Abstract

Luminescent yttrium phosphate is engineered into an environmentally benign near infrared (NIR) reflective yellow pigment by the substitution of bismuth and vanadium metals in the host lattice. A series of YP(1–x)VxO4 (x = 0, 0.05, 0.1, 0.15, 0.2, and 0.4), Y(1–y)BiyPO4 (y = 0.1 and 0.3), and Y(1–y)BiyP(1–x)VxO4 (x = y = 0.2, 0.4, and 0.6) were prepared by the precipitation method. Secondary phase was noticed at x = 0.2 and y = 0.2 while substituting vanadium and bismuth, respectively, due to high ionic radii of the dopant ions. Co substitution of vanadium and bismuth in the YPO4 lattice enhanced both NIR reflectance and yellow color of all the fabricated materials. XPS spectra proved the presence of trivalent bismuth and pentavalent vanadium in Y0.4Bi0.6P0.4V0.6O4. Due to the substitution effect, a more defined morphology was noticed, which enhanced the scattering co-efficient of the fabricated materials; hence, the NIR reflectance of the materials was increased from 68% (YPO4) to 83% (Y0.4Bi0.6P0.4V0.6O4). Chemical and thermal stability test of Y0.4Bi0.6P0.4V0.6O4 confirmed the color and strength of the designed pigment. With good yellow hue (b* = +56.06), high NIR solar reflectance (R* = 83%), and good stability, Y0.4Bi0.6P0.4V0.6O4 can act as an environmentally benign cool yellow pigment.
1. Introduction
The urban heat island (UHI) effect is one of the most serious discomforts experienced by the people who live in cities. Due to this effect, the interior of the built up area experiences more heat than its external surroundings. The UHI effect is generally caused due to the absorption of solar radiation by the buildings and other architectures in the urban areas during the long day time, and no dissipation of the absorbed energy in the night time.1 This phenomenon significantly increases the difference in temperature between urban and rural areas (the radiation is more rapid due to natural greeneries). In order to reduce the high interior temperature, use of air conditioners are preferred which consumes around 40% of electricity.2 Heat absorption is quite high when the buildings are made of asphalt sheets, cement, concretes, and metallic structures.3 Since approximately 53% of the solar radiation consists of the near-infrared region (NIR), it is obvious that building infrastructures coated with NIR-reflective materials could effectively reduce the interior temperature, in turn keeping the building cool. Hence, these NIR-reflective materials are coined as cool coating materials.4
Neoteric investigators are focused on the synthesis of NIR reflective materials based on rare earth materials as host/dopant compound; hence one could increase the reflectivity of the material while maintaining the aesthetic color.5 Sky blue color of SrCuSi4O10 was modified into dark blue by doping the Eu3+ ion into Sr2+ and the NIR reflectance of the fabricated materials increased significantly by the substitution of 20% dopant in the host system.6 (Y1–xRx)2Cu2O5 (R = Gd3+, Sc3+, Yb3+ Lu3+ and Tm3+) green pigments7 were synthesized by the solid state reaction method and proved to be environmentally benevolent pigments compared to toxic Cr and Co green pigments. Pale yellow-colored Y6Mo6O12 [Reflectance (R*) = 92%] is converted into dark yellow-colored SixY6–xMo6O12 with R* = 98% and dark brown color PrxY6–xMo6O12 with R* = 60%.8 Fe3+-substituted La2Mo2O9 nanocrystalline material with brilliant yellow color was developed with 92% reflectance, using the poly acryl amide method.9 Red Li3InB2O610 was prepared using the solid-state reaction method to obstruct the infrared radiations. The incorporation of V5+ (3d0) in the pentavalent or the tetravalent site has been a promising strategy in recent years to obtain a yellow hue from white materials. In particular, vanadium in the pentavalent state (V2O5) is yellow due to the charge transfer transition from O2p (II) to V3d (V).11 V-ZrO2 is a well-known inorganic yellow pigment prepared by various conventional and non-conventional methods, evidencing the incorporation of V into ZrO2 that induces the yellow hue in the resultant pigment material.12 Gopal et al. reported the abnormal perception in the optical property of V5+-substituted BPO4. Open framework BPO4 (white) was converted into yellow phosphovanadate by substituting the VO4 site into the PO4 site.13 BiVO4 was also proved as a promising yellow pigment in the industrial sector. Compiling all the preceding results, substitution or doping of bismuth and vanadium ions could be one of the iconic ideas to develop eco-friendly yellow pigments.
Lanthanide-based orthophosphates (LnPO4) are recently materialized as an important class of inorganic substances, whose applications are widespread in various fields such as laser host materials, luminescent and phosphor materials, upconversion of photons, and bio-imaging.14 These unique phosphates exhibit a variety of polymorphic forms, namely, monoclinic monazite, tetragonal xenotime, tetragonal zircon, monoclinic churchite, and hexagonal rhabdophane.15 The formation of the foregoing crystal structures are decided by the synthesis methodology and the nature of the metal cation.16 Some of the important rare earth phosphates that fit into the abovementioned profile are CePO4, PrPO4, YPO4, LuPO4, LaPO4, GaPO4, and EuPO4. The physical, chemical, and optical properties of these materials were engineered by introducing a suitable substituent in the host lattice; hence the desired properties were achieved. Among various rare earth phosphates, yttrium phosphate (YPO4) has versatile applications due to its sturdy nature and high chemical stability.17 It exists in two distinctive crystal structures, namely, hexagonal (D2 symmetry) and tetragonal (D2d symmetry).18 The tetragonal YPO4 coexists with other elements in the mineral called xenotime,19 which belongs to the space group of I41/amd, wherein the Y3+ ion follows D2d symmetry. The structure of YPO4 is formed by chains of corner-shared dodecahedron YO8 and tetrahedron PO4.20
YPO4 is considered as one of the best host material to prepare luminescent compounds, since they acquire high quantum efficiency and because of the existence of explicit absorption and emission spectral regions.21 High absorption and high vacuum stability during the radiation process in the ultraviolet (UV) domain make them as an efficient source of red light in normal and plasma display process.22 Eu3+-doped YPO4 was reported to have the aforementioned quality; hence the possibility of using the material as a red light-emissive source was proposed by Chanchan and Singh.21 Due to its high thermal expansion coefficient, it was explored as green barrier coatings to protect silicon ceramics. Ce3+ and Tb3+ co-doped yttrium phosphate was reported as potent glass ceramic material for WLED application due its adorable luminescent property on doping of a rare earth cation into yttrium lattice.23 Dy3+/YPO4 and Sm3+:YPO4 co-doped with Bi3+ was reported as luminescent material with better emission intensity.24,25
As it is widely accepted that the optical, magnetic, and other physical/chemical properties are greatly influenced by the preparation methodology, it is necessary to carry out a detailed research so as to select a suitable method of synthesis to achieve the desired properties (exclusive size, shape, particle distribution, and crystallinity) on the resultant material.26 Water-borne methods are much preferred over organic solvent-based methods due their non-toxic nature. A lot of research is being undertaken to synthesize micro- and nano-structured YPO4 with desired properties, from the traditional solid state reaction route to versatile wet chemical routes, including precipitation method, hydrothermal method, solvo-thermal methods, and ion-exchange reactions.27,28 Among these methods, the chemical co-precipitation method has attracted much attention due to high production and purity of the desired product, use of less or no organic solvents, good reproducibility of the product, and low production cost. In a typical precipitation method, expected products are precipitated as their hydroxides using appropriate precipitating agents such as ammonium hydroxide, NaOH, and tetra methyl ammonium hydroxide from the metal precursors, following the calcination process. Hence, in the current research, the precipitation method is preferred.29
Though YPO4-based materials are studied for their luminescent property, it was not explored as a NIR-reflective pigment. To prove its application in another interesting field, herein, we investigated the NIR-reflective properties and color properties of YPO4-, bismuth-, and vanadium-substituted YPO4 synthesized by a simple precipitation method and demonstrated its excellence in cool coating applications.
2. Materials and Methods
2.1. Materials
Yttrium oxide (99%), di-ammonium hydrogen phosphate (98%), aqueous ammonia (NH4OH), bismuth nitrate penta hydrate (99%), and ammonium meta vanadate (98%) are used for the synthesis. All the chemicals were purchased from S.D. Fine Chemicals Private Limited, India, and used without further purification.
2.2. Method of Synthesis
YP(1–x)VxO4 (x = 0, 0.05, 0.1, 0.15, 0.2, and 0.4), Y(1–y)BiyPO4 (y = 0.1 and 0.3), and Y(1–x)BixP(1–y)VyO4 (x = y = 0.2, 0.4, and 0.6) were prepared following the precipitation method. Details of the prepared samples and the corresponding sample are given in Table 1. To synthesize yttrium phosphate, 1 M of yttrium oxide was dissolved 1:1 HNO3. To this, 1 M of di-ammonium hydrogen phosphate was added drop by drop, after which the pH of the resulting solution was adjusted to 4. Once the precipitation is completed, it was washed with distilled water, filtered, and dried in an oven at 100 °C. The dried precursor powder was ground well and subjected to calcinations at 300, 500, and 900 °C. To synthesize bismuth-substituted yttrium phosphate, bismuth nitrate solution was added along with yttrium nitrate solution and the abovementioned process was repeated. To synthesize vanadium-substituted materials, yellow ammonium meta vanadate solution which was pre-dissolved using 1:1 HNO3 was added after adding ammonium phosphate solution.
Table 1. Prepared Samples and Its Acronyms.
| acronym | composition |
|---|---|
| YP | YPO4 |
| YV05 | YP0.95V0.05O4 |
| YV1 | YP0.90V0.10O4 |
| YV15 | YP0.85V0.15O4 |
| YV2 | YP0.8V0.2O4 |
| YV4 | YP0.6V0.4O4 |
| YB1 | Y0.9Bi0.1PO4 |
| YB3 | Y0.7Bi0.3PO4 |
| YB2V2 | Y0.9Bi0.2P0.8V0.2O4 |
| YB4V4 | Y0.6Bi0.4P0.6V0.4O4 |
| YB6V6 | Y0.4Bi0.6P0.4V0.6O4 |
2.3. Chemical Stability Test and Thermal Stability Test
Concerning the stability of the pigment toward aggressive chemicals, 100 mg of specific pigment was soaked in 2% HNO3, 2% H2SO4, 2% NaOH, and 2% HCl individually and stirred for an hour. Treated pigments were strained using filter paper, rinsed using distilled water, and air dried, and their weight was re-examined. Thermal stability of the particular pigment was studied by subjecting it to high calcination temperatures from 900 to 1100 °C for 6 h.
2.4. Characterization Techniques
The powder
X-ray diffraction (PXRD) technique was employed to identify the phase
formation of synthesized materials. Structural analysis was accomplished
using the BRUKER D8 Advanced instrument via a Cu K alpha radiation
with 1.54056 Å handling wavelength. The lattice parameter of
the synthesized materials was calculated using powder X software.
AlO4, PO4, and Al–O stretching and bending
vibrations and the functional groups present in the fabricated pigment
components were recognized with the Fourier transform infrared spectroscopy
(FT-IR) technique in the 400 and 4000 cm–1 range
with the JASCO FT-IR instrument. The modification in the morphology
due to substitution and the presence of elements were established
using the ZESIS EVO18 scanning electron microscopy (SEM) instrument.
HR-TEM (FEI-TECHNAI, G2-20 TWIN) with the operating voltage of 200
kV was employed to analyze the detailed morphology of the selected
pigment particle, and the selected area diffraction pattern (SAED)
was also employed with the same instrument. Absorbance and diffuse
reflectance spectra were recorded using the JASCO-V670 UV-DRS spectrophotometer
in the 200 to 2500 nm operating range with the step size of 1 nm.
The color coordinates such as L* = lightness or darkness
of the pigment (0–100), a* value representing
the greenness (−a*) or redness (+a*), and b* values representing the yellowness (+b*) and blueness (−b*) of any synthesized
material were calculated according to the recommendation by Commission
Internationale de l’Eclairage (CIE) using D 65 illuminant through
a 10° viewer angle. The amount of saturation is calculated with
the term Chroma (C), which can be calculated by
. The
hue in the cylindrical color space
extends from 0 to 360°, which could be found by the formula
. NIR reflectance (R*)
of the fabricated materials was found via the following formula
![]() |
At this juncture, the reflectance acquired from the experimental data is noted as r(λ) and the solar irradiance is termed as i(λ) with the unit of (W/m2·nm).
3. Results and Discussion
3.1. TGA/DTA Analysis
Simultaneous TGA/DTA was carried out to identify the thermal properties of YPO4. Two stages of weight losses were observed from thermogravimetric analysis, as shown in Figure 1. According to Figure 1, a total weight loss of 11.5% was observed after thermal treatment from ambient temperature to 1200 °C. Initial weight loss of about 8.7% was noticed from 100 to 400 °C, and further increase in temperature till 800 °C brought a slight weight loss of around 2.8%,30 after which no significant weight loss was observed, which showed that the system is stable above 800 °C due to the loss of lattice water. The DTA curve showed an endothermic peak below 100 °C, which was attributed to the disappearance of water molecules and nitrates.
Figure 1.

TGA/DTA analysis of YPO4.
3.2. PXRD Analysis
The PXRD technique was employed to identify the crystal structure, phase purity, lattice details, and the average crystallite size of the synthesized materials. Figure 2 shows the PXRD pattern of YPO4. Peaks corresponding to 2θ angle observed are well harmonized with the standard structure (JCPDS. no. 74-2429) with the space group of I41/amd,31 and the lattice values are found to be a = b = 6.913, c = 6.174 with the average crystallite size of 34 nm. Table 1 shows the prepared pigments and its acronyms.
Figure 2.
PXRD pattern of YPO4.
3.2.1. Effect of Vanadium Substitution on the Structure of YPO4
Figure 3a shows the diffraction pattern of YP1–xVxO4, x = 0.05, 0.1, 0.15, and 0.2 prepared by the chemical precipitation method. One could affirm the formation of the tetragonal structure after the substitution of V5+ into P5+ lattice with the shift in the diffraction peak to lower theta angle (Figure 3b). The shift is obvious due to the substitution of higher ionic radii (0.36 Å) V5+ with lower ionic radii (0.17 Å) P5+ having the co-ordination number of 4. Due to this substitution, lattice strain in the crystal structure occurs owing to inadequate ionic radii of P5+, which could not accommodate bigger V5+ in its lattice and possibly shifted the high intense peak in the direction of the lower theta angle.32 Further addition of V5+ (x = 0.4) leads to the formation of secondary phase V2O5 in the YPO4 structure (Figure 4a). Comparable finding was discerned by Wujczyk et al.20 In the abovementioned research, a series of luminescent materials were formed by doping Er3+ and Tm3+ co-doped with Yb3+ in YP1–xVxO4 (x = 0–1.0). They concluded that lattice strain and shift in diffraction peaks toward the lower angle is due to V5+ replacement and not due to the other rare earth ions present.
Figure 3.
(a) YP1–xVxO4, x = 0.05, 0.1, 0.15, and 0.2. (b) Expanded view.
Figure 4.
PXRD pattern of (a) YP1–xVxO4 (x = 0, 0.2, and 0.4). (b) Y1–xBixPO4 (x = 0, 0.1, and 0.2).
3.2.2. Effect of Bismuth Substitution on the Structure of YPO4
Figure 4b depicts the PXRD spectra of Y1–yBiyPO4 (y = 0, 0.1, and 0.2). When y = 0.1, that is, Y0.9Bi0.1PO4, we could observe the formation of the tetragonal crystal structure without the formation of secondary phase. However, at y = 0.2, Y0.8Bi0.2O4, bismuth phosphate as BiPO4 appeared as a secondary phase. Emergence of the secondary phase was noticed due to the substitution of larger Bi3+ (ionic radii of bismuth = 1.03 Å) into smaller Y3+ (ionic radii of yttrium = 0.9 Å) lattice, which leads to lattice expansion.33 Higher loading of bismuth ion resulted in concoction of YPO4 and BiPO4 phases. Lower and higher substitution of bismuth into YPO4 lattice did not result in the expected color; therefore, synthesis of this series was not continued. Table 2 shows 2 theta and relative peak intensity (R.P.I) values of YP1–xVxO4, x = 0, 0.05, 0.1, 0.15, and 0.2) and Y0.9Bi0.1PO4.
Table 2. 2 Theta and R.P.I Values of YP1–xVxO4, x = 0, 0.05, 0.1, 0.15, and 0.2) and Y0.9Bi0.1PO4.
| YP |
YV05 |
YV1 |
YV15 |
YV2 |
YB1 |
||||||
|---|---|---|---|---|---|---|---|---|---|---|---|
| 2θ | aR.P.I | 2θ | aR.P.I | 2θ | aR.P.I | 2θ | aR.P.I | 2θ | aR.P.I | 2θ | aR.P.I |
| 19.751 | 973 | 19.723 | 969 | 19.719 | 978 | 19.711 | 945 | 19.698 | 986 | 19.715 | 792 |
| 25.932 | 3465 | 25.927 | 2920 | 25.910 | 2858 | 25.865 | 2789 | 25.804 | 2788 | 25.807 | 2756 |
| 32.633 | 385 | 32.629 | 450 | 32.620 | 378 | 32.608 | 365 | 32.596 | 340 | 32.760 | 333 |
| 35.275 | 1757 | 35.268 | 1749 | 35.256 | 1787 | 35.289 | 1723 | 35.265 | 1748 | 35.065 | 984 |
| 37.138 | 583 | 37.136 | 580 | 37.128 | 575 | 37.121 | 563 | 37.117 | 598 | 36.913 | 612 |
| 39.867 | 488 | 39.765 | 476 | 39.667 | 497 | 39.546 | 465 | 39.446 | 472 | 39.804 | 447 |
| 42.323 | 790 | 42.312 | 795 | 42.309 | 778 | 42.324 | 765 | 42.319 | 758 | 42.237 | 588 |
| 47.103 | 537 | 47.97 | 528 | 47.94 | 549 | 47.925 | 556 | 47.893 | 543 | 47.086 | 497 |
| 50.294 | 530 | 50.281 | 514 | 50.276 | 528 | 50.286 | 535 | 50.283 | 546 | 50.178 | 497 |
| 51.911 | 1291 | 51.891 | 1285 | 51.880 | 1274 | 51.876 | 1269 | 51.860 | 1245 | 51.845 | 734 |
| 53.385 | 576 | 53.378 | 586 | 53.369 | 592 | 53.360 | 567 | 53.367 | 583 | 53.365 | 576 |
| 60.348 | 375 | 60.343 | 363 | 60.339 | 397 | 60.326 | 384 | 60.321 | 384 | 59.923 | 340 |
| 65.544 | 392 | 65.539 | 373 | 65.520 | 365 | 65.513 | 386 | 65.497 | 395 | 65.512 | 269 |
| 67.999 | 392 | 67.986 | 385 | 67.964 | 374 | 67.986 | 358 | 67.976 | 376 | 67.839 | 400 |
| 74.281 | 441 | 74.276 | 435 | 74.276 | 467 | 74.275 | 456 | 74.263 | 472 | 74.042 | 400 |
Relative peak intensity.
3.2.3. Effect of Simultaneous Substitution of Both V5+ and Bi3+ into YPO4 Lattice
Figure 5 shows the PXRD spectra of Y1–xBixP1–yVyO4 (x = y = 0, 0.2, 0.4, and 0.6). Co-substitution of bismuth and vanadium in the YPO4 lattice was proposed to improve its color properties. At x = y = 0.2, one could not notice any secondary phase formation. Higher substitution leads to the formation of mixed phases of YPO4 and BiPO4. Absence of V2O5 as the secondary phase is due to the possibility of iso-structural formation of YPO4 and YVO4.29
Figure 5.

PXRD pattern of Y1–xBixP1–yVyO4 (x = y = 0, 0.2, 0.4, and 0.6).
3.3. Investigation on the FT-IR and Raman Spectra of Synthesized Materials
The interaction between the metal ions present (Y, V, P, Bi, and O) in the fabricated materials with oxide ion (O2–) is identified by decoding the FT-IR spectra of the materials YPO4, YP0.8V0.2O4, Y0.9Bi0.1PO4, and Y0.8Bi0.2P0.8V0.2O4, as illustrated in Figure 6. In all the portrayed samples, two different significant vibrations were noticed at 637 and 518 cm–1 corresponding to (PO4)3– phosphate moiety, due to distorted cockeyed vibrations. A peak at 987 cm–1 is due to P–O stretching vibrations.34−36 An additional vibration at 835 and 800 cm–1 was noticed for YV2 and YB2V2 due to the presence of VO43–. A similar observation was noticed when Eu3+ and V5+ were doped into YPO4 lattice.34 The presence of metal oxide vibrations such as Bi–O and Y–O is noticed between 400 and 450 cm–1.37 The abovementioned results emphasized the formation of yttrium phosphate-, vanadium-, and bismuth-doped yttrium phosphates.
Figure 6.

FT-IR spectra of YPO4, YP0.8V0.2O4, Y0.9Bi0.1PO4, Y0.8Bi0.2P0.8V0.2O4.
Raman spectral analysis of Y0.8Bi0.2P0.8V0.2O4 was carried out to identify the necessary vibration in the material, and the spectrum is depicted in Figure 7. As shown in Figure 7, the spectrum consists of various Raman vibrations that correspond to VO4, PO4, and Bi–O vibrations. The Raman shifts observed at 1011 cm–1 refer to asymmetric stretching vibrations of the PO4 group. The peaks at 501 cm–1 corresponding to bending mode of the PO4 group and the band at 196 cm–1 could be due to O–Bi–O bending mode of vibrations.38,39 VO4 stretching vibrations are observed at 841 cm–1.40 Significant vibrations indicate the presence of metal oxide bonding, phosphate, and vanadate moiety in the YB2V2 material.
Figure 7.

Raman spectrum of Y0.8Bi0.2P0.8V0.2O4.
3.4. X-ray Photoelectron Spectroscopy Analysis
To appreciate the chemical state of all the metal cations present in the prepared sample, XPS was carried out. Figure 8 depicts the XPS spectra of Y0.4Bi0.6P0.4V0.6O4 (as a representation) synthesized by the chemical precipitation method. Standard C 1s binding energy was compared with C 1s energy obtained by the experiment, and the results are used to calibrate of binding energy of elements present in the system. The high resolution peak are fitted to show Y 3d, P 2p, V 2p, O 1s, and Bi 4f states present in the material. The valence energy state of metal ions is identified based on their binding energy locations. The Y 3d state was recognized by the presence of peaks at 157.17 and 158.88 eV corresponding to 3d5/2 and 3d3/2.41 Phosphorous in its +5 oxidation state was identified by the existence of a peak at 133.32 eV, suggesting the presence of the P 2p energy level of the cation. From the figure, one could confirm the presence of the V 2p state by observing the binding energy at 517 eV corresponding to the V 2p1/2 level and 524.53 eV corresponding to the V 2p3/2 level. Parhi and Manivannan42 reported the existence of the V 2p1/2 energy level at 518.4 eV in YVO4. A shift in the peak position to a lower energy level is noticed in the current investigation, which could be due to the substitution effect. The deconvoluted spectrum of oxygen also presented in the same illustration. Here, the O 1s core level is fitted by the presence of three binding energies centered at 531.64, 530.57, and 529.61 eV corresponding to the surface hydroxyl group, oxygen in the surface lattice, and the metal–oxygen bond present in the material.43 Significant spin orbit doublets, namely, Bi 4f5/2 and Bi 4f7/2 were centered at 158.71 and 164.12 eV, which confirmed the presence of bismuth in its +3 oxidation state. These doublet peaks are further deconvoluted, resulting in four different binding energies found at 158.81, 158.29, 163.76, and 164.29 eV.44
Figure 8.
XPS spectra of Y0.4Bi0.6P0.4V0.6O4. (a) Y 3d, (b) P 2p, (c) V 2p and O 1s, (d) Bi 4f.
3.5. SEM, TEM, and EDAX Analyses
SEM micrographs of YPO4, YP0.8V0.2O4, Y0.9Bi0.1PO4, and Y0.4Bi0.6P0.4V0.6O4 are illustrated in Figure 9. The images clearly show that the particles are highly aggregated due to the rapid growth of particles during the precipitation reaction. This process of aggregation of nanoparticles begins at the stage of pre-nuclei formation after the dissolution of precursor materials. The nanoparticles are formed after the formation of a new boundary between the nuclei due to the free energy supplied to them. However, in some cases, there is a possibility of formation of pre-nucleation clusters, which leads to aggregation of small particles. Such aggregated particles are very stable and difficult to separate during mechanical grinding.45 No significant change in the morphology was identified due to the introduction of vanadium and bismuth into tetragonal YPO4, whereas co-substitution of both vanadium and bismuth in the YPO4 lattice induced the change in morphology as a well-defined rectangular particle, which is clearly seen in Figure 9. The change in morphology could be due to the substitution effect.46 The elemental composition obtained from EDAX analysis of selected elements is tabulated in Table 3, which confirmed the presence of Y, P, V, Al, Bi, and O in the currently reported materials. In order to arrive at a better conclusion about the morphology of the fabricated materials, TEM analysis was carried out. Figure 10 shows the TEM and EDAX images of Y0.4Bi0.6P0.4V0.6O4. Rectangular-shaped particles with particle size ranging between 85 and 100 nm are obtained, which are in good agreement with the results obtained by PXRD. Dark spots from the SAED pattern indicates the crystalline nature of the sample.
Figure 9.
SEM images of (a) YPO4, (b) YP0.8V0.2O4, (c) Y0.9Bi0.1PO4, and (d) Y0.4Bi0.6P0.4V0.6O4 pigments.
Table 3. Elemental Composition of the Selected Materials Obtained from EDAX Analysis.
| composition | element | weight % | atomic % |
|---|---|---|---|
| Y0.9Bi0.1PO4 | Y | 17.23 | 3.53 |
| Bi | 3.81 | 0.33 | |
| P | 3.13 | 1.84 | |
| O | 83.44 | 94.96 | |
| total | 100 | ||
| Y0.4Bi0.6P0.4V0.6O4 | Y | 14.28 | 3.47 |
| P | 1.58 | 1.10 | |
| Bi | 12.14 | 1.26 | |
| V | 3.40 | 1.44 | |
| O | 68.60 | 92.73 | |
| total | 100 |
Figure 10.
TEM images and EDAX spectrum of the Y0.4Bi0.6P0.4V0.6O4 pigment.
3.6. UV-DRS Analysis
Figure 11 shows the UV-DRS spectra of YPO4, YP1–xVxO4 (x = 0.05, 0.1, 0.15, 0.2, and 0.4), and Y1–xBixP1–yVyO4 (x = y = 0, 0.2, 0.4, and 0.6) prepared using the precipitation method. As mentioned in another study, YPO4 shows a weak absorption peak between 230 and 330 nm due to charge transfer transition between the divalent oxygen anion (O22–) and the pentavalent phosphorus cation (P5+),47 and the color of material was found to be white. An additional transition between 350 and 470 nm was perceived because of the increase in vanadium content and due to the charge transfer between V5+–O2– rather than d–d transition, since V5+ had a d0 electronic configuration.48 As a result, yellow tinge was developed for the engineered materials. Bismuth and vanadium co-substituted materials slowly shifted the absorption maximum to higher wavelength, and it reached to 450 nm while substituting 60% of bismuth and vanadium in the YPO4 lattice. Since 400–450 nm is a violet region, its complementary yellow color is observed. Here, introduction of vanadium in the phosphorous lattice produces VO43– transition. Along with this VO43– transition, a hybrid orbital transition is also possible from the mixture of Bi 6s and O 2p orbitals to V 3d orbitals.49 The abovementioned phenomenon is responsible for the intense yellow color of the co-doped materials.
Figure 11.

UV-DRS spectrum of YPO4, YP1–xVxO4 (x = 0.05, 0.1, 0.15, 0.2, and 0.4), and Y1–xBixP1–yVyO4 (x = y = 0, 0.2, 0.4, and 0.6).
The digital images of vanadium-substituted YPO4 and bismuth and vanadium co-substituted YPO4 are presented in Figures 12 and 13, respectively, and their corresponding color values are tabulated in Table 4. When x = 0, a negative b* (−1.29) value was obtained, confirming their white color. Substitution of bismuth into yttrium lattice does not significantly affect the color; hence they are white, which is further confirmed by its corresponding color values. Substitution of 0.2 M of vanadium slightly increased the yellow tinge of material to +15.71. 0.4 M of vanadium in the P5+ increased the yellowness of the material to +26.6 with a slight decrease in lightness of the material. Here, the amount of V5+ content was doubled from 0.2 to 0.4 M so as to increase the yellow hue of the pigment. Further addition of V5+ in the YPO4 structure did not increase the b* value. Simultaneous addition of both bismuth and vanadium in the YPO4 steadily increased the yellowness value from +26.16 to +56.06 (Table 4). The hue angle represents the color of the system in terms of angular position in the cylindrical color space. It is observed from the literature, if the hue angle was found to be between 70 and 105°, the material will exhibit yellow color. All the optically engineered materials in the current study are having the hue angle between 70 and 90°, which confirmed the yellow color of the pigments.
Figure 12.

Digital images of YPO4, YP1–xVxO4 (x = 0.2 and 0.4), and Y1–xBixPO4 (x = 0.1, and 0.3).
Figure 13.

Digital images of Y1–xBixP1–yVyO4 (x = y = 0, 0.2, 0.4, and 0.6).
Table 4. Color Properties and NIR Solar Reflectance of Synthesized Compounds.
| color co-ordinates |
||||||
|---|---|---|---|---|---|---|
| compounds | L | a* | b* | C* | h° | NIR solar reflectance R* (%) |
| YPO4 | 95.09 | 0.56 | –1.29 | 1.41 | 26.53 | 68 |
| Y0.9Bi0.1PO4 | 96.23 | 2.00 | 1.26 | 1.72 | 32.21 | 81 |
| Y0.7Bi0.3PO4 | 94.78 | 1.78 | 1.90 | 1.89 | 46.86 | 65 |
| YP0.8V0.2O4 | 95.97 | –1.02 | 5.71 | 5.23 | 77.96 | 78 |
| YP0.6V0.4O4 | 89.36 | –2.86 | 25.62 | 26.16 | 87.48 | 64 |
| Y0.8Bi0.2P0.8V0.2O4 | 85 | 4.20 | 30.14 | 31.54 | 82.06 | 72 |
| Y0.6Bi0.4P0.6V0.4O4 | 91.35 | –5.81 | 30.55 | 35.76 | 87.68 | 82 |
| Y0.4Bi0.6P0.4V0.6O4 | 81.36 | 1.87 | 56.06 | 60.25 | 88.08 | 83 |
3.7. NIR Reflectance Analysis
Figures 14 and 15 show the NIR reflectance of YP1–xVxO4 (x = 0, 0.2, and 0.4) and Y1–xBixP1–yVyO4 (x = y = 0, 0.2, 0.4, and 0.6). The NIR solar reflectance of white YPO4 in the 800–2500 nm NIR region was found to be 68%. Significant increase in the NIR reflectance was observed after the addition of 0.2 M of V5+ ion in the YPO4 structure. Nearly 10% increase in the reflectance was noticed; hence, the NIR reflectance increased from 68 to 78%. Further addition of V5+, that is, 0.4 M addition of vanadium content significantly reduced the NIR reflectance from 68 to 64%. Concurrent addition of both bismuth and vanadium in the YPO4 structure increased the NIR reflectance of the hybrid material gradually and it reached to 82% for Y0.4Bi0.6P0.4V0.6O4. The reason for the increase in reflectance of material is explained due to the morphology change that occurred during co-substitution of both bismuth and vanadium in the YPO4 structure. According to the Kubelka–Munk (KM) theory, the mean particle size (d) is inversely related to the scattering coefficient (S). Hence, if the mean particle size is less (d < 1 μm), the scattering efficiency of the material increases, which in turn increases the reflectance.50 In the current study, the co-substituted materials have the least mean particle size of 98 nm (YB6V6), 83 nm (YB4V4), 112 nm (YB2V2) compared to vanadium-substituted YPO4 materials. According to KM theory, more regular particles with a definite morphology are also important for high NIR reflectance. Rectangular-shaped particles with more regular distribution could be one of the important reasons for high reflectance of YB6V6. The abovementioned result was supported by Tao et al., who stated that the high reflectance could be due to the uniformed spherical particles and low particle size.51
Figure 14.

NIR solar reflectance spectrum of YP1–xVxO4 (x = 0, 0.2, and 0.4); reflectance spectra are provided in the inset.
Figure 15.

NIR solar reflectance spectrum of Y1–xBixP1–yVyO4 (x = y = 0, 0.2, 0.4, and 0.6).
3.8. Chemical and Thermal Stability Test
Table 5 shows the CIE Lab color values of Y0.4Bi0.6P0.4V0.6O4 after it was subjected to chemical and thermal treatment. From the results, one could observe very less change in the b* value after treating with 2% acid and base medium and a negligible amount of weight loss, indicating that they are chemically stable and can be used as metallic coating materials. To check the suitability of the material for the application in ceramic and glaze substances, it was calcined to higher temperatures from 900 to 1000 °C and 1200 °C. As the temperature increases, we could observe that the b* value increases from 56.06 to 70.64, which shows that the yellowness value increases as the temperature increases.
Table 5. Color Values of Y0.4Bi0.6P0.4V0.6O4 after Chemical and Thermal Treatment.
| color
parameters |
|||
|---|---|---|---|
| 2% acid or alkali | L* | a* | b* |
| 81.36 | 1.87 | 56.06 | |
| NaOH | 82.40 | 1.73 | 55.93 |
| HNO3 | 83.98 | 1.69 | 54.82 |
| HCl | 81.28 | 1.85 | 55.92 |
| H2SO4 | 81.40 | 1.89 | 56.04 |
| Thermal Stability | |||
| 900 °C | 81.36 | 1.87 | 56.06 |
| 1000 °C | 78.46 | 1.12 | 63.49 |
| 1200 °C | 50.65 | 0.65 | 70.64 |
4. Conclusions
Tetragonal YPO4-, vanadium-, and bismuth-substituted YPO4 and Bi and V co-substituted YPO4 are prepared by a simple precipitation method and authenticated by PXRD, FTIR, and Raman spectroscopy. From XRD, we could notice that 0.4 M of vanadium substitution into the YPO4 structure formed a secondary phase structure V2O5, whereas the color of the material changed from white to yellow. Substitution of bismuth into the YPO4 structure does not increase the b* value, which proved that the bismuth ion does not play a significant role in the formation of yellow color of the pigment, while influence of vanadium in the development of yellow color is noticed. On the other hand, co-substitution of both vanadium and bismuth in the YPO4 lattice enhanced the b* value and NIR reflectance. Y0.4Bi0.6P0.4V0.6O4 showed 83% NIR solar reflectance, and its b* value was also found to be higher (56.06). The reason for higher reflectance could be due to the formation of more uniform particles, which was confirmed by TEM images. The fabricated material was found to be stable while treating with 2% H2SO4. Higher thermal treatments (subjected to calcination at 1000 and 1200 °C) increased the b* value of YB6V6 to 63.49 and 70.64. Hence, with higher b* and NIR solar reflectance, Y0.4Bi0.6P0.4V0.6O could serve as an environmentally benign cool yellow pigment.
Acknowledgments
The authors acknowledge VIT, Vellore, for providing lab facility to perform research. S.S. and V.E. would like to acknowledge Council of Scientific and Industrial Research (CSIR), New Delhi, Government of India, for the financial support toward Extramural research fund [01(3085)21/EMR/-II] and Senior Research Fellowship [file no. 09/844/(0098)/2020-EMR-1].
The authors declare no competing financial interest.
References
- Li Q.; Zhang H.; Liu X.; Huang J. Urban heat island effect on annual mean temperature during the last 50 years in China. Theor. Appl. Climatol. 2004, 79, 165–174. 10.1007/s00704-004-0065-4. [DOI] [Google Scholar]
- Chua K. J.; Chou S. K.; Yang W. M.; Yan J. Achieving better energy-efficient air conditioning - A review of technologies and strategies. Appl. Energy 2013, 104, 87–104. 10.1016/j.apenergy.2012.10.037. [DOI] [Google Scholar]
- Bretz S.; Akbari H.; Rosenfeld A. Practical issues for using solar-reflective materials to mitigate urban heat islands. Atmos. Environ. 1998, 32, 95–101. 10.1016/s1352-2310(97)00182-9. [DOI] [Google Scholar]
- Pisello A. L.; Cotana F.; Brinchi L. On a cool coating for roof clay tiles: development of the prototype and thermal-energy assessment. Energy Proc. 2014, 45, 453–462. 10.1016/j.egypro.2014.01.049. [DOI] [Google Scholar]
- Sreeram K. J.; Aby C. P.; Nair B. U.; Ramasami T. Colored cool colorants based on rare earth metal ions. Sol. Energy Mater. Sol. Cells 2008, 92, 1462–1467. 10.1016/j.solmat.2008.06.008. [DOI] [Google Scholar]
- Zhang Y.; Zhang Y.; Zhao X.; Zhang Y. Sol-gel synthesis and properties of europium-strontium copper silicates blue pigments with high near-infrared reflectance. Dyes Pigm. 2016, 131, 154–159. 10.1016/j.dyepig.2016.04.011. [DOI] [Google Scholar]
- Masui T.; Takeuchi N.; Nakado H.; Imanaka N. Novel environment-friendly green pigments based on rare earth cuprate. Dyes Pigm. 2015, 113, 336–340. 10.1016/j.dyepig.2014.09.007. [DOI] [Google Scholar]
- George G.; Vishnu V. S.; Reddy M. L. P. The synthesis, characterization and optical properties of silicon and praseodymium doped Y6MoO12 compounds: Environmentally benign inorganic pigments with high NIR reflectance. Dyes Pigm. 2011, 88, 109–115. 10.1016/j.dyepig.2010.05.010. [DOI] [Google Scholar]
- Han A.; Ye M.; Liu L.; Feng W.; Zhao M. Estimating thermal performance of cool coatings colored with high near-infrared reflective inorganic pigments: Iron doped La 2 Mo 2 O 7 compounds. Energy Build. 2014, 84, 698–703. 10.1016/j.enbuild.2014.08.024. [DOI] [Google Scholar]
- Divya S.; Das S. Eco-friendly Li3InB2O6 based red pigments for various IR blocking cool coating applications. Opt. Mater. 2020, 109, 110410. 10.1016/j.optmat.2020.110410. [DOI] [Google Scholar]
- Pailhé N.; Gaudon M.; Demourgues A. (Ca2+, V5+) co-doped Y2Ti2O7 yellow pigment. Mater. Res. Bull. 2009, 44, 1771–1777. 10.1016/j.materresbull.2009.03.009. [DOI] [Google Scholar]
- Calatayud J. M.; Alarcón J. V-containing ZrO 2 inorganic yellow nano-pigments prepared by hydrothermal approach. Dyes Pigm. 2017, 146, 178–188. 10.1016/j.dyepig.2017.07.003. [DOI] [Google Scholar]
- Gopal B.; Muralidharan A.; Bakthavatsalam R.; Nellaiappan S.; Kumar A. S. Unusual observation of optical property of V 5+ substituted BPO 4 and its tunable redox features. Mater. Res. Bull. 2017, 91, 122–126. 10.1016/j.materresbull.2017.03.020. [DOI] [Google Scholar]
- Lai H.; Du Y.; Zhao M.; Sun K.; Yang L. CTAB assisted hydrothermal preparation of YPO4:Tb3+ with controlled morphology, structure and enhanced photoluminescence. Mater. Sci. Eng., B 2014, 179, 66–70. 10.1016/j.mseb.2013.10.007. [DOI] [Google Scholar]
- Zhang Y. W.; Yan Z. G.; You L. P.; Si R.; Yan C. H. General synthesis and characterization of monocrystalline lanthanide orthophosphate nanowires. Eur. J. Inorg. Chem. 2003, 4099–4104. 10.1002/ejic.200300399. [DOI] [Google Scholar]
- Yan Z. G.; Zhang Y. W.; You L. P.; Si R.; Yan C. H. General synthesis and characterization of monocrystalline 1D-nanomaterials of hexagonal and orthorhombic lanthanide orthophosphate hydrate. J. Cryst. Growth 2004, 262, 408–414. 10.1016/j.jcrysgro.2003.10.058. [DOI] [Google Scholar]
- Ghosh P.; de la Rosa E.; Oliva J.; Solis D.; Kar A.; Patra A. Influence of surface coating on the upconversion emission properties of LaPO4:Yb/Tm core-shell nanorods. J. Appl. Phys. 2009, 105, 113532. 10.1063/1.3143890. [DOI] [Google Scholar]
- Ma C. G.; Popov A. V.; Vanetsev A. S.; Gaitko O. M.; Orlovskaya E. O.; Lange S.; Sildos Y. V.; Orlovskii Y. V. Vacuum ultraviolet spectroscopic analysis of Ce3+-doped hexagonal YPO4·0.8H2O based on exchange charge model. J. Lumin. 2014, 152, 70–74. 10.1016/j.jlumin.2013.10.069. [DOI] [Google Scholar]
- Vanetsev A. S.; Samsonova E. V.; Gaitko O. M.; Keevend K.; Popov A. V.; Mäeorg U.; Mändar Y. V.; Sildos I.; Orlovskii Y. V. Phase composition and morphology of nanoparticles of yttrium orthophosphates synthesized by microwave-hydrothermal treatment: the influence of synthetic conditions. J. Alloys Compd. 2015, 639, 415–421. 10.1016/j.jallcom.2015.03.125. [DOI] [Google Scholar]
- Wujczyk M.; Watras A.; Szyszka K.; Wiglusz R. J. Influence of vanadium concentration on up-conversion luminescence in Er3+-Yb3+ and Tm3+-Yb3+ ions pair co-doped YVxP1–xO4 solid state solution. J. Alloys Compd. 2021, 884, 161022. 10.1016/j.jallcom.2021.161022. [DOI] [Google Scholar]
- Chanchan K.; Singh S. D. Concentration effect on the luminescent and structural properties of YPO4: Eu3+. Mater. Today: Proc. 2022, 65, 2480–2483. 10.1016/j.matpr.2022.04.055. [DOI] [Google Scholar]
- Minakova T.; Mjakin S.; Bakhmetyev V.; Sychov M.; Zyatikov I.; Ekimova I.; Kozik I.; Chen Y.-W.; Kurzina I. High Efficient YVPO4 Luminescent Materials Activated by Europium. Crystals 2019, 9, 658. 10.3390/cryst9120658. [DOI] [Google Scholar]
- Zhang Z.; Zhang Y.; Feng F.; Cheng W.; Xia X.; Zhang X. Luminescent properties of Ce3+/Tb3+co-doped glass ceramics containing YPO4 nanocrystals for W-LEDs. J. Rare Earths 2016, 34, 464–469. 10.1016/s1002-0721(16)60050-9. [DOI] [Google Scholar]
- Cao Y.; Liu Y.; Feng H.; Yang Y. Effects of Bi3+ co-doping on luminescence of YPO4:Dy3+ powders. Ceram. Int. 2014, 40, 15319–15323. 10.1016/j.ceramint.2014.06.127. [DOI] [Google Scholar]
- Angiuli F.; Cavalli E.; Boutinaud P.; Mahiou R. Emission properties of Sm3+/Bi3+-doped YPO4 phosphors. J. Lumin. 2013, 135, 239–242. 10.1016/j.jlumin.2012.10.005. [DOI] [Google Scholar]
- Elakkiya V.; Sumathi S. Effect of preparation methodology on the colour and the NIR reflectance of nickel doped manganese pyrophosphate. J. Alloys Compd. 2018, 768, 535–544. 10.1016/j.jallcom.2018.07.236. [DOI] [Google Scholar]
- Deng T.; Yan S.; Hu J. A novel narrow band UV-B emitting phosphor-YPO4:Sb3+,Gd3+. J. Rare Earths 2016, 34, 137–142. 10.1016/s1002-0721(16)60005-4. [DOI] [Google Scholar]
- Ghosh D.; Luwang M. N. Arsenic detection in water: YPO4:Eu3+ nanoparticles. J. Solid State Chem. 2015, 232, 83–90. 10.1016/j.jssc.2015.09.009. [DOI] [Google Scholar]
- Abid N.; Khan A. M.; Shujait S.; Chaudhary K.; Ikram M.; Imran M.; Haider M.; Khan M.; Khan Q.; Maqbool M. Synthesis of nanomaterials using various top-down and bottom-up approaches, influencing factors, advantages, and disadvantages: A review. Adv. Colloid Interface Sci. 2022, 300, 102597. 10.1016/j.cis.2021.102597. [DOI] [PubMed] [Google Scholar]
- Di W.; Zhao X.; Lu S.; Wang X.; Zhao H. Thermal and photoluminescence properties of hydrated YPO4:Eu3+ nanowires. J. Solid State Chem. 2007, 180, 2478–2484. 10.1016/j.jssc.2007.06.025. [DOI] [Google Scholar]
- Kahouadji B.; Guerbous L.; Jovanović D. J.; Dramićanin M. D. Temperature dependence of red emission in YPO4:Pr3+ nanopowders. J. Lumin. 2022, 241, 118499. 10.1016/j.jlumin.2021.118499. [DOI] [Google Scholar]
- Ding C.; Tian M.; Han A.; Ye M.; Chen X. V-doped LaPO4 new solar heat-reflective pigments and its improvement on the aging resistance of poly-methyl methacrylate. Sol. Energy 2020, 195, 660–669. 10.1016/j.solener.2019.12.002. [DOI] [Google Scholar]
- Kirankumar V. S.; Sumathi S. Catalytic activity of bismuth doped zinc aluminate nanoparticles towards environmental remediation. Mater. Res. Bull. 2017, 93, 74–82. 10.1016/j.materresbull.2017.04.022. [DOI] [Google Scholar]
- Batista J. C.; de Sousa Filho P. C.; Serra O. A. Effect of the vanadium(v) concentration on the spectroscopic properties of nanosized europium-doped yttrium phosphates. Dalton Trans. 2012, 41, 6310–6318. 10.1039/c2dt30380a. [DOI] [PubMed] [Google Scholar]
- Kijkowska R.; Cholewka E.; Duszak B. X-ray diffraction and Ir-absorption characteristics of lanthanide orthophosphates obtained by crystallisation from phosphoric acid solution. J. Mater. Sci. 2003, 38, 223–228. 10.1023/a:1021188810349. [DOI] [Google Scholar]
- Nakamoto K.Infrared Spectra of Inorganic and Coordination Compounds; John Wiley & Sons: New York–London. 328 Seiten, Zahlreiche Abbildungen und Tabellen. Preis: 72s, 1963; Vol. 67, pp 996. [Google Scholar]
- Luwang M. N.; Ningthoujam R. S.; Srivastava S. K.; Vatsa R. K. Disappearance and Recovery of Luminescence in Bi3+, Eu3+ Codoped YPO4 Nanoparticles Due to the Presence of Water Molecules Up to 800 °C. J. Am. Chem. Soc. 2011, 133, 2998–3004. 10.1021/ja1092437. [DOI] [PubMed] [Google Scholar]
- Azzam A. B.; El-Sheikh S. M.; Geioushy R. A.; Salah B. A.; El-Dars M.; Helal A. S. Facile fabrication of a novel BiPO4 phase junction with enhanced photocatalytic performance towards aniline blue degradation. RSC Adv. 2019, 9, 17246–17253. 10.1039/c9ra02315a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bouddouch A.; Amaterz E.; Bakiz B.; Guinneton F.; Taoufyq A.; Villain S.; Gavarri J. R.; Mansori M.; Valmalette J. C.; Benlhachemi A. High photocatalytic performance of bismuth phosphate and corresponding photodegradation mechanism of Rhodamine B. Res. Chem. Intermed. 2022, 48, 3315–3334. 10.1007/s11164-022-04762-1. [DOI] [Google Scholar]
- Kolesnikov I. E.; Kalinichev A. A.; Kurochkin M. A.; Golyeva E. V.; Terentyeva A. S.; Kolesnikov E. Y.; Lähderanta E. Structural, luminescence and thermometric properties of nanocrystalline YVO4:Dy3+ temperature and concentration series. Sci. Rep. 2019, 9, 2043. 10.1038/s41598-019-38774-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang W. K.; Wang S. Y.; Liu K. F.; Tsai P. C.; Zhang Y. H.; Huang S. Y. Plasma Etching Behavior of SF6 Plasma Pre-Treatment Sputter-Deposited Yttrium Oxide Films. Coatings 2020, 10, 637. 10.3390/coatings10070637. [DOI] [Google Scholar]
- Parhi P.; Manivannan V. Novel microwave initiated solid-state metathesis synthesis and characterization of lanthanide phosphates and vanadates, LMO4 (L=Y, La and M=V, P). Solid State Sci. 2008, 10, 1012–1019. 10.1016/j.solidstatesciences.2007.11.038. [DOI] [Google Scholar]
- Elakkiya V.; Abhishekram R.; Sumathi S. Copper doped nickel aluminate: Synthesis, characterisation, optical and colour properties. Chin. J. Chem. Eng. 2019, 27, 2596–2605. 10.1016/j.cjche.2019.01.008. [DOI] [Google Scholar]
- Savunthari K. V.; Shanmugam S. Effect of co-doping of bismuth, copper and cerium in zinc ferrite on the photocatalytic degradation of bisphenol A. J. Taiwan Inst. Chem. Eng. 2019, 101, 105–118. 10.1016/j.jtice.2019.04.042. [DOI] [Google Scholar]
- Li D.; Kaner R. B. How nucleation affects the aggregation of nanoparticles. J. Mater. Chem. 2007, 17, 2279–2282. 10.1039/b700699c. [DOI] [Google Scholar]
- Rao P. A.; Raghavendra V.; Suryanarayana B.; Paulos T.; Murali N.; Varma P. P.; Chandramouli K. Cadmium substitution effect on structural, electrical and magnetic properties of Ni-Zn nano ferrites. Results Phys. 2020, 19, 103487. 10.1016/j.rinp.2020.103487. [DOI] [Google Scholar]
- Wu J.; Jia H.; Li M.; Jia H.; Liu Z. Influence of pH on nano-phosphor YPO4: 2% Sm3+ and luminescent properties. Appl. Phys. A 2020, 126, 87. 10.1007/s00339-019-3257-6. [DOI] [Google Scholar]
- Wang H.; Odawara O.; Wada H. One-step preparation of YVO4:Eu3+ nanoparticles by pulsed laser ablation. Sci. Rep. 2016, 683, 1–6. 10.1016/j.jallcom.2016.05.071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sameera S.; Prabhakar Rao P. P.; James V.; Divya S.; Raj A. K. Influence of (LiLa)1/2MoO4 substitution on the pigmentary properties of BiVO4. Dyes Pigm. 2014, 104, 41–47. 10.1016/j.dyepig.2013.12.029. [DOI] [Google Scholar]
- Pasikatan M. C.; Steele J. L.; Spillman C. K.; Haque E. Near infrared reflectance spectroscopy for online particle size analysis of powders and ground materials. J. Near Infrared Spectrosc. 2001, 9, 153–164. 10.1255/jnirs.303. [DOI] [Google Scholar]
- Tao Z.; Zhang W.; Huang Y.; Wei D.; Seo H. J. A novel pyrophosphate BaCr2(P2O7)2 as green pigment with high NIR solar reflectance and durable chemical stability. Solid State Sci. 2014, 34, 78–84. 10.1016/j.solidstatesciences.2014.05.016. [DOI] [Google Scholar]







