Next Article in Journal
A Novel Measurement Approach to Experimentally Determine the Thermomechanical Properties of a Gas Foil Bearing Using a Specialized Sensing Foil Made of Inconel Alloy
Previous Article in Journal
Effects of Solvent Additive and Micro-Patterned Substrate on the Properties of Thin Films Based on P3HT:PC70BM Blends Deposited by MAPLE
Previous Article in Special Issue
Eu3+ and Tb3+ @ PSQ: Dual Luminescent Polyhedral Oligomeric Polysilsesquioxanes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Luminescence Properties and Energy Transfer of Eu3+, Bi3+ Co-Doped LuVO4 Films Modified with Pluronic F-127 Obtained by Sol–Gel

by
Brenely González-Penguelly
1,2,
Grethell Georgina Pérez-Sánchez
1,
Dulce Yolotzin Medina-Velázquez
1,*,
Paulina Martínez-Falcón
1 and
Angel de Jesús Morales-Ramírez
3,4
1
División de Ciencias Básicas e Ingeniería, Universidad Autónoma Metropolitana-Azcapotzalco, Ciencias Básicas e Ingeniería, Av. San Pablo No. 180, CDMX 02200, Mexico
2
Centro Universitario UAEM Valle de Mexico, Boulevard Universitario S/N Valle Escondido, Río San Javier, Cd López Mateos 54500, Mexico
3
Instituto Politécnico Nacional, CIITEC IPN, Cerrada de Cecati S/N Col. Santa Catarina, Azcapotzalco, CDMX 02250, Mexico
4
Instituto Politécnico Nacional, ESIQIE IPN, Av. Luis Enrique Erro S/N, Nueva Industrial Vallejo, Gustavo A. Madero, CDMX 07738, Mexico
*
Author to whom correspondence should be addressed.
Materials 2023, 16(1), 146; https://doi.org/10.3390/ma16010146
Submission received: 8 September 2022 / Revised: 28 November 2022 / Accepted: 1 December 2022 / Published: 23 December 2022
(This article belongs to the Special Issue Structural and Optical Studies of Eu3+ Doped Materials)

Abstract

:
Nowadays, orthovanadates are studied because of their unique properties for optoelectronic applications. In this work, the LuVO4:Eu3+, Bi3+ films were prepared by the sol–gel method, using a new simple route, and deposited by the dip-coating technique. The obtained films are transparent, fracture-free, and homogenous. The sol–gel process was monitored by Fourier-transform infrared spectroscopy (FTIR), and according to X-ray diffraction (XRD) results, the crystal structure was tetragonal, and films that were highly oriented along the (200) low-energy direction were obtained. The morphological studies by scanning electron microscopy (SEM) showed uniformly distributed circular agglomerations of rice-like particles with nanometric sizes. The luminescence properties of the films were analyzed using a fixed concentration of 2.5 at. % Eu3+ and different concentrations of Bi3+ (0.5, 1.0, and 1.5 at. %); all the samples emit in red, and it has been observed that the light yield of Eu3+ is enhanced as the Bi3+ content increases when the films are excited at 350 nm, which corresponds to the 1S03P1 transition of Bi3+. Therefore, a highly efficient energy transfer mechanism between Bi3+ and Eu3+ has been observed, reaching up to 71%. Finally, it was established that this energy transfer process occurs via a quadrupole–quadrupole interaction.

1. Introduction

Nowadays, phosphor materials with luminescent properties prepared by various techniques are extensively studied because of their great ability to convert UV radiation to the visible spectrum and its multiple applications in different fields, such as integrated photonics [1], optics [2,3], and electronics [4,5]. Most luminescent materials are usually oxides, sulfides, or oxysulfides doped with some transition metal or rare-earth element [6,7,8], but in recent years, some lanthanides known as orthovanadates are currently used as matrices due to their high emission efficiency [9,10,11,12]. Additionally, these materials have unique properties due to the stoichiometric combination of their components and their symmetry with the crystal of the complex oxides; these characteristics increase the luminescent activity in comparison with the simple materials [13,14,15,16,17,18]. Likewise, rare-earth ions that are used as impurities in orthovanadates have shown strong emissions as well as variations in the emission color corresponding to the different activating ions under the excitation of an electron beam or ultraviolet. The luminescent properties presented in these materials depend largely on the location of the 4f energy levels of the dopants and the ratio of the valence and conduction bands (VB and CB, respectively) of the matrix [3,12,19,20,21]. In the energy transfer, the co-doped materials are interesting for optoelectronic applications because of increases in their luminescent effect. Particularly for the Eu3+ emission, several ions have been proposed to increase its well-known reddish emission, for example, usually by the use of other rare earths, such as Dy3+ [22], Tb3+ [23], Tm3+ [24], or Sm3+ [25]. However, the search of other non-rare earths sensitizers has attracted attention in order not only to increase the light yield, but also to modify the excitation wavelength. Therefore, other ions have been studied, like Sb3+ [26] or Li+ [27]. Another possibility is the use of Bi3+, which increases the luminescent properties and broadens the excitation spectrum of europium ions in LuVO4 [28] or Lu2O3 [29]. In the case of LuVO4, since it has been shown to have better optical properties than yttrium vanadate [2,8,30], several routes of synthesis have been studied to obtain higher optical properties. However, most of the synthesis routes that have been developed require high temperatures and highly controlled environments to carry out the reaction, for example, in solid state or hydrothermal reaction synthesis. Consequently, new energy-saving simple-synthesis processes are required that can be scalable to industry. In the case of sol–gel synthesis, it allows one to prepare homogeneous, high-purity materials with the possibility of controlling morphologies and impurities at low preparation temperatures without extremely controlled environments. On the other hand, Pluronic F-127 acid was added to the synthesis because it changes the properties of the film, as was shown in Tsotetsi et al. in 2022 [31] and Arconada et al. in 2010 [32], where it was used to increase the thickness because of its viscosity, for providing porosity to the film, and as a morphology controller. Another characteristic of Pluronic F-127 is that it acts as a structure-directing agent in TiO2 in catalytic and solar cell applications [33,34]. In the present work, transparent films of the LuVO4:Eu3+, Bi3+ phosphor were obtained using sol–gel synthesis and the dip-coating method, which can be used in white light emission LEDs, using Bi3+ as a sensitizer of Eu3+ to improve red emission intensity under UV radiation. The films were characterized in terms of their crystalline structure and excitation/emission intensity spectra, as well as in relation to the effect of the annealing temperature and as a function of the Bi3+ content.

2. Materials and Methods

Synthesis

LuVO4:Eu3+, Bi3+ films were prepared using the sol–gel method and dip-coating technique. First, lutetium acetate (CH3CO2)3Lu•xH2O (99.9%, Sigma-Aldrich, Saint Louis, MO, USA) was dissolved in ethanol (99.5, C2H6O, Fermont, Monterrey, México) under vigorous magnetic stirring at 60 °C, obtaining a sol with a 0.15 M lutetium concentration. Later, a second sol was prepared from ammonium metavanadate NH4VO3 (99.0%, Fermont, Monterrey, México) dissolved in an ammonium hydroxide NH4OH (30%, Fermont, Monterrey, México) solution in distilled water (36 M), obtaining a sol with a 0.09 M vanadium concentration after vigorous magnetic stirring at 80 °C for 1 h. Later, the two sols were combined to obtain the LuVO4 sol-precursor, and nitric acid was added (0.5 M) to obtain a clear, homogeneous, and bluish solution after 2 h of stirring at 80 °C. On the other hand, to add the Eu3+ and Bi3+ ions to the lutetium vanadate sol, europium nitrate Eu(NO3)3·5H2O (99.5%, Alfa Aesar, Tewksbury, MA, USA) was incorporated in order to obtain 2.5 mol % Eu3+ samples, and bismuth was incorporated from a solution prepared by the dissolution of bismuth nitrate Bi(NO3)3·5H2O (98%, Sigma-Aldrich, Saint Louis, MO, USA) at a 1:1 molar ratio with ethanol:diethyleneglycol (C4H10O3, 98%, Sigma-Aldrich, Saint Louis, MO, USA), obtaining a Bi3+ 0.05 M concentration. Later, a fixed volume was incorporated into the lutetium vanadate sol to obtain a Bi3+ 0.5, 1, and 1.5 at. % doping-level. Finally, to modify the viscosity of the previously obtained sol and therefore obtain homogeneous films, diethylene glycol (HOCH2CH2)2O (99% Sigma-Aldrich, Saint Louis, MO, USA) was first added (5.7 M), followed by acetylacetone (6.8 M) CH3COCH2COCH3 (99% Sigma-Aldrich, Saint Louis, MO, USA), 0.8 M of HNO3, and 0.006 M Pluronic F-127 (C3H6O·C2H4O, 99% Sigma-Aldrich, Saint Louis, MO, USA). The final sol can be used to obtain powders or films; for powders, the sol was dried at 100 °C for 24 h and then annealed at several temperatures from 200 to 1000 °C for 3 h for each temperature.
For the dip-coating procedure, the sols were filtered using a 0.2 µm filter, and carefully cleaned silica glass substrates (QSI quartz, Quartz Scientific Inc., Lake County, OH, USA, refractive index = 1.417) were dipped into the prepared sol and pulled up at a constant rate of 4 cm·s−1. After each dipping, the films were dried at 100 °C for 10 min to remove the water and the most volatile organics components. Subsequently, the film was annealed at 300 °C and 500 °C for 10 min each to remove the organic remnants. The dipping cycle was repeated 3 times. Finally, to crystallize the LuVO4 phase, the films were annealed from 600 to 1000°C for 3 h.
The Fourier-transform infrared spectroscopy (FTIR) spectra of the samples were recorded in the range of 4000–400 cm−1, using FTIR (FTIR 2000, Perkin Elmer, Waltham, MA, USA) and the KBr pelleting technique. The phase composition of the powders was identified by X-ray diffraction (XRD) at room temperature on a powder diffractometer (Bruker D8Advance, Billerica, MA, USA), using Cu Kα radiation (1.5418 Å). The morphology studies were carried out with a Zeiss Supra 55VP scanning electron microscope. The luminescence study was carried out at room temperature with a fluorescence spectrophotometer Hitachi F-7000, equipped with a 150 W xenon lamp and an R955 photomultiplier tube.

3. Results and Discussion

Chemical Evolution and Structural Properties

To observe the xerogel chemical evolution during the annealing process in the sol–gel process for the synthesis of LuVO4: 2.5% at. Eu3+ and 1.5% at. Bi3+ sample, the FTIR study at different temperatures (from 200 to 1000 °C) was carried out (Figure 1). From 200 to 400 °C, strong bands at 3300 cm−1 (ν), 1650 cm−1 (δ), and 750 cm−1 (δ) are related to the O-H groups due to the presence of water and alcohols (ethanol and diethilenglycol). All these bands disappear at 600 °C, which indicates that this temperature is adequate to eliminate the O-H groups present in the samples.
In addition, bands associated with the carbonyl groups (-COOH) ~1725–1700 cm−1 are observed and attributed to the stretching bond of the lutetium acetate precursor employed, and are completely removed at 600 °C, which is an indicator of the moment of the beginning of crystallization process [8]. Bands around ~1430–1100 cm−1 are observed, which correspond to the symmetric and asymmetric vibrations of the C-H bonds, probably from the use of Pluronic F-127; however, as can be observed, they disappear at 700 °C [35,36]. Next, at 500 °C, bands related to the crystallization of LuVO4 can be observed at ~810–830 cm−1 and ~446–455 cm−1, which correspond to the V-O bonds of the vanadate group (VO23−) and Lu-O, respectively [2]. It is important to notice that, from 700 °C, all the organic groups have been completely removed from the LuVO4 system, and, therefore, all the luminescent properties correspond only to the ceramic film. The same results have been obtained for the 0, 0.5, and 1.0 at. % of Bi3+ (not shown). All the previous observations can be verified by the diffraction patterns obtained by XRD.
First, the XRD patterns of LuVO4: 2.5 at. % Eu3+, X at. % Bi3+ (X = 0.5, 1, 1.5) powders derived from the same sol used to prepare the films, annealed at 1000 °C, are presented in Figure 2. It can be observed that all samples are exactly in agreement with the corresponding 98-041-9281 ICSD from LuVO4, and the only appreciable impurity, observed at ≈ 26.05°, could be related to the formation of a secondary vanadate phase, LuVO3 (98-018-5830 ICSD). It is important to notice that no evidence of the possible formation of impurities from Bi3+ or Eu3+ ions has been detected. This result is expected for Eu3+ because, being a rare earth, it can occupy the position of Lu in the vanadate structure. On the other hand, regarding the case of bismuth, since it would also present a tetragonal structure forming BiVO4 and the same valence as Lu, it can be possible to have at least a partial dissolution of Bi in the structure. Furthermore, the inset in Figure 2 shows that the principal LuVO4 peak presents a gradual shift to the lower degree with the increase in Bi3+ content, implying the lattice expansion of LuVO4 when the Lu atoms in the lattices are gradually substituted by Bi, which has a bigger radius. Similar results have been reported previously by for the YVO4 system [37], who reported the appearance of the BiVO4 phase until the 4.0 at. % of Bi3+. Lu M. et al. [38]. also reported the absence of impurities for YVO4: 8% Eu, 10% Bi nanoparticles Finally, Zeng L. et al. [39] also reported the substitution of Lu3+ by Bi3+ ions in nanoparticles of LuVO4 obtained by a hydrothermal method.
On the other hand, Figure 3 shows the structural evolution of the LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ film as the annealing temperature increases. As observed, the crystallization process begins at 650 °C, which confirms the FTIR observation. The diffraction peaks can be ascribed to the diffraction pattern of tetragonal lutetium vanadate LuVO4 (98-041-9281, ICSD), space group I41/amd, and with the lattice systems a (7.01 Å), b (7.01 Å), and c (6.19 Å).
It is evident that the increment in the annealing temperature increases the crystallization of the films. At 600–650 °C, the broad peak centered at ~25–25.5° is typical for the amorphous structure obtained by the xerogel formation by the sol–gel process [40]. This amorphous structure disappears at 700 °C and with the increment in temperature, other peaks characteristic of the lutetium vanadate structure do not appear, which indicates that the films are highly oriented towards the (200) direction. In general, this result can probably be attributed to the films which, when deposited, tend to be structured first in the direction of lower surface energy, in this case, the (200) direction. Additionally, the presence of F127 probably tends to decrease the surface energy of the deposited xerogel; thus, the films maintain this preferential orientation. Furthermore, it has been stated [41] that the slow evaporation of high-molecular-weight alcohol, upon heating, like the diethylene glycol used in the present work, allows the structural orientation of the film before crystallization, preferable in the kinetically favored orientation along the c-axis (200). In order to quantify this observation, the preferential orientation parameter αhkl was calculated, which is defined as [42]:
h k l = I h k l I h k l
where Ihkl is the relative intensity of the corresponding diffraction. Table 1 shows the (200) preferential orientation-index α200 of the prepared films, and as observed, it practically does not present changes due to the effect of temperature increase, remaining at approximately 0.90, which shows that the orientation process was carried out from low temperatures and, once the material crystallized, the system tends to maintain that orientation. It is worth mentioning that when Pluronic F-127 acid is used, in some studies, it has been shown that it changes the size of the particles and their orientations [36,37,38,39,40,41,42,43], which is a very desirable property for luminescent films since oriented films tend to minimize the scattering of light during emission process, therefore incrementing its quantum efficiency [44,45]. On the other hand, the crystallite size does not present a significant difference with the increment in the annealing temperature, as can be observed in Table 1. In this regard, the crystallite size was computed by the Scherrer equation:
D = K λ β cos θ
where D is the crystallite size, K is a constant (0.9), λ is the X-ray wavelength (Cu-Kα = 0.15418 nm), β is the full width at half maximum (FWHM), and θ is the half diffraction angle of the peak centroid.

4. Morphological Studies

The obtained films were transparent and fracture-free, as can be observed in Figure 4a. Additionally, the luminescent macroscopic behavior under short and large UV- Vis excitation are presented in Figure 4b,c, respectively.
The scanning micrographs of the films, shown in Figure 5, correspond to LuVO4: 2.5 at. % Eu3+, 1.5 at. % Bi3+ film, deposited and annealed at 700 °C (Figure 5a) and 1000 °C (Figure 5b) at 10,000×.
Micrographs of films containing LuVO4: 2.5 at. % Eu3+, 1.5 at. % Bi3+ show a distribution of circular agglomerations of rice-like particles (according to NIST), morphology that is typical of vanadate [46,47,48], and as observed, these particles are distributed over the whole surface. The particle size was measured directly from the micrographs of 180 individual particles for both samples in such a way that it was possible to determine the average size is 579.3 ± 96.7 nm and 716.4 ± 123.8 nm, for the 700 °C and 1000 °C, respectively. The increment in the particle size can be explained since the energy excess with the higher temperature promotes the growth of particles. From Figure 4c, which corresponds to the film annealed at 1000 °C at 500×, it is possible to observe that these particle growth zones are distributed over the surface. Additionally, from the results shown in XRD, it was determined that the films grow preferentially in the (200) direction; therefore, the morphology of the particles corresponds to the growth in this direction. Similar morphologies have been observed in other yttrium vanadate films [49]. Finally, it is important to note that this preferential growth of the films is probably the product of the presence of Pluronic F-127 and diethylene glycol, which, as evaporate, would promote the formation of some pores [50,51].

5. Photoluminescence Study

Figure 6 shows the excitation and emission spectra corresponding to the LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films annealed at 1000 °C. The excitation spectrum shows a wide band at 250–380 nm, where the absorption of vanadate groups at λem = 615 nm can be assigned (VO43−) according to Liang et al. [22]. On the other hand, the emission spectrum of the films of the LuVO4:Eu3+, Bi3+ system is shown under excitation at 350 nm, and peaks can be observed at 594, 615, 650, and 700 nm that correspond to the transitions of Eu3+ ions: 5 D 0 7 F J (J = 1, 2, 3, 4), respectively, of which the one located at 615 nm is the most acute, therefore it is the one with the highest emission. The band of bismuth emission, located at 550 nm and generated with 350 nm excitation, decreases, as shown by the curve in the blue color, which has been scaled by ten. The best luminescent properties were reported in the case of the higher bismuth content, which could be explained by the charge transfer of the metal-metal bond (Bi3+, V5+) and the subsequent energy transfer to Eu3+ [20], and a better index of refraction, density, possibly a reduction of the pores, and a structural alignment of the films, which reduces the dispersion of the emissions and for that reason, better patterns are obtained [52].
On the other hand, the luminescence properties of bismuth ions were analyzed. Figure 7a presents the excitation and emission spectra of Bi3+ for a mono-doped LuVO4: 1.0 at. % Bi3+ film. As can be observed, congruent results are obtained when comparing with the observations of the co-doped sample, being the wavelength excitation at 332 nm and the greenish emission at 550 nm, products of the energy transition 1S03P1 for the excitation process and vice versa for the emission one. However, as also observed in Figure 6, for the co-doped samples, the greenish emission almost disappears, which clearly indicates an energy transfer process from Bi to Eu. Figure 7b presents the reduction of the 550 nm emission from Bi3+ in the co-doped samples, compared with the mono-doped one. As the Bi3+ content increases, the 550 nm emission almost disappears because of the energy transfer.
Therefore, the previous results establish an energy transfer process from Bi3+ to Eu3+. Similar behavior has been observed by Zhu et al. (2013) [53], and it is possible to propose the mechanism for LuVO4:Eu3+, Bi3+ presented in Figure 8. First, when LuVO4 is excited by the UV light, the energy could be absorbed by Bi3+ via the 1S03P1 transition. Next, there are two different types of excited states of Bi3+, 3P1, which relaxes to the 1S0 ground state, producing a broad luminescence band that could transfer energy to the Eu3+. Thereafter, the absorbed energy can be transferred to the 5D2 level of Eu3+, and the emission of Bi3+ partly quenches. At the same time, Eu3+ ions can also excite to 5D2 from the ground state, 7F0, and all the energy falls to the ground state, 5D0. Immediately, red light emission peaks at 615 nm occur because of the 5D0→F2 characteristic transitions of Eu3+ ions.
The influence of the Bi3+ content on the luminescence of Eu3+ is observed in Figure 9a at λexc = 350 nm. In this case, the intensity at 615 nm increases when the atomic content of bismuth is bigger, while the emission observed at 538 nm (corresponding with the bismuth emission) reduces in ratio with the increase in europium emission because an energy transfer occurs according to the energy diagram (Figure 8). For comparison purposes, the emission spectrum of an Eu-mono-doped sample has been added, with λexc = 308 nm, to observe that there is indeed a process of increment in the emission of Eu3+ due to the presence of bismuth. Furthermore, the emission of Eu3+ is highly dependent on the surroundings of the cation. It is possible to set the ratio R = I (5D07F2)/I (5D07F1) as a reference to the coordination state and symmetry since the intensity of the 5D07F1 magnetic dipole transition (centered at 594 nm) is independent of the surroundings of the Eu3+, whereas the intensity of the 5D07F2 electric dipole transition (centered at 618 nm) is very sensitive to site symmetry. If the R value is low, the Eu3+ cation tends to localize at a centrosymmetric site or a high-symmetry site. In contrast, when the R value is high, the Eu3+ cation tends to localize at a non-centrosymmetric site or a low-symmetry site [54]. For the LuVO4: 2.5% at. Eu3+, X% at. Bi3+, the R factor increases from 3.91→4.73→4.81 for X = 0, 0.5, 1, and 1.5, respectively, showing that the increment in the Bi3+ content promotes the emission of the low-symmetry sites.
On the other hand, it is possible to calculate the energy transfer efficiency (ET) between Bi3+→Eu3+ [55]:
ŋ E T = 1 I S I S 0
where I𝑆 is the luminous intensity of Bi3+ ions in the presence of Eu3+ ions, and I𝑆0 is the luminous intensity of Bi3+ ions in the absence of Eu3+ ions. The energy transfer efficiency, as observed in Figure 9b, shows that as the Bi3+ content increases, the energy transfer efficiency from Bi3+→Eu3+ also increases, reaching high values up to 71% in the 1.5 at. % sample. The ET values vary depending on the matrix and the co-doped ion. For comparison, for other proposed sensitizers for Eu3+, an ET of 68% was obtained for Tm3+→Eu3+ in LiLaSiO4:Tm3+, Eu3+ [56]; 78% for Tb3+→Eu3+ in Ba2SiO4 [57]; or 17% for Sm3+→Eu3+ on TeO2−GeO2−ZnO glasses [24].
On the other hand, the energy transfer from the sensitizers to the activators could be achieved through electric multipole interactions. According to the Dexter energy level transfer, the electric multipole interactions can be related to [3]:
I S I S 0 ŋ 0 ŋ   C n / 3
where ŋ0 and ŋ represent the quantum efficiencies of Bi3+ ions in the absence and presence of Eu3+ ions, respectively, the ratio 𝜂0/𝜂 can be calculated by the ratio of luminous intensities (I𝑆0/IS), and C is the total molar concentration of Bi3+ and Eu3+, whereas n = 6, 8, or 10 correspond to electric dipole–electric dipole, electric dipole–electric quadrupole, or electric quadrupole–electric quadrupole interactions, respectively. The corresponding plots are presented in Figure 10, and as observed, the best linear fit is exhibited when n = 10/3, implying that the most probable interaction for the energy transfer from Bi3+ to Eu3+ in the LuVO4 system occurs via quadrupole–quadrupole interactions. This electronic interaction is in good agreement with other Bi3+→Eu3+ energy transfer processes on other systems; for example, the same process has been observed in a Bi3+/Eu3+-doped SiO2–Al2O3–CaO–B2O3–La2O3–K2O glass, with an ET efficiency of 86% [58], and also on the Ca10(PO4)6F2:Bi3+, Eu3+, with an ET efficiency of 50% [53].
To corroborate the energy-transfer process, the Bi3+ fluorescence decay behavior has been studied. Figure 11 shows the fluorescence decay curves of LuVO4: Eu3+, Bi3+ films at 550 nm when the systems are excited at 350 nm. The best fit of the experimental data was obtained using the double exponential equation [59]:
I ( t ) = I 0 + A 1 e t τ 1 + A 2 e t τ 2
where I(t) is the luminescence intensity at time t, τ1 and τ2 are rapid and slow decay times for the exponential components, respectively, and A1 and A2 are constants.
That the double exponential equation presented the best fit can be explained since, from classical kinetics, there are two emitting species [60], which is the case in co-doped systems. S Dutta et al. [61] established that this can occur for systems with low concentrations of lanthanides and that it can be due to three factors: (1) the difference in the nonradiative probability of decays for lanthanide ions at or near the surface and lanthanide ions in the core; (2) an inhomogeneous distribution of the doping ions in the host material, leading to the variation in the local concentration; and (3) the transfer of excitation energy from the donor to lanthanide activators [62,63,64].
From the calculation of the decay values of τ1 and τ2, it can be observed that the values of the co-doped samples are smaller compared with the mono-doped Bi3+ sample, which corroborates the energy mechanism process between Bi3+ ions to Eu3+ ions. Furthermore, as the relation of Eu3+/Bi3+ increases (from 1.5 to 0.5 Bi3+ samples), the fluorescence lifetime decreases. This effect can be explained by the fact that, as this Eu3+/Bi3+ ratio increases, the Bi3+→Eu3+ ion distance must shorten, and the interaction gradually increases, so the fluorescence lifetime of Bi3+ ions decreases. The fast decay component (τ1) is derived from the fluorescence decay, whereas the slow decay component (τ2) possibly results from the luminescence caused by defects of phosphors on the surface and in the bulk structure [65].
On the other hand, the quality of the light emission is evaluated by the CIE 1931 color matching function, the correlated color temperature (CCT), and the color purity (CP), to understand the fluoresce center [66,67,68,69]. Table 2 presents the CIE color coordinates for different concentrations of bismuth; the mono-doped Bi film presents coordinates (0.3611, 0.5124) that are congruent with the results of Krishnan (2018), which studied bismuth silicate glasses for white light generation and bismuth silicate oxyfluoride glasses for LED applications. The calculated CCT values are below 3000 K in the co-doped films and, therefore, could be useful for warm light sources. When Bi is incorporated with 2.5 at. % Eu, the CP improved to 0.86. This highest CP and low CCT with the CIE coordinates moving to the red emission region is due to the energy transfer, as explained earlier, which is congruent with similar reports [70,71].
Table 2 also shows the RGB coordinates, calculated from the x and y coordinate values, and the hex values, from the RGB coordinates [72,73]. The hex code represents a useful tool to present the actual color display used in LED displays. As observed, all the co-doped samples present reddish color, whereas the Bi mono-doped sample has a greenish color.
Finally, the CIE chromaticity coordinates (Figure 12) proved that when the bismuth concentration is 1.5 at. %, the color of the emission is reddish, according to the diagram. A displacement from emitting a yellow-green color to a red color appears when the pure bismuth (100%) is combined with a constant europium concentration (2.5 at, %); however, it is evident that the best energy transfer occurs with a greater amount of bismuth because in the case of 0.5 and 1% europium the displacement is minimal between them, but emission changes to the color red. This result is congruent with that obtained by Zhang (2018) [28].

6. Conclusions

A new sol–gel synthesis was carried out to easily obtain films of the LuVO4 system at different concentrations of bismuth, LuVO4: 2.5% at. Eu3+, X% at. Bi3+ (X = 0, 0.5, 1, and 1.5), using the dip-coating method at low temperatures. The results show that the first bond of Lu-O is obtained from 500 °C, and it is possible to obtain completely structured films from 700 °C. Furthermore, it has been stated that, due to the presence of Pluronic F-123 and diethylene glycol during the sol–gel process, it has been possible to obtain highly oriented films along the (200) direction. Luminescence studies show that there is an energy transfer mechanism from Bi3+ to Eu3+, which effectively increases the light yield of the 5D07F2 transition of Eu3+. Furthermore, it was shown that the energy transfer increases as the Bi3+ concentration increases, reaching an energy transfer efficiency for Bi3+→Eu3+ of 71% with 1.5 at. % Bi3+, and it was observed that the energy transfer takes place via a quadrupole–quadrupole interaction. Finally, the actual real color from the co-doped samples is reddish, which is typical for the presence of Eu3+. The produced films could have different possible applications in white emission LEDs, mobile devices screens, lasers, and optoelectronic devices.

Author Contributions

Conceptualization, A.d.J.M.-R. and P.M.-F.; methodology, D.Y.M.-V. and G.G.P.-S.; investigation, P.M.-F., G.G.P.-S. and B.G.-P.; writing-original draft preparation, B.G.-P.; writing-reviewing and editing, A.d.J.M.-R., and D.Y.M.-V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CONACYT grant number 285600, and SIP-IPN grant number 20221369.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would also like to thank Luis Garces and A. Anahid Morales Hernandez for the editing work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zou, Y.; Hu, J.; Li, B.; Lin, L.; Li, Y.; Liu, F.; Li, X.Y. Tailoring the Coordination Environment of Cobalt in a Single-Atom Catalyst through Phosphorus Doping for Enhanced Activation of Peroxymonosulfate and thus Efficient Degradation of Sulfadiazine. Appl. Catal. B 2022, 312, 121408. [Google Scholar] [CrossRef]
  2. Chen, D.; Yu, Y.; Huang, P.; Lin, H.; Shan, Z.; Zeng, L.; Yang, A.; Wang, Y. Color-tunable luminescence for Bi3+/Ln3+: YVO4 (Ln = Eu, Sm, Dy, Ho) nanophosphors excitable by near-ultraviolet light. J. Eur. Ceram. Soc. 2010, 35, 859–869. [Google Scholar] [CrossRef]
  3. Chen, D.; Xiang, W.; Liang, X.; Zhong, J.; Yu, H.; Ding, M.; Lu, H.; Ji, Z. Advances in transparent glass-ceramic phosphors for white light-emitting diodes—A review. J. Eur. Ceram. Soc. 2014, 35, 859–869. [Google Scholar] [CrossRef]
  4. Zhang, G.; Huang, S.; Wang, F.; Yan, H. Layer-Dependent Electronic and Optical Properties of 2D Black Phosphorus: Fundamentals and Engineering. Laser Photonics Rev. 2021, 15, 2000399. [Google Scholar] [CrossRef]
  5. Ueda, J.; Tanabe, S. Review of luminescent properties of Ce3+-doped garnet phosphors: New insight into the effect of crystal and electronic structure. Opt. Mater. 2019, 1, 100018. [Google Scholar] [CrossRef]
  6. Gupta, A.D.; Rene, E.R.; Giri, B.S.; Pandey, A.; Singh, H. Adsorptive and photocatalytic properties of metal oxides towards arsenic remediation from water: A review. J. Environ. Chem. Eng. 2021, 9, 106376. [Google Scholar] [CrossRef]
  7. Smet, P.F.; Moreels, I.; Hens, Z.; Poelman, D. Luminescence in sulfides: A rich history and a bright future. Materials 2010, 3, 2834–2883. [Google Scholar] [CrossRef] [Green Version]
  8. Yang, L.; Li, G.; Zhao, M.; Yang, E.; Li, L. Lattice defect quenching effects on luminescence properties of Eu3+ doped YVO4 nanoparticles. J. Nanopart. Res. 2013, 1, 1996. [Google Scholar] [CrossRef]
  9. Stopikowska, N.; Runowski, M.; Woźny, P.; Lis, S.; Du, P. Generation of Pure Green Up-Conversion Luminescence in Er3+ Doped and Yb3+-Er3+ Co-Doped YVO4 Nanomaterials under 785 and 975 nm Excitation. J. Nanomater. 2022, 12, 799. [Google Scholar] [CrossRef]
  10. Piotrowski, W.; Dalipi, L.; Elzbieciak-Piecka, K.; Bednarkiewicz, A.; Fond, B.; Marciniak, L. Self-Referenced Temperature Imaging with Dual Light Emitting Diode Excitation and Single-Band Emission of AVO4: Eu3+ (A = Y, La, Lu, Gd) Nanophosphors. Adv. Photonics 2022, 3, 2100139. [Google Scholar] [CrossRef]
  11. Getz, M.N.; Nilsen, O.; Hansen, P.-A. Sensors for optical thermometry based on luminescence from layered YVO4: Ln3+ (Ln = Nd, Sm, Eu, Dy, Ho, Er, Tm, Yb) thin films made by atomic layer deposition. Sci. Rep. 2019, 9, 10247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Zhang, Y.; He, H.; Zhu, W.; Zheng, A. LuVO4:Ln3+ (Ln = Sm, Eu, Dy, Er, and Tm) with high uniform size and morphology: Controlled synthesis, growth mechanism and optical properties. CrystEngComm 2011, 13, 6471–6480. [Google Scholar] [CrossRef]
  13. Fang, C.; Wu, D.; Xie, J.; Li, Y.; Lou, Y.; Pu, Y. Preparation and luminescence properties of Ca2+-doped LaVO4: Dy3+ phosphors. J. Mol. Struct. 2022, 1266, 133497. [Google Scholar] [CrossRef]
  14. Cao, L.; Wang, Z.; Cai, P.; Hu, X.; Zhang, B.; Chu, X.; Li, S. Up-conversion luminescence properties of LuVO4: Yb3+/Tm3+/Er3+ submicron materials for high sensitivity temperature probing. J. Mater. Sci. Mater. Electron. 2021, 32, 28088–28097. [Google Scholar] [CrossRef]
  15. Tymiński, A.; Grzyb, T. Enhancement of the up-conversion luminescence in LaVO4 nanomaterials by doping with M2+, M4+ (M2+ = Sr2+, Ba2+, Mg2+; M4+= Sn4+) ions. J. Alloys Compd. 2019, 782, 69–80. [Google Scholar] [CrossRef]
  16. Bai, Y.; Bai, C.; Mo, G.; Guo, Y.; Jia, M.; Huo, X. LuVO4: Eu3+ nanorods: Synthesis and luminescence. J. Mater. Sci. Mater. Electron. 2018, 29, 3189–3193. [Google Scholar] [CrossRef]
  17. de Morales-Ramírez, A.J.; Domínguez, F.S.; Medina, D.; David, J.-V.; Margarita, G.-H.; Dorantes-Rosales, H. Synthesis and photoluminescent properties of Y2O3:Eu3+ thin films prepared from F127-containing solution. J. Ceram. Soc. Jpn. 2014, 122, 701–707. [Google Scholar] [CrossRef] [Green Version]
  18. de Morales-Ramírez, A.J.; García Hernández, M.; Yepez Ávila, J.; Carrillo Romo, F.; García Murillo, A.; de la Rosa, E.; Garibay Febles, V.; Reyes Miranda, J. Eu3+, Bi3+ codoped Lu2O3 nanopowders: Synthesis and luminescent properties. Mater. Res. 2013, 18, 1365–1371. [Google Scholar] [CrossRef]
  19. Li, B.; Li, H.; Yang, C.; Ji, B.; Lin, J.; Tomie, T. Theory of multiphoton photoemission disclosing excited states in conduction band of individual TiO2 nanoparticles. Chin. Phys. B 2021, 30, 114214. [Google Scholar] [CrossRef]
  20. Li, D.; Liu, C.; Jiang, L. Luminescent LuVO4:Ln3+ (Ln = Eu, Sm, Dy, Er) hollow porous spheres for encapsulation of biomolecules. Opt. Mater. 2015, 48, 18–24. [Google Scholar] [CrossRef]
  21. Krumpel, H.A.; Van der Kolk, E.; Cavalli, E.; Boutinaud, P.; Bettinelli, M.; Dorenbos, P. Lanthanide 4f-level location in AVO4:Ln3+ (A = La, Gd, Lu) crystals. J. Condens. Matter Phys. 2009, 21, 11550–11558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Leow, T.; Liu, H.; Hussin, R.; Ibrahim, Z.; Deraman, K.; Lintang, H.O.; Shamsuri, W.N.W. Effects of Eu3+ and Dy3+ doping or co-doping on optical and structural properties of BaB2Si2O8 phosphor for white LED applications. J. Rare Earths 2016, 34, 21–29. [Google Scholar] [CrossRef]
  23. Babu, J.K.; Rao, B.S.; Suresh, K.; Sridhar, M.; Murthy, K.V.R. 3Photoluminescence study of activator ions (Eu, Tb) co-doped in different host environments (CaO, CaSiO3, CaAl2O4 and CaSiAl2O6). Mater. Today Proc. 2019, 18, 2530–2539. [Google Scholar] [CrossRef]
  24. Alvarez-Ramos, M.E.; Carrillo-Torres, R.C.; Sánchez-Zeferino, R.; Caldiño, U.; Alvarado-Rivera, J. Co-emission and energy transfer of Sm3+ and/or Eu3+ activated zinc-germanate-tellurite glass as a potential tunable orange to reddish-orange phosphor. J. Non-Cryst. Solids 2019, 521, 119462. [Google Scholar] [CrossRef]
  25. Ren, Q.; Zhao, Y.; Wu, X.; Du, L.; Pei, M.; Hai, O. Luminescence characteristics of a novel Tm/Eu co-doped polychromatic tunable phosphor. J. Solid State Chem. 2021, 294, 121869. [Google Scholar] [CrossRef]
  26. Wen, F.; Sun, L.; Kim, J. Photoluminescent properties of Sb3+-doped and (Sb3+, Eu3+) co-doped YBO3 phosphors prepared via hydrothermal method and solid-state process. Front. Chem. Sci. Eng. 2011, 5, 429–434. [Google Scholar] [CrossRef]
  27. Yashodha, S.R.; Dhananjaya, N.; Yogananda, H.S.; Vinutha, K.; Ravikumar, C.R. Preparation and characterization of multifunctional Li+ co-doped LaOCl: Eu3+ for sensor, phosphor and catalytic applications. Inorg. Chem. Commun. 2022, 139, 109402. [Google Scholar] [CrossRef]
  28. Zhang, X.; Zhu, Z.; Guo, Z.; Mo, F.; Wu, Z.C. A zero-thermal-quenching and color-tunable phosphor LuVO4: Bi3+, Eu3+ for NUV LEDs. Dyes Pigm. 2018, 156, 67–73. [Google Scholar] [CrossRef]
  29. de Morales-Ramírez, A.J.; Margarita, G.-H.; Medina-Velázquez, D.; Ruiz-Guerrero, R.; Juárez-López, F.; Reyes-Miranda, J. Luminescence properties of co-doped Eu3+, Bi3+ Lu2O3/polyvinylpyrrolidone films. Coatings 2018, 8, 434. [Google Scholar] [CrossRef] [Green Version]
  30. Chen, T.; Zhang, H.; Luo, Z.; Liang, J.; Wu, X. Facile Preparation of YVO4: RE Films and the Investigation of Photoluminescence. Coatings 2022, 12, 461. [Google Scholar] [CrossRef]
  31. Tsotetsi, D.; Dhlamini, M.; Mbule, P. Sol-gel derived mesoporous TiO2: Effects of non-ionic co-polymers on the pore size, morphology, specific surface area and optical properties analysis. Results Mater. 2022, 14, 100266. [Google Scholar] [CrossRef]
  32. Arconada, N.; Castro, Y.; Durán, A. Photocatalytic properties in aqueous solution of porous TiO2-anatase films prepared by sol-gel process. Appl. Catal. A Gen. 2010, 385, 101–107. [Google Scholar] [CrossRef]
  33. Obregón, S.; Rodríguez-González, V. Photocatalytic TiO2 thin films and coatings prepared by sol-gel processing: A brief review. J. Sol-Gel Sci. Technol. 2021, 102, 125–141. [Google Scholar] [CrossRef]
  34. Yang, J.; Peterlik, H.; Lomoschitz, M.; Schubert, U. Preparation of mesoporous titania by surfactant-assisted sol-gel processing of acetaldoxime-modified titanium alkoxides. J. Non-Cryst. Solids 2010, 356, 1217–1227. [Google Scholar] [CrossRef]
  35. Youn, J.; Choi, J.H.; Lee, S.; Lee, S.W.; Moon, B.K.; Song, J.E.; Khang, G. Pluronic F-127/silk fibroin for enhanced mechanical property and sustained release drug for tissue engineering biomaterial. Materials 2021, 14, 1287. [Google Scholar] [CrossRef]
  36. de Morales-Ramírez, A.J.; García-Hernández, M.; Murillo, A.G.; Carrillo-Romo, F.; Palmerín, J.M.; Medina-Velázquez, D.; Carrera-Jota, M.L. Structural and Luminescence Properties of Lu2O3:Eu3+ F127 Tri-Block Copolymer Modified Thin Films Prepared by Sol-Gel Method. Materials 2013, 6, 713–725. [Google Scholar] [CrossRef] [Green Version]
  37. Xu, W.; Song, H.; Yan, D.; Zhu, H.; Wang, Y.; Xu, S.; Biao, D.; Liu, Y. YVO4: Eu 3+, Bi 3+ UV to visible conversion nano-films used for organic photovoltaic solar cells. J. Mater. Chem. 2011, 21, 12331–12336. [Google Scholar] [CrossRef]
  38. Luo, M.; Sun, T.Y.; Wang, J.H.; Yang, P.; Gan, L.; Liang, L.L.; Yu, X.F.; Gong, X.H. Synthesis of carboxyl-capped and bright YVO4: Eu, Bi nanoparticles and their applications in immunochromatographic test strip assay. Mater. Res. Bull. 2013, 48, 4454–4459. [Google Scholar] [CrossRef]
  39. Zeng, L.; Tang, Q.; Lin, B.; Zhou, H.; Liu, G.; Li, Y.; Chen, D. Fabrication of Bi3+/Ln3+: LuVO4 (Ln = Eu, Sm, Dy, Ho) nanophosphors and its color-tunable optical performance. J. Lumin. 2018, 194, 667–674. [Google Scholar] [CrossRef]
  40. Pawlik, N.; Szpikowska-Sroka, B.; Pisarska, J.; Goryczka, T.; Pisarski, W.A. Reddish-Orange Luminescence from BaF2: Eu3+ Fluoride Nanocrystals Dispersed in Sol-Gel Materials. Materials 2019, 12, 3735. [Google Scholar] [CrossRef]
  41. Amakali, T.; Daniel, L.S.; Uahengo, V.; Dzade, N.Y.; De Leeuw, N.H. Structural and optical properties of ZnO thin films prepared by molecular precursor and sol-gel methods. Crystals 2020, 10, 132. [Google Scholar] [CrossRef] [Green Version]
  42. Lee, J.H.; Hwang, K.S.; Kim, T.S. The microscopic origin of residual stress for flat self-actuating piezoelectric cantilevers. Nanoscale Res. Lett. 2011, 6, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Soli, J.; Kachbouri, S.; Elaloui, E.; Charnay, C. Role of surfactant type on morphological, textural, optical, and photocatalytic properties of ZnO nanoparticles obtained by modified sol-gel. J. Sol-Gel Sci. Technol. 2021, 100, 271–285. [Google Scholar] [CrossRef]
  44. Feng, Q.; Qian, Y.; Wang, H.; Hou, W.; Peng, X.; Xie, S.; Wang, S.; Xie, L. Donor arylmethylation toward horizontally oriented TADF emitters for efficient electroluminescence with 37% external quantum efficiency. Adv. Opt. Mater. 2022, 10, 2102441. [Google Scholar] [CrossRef]
  45. Song, J.; Kim, K.H.; Kim, E.; Moon, C.K.; Kim, Y.H.; Kim, J.J.; Yoo, S. Lensfree OLEDs with over 50% external quantum efficiency via external scattering and horizontally oriented emitters. Nat. Commun. 2018, 9, 3207. [Google Scholar] [CrossRef]
  46. Murugan, C.; Subramani, K.; Subash, R.; Sathish, M.; Pandikumar, A. High-Performance High-Voltage Symmetric Supercapattery Based on a Graphitic Carbon Nitride/Bismuth Vanadate Nanocomposite. Energy Fuels 2020, 34, 16858–16869. [Google Scholar] [CrossRef]
  47. Liang, Z.; Cao, Y.; Qin, H.; Jia, D. Low-heating solid-state chemical synthesis of monoclinic scheelite BiVO4 with different morphologies and their enhanced photocatalytic property under visible light. Mater. Res. Bull. 2016, 84, 397–402. [Google Scholar] [CrossRef]
  48. Kumar, V.; Khan, A.F.; Chawla, S. Intense red-emitting multi-rare-earth doped nanoparticles of YVO4 for spectrum conversion towards improved energy harvesting by solar cells. J. Phys. D 2013, 46, 365101. [Google Scholar] [CrossRef]
  49. Brack, P.; Sagu, J.S.; Peiris, T.N.; McInnes, A.; Senili, M.; Wijayantha, K.U.; Marken, F.; Selli, E. Aerosol—assisted CVD of bismuth vanadate thin films and their photoelectrochemical properties. Chem. Vap. Depos. 2015, 21, 41–45. [Google Scholar] [CrossRef] [Green Version]
  50. Roy, S.; Ghosh, S.P.; Pradhan, D.; Sahu, P.K.; Kar, J.P. Morphological and electrical study of porous TiO2 films with various concentrations of Pluronic F-127 additive. J. Porous Mater. 2021, 28, 231–238. [Google Scholar] [CrossRef]
  51. González-Penguelly, B.; de Morales-Ramírez, A.J.; Rodríguez-Rosales, M.G.; Rodríguez-Nava, O.; Carrera-Jota, M.L. New infrared-assisted method for sol-gel derived ZnO: Ag thin films: Structural and bacterial inhibition properties. Mater. Sci. Eng. C 2017, 78, 833–841. [Google Scholar] [CrossRef] [PubMed]
  52. Kang, F.; Peng, M.; Zhang, Q.; Qiu, J. Abnormal Anti-Quenching and Controllable Multi-Transitions of Bi3+ Luminescence by Temperature in a Yellow-Emitting LuVO4:Bi3+ Phosphor for UV-Converted White LEDs. Chem. Eur. J. 2014, 20, 11522–11530. [Google Scholar] [CrossRef] [PubMed]
  53. Zhu, H.; Xia, Z.; Liu, H.; Mi, R.; Hui, Z. Luminescence properties and energy transfer of Bi3+/Eu3+ codoped Ca10(PO4)6F2 phosphors. Mater. Res. Bull. 2013, 48, 3513–3517. [Google Scholar] [CrossRef]
  54. Narro-García, R.; Desirena, H.; López-Luke, T.; Guerrero-Contreras, J.; Jayasankar, C.K.; Quintero-Torres, R.; Rosa, E.D.L. Spectroscopic properties of Eu3+/Nd3+ co-doped phosphate glasses and opaque glass-ceramics. Opt. Mater. 2015, 46, 34–39. [Google Scholar] [CrossRef]
  55. Deopa, N.; Kaur, S.; Prasad, A.; Joshi, B.; Rao, A.S. Spectral studies of Eu3+ doped lithium lead alumino borate glasses for visible photonic applications. Opt. Laser Technol. 2018, 108, 434–440. [Google Scholar] [CrossRef]
  56. Wu, X.; Du, L.; Ren, Q.; Hai, O. Luminescence properties and energy transfer of Tm3+–Eu3+ double-doped LiLaSiO4 phosphors. J. Mater. Sci. Mater. 2021, 32, 17662–17673. [Google Scholar] [CrossRef]
  57. Bispo-Jr, A.G.; Lima, S.A.M.; Pires, A.M. Energy transfer between terbium and europium ions in barium orthosilicate phosphors obtained from sol-gel route. J. Lumin. 2018, 199, 372–378. [Google Scholar] [CrossRef] [Green Version]
  58. Giraldo, O.G.; Fei, M.; Wei, R.; Teng, L.; Zheng, Z.; Guo, H. Energy transfer and white luminescence in Bi3+/Eu3+ co-doped oxide glasses. J. Lumin. 2020, 219, 116918. [Google Scholar] [CrossRef]
  59. Ruelle, N.; Fouassier, M.P.T. Cathodoluminescent properties and energy transfer in red calcium sulfide phosphors (CaS: Eu, Mn). Jpn. J. Appl. Phys. 1992, 31, 2786. [Google Scholar] [CrossRef]
  60. Delgado, T.; Gartmann, N.; Waltfort, B.; LaMattina, F.; Pollnau, M.; Rosspeintne, A.; Afshani, J.; Olchowka, J.; Hageman, H. Fundamental Loading-Curve Characteristics of the Persistent Phosphor SrAl2O4:Eu2+,Dy3+,B3+: The Effect of Temperature and Excitation Density. Adv. Photonics Res. 2022, 4, 2100179. [Google Scholar] [CrossRef]
  61. Dutta, S.; Soma, S.; Sharma, K. Excitation spectra and luminescence decay analysis of K+ compensated Dy3+ doped CaMoO4 phosphors. RSC Adv. 2015, 10, 7380–7387. [Google Scholar] [CrossRef]
  62. Ninjbadgar, T.; Garnweitner, G.; Börger, A.; Goldenberg, L.; Sakhno, O.; Stumpe, J. Synthesis of Luminescent ZrO2:Eu3+ Nanoparticles and Their Holographic Sub-Micrometer Patterning in Polymer Composites. Adv. Funct. Mater. 2009, 19, 1819–1825. [Google Scholar] [CrossRef]
  63. Byrd, R.H.; Hribar, M.E.; Nocedal, J. An interior point algorithm for large-scale nonlinear programming. SIAM J. Optim. 1999, 9, 877–900. [Google Scholar] [CrossRef]
  64. Coleman, T.F.; Li, Y. On the convergence of interior-reflective Newton methods for nonlinear minimization subject to bounds. Math. Program. 1994, 67, 189–224. [Google Scholar] [CrossRef]
  65. Zhou, W.; Deng, S.; Rong, C.; Xie, Q.; Lian, S.; Zhang, J.; Lia, C.; Yu, L. Synthesis, crystal structure and luminescence of a near ultraviolet-green to red spectral converter BaY2S4:Eu2+, Er3+. RSC Adv. 2013, 3, 16781–16787. [Google Scholar] [CrossRef]
  66. Broadbent, A.D. A critical review of the development of the CIE1931 RGB colormatching functions. Color Res. Appl. 2004, 29, 267–272. [Google Scholar] [CrossRef]
  67. Schanda, J. Colorimetry: Understanding the CIE System; Wiley-Interscience: Hoboken, NJ, USA, 2007. [Google Scholar]
  68. Jin, M.; Li, X.; Yan, F.; Chen, W.; Jiang, L.; Zhang, X. The Effects of Low-Color-Temperature Dual-Primary-Color Light-Emitting Diodes on Three Kinds of Retinal Cells. J. Photochem. Photobiol. 2021, 214, 112099. Available online: https://n9.cl/nta6t (accessed on 8 December 2022). [CrossRef]
  69. Wang, S.; Wang, G.; Liang, X.; Li, D.; Zhang, Z.; Guo, K.; Li, J.; Miao, Y.; Wang, H. Dyes and pigments: Introduction of chlorine into spiro[fluorene-9,9′-xanthene] based luminophore for high color purity single-molecule white emitter. Dye. Pigment. 2022, 204, 110450. Available online: https://n9.cl/jb25m (accessed on 8 December 2022). [CrossRef]
  70. Krishnan, M.L.; Neethish, M.M.; Kumar, V.R.K. Structural and optical studies of rare earth-free bismuth silicate glasses for white light generation. J. Lumin. 2018, 201, 442–450. [Google Scholar] [CrossRef]
  71. Yin, Y.; Wang, Z.; Zhu, M.; Zhang, Y.; Li, J.; Dou, C.; Sun, S.; Teng, B.; Che, Y.; Zhong, D. Full-visible-spectrum emission with high color rendering index and low correlated color temperature enabled by a sin-gle-phased phosphor of α-Sr2V1.98P0.02O7: 0.5% Eu3+. Mater. Res. Bull. 2021, 141, 111344. Available online: https://n9.cl/lwvsx (accessed on 8 December 2022). [CrossRef]
  72. Ohta, N.; Robertson, A. Colorimetry Fundamentals and Applications; John Wiley & Sons Ltd: Hoboken, NJ, USA, 2005; pp. 63–114. [Google Scholar]
  73. Fu, K.; He, J.; Yang, X. Few-shot learning-based RGB-D salient object detection: A case study. Neurocomputing 2022, 512, 142–152. Available online: https://n9.cl/j0s2zy (accessed on 8 December 2022). [CrossRef]
Figure 1. FTIR spectra of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ chemical evolution.
Figure 1. FTIR spectra of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ chemical evolution.
Materials 16 00146 g001
Figure 2. The XRD patterns of LuVO4: 2.5 at. % Eu3+, X at. % Bi3+ (X = 0.5, 1, 1.5) powders prepared from the same sol of the films. The inset is the partial magnification of the (200) peak. On X= 1.5 sample, at ≈ 26.05° (*) the secondary phase LuVO3 it’s observed.
Figure 2. The XRD patterns of LuVO4: 2.5 at. % Eu3+, X at. % Bi3+ (X = 0.5, 1, 1.5) powders prepared from the same sol of the films. The inset is the partial magnification of the (200) peak. On X= 1.5 sample, at ≈ 26.05° (*) the secondary phase LuVO3 it’s observed.
Materials 16 00146 g002
Figure 3. Structural evolution of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films by XRD analysis.
Figure 3. Structural evolution of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films by XRD analysis.
Materials 16 00146 g003
Figure 4. Photography of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films (a) without, (b) with UV-Vis short excitation (254 nm), and (c) with long UV-Vis on excitation (380 nm).
Figure 4. Photography of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films (a) without, (b) with UV-Vis short excitation (254 nm), and (c) with long UV-Vis on excitation (380 nm).
Materials 16 00146 g004
Figure 5. SEM micrographs of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films films: (a) after annealing for 3 h at 700 °C, 10,000×, (b) after annealing at 1000 °C for 3 h, 10,000×, and (c) after annealing at 1000 °C for 3 h, 500×.
Figure 5. SEM micrographs of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films films: (a) after annealing for 3 h at 700 °C, 10,000×, (b) after annealing at 1000 °C for 3 h, 10,000×, and (c) after annealing at 1000 °C for 3 h, 500×.
Materials 16 00146 g005
Figure 6. Excitation and emission spectra of LuVO4 films with 2.5 at. % Eu3+ and 1.5 at. % Bi3+ after 3 h of annealing at 1000 °C.
Figure 6. Excitation and emission spectra of LuVO4 films with 2.5 at. % Eu3+ and 1.5 at. % Bi3+ after 3 h of annealing at 1000 °C.
Materials 16 00146 g006
Figure 7. (a) Excitation and emission spectra of Bi3+ and (b) Intensity of 550 nm Bi3+ 1S0 3P1 transition.
Figure 7. (a) Excitation and emission spectra of Bi3+ and (b) Intensity of 550 nm Bi3+ 1S0 3P1 transition.
Materials 16 00146 g007
Figure 8. Energy level diagram of Bi3+ and Eu3+ in LuVO4.
Figure 8. Energy level diagram of Bi3+ and Eu3+ in LuVO4.
Materials 16 00146 g008
Figure 9. (a) Emission spectra of LuVO4: 2.5 at. % Eu3+, X Bi at. % (X = 0.5, 1.0 and 1.5) at λexc = 350 nm. For the mono-doped sample, λexc = 302 nm for comparison purposes. (b) Influence of the Bi3+ content on the luminescence of Eu3+ and energy transfer efficiency of Bi3+→Eu3+.
Figure 9. (a) Emission spectra of LuVO4: 2.5 at. % Eu3+, X Bi at. % (X = 0.5, 1.0 and 1.5) at λexc = 350 nm. For the mono-doped sample, λexc = 302 nm for comparison purposes. (b) Influence of the Bi3+ content on the luminescence of Eu3+ and energy transfer efficiency of Bi3+→Eu3+.
Materials 16 00146 g009
Figure 10. Fitting relationship of IS0/IS of Bi3+ on C(Bi3++Eu3+)6/3 for a dipole–dipole interaction; C (Bi3++Eu3+)8/3 for a dipole–quadrupole interaction, and C(Bi3++Eu3+)10/3 for a quadrupole–quadrupole interaction.
Figure 10. Fitting relationship of IS0/IS of Bi3+ on C(Bi3++Eu3+)6/3 for a dipole–dipole interaction; C (Bi3++Eu3+)8/3 for a dipole–quadrupole interaction, and C(Bi3++Eu3+)10/3 for a quadrupole–quadrupole interaction.
Materials 16 00146 g010
Figure 11. Decay curves of LuVO4:Eu3+, Bi3+ films.
Figure 11. Decay curves of LuVO4:Eu3+, Bi3+ films.
Materials 16 00146 g011
Figure 12. CIE chromaticity coordinates for the LuVO4: 2.5 at. % Eu3+, X at. % Bi3+ (X = 0.5, 1, 1.5) films and a Bi3+ mono-doped sample.
Figure 12. CIE chromaticity coordinates for the LuVO4: 2.5 at. % Eu3+, X at. % Bi3+ (X = 0.5, 1, 1.5) films and a Bi3+ mono-doped sample.
Materials 16 00146 g012
Table 1. Crystallite size and preferential orientation index of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films as a function of annealing temperature.
Table 1. Crystallite size and preferential orientation index of LuVO4: 2.5% at. Eu3+, 1.5% at. Bi3+ films as a function of annealing temperature.
SystemAnnealing
Temperature (°C)
Crystallite Size (nm)Preferential Orientation Index α200
LuVO4: Eu3+ 2.5 at. %, Bi3+ 1.5 at. %70068.2 ± 6.20.92
80071.6 ± 7.80.91
90066.9 ± 4.20.91
100080.8 ± 6.10.90
Table 2. Chromaticity coordinates (x, y), correlated color temperature (CCT), color purity (CP), RGB coordinates, and Hex color.
Table 2. Chromaticity coordinates (x, y), correlated color temperature (CCT), color purity (CP), RGB coordinates, and Hex color.
SampleCIEE Color
Coordinates (x, y)
CTTKCPRGBHexHex Color
RGB
Eu3+
mono-doped
(0.6423, 0.3416)10260.92255340#FF2200Materials 16 00146 i001
Bi3+ mono-doped(0.3611, 0.5124)49390.5918925580#BDFF50Materials 16 00146 i002
Eu3+ 2.5 at. %, Bi3+ 0.5 at. %(0.5713, 0.3734)16200.782559733#FF6121Materials 16 00146 i003
Eu3+ 2.5 at. %, Bi3+ 1.0 at. %(0.5721, 0.3724)16200.792559633#FF0021Materials 16 00146 i004
Eu3+ 2.5 at. %, Bi3+ 1.5 at. %(0.5999, 0.3684)16800.86255810#FF5100Materials 16 00146 i005
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

González-Penguelly, B.; Pérez-Sánchez, G.G.; Medina-Velázquez, D.Y.; Martínez-Falcón, P.; Morales-Ramírez, A.d.J. Luminescence Properties and Energy Transfer of Eu3+, Bi3+ Co-Doped LuVO4 Films Modified with Pluronic F-127 Obtained by Sol–Gel. Materials 2023, 16, 146. https://doi.org/10.3390/ma16010146

AMA Style

González-Penguelly B, Pérez-Sánchez GG, Medina-Velázquez DY, Martínez-Falcón P, Morales-Ramírez AdJ. Luminescence Properties and Energy Transfer of Eu3+, Bi3+ Co-Doped LuVO4 Films Modified with Pluronic F-127 Obtained by Sol–Gel. Materials. 2023; 16(1):146. https://doi.org/10.3390/ma16010146

Chicago/Turabian Style

González-Penguelly, Brenely, Grethell Georgina Pérez-Sánchez, Dulce Yolotzin Medina-Velázquez, Paulina Martínez-Falcón, and Angel de Jesús Morales-Ramírez. 2023. "Luminescence Properties and Energy Transfer of Eu3+, Bi3+ Co-Doped LuVO4 Films Modified with Pluronic F-127 Obtained by Sol–Gel" Materials 16, no. 1: 146. https://doi.org/10.3390/ma16010146

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