11.17 image file: c5ra04734j-t1.tif lattice positively affects the colour property. The oxide anion deficiencies were occupied by the charge compensation mechanism, leading to the elimination of the impurity energy levels between the valence and the conduction bands in the band structure. However, the introduction of an excess amount of Zr4+ ions induces the formation of yellowish monoclinic BiVO4 as an impurity phase to decrease the red hue.
Doping effect of Al3+. Dissolution of Al3+ reduces the bandgap energy of Bi4V2O11.17 by broadening the width of V3d conduction band, although it increases the amount of the oxide anion deficiencies (oxide anion vacancies). Accordingly, both positive and negative effects are produced simultaneously. It is considered that the positive effect is dominant when a small amount of Al3+ is doped into the lattice, but adversely the negative effect will become prominent when the amount of Al3+ is excessive.

Consequently, the synergetic effects were observed when both Zr4+ and Al3+ were doped into the Bi4V2O11.17 lattice. As a result, the band structure was optimized at the composition of

image file: c5ra04734j-t2.tif

Table 11 summarizes the CIE L*a*b*CH° colour coordinate data and bandgap energies (Eg) of the (Bi0.92Zr0.07Al0.01)4V2O11.34 pigment, the commercial cadmium selenide red (Holbein works, Ltd, PG002), commercial mercuric sulfide red (Holbein works, Ltd, PG009), and commercial iron oxide (Morishita Bengara Kogyo Co. Ltd, MR-320A) pigments. Although the redness value (a*) and the colour saturation parameter (C) for the (Bi0.92Zr0.07Al0.01)4V2O11.34 pigment (a* = +41.9, C = 53.9) fell short of those for the toxic cadmium selenide red (a* = +63.7, C = 84.6) and mercuric sulfide red (a* = +56.5, C = 69.5) pigments, they were significantly higher than those for the environmental friendly iron oxide pigment (a* = +28.9, C = 38.4). In addition, the brightness value (L*) of this pigment (L* = 49.6) was also larger than that of the iron oxide pigment (L* = 38.9), thus, the (Bi0.92Zr0.07Al0.01)4V2O11.34 pigment produced a light red hue.

Table 11 The CIE L*a*b*CH° colour coordinate data and bandgap energies (Eg) of the pigments
Pigments L* a* b* C H° Eg/eV
(Bi0.92Zr0.07Al0.01)4V2O11.34 49.6 +41.9 +34.0 53.9 39.0 2.18
Commercial cadmium selenide red (CdS·CdSe) 51.9 +63.7 +55.8 84.6 41.2 2.11
Commercial mercuric sulfide red (HgS) 52.0 +56.5 +40.5 69.5 35.6 2.13
Commercial iron oxide (Fe2O3) 38.9 +28.9 +25.3 38.4 41.2 2.21


Furthermore, the (Bi0.92Zr0.07Al0.01)4V2O11.34 sample was heated at 100 °C for 48 h in a flow of the moist air, and the colour coordinate data were measured after this treatment to evaluate the humidity resistance. The colour coordinate data were measured and the colour difference (ΔE) was calculated from the data before and after the test using the following equation:

ΔE = [(L*afterL*before)2 + (a*aftera*before)2 + (b*afterb*before)2]1/2

A small ΔE indicates that the colour of the pigment is stable. From the result listed in Table 12, it is evident that the present (Bi0.92Zr0.07Al0.01)4V2O11.34 pigment possesses high humidity resistance.

Table 12 The CIE L*a*b* colour difference (ΔE) of the (Bi0.92Zr0.07Al0.01)4V2O11.34 samples before and after the humidity resistance test
Samples L* a* b* ΔE
Before the test 49.6 +41.9 +34.0
After the test 49.9 +41.4 +37.6 3.6


Conclusions

Novel environmental friendly red pigments, (Bi1−xyZrxAly)4V2O11+δ (0 ≤ x ≤ 0.15; 0 ≤ y ≤ 0.10), were successfully synthesized by the conventional solid state reaction method. The colour of these pigments depended on the calcination condition and the composition, and the most vivid red hue was obtained for the (Bi0.92Zr0.07Al0.01)4V2O11.34 pigment calcined at 800 °C for 10 h in a flow of pure O2, which has CIE L*a*b* colour parameters of L* = 49.6, a* = +41.9, and b* = +34.0. The a* value, which corresponds to the redness, was significantly higher than that of a commercial iron oxide pigment (a* = +28.9). Since Bi4V2O11 is considered to be a nontoxic compound, and Zr and Al are also safe elements, the present pigment should be an effective alternative for the conventional iron oxide red pigment.

Acknowledgements

The authors wish to thank Mr Akihiro Hirano and Mr Satoshi Yoshida, HORIBA, Ltd for the oxygen/nitrogen/hydrogen analysis. This work was supported by the Development of Alternative Technology for Hazardous Chemical Substances and Development of Novel Environment- and Human-friendly Inorganic Pigments for Three Primary Colors (FY2010-2014) programs of the New Energy and Industrial Technology Development Organization (NEDO) and the Ministry of Economy, Trade and Industry, Japan (METI).

Notes and references

  1. E. B. Faulkner and R. J. Schwartz, High Performance Pigments, Wiley-VCH, Weinheim, 2nd edn, 2009 Search PubMed.
  2. L. S. Kumari, P. P. Rao and P. Koshy, J. Am. Ceram. Soc., 2010, 93, 1402 CAS.
  3. R. A. Candeia, M. I. B. Bernardi, E. Longo, I. M. G. Santos and A. G. Souza, Mater. Lett., 2004, 58, 569 CrossRef CAS.
  4. P. Šulcová, L. Vitásková and M. Trojan, J. Therm. Anal. Calorim., 2010, 99, 409 CrossRef.
  5. M. Llusar, L. Vitásková, P. Šulcová, M. A. Tena, J. A. Badenes and G. Monrós, J. Eur. Ceram. Soc., 2010, 30, 37 CrossRef CAS PubMed.
  6. S. T. Aruna, S. Ghosh and K. C. Patil, Int. J. Inorg. Mater., 2001, 3, 387 CrossRef CAS.
  7. G. George, V. S. Vishnu and M. L. P. Reddy, Dyes Pigm., 2011, 88, 109 CrossRef CAS PubMed.
  8. V. S. Vishnu, G. George and M. L. P. Reddy, Dyes Pigm., 2010, 85, 117 CrossRef CAS PubMed.
  9. V. S. Vishnu and M. L. P. Reddy, Sol. Energy Mater. Sol. Cells, 2011, 95, 2685 CrossRef CAS PubMed.
  10. V. James, P. P. Rao, S. Sameera and S. Divya, Ceram. Int., 2014, 40, 2229 CrossRef CAS PubMed.
  11. M. Jansen and H. P. Letschert, Nature, 2000, 404, 980 CrossRef CAS PubMed.
  12. K. A. Winship, Adverse Drug React. Acute Poisoning Rev., 1983, 2, 103 Search PubMed.
  13. Wendusu, T. Masui and N. Imanaka, Chem. Lett., 2012, 41, 1616 CrossRef CAS.
  14. Wendusu, T. Masui and N. Imanaka, J. Asian Ceram. Soc., 2014, 2, 195 CrossRef PubMed.
  15. S. J. Patwe, A. Patra, R. Dey, A. Roy, R. M. Kadam, S. N. Achary and A. K. Tyagi, J. Am. Ceram. Soc., 2013, 96(11), 3448 CrossRef CAS PubMed.
  16. K. Sooryanarayana, T. N. Guru Row and K. B. R. Varma, Mater. Res. Bull., 1997, 32, 1651 CrossRef CAS.
  17. Wendusu, K. Ikawa, T. Masui and N. Imanaka, Chem. Lett., 2011, 40, 792 CrossRef CAS.
  18. T. Masui, T. Honda, Wendusu and N. Imanaka, Dyes Pigm., 2013, 99, 636 CrossRef CAS PubMed.
  19. Wendusu, T. Honda, T. Masui and N. Imanaka, RSC Adv., 2013, 3, 24941 RSC.
  20. R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751 CrossRef.
  21. D. R. Eppler and R. A. Eppler, Ceram. Eng. Sci. Proc., 1996, 17, 77 CAS.
  22. M. Kato and M. Takahashi, J. Mater. Sci. Lett., 2001, 20, 413 CrossRef CAS.
  23. S. Poulston, N. J. Price, C. Weeks, M. D. Allen, P. Parlett, M. Steinberg and M. Bowker, J. Catal., 1998, 178, 658 CrossRef CAS.
  24. H. Fan, G. Wang and L. Hu, Solid State Sci., 2009, 11, 2065 CrossRef CAS PubMed.
  25. J. Sun, X. Li, Q. Zhao, J. Ke and D. Zhang, J. Phys. Chem. C, 2014, 118, 10113 CAS.
  26. A. Demourgues, C. Dussarrat, R. Bontchev, B. Darriet, F. Weill and J. Darriet, Nucl. Instrum. Methods Phys. Res., 1995, 97, 82 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.