Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Judd–Ofelt analysis and the spectroscopic elucidation of β-BaB2O4:Eu3+ phosphors for optoelectronic applications

R. Kiran a, Vaishnavi Shenoya, M. I. Sayyedbcd, Aljawhara H. Almuqrine and Sudha D. Kamath*a
aDepartment of Physics, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal, Karnataka, India. E-mail: sudha.kamath@manipal.edu
bRenewable Energy and Environmental Technology Center, University of Tabuk, Tabuk, 47913, Saudi Arabia
cDepartment of Physics, Faculty of Science, Isra University, Amman, Jordan
dDepartment of Physics and Technical Sciences, Western Caspian University, Baku, Azerbaijan
eDepartment of Physics, College of Science, Princess Nourah Bint Abdulrahman University, P.O. Box 84428, Riyadh, 11671, Saudi Arabia

Received 7th August 2025 , Accepted 1st November 2025

First published on 7th November 2025


Abstract

We have used the solid-state reaction method to prepare Eu3+ (1, 1.5, 2, 2.5, and 3 mol%) doped β-BaB2O4 phosphors. Phase purity of the prepared phosphors was confirmed using X-ray diffraction (XRD) studies. Photoluminescence (PL) measurements using 395 nm excitation exhibited three intense peaks at 593 nm, 615 nm, 653 nm, and 702 nm, respectively, due to the 5D07Fj (j = 0, 1, and 2) transitions. The PL emission intensity will increase up to 2 mol%, and thereafter, there will be a decline due to concentration quenching. Chromaticity analysis, along with correlated colour temperature (1796 K to 1976 K), confirmed the material's potential as a warm light emitter. Diffuse reflectance spectroscopy revealed a direct bandgap energy of 4.30 eV for the optimized sample. Finally, using Judd–Ofelt intensity parameters, the calculation of critical radiative parameters, including transition probability, lifetime, and branching ratio, was carried out. The results demonstrate the efficient luminescence and excellent colour stability of β-BaB2O4:Eu3+, highlighting its potential for optoelectronic applications.


1 Introduction

In the span of the past century, advances in materials science have transformed technologies from silicon-based electronics and lightweight composites to high-performance polymers, reshaping various aspects of life.1–4 These innovations have enabled faster computing, sustainable energy solutions, fundamentally altering how we live and interact with the world.5–7 Among the various novel materials, phosphors doped with rare-earth (RE) elements have revolutionized lighting and display technologies by converting ultraviolet or blue excitation into vibrant visible emissions. From energy-efficient fluorescent lamps and high-definition plasma screens to advanced bioimaging probes, phosphors have enabled more sustainable and more versatile applications that touch virtually every aspect of modern life.8–11 Among the various phosphors, RE-doped borate phosphors have emerged as highly attractive materials, owing to their robust physicochemical stability, broad vacuum-ultraviolet transparency, and elevated resistance to photodamage. The flexible crystal lattices of the borate group readily incorporate a variety of RE activator ions, which enables the fine-tuning of emission properties. Furthermore, the emissive behaviour of these phosphors is known to be sensitive to parameters such as synthesis route, particle size and morphology, host lattice symmetry, and activator site occupancy. Such tunability makes RE borate phosphors an ideal candidate for plasma display panels, mercury-free fluorescent lamps, and next-generation solid-state lighting devices.12–14

In the present study, β-BaB2O4 was selected as the host matrix for phosphor development and doped with varying concentrations of RE ions. Previous investigations involving this host have demonstrated its suitability for luminescent applications, including systems such as β-BaB2O4:Dy3+,15 β-BaB2O4:Sm3+,16 β-BaB2O4:Nd3+, Yb3+,17 and β-BaB2O4:Pb2+, Cu2+.18 In this work, Eu3+ was selected as the activator due to its high chemical reactivity and stable trivalent oxidation state under ambient conditions. The Eu3+ ion exhibits sharp emissions due to the 5D07FJ (J = 0, 1, 2, 3, 4, 5, 6) electronic transitions, making it highly suitable for red luminescence.19 As a result, Eu3+ activated phosphors are considered to be a promising contender for efficient red spectral component in white light-emitting diode (WLED) technologies.20–23 Prior studies by Jie Liu et al. and Zhihua Li et al. have mainly focused on the room temperature optical properties of Eu3+ doped β-BaB2O4 phosphors. Jie Liu et al. synthesized β-BaB2O4:xEu3+ (x = 0.00, 0.02, 0.04, 0.06, 0.08, 0.10) phosphors and evaluated the impact of various charge compensating ions (K+, Na+, and Li+) on PL emission intensity.24 Their findings indicated that incorporating charge compensators enhanced PL intensity, with K+ yielding the most significant improvement. However, this study did not include a comprehensive analysis of the crystal structure via XRD, the reflectance measurements, and the thermal stability. In another effort, Zhihua Li et al. synthesized BaB2O4:Eu3+ phosphors and explored similar optical properties.25 Nonetheless, the reported XRD data exhibited high background noise and secondary phase peaks, suggesting the presence of impurities. Additionally, the selected dopant concentrations (5.69, 7.48, 10.30, 12.00, 16.10, 18.90, 20.60, and 21.30 mol%) deviate drastically from the range reported by Jie Liu et al., and the study lacked both reflectance analysis and the stability of the phosphor under elevated temperature conditions. Also, none of the reported studies carried out Judd–Ofelt (JO) analysis and the evaluation of the radiative parameters.

Therefore, in the present investigation, β-BaB2O4:xEu3+ (x = 1, 1.5, 2, 2.5, and 3 mol%) phosphors were prepared via the solid-state reaction method. This work aims to bridge existing research gaps by offering a systematic and comprehensive investigation of the material's optical, reflectance, colour tunability, structural properties, and thermal stability. In addition, using the JO theory, radiative parameters were evaluated for the prepared phosphors to check their suitability for various optoelectronic applications.

2 Experimental

2.1 Phosphor synthesis

We have prepared β-BaB2O4:Eu3+ phosphors via the solid-state reaction method. Initially, precisely weighted BaCO3 (99%), H3BO3 (99%), and Eu2O3 (99.99%) are taken in stoichiometric amounts and carefully ground to ensure uniform distribution and improved homogeneity throughout the mixture. Following this, the resulting mixtures were transferred to an alumina crucible and calcinated at 800 °C for three hours. Finally, the samples were cooled to room temperature and readied for further characterization.

2.2 Material analysis techniques

Crystal structure and the phase purity of the synthesized phosphors were verified through XRD analysis, conducted over a 2θ range of 15–70° using a Rigaku Miniflex 600 diffractometer. PL excitation and emission spectra were subsequently recorded using a JASCO FP-800 spectrofluorometer. Optical reflectance and absorption spectra were evaluated via the diffuse reflectance spectroscopy (DRS) mode with the help of a PerkinElmer Lambda 900 spectrophotometer. Finally, the thermal stability of the sample was assessed through thermogravimetric analysis (TGA) using a PerkinElmer TGA 4000, with measurements conducted up to 500 °C.

3 Results and discussion

3.1 Analysis of phosphor structure

XRD analysis confirmed the crystalline phase of the β-BaB2O4:Eu3+ phosphors. As shown in Fig. 1, the diffraction pattern exhibits peak positions that are well-aligned with those of the standard reference (JCPDS card no. 80-1049) patterns, indicating the successful formation of pure β-BaB2O4:Eu3+ phase. The absence of secondary or impurity peaks suggests that the incorporation of Eu3+ ions has a negligible impact on the structure of the host crystal lattice. A closer examination of the XRD pattern between 24° and 27° in Fig. 1 shows that, at 1.5 mol% Eu3+ doping, the XRD peaks shift slightly toward higher 2θ values. This behaviour can be attributed to the substitution of larger Ba2+ ions by smaller Eu3+ ions, leading to a contraction of the unit cell.26 Interestingly, at higher doping concentrations of 2.5 mol% and 3.0 mol% Eu3+, the peaks shift toward lower 2θ values. This anomalous behaviour suggests that, beyond a certain concentration, Eu3+ ions preferentially occupy interstitial sites rather than substituting Ba2+ ions. Occupation of interstitial sites expands the interplanar spacing and increases the unit cell volume, resulting in the observed shift of diffraction peaks to lower angles.27
image file: d5ra05778g-f1.tif
Fig. 1 Experimental XRD patterns overlaid with standard reference peaks confirming the crystalline phase of the phosphors.

This can be understood using the acceptable percentage difference (Ra) between the host and dopant ions should generally remain below 30% to ensure successful substitution without significant lattice distortion.28 It can be evaluated through the expression given below29

 
image file: d5ra05778g-t1.tif(1)
Here, Rh represents the ionic radius of the host cations, REu is that of the Eu3+ ion, and CN stands for the coordination number. The values were calculated based on Shannon ionic radius data, with the outcomes presented in Fig. 2(a). As shown, the estimated Ra value for the Eu3+–Ba2+ ion pair falls within the permissible threshold, indicating a compatible ionic size match. Therefore, Eu3+ ions will successfully substitute Ba2+ sites within the host lattice without causing significant structural distortion. Additionally, we have drawn the visual representation of the unit using the reference file, and it is provided as Fig. 2(b).


image file: d5ra05778g-f2.tif
Fig. 2 (a) Acceptable percentage difference in radii (b) pictorial representation of the unit cell.

Subsequently, the crystallite size (D) of the synthesized phosphor was estimated using the Size–Strain Plot (SSP) method. In this approach, it is proposed that the broadening due to the crystallite size follows a Lorentzian distribution, while the strain-induced broadening by a Gaussian function. Based on these assumptions, the following relationship was applied30

 
image file: d5ra05778g-t2.tif(2)

In this equation, λ denotes the wavelength of the X-ray, β refers to the FWHM of the peak, and d indicates the spacing between lattice planes. To evaluate the crystallite size, a plot of (βd[thin space (1/6-em)]cos[thin space (1/6-em)]θ)2 versus (βd2[thin space (1/6-em)]cos[thin space (1/6-em)]θ) was constructed for all major diffraction peaks for β-BaB2O4:xEu3+ (x = 1, 1.5, 2, 2.5, and 3 mol%) phosphors, and it is shown in Fig. 3(a)–(e). From the linear fit of each dataset, the crystallite sizes were calculated using the slope of the fitted line. The estimated crystallite sizes were found to be 14.4 nm, 11.6 nm, 14.4 nm, 16.1 nm, and 16.1 nm, respectively, when the concentration of Eu3+ varies from 1 mol% to 3 mol%.


image file: d5ra05778g-f3.tif
Fig. 3 (a)–(e) SSP plot for the phosphor with dopant concentration.

3.2 Phosphor optimization and analysis of PL properties

To comprehensively understand the optical characteristics of the phosphors, the excitation spectrum was recorded at room temperature. The PL excitation spectrum corresponding to 2 mol% Eu3+ is illustrated in Fig. 4 using an emission wavelength at 615 nm across the 240–500 nm wavelength range. The spectrum reveals key insights into the electronic transitions involved in the luminescence mechanism. A prominent broad excitation band is observed in the 240 to 290 nm region, which is attributed to the charge transfer band (CTB) associated with the Eu3+ ions. This CTB originates from an electron transition between the 2p orbital of the oxygen anion to the partially filled 4f orbitals of the Eu3+ ion.
image file: d5ra05778g-f4.tif
Fig. 4 Excitation spectrum corresponding to the β-BaB2O4:xEu3+ (x = 2 mol%) phosphors.

This type of transition is characteristic of RE-doped oxide materials, and its presence indicates strong covalent interactions between the host lattice and the dopant ion.31 In addition to the CTB, sharp excitation peaks are detected in the 295–480 nm range, signifying the intra-4f transitions. These peaks are observed at 299, 321, 363, 384, 395, 416, and 467 nm, corresponding to the intra-4f transitions from the ground state (7F0) to the excited states (5F4, 5H6, 5L8, 5L7, 5L6, 5D3, and 5D2).32 Although f–f electronic transitions are generally forbidden by parity selection rules, the local asymmetry of the crystal field surrounding the Eu3+ ions relaxes these restrictions.33 Among the observed excitation peaks, the one centered at 395 nm displayed the strongest intensity and was consequently selected as the optimal excitation wavelength.

As illustrated in Fig. 5, the emission spectra display multiple sharp and well-defined peaks in the 550–750 nm wavelength range, characteristic of Eu3+ transitions. These peaks originate from the intra-configurational f–f transitions of Eu3+ ions, specifically from the excited 5D0 level to the various sublevels of the 7FJ ground state (J = 0–4). Among these transitions, the most intense emission, centered around 615 nm (5D07F2), corresponds to the 5D07F2 transition. Additional emission bands are observed at approximately 581 nm (5D07F0), 593 nm (5D07F1), 653 nm (5D07F3), and 702 nm (5D07F4), respectively.34 Fig. 6 illustrates the simplified energy transfer pathway from O2− → Eu3+ along with the detailed emission mechanism of Eu3+ ions. The corresponding intensity variations of the emission peaks at 593, 615, and 702 nm are presented in Fig. 7(a).35–37


image file: d5ra05778g-f5.tif
Fig. 5 PL emission spectra for β-BaB2O4:xEu3+ with excitation at 395 nm.

image file: d5ra05778g-f6.tif
Fig. 6 Energy level diagram of Eu3+ describing various emission peaks.

image file: d5ra05778g-f7.tif
Fig. 7 (a) Variation of intensity for the peaks at 593, 615, and 702 with the concentration of Eu3+. (b) Concentration dependent asymmetry ratio.

Notably, the 5D07F2 transition is an electric dipole (ΔJ = 2) transition, often termed a hypersensitive transition due to its pronounced sensitivity to the local crystal field surrounding Eu3+ ions. In contrast, the 5D07F1 transition, identified as a magnetic dipole transition, is relatively insensitive to variations in the local crystal field environment. Based on JO theory, such dominant electric dipole transition suggests that Eu3+ ions are situated at non-centrosymmetric sites within the host lattice. In non-centrosymmetric environments, the electric dipole transitions become parity-allowed, particularly evident in the 610–630 nm emission range, while magnetic dipole transitions are parity-forbidden. Conversely, if Eu3+ ions are situated at the inversion symmetry sites, the magnetic dipole transitions prevail due to the suppression of the electric dipole pathway.38 To quantitatively evaluate the local site symmetry of Eu3+, the asymmetry ratio, defined as IED/IMD, was determined.39 Values of this ratio are provided in Fig. 7(b), and it ranges between 1.18 and 1.74, indicating a significant degree of local asymmetry around the Eu3+ ion, corroborating its incorporation into non-centrosymmetric positions within the host structure.40 Notably, the emission intensity increases steadily with higher Eu3+ doping levels, reaching a maximum at x = 2 mol%. Above this concentration, the intensity diminishes due to the concentration quenching. This is due to the enhanced non-radiative energy transfer among neighboring Eu3+ ions, which leads to diminished radiative recombination efficiency.

To better understand the nature of the interaction mechanism at higher Eu3+ concentrations, the critical distance (Rc) is computed. It represents the average distance at which non-radiative energy transfer becomes significant, and the following relation is employed to compute Rc41

 
image file: d5ra05778g-t3.tif(3)

In this equation, V and N represent the volume of the unit cell and the number of available cationic sites per unit cell that can be substituted by Eu3+ ions. Furthermore, xc denotes the optimum dopant concentration. The Rc value was found to be approximately 23.96 Å, and according to theoretical models, when Rc > 5 Å, the dominant quenching mechanism is attributed to multipole–multipole interactions rather than short-range exchange interaction.42 Thus, the concentration quenching observed is mainly attributed to long-range multipolar interactions between Eu3+ ions.

From Dexter's theory, the emission intensity (I) of each Eu3+ ion in a luminescent material can be described as a function dependent on dopant concentration (x) using the following equation43,44

 
image file: d5ra05778g-t4.tif(4)

In this expression, β and K are the numerical constants influenced by the host β-BaB2O4 as well as the excitation parameters. The parameter Q characterizes the nature of the non-radiative interaction mechanism between neighbouring Eu3+ ions. The value of Q may be 6, 8, or 10, corresponding respectively to dipole–dipole, dipole–quadrupole, and quadrupole–quadrupole interactions.45 Under the assumption that βxQ/3 ≫ 1, we get

 
image file: d5ra05778g-t5.tif(5)

To determine the precise value of the interaction parameter Q, a plot of image file: d5ra05778g-t6.tif versus loge(x) is presented in Fig. 8, the linear graph has a slope of −2.08, which is associated with a Q value of 6.24. This result is closely aligned with Q = 6, and hence the concentration quenching of Eu3+ luminescence primarily arises from dipole–dipole energy transfer interactions in the investigated host material.


image file: d5ra05778g-f8.tif
Fig. 8 Dexter plot for the prepared phosphors with variation of PL intensity for 615 nm as the inset.

Furthermore, the decay profiles of the Eu3+ doped phosphors were well fitted using a mono-exponential function of the form46

I(t) = I0 + A exp(−t/τ)
where I0 is the initial luminescence intensity, τ is the fluorescence decay lifetime, and A represents the constant. The fitted decay curves are presented in Fig. 9(a)–(e). The variation of lifetime with Eu3+ concentration follows the same trend as the emission intensity. Specifically, the lifetime increases with Eu3+ content and reaches a maximum value of 1.988 ms at 2 mol%, beyond which a progressive decrease is observed. This reduction at higher concentrations is attributed to the enhanced probability of non-radiative energy transfer among closely spaced Eu3+ ions, leading to concentration quenching.47


image file: d5ra05778g-f9.tif
Fig. 9 (a)–(e) Decay curves corresponding to the doped phosphors.

3.3 Analysis of colorimetric properties

The emission colour of the phosphors was assessed through colorimetric analysis, employing the Commission Internationale de l'Éclairage (CIE) chromaticity coordinate system to quantitatively assess the perceived colour of the emitted light. The CIE 1931 chromaticity coordinates (xp, yp) for all samples were derived from their emission spectra, using an excitation wavelength of 395 nm. The coordinates are summarized in Table 1, and they were found to exhibit minimal variation, indicating consistency in emission colour across different compositions. The CIE coordinates of the prepared phosphors are illustrated in the CIE diagram shown in Fig. 10.48
Table 1 CIE 1931 coordinates with CCT and C P for the prepared phosphors
Eu3+ (mol%) xp yp x0 y0 CCT (K) Colour purity (%)
1.0 0.6376 0.3620 0.3320 0.1858 1948 100
1.5 0.6369 0.3627 0.3320 0.1858 1933 100
2.0 0.6391 0.3606 0.3320 0.1858 1979 100
2.5 0.6352 0.3644 0.3320 0.1858 1900 100
3.0 0.6290 0.3706 0.3320 0.1858 1796 100



image file: d5ra05778g-f10.tif
Fig. 10 Chromaticity coordinates of the synthesized phosphors in the CIE 1931 colour coordinate diagram.

Subsequently, the correlated colour temperature (CCT) was calculated to quantitatively describe the perceived warmth or coolness of the emitted light. Warm white, typically preferred for residential and ambient lighting light is generally characterized by CCT values of below 3200 K, whereas CCT values exceeding 4000 K correspond to cooler light, more suitable for commercial or industrial environments.49 To estimate the CCT, McCamy's empirical formula was employed50

 
CCT = −449n3 + 3525n2 − 6823.3n + 5520.33 (6)
Here, image file: d5ra05778g-t7.tif with (x0, y0) = (0.3320, 0.1858), representing the chromaticity epicenter of white light convergence. Using the CIE coordinates obtained under 395 nm excitation, the optimal phosphor sample was found to have a CCT value of 1979 K, placing it firmly within the warm light region. The observed variation in CCT values arises from the fact that CCT is calculated based on the chromaticity coordinates (x, y) in the CIE diagram. These coordinates depend on the relative spectral power distribution, that is, the relative intensities of the individual emission peaks. Consequently, even if the overall PL spectrum shape remains largely unchanged, small variations in the relative intensities of peaks can shift the chromaticity coordinates, resulting in changes in the calculated CCT values. The variations in CCT as a function of doping concentration are systematically compiled in Table 1.

Finally, the CIE coordinates were used to compute the colour purity (C P) of the synthesized phosphor materials based on the following standard formula51

 
image file: d5ra05778g-t8.tif(7)
Here (xd, yd) indicates the CIE coordinates for the dominant wavelength on the chromaticity diagram. Using ColorCalculator v7.77 analysis software, and PL data acquired under 395 nm excitation in the 550–750 nm emission range, the CP of all samples was calculated to be nearly 100%. This exceptionally high purity results from the fact that both the sample coordinates and the dominant wavelength lie nearly coincident at the periphery of the CIE chromaticity diagram. This positioning confirms that the emitted light corresponds to pure reddish-orange emission, with minimal deviation from the dominant wavelength. For clarity, the colorimetric properties are also compared with previously reported phosphors in Table 2, thereby emphasizing the consistency of the observed trends with earlier studies. The evaluation reveals that the β-BaB2O4:Eu3+ phosphor exhibits notably higher CP relative to its counterparts. This enhanced chromatic performance underscores the material's effectiveness as a high-quality orange-red emitter. As a result, the synthesized phosphor exhibits significant potential as a viable candidate for its potential applicability in warm colour emitting devices.

Table 2 Comparison of colorimetric parameters of β-BaB2O4:xEu3+ (x = 2 mol%) phosphor with various previously reported Eu3+ doped phosphorsa
Phosphor xc (xp, yp) CP (%) CCT (K) Ref.
a xc = optimized Eu3+ concentration.
Li2SiO3:Eu3+ 4 mol% (0.5800, 0.3375) 75 1240 52
Ba3Lu4O9:Eu3+ 5 mol% (0.6535, 0.3461) 95 2710 53
SrZrO3:Eu3+ 5 mol% (0.6262, 0.3333) 92 4324 54
Ba2LaTaO6:Eu3+ 35 mol% (0.5872, 0.4064) 98.16 1723 55
Sr3LiSbO6:Eu3+ 2 mol% (0.6166, 0.3802) 99.2 1818 56
Ba5P6O20:Eu3+ 12 mol% (0.6471, 0.3494) 91.4 4336 57
SrLaGaO4:Eu3+ 30 mol% (0.647, 0.353) 99.56 4431 58
LiYO2:Eu3+ 20 mol% (0.662, 0.335) 98.9 4354 59
Na2ZrO3:Eu3+ 2 mol% (0.65, 0.35) 91.57 2524 60
KCaF3:Eu3+ 6 mol% (0.5736, 0.4224) 95.56 1565 61
β-BaB2O4:Eu3+ 2 mol% (0.6391, 0.3606) 100 1979 This work


3.4 Analysis of DRS spectra

The combination of diffuse reflectance spectrum and absorbance spectrum for the β-BaB2O4:xEu3+ (x = 2 mol%) phosphor is presented in Fig. 11. Within the wavelength range of 250 to 350 nm, the spectra show prominent absorption, primarily resulting from CTB transitions between O2− and Eu3+ ions, features consistent with those observed in their excitation spectra. Additionally, two weaker absorption peaks are identified near 390 nm and 470 nm, due to the 4f–4f transitions of RE ions.62
image file: d5ra05778g-f11.tif
Fig. 11 Absorbance and reflectance spectra in the UV-Vis-NIR range for the optimized phosphor.

Next, the optical band gap (Eg) was calculated by transforming the DRS measurements into the Kubelka–Munk function F(R), as described by the equation below63

 
image file: d5ra05778g-t9.tif(8)
where α, R, and S denote the absorption coefficient, reflectance, and the scattering coefficient, respectively, with S considered to be a constant across the measured wavelength range.64 To further evaluate Eg, the Tauc relation was applied, which links α to the photon energy () as follows65
 
αhν = C(Eg)n (9)

In this equation, C denotes a constant, and the numerical exponent n is equal to 2 for indirect allowed transitions and 1/2 for direct allowed transitions. Given the direct proportionality between α and F(R), the Tauc equation can be rearranged as

 
[F(R)]1/n = B(Eg) (10)

with B being the constant of proportionality.

To estimate the band gap value of optimized phosphors, a graph of [F(R)]1/n and [F(R)]nagainst was plotted. The Eg was then obtained by extrapolating the linear segment of the graph to the photon energy axis. The Tauc's plots corresponding to the optimized sample are shown in Fig. 12(a) and (b). Here, n = 1/2 shows the better linear fit, and hence it can be concluded that the β-BaB2O4:xEu3+ (x = 2 mol%) phosphor shows a direct band gap, and the Eg value was calculated to be 4.30 eV.


image file: d5ra05778g-f12.tif
Fig. 12 Tauc plot for the optimized phosphor with (a) n = 2, (b) n = 1/2.

3.5 Analysis of TGA curve

TGA studies were performed to assess the thermal stability of the β-BaB2O4:xEu3+ (x = 2 mol%) phosphor at elevated temperatures. The obtained TGA curves for both the undoped and Eu3+ doped samples, presented in Fig. 13, exhibit an initial weight increase followed by a gradual decrease with rising temperature. The slight initial mass gain is attributed to the adsorption of atmospheric gases on the sample surface, while the subsequent weight loss primarily results from the desorption and evaporation of physically adsorbed moisture and residual volatile species.66 Importantly, no significant mass loss or decomposition event was observed throughout the heating process, indicating the strong thermal robustness of the host lattice. The optimized β-BaB2O4:2 mol% Eu3+ phosphor retained its structural integrity up to 500 °C, exhibiting only a minor mass loss of about 2.16 wt%, compared to 4.06 wt% for the pure host. This demonstrates that Eu3+ incorporation not only enhances the luminescent properties but also slightly improves the thermal stability of the material, confirming its suitability for high-temperature optoelectronic and photonic applications.
image file: d5ra05778g-f13.tif
Fig. 13 TGA curve for β-BaB2O4:xEu3+ (x = 2 mol%) phosphor.

3.6 Analysis of JO parameters

To assess the radiative parameters of RE ions in the host matrix, it is important to examine the JO parameters, namely Ωk (k = 2, 4, 6).67,68 These parameters also provide valuable information regarding the local symmetry, bonding characteristics, and covalency around the RE ion within the host lattice. Among them, Ω2 is particularly sensitive to changes in the local symmetry and ligand field, making it a reliable indicator of covalency and the structural configuration around the dopant ion. Meanwhile, Ω4 and Ω6 are primarily linked to the macroscopic properties of the host, including rigidity, viscosity, and dielectric behaviour.69 JO parameters are usually obtained from absorption spectra; however, for Eu3+ doped phosphors, the 5D07F1 transition remains largely unaffected by local surroundings and is therefore used as a reliable reference to calculate electric dipole transition probabilities for other transitions. These include transitions 5D07FJ (J′ = 2, 4, 6), which are influenced by the local symmetry and are used to extract the JO parameters for the prepared phosphors.70 In our work, the Ω6 parameter could not be calculated because the 5D07F6 emission transition was too weak to be observed. As is commonly reported for Eu3+ systems, Ω6 was therefore considered to be zero in the determination of the luminescence radiative properties.71

First, we define the spontaneous emission probability of a magnetic dipole72

 
image file: d5ra05778g-t10.tif(11)

Similarly, for an electric dipole

 
image file: d5ra05778g-t11.tif(12)
Here, n represents the refractive index of the host material, while Smd is equal to 7.83 × 10−42 esu2 cm2, corresponding to the magnetic dipole line strength for Eu3+ ions, which remains unaffected by the surrounding host matrix.73 The parameter λmd and λed denotes the wavelengths, e is the charge of an electron, h denotes Planck's constant, and (2J′ + 1) refers to the degeneracy of the excited state. The squared reduced matrix elements ‖〈5D0|U(J)|7F2〉‖2 = 0.0023 and ‖〈5D0|U(J)|7F4〉‖2 = 0.0032, respectively.74

First, we take the intensity ratio of emission bands originating from the magnetic dipole (Imd) and electric dipole (Ied) transitions can be represented as follows75

image file: d5ra05778g-t12.tif

Substituting for image file: d5ra05778g-t13.tif and image file: d5ra05778g-t14.tif from eqn (11) and (12) will give us

image file: d5ra05778g-t15.tif

On simplifying, we get

 
image file: d5ra05778g-t16.tif(13)

This ratio can also be expressed as

 
image file: d5ra05778g-t17.tif(14)

On comparing eqn (13) and (14), we get

image file: d5ra05778g-t18.tif
Here, the term on the light hand side denotes that they are under the PL emission peaks, and hence, we get
 
image file: d5ra05778g-t19.tif(15)
 
image file: d5ra05778g-t20.tif(16)

Table 3 presents the results of calculations and the ratio Ω2/Ω4 for the prepared phosphors. For all the prepared phosphors, JO parameters follow Ω2 > Ω4 trends. This condition reflects the enhanced covalency in the Eu–O bonds, which is most prominent at all the doping levels. The elevated Ω2 parameter in Table 3 emphasizes the hypersensitive nature of the 5D07F2 transition and implies that Eu3+ ions occupy sites with a highly polarizable and asymmetric chemical environment.

Table 3 Summary of JO intensity parameters calculation for the prepared phosphors
Concentration of Eu3+ (mol%) Ω2 (10−20 cm2) Ω4 (10−20 cm2) Trend Ω2[thin space (1/6-em)]:[thin space (1/6-em)]Ω4
1 2.58 0.76 Ω2 > Ω4 3.40
1.5 2.33 0.74 Ω2 > Ω4 3.14
2 2.32 0.74 Ω2 > Ω4 3.14
2.5 2.06 0.57 Ω2 > Ω4 3.59
3 1.82 0.39 Ω2 > Ω4 4.70


Additionally, the observed trend of Ω2 > Ω4 further supports the presence of strong metal–ligand covalency and a non-centrosymmetric environment surrounding the Eu3+ ions.76 Additionally, we have presented the comparison of JO parameters and the trend for Ω2[thin space (1/6-em)]:[thin space (1/6-em)]Ω4 in Table 4.

Table 4 Comparison of JO parameters of the optimized phosphor with the previously reported works
Phosphor Ω2 (10−20 cm2) Ω4 (10−20 cm2) Ω4 (10−20 cm2) Trend

image file: d5ra05778g-t21.tif

Ref.
KBaScSi3O9:Eu3+ 1.25 0.38 0 Ω2 > Ω4 > Ω6 3.29 77
CaMoO4:Eu3+ 11.50 2.65 0 Ω2 > Ω4 > Ω6 4.34 78
Ca2MgSi2O7:Eu3+ 2.80 2.77 0 Ω2 > Ω4 > Ω6 1.01 79
CaZrO3:Eu3+ 4.36 0.89 0 Ω2 > Ω4 > Ω6 4.90 80
Ba2GdSbO6:Eu3+ 1.52 0.53 0 Ω2 > Ω4 > Ω6 2.87 81
Y2Si2O7:Eu3+ 5.91 0.98 0 Ω2 > Ω4 > Ω6 6.03 82
Na2ZrO3:Eu3+ 5.12 1.59 0 Ω2 > Ω4 > Ω6 3.25 83
CaGeO3:Eu3+ 4.59 1.70 0 Ω2 > Ω4 > Ω6 2.70 84
Y7O6F9:Eu3+ 5.70 1.80 0 Ω2 > Ω4 > Ω6 3.17 85
β-BaB2O4:Eu3+ 2.32 0.74 0 Ω2 > Ω4 > Ω6 3.14 This work


3.7 Analysis of radiative parameters

Ω2 and Ω4 can be used to estimate key radiative properties of Eu3+ doped β-BaB2O4 phosphors, including total radiative transition probability (AT), mean lifetime (τmean), effective bandwidth (Δλeff), branching ratio (βJ′–J) and stimulated emission cross-section (σJ′–J), and for the corresponding emission transitions.

For the ith transition, the radiative transition probability86

image file: d5ra05778g-t22.tif

Therefore, the total radiative transition probability is given by

 
image file: d5ra05778g-t23.tif(17)

Using the AT, we can determine the mean lifetime and branching ratio for a given transition using equations87

 
image file: d5ra05778g-t24.tif(18)
 
image file: d5ra05778g-t25.tif(19)

Finally, the cross-section of stimulated emission is calculated using88

 
image file: d5ra05778g-t26.tif(20)
where, Δλeff, c, image file: d5ra05778g-t27.tif, and λJ′–J, represents the FWHM of the emission peak, speed of light, radiative transition probability, and the peak emission wavelength, respectively. Table 5 represents the radiative parameters corresponding to the prepared phosphors. For efficient laser performance, it is preferable for the emission peak to exhibit a narrow FWHM, as this enhances σJ′–J. A high σJ′–J value associated with a strong emission transition is particularly advantageous for applications requiring low lasing thresholds and high optical gain, making the material suitable for laser-based technologies.89

Table 5 Radiative parameters of the prepared phosphors
Eu3+ concentration Transition

image file: d5ra05778g-t28.tif

image file: d5ra05778g-t29.tif

AT (s−1) τmean (s) β Δλeff (nm) σJ′–J (10−23 cm2)
1 mol% 5D07F2 48.40 88.09 0.0114 0.55 14.80 3.25
5D07F4 3.82 0.04 11.69 5.51
5D07F1 35.87 0.41 11.61 2.24
1.5 mol% 5D07F2 43.76 83.37 0.0120 0.52 15.95 2.73
5D07F4 3.74 0.04 11.61 5.44
5D07F1 35.87 0.43 10.64 2.90
2 mol% 5D07F2 43.51 83.10 0.0120 0.52 15.15 2.85
5D07F4 3.72 0.04 11.48 5.47
5D07F1 35.87 0.43 11.25 2.74
2.5 mol% 5D07F2 38.69 77.45 0.0129 0.50 14.57 2.64
5D07F4 2.89 0.04 9.99 4.88
5D07F1 35.87 0.46 11.73 2.63
3 mol% 5D07F2 34.10 71.92 0.0139 0.47 14.83 2.29
5D07F4 1.95 0.03 9.91 3.32
5D07F1 35.87 0.50 11.37 2.71


In our case, we observe that the branching ratio associated with the 5D07F2 transition is significantly greater than that of the 5D07F1 and 5D07F4 transitions, indicating enhanced colour purity in the studied phosphor material. Furthermore, the elevated values of σJ′–J highlight the strong radiative characteristics of the phosphor, making it a promising candidate for laser applications.

4 Conclusions

Eu3+ doped β-BaB2O4 phosphors were successfully synthesized via the solid-state reaction route, and their structural, optical, and thermal properties were systematically investigated. XRD analysis confirmed the formation of a single-phase β-BaB2O4 structure without any secondary peaks or lattice distortions, indicating the effective incorporation of Eu3+ ions into the host lattice without altering its crystal symmetry. Under 395 nm excitation, the photoluminescence spectra revealed dominant emission bands centered at 581, 593, 615, 653, and 702 nm, corresponding to the 5D07F0, 5D07F1, 5D07F2, 5D07F3, and 5D07F4 transitions, respectively, exhibiting excellent colour purity (≈100%). The concentration quenching behaviour was attributed to non-radiative energy transfer via dipole–dipole interactions among adjacent Eu3+ ions. The material exhibited a direct band gap of 4.30 eV and exceptional thermal stability, retaining structural integrity with only ∼2.18% mass loss at 500 °C. Judd–Ofelt analysis indicated Ω2 > Ω4 for all Eu3+ concentrations, implying strong covalent bonding between the Eu3+ ions and the host lattice. Furthermore, the asymmetry ratio greater than unity confirmed that Eu3+ ions occupy low-symmetry sites within the β-BaB2O4 matrix. Among the observed transitions, the 5D07F2 transition exhibited the highest branching ratio and radiative probability, signifying its dominance in the red emission region. Collectively, these findings demonstrate that β-BaB2O4:Eu3+ phosphors possess superior structural stability, efficient red emission, and favorable electronic characteristics, making them promising candidates for advanced solid-state lighting and laser applications.

Conflicts of interest

There are no conflicts of interest to declare.

Data availability

The data sets generated and/or analyzed in this study are available from the corresponding author on reasonable request.

Acknowledgements

The authors acknowledge the financial support from the Manipal Academy of Higher Education, Manipal, India. The authors express their gratitude to Princess Nourah Bint Abdulrahman University Researchers Supporting Project number (PNURSP2025R2), Princess Nourah Bint Abdulrahman University, Riyadh, Saudi Arabia.

References

  1. Y. Tao, C. D. Rahn, L. A. Archer and F. You, Sci. Adv., 2021, 7, eabi7633 CrossRef CAS PubMed.
  2. R. Yadav, M. Singh, D. Shekhawat, S.-Y. Lee and S.-J. Park, Compos. Appl. Sci. Manuf., 2023, 175, 107775 CrossRef CAS.
  3. F. Khan, N. Hossain, J. J. Mim, S. M. Rahman, Md. J. Iqbal, M. Billah and M. A. Chowdhury, J. Eng. Res., 2025, 13, 1001–1023 CrossRef.
  4. Q. Shi, J. Zhou, S. Ullah, X. Yang, K. Tokarska, B. Trzebicka, H. Q. Ta and M. H. Rümmeli, Energy Storage Mater., 2021, 34, 735–754 CrossRef.
  5. J. Zhang, D. Shao, L. Jiang, G. Zhang, H. Wu, R. Day and W. Jiang, Renew. Sustain. Energy Rev., 2022, 159, 112207 CrossRef CAS.
  6. R. Aksakal, C. Mertens, M. Soete, N. Badi and F. Du Prez, Adv. Sci., 2021, 8, 2004038 CrossRef CAS PubMed.
  7. V. Erduran, M. Bekmezci, R. Bayat and F. Sen, in Functionalized Nanomaterial-Based Electrochemical Sensors, Elsevier, 2022, pp. 97–111 Search PubMed.
  8. A. Dwivedi, A. Roy and S. B. Rai, RSC Adv., 2023, 13, 16260–16271 RSC.
  9. R. Mahajan and R. Prakash, J. Mater. Sci.: Mater. Electron., 2022, 33, 25491–25517 CrossRef CAS.
  10. D. K. Patel and V. K. Sharma, Trans. Indian Inst. Met., 2025, 78, 165 CrossRef.
  11. R. Kiran, N. Kamath, M. I. Sayyed, A. H. Almuqrin and S. D. Kamath, RSC Adv., 2025, 15, 20040–20060 RSC.
  12. M. M. Lanje, M. M. Yawalkar, J. S. Dahegaonkar and S. J. Dhoble, J. Phys.: Conf. Ser., 2021, 1913, 012031 CrossRef CAS.
  13. P. K. Tawalare, Luminescence, 2022, 37, 1226–1245 CrossRef CAS PubMed.
  14. A. N. Yerpude and S. J. Dhoble, in Phosphor Handbook, Elsevier, 2023, pp. 155–176 Search PubMed.
  15. K. R, P. A, S. M. M. Kennedy, M. I. Sayyed, T. A. Hanafy, V. Mishra and S. D. Kamath, J. Mol. Struct., 2025, 1322, 140381 CrossRef CAS.
  16. R. Kiran, H. M. Pratheeksha, V. Saraswathi A, A. Princy, S. M. M. Kennedy, A. S. Altowyan, M. I. Sayyed and S. D. Kamath, J. Solid State Chem., 2024, 337, 124792 CrossRef CAS.
  17. L. H. A. R. Ferreira, G. Dantelle, A. Ibanez and L. J. Q. Maia, Phys. B, 2022, 644, 414193 CrossRef CAS.
  18. W. G. Zou, M. K. Lü, F. Gu, S. Wang, Z. Xiu and G. Zhou, Mater. Sci. Eng. B, 2006, 127, 134–137 CrossRef CAS.
  19. J. Lakde, C. M. Mehare, K. K. Pandey, N. S. Dhoble and S. J. Dhoble, J. Phys.: Conf. Ser., 2021, 1913, 012029 CrossRef CAS.
  20. R. S. Yadav and S. B. Rai, J. Phys. Chem. Solids, 2017, 110, 211–217 CrossRef CAS.
  21. S. Yadav, D. Kumar, R. S. Yadav, S. B. Rai and A. K. Singh, Ceram. Int., 2022, 48, 30754–30766 CrossRef CAS.
  22. Monika, R. S. Yadav, A. Bahadur and S. B. Rai, RSC Adv., 2023, 13, 20164–20178 RSC.
  23. E. Rai, R. S. Yadav, D. Kumar, A. K. Singh, V. J. Fulari and S. B. Rai, RSC Adv., 2023, 13, 4182–4194 RSC.
  24. J. Liu, X.-D. Wang, Z.-C. Wu and S.-P. Kuang, Spectrochim. Acta Mol. Biomol. Spectrosc., 2011, 79, 1520–1523 CrossRef CAS PubMed.
  25. Z. Li, H. Ma, N. Li, Y. Du and Q. Shao, J. Alloys Compd., 2018, 747, 340–347 CrossRef CAS.
  26. R. S. Yadav and S. B. Rai, J. Alloys Compd., 2017, 700, 228–237 CrossRef CAS.
  27. E. Rai, R. S. Yadav, D. Kumar, A. K. Singh, V. J. Fulari and S. B. Rai, J. Lumin., 2022, 241, 118519 CrossRef CAS.
  28. P. Du, X. Huang and J. S. Yu, Inorg. Chem. Front., 2017, 4, 1987–1995 RSC.
  29. X. Tian, L. Guo, J. Wen, L. Zhu, C. Ji, Z. Huang, H. Qiu, F. Luo, X. Liu, J. Li, C. Li, Y. Peng, J. Cao, Z. He and H. Zhong, J. Alloys Compd., 2023, 959, 170574 CrossRef CAS.
  30. A. Khorsand Zak, W. H. A. Majid, M. E. Abrishami and R. Yousefi, Solid State Sci., 2011, 13, 251–256 CrossRef CAS.
  31. P. Rohilla and A. S. Rao, Mater. Res. Bull., 2022, 150, 111753 CrossRef CAS.
  32. K. Ganesh Kumar, P. Balaji Bhargav, K. Aravinth, N. Ahmed and C. Balaji, Ceram. Int., 2022, 48, 36038–36045 CrossRef CAS.
  33. E. Rai, R. S. Yadav, D. Kumar, A. K. Singh, V. J. Fulari and S. B. Rai, J. Lumin., 2022, 241, 118519 CrossRef CAS.
  34. Q. Zhao, J. Chen, X. Jing, T. Lang, M. Cai, L. Peng, Q. Qiang, W. Chen, E. F. Polisadova and B. Liu, J. Lumin., 2024, 276, 120878 CrossRef CAS.
  35. W. T. Carnall, P. R. Fields and K. Rajnak, J. Chem. Phys., 1968, 49, 4450–4455 CrossRef CAS.
  36. C. Hernández-Fuentes, R. Ruiz-Guerrero, A. Morales-Ramírez, P. Molina-Maldonado and D. Medina-Velazquez, Crystals, 2020, 10, 674 CrossRef.
  37. A. Hooda, S. P. Khatkar, A. Khatkar, R. K. Malik, S. Devi, J. Dalal and V. B. Taxak, J. Mater. Sci.: Mater. Electron., 2019, 30, 8751–8762 CrossRef CAS.
  38. N. Degda, N. Patel, V. Verma, K. V. R. Murthy, N. Chauhan, M. Singhal and M. Srinivas, Opt. Mater., 2023, 142, 114019 CrossRef CAS.
  39. V. Ghenea, I. Culeac and A. Buzdugan, J. Eng. Sci., 2024, 31, 28–38 CrossRef.
  40. P. J. Chaware, Y. D. Choudhari, D. M. Borikar and K. G. Rewatkar, Opt. Mater., 2022, 133, 112945 CrossRef CAS.
  41. Y. Du, Y. Li, Y. Zhao, X. Zhang, C. Li, R. Yang, Z. Zou, J. Lian, J. Duan, H. Lin and R. Yu, J. Lumin., 2023, 260, 119893 CrossRef CAS.
  42. Q. Lv, R. Deng, J. Guo, Z. Zhou, Z. Ma, X. Hu, W. Shi, B. Deng, Y. Yu and R. Yu, J. Lumin., 2022, 252, 119320 CrossRef CAS.
  43. D. L. Dexter, J. Chem. Phys., 1953, 21, 836–850 CrossRef CAS.
  44. S. Kaur, A. S. Rao and M. Jayasimhadri, Ceram. Int., 2017, 43, 7401–7407 CrossRef CAS.
  45. T. A. Safeera and E. I. Anila, J. Lumin., 2019, 205, 277–281 CrossRef CAS.
  46. P. Du, J. Tang, W. Li, L. Luo and M. Runowski, Mater. Today Chem., 2022, 26, 101013 CrossRef CAS.
  47. H. Li, J. Zhu, Z. Fang, X. Xiang, J. Jiao, H. Zhang, W. Hu and J. Zhu, Mater. Today Chem., 2023, 30, 101558 CrossRef CAS.
  48. Y. Ma, S. Tang, C. Ji, D. Wu, S. Li, J. Xu, T. Zeng, Z. Huang, H. He and Y. Peng, J. Lumin., 2022, 242, 118530 CrossRef CAS.
  49. B. Das, S. Bardhan, T. Maity and S. Mazumdar, Results Opt., 2020, 1, 100013 CrossRef.
  50. C. S. McCamy, Color Res. Appl., 1992, 17, 142–144 CrossRef.
  51. H. Gao, J. Zhao, Y. Zhang, X. Zhang, H. Bu, Z. Zhao, X. Song, Z. Yang and J. Sun, J. Appl. Phys., 2021, 129, 143102 CrossRef CAS.
  52. P. Barik, A. Verma, R. Kumar, V. Kumar and I. P. Sahu, Appl. Phys. A, 2023, 129, 677 CrossRef CAS.
  53. A. Khatkar, D. Kumar, R. Kumar and S. Lata, J. Fluoresc., 2025 DOI:10.1007/s10895-025-04338-3.
  54. K. C. Sushma, R. B. Basavaraj, D. P. Aarti, M. B. M. Reddy, G. Nagaraju, M. S. Rudresha, H. M. S. Kumar and K. N. Venkatachalaiah, J. Mol. Struct., 2023, 1283, 135192 CrossRef CAS.
  55. F. Li, R. Cui, G. Yuan, X. Zhang, M. Zhang and C. Deng, J. Rare Earths, 2023, 41, 1678–1688 CrossRef CAS.
  56. C. Kumari, J. Manam and S. K. Sharma, Mater. Sci. Semicond. Process., 2023, 158, 107385 CrossRef CAS.
  57. Z. Guo, H. Jiang, H. Li, H. Zhang, C. Liu, R. Zhao, Z. Yang, H. Tang, J. Li, J. Zhang and J. Zhu, Appl. Mater. Today, 2024, 37, 102095 CrossRef.
  58. Y. Qiu, R. Cui, J. Zhang and C. Deng, J. Solid State Chem., 2023, 327, 124265 CrossRef CAS.
  59. L. Zhang, Y. Xu, X. Wu, S. Yin and H. You, Mater. Adv., 2022, 3, 2591–2597 RSC.
  60. P. Khajuria, V. D. Sharma, I. Kumar, A. Khajuria, R. Prakash and R. J. Choudhary, J. Alloys Compd., 2025, 1025, 180268 CrossRef CAS.
  61. R. Kameshwaran, A. Raja, R. R. Kumar, D. J. Daniel, D. O. Annalakshmi, K. Aravinth, P. B. Bhargav and P. Ramasamy, Appl. Radiat. Isot., 2023, 191, 110520 CrossRef CAS PubMed.
  62. S. Verma, I. Ayoub, S. Som, V. Sharma, G. Kumar, H. C. Swart and V. Kumar, Opt. Mater., 2023, 136, 113416 CrossRef CAS.
  63. A. A. Christy, O. M. Kvalheim and R. A. Velapoldi, Vib. Spectrosc., 1995, 9, 19–27 CrossRef CAS.
  64. L. F. Gate, Appl. Opt., 1974, 13, 236 CrossRef CAS PubMed.
  65. Anu and A. S. Rao, Opt. Mater., 2023, 145, 114476 CrossRef CAS.
  66. N. S. Bajaj and S. K. Omanwar, J. Lumin., 2014, 148, 169–173 CrossRef CAS.
  67. B. R. Judd, Phys. Rev., 1962, 127, 750–761 CrossRef CAS.
  68. G. S. Ofelt, J. Chem. Phys., 1962, 37, 511–520 CrossRef CAS.
  69. B. Verma, R. N. Baghel, D. P. Bisen, N. Brahme and V. Jena, Opt. Mater., 2022, 123, 111787 CrossRef CAS.
  70. A. Ćirić, S. Stojadinović, M. Sekulić and M. D. Dramićanin, J. Lumin., 2019, 205, 351–356 CrossRef.
  71. Q. Chen, B. Miao, P. S. Kumar and S. Xu, Opt. Mater., 2021, 116, 111093 CrossRef CAS.
  72. S. R. Yashodha, N. Dhananjaya, S. R. Manohara and H. S. Yogananda, J. Mater. Sci.: Mater. Electron., 2021, 32, 11511–11523 CrossRef CAS.
  73. W. F. Krupke, Phys. Rev., 1966, 145, 325–337 CrossRef CAS.
  74. W. T. Carnall, P. R. Fields and K. Rajnak, J. Chem. Phys., 1968, 49, 4450–4455 CrossRef CAS.
  75. R. Venkatesh, N. Dhananjaya, M. K. Sateesh, J. P. Shabaaz Begum, S. R. Yashodha, H. Nagabhushana and C. Shivakumara, J. Alloys Compd., 2018, 732, 725–739 CrossRef CAS.
  76. S. Tanabe, T. Ohyagi, N. Soga and T. Hanada, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 46, 3305–3310 CrossRef CAS PubMed.
  77. R. Nagaraj, A. Raja and S. Ranjith, J. Alloys Compd., 2020, 827, 154289 CrossRef CAS.
  78. H. N. Van, M. T. Yen Thanh, V.-H. Pham, P. Van Huan, V. T. Ngoc Minh, P. A. Tuan and H. T. Duy, Opt. Mater., 2022, 132, 112831 CrossRef CAS.
  79. P. J. Chaware, Y. D. Choudhari, D. M. Borikar and K. G. Rewatkar, Opt. Mater., 2022, 133, 112945 CrossRef CAS.
  80. S. Jang, J. Lee, S. W. Wi, H. Lim, Y. J. Jeong, J.-S. Chung, W. K. Kang and Y. S. Lee, J. Lumin., 2021, 240, 118433 CrossRef CAS.
  81. C. Kumari, J. Manam and S. K. Sharma, J. Lumin., 2023, 263, 119983 CrossRef CAS.
  82. P. Kumar, D. Singh, S. Kadyan, H. Kumar and R. Kumar, Ceram. Int., 2024, 50, 34596–34608 CrossRef CAS.
  83. P. Khajuria, V. D. Sharma, I. Kumar, A. Khajuria, R. Prakash and R. J. Choudhary, J. Alloys Compd., 2025, 1025, 180268 CrossRef CAS.
  84. W. Hu, Y. Jin, Y. Chen, Y. Wei, C. Wang, Z. Zhang, F. Meng and H. Ren, J. Alloys Compd., 2024, 1008, 176702 CrossRef CAS.
  85. N. Rakov, F. Matias and G. S. Maciel, Phys. B, 2023, 652, 414625 CrossRef CAS.
  86. H. N. Van, M. T. Yen Thanh, V.-H. Pham, P. Van Huan, V. T. Ngoc Minh, P. A. Tuan and H. T. Duy, Opt. Mater., 2022, 132, 112831 CrossRef CAS.
  87. S. S. Babu, P. Babu, C. K. Jayasankar, W. Sievers, Th. Tröster and G. Wortmann, J. Lumin., 2007, 126, 109–120 CrossRef CAS.
  88. S. B. Mallur, T. C. Khoo, S. Rijal, O. R. Huff and P. K. Babu, Mater. Chem. Phys., 2021, 258, 123886 CrossRef CAS.
  89. T. Krishnapriya, A. Jose, T. A. Jose, E. Sreeja, N. V. Unnikrishnan and P. R. Biju, J. Mater. Sci.: Mater. Electron., 2020, 31, 22452–22466 CrossRef CAS.

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.