Next Article in Journal
From Ethnobotany to Biotechnology: Wound Healing and Anti-Inflammatory Properties of Sedum telephium L. In Vitro Cultures
Previous Article in Journal
Development of a Convenient and Quantitative Method for Evaluating Photosensitizing Activity Using Thiazolyl Blue Formazan Dye
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of the Addition of WO3 on the Structure and Luminescent Properties of ZnO-B2O3:Eu3+ Glass

1
Institute of General and Inorganic Chemistry, Bulgarian Academy of Sciences, G. Bonchev, Str., bld. 11, 1113 Sofia, Bulgaria
2
Institute of Optical Materials and Technologies “Acad. Jordan Malinowski”, Bulgarian Academy of Sciences, Blvd. Akad. G. Bonchev 109, 1113 Sofia, Bulgaria
3
Institute of Electronics, Bulgarian Academy of Sciences, Tzarigradsko Shousse 72, 1784 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(11), 2470; https://doi.org/10.3390/molecules29112470
Submission received: 29 April 2024 / Revised: 20 May 2024 / Accepted: 22 May 2024 / Published: 24 May 2024

Abstract

:
Glasses with the compositions in mol % of 50ZnO:(50 − x)B2O3:0.5Eu2O3:xWO3, x = 0, 1, 3, 5 and 10 were obtained by applying the melt-quenching method and investigated by Raman spectroscopy, DSC analysis and photoluminescence (PL) spectroscopy. Raman spectra revealed that tungstate ions incorporate into the base zinc borate glass as tetrahedral [WO4]2− groups, and octahedral [WØ4O2]2− species with four bridging and two non-bridging oxygen atoms. There are also metaborate, [BØ2O] and pyroborate units, [B2O5]4−, in the glass networks. The glasses are characterized by good transmission in the visible region, at about 80%. Photoluminescence (PL) spectra evidenced that WO3 is an appropriate constituent for the modification of zinc borate glass structure and for enhancing the Eu3+ luminescent intensity. The most intense luminescence peak observed, at 612 nm, suggests that the glasses are potential materials for red emission.

Graphical Abstract

1. Introduction

Distinct visible emission of glasses doped with rare earth (RE) ions has attracted researchers to the development of glasses for various optical devices such as lasers, upconverters, stimulated phosphors, white light-emitting diodes (WLEDs) and optical amplifiers [1]. The luminescent and absorption properties of RE ions in glasses greatly depend on the chemical composition, structure and nature of the bonds of the host glass [2].
B2O3 is a good glass former and can form glass alone with good transparency, high chemical durability, thermal stability and good rare earth ion solubility [3]. Borate-based glasses are good hosts for RE ions, having flexibility for both the size and composition of the materials [4]. In the glass network, ZnO acts as a glass former as well as modifier, which depends on its mole concentration [5]. Its presence in the glass composition shortens the time taken for the solidification of glasses during the quenching process. Glasses containing ZnO have high chemical stability and less thermal expansion. Their wide band gap and intrinsic emitting property make them promising candidates for the development of optoelectronic devices, solar energy concentrators, ultraviolet emitting lasers, and gas sensors [6]. ZnO is thermally stable and appreciably covalent in character [7]. WO3 is a semi-glass former with various structural units, like tetrahedral and octahedral units (WO4 and WO3) of W6+ and W5+ ions in the glass network [8]. Particularly, the addition of WO3 to glasses enables special features such as the enhancement of the devitrification resistance and chemical durability of the glasses and increases in the solubility of rare earth ions in the host glass network [9]. Tungsten ions are well known for their unusual influence on the optical and electrochemical properties of glasses [10,11]. Due its high polarizability, WO3 could contribute to a great extent to the obtaining of high-refractive-index glasses, the enhancement of non-linear optical properties and improvements in the luminescent properties of RE ions [10].
In our recent articles, we have reported about the synthesis of tungsten-modified zinc borate glasses and glass crystalline materials of compositions 50ZnO:(50 − x)B2O3:xWO3, x = 0, 10, 15, 20 mol % doped with different amounts of Eu2O3 (0.1; 0.5; 1; 2; 5 and 10 mol %) [12,13]. It was found that the addition of 10 mol % WO3, at the expense of B2O3, increases glass density, providing clear and homogeneous bulk glass samples and exhibiting a better emission performance compared to the base binary zinc borate glass. Partially crystallized specimens comprising ZnWO4 as the crystalline phase were obtained from the compositions having 15 and 20 mol % WO3.
As a continuation of our previous studies, mentioned above, in this work, we investigate the effect of the addition of smaller amounts of WO3 (up to 5 mol %) on the structure and luminescent properties of ZnO-B2O3 glass doped with 0.5 mol % Eu2O3 by using Raman spectroscopy, differential scanning calorimetry and photoluminescent spectroscopy. The aim is to find the optimal glass composition, ensuring the most appropriate structure for accommodating the active rare earth ion, that will improve its luminescence properties.

2. Results

2.1. Thermal Analysis

Bulk transparent glasses were obtained from the nominal compositions in mol % of 50ZnO:(50 − x)B2O3:0.5Eu2O3:xWO3, x = 0, 1, 3, 5. Glass samples having 10 mol % WO3 (50ZnO:40:B2O3:0.5Eu2O3:10WO3), previously reported by us in ref. [13], were prepared and investigated again in the present work to compare their structure and luminescence behavior with those of glasses having lower WO3 content. Having in mind the high vapor pressure of WO3 oxide, which is the doping component, we accept that its actual concentration does not change significantly from the nominal one after the melting of the compositions. The X-ray diffractograms of the investigated glasses, along with the photographic images of the samples, are shown in references [13,14]. All investigated samples were X-ray amorphous. The prepared WO3--free glass was colorless, while WO3--containing glasses were brownish, due to the presence of some amount of Eu2+ ions, established by EPR measurements in ref. [13].
50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3, (x = 0, 1, 3, 5 and 10 mol %) glasses have been also investigated by DSC analysis in order to obtain information for some thermal parameters and for structural changes that take place due to the compositional changes [15]. The glass transition temperature, Tg, has been determined, since it is connected with both the strength of inter-atomic bonds and glass network connectivity. A higher Tg corresponds to a more rigid structure, whereas the glasses having a loose-packed structure have a lower Tg [16,17,18]. Figure 1 compares the DSC curves of the glasses investigated in this work. Two humps, corresponding to the two glass transition temperatures Tg1 and Tg2, are observed. Their values are listed in Table 1. The presence of two glass transition effects is connected with the formation of two amorphous phases with different compositions in the investigated glasses.
As one can see, the addition of WO3 into zinc borate glass causes a slight decrease in the glass transition temperature values. Having in mind our previous IR data and as well as the established values of structurally sensitive physical parameters [14], we explain the observed reduction in Tg1 and Tg2 as a result of increasing non-bridging atoms with the addition of WO3.
The values of Tg1 and Tg2 were correlated with average single-bond enthalpy, EB, of glasses using the following relationship proposed in ref. [19]:
E B = E 50 Z n - O + E ( 50 x ) B - O + E x W - O + E 0.5 E u - O   100
where EW-O; EB-O, EEu-O and EZn-O are the bond dissociation energies for the single bonds: W-O; B-O, Eu-O and Zn-O, respectively [20]. EB values gradually decrease with WO3 loading because of the gradual replacement of stronger B-O bonds with higher bond dissociation energy (808 kJ mol−1) by weaker W-O bonds with smaller bond dissociation energy (653 kJ mol−1) [20].

2.2. Structural Analysis

In our previous studies, we have investigated the structure of these glasses by applying IR spectroscopy [13,14]. The analysis revealed the presence of several borate structural units in the glass network as [BØ2O] and [BØ4] metaborate groups, and pyroborate [B2O5]4− entities. WO3 modifies the borate network, resulting in a decrease in [BØ4] units and the favoring of the formation of pyroborate dimers, [B2O5]4−. Tungstate ions incorporate into base zinc borate glass as tetrahedral [WO4]2− groups, and octahedral [WØ4O2]2− species with four bridging and two non-bridging oxygen atoms. The densities of these glasses have been measured, and on this basis, several structurally sensitive parameters such as molar volume (Vm), oxygen volume (Vo) and oxygen packing density (OPD) have been also established and are reported in refs. [13,14]. Their values also showed the depolymerization of the glass structure of the base zinc borate glass (increasing NBOs) when WO3 is added.
In the present study, we have used, additionally, Raman spectroscopy in order to obtain more complete structural information about the investigated glasses. The Raman spectra obtained are shown in Figure 2. The Raman spectrum of the base zinc borate glass, without WO3 (Figure 2; x = 0), is in good agreement with what has been reported by other authors for similar compositions [21,22,23]. It contains a strong and broad band centered at about 870 cm1, bands at 800 cm1 and 775 cm1, a broad shoulder at about 705 cm1, features at about 260 cm1 and 315 cm1 and a high-frequency envelope from about 1200 to 1550 cm−1.
The most prominent band, at 870 cm−1, observed only in the Raman spectrum of glass x = 0, is due to the symmetric stretching of B-O-B bridges in pyroborate dimers, [B2O5]4− [21,22,23]. The band at 800 cm1 is connected with the ring breathing of the boroxol rings, while the not-well-resolved band at 775 cm−1 signals the presence of six-membered borate rings consisting of two [BO0] triangles and one tetrahedral unit [BO4] (i.e., triborate rings) [21,22,23]. The broad shoulder at about 705 cm−1 contains contributions of at least four borate arrangements: deformation modes of metaborate chains, [BØ2O] (Ø = bridging oxygen, O = nonbridging oxygen), in-plane and out-of-plane bending modes of both polymerized (BØ0) species and isolated orthoborate units (BO3)3−, and bending of the B-O-B connection in the pyroborate dimers, [B2O5]4−) [21,22,23]. The higher-frequency activity in the 1200 to 1550 cm−1 range is connected with the stretching vibrations of B-O bonds, involving non-bridging oxygen (NBO) in the pyroborate dimers (1240 cm1), and in metaborate triangular units, BØ2O (1380 cm1) [21,22,23]. The lower-frequency features at 260 and 315 cm−1 are related to the Zn-O vibrations and Eu-O vibrations, respectively [13]. The addition of WO3 influences the Raman spectrum of the glass x = 0 considerably. A new and strong band at 970 cm1 appears, due to the ν1 symmetric stretching mode of isolated [WO4]2− tetrahedra, charge balanced by Zn2+ and Eu3+ ions [13]. Another band, characteristic of the [WO4]2− tetrahedral units, is the band at 800 cm1, which is due to the asymmetric stretching ν3 mode of tetrahedral [WO4]2− groups. This band significantly overlaps with the ring breathing mode of the boroxol rings. The band at 775 cm1, due to the six-membered borate rings with one [BO4]- unit present in the spectrum of WO3-free zinc borate glass (Figure 2, x = 0), disappears in the spectra of glasses containing WO3, indicating the destruction of these structural groups when WO3 is added. Thus, depolymerization of the borate oxygen network took place, which is in agreement with the previously reported IR data [13,14]. The low-frequency band at 315 cm1, attributed to the Eu-O vibration in the glass x = 0, overlaps with the ν2 [WO4]2− mode, which explains its increasing intensity upon WO3 loading. Another new low-frequency band at 355 cm1 is attributed to the ν4 bending mode of the [WO4]2− tetrahedra [13]. In addition to tungstate tetrahedral groups, there are also tungstate octahedral [WØ4O2]2− species with four bridging and two non-bridging oxygen atoms in the structure of the studied glasses, which by edge-sharing form [W2O84−] polymeric anions [13]. These tungstate species are charge balanced by Zn2+ and Eu3+ cations [13]. The presence of several new bands in the spectra of WO3-containing glasses at 840 cm1, 865 cm1, 663 cm1, 690 cm1 and 400 cm1 can be discussed in terms of the vibrations of the WO2 terminal units and [W2O4]n chains of the [W2O84-]n anions, as was reported in the literature for the ZnWO4 compound [13,24]. In particular, Raman bands at 840 and at 865 cm1 are connected with νas(WO2) and νs(WO2) vibrations of terminal WO2 units. The bands at 663, 690 cm1 and 400 cm1 are related to νas[W2O4]n and νs[W2O4]n vibrations, which involve mainly two-oxygen bridges (W2O2) of the chain structure [W2O4]n [13]. On the other hand, the band at 840 cm1 in the spectra of WO3-containing glasses can be also connected with the symmetric B-O-B stretching of pyroborate dimmers, [B2O5]4− [21,23]. Because of its complex character, the band growth with increasing WO3 loads observed is not possible to explain in a straightforward manner. The Raman activity in the high-frequency region, 1200–1550 cm1, is connected with the vibration of B-O-containing borate arrangements. More particularly, the higher-frequency activity at 1240 cm−1 reflects the symmetric stretch of boron-non-bridging oxygen bonds, ν(B-O) of the pyroborate dimers, while the other two features at 1380 and 1400 cm−1 are due to the B-O stretching in metaborate triangular units, [BØ2O] [21,22,23]. There is a red shift in the frequency of the band at 1380 cm1 to 1320 cm1 in the spectra of glasses x = 5 and x = 10. As is reported in ref. [21], the frequency of the boron oxygen stretching in trigonal metaborate units is highly sensitive to the surrounding environment, i.e., the degree of the covalency of the non-bridging oxygen with the charge-balancing cations. The red shift observed indicates that in the glasses having higher WO3 content (5 and 10 mol %), [BØ2O] entities are charge balanced mainly by Zn2+ and Eu3+ ions. The proposed band assignments are summarized in Table 2.

2.3. Optical Transmission Spectra

Figure 3a,b presents the optical transmission spectra and absorption coefficient data of the investigated glasses. As is seen from Figure 3a, glasses are characterized as having good transmission in the visible region of the electromagnetic spectrum at around 80%; this slightly decreases with the WO3 loading. The changes in the optical transmission observed are due to the low coloration of the glasses containing WO3. In addition, the absence of any absorption bands in the visible range indicates that there are no tungstate ions in the lower-than-W+6 valance state, since the reduced W+5 and W+4 ions produce very intensive absorption bands in the visible range due to d-d transition [25].
Two weak absorption bands at 392 and 463 nm are observed, corresponding to the f-f transitions of Eu3+ ions between the ground and the excited states. We have calculated the absorption coefficient (α) using the following equation:
α = ln 100 T / d
where T is the percentage transmission and d is the thickness of the glass. The absorption coefficients versus wavelength spectra are presented in Figure 3b. The maximum absorption values of the glasses increase with the increase in WO3 content and vary between 280 and 311 nm.

2.4. Luminescent Properties

The optical properties of the obtained glasses were further studied. The excitation spectra of 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10) glasses, shown in Figure 4, are recorded at room temperature, monitoring the most prominent Eu3+ emission at 612 nm (5D07F2 transition). The spectra consist of two distinct parts. The first part is composed of a broad band in the spectral range from 250 nm to 350 nm due to the ligand-to-metal charge transfer transitions (LMCT) of O2− → W6+ [26] and O2− → Zn2+ [27] inside the WOn (WOn = WO4 and WO6) and ZnOn (ZnOn = ZnO4) host absorbing groups, respectively, and from the oxygen 2p orbital to the empty 4f orbital of europium (O2−→ Eu3+) [28,29,30,31]. Additionally, several sharp intra-configurational 4f → 4f transitions of Eu3+ can be observed in the spectra at 317 nm, 375 nm, 380 nm, 392 nm, 413 nm, 463 nm, 524 nm, 531 nm and 576 nm due to 7F05H3, 7F05D4, 7F05G2, 7F15L7, 7F05L6, 7F05D3, 7F05D2, 7F05D1, 7F15D1, 7F05D0, respectively [32]. The 7F05H3 excitation peak at 317 nm is superposed over the broad charge transfer band. The strongest excitation lines are observed at 392 nm and 463 nm, assigned to the 7F05L6 and 7F05D2 transitions in the near-UV and visible blue regions. These wavelengths are suitable for excitation with commercial near-ultraviolet light-emitting diodes (LEDs) (250–400 nm) and blue LED chips (430–470 nm).
As can be seen from Figure 4, with increasing concentration of WO3 in the glass composition, the intensity of Eu3+ f-f transitions increases for up to 5 mol % WO3. A further increase in WO3 leads to a decrease in the excitation intensity. Accordingly, the addition of WO3 into Eu3+-doped 50ZnO:50B2O3 host glass up to 5 mol % is beneficial to achieve good excitation, because as a rule, Eu3+ excitation peaks are characterized by low intensity due to the parity-forbidden law. The appearance of absorption of the host matrix, when monitoring the Eu3+ emission at 612 nm, has been shown to play an important role in the enhancement of the rare earth emission intensity through the occurrence of non-radiative energy transfer, in particular from WOn and ZnOn structural polyhedra to the Eu3+ ion [13,14,26,33,34,35,36,37,38]. This process is known as “host sensitized” energy transfer. Additionally, according to data from the literature, the 350 nm–700 nm region is registered to be the broad emission band (red line, Figure 5) of WO3 [39] and ZnO [27]. In the same spectral region are also located the excitation bands of Eu3+ (black line) [32]. The other requirement for energy transfer is the overlap of the host group emission and the excitation levels of the active ion. As can be seen from Figure 5, in our case, this condition is satisfied.
In Figure 6, it can be seen the emission spectra of pure and Eu3+-doped 50ZnO:40B2O3:10WO3 glass matrix were acquired upon excitation at the maximum value of the host charge transfer band at 248 nm (WOn and ZnOn absorbing groups). The observed decrease in the emission of the host absorbing groups in Eu3+-doped glass (red line), as compared to pure host emission (black line) [27,39], evidences that energy transfer has occurred. In other words, the energy absorbed by tungstate and zincate groups is further transferred non-radiatively to the active Eu3+ ion. The results obtained above also imply that Eu3+ and the WOn and ZnOn groups are closely coordinated in the structure [33]. It is known that the probability of energy transfer increases when the host absorbing groups, in our case WOn or ZnOn and the Eu3+ active ion, are nearest neighbors in the structure [40]. The emission spectra of Eu3+-doped glasses are also obtained by monitoring the 7F0  5L6 transition at 392 nm (Figure 7).
The characteristic intra-configurational transition of Eu3+ 5D07FJ, where J = 0, 1, 2, 3, 4 at 578 nm, 591 nm, 642 nm, 652 nm and 700 nm, appeared in the spectra [32]. The addition of WO3 into the glass composition leads to an increase in the emission intensity for up to 5 mol %. At 10 mol % WO3, a luminous quenching is observed in the spectra. Among the Eu3+ sharp emission bands, the most intense one, located at 612 nm, was allowed by the electric dipole and was sensitive to the changes in the surrounding 5D07F2 transition (red color of emission), followed by the magnetic dipole, which allowed the 5D07F1 transition (orange emitting color), which is insensitive to the coordination environment around Eu3+ ions. The ratio between these emissions, known as the asymmetric ratio R, can reveal the degree of asymmetry in the local environment around the Eu3+ and the strength of the Eu–O covalence for various Eu3+-doped compounds. The lower the value of the asymmetry parameter, the higher the symmetry around the active ion, and the lower the Eu–O covalency and emission intensity. The increase in R value is due to the increase in asymmetry and covalency between the Eu3+ ion and the ligands [41]. The R values of the synthesized glasses are listed in Table 3 along with other data reported in the literature about Eu3+-doped oxide glasses [6,13,34,42,43,44,45,46,47,48]. The relatively higher values compared to other reported Eu3+-doped oxide glasses indicate that Eu3+ occupies crystallographic sites with low symmetry and also provide evidence of the high Eu3+-O2− covalency. The highest obtained R value of 5.82, attributed to 50ZnO:45B2O3:5WO3:0.5Eu2O3, is an indication of the highest emission intensity. Additionally, the Eu3+ 5D07F1 transition splits into three emission peaks, centered at 586 nm, 591 nm and 596 nm. This effect probably arises from crystal field splitting, which causes a single transition to produce multiple emission peaks [49]. Furthermore, the appearance of the 5D07F0 transition, which is sensitive to the crystal field and is forbidden based on the standard Judd–Ofelt theory, indicates that the Eu3+ ion occupies non-centrosymmetric sites with C, Cn or Cs symmetry [50].

CIE Color Coordinates and CCT (K) Values

To better understand the luminescent behavior and the actual color of emissions, the standard Commission Internationale de l’Eclairage (CIE) 1931 chromaticity diagram was used [51]. The color chromaticity coordinates of the synthesized glasses are calculated from the PL spectra (Figure 7), using SpectraChroma software, Version 1.0.1 (CIE coordinate calculator) [52] and are shown at Figure 8. The obtained values of glasses with different WO3 content, enlisted at Table 4, are almost similar and are very close to the CIE coordinate of standard red light (0.67, 0.33) and to the color coordinates of the commercial red phosphor Y2O2S:Eu3+ (0.658; 0.340) [53].
The values of correlated color temperature (CCT) were calculated by the McCamy empirical expression [54]:
C C T = 449 n 3 + 3525 n 2 6823 n + 5520.33 ;
n = ( x x e ) ( y y e )
Here, xe = 0.332, ye = 0.186 are the epicenter coordinates and x and y are the calculated color chromaticity coordinates. In our case, the calculated CCT values shown in Table 4 are in the range of 2270.30 K ÷ 2439.74 K, and the Eu3+-doped glasses can be referred to as warm light-emitting materials.

3. Discussion

Raman analysis revealed that tungstate ions incorporate into the base zinc borate glass as tetrahedral [WO4]2− groups, and octahedral [WØ4O2]2− species with four bridging and two non-bridging oxygen atoms. There are also metaborate, [BØ2O] and pyroborate units, [B2O5]4− in the glass networks. Tungstate, metaborate and pyroborate units are charge balanced by Zn2+ and Eu3+ ions via Zn-O-W, Zn-O-B, Eu-O-W and Eu-O-B bonding. As it was proved in our previous paper, W-O-B bonding is not possible [25]. Thus, the investigated glasses are characterized by a heterogeneous structure, because in the main zinc borate glass network, regions rich in tungsten are formed. Structural heterogeneity in these glasses increases with increasing WO3 content, as the number of tungstate structural units, as well as Zn-O-W bonds, increase. This observation has been also confirmed by DCS analysis, wherein two glass transition effects have been noted in the DSC curve of glasses, due to the presence of two amorphous phases with different compositions. On the other hand, the addition of WO3 into the base zinc borate glass causes depolymerization of the borate oxygen network, i.e., increasing the number of non-bridging oxygen atoms. In this way, with increasing WO3 concentrations, the structural disorder in the amorphous network increases, and also, a more loosened glass structure is formed. The established structural features of the investigated glasses determine their good emission properties. All glasses are characterized by a high-intensity red luminescence band at 612 nm corresponding to the forced electric dipole transition (ED) 5D0  7F2, which is stronger for the WO3-containing glasses as compared with the WO3-free glass. The addition of WO3 to the base zinc borate glass improves the Eu3+’s luminescence properties, which can be explained by the presence of both borate and tungstate units in the active ions’ surroundings, which increases the Eu3+’s site asymmetry and hence the Eu3+’s emission intensity. Also, the presence of tungstate groups around Eu3+ ions ensures an occurrence of non-radiative energy transfer from tungstate units to the active ions, which additionally improves the Eu3+’s luminescence behavior. The observed maximum of Eu3+ emission for glass x = 5 shows that this glass composition ensures the most appropriate glass structure for accommodating the active ion. With a further increase in WO3 content of 10 mol %, glass network heterogeneity is too high to promote the good distribution of the rare earth ions in the matrix, and, thus, the emission properties of europium are reduced. On the other hand, with the incorporation of a higher amount of WO3, the average distance between tungstate groups would be decreased, which would make the energy transfer between them more sufficient and, thus less energy would be expected to be transferred to the Eu3+ ions [55].

4. Materials and Methods

Glasses with nominal compositions 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3, (x = 0, 1, 3, 5 and 10 mol %) were obtained by applying the conventional melt-quenching method, using commercial powders of reagent-grade WO3 (Merck KGaA, Darmstadt, Germany), ZnO (Merck KGaA, Amsterdam, The Netherlands), H3BO3 (SIGMA-ALDRICH, St. Louis, MO, USA), and Eu2O3 (SIGMA-ALDRICH, St. Louis, MO, USA) as starting materials. The details of glass synthesis were given in refs. [13,14]. The glass transition (Tg) temperatures of the glasses were determined by differential scanning calorimetry (DSC) using a Netzsch 404 Pegasus instrument, 2021 Selb, Germany, at a heating rate of 10 K/min in an Ar flow of 10 mL/s, using corundum crucibles with lids. Optical transmission spectra at room temperature for the glasses were measured by a spectrometer (Ocean Optics, HR 4000, Duiven, 2010, The Netherlands) using a UV LED light source at 385 nm. Photoluminescence (PL) excitation and emission spectra at room temperature for all glasses were measured with a Spectrofluorometer FluoroLog3-22, 2014 (Horiba JobinYvon, Longjumeau, France). Raman spectra were recorded with a Raman spectrometer (Delta NU, Advantage NIR 785 nm, 2010, Midland, ON, Canada).

5. Conclusions

The present investigation demonstrates the relationship between the host glass structure and the optical properties of Eu3+-doped glasses 50ZnO:(50 − x)B2O3:xWO3, x = 0 1, 3, 5 and 10 mol %. The result indicates that the addition of WO3 of up to 5 mol % to the base zinc borate glass improves the luminescence intensity of doped rare earth ions, which is attributed to the formation of a more disordered glass network where Eu3+ ions are surrounded by both borate and tungstate units, which ensures a highly asymmetric local structure around Eu3+ ions sites, accordingly yielding a strong red emission of the active ions. A further increase in the tungsten content leads to luminescent quenching. All findings obtained here are favorable for the elaboration of novel red-emitting glass materials.

Author Contributions

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

Funding

The authors are grateful for the financial support of project BG05M2OP001-1.001-0008, “National Center of mechatronics and clean technologies”, funded by the Operational Programme Science and Education for Smart Growth and co-financed by the European Union through the European Regional Development Fund.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nagami, M.; Abe, Y. Properties of sol—Gel-derived Al2O3 SiO2 glasses using Eu3+ ion fluorescence spectra. J. Non-Cryst. Solids 1996, 197, 73–78. [Google Scholar] [CrossRef]
  2. Weber, M.J. Science and technology of laser glass. J. Non-Cryst. Solids 1990, 123, 208–222. [Google Scholar] [CrossRef]
  3. Bengisu, M. Borate glasses for scientific and industrial applications: A review. J. Mater. Sci. 2016, 51, 2199–2242. [Google Scholar] [CrossRef]
  4. Mitailipu, M.; Poeppelmeier, K.; Pan, S. Borates: A Rich Source for Optical Materials. Chem. Rev. 2021, 121, 1130–1202. [Google Scholar] [CrossRef] [PubMed]
  5. Colak, C.S.; Akyuz, I.; Atay, F. On the dual role of ZnO in zinc-borate glasses. J. Non-Cryst. Solids 2016, 432, 406–412. [Google Scholar] [CrossRef]
  6. Annapurna, K.; Das, M.; Kundu, M.; Dwivedhi, R.N.; Buddhudu, S. Spectral properties of Eu3+: ZnO–B2O3–SiO2 glasses. J. Mol. Struct. 2005, 741, 53–60. [Google Scholar] [CrossRef]
  7. Lee, J.D. Concise Inorganic Chemistry; Blackwell Science: Oxford, UK, 1996. [Google Scholar]
  8. Prasad, R.N.A.; Siva, B.V.; Neeraja, K.; Mohanq, N.K.; Rojas, J.I. Effect of modifier oxides on spectroscopic and optical properties of Pr3+doped PbO-Ro2O3–WO3–B2O3 glasses (with Ro2O = Sb2O3, Al2O3, and Bi2O3). J. Lumin. 2021, 230, 117666. [Google Scholar] [CrossRef]
  9. Prasad, R.N.A.; Siva, B.V.; Neeraja, K.; Mohanq, N.K.; Rojas, J.I. Influence of modifier oxides on spectroscopic features of Nd2O3 doped PbO-Ro2O3–WO3–B2O3 glasses (with Ro2O3 = Sb2O3, Al2O3, and Bi2O3). J. Lumin. 2020, 223, 117171. [Google Scholar] [CrossRef]
  10. Dousti, M.R.; Poirier, G.Y.; Camargo, A.S.S. Structural and spectroscopic characteristics of Eu3+-doped tungsten phosphate glasses. Opt. Mater. 2015, 45, 185–190. [Google Scholar] [CrossRef]
  11. Ataalla, M.A.; Afify, S.; Hassan, M.; Abdallah, M.; Milanova, M.; Aboul-Enein, H.Y.; Mohamed, A. Tungsten-based glasses for photochromic, electrochromic, gas sensors, and related applications: A. review. J. Non-Cryst. Solids 2018, 491, 43–54. [Google Scholar] [CrossRef]
  12. 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]
  13. 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]
  14. 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. [Google Scholar] [CrossRef] [PubMed]
  15. Swapna, G.; Upender, M.; Prasad, M. Raman FTIR, thermal and optical properties of TeO2-Nb2O5-B2O3-V2O5 quaternary glass system. J. Taibah Univ. Sci. 2017, 11, 583–592. [Google Scholar] [CrossRef]
  16. Ray, N.H. Composition-properties relationship in Inorganic Oxide Glasses. J. Non-Cryst. Solids 1974, 15, 423–434. [Google Scholar] [CrossRef]
  17. Saddeek, Y.; Azooz, M.; Saddek, A. Ultrasonic investigations of some bismuth borate glasses doped with Al2O3. Bull. Mater. Sci. 2015, 38, 241–246. [Google Scholar] [CrossRef]
  18. Saddeek, Y. Effect of B2O3 on the structure and properties of tungsten-tellurite glasses. Philos. Mag. 2009, 89, 41–54. [Google Scholar] [CrossRef]
  19. Kaur, A.; Khanna, A.; González-Barriuso, M.; González, F.; Chen, B. Short-range structure and thermal properties of alumino-tellurite glasses. J. Non-Cryst. Solids 2017, 470, 14–18. [Google Scholar] [CrossRef]
  20. Cottrell, T.L. The Strength of Chemical Bonds, 2nd ed.; Butterworth: London, UK, 1958. [Google Scholar]
  21. Topper, B.; Möncke, D.; Youngman, R.E.; Valvi, C.; Kamitsos, E.I.; Varsamis, C.P. Zinc borate glasses: Properties, structure and modelling of the composition-dependence of borate speciation. Phys. Chem. Chem. Phys. 2023, 25, 5967–5988. [Google Scholar] [CrossRef]
  22. Yao, Z.Y.; Möncke, D.; Kamitsos, E.I.; Houizot, P.; Célarié, 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]
  23. Kamitsos, E.I.; Karakassides, M.A.; Chryssikos, G.D. Vibrational Spectra of Magnesium-Sodium-Borate Glasses. 2. Raman and Mid-Infrared Investigation of the Network Structure. J. Phys. Chem. 1987, 91, 1073–1079. [Google Scholar] [CrossRef]
  24. Mancheva, M.; Iordanova, R.; Dimitriev, Y. Mechanochemical synthesis of nanocrystalline ZnWO4 at room temperature. J. Alloys Compd. 2011, 509, 15–20. [Google Scholar] [CrossRef]
  25. Aleksandrov, L.; Komatsu, T.; Shinozaki, K.; Honma, T.; Iordanova, R. Structure of MoO3–WO3–La2O3–B2O3 glasses and crystallization of LaMo1 − xWxBO6 solid solutions. J. Non-Cryst. Solids 2015, 429, 171–177. [Google Scholar] [CrossRef]
  26. 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]
  27. Nimpoeno, W.A.; Lintang, H.O.; Yuliati, L. Zinc oxide with visible light photocatalytic activity originated from oxygen vacancy defects. IOP Conf. Ser. Mater. Sci. Eng. 2020, 833, 012080. [Google Scholar] [CrossRef]
  28. Blasse, G.; Grabmaier, B.C. Luminescent Materials, 1st ed.; Springer: Berlin/Heidelber, Germany, 1994; p. 18. [Google Scholar]
  29. Hoefdraad, H.E. The charge-transfer absorption band of Eu3+ in oxides. J. Solid State Chem. 1975, 15, 175–177. [Google Scholar] [CrossRef]
  30. Parchur, A.K.; Ningthoujam, R.S. Behaviour of electric and magnetic dipole transitions of Eu3+, 5D0-7F0 and Eu-O charge transfer band in Li+ co-doped YPO4:Eu3+. RSC Adv. 2012, 2, 10859–10868. [Google Scholar] [CrossRef]
  31. Mariselvam, K.; Liu, J. Synthesis and luminescence properties of Eu3+ doped potassium titano telluroborate (KTTB) glasses for red laser applications. J. Lumin. 2021, 230, 117735. [Google Scholar] [CrossRef]
  32. Binnemans, K. Interpretation of europium (III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef]
  33. Thieme, C.; Herrmann, A.; Kracker, M.; Patzig, C.; H¨oche, 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]
  34. 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]
  35. Sreena, T.S.; Raj, A.K.; Rao, P.P. Effects of charge transfer band position and intensity on the photoluminescence properties of Ca1.9M2O7: 0.1Eu3+ (M = Nb, Sb and Ta). Solid State Sci. 2022, 123, 106783. [Google Scholar] [CrossRef]
  36. Lei, F.; Yan, B.; Chen, H.H. Solid-state synthesis, characterization and luminescent properties of Eu3+-doped gadolinium tungstate and molybdate phosphors: Gd(2−x)MO6:Eux3+(M = W, Mo). J. Solid State Chem. 2008, 181, 2845–2851. [Google Scholar] [CrossRef]
  37. Sun, X.Y.; Jiang, D.G.; Chen, S.W.; Zheng, G.T.; Huang, S.M.; Gu, M.; Zhao, J.T. Eu3+- activated borogermanate scintillating glass with a high Gd2O3 content. J. Am. Ceram. Soc. 2013, 96, 1483–1489. [Google Scholar] [CrossRef]
  38. Pang, M.; Liu, X.; Lin, J. Luminescence properties of R2MoO6: Eu3+ (R= Gd, Y, La) phosphors prepared by Pechini sol-gel process. J. Mater. Res. 2005, 20, 2676–2681. [Google Scholar] [CrossRef]
  39. 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]
  40. Blasse, G. On the Eu3+ fluorescence of mixed metal oxides. IV. The photoluminescent efficiency of Eu3+-activated oxides. J. Chem. Phys. 1996, 45, 2356–2360. [Google Scholar] [CrossRef]
  41. Dejneka, M.; Snitzer, E.; Riman, R.E. Blue, green and red fluorescence and energy transfer of Eu3+ in fluoride glasses. J. Lumin. 1995, 65, 227–245. [Google Scholar] [CrossRef]
  42. Iordanova, R.; Milanova, M.; Yordanova, A.; Aleksandrov, L.; Nedylkov, N.; Kukeva, R.; Petrova, P. Structure and Luminescent Properties of Niobium Modified ZnO-B2O3: Eu3+ Glass. Materials 2024, 17, 1415. [Google Scholar] [CrossRef]
  43. 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]
  44. Babu, S.S.; Jang, K.; Cho, E.J.; Lee, H.; Jayasankar, C.K. Thermal, structural and optical properties of Eu3+-doped zinc-tellurite glasses. J. Phys. D Appl. Phys. 2007, 40, 5767. [Google Scholar] [CrossRef]
  45. Adamu, S.B.; Halimah, M.K.; Chan, K.T.; Muhammad, F.D.; Nazrin, S.N.; Tafida, R.A. Eu3+ ions doped zinc borotellurite glass system for white light laser application: Structural, physical, optical properties, and Judd-Ofelt theory. J. Lumin. 2022, 250, 119099. [Google Scholar] [CrossRef]
  46. Babu, P.; Jayasankar, C.K. Optical spectroscopy of Eu3+ ions in lithium borate and lithium fluoroborate glasses. Phys. B Condens. Matter. 2000, 279, 262–281. [Google Scholar] [CrossRef]
  47. Nageno, Y.; Takebe, H.; Morinaga, K.; Izumitani, T.J. Effect of modifier ions on fluorescence and absorption of Eu3+ in alkali and alkaline earth silicate glasses. Non-Cryst. Solids 1994, 169, 288. [Google Scholar] [CrossRef]
  48. Kawano, N.; Shinozaki, K.; Kato, T.; Onoda, D.; Takebuchi, Y.; Fukushima, H.; Yanagida, T. Scintillation characteristics of Eu2O3-doped WO3–Al2O3–TeO2 glasses. J. Lumin. 2022, 249, 119003. [Google Scholar] [CrossRef]
  49. Wu, L.; Zhang, F.; Wu, L.; Yi, H.; Wang, H.; Zhang, Y.; Xu, J. Structure refinement and one-center luminescence of Eu3+ activated ZnBi2B2O7 under UV excitation. J. Alloys Compd. 2015, 648, 500–506. [Google Scholar] [CrossRef]
  50. Binnemans, K.; Görller-Walrand, C. Application of the Eu3+ ion for site symmetry determination. J. Rare Earths 1996, 14, 173–180. [Google Scholar]
  51. Smith, T.; Guild, J. The CIE colorimetric standards and their use. Trans. Opt. Soc. 1931, 33, 73. [Google Scholar] [CrossRef]
  52. Paolini, T.B. SpectraChroma (Version 1.0.1) [Computer Software]. 2021. Available online: https://zenodo.org/records/4906590 (accessed on 7 June 2021).
  53. 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]
  54. McCamy, C.S. Correlated color temperature as an explicit function of chromaticity coordinates. Color Res. Appl. 1992, 17, 142–144. [Google Scholar] [CrossRef]
  55. Xue, J.; Noh, H.; Choi, B.; Park, S.; Jeong, J.; Kim, J. Molybdenum substitution induced luminescence enhancement in Gd2W1-xMoxO6Eu3+ phosphors for near ultraviolet based solid state lighting. J. Lunin. 2018, 202, 97–106. [Google Scholar]
Figure 1. DSC curves of the obtained glasses.
Figure 1. DSC curves of the obtained glasses.
Molecules 29 02470 g001
Figure 2. Raman spectra of the obtained glasses.
Figure 2. Raman spectra of the obtained glasses.
Molecules 29 02470 g002
Figure 3. Optical transmission spectra at room temperature, (a), and absorption coefficient in the range of 150 nm–900 nm, (b) for studied glasses.
Figure 3. Optical transmission spectra at room temperature, (a), and absorption coefficient in the range of 150 nm–900 nm, (b) for studied glasses.
Molecules 29 02470 g003
Figure 4. Excitation spectra of 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10) glasses.
Figure 4. Excitation spectra of 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10) glasses.
Molecules 29 02470 g004
Figure 5. Excitation (black line) and emission (red line) spectra of 0.5 mol % Eu3+-doped 50ZnO:45B2O3:5WO3 glass.
Figure 5. Excitation (black line) and emission (red line) spectra of 0.5 mol % Eu3+-doped 50ZnO:45B2O3:5WO3 glass.
Molecules 29 02470 g005
Figure 6. Emission spectra of host matrix 50ZnO:40B2O3:10WO3 (black line) and Eu3+-doped 50ZnO:40B2O3:10WO3 glass (red line) at 248 nm excitation.
Figure 6. Emission spectra of host matrix 50ZnO:40B2O3:10WO3 (black line) and Eu3+-doped 50ZnO:40B2O3:10WO3 glass (red line) at 248 nm excitation.
Molecules 29 02470 g006
Figure 7. PL emission spectra of glasses 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10).
Figure 7. PL emission spectra of glasses 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10).
Molecules 29 02470 g007
Figure 8. CIE chromaticity diagram of the 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (a) x = 0, (b) x = 1, (c) x = 3, (d) x = 5, (e) x = 10 glasses.
Figure 8. CIE chromaticity diagram of the 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (a) x = 0, (b) x = 1, (c) x = 3, (d) x = 5, (e) x = 10 glasses.
Molecules 29 02470 g008
Table 1. Values of glass transition temperatures (Tg) and average single-bond enthalpy (EB) of the investigated glasses.
Table 1. Values of glass transition temperatures (Tg) and average single-bond enthalpy (EB) of the investigated glasses.
Sample IDTg1/°CTg2/°CEB/kJ mol−1
x = 0570 (843 K)-548
x = 1568 (841 K)688 (961 K)546
x = 3565 (838 K)680 (953 K)543
x = 5543 (816 K)653 (926 K)540
x = 10532 (805 K)626 (899 K)532
Table 2. Peak positions in the Raman spectra of glasses 50ZnO:(50 − x)B2O3:0.5Eu2O3:xWO3, (x = 0, 1, 3, 5 and 10 mol %) and their assignments.
Table 2. Peak positions in the Raman spectra of glasses 50ZnO:(50 − x)B2O3:0.5Eu2O3:xWO3, (x = 0, 1, 3, 5 and 10 mol %) and their assignments.
Peak PositionsAssignmentsRef.
260ν (Zn-O)[21,22]
315ν (Eu-O) + ν2 [WO4]2[13]
355ν4 [WO4]2−[13]
400νs[W2O4]n[13]
663νas[W2O4]n[13]
690νas[W2O4]n[13]
705δ[BØ2O] + in-plane and out-of-plane bending modes of both polymerized (BØ0) species and isolated orthoborate units (BO3)3-+ bending of the B-O-B connection in the pyroborate dimers, [B2O5]4−[21,22,23]
800ring breathing of the boroxol rings+ ν3[WO4]2−[13,21,22,23]
840νas(WO2) + B-O-B stretching of [B2O5]4−[13]
865νs(WO2)[13]
870B-O-B bridges in pyroborate dimers, [B2O5]4−[21,22,23]
970ν1[WO4]2−[13]
1240B-O stretch in pyroborate units[21,22,23]
1380–1320B-O stretch in metaborate units[21,22,23]
1400B-O stretch in metaborate units[21,22,23]
Table 3. Values of the relative intensity ratio (R) for 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10) glasses.
Table 3. Values of the relative intensity ratio (R) for 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10) glasses.
Glass CompositionRelative Luminescent Intensity Ratio, RReference
50ZnO:50B2O3:0.5Eu2O34.34Current work
50ZnO:49B2O3:1WO3:0.5Eu2O35.67Current work
50ZnO:47B2O3:3WO3:0.5Eu2O35.71Current work
50ZnO:45B2O3:5WO3:0.5Eu2O35.82Current work
50ZnO:40B2O3:10WO3:0.5Eu2O35.57Current work + [13]
50ZnO:(50 − x)B2O3: xNb2O5:0.5Eu2O3:, x = 0, 1, 3 and 5 mol %4.31–5.16[42]
50ZnO:40B2O3:10WO3:xEu2O3 (0 ≤ x ≤ 10)4.54÷5.77[13]
50ZnO:40B2O3:5WO3:5Nb2O5:xEu2O3 (0 ≤ x ≤ 10)5.09÷5.76[34]
4ZnO:3B2O3 0.5–2.5 mol % Eu2O32.74–3.94[43]
60TeO2:39ZnO:1Eu2O33.25[44]
60TeO2:20ZnO:19LiF:1Eu2O33.70[44]
60TeO2:19ZnO:10Na2O:10Li2O:1Eu2O33.73[44]
60ZnO:20B2O3:(20 x)SiO2−xEu2O3 (x = 0 and 1)3.166[6]
[{(TeO2)0.7:(B2O3)0.3}0.7:(ZnO)0.3](1−y)(Eu2O3)y with 0.01 = y ≤ 0.052.15–3.08[45]
59.5Li2O:39.5B2O3:1Eu2O34.18[46]
19.5Na2O:20MgO:59.5SiO2:1Eu2O34.43[47]
(15 − x)WO3–5Al2O3–80TeO2–xEu2O3; x = 0.1÷5 mol %5.4÷6[48]
Table 4. CIE chromaticity coordinates and correlated color temperatures (CCTs, K) of 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10).
Table 4. CIE chromaticity coordinates and correlated color temperatures (CCTs, K) of 50ZnO:(50 − x)B2O3:xWO3:0.5Eu2O3 (x = 0, 1, 3, 5 and 10).
Glass CompositionChromaticity Coordinates (x, y)CCT (K)
50ZnO:50B2O3:0.5Eu2O30.645, 0.3462301.26
50ZnO:49B2O3:1WO3:0.5Eu2O30.651, 0.3482324.59
50ZnO:47B2O3:3WO3:0.5Eu2O30.650, 0.3502270.30
50ZnO:45B2O3:5WO3:0.5Eu2O30.652, 0.3472359.79
50ZnO:40B2O3:10WO3:0.5Eu2O30.654, 0.3452439.74
NTSC standard for red phosphors0.670, 0.330
Y2O2S:Eu3+0.658, 0.340
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.; Aleksandrov, L.; Milanova, M.; Iordanova, R.; Petrova, P.; Nedyalkov, N. Effect of the Addition of WO3 on the Structure and Luminescent Properties of ZnO-B2O3:Eu3+ Glass. Molecules 2024, 29, 2470. https://doi.org/10.3390/molecules29112470

AMA Style

Yordanova A, Aleksandrov L, Milanova M, Iordanova R, Petrova P, Nedyalkov N. Effect of the Addition of WO3 on the Structure and Luminescent Properties of ZnO-B2O3:Eu3+ Glass. Molecules. 2024; 29(11):2470. https://doi.org/10.3390/molecules29112470

Chicago/Turabian Style

Yordanova, Aneliya, Lyubomir Aleksandrov, Margarita Milanova, Reni Iordanova, Petia Petrova, and Nikolay Nedyalkov. 2024. "Effect of the Addition of WO3 on the Structure and Luminescent Properties of ZnO-B2O3:Eu3+ Glass" Molecules 29, no. 11: 2470. https://doi.org/10.3390/molecules29112470

Article Metrics

Back to TopTop