Next Article in Journal
Samarium-Doped Lead Magnesium Niobate-Lead Titanate Ceramics Fabricated by Sintering the Mixture of Two Different Crystalline Phases
Next Article in Special Issue
Effect of TiO2 on Pd/La2O3-CeO2-Al2O3 Systems during Catalytic Oxidation of Methane in the Presence of H2O and SO2
Previous Article in Journal
The Microstructure, Solidification Path, and Microhardness of As-Cast Ni-Al-Cr-Os Alloys in a Ni-Rich Region
Previous Article in Special Issue
Machine Learning Prediction of the Redox Activity of Quinones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Network Structure and Luminescent Properties of ZnO–B2O3–Bi2O3–WO3:Eu3+ Glasses

1
Institute of General and Inorganic Chemistry, Bulgarian Academy of Sciences, Acad. G. Bonchev Str., bld. 11, 1113 Sofia, Bulgaria
2
Centre for Energy Research, 29-33 Konkoly Thege Street, 1121 Budapest, Hungary
3
Institute of Optical Materials and Technologies “Acad. Jordan Malinowski”, Bulgarian Academy of Sciences, Acad. G. Bonchev Str., bld. 109, 1113 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Materials 2023, 16(20), 6779; https://doi.org/10.3390/ma16206779
Submission received: 21 September 2023 / Revised: 13 October 2023 / Accepted: 16 October 2023 / Published: 20 October 2023

Abstract

:
In this study, we investigated the influence of Bi2O3 and WO3 on both structure and optical properties of 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3; x = 1, 5, 10 glasses doped with 0.5 mol% Eu2O3. IR spectroscopy revealed the presence of trigonal BØ3 units connecting superstructural groups, [BØ2O] metaborate groups, tetrahedral BØ4 units in superstructural groupings (Ø = bridging oxygen atom), borate triangles with nonbridging oxygen atoms, [WO4]2− tetrahedral, and octahedral WO6 species. Neutron diffraction experimental data were simulated by reverse Monte Carlo modeling. The atomic distances and coordination numbers were established, confirming the short-range order found by IR spectra. The synthesized glasses were characterized by red emission at 612 nm. All findings suggest that Eu3+ doped zinc borate glasses containing both WO3 and Bi2O3 have the potential to serve as a substitute for red phosphor with high color purity.

1. Introduction

Currently, white-light-emitting diodes are being investigated extensively as the next-generation solid-state light source owing to the advantages they bring to the table, including safety, an environmentally friendly nature, high stability, low power consumption, and long operational lifetime [1]. One of the ways in which white light can be obtained is by combining tri-color phosphors, such as ZnS:Cu+, Al3+ (green) [2], BaMgAl10O17:Eu2+ (blue) [3], and Y2O2S: Eu3+ (red) [4], coated on InGaN-based LED chip, emitting around 400 nm (near-UV). However, the red-emitting phosphor shows lower efficiency (eight times lower) compared to the blue and green phosphors, as well as exhibiting chemical instability under UV radiation which may cause environmental pollution due to the release of sulfide gas. The other commercially applied red phosphors, such as Y2O3:Eu3+ and YVO4:Eu3+, also cannot achieve high emission efficiency [5]. Therefore, red-emitting phosphor with chemical and thermal stability and high efficiency upon near-UV excitation remains to be found.
Europium (III) ion is being considered as a suitable activator for red emission, resulting from its 5D07Fj (j = 0–4) transitions in the visible range [6]. Unfortunately, Eu3+-doped materials cannot be efficiently excited by the present LED chips, because its excitation peaks are weak in nature due to parity-forbidden f–f transitions. Searching for host materials that can overcome the weak Eu3+ absorptions is important for achieving high excitation and emission efficiency of the red luminescence. A possibility is introducing sensitizers into the host composition, such as Ce3+, Bi3+, Tb3+, etc. [7,8]. It is well known that the luminescent properties of Eu3+-doped materials can be modified by changing the host structure and composition. Glass materials are suitable matrixes for doping with lanthanide ions due to their chemical stability, high optical homogeneity, absence of absorbing particles, and low nonlinear refractive indices. Among many potential glass materials for luminescence applications, binary ZnO–B2O3 glasses have been attracting continuous scientific interest. Homogeneous binary zinc–borate glasses are formed in a very narrow range of compositions because of the existence of a very large region of immiscibility of two liquids in a ZnO–B2O3 system [9]. However, these glasses are characterized by good chemical and thermal stability, high mechanical strength, low dispersion, and low glass transition temperature. They possess high transparency (up to 90%) from the visible to mid-infrared region of the spectrum [10]. Eu3+-doped zinc borate glasses yield very strong orange/red photoluminescence by UV excitation, especially for low europium concentrations (<1019 cm3) [11]. It has been reported that the addition of WO3 and/or Bi2O3 to ZnO–B2O3 glasses induces the expansion of the glass-forming region and also lowers the phonon energy [12,13]. In our recent works, we reported, for the first time, the preparation of tungsten-containing ZnO–B2O3 glasses doped with Eu3+ active ion and their luminescent properties [14,15]. The obtained results from glass structure, physical, thermal, and optical properties indicate the suitability of the 50ZnO:40B2O3:10WO3 glass network for the luminescence performance of Eu3+ ions. The positive effect of the addition of WO3 on the luminescence intensity is proven by the stronger Eu3+ emission of the zinc–borate glass containing WO3 compared to the WO3-free zinc–borate glass, a phenomenon engendered mainly by the energy transfer from tungstate groups to the Eu3+ ions (sensitizing effect). The most intense luminescence peak observed at 612 nm and the high-integrated emission intensity ratio (R) of the 5D07F2/5D07F1 transitions at 612 nm and 590 nm of 5.77 suggest that the glasses have the potential for red emission materials.
Another desirable component for luminescent glass hosts is Bi2O3 oxide, commonly used as an activator, emitting in the spectral region of 380–700 nm due to 3P11S0 transition upon NUV excitation. Among many studies, the Bi3+ ion is also recognized as a favored sensitizer, which can greatly enhance the luminescence of the rare-earth ions (Eu3+, Sm3+, Tb3+) through resonant energy transfer [16,17]. High bright red emission in Eu3+ containing zinc–borate glasses codoped with Bi3+ was observed, enhanced by 346 nm excitation (1S0-3P1 of Bi3+ ions) due to the sensitization effect of Bi3+ codopant [18]. Zinc bismuth borate glasses doped with different Eu3+ concentrations (1, 3, 5, 7, and 9 mol%) were prepared, and the systematic analysis of the results suggested that the glass doped with a Eu3+ concentration of 5 mol% is suitable for LED and display device applications [19].
More recently, we prepared zinc–borate glasses modified with Bi2O3. Bulk, transparent, dark brownish glasses with composition 50ZnO:(40 − x)B2O3:10Bi2O3:0.5Eu2O3:xWO3, x = 0 and 0.5, were synthesized. The obtained structural and optical data indicate that a zinc–borate glass network containing Bi2O3 provides highly asymmetric sites of Eu3+ ions, leading to high emission intensity. Moreover, the presence of WO3 also leads to the increase in emission intensity of the rare-earth Eu3+ ion, as a result of the nonradiative energy transfer from the glass host to the active ion [20]. These data above show that the ZnO–B2O3 glass system containing both bismuth and tungstate oxides is a particularly interesting host for the europium ions in red phosphors applications.
Here, we continued our investigations by preparing such a glass composition with increasing WO3 content and with the addition of low Bi2O3 concentrations (1 mol%) in order to meet the requirement of colorless glasses for optical application. The aim was to obtain bulk, colorless glasses with compositions 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%, and to establish the influence of Bi2O3 and WO3 on glass formation, structure, and optical properties.

2. Materials and Methods

2.1. Sample Preparation

Glasses of the compositions in mol% 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3; x = 1, 5, 10, doped with 0.5 mol% Eu2O3 were obtained by applying the melt quenching method, using reagent-grade ZnO (Merck KGaA, Amsterdam, The Netherlands), WO3 (Merck KGaA, Darmstadt, Germany), B2O3 (SIGMA-ALDRICH, St. Louis, MO, USA), Bi2O3 (Alfa Aesar, Karlsruhe, Germany), and Eu2O3 (SIGMA-ALDRICH, St. Louis, MO, USA) as raw materials. B2O3 enriched with 11B isotope (99.6%) was used in order to avoid the high neutron absorption cross-section of the 10B isotope. Further in the text, samples’ names are abbreviated to ZBBW1:Eu, ZBBW5:Eu, and ZBBW10:Eu, where the number refers to the WO3 content in the compositions. The homogenized batches were melted at 1200 °C for 20 min in a platinum crucible in air. The melts were cast into graphite molds to obtain bulk glass samples. Glasses obtained by us earlier with the same compositions (50ZnO:(50 − x)B2O3:xWO3; x = 1, 5, 10) doped with 0.5 mol% Eu2O3 without Bi2O3, and denoted as ZBW1:Eu, ZBW5:Eu, and ZBW10:Eu, were also included here in order to establish the influence of bismuth addition on the structure and luminescence properties of the new glass compositions. The glasses without Bi2O3 containing 1 and 5 mol% WO3 were reported in the 6th International Conference on Optics, Photonics, and Lasers (OPAL’ 2023), while the glass with the highest WO3 amount of 10 mol% was represented in ref. [15].

2.2. Characterization Techniques

The phase formation of the samples was established by X-ray phase analysis with a Bruker D8 advance diffractometer, Karlsruhe, Germany, using Cu Kα radiation in the 10 < 2θ < 60 range. The thermal stability of the obtained glasses was examined by differential scanning calorimetry (DSC) with a Netzsch 404 F3 Pegasus instrument, Selb, Germany, in the temperature range 25–750 °C at a heating rate of 10 K/min in an argon atmosphere. The density of the obtained glasses at room temperature was measured by the Archimedes principle using toluene (ρ = 0.867 g/cm3) as an immersion liquid on a Mettler Toledo electronic balance of sensitivity 10−4 g. The IR spectra of the glasses were measured using the KBr pellet technique on a Nicolet-320 FTIR spectrometer, Madison, WI, USA, with a resolution of ±4 cm−1, by collecting 64 scans in the range 1600–400 cm−1. A random error in the center of the IR bands was found to be ±3 cm−1. The EPR analyses were carried out in the temperature range 120–295 K in X band at frequency 9.4 GHz on a spectrometer Bruker EMX Premium, Karlsruhe, Germany. Optical absorption spectra (UV–VIS–NIR) in the range 190–1500 nm were obtained with an error ˂ 1% using a commercial double-beam spectrometer (UV-3102PC, Shimadzu, Kyoto, Japan). Photoluminescence (PL) excitation and emission spectra at room temperature for all glasses were measured with a Spectrofluorometer FluoroLog3–22, Horiba JobinYvon, Longjumeau, France. Neutron diffraction measurements were carried out in the momentum transfer range, Q = 0.45–9.8 Å−1, for 24 h using neutrons of de Broglie wavelength, λ = 1.069 Å, at the 2-axis PSD diffractometer of Budapest Neutron Centre. The powder glass samples were mounted in a thin-walled cylindrical vanadium can with a diameter of 8 mm for the neutron diffraction experiments. The neutron diffraction data were corrected for detector efficiency, background scattering, and absorption effects, and normalized with vanadium [21]. The total structure factor, S(Q), was calculated using local software packages.

2.3. The Reverse Monte Carlo Simulation

Reverse Monte Carlo (RMC) simulations were performed on neutron diffraction datasets of the experimental total structure factor, S(Q), to determine the short-range structural properties of glasses by using RMC++ software (https://www.szfki.hu/~nphys/rmc++/downloads.html, accessed on 15 October 2023) [22]. The RMC technique minimizes the squared difference between the experimental S(Q) and the simulated one from a 3-dimensional atomic configuration by using the following equations:
S ( Q ) = i , j k w i j S i j ( Q )
S i j ( Q ) = 1 + 4 π ρ 0 Q 0 r m a x r g i j ( r ) 1 sin Q r d r
w i j = c i c j b i b j i , j k c i b j 2
where ci and bi are the molar fraction and coherent neutron scattering length for atoms of type i, the Sij(Q) denotes the partial structure factors, and wij are the neutron scattering weight factors for the 21 atomic pairs for the ZBBW:Eu series (explanation: k = 5, thus k(k + 1)/2 = 15 different atomic pairs are present). RMC simulations were used to generate partial atomic pair correlation function, gij(r), and coordination number distributions. The simulation was started with an initial random configuration by building a box that contained 10.000 atoms of Zn, B, Bi, W, Eu, and O, with the atomic density, ρo, values of 0.0947 Å3, 0.0919 Å3, and 0.0888 Å3 for the samples ZBBW1:Eu, ZBBW5:Eu, and ZBBW10:Eu, respectively. The RMC model box lengths for the three samples were 23.63 Å, 23.87 Å, and 24.14 Å for ZBBW1:Eu, ZBBW5:Eu, and ZBBW10:Eu samples, respectively.
In the RMC simulation procedure, constraints were used for the minimum interatomic distances between atom pairs (cut-off distances) to avoid unreasonable atom contacts. For each sample, about fifty RMC configurations were obtained with more than 2,600,000 accepted configurations of atoms.

3. Results

3.1. XRD Analysis and DSC Studies

Bulk, transparent, slightly colored glasses (insets, Figure 1a) of 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%, were obtained in this study. The measured X-ray diffraction patterns are shown in Figure 1a, and confirm the amorphous nature of the prepared materials. Glasses without Bi2O3 (50ZnO:(50 − x)B2O3:xWO3:0.5Eu, x = 1, 5, and 10 mol%) were obtained earlier. The XRD patterns of the glass samples having 1 and 5 mol% WO3 were present in the 6th International Conference on Optics, Photonics, and Lasers (OPAL’ 2023). The photograph of the glass with the highest WO3 amount of 10 mol% was represented in ref. [15].
DSC curves of the glass samples 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%, obtained are presented in Figure 2. Glasses are characterized with two humps corresponding to the two glass transition temperatures, Tg1 and Tg2. The two glass transition effects observed are connected with the presence of two amorphous phases with different compositions in the investigated glasses. In the DSC curve of the glass having the highest WO3 concentration of 10 mol%, an exothermic peak due to the glass crystallization at temperature Tc = 683 °C was observed. For the glasses containing lower WO3 concentration of 5 and 1 mol%, glass crystallization effects did not appear, evidencing that these glasses possess higher thermal stability, which decreases with increasing WO3 content. The DSC analysis shows that the thermal parameters of glasses present here do not differ significantly from those obtained for the glasses with the similar compositions without Bi2O3 reported in Ref. [15]. However, the thermal stability of glasses containing 1 mol% Bi2O3 was slightly lowered, most probably because of the increased structural heterogeneity and, hence, higher crystallization ability of the composition.

3.2. IR Spectral Analysis

Structural information of the studied glasses was obtained by comparative analysis of IR spectra of glasses with the same compositions without Bi2O3 and containing 1 mol% Bi2O3, which are shown in Figure 3.
The IR spectra of Bi2O3-free glasses with the lower WO3 concentration of 5 and 1 mol% were previously reported by us in the 6th International Conference on Optics, Photonics, and Lasers (OPAL’ 2023), while the glass with the highest WO3 amount of 10 mol% was discussed earlier in ref. [15]. The IR spectra of the glasses without Bi2O3 contain bands characteristic to the metaborate groups BØ2O (shoulder at 1460 cm−1; band at 625–640 cm−1), pyroborate dimmers, B2O54− (weak band at 1110 cm−1; band at 625–640 cm−1), and superstructural groupings with BØ3 and BØ4 species (bands at 1250 cm−1 and at 1040–1050 cm−1, band at 680 cm−1) [14,23,24,25]. There are also WO6 (bands at 870 cm−1 and at 680 cm−1) and [WO4]2− tetrahedra (band at 480 cm−1) [23]. The addition of Bi2O3 to the glass compositions causes the BO3 →BO4 transformation, resulting in the increase in the number of superstructural units (increased intensity of the bands at 1240 and 1045–1035 cm−1 and band at 680 cm−1) [23]. For both the glass series, a partial [WO4]2− → WO6 transformation with increasing WO3 concentration occurs, manifested by the disappearance of the bands at 940 cm−1 and 880 cm−13[WO4]2−), and formation of an intense and well-formed band at 860–870 cm−1 (νWO6). The tungstate octahedral species sharing common corners (W–O–W) and edges (W2O2) are supposed, having in mind the structural and IR data of crystalline Bi2WO6, Bi2W2O9, and ZnWO4 phases [26,27,28,29,30,31]. These compounds consist of corner- or edge-shared WO6, whose IR spectra contain strong bands situated in the same spectral regions, from 870–800 cm−1 and 700–600 cm−1. Tungstate and borate species are charge-balanced by Bi3+, Zn2+, and Eu3+ ions via Bi–O–W, B–O–Bi, Zn–O–W, Zn–O–B, Eu–O–W, and Eu–O–B bonding. Additionally, the new high frequency band at 1350 cm−1 observed in glasses having higher WO3 of 5 and 10 mol% (Figure 2, ZBBW5:Eu and ZBBW10:Eu) is attributed to stretching of B–O bonds in BØ2O triangles, and its presence suggests stronger interaction between Bi3+ ions and nonbridging oxygens [32]. Bi3+ ions are incorporated in the structure of investigated glasses as BiO6 octahedra (bands at 640 and at 480 cm−1). Thus, the addition of Bi2O3 to glasses 50ZnO:(49 − x)B2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%, leads to the formation of more stable and reticulated glass structure, compared with the glasses with the same composition without Bi2O3. The Zn2+ ions generally participate in the borate glasses as ZnO4 tetrahedra with characteristic Zn2+ motion at 225 cm−1 [33]. The frequency of the Eu–O vibration, ν(Eu–O), has been measured at about 280 cm−1 for glasses (1 − 2x)Eu2O3 − x(SrO − B2O3) [34].
The detailed assignments of the bands observed in the IR spectra of the present glasses are summarized in Table 1.

3.3. Density, Molar Volume, Oxygen Packing Density, and Oxygen Molar Volume

A structural information of two series of glasses was also gained by density (ρg) measurement, on which basis the values of several physical parameters listed in Table 2 (molar volume (Vm), oxygen molar volume (Vo), and oxygen packing density (OPD)) are evaluated, using the conventional formulae [37]. Bi2O3 containing glasses are characterized with the higher density as compared with the respective Bi2O3-free glasses because of the replacement of lighter B2O3 (molecular weight 69.62 g/mol) with heavier Bi2O3 (molecular weight 465.96 g/mol). The Vm and Vo values of glasses having 1 mol% Bi2O3 are lower, while their OPD values are higher as compared with the values of the same parameters established for the glasses without Bi2O3, evidencing better packing and bonding in the Bi2O3-containing glass network and lower number of nonbridging oxygens (NBOs) [38].

3.4. RMC Modeling and Results

The RMC technique provided an excellent fit of the simulated structure factors (S(Q)-1) with the experimental one for 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol% (Figure 4).
From the RMC simulation, partial atomic pair-correlation functions, gij(r), and average coordination number distributions, CNij, were revealed, with good stability and statistics. The Zn–O distribution functions show symmetrical peaks centered in the range of 1.95 ± 0.01 Å (Table 3).
In function of gij(r), we specify a range in r over which atoms are counted as neighbors. This can be understood as defining coordination shells. Introducing a min point (positions of minimum values on the lower) and max point (the upper side of the corresponding peak), these are presented in Table 4, where we present the average coordination numbers (summarized in Table 4).
The average coordination number of the Zn–O was obtained from the RMC analysis, and it was found that Zn4+ was tetrahedrally coordinated with oxygens in the glassy network for all studied samples (Table 4). The B–O distribution function showed a relatively broad first neighbor distance at 1.40 ± 0.05 Å, and a slight shoulder at 1.80 ± 0.1 Å appeared in function of concentration in the 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%, glass samples. The B–O coordination was in the range of 3.48 ± 0.05 to 4.00 ± 0.05, and obtained the changes in BO3/BO4 ratio. The boron atoms were coordinated mostly by three and four oxygen atoms, forming trigonal BO3 and tetrahedral BO4 units, in agreement with coordination numbers in SiO2–Na2O–B2O3 glasses [39], in MoO3–ZnO–B2O3 glasses [40], and in ZnO–B2O3–Li2O–Al2O3 glasses [41]. The nearest W–O distances showed characteristic peaks at 1.75 ± 0.05 (Table 3) for both series. The average coordination number of W–O was in a wide range, from 6.20 ± 0.1 to 6.73 ± 0.1 (see Table 3) within the limits of experimental uncertainty. Based on the coordination numbers, we can predict that the W–O network consists of WO4 and WO6 units. The WO4/WO6 ratio changed with the WO3 concentration, and samples with highest WO3 concentration (10 mol%) had mostly WO6 units with fewer WO4 units. In the case of Bi–O and Eu–O, thanks to the very low Bi2O3 and Eu2O3 concentration, it was not relevant to obtain reliable numbers for the coordination.

3.5. EPR Spectroscopy

The EPR analyses of ZBBW5:Eu were carried out at 295 K and 120, and in Figure 5 the obtained spectra are shown. As seen, the spectra contain multiple signals with different intensities and g-factors. The most prominent features are assigned to impurities of isolated Fe3+ ions (the signal with g = 4.25) and isolated Mn2+ ions (six hyperfine structure lines marked with *, inset).
Eu2+ ions possess electron spin S = 7/2, and in their EPR spectrum seven signals were observed, corresponding to seven allowed transitions occurring between four Kramers’ doublets (ms = ±1/2, ms = ±3/2, ms= ±5/2, and ms = ±7/2) according to the selection rule [42]. Depending on the zero-field splitting parameters (D, E) and crystal field symmetry, the EPR spectra of Eu2+ usually include a part of these signals. In the spectra of ZBBW5:Eu are discernable a set of not-well-resolved and low-intensive signals with g factors about 6.0, 4.7, 3.4, and 2.8 located in the range 0–300 mT. These signals could be assigned to Eu2+ ions [42,43] localized in a low-symmetry crystal field with large zero-field splitting (D > hν).
In addition, the central region of the spectrum shows a signal with g = 2.00. The assignment of this signal is somewhat difficult, as it could derive both from Fe3+ ion impurities existing in the sample and Gd3+ ions in a highly symmetric environment. That is why its attribution remains unclear.
To summarize, the EPR spectra recorded for sample ZBBW5:Eu confirm the presence of Eu2+ ions, with their concentration being assessed as extremely low based on the comparison between the background spectrum and the analyzed spectrum.

3.6. Luminescent Properties

Figure 6 shows the photoluminescent excitation spectra of Eu3+-doped 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3, x = 1, 5, and 10 mol%, glasses monitoring the 5D07F2 red emission of Eu3+ at 612 nm [6].
The low-intensive broad band below 350 nm is ascribed to the characteristic absorption of Bi3+ (1S03P1) [17], the charge transfer bands (CTB), resulting from energy transitions from O2− to W6+ in WO4 and WO6 groups and O2− to Eu3+, as well as the ground 4f state of Eu3+ to W6+ [6,44,45,46,47,48]. The existence of the excitation band of host lattice absorption at Eu3+ emission (612 nm) implies the existence of nonradiative energy transfer from Bi3+ and WOn groups to the active rare-earth ion [48,49]. The sharp lines in 350–600 nm range correspond to the f → f intraconfigurational forbidden transitions of Eu3+ from the ground state (7F0) and from the first excited state (7F1): 7F05D4 (360 nm), 7F05G2 (375 nm), 7F15L7 (381 nm), 7F05L6 (392 nm), 7F05D3 (412 nm), 7F05D2 (463 nm), 7F05D1 (524 nm), 7F15D1 (531 nm), and 7F05D0 (576 nm) [6]. Among them, the electronic transition at 392 nm is the strongest one and was used as an excitation wavelength. Compared to the CTB, the intensity of the narrow f–f lines is stronger. This is favorable for Eu3+-doped luminescent materials, since, in general, the intensity of these Eu3+ transitions is weak due to the parity-forbidden law. Thus, the obtained glasses can be effectively excited by near-UV and blue light, which is compatible with the present LED chips.
The emission spectra of Eu3+-doped 50ZnO:(49 − x)B2O3: 1Bi2O3:xWO3, x = 1, 5, and 10 mol%, glasses under the excitation of λex = 392 nm consist of five emission peaks centered at 578, 592, 612, 651, and 700 nm, originating from 5D07FJ (J = 0, 1, 2, 3, 4) intraconfigurational transitions of Eu3+ (Figure 7) [6].
The characteristic single broad band emission of Bi3+, originating from 3P11S0 transition is located at 380–700 nm [50]. In the same spectral region, we also registered the broad emission band of the WOn group [51]. The general requirement for energy transfer from both WO3 and Bi2O3 to the rare-earth ion is satisfied, i.e., there exists a spectral overlap between the excitation peaks of Eu3+ (Figure 6) and the emission band of Bi3+ and WO3 (Figure 7). As a result, both oxides can act as sensitizers, transferring the emission energy nonradiatively to the activator Eu3+ by quenching their luminescence. Moreover, an indication of energy transfer is the absence of characteristics of WO3 and Bi2O3 emission bands [15,20,48]. As can be seen, with the increase of WO3 up to 10 mol% (Figure 7) and with the introduction of Bi2O3 in the 50ZnO:40B2O3:10WO3:0.5Eu2O3 glass composition (Figure 8), a significant enhancement of the emission intensities was achieved.
The most intensive emission peak, observed at 612 nm, corresponds to the hypersensitive to the site symmetry electric dipole (ED) transition 5D07F2, while the second-most intensive magnetic dipole (MD) 5D07F1 one is insensitive to the site symmetry and is considered almost constant [6,44,52,53]. The integrated emission intensity ratio (R) of these two transitions 5D07F2/5D07F1 is used to estimate the degree of symmetry around Eu3+ ions and the strength of covalence of the europium–oxygen bond. The higher R values indicate more site asymmetry of the rare-earth ion, a high covalency between Eu3+ and O2− ions, and an enhanced emission intensity [6,54,55]. The intensity ratios, R, of the present glasses (R = 4.7–5.7) (Table 5) are higher than most of the other reported Eu3+-doped glasses and have close values to Eu3+:Y2O3 and Eu3+:Y2O2S, indicating that the synthesized glasses are characterized by a more distorted environment of the Eu3+ ion and a high covalent bonding between Eu3+ and the surrounding ligands, thus achieving an enhanced Eu3+ emission intensity [15,18,20,36,56,57,58,59,60,61,62,63,64].
The R values were found to increase from 4.61 to 5.73 (Table 5) as the WO3 concentration raised from 1 to 10 mol%. The incorporation of small amounts of Bi2O3 (1 mol%) into the glass structure also led to an increase in the asymmetric ratio from 5.57 for the glass 50ZnO:40B2O3:10WO3:0.5Eu2O3 to 5.73 for the glass 50ZnO:39B2O3:1Bi2O3:10WO3:0.5Eu2O3. These R values were much higher as compared to the R value for glass with high Bi2O3 content (10 mol%) (50ZnO:(40 - x)B2O3:10Bi2O3:0.5Eu2O3:xWO3, x = 0 and 0.5) (R = 3.79) [20]. This result shows that we have found an appropriate glass composition for hosting an active rare-earth ion that provide a high Eu3+ emission intensity.

4. Discussion

In this study, by analyzing the IR spectra of zinc–borate glasses containing 1, 5, and 10 mol% WO3, with and without Bi2O3, it was found that 1 mol% Bi2O3 leads to an increase in the number of the BO4 involved in superstructural groupings and formation of Bi–O–B, Bi–O–W, Zn–O–B, Zn–O–W, Eu–O–W, and Eu–O–B cross-links in the glass structure. With WO3 loading in the Bi3+-containing glasses, a partial [WO4]2− → WO6 transformation took place. The tungstate octahedra shared common corners and/or edges, forming W–O–W and W2O2 bonds. There were also B–O–B bonds with different numbers with the WO3 content. The RMC modeling also revealed the presence of both WO6 and WO4 units and trigonal BO3 and tetrahedral BO4 units varying in amount with the composition. Thus, the bismuth-containing glasses were characterized by a more reticulated and rigid network that ensures low symmetry sites of Eu3+ ions, which is favorable for the luminescence emission of the active Eu3+ ion. The structural features revealed by IR analysis agreed well with the measured density and calculated physical parameters. Bi2O3-containing glasses were characterized with the lower molar and oxygen molar volume and higher oxygen packing density, as compared with Bi2O3-free glasses with the same compositions, as a result of the formation of highly cross-linked structure and the presence of new mixed bonding with participation of Bi3+ ions. A significant enhancement of the Eu3+ emission was established in the glasses 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, (a) x = 1 mol%, (b) x = 5 mol%, (c) x = 10 mol%, in the presence of low Bi2O3 content (1 mol%) and with the increase of WO3 content (up to 10 mol%). The photoluminescent spectra of the new glasses showed an intensive red luminescence at 612 nm as well as a very large value of the luminesce ratio R (over 5), both evidencing that Eu3+ ions occupied distorted sites in the created glass network. Particularly, the glass with 10 mol% WO3 showed the strongest emission of the active ions as a result of the structural features established and also because of an energy transfer from tungstate and bismuthate groups to the active ion.

5. Conclusions

The results of this investigation show that the zinc borate glass matrix with the simultaneous presence of both Bi2O3 and WO3 is very suitable for implementing the active Eu3+ as it possesses a reticulated and rigid glass structure, ensuring a more asymmetrical local structure around Eu3+ sites, accordingly yielding a higher luminescence of the incorporated Eu3+ ions. On the other hand, both bismuth and tungsten oxides have a synthesizer effect by transferring the emission energy nonradiatively to the activator Eu3+, which additionally improves its luminesce properties. This suggests that the obtained glasses are potential candidates for red-light-emitting phosphors.

Author Contributions

Conceptualization, A.Y., M.M. and R.I.; methodology, A.Y., M.M. and R.I.; software, A.Y., M.M. and M.F.; validation, A.Y., M.M. and M.F.; formal analysis, A.Y., M.M. and M.F.; investigation, A.Y., M.M., M.F., L.A. and P.P.; resources, R.I. and M.F.; data curation, A.Y., M.M., R.I. and M.F.; writing—original draft preparation, A.Y., M.M. and M.F.; writing—review and editing, R.I.; visualization, A.Y., M.M. and R.I.; supervision, R.I.; project administration, R.I. and M.F.; funding acquisition, R.I. and M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Science Programme “European research network” under project TwinTeam Д01-272/02.10.2020 г. and by joint research project within the framework of an international scientific cooperation between BAS and MTA IC-HU/01/2022-2023.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Special thanks are due to R. Kukeva for EPR measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cho, J.; Park, J.H.; Kim, J.K.; Schubert, E.F. White light-emitting diodes: History, progress, and future. Laser Photonics Rev. 2017, 11, 1600147. [Google Scholar] [CrossRef]
  2. Chen, Y.Y.; Duh, J.G.; Chiou, B.S.; Peng, C.G. Luminescent mechanisms of ZnS: Cu: Cl and ZnS: Cu: Al phosphors. Thin Solid Films 2001, 392, 50–55. [Google Scholar] [CrossRef]
  3. Kim, K.B.; Kim, Y.I.; Chun, H.G.; Cho, T.Y.; Jung, J.S.; Kang, J.G. Structural and optical properties of BaMgAl10O17: Eu2+ phosphor. Chem. Mater. 2002, 14, 5045–5052. [Google Scholar] [CrossRef]
  4. Trond, S.S.; Martin, J.S.; Stanavage, J.P.; Smith, A.L. Properties of Some Selected Europium-Activated Red Phosphors. J. Electrochem. Soc. 1969, 116, 1047–1050. [Google Scholar] [CrossRef]
  5. Feldmann, C.; Jüstel, T.; Ronda, C.R.; Schmidt, P.J. Inorganic luminescent materials: 100 years of research and application. Adv. Funct. Mater. 2003, 13, 511–516. [Google Scholar] [CrossRef]
  6. Binnemans, K. Interpretation of europium (III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef]
  7. Setlur, A.A. Sensitizing Eu3+ with Ce3+ and Tb3+ to make narrow-line red phosphors for light emitting diodes. Electrochem. Solid-State Lett. 2012, 15, J25. [Google Scholar] [CrossRef]
  8. Hong, J.H.; Zhang, Z.G.; Cong, C.J.; Zhang, K.L. Energy transfer from Bi3+ sensitizing the luminescence of Eu3+ in clusters embedded into sol–gel silica glasses. J. Non-Cryst. Solids 2007, 353, 2431–2435. [Google Scholar] [CrossRef]
  9. Spadaro, F.; Rossi, A.; Laine, E.; Hartley, J.; Spencer, N.D. Mechanical and tribological properties of boron oxide and zinc borate glasses. Phys. Chem. Glas. Eur. J. Glass Sci. Technol. B 2016, 57, 233–244. [Google Scholar] [CrossRef]
  10. Ehrt, D. Zinc and manganese borate glasses-phase separation, crystallization, photoluminescence and structure. Phys. Chem. Glas. Eur. J. Glass Sci. Technol. B 2013, 54, 65–75. [Google Scholar]
  11. Ivankov, A.; Seekamp, J.; Bauhofer, W. Optical properties of Eu3+-doped zinc borate glasses. J. Lumin. 2006, 121, 123–131. [Google Scholar] [CrossRef]
  12. Taki, Y.; Shinozaki, K.; Honma, T.; Dimitrov, V.; Komatsu, T. Electronic polarizability and interaction parameter of gadolinium tungsten borate glasses with high WO3 content. J. Solid. State Chem. 2014, 220, 191–197. [Google Scholar] [CrossRef]
  13. Ali, A.A.; Fathi, A.M.; Ibrahim, S. Material characteristics of WO3/Bi2O3 substitution on the thermal, structural, and electrical properties of lithium calcium borate glasses. Appl. Phys. A 2023, 129, 299. [Google Scholar] [CrossRef]
  14. Milanova, M.; Aleksandrov, L.; Iordanova, R. Structure and luminescence properties of tungsten modified zinc borate glasses doped with Eu3+ ions. Mater. Today Proc. 2022, 61, 1206–1211. [Google Scholar] [CrossRef]
  15. Milanova, M.; Aleksandrov, L.; Yordanova, A.; Iordanova, R.; Tagiara, N.S.; Herrmann, A.; Gao, G.; Wondraczek, L.; Kamitsos, E.I. Structural and luminescence behavior of Eu3+ ions in ZnO-B2O3-WO3 glasses. J. Non-Cryst. Solids 2023, 600, 122006. [Google Scholar] [CrossRef]
  16. Yu, B.; Zhou, X.; Xia, H.; Chen, B.; Song, H. Novel Bi3+/Eu3+ co-doped oxyfluoride transparent KY3F10 glass ceramics with wide tunable emission and high optical temperature sensitivity. J. Lumin. 2021, 239, 118366. [Google Scholar] [CrossRef]
  17. 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]
  18. Sontakke, A.D.; Tarafder, A.; Biswas, K.; Annapurna, K. Sensitized red luminescence from Bi3+ co-doped Eu3+: ZnO–B2O3 glasses. Phys. B Condens. Matter 2009, 404, 3525–3529. [Google Scholar] [CrossRef]
  19. Kaewkhao, J.; Boonin, K.; Yasaka, P.; Kim, H.J. Optical and luminescence characteristics of Eu3+ doped zinc bismuth borate (ZBB) glasses for red emitting device. Mater. Res. Bull. 2015, 71, 37–41. [Google Scholar] [CrossRef]
  20. Aleksandrov, L.; Yordanova, A.; Milanova, M.; Iordanova, R.; Fabian, M. Doping effect of WO3 on the structure and luminescent properties of Zno-B2O3-Bi2O3:Eu3+ glass. J. Chem. Technol. Metall. 2023, 58, 707–715. [Google Scholar]
  21. Svab, E.; Meszaros, G.; Deak, F. Neutron powder diffractometer at the Budapest research reactor. Mat. Sci. Forum 1996, 228, 247–252. [Google Scholar] [CrossRef]
  22. Gereben, O.; Jovari, P.; Temleitner, L.; Pusztai, L. A new version of the RMC++ Reverse Monte Carlo programme, aimed at investigating the structure of covalent glasses. J. Optoelec. Adv. Mater. 2007, 9, 3021–3027. [Google Scholar]
  23. Iordanova, R.; Milanova, M.; Aleksandrov, L.; Khanna, A. Structural study of glasses in the system B2O3-Bi2O3-La2O3-WO3. J. Non-Cryst. Solids 2018, 481, 254–259. [Google Scholar] [CrossRef]
  24. Kamitsos, E.I.; Patsis, A.P.; Karakassides, M.A.; Chryssikos, G.D. Infrared reflectance spectra of lithium borate glasses. J. Non-Cryst. Solids 1990, 126, 52–67. [Google Scholar] [CrossRef]
  25. Ouis, M.A.; El Batal, F.H.; Azooz, M.A. FTIR, optical and thermal studies of gadmium borate glass doped with Bi2O3 and effect of gamma irradiation. J. Aust. Ceram. Soc. 2020, 56, 283–290. [Google Scholar] [CrossRef]
  26. Islam, M.; Lazure, S.; Vannier, R.; Nowgorodski, G.; Mairesse, G. Structural and computational studies of Bi2WO6 based oxygen ion conductors. J. Mater. Chem. 1998, 8, 655–660. [Google Scholar] [CrossRef]
  27. Buttrey, D.; Vogt, T.; White, B. High-temperature incommensurate-to-commensurate phase transition in the Bi2MoO6 catalyst. J. Solid State Chem. 2000, 155, 206–208. [Google Scholar] [CrossRef]
  28. Sleight, A.W. Accurate cell dimensions for ABO4 molybdates and tungstates. Acta Cryst. 1972, 28, 2899–2991. [Google Scholar] [CrossRef]
  29. Mączka, M.; Macalik, L.; Hermanovicz, K.; Kepinski, L.; Tomaszewski, P. Phonon properties of nanosized bismuth layered ferroelectric material—Bi2WO6. J. Raman Spectrosc. 2010, 41, 1059–1066. [Google Scholar] [CrossRef]
  30. Ma<monospace>̧</monospace>czka, M.; Macalik, L.; Hanuza, J. Raman and IR spectra of the cation-deficient Aurivillius layered crystal Bi2W2O9. J. Raman Spectrosc. 2009, 40, 2099–2103. [Google Scholar]
  31. Zhao, X.; Yao, W.; Wu, Y.; Zhang, S.; Yan, H.; Zhi, Y. Fabrication and photoelectrochemical properties of porous ZnWO4 film. J. Solid State Chem. 2006, 179, 2562–2570. [Google Scholar] [CrossRef]
  32. Varsamis, C.P.; Makris, E.N.; Valvi, C.; Kamitsos, E.I. Short-range structure, the role of bismuth and property-structure correlations in bismuth borate glasses. Phys. Chem. Chem. Phys. 2021, 23, 10006–10020. [Google Scholar] [CrossRef] [PubMed]
  33. Yao, Z.Y.; Möncke, D.; Kamitsos, E.I.; Houizot, P.; Celarie, F.; Rouxel, T.; Wondraczek, L. Structure and mechanical properties of copper-lead and copper-zinc borate glasses. J. Non-Cryst. Solids 2016, 435, 55–68. [Google Scholar] [CrossRef]
  34. Winterstein-Beckmann, A.; Moncke, D.; Palles, D.; Kamitsos, E.I.; Wondraczek, L. Structure and Properties of Orthoborate Glasses in the Eu2O3–(Sr,Eu)O–B2O3 Quaternary. J. Phys. Chem. B 2015, 119, 3259–3272. [Google Scholar] [CrossRef] [PubMed]
  35. Iordanova, R.; Milanova, M.; Aleksandrov, L.; Shinozaki, K.; Komatsu, T. Structural study of WO3-La2O3-B2O3-Nb2O5 glasses. J. Non-Cryst. Solids 2020, 543, 120132. [Google Scholar] [CrossRef]
  36. Aleksandrov, L.; Milanova, M.; Yordanova, A.; Iordanova, R.; Nedyalkov, N.; Petrova, P.; Tagiara, N.S.; Palles, D.; Kamitsos, E.I. Synthesis, structure and luminescence properties of Eu3+-doped 50ZnO.40B2O3.5WO3.5Nb2O5 glass. Phys. Chem. Glas. Eur. J. Glass Sci. Technol. B 2023, 64, 101–109. [Google Scholar]
  37. El-Fayoumi, M.A.K.; Farouk, M. Structural properties of Li-borate glasses doped with Sm3+ and Eu3+ ions. J. Alloys Compd. 2009, 482, 356–360. [Google Scholar] [CrossRef]
  38. Villegas, M.A.; Fernández Navarro, J.M. Physical and structural properties of glasses in the TeO2–TiO2–Nb2O5 system. J. Eur. Ceram. Soc. 2007, 27, 2715–2723. [Google Scholar] [CrossRef]
  39. Fabian, M.; Araczki, C. Basic network structure of SiO2-Na2O-B2O3 glasses from diffraction and reverse Monte Carlo simulation. Phys. Scrip. 2016, 91, 054004. [Google Scholar] [CrossRef]
  40. Fabian, M.; Svab, E.; Krezhov, K. Network structure with mixed bond-angle linkages in MoO3-ZnO-B2O3 glasses: Neutron diffraction and reverse Monte Carlo modelling. J. Non-Cryst. Solids 2016, 433, 6–13. [Google Scholar] [CrossRef]
  41. Othman, H.; Valiev, D.; Polisadova, E. Structural and mechanical properties of zinc aluminoborate glasses with different content of aluminium oxide. J. Mater. Sci Mater. Electron. 2017, 28, 4647–4653. [Google Scholar] [CrossRef]
  42. Brodbeck, M.; Iton, L.E. The EPR spectra of Gd3+ and Eu3+ in glassy systems. J. Chem. Phys. 1985, 83, 4285–4299. [Google Scholar] [CrossRef]
  43. Nandyala, S.; Hungerford, G.; Babu, S.; Rao, J.L.; Leonor, I.B.; Pires, R.; Reis, R.L. Time resolved emission and electron paramagnetic resonance studies of Gd3+ doped calcium phosphate glasses. Adv. Mater. Lett. 2016, 7, 277–281. [Google Scholar] [CrossRef]
  44. Blasse, G.; Grabmaier, B.C. Luminescent Materials, 1st ed.; Springer: Berlin/Heidelber, Germany, 1994; p. 18. [Google Scholar]
  45. Hoefdraad, H.E. The charge-transfer absorption band of Eu3+ in oxides. J. Solid State Chem. 1975, 15, 175–177. [Google Scholar] [CrossRef]
  46. Parchur, A.K.; Ningthoujam, R.S. Behaviour of electric and magnetic dipole transitions of Eu3+, 5D07F0 and Eu–O charge transfer band in Li+ co-doped YPO4:Eu3+. RSC Adv. 2012, 2, 10859–10868. [Google Scholar] [CrossRef]
  47. Mariselvam, K.; Juncheng, L. Synthesis and luminescence properties of Eu3+ doped potassium titano telluroborate (KTTB) glasses for red laser applications. J. Lumin. 2021, 230, 117735. [Google Scholar] [CrossRef]
  48. Dutta, P.S.; Khanna, A. Eu3+ activated molybdate and tungstate based red phosphors with charge transfer band in blue region. ECS J. Solid State Sci. Technol. 2013, 2, R3153–R3167. [Google Scholar] [CrossRef]
  49. Thieme, C.; Herrmann, A.; Kracker, M.; Patzig, C.; Höche, T.; Rüssel, C. Microstructure investigation and fluorescence properties of Europium-doped scheelite crystals in glass-ceramics made under different synthesis conditions. J. Lumin. 2021, 238, 118244. [Google Scholar] [CrossRef]
  50. Minquan, W.; Xianping, F.; Guohong, X. Luminescence of Bi3+ ions and energy transfer from Bi3+ ions to Eu3+ ions in silica glasses prepared by the sol-gel process. J. Phys. Chem. Solids 1995, 56, 859–862. [Google Scholar] [CrossRef]
  51. Sungpanich, J.; Thongtem, T.; Thongtem, S. Large-scale synthesis of WO3 nanoplates by a microwave-hydrothermal method. Ceram. Int. 2012, 38, 1051–1055. [Google Scholar] [CrossRef]
  52. Shionoya, S.; Yen, W.M.; Yamamoto, H. Phosphor Handbook, 2nd ed.; CRC Press: London, UK; New York, NY, USA, 2018; p. 177. [Google Scholar]
  53. Gandhi, Y.; Kityk, I.V.; Brik, M.G.; Rao, P.R.; Veeraiah, N. Influence of tungsten on the emission features of Nd3+, Sm3+ and Eu3+ ions in ZnF2–WO3–TeO2 glasses. J. Alloys Compd. 2010, 508, 278–291. [Google Scholar] [CrossRef]
  54. Devi, C.H.B.; Mahamuda, S.; Swapna, K.; Venkateswarlu, M.; Rao, A.S.; Prakash, G.V. Compositional dependence of red luminescence from Eu3+ ions doped single and mixed alkali fluoro tungsten tellurite glasses. Opt. Mater. 2017, 73, 260–267. [Google Scholar] [CrossRef]
  55. Nogami, M.; Umehara, N.; Hayakawa, T. Effect of hydroxyl bonds on persistent spectral hole burning in Eu3+ doped BaO-P2O5 glasses. Phys. Rev. B 1998, 58, 6166–6171. [Google Scholar] [CrossRef]
  56. Swapna, K.; Mahamuda, S.; Rao, A.S.; Sasikala, T.; Packiyaraj, P.; Moorthy, L.R.; Vijayaprakash, G. Luminescence characterization of Eu3+ doped Zinc Alumino Bismuth Borate glasses for visible red emission applications. J. Lumin. 2014, 156, 80–86. [Google Scholar] [CrossRef]
  57. Raju, B.D.; Reddy, C.M. Structural and optical investigations of Eu3+ ions in lead containing alkali fluoroborate glasses. Opt. Mater. 2012, 34, 1251–1260. [Google Scholar] [CrossRef]
  58. Bettinelli, M.; Speghini, A.; Ferrari, M.; Montagna, M. Spectroscopic investigation of zinc borate glasses doped with trivalent europium ions. J. Non-Cryst. Solids 1996, 201, 211–221. [Google Scholar] [CrossRef]
  59. Lakshminarayana, G.; Buddhudu, S. Spectral analysis of Eu3+ and Tb3+:B2O3-ZnO-PbO glasses. Mater. Chem. Phys. 2007, 102, 181–186. [Google Scholar] [CrossRef]
  60. Babu, A.M.; Jamalaiah, B.C.; Suhasini, T.; Rao, T.S.; Moorthy, L.R. Optical properties of Eu3+ ions in lead tungstate tellurite glasses. Solid State Sci. 2011, 13, 574–578. [Google Scholar] [CrossRef]
  61. Bindu, S.H.; Raju, D.S.; Krishna, V.V.; Rao, T.R.; Veerabrahmam, K.; Raju, C.L. UV light induced red emission in Eu3+-doped zincborophosphate glasses. Opt. Mater. 2016, 62, 655–665. [Google Scholar] [CrossRef]
  62. Gökçe, M. Development of Eu3+ doped bismuth germanate glasses for red laser applications. J. Non-Cryst. Solids 2019, 505, 272–278. [Google Scholar] [CrossRef]
  63. Kabir, M.; Ghahari, M.; Afarani, M.S. Co-precipitation synthesis of nano Y2O3: Eu3+ with different morphologies and its photoluminescence properties. Ceram. Int. 2014, 40, 10877–10885. [Google Scholar] [CrossRef]
  64. Som, S.; Mitra, P.; Kumar, V.; Kumar, V.; Terblans, J.J.; Swart, H.C.; Sharma, S.K. The energy transfer phenomena and colour tunability in Y2O2S: Eu3+/Dy3+ micro-fibers for white emission in solid state lighting applications. Dalton Trans. 2014, 43, 9860–9871. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of (a) 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%; (b) 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%.
Figure 1. XRD patterns of (a) 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%; (b) 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%.
Materials 16 06779 g001
Figure 2. DSC curves of glasses 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%.
Figure 2. DSC curves of glasses 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%.
Materials 16 06779 g002
Figure 3. IR spectra of 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol% (in red); and 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol% (in black).
Figure 3. IR spectra of 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol% (in red); and 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol% (in black).
Materials 16 06779 g003
Figure 4. Experimental (color) and RMC (black line)-simulated neutron scattering structure factors for 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol% glasses.
Figure 4. Experimental (color) and RMC (black line)-simulated neutron scattering structure factors for 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol% glasses.
Materials 16 06779 g004
Figure 5. EPR spectra of ZBBW5:Eu, recorded at 120 and 295 K. The quartz tube background is represented at the bottom. EPR measurement conditions: Att = 16 (5 mW); MA = 20.
Figure 5. EPR spectra of ZBBW5:Eu, recorded at 120 and 295 K. The quartz tube background is represented at the bottom. EPR measurement conditions: Att = 16 (5 mW); MA = 20.
Materials 16 06779 g005
Figure 6. Excitation spectra of 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3 (x = 1, 5, and 10 mol%) glasses.
Figure 6. Excitation spectra of 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3 (x = 1, 5, and 10 mol%) glasses.
Materials 16 06779 g006
Figure 7. Emission spectra of 50ZnO:(49 − x)B2O3: 1Bi2O3:xWO3:0.5Eu2O3 (x = 1, 5, and 10 mol%) glasses.
Figure 7. Emission spectra of 50ZnO:(49 − x)B2O3: 1Bi2O3:xWO3:0.5Eu2O3 (x = 1, 5, and 10 mol%) glasses.
Materials 16 06779 g007
Figure 8. Emission spectra of 50ZnO:(40 − x)B2O3: xBi2O3:10WO3:0.5Eu2O3 (x = 0 and 1 mol%).
Figure 8. Emission spectra of 50ZnO:(40 − x)B2O3: xBi2O3:10WO3:0.5Eu2O3 (x = 0 and 1 mol%).
Materials 16 06779 g008
Table 1. Infrared bands (in cm−1) and their assignments for glasses 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%.
Table 1. Infrared bands (in cm−1) and their assignments for glasses 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%.
Infrared Bands Position (cm−1)AssignmentRef.
475ν4[WO4]2− + Bi–O vibrations in the BiO6 groups[32,35]
680Bending vibrations of B–O–B bonds in superstructural [32]
640–625Bending vibrations of B–O–B bonds in meta- and pyroborates + Bi–O vibrations in the BiO6 groups[32]
860–870νWO6[14,23]
940; 880ν3[WO4]2− in distorted tetrahedra[23]
1050–1035νas4 involved in superstructural units[14,23]
1100νas(B–O–B); B–O–B bridge in pyroborate units, B2O54−[24,35]
1245νas(B–O–B); B–O–B bridges connect BO3 units + BO3 stretch in meta-, pyro-, orthoborate units[15,36]
1350ν(B–O) stretch in BØ2O units charge balanced by Bi3+[32]
1460ν(B–O) stretch in BØ2O units[23]
Table 2. Values of physical parameters of glasses 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%: density (ρg), molar volume (Vm), oxygen molar volume (Vo), oxygen packing density (OPD).
Table 2. Values of physical parameters of glasses 50ZnO:(49 − x)B2O3:1Bi2O3:xWO3:0.5Eu2O3, x = 1, 5, and 10 mol%: density (ρg), molar volume (Vm), oxygen molar volume (Vo), oxygen packing density (OPD).
Sample
ID
ρg
(g/cm3)
Vm
(cm3/mol)
Vo
(cm3/mol)
OPD
(g atom/L)
ZBW1:Eu3.475± 0.00222.7011.2988.55
ZBW5:Eu3.689± 0.00223.1411.5186.86
ZBW10:Eu *3.910 ± 0.002 *23.91 *11.73 *84.27 *
ZBBW1:Eu3.679 ± 0.00122.5711.2089.28
ZBBW5:Eu3.889 ± 0.00523.0111.4287.57
ZBW10:Eu4.175 ± 0.00123.3811.6086.18
* The physical parameters of glass ZBW10:Eu were previously reported in Ref. [15].
Table 3. Interatomic X–O distances, gij(r), obtained from RMC simulation. The errors are estimated from the reproducibility of various RMC runs.
Table 3. Interatomic X–O distances, gij(r), obtained from RMC simulation. The errors are estimated from the reproducibility of various RMC runs.
Title 1Zn–O
gij(r) (Å)
B–O
gij(r) (Å)
Bi–O
gij(r) (Å)
W–O
gij(r) (Å)
Eu–O
gij(r) (Å)
O–O
gij(r) (Å)
ZBBW1:Eu1.95 ± 0.011.40/1.80 ± 0.052.00 ± 0.051.75 ± 0.052.20 ± 0.052.35 ± 0.03
ZBBW5:Eu1.95 ± 0.011.40/1.80 ± 0.052.00 ± 0.051.75 ± 0.052.20 ± 0.052.35 ± 0.03
ZBB10:Eu1.95 ± 0.011.40/1.80 ± 0.052.00 ± 0.051.75 ± 0.052.20 ± 0.052.35 ± 0.03
Table 4. Average coordination numbers, CNij, calculated from RMC simulation. In brackets, the interval is indicated, where the actual coordination number was calculated.
Table 4. Average coordination numbers, CNij, calculated from RMC simulation. In brackets, the interval is indicated, where the actual coordination number was calculated.
SampleZn–O
CNij
B–O
CNij
W–O
CNij
O–O
CNij
ZBBW1:Eu4.01 ± 0.05
(min: 1.80–max: 2.20)
3.90 ± 0.05
(min: 1.20–max: 1.65)
6.20 ± 0.1
(min: 1.65–max: 2.23)
5.63 ± 0.1
(min: 2.20–max: 2.60)
ZBBW5:Eu3.99 ± 0.05
(min: 1.80–max: 2.20)
3.52 ± 0.05
(min: 1.20–max: 1.65)
6.42 ± 0.1
(min: 1.60–max: 2.25)
5.32 ± 0.1
(min: 2.20–max: 2.60)
ZBBW10:Eu3.97 ± 0.05
(min: 1.80–max: 2.20)
3.48 ± 0.05
(min: 1.20–max: 1.65)
6.73 ± 0.1
(min: 1.60–max: 2.25)
5.54 ± 0.1
(min: 2.20–max: 2.60)
Table 5. Comparison of the luminescence intensity ratio (R) of 5D07F2 to 5D07F1 transition of Eu3+-doped oxide glasses.
Table 5. Comparison of the luminescence intensity ratio (R) of 5D07F2 to 5D07F1 transition of Eu3+-doped oxide glasses.
Glass CompositionR ValuesRef.
50ZnO:48B2O3:1Bi2O3:1WO3:0.5Eu2O34.61Present work
50ZnO:44B2O3:1Bi2O3:5WO3:0.5Eu2O35.04Present work
50ZnO: 40B2O3:10WO3:0.5Eu2O35.57[12]
50ZnO:39B2O3:1Bi2O3:10WO3:0.5Eu2O35.73Present work
50ZnO:40B2O3:10WO3:xEu2O3 (0 ≤ x ≤ 10)4.54–5.77[15]
50ZnO:40B2O3:5WO3:5Nb2O5:xEu2O3 (x = 0, 0·1, 0·5, 1, 2, 5 and 10)5.09–5.76[36]
50ZnO:(40 − x)B2O3:10Bi2O3:0.5Eu2O3:xWO3,
x = 0 and 0.5
3.58; 3.79[20]
20ZnO:8Al2O3:(12 − x)Bi2O3:60B2O3:xEu2O31.951–2.78[56]
39.5Li2O:59.5SiO2:1Eu2O33.20[57]
4ZnO:3B2O3 0.5 ÷ 2.5 mol% Eu3+3.94–2.74[58]
Eu3+: 45B2O3-5ZnO-49PbO3.03[59]
15PbF2:25WO3:(60 − x)TeO2:xEu2O3 x = 0.1,
0.5, 1.0 and 2.0 mol%
2.37–2.78[60]
40ZnO:(30 − x) B2O3:30P2O5:xEu2O3
(0.1 ≤ x ≤ 0.9)
2.96–3.65[61]
60ZnO:(40x)B2O3:0.2Eu2O3:xBi2O3 (x = 0,
0.1, 0.2, 0.5, 1.0)
2.98[18]
(100 − x):(0.2Bi2O3–0.8GeO2):xEu2O3
(x = 0.5, 1, 1.5, 2 mol%)
3.94–4.21[62]
Eu3+:Y2O3 3.8–5.2[63]
Eu3+ doped Y2O2S6.45–6.62[64]
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

Yordanova, A.; Milanova, M.; Iordanova, R.; Fabian, M.; Aleksandrov, L.; Petrova, P. Network Structure and Luminescent Properties of ZnO–B2O3–Bi2O3–WO3:Eu3+ Glasses. Materials 2023, 16, 6779. https://doi.org/10.3390/ma16206779

AMA Style

Yordanova A, Milanova M, Iordanova R, Fabian M, Aleksandrov L, Petrova P. Network Structure and Luminescent Properties of ZnO–B2O3–Bi2O3–WO3:Eu3+ Glasses. Materials. 2023; 16(20):6779. https://doi.org/10.3390/ma16206779

Chicago/Turabian Style

Yordanova, Aneliya, Margarita Milanova, Reni Iordanova, Margit Fabian, Lyubomir Aleksandrov, and Petia Petrova. 2023. "Network Structure and Luminescent Properties of ZnO–B2O3–Bi2O3–WO3:Eu3+ Glasses" Materials 16, no. 20: 6779. https://doi.org/10.3390/ma16206779

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