DOI:
10.1039/D5RA04226G
(Paper)
RSC Adv., 2025,
15, 29238-29253
Luminescence and energy transfer processes in Gd0.99Er0.01Al0.995Cr0.05O3
Received
14th June 2025
, Accepted 5th August 2025
First published on 19th August 2025
Abstract
Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples were synthesized using a solid-state reaction method. Structural analysis revealed that the samples crystallized in an orthorhombic structure phase with a Pbnm space group. The average crystallite sizes were around 283 nm and 574 nm for Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3, respectively. Derivative absorption spectrum fitting (DASF) and first-derivative reflectance (dR /dλ) methods confirmed that the samples possess a direct wide band gap, with energies of 5.93 eV and 5.90 eV, respectively. The photoluminescence (PL) spectrum of Gd0.99Er0.01AlO3 under λex = 377 nm excitation exhibits a green emission and intense sharp red lines at 680 nm, 697 nm, 705 nm, 717 nm and 758 nm. The green emission corresponds to the transitions 2H11/2 → 4I15/2 and 4S3/2 → 4I15/2 of Er3+ ions, while the sharp red lines are attributed to transitions between intrinsic defect centers related to the GdAlO3 host coupled to B3g (4) and B1g (7) vibrational modes. Efficient energy transfer via resonant phonon-assisted and cross-relaxation processes from Er3+ and intrinsic defect centers to Cr3+ is responsible for the decrease in green and red emission line intensities in Gd0.99Er0.01Al0.995Cr0.05O3. The energy transfer from Er3+ and intrinsic defect centers indicates that red emission lines at 697 nm and 726 nm in Gd0.99Er0.01Al0.995Cr0.05O3 mainly originate from the 2T1 (2G) → 4A2 (4F) and 2Eg (2G) → 4A2g (4F) transitions of Cr3+ ions.
1. Introduction
Perovskite compounds serve as excellent host materials for various optical applications due to their chemical and thermal stability.1,2 They follow the generic formula ABO3. Rare-earth orthoaluminates (REAlO3), such as gadolinium aluminate (GdAlO3), possess significant optical, thermal, and mechanical properties, making them appropriate as solid-state laser hosts.3 GdAlO3 is also known for its relatively high dielectric constant, making it valuable for electronic applications, and is being developed as a potential material for neutron absorption and control rod applications.4 In cubic perovskites, the tolerance factor is tobs=1,5 whereas for GdAlO3 with the Pbnm space group, tobs = 0.986,6 indicating a slight distortion from the cubic structure. GdAlO3, formed with Gd3+ ions having a relatively large ion radius (180.4 pm), closely approaches the ideal cubic perovskite crystal cell (Pm
m). GdAlO3 with the Pbnm space group demonstrates a high accommodation capacity within the perovskite structure and assists in the modulation of its electronic and spectroscopic properties through the substitution of Gd3+ with various rare-earth activators (e.g., Eu3+, Er3+, Tb3+, Ce3+, Yb3+, Dy3+) and Al3+ with a transition metal activator ion such as Cr3+ and Mn4+. In this structure, the Gd3+ ion occupies a non-centrosymmetric site, leading to mixing of the 4fn states with the first excited configuration 4fn−15d. This mixing is caused by the odd terms in the crystal field, and is responsible for the strength of induced electric dipole transitions. GdAlO3 with the Pbnm space group is appropriate for generating intense 4f electric dipole transitions, thereby enabling efficient luminescence. In recent years, trivalent rare-earth ions (RE3+) and transition-metal fluorescence in diverse host matrices have attracted significant attention due to their applications in persistent luminescent materials, photo-functional materials and luminescence thermometry.7–10 Doping GdAlO3 with transition metals and rare-earth ions is therefore of great interest for the development of advanced optical materials. Cr3+, in particular, is a transition-metal ion that acts as both a trapping and recombination center and has been widely studied in persistent luminescence research. Its unique properties can enhance the performance of imaging techniques, providing valuable insights into biological systems for in vivo bioimaging.11,12 However, the concentration of Cr3+ must be carefully optimized: a low concentration results in weak luminescence, whereas a high concentration leads to quenching, thereby reducing both the intensity and afterglow duration. A recent study by Jinan Xu et al.13 demonstrated that La0.9898Er0.01Sm0.0002Al0.995Cr0.005O3 exhibits long-term persistent luminescence at 1553 nm due to the Er3+: (4I13/2 → 4I15/2) transition, as well as at 734 nm, due to the Cr3+: (2E (2G) → 4A2 (4F)) transition. In ZnGa2O4
:
Cr3+14 the persistent luminescence intensity increases with Cr3+ concentration up to 0.4–0.6%, after which concentration quenching reduces both intensity and lifetime. Moreover, in recent studies, the emission intensity of Cr3+ reaches its optimum at 0.5 mol% Cr3+ concentration and can be further enhanced by Li+ ion in Cr3+/Li+ co-doping ZnGa2O4 phosphor 15. The lifetimes of the 4T2 (4F) and 2E states of Cr3+ decrease with increasing concentrations of Cr3+ and Cr3+/Li+ ions.15 Similarly, a recent study reported by Ekta Rai et al.16 demonstrated that in Cr3+ and Eu3+ co-doped LaVO4, the emission intensity is optimal at 0.5 mol% Cr3+ and 1 mol% Eu3+ concentration. The emission intensity at 614 nm, corresponding to the 5D0 → 7F2 transition in the Eu3+ doped LaVO4 phosphor, reduces when Cr3+ ion is co-doped due to energy transfer between Cr3+ and Eu3+.16 This energy transfer was confirmed by the decrease of the lifetime of the 5D0 level of Eu3+ ions in Eu3+, Cr3+ co-doped LaVO4 phosphor.16 Understanding the energy levels of dopant ions, traps states, in GdAlO3 host and the energy transfer process between them is crucial for evaluating the suitability of material for optical applications such as LEDs, plant growth lighting, and in vivo optical imaging. The experimental origin of luminescence in GdAlO3 is studied by K Dhahri et al.17 The energy levels of Cr3+ combined with various trivalent lanthanides in GdAlO3 have been studied by Hongde Luo and Pieter Dorenbos 18. However, to the best of our knowledge, Er3+, Cr3+ Co-doped GdAlO3 has not yet been explored. Taking this into account, the present work reports for the first time the synthesis and investigation of a Er3+, Cr3+ co-doped GdAlO3. This study aims to elucidate the energy transfer process occurring between Er3+, Cr3+, and traps states (intrinsic defects). Furthermore, based on both experimental results and theoretical optical considerations, we propose a detailed mechanism for the energy transfer involving Er3+, Cr3+, and the trap states.
2. Experimental procedures (synthesis and characterization)
The GdAlO3, Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples were prepared using a conventional solid-state reaction method. Gd2O3 (99%), Al2O3 (99%), Cr2O3 (99%), and Er2O3 (99%) were used as starting raw materials in stoichiometric amounts. The precursor materials were ground into fine powders using an agate mortar. The powders were initially annealed at 700 °C and then reground, pestled, and gradually heated to 1200 °C in an alumina crucible, where they were sintered for four hours. Finally, the powders were pressed into pellets with an 8 mm diameter. Several techniques were employed to characterize the physical and structural properties of the compounds. The phase compositions of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 were identified by X-ray diffraction (XRD) measurements using a Siemens D5000 X-ray powder diffractometer utilizing CuKα radiation (λ = 1.5406 Å) over a range of 20°–100°, with a step size of 0.02°. The powder morphology and chemical homogeneity were studied by scanning electron microscopy (SEM) using a TESCAN VEGA3 SBH instrument equipped with an energy dispersive microscopy (EDS) detector. Raman spectra were recorded in the range of 50–1100 cm−1 using a Horiba LabRam HR Evolution micro-Raman confocal system, with wavelength laser excitations at λ = 532 nm, 633 nm, and 785 nm. Absorption and reflectance spectra were recorded using a (SHIMADZU, UV-3101PC) UV–vis–IR spectrophotometer. Photoluminescence emission (PL) and excitation (PLE) measurements were recorded using a Horiba–Jobin–Yvon Fluorolog 322 spectrometer in time-resolved mode, using a pulsed lamp with a 0.05 ms post-flash delay.
3. Results and discussions
3.1. Structural analysis
3.1.1. Crystal structure and X-ray diffraction patterns. The XRD pattern analysis of the GdAlO3, Gd0.99Er0.01AlO3, and Gd0.99Er0.01Al0.995Cr0.05O3 samples was performed at room temperature and is shown in Fig. 1(a)–(c). The XRD data were refined using the Rietveld method in the Full Prof software suite.19 The diffraction peaks align closely with the crystal planes of the orthorhombic GdAlO3 structure all diffraction peaks are indexed according to the PDF card no. 46-0395.20 Refinement results indicate that all samples crystallize in the orthorhombic GdAlO3 structure phase with the Pbnm space group.20 The estimated Rietveld refinement parameters, including the goodness of fit (χ2), reliability factors (R-profile factor, R-Bragg factor, and R-crystallographic factor), lattice parameters, cell volumes (V), and interatomic distances, are listed in Table 1. The crystal structure of Gd0.99Er0.01Al0.995Cr0.05O3 and Gd0.99Er0.01AlO3 compounds using VESTA software is presented in Fig. 2. The average crystallite size (Dsc) was estimated from the line broadening of the peak with the highest intensity associated with the plane (112), using the Debye–Scherrer formula.21 |
 | (1) |
where K = 0.9 for spherical shape, λ is the wavelength of X-ray used, β is the full width at half maximum (FWHM) of the diffraction peak, and λ is the Bragg angle for the most intense peak. Furthermore, XRD peak broadening also has a contribution from the self-induced strain (ε) developed in crystallites during the growth that is effective in the nanoparticles.22 We additionally used the Williamson–Hall equation:23 |
 | (2) |
to determine the crystallite size and strain, taking into account the contribution of crystallites and strain to peak broadening. Where K is a constant (K = 0.9 for spherical shape), λ is the wavelength of the used X-ray, β is the full width at half maximum (FWHM) of the diffraction peak, and ε is the effective strain and θ is the Bragg angle for the most intense peak. Eqn (2) represents the Uniform Deformation Model (UDM), which assumes uniform strain in all crystallographic directions. The term (β
cos
θ) is plotted with respect to (4
sin
θ) in Fig. 3 for the preferred orientation peaks (hkl) of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples showing that with the y-intercept and slope of the fitted line determining the crystallite size and related strain, respectively. The lattice strain observed is attributed to defects concentrated along the amorphous grain boundaries. These defects create a stress field within the grain boundary region, thereby inducing strain in the system.24 Table 1 displays the crystallite size estimated from Debye Scherrer's formula and W–H plot as well as the related strain. The crystallite size increases considerably with Cr3+ doping. Cr3+ incorporation can increase both crystallite and particle size in certain oxide materials. When small amounts of Cr3+ ions substitute the cation Al3+ in the Gd0.99AlEr0.01O3 host lattice, they can induce lattice strain, modify the crystal growth process, and reduce the number of nucleation sites, leading to larger crystallites and particles. This effect is noticeable at low doping levels, as seen in Cr3+-doped gadolinium aluminum garnet and doped Mn3O4 systems.25–28 This fact explains the considerable increase of both crystallite and particle size of Gd0.99AlEr0.01O3 by Cr3+ doping with low concentrations. However, as the doping concentration increases further, excessive lattice distortion can inhibit growth, resulting in smaller crystallites and particles, a trend seen in several oxide systems.25,26,28
 |
| Fig. 1 Rietveld refinement of X-ray diffraction pattern of (a) GdAlO3 (b) Gd0.99Er0.01AlO3 and (c) Gd0.99Er0.01Al0.995Cr0.05O3. | |
Table 1 Refined crystallographic parameters, average particle size and average strain value of GdAlO3, Gd0.99AlEr0.01O3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples
Compounds |
GdAlO3 |
Gd0.99AlEr0.01O3 |
Gd0.99Er0.01Al0.995Cr0.05O3 |
a (Å) |
5.253(2) |
5.253(8) |
5.253(6) |
b (Å) |
5.302(5) |
5.303(7) |
5.302(5) |
c (Å) |
7.447(2) |
7.448(6) |
7.447(8) |
v (Å3) |
207.442(4) |
207.552(1) |
207.474(9) |
d(Gd–Gd) |
— |
3.736(7) |
3.804(8) |
d(Gd–Al) |
— |
3.072(3) |
3.266(8) |
χ2 |
1.853 |
1.389 |
1.46 |
RP (%) |
12.3 |
11.7 |
14.8 |
Rwp (%) |
11.5 |
10.4 |
12.0 |
Re (%) |
8.43 |
8.81 |
9.93 |
Dsc (nm) |
— |
108.5(6) |
111.0(2) |
DW–H (nm) |
— |
283.1(4) |
574.3(8) |
DSEM (nm) |
— |
300 |
639 |
ε |
— |
0.00076(1) |
0.00086(6) |
 |
| Fig. 2 The Crystal structure using VESTA software for (a) Gd0.99Er0.01AlO3 and (b) Gd0.99Er0.01Al0.995Cr0.05O3 compounds. | |
 |
| Fig. 3 W–H plots of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 compounds. | |
3.1.2. SEM and EDS analysis. The morphological characterization of Gd0.99Er0.01Al0.995Cr0.05O3 and Gd0.99Er0.01AlO3 compounds was carried out using scanning electron microscopy (SEM), as illustrated in Fig. 4(a) and (b), respectively. The SEM images show that the particles are approximately spherical. Due to the high surface energy of the nanoparticles, the synthesized samples exhibit noticeable aggregation at the annealing temperature.29 The grain size distribution, shown in the inset of Fig. 4(a) and (b), was analyzed using Image J software, and the resulting histograms were fitted to a Lorentzian function. The average grain size distribution revealed peaks around 300 nm for Gd0.99Er0.01AlO3 and 639 nm for Gd0.99Er0.01Al0.995Cr0.05O3 as presented in Fig. 4(a) and (b), and summarized in Table 1. Furthermore, energy dispersive spectra (EDS) were recorded for both samples, as illustrated in Fig. 4(c) and (d). The EDS spectra confirm the existence of the expected constituent elements: Gd, Er, Al, Cr, and O. These results further confirm the compositional purity of the synthesized compound.
 |
| Fig. 4 (a and b) SEM micrographs, with the inset showing the size distribution histogram for Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 compounds, (c and d) spectra of chemical analysis for Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 compounds. | |
3.1.3. Raman spectra analysis of Gd0.99AlEr0.0O3 and Gd0.99Er0.01Al0.995Cr0.05O3 compounds. Raman spectroscopy is a powerful technique that extracts information on the development of the desired phase, detecting impurities, and identifying structural defects by examining Raman active phonon modes. The Raman spectra of GdAlO3 in the orthorhombic (Pbnm) perovskite structure have been studied both theoretically and experimentally by Anastasia Chopelas.30 For GdAlO3 with the orthorhombic (Pbnm) structure, group theory predicts the following optical modes in the Brillouin zone center.30 |
Γ′ = 7Ag® + 5B1g® + 7B2g® + 5B3g® + 8Au(1) + 7B1u® + 9B2u(IR) + 9B3u(IR)
| (3) |
where R and IR denote, respectively, their Raman and infrared spectral activity. Fig. 5 and 6 display the Raman spectra of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 respectively, recorded using excitation wavelengths of 532 nm, 633 nm, and 785 nm. By comparing the spectra obtained with various excitation wavelengths, it is possible to distinguish between Raman scattering and luminescence based on their distinct natures: bands with fixed locations are true Raman bands, whereas bands that shift in position are associated with luminescence.31 By extending the collection range to 1100 cm−1, our measurement revealed several intense and clearly non-vibrational extra bands above 579 cm−1. These bands were attributed to fluorescence as they resemble the characteristic f–f transitions of trivalent lanthanide ions. The appearance of resonance Raman, resonance fluorescence and relaxed fluorescence can be attributed to the excitation energies of the wavelength's excitation 532 nm, 633 nm, and 785 nm which are resonant with the transitions 4I15/2 → 4S3/2, 4I15/2 → 4F9/2, 4I15/2 → 4I9/2 of Er3+, respectively.32 Raman spectra of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 collected with 532 nm laser excitation are shown in Fig. 5 (a) and 6(a). They present prominent peaks at 330 cm−1, 355 cm−1, 403 cm−1, 468 cm−1, 507 cm−1, 545 cm−1, 579 cm−1 assigned respectively to Resonance Raman and resonance fluorescence associated to the vibrations modes B3g(3), Ag(5), B3g (4), B2g (4), B1g(6), B3g(5), B1g(7).30 The peaks at 281 cm−1, 625 cm−1, 677 cm−1, 734 cm−1, 807 cm−1, 892 cm−1, 947 cm−1, 1026 cm−1 are assigned to the vibration modes 2B2g(1), 2Ag(4), 2B3g(3), 2Ag(5), 2B3g(4), 2Ag(6), 2B2g (4), 2B1g(6)30 related to relaxed fluorescence. The bands of Raman spectra of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 recorded with 785 nm laser excitation in the vicinity of 1000 cm−1 are clearly due to relaxed fluorescence. Some vibration modes such Ag(3) at 235 cm−1 (ref. 30) occur only under 785 nm excitation, the vibrations mode B1g(3) at 220 cm−1 appear only under 633 nm excitation, whereas the vibrations mode at Ag(4) at 314 cm−1 appear only under 633 nm excitation. The bands at 403 cm−1 and 579 cm−1 due to B3g(4) and B1g(7) vibration modes are true Raman bands since they have fixed locations and do not depend on the wavelength excitation. They appear under all the wavelengths excitations of 532 nm, 633 nm, and 785 nm.
 |
| Fig. 5 Raman spectra of Gd0.99AlEr0.01O3 under different excitation wavelengths: (a) 532 nm, (b) 633 nm, and (c) 785 nm. | |
 |
| Fig. 6 Raman spectra of Gd0.99Er0.01Al0.995Cr0.05O3 under different excitation wavelengths: (a) 532 nm, (b) 633 nm, and (c) 785 nm. | |
3.2. Optical properties
3.2.1. Absorbance, reflectance spectra and band gap determination. The wavelength–dependent absorbance spectra of Gd0.99Er0.0AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3, recorded in the 200–2500 nm range, are illustrated in Fig. 7. Both spectra exhibit an intense absorption band at 246 nm and two weaker bands at 975 nm and 1535 nm. With Cr3+ co-doping, an additional broad band appears at 565 nm. The absorption band at 246 nm is assigned to 8S7/2 → 6D7/2 transition of Gd3+ ions,33 whereas the peaks at 975 nm and 1535 nm correspond to the 4I15/2 →4I11/2 and 4I15/2 →4I13/2, transitions of Er3+ ions,32 respectively. The additional broad band in the absorbance spectrum of Gd0.99Er0.01Al0.995Cr0.05O3 at 565 nm is assigned to the 4A2 (4F) → 4T2 (4F) transition of Cr3+ ions. The band gap energy of the samples needs to be correctly determined in order to predict semiconductor optical properties. The derivation of absorption spectrum fitting (DASF), a precise method developed by Souri and Tahan,34 was used to ascertain the band gap's value and nature. The main advantage of this method is that it does not require any presumption of the nature of the optical transition and linear extrapolation. The absorption coefficient can be expressed as a function of the optical gap and the energy of photons as follows:35,36
 |
| Fig. 7 Absorbance spectra of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples at room temperature in the wavelength range 200–2500 nm. | |
By rewriting the eqn (1) as a function of the wavelength (λ): eqn (4) become:
|
 | (5) |
where
α(
λ), is the absorption coefficient defined by the Beer–Lambert's law as:
|
 | (6) |
where
d and
A as film thickness and film absorbance. Using
(5) and
(6); the absorbance can be rewritten as:
|
 | (7) |
where:
According to eqn (7) we have:
|
 | (8) |
Eqn (8) can be reformulated as follows:
|
 | (9) |
By differentiating
with respect to
we obtained the following equation:
|
 | (10) |
The band gap's value can be determined using the following expression for absorbance.37
|
 | (11) |
A(λ), λ, and λg are, respectively, absorbance, the incident wavelength, and the wavelength corresponding to the band gap energy. m and D are constants. The plot left side of eqn (11)
for Gd0.99AlEr0.01O3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples is shown in Fig. 8. The peak maxima can be used to determine the band gap energy, as seen in Fig. 8, at
the peak maximum discontinuity occurs. The optical band gap is computed as
Using the obtained λg, the resulting values are (5.93 ± 0.01) eV and (5.90 ± 0.01) eV for Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 respectively. Marotti et al. showed that for direct band gap semiconductors, dR/dλ peaks close to Eg, whereas for indirect band gap compounds, it approaches zero. Fig. 9 confirms the direct character of the optical band gap Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3, indicating that dR/dλ reaches a maximum at about 5.93 eV and 5.90 eV for Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples. The optical band gap values determined via the derivation of absorption spectrum fitting (DASF) and the first derivative of reflectance, dR/dλ, are the same. This supports the correctness of the band gap energy values found. The decrease in the band gap of Gd0.99AlEr0.01O3 when doped with low concentration of Cr3+ ions introduces localized energy levels within the band gap which act as both deep electron and deep hole traps within the Gd0.99AlEr0.01O3 band gap. These traps are localized and do not merge with the conduction or valence bands at low doping levels. At low concentrations, these levels do not significantly alter the overall electronic structure or the positions of the conduction and valence bands.18 The increased hybridization Cr–O between the Cr-3d and O-2p orbitals due Cr3+ incorporation modifies slightly the top of the valence band which may shift upward (due to Cr 3d–O 2p interactions) or the bottom of the conduction band may shift downward.38 So, the band gap remains nearly unchanged.
 |
| Fig. 8 The variation of for Gd0.99AlEr0.01O3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples. | |
 |
| Fig. 9 The room temperature reflectance spectrum R(λ) of (a) Gd0.99AlEr0.01O3 and (b) Gd0.99Er0.01Al0.995Cr0.05O3. The inset shows the evolution of dR/dλ with λ. | |
3.2.2. Photoluminescence (PL) and photoluminescence excitation (PLE) spectra. The PL spectra of the GdAlO3 and Gd0.99Er0.01AlO3 samples, excited at 377 nm with a 0.05 ms delay after flash, in the wavelength range of 400–800 nm, are presented in Fig. 10. At room temperature, the PL spectrum of undoped GdAlO3 exhibits distinct sharp red emission lines at 680 nm (14
705 cm−1), 697 nm (14
347 cm−1), 705 nm (14
184 cm−1) 717 nm (13
947 cm−1), and 758 nm (13
192 cm−1). Upon Er3+ doping, these red emission lines persist, while additional bands appear in the wavelength ranges of 521–535 nm, 542–562 nm, and 655–657 nm, corresponding to the 2H11/2 → 4I15/2, 4S3/2 → 4I15/2 and 4F9/2 → 4I15/2 transitions of Er3+ ions, respectively.32 These bands are the result of intra-configurational 4f–4f transitions of Er3+ ions that appear within the 4f shell. According to the Laporte selection rule, 4f–4f transitions in rare-earth ions are parity-forbidden. The appearance of sharp spectral characteristics in Er3+ spectra could be explained by a non–central crystalline field's odd-order terms, which can create a coupling between odd and even states. Thus, resulting in mixed states of the 4fn with the first excited 4fn−1 5d configuration, which mitigates Laporte's rule. This fact is responsible for the high intensity of induced electric dipole transitions observed in the PL and PLE spectra.39 The red lines at 680, 697, 717, 705 and 758 nm are superimposed in both GdAlO3 and Gd0.99Er0.01AlO3, indicating that Er3+ doping does not modify the structure of the sharp lines and their energy locations confirms that these sharp red lines originate from the host GdAlO3 matrix. The red lines cannot originate from intraconfigurational transitions of Gd3+ ions since the excitation wavelength is 377 nm and the first excited state is 6P7/2 at 314 nm. In order for 4f–4f transitions of Gd3+ to take place under 377 nm excitation, two photons of 377 nm absorption must happen. Since we used a pulsed lamp with low intensity rather than a high-power laser, this is not possible. Similar sharp red emission lines under 320 nm excitation were also reported in undoped GdAlO3 by Kh. Dhahri et al.17 Based on their findings, the red emission in GdAlO3 is attributed to the presence of oxygen vacancies, singly ionized Vo+. In CaGdAlO4-type layered perovskites,40 deep red luminescence (emission around 711 nm) under 338 nm excitation is attributed to oxygen defects, especially oxygen interstitials. These defects create localized energy states within the bandgap, enabling radiative recombination that results in red light emission when the material is excited by UV or visible light. The origin of red luminescence in undoped GdAlO3 is primarily linked to intrinsic crystal defects, specifically oxygen-related defects, rather than the presence of intentional dopants. In addition, under 532 nm excitation, undoped YAlO3 single crystals exhibit emission bands in the wavelength range 670–800 nm, including peaks at ∼688, 703, 715, 732, and 750 nm,41 similar to those observed in GdAlO3 (ref. 17) and to the red lines found in the present work. The PLE spectrum monitored at the emission wavelength 715 nm in YAlO3 shows the strong excitation bands at ∼320 nm and 315 nm.41 In the distorted perovskite structure of GdAlO3, cation vacancies such as Vo+ and Vo++ are the dominant intrinsic defects to neutralize the minor amount of Cr3+ and Er3+.42 Taking this into account, the photoluminescence (PL) process responsible for the red emission under 377 nm excitation can be described as follows: under excitation at 377 nm, electrons are excited from the valence band and subsequently trapped by intrinsic defect. These trapped electrons then relax and are captured by deep acceptor states associated with intrinsic defects. According to the configuration coordinate model, the resulting red emission peaks can be ascribed to electron transitions between donor and acceptor levels associated with vibrational modes B3g(4) and B1g(7) as shown in Fig. 11. With chrome co-doping, an additional intense red emission line appears at 726 nm, and a weak peak at 693 nm, as shown in the room-temperature PL spectrum of Gd0.99Er0.01Al0.995Cr0.05O3 in Fig. 12. The 726 nm emission corresponds to the 2E (2G) → 4A2(4F) transition of Cr3+ ions.43 The co-doping by chrome induces a significant decrease in the emission intensity of the Er3+ ions, a dramatic decrease in the emission line intensity at 680 nm, 697 nm, 717 nm, 705 nm. A low intensity peak at 693 nm occurs assigned to the transition from the fundamental state 4A2(4F) to the sublevel of 2T1(2G) split by spin–orbit coupling. An intense emission line emerges at 726 nm Fig. 13. Photoluminescence excitation (PLE) spectra of Gd0.99Er0.01Al0.995Cr0.05O3 and Gd0.99Er0.01AlO3 monitored at 542 nm are presented in Fig. 14a and b, respectively. The room-temperature PLE spectrum of Gd0.99Er0.01Al0.995Cr0.05O3 monitored at 542 nm, corresponding to an Er3+ transition, shows typical Er3+ excitation lines peaking at around (357 nm, 366 nm), 377 nm, 406 nm, 443 nm, 450 nm, 487 nm, and 521 nm. These lines are attributed to the following transitions: 4I15/2 → 4G7/2, 4I15/2 → 4G11/2, 4I15/2 → 2H9/2, 4I15/2 → 4F3/2, 4I15/2 → 4F5/2, 4I15/2 → 4F7/2, 4I15/2 → 2H11/2, respectively.32 The intensity I (N, τ) of the emission lines depends on both the population density N of the excited state and the radiative lifetime τ of the emitting level. The strongest excitation peak at 377 nm, corresponding to the transition4I15/2 → 4G11/2, can be justified by the fact that the transition 4I15/2 → 4G11/2 obeys the selection rule for an electronic–dipole transition in the context of Judd–Oeffelt theory.39
 |
| Fig. 10 Room temperature PL emission spectra of the GdAlO3 and Gd0.99Er0.01AlO3 samples collected with excitation at 377 nm with flash-lamp and 0.05 ms delay after flash. | |
 |
| Fig. 11 Configuration coordinate diagram showing the transition responsible for red emission lines which takes place between vibrational levels at (T = 300 K) in Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3. | |
 |
| Fig. 12 Room temperature PL spectrum of Gd0.99Er0.01Al0.995Cr0.05O3 in the wavelength range of 400–740 nm under 377 nm excitation. | |
 |
| Fig. 13 Room temperature PL spectra of Gd0.99Er0.01Al0.995Cr0.05O3 and Gd0.99Er0.01AlO3 samples under 377 nm excitation. | |
 |
| Fig. 14 Room-temperature photoluminescence excitation (PLE) spectra monitored at 542 nm of (a) Gd0.99Er0.01AlO3 and (b) Gd0.99Er0.01Al0.995Cr0.05O3 (4S3/2 →4I15/2: Er3+). | |
The intensity of 4f–4f transitions of rare-earth elements within a host matrix can be described using the standard Judd–Ofelt (J–O) theory. According to this theory, the expressions for the electric dipolar line strength SJJ′ED and electric dipolar oscillator strength fcal(J,J′) of transitions from the state |S, L, J> to the state |S′, L′, J′> are given by:39
|
 | (12) |
|
 | (13) |
where
Ωλ are the Judd–Ofelt intensity parameters. The terms in brackets represent the doubly reduced matrix elements in intermediate coupling.
J is the total angular momentum of the initial state,
h is the Planck constant,
c is the speed of light,
![[small lambda, Greek, macron]](https://www.rsc.org/images/entities/i_char_e0cc.gif)
is the mean wavelength corresponding to the specific absorption band of a transition |
S,
L,
J> to the state|
S′,
L′,
J′> and
n is the refractive index of GdAlO
3. Assuming that the host matrix has minimal influence on these values, we take this value as the value of Er
3+ in aqueous solutions (aq), or Er
3+ in LaF
3 crystal as mentioned in ref.
44. The transitions from the ground state
4I
15/2 to the excited states
2H
11/2 and
4G
11/2 are characterized by large reduced matrix elements of the unit tensor.
32 Thus, they present a high population densities of Er
3+ ions in these excited state
4G
11/2 and
2H
11/2. According to
eqn (5) and
(6), this may result in strong absorption exhibited in the PLE spectrum.
32 The existence of a significant electric dipole transition
4I
15/2 →
4G
11/2 implies that Er
3+ ions occupy non-centrosymmetric sites in the GdAlO
3 lattice. The appearance of the emission line at 542 nm (
4S
3/2 →
4I
15/2) transition under 377 nm excitation can be attributed to three factors: firstly, strong absorption to the
4G
11/2 state since
λex = 377 nm is a resonant excitation, secondly the rate of multiphonon relaxation
4G
11/2 →
2H
9/2 →
4F
3/2 →
4F
5/2 →
4F
7/2 →
2H
11/2 →
4S
3/2 exceeds the probability of radiative decay transitions to the ground state
4I
15/2; finally, the high energy separation between the emitted level
4S
3/2 and
4F
9/2 level. However, the high intensity of the emission line at 542 nm under 521 nm excitation (
4I
15/2 →
2H
11/2) can be attributed to the large reduced matrix elements of the unit tensor of this transition, leading to high cross-absorption to the
2H
11/2 state. Moreover, the lowest energy separation (approximately 1000 cm
−1) between the emitted level
4S
3/2 and
2H
11/2 levels increases the multiphonon relaxation. Hence, the multiphonon relaxation from
2H
11/2 to
4S
3/2 level is efficient, which induces the population of
4S
3/2 level. The increase of the intensity of the 542 nm emission line (
4S
3/2 →
4I
15/2) with decreasing wavelength excitation from 406 nm to 487 nm can be explained by the multi-phonon relaxation between the excited level and the emitting level
4S
3/2, which is governed by the energy-gap law or phonon law.
45 The multi-phonon relaxation rate (
Wnr) increases with decreasing energy separation between the excited levels and the emitting level
4S
3/2. Hence, the emission at 542 nm increases with decreasing wavelength excitation from 406 nm (
4I
15/2 →
2H
9/2) to 487 nm (
4I
15/2 →
4F
7/2). The energy separation between the
4S
3/2 and
4F
9/2 levels is approximately 3100 cm
−1,
32 requiring five phonons (579 cm
−1 each) to bridge the gap. Therefore, the non-radiative relaxation from
4S
3/2 to
4F
9/2 highly inefficient. As a result, the red emission intensity at 655 nm is lower than the green emission intensity at 542 nm under 377 nm excitation in Gd
0.99Er
0.01AlO
3 (
Fig. 10). The Photoluminescence process at 542 nm under 377 nm excitation in Gd
0.99Er
0.01AlO
3 is shown in
Fig. 15.
 |
| Fig. 15 The Photoluminescence process at 542 nm under 377 nm in Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3. | |
The room-temperature PLE spectrum of Gd0.99Er0.01AlO3 monitored at 542 nm, shows the same characteristic excitation lines as those observed in Gd0.99Er0.01Al0.995Cr0.05O3 with two additional bands at 241 nm and 275 nm. These bands are attributed to the 8S7/2 → 6D7/2 and 8S7/2 → 6I7/2 transitions of Gd3+ ions, respectively.46 The disappearance of these emission bands in the Cr-doped sample (Gd0.99Er0.01Al0.995Cr0.05O3), under 542 nm monitoring, suggests the absence of energy transfer from Gd3+ to Er3+ in the presence of Cr3+. The room-temperature PLE spectrum of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 monitored at 697 nm are shown in (Fig. 16). The spectrum of Gd0.99Er0.01AlO3 exhibits two intense peaks at 322 nm and 275 nm, which assigned to the excitation of the electron from the valence band which trapped by the defect within the forbidden bandgap and the 8S7/2 → 6I7/2 transitions of Gd3+ ions, respectively. In contrast, the photoluminescence spectrum of Gd0.99Er0.01Al0.995Cr0.05O3 monitored at 697 nm reveals intense peaks at 329 nm and 276 nm, assigned to the excitation of the electron from the valence band which trapped by the defect and the 8S7/2 → 6I7/2 transitions of Gd3+ ions, respectively. Additionally, four extra bands are observed at 565 nm, 413 nm, 314 nm, and 241 nm. The broad bands around 565 nm and 413 nm are assigned to transitions from the ground state 4A2 (4F) to the excited states 4T2 (4F) and 4T1(4F) of Cr3+.47 The peaks at 314 nm and 241 nm are assigned to the (8S7/2 → 6P7/2:Gd3+) and (8S7/2 → 6D7/2: Gd3+) transitions, respectively.46
 |
| Fig. 16 Room temperature PLE spectra in Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 monitored at 697 nm. | |
3.2.3. Crystal field analysis and energy level schemes of Cr3+ ions in Gd0.99Er0.01Al0.995Cr0.05O3 nanoparticles. The energy levels of Cr3+ ions in Gd0.99Er0.01Al0.995Cr0.05O3 nanoparticles were calculated using the total Hamiltonian48 |
H = H0 + Hee(B, C) + HTrees(α) + HCF(Dq) + HSO(ξ)
| (14) |
Eqn (14) describes the entire Hamiltonian H, where H0 is the configuration Hamiltonian term, and Hee(B, C), representing the electron–electron repulsion Hamiltonian. This term gives rise to the eight Russell–Saunders terms 2S+1L, including 4F, 4P, 2G, 2P, 2H, 2F, (2aD) and (2bD) for Cr3+ ions with 3 d3 configuration. HCF(Dq) is the crystal field Hamiltonian, and HSO (ξ) represents the spin–orbit coupling Hamiltonian. Using Racah algebraic techniques, the energy levels of the Russell–Saunders terms for the 3 d3 configuration are expressed in terms of Racah parameters A, B, and C, which depend on the double radial integrals F and G. The relative energies are those measured by optical spectroscopy. The quantity related to the A parameter is eliminated since it is the same for all the Russell–Saunders terms. Cr3+ ions (3 d3) are assumed to substitute Al3+ ions at the octahedral [AlO6] site in an intermediate crystal field (CF) strength. The basic function in the LS coupling scheme are expressed as:49,50
The crystal field energy levels can then be obtained by diagonalizing the entire Hamiltonian H = H0 +Hee (B,C) + HCF (Dq) + HSO (ξ). The crystal field Hamiltonian (HCF) in Wybourne notation is the spin–orbit hamiltonian and the Trees hamiltonian are expressed as follow:
|
 | (16) |
|
 | (17) |
where
ξ3d is the spin–orbit coupling constant,
α is the Trees parameter. The Racah and crystal field parameters
B,
C, and
Dq are determined using the Newton–Raphson method by fitting the experimental energies levels to the theoretical ones
4A
2 (
4F) →
4T
2 (
4F) (565 nm),
4A
2(
4F) →
4T
1(
4F) (413 nm), and
2E(
2G) →
4A
2(
4F) (726 nm). The adjusted spin–orbit coupling
ξ3d and Trees parameter
α are calculated as follows:
|
 | (19) |
|
 | (20) |
B0 = 918 cm−1, C0 = 4133 cm−1,48 ξ0 = 275 cm−1 and α0 = 30 cm−1,48 which refer to the free ion parameters of Cr3+. The matrix elements of the crystal field, spin–orbit, and Trees Hamiltonians in the basics are provided by Y. Y. Yeung and C. Rudowicz.50 The full Hamiltonian matrix H (as defined in eqn (14)) was diagonalized to derive the energy levels as a function of the Racah parameters B and C, the crystal field parameter Dq, and the spin–orbit coupling constant. This diagonalization was performed using unique code developed in our lab with the Maple program. The theoretical computed values are B = 635 cm−1, C = 3008 cm−1 and Dq = 1776 cm−1 with (Dq/B = 2.79). The calculated parameters were used to calculate the energy levels at room temperature, as listed in Table 2. The Tanabe–Sugano diagram for Cr3+ ions in octahedral site symmetry, shown for the ratio C/B = 4.73 in Fig. 17, illustrates the overall behavior of Cr3+ energy levels in terms of Dq/B relative to the local field intensity. The vertical line corresponds to the calculated Dq/B value from our theoretical computation of Cr3+ levels in Gd0.99Er0.01Al0.995Cr0.05O3. It is well known that when Dq/B < 2.3, Cr3+ ions experience a weak crystal field, resulting in broad-band emission. However, when Dq/B > 2.3, the ions exhibit strong and narrow peak emission through the 2E(2G) → 4A2(4F) transitions.51 In our case, the calculated Dq/B = 2.9, confirms that the energy of the 2E state is the lowest excited energy level. These results demonstrate that Cr3+ ions experience a strong crystal field, exhibiting sharp 2E(2G) → 4A2(4F) emission at 726 nm.
Table 2 Experimental and calculated energies levels (cm−1) of Cr3+ ion in octahedral symmetry in Gd0.99Er0.01Al0.995Cr0.05O3 nanoparticles
 |
| Fig. 17 Tanabe–Sugano diagram for Cr3+ ions with C/B = 4.73. The vertical line at Dq/B = 2.79 represents the energy levels identified for Cr3+in Gd0.99Er0.01Al0.995Cr0.05O3. | |
3.2.4. Energy transfer process from Gd3+, Er3+ and oxygen defects to Cr3+ ions in Gd0.99Er0.01Al0.995Cr0.05O3. The lack of Gd3+ transitions in the PLE spectrum of the Gd0.99Er0.01Al0.995Cr0.05O3 sample monitored at 542 nm proves the weak efficiency of energy transfer between Gd3+ and Er3+ ions. This can be attributed, first, to the shorter distance between Gd3+ and Cr3+ ions (dGd–Cr = 3.26683 Å, Table 1) compared to the distance between Gd3+ and Er3+ ions (dGd–Er = 3.80481 Å, Table 1) in Gd0.99Er0.01Al0.995Cr0.05O3. Second, the dominance of the higher energy transfer efficacy from Gd3+ to Cr3+ ions, which is explained by the higher trapping efficiency of the migrating excitation energy levels 6PJ and 6IJ of Gd3+ by Cr3+ activators in Gd0.99Er0.01Al0.995Cr0.05O3. Which is greater than by Er3+.52 This higher trapping efficiency is further explained by the spectral overlaps between the Cr3+ excitation band (4T1 (4P)) and the 6I7/2 level of Gd3+ ions.52 Notably, the peaks at 314 nm and 275 nm in the PLE spectrum of Gd0.99Er0.01Al0.995Cr0.05O3 monitored at 697 nm, present high intensities comparable to the peak at 329 nm. This experimental fact points out that under 314 nm and 275 nm, Gd3+ ions also act as donors of energy to the intrinsic defects via resonant phonon-assisted energy transfer. The decay PL curves at 697 nm and 542 nm under 377 nm excitation are shown in Fig. 18 (a) and (b) for Gd0.99Er0.01AlO3 and in Fig. 18 (c) and (d), for Gd0.99Er0.01Al0.995Cr0.05O3. The decay PL curves at 542 nm under λex = 377 nm in Gd0.99Er0.01AlO3 is fitted to a monoexponential equation I(t) = Aexp(−t/τ) with fluorescence lifetimes
The decay of the emission at 542 nm under 377 nm excitation in Gd0.99Er0.01Al0.995Cr0.05O3 is bi-exponential with lifetimes τ′ = 0.13 ms and
Kr is the emissive rate constant, Knr is the non-radiative rate and KT is the transfer rate constant. The decay curve shows two lifetimes: an unchanged longer time 0.13 ms (no transfer), a shorter time 0.03 ms (efficient transfer) which indicates an efficient energy transfer from a part of Er3+ ions to Cr3+ is taken place. The decay rate of Er3+ emission in Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples as well as the ET efficiency from Er3+ to Cr3+ can be calculated using eqn (21)–(23), |
Kr + Knr + KT = (τEr,Cr)−1
| (22) |
|
 | (23) |
 |
| Fig. 18 Decay PL curves under 377 nm excitation for Gd0.99Er0.01AlO3 at (a) 697 nm and (b) 542 nm, and for Gd0.99Er0.01Al0.995Cr0.05O3 at (c) 697 nm and (d) 542 nm. | |
τEr and τEr,Cr, are the life times of the Er3+
:
4S3/2 level in Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 respectively. The ET efficiency (ηET) from Er3+
:
4S3/2 to the Cr3+: as indicated from the energy level diagram in Fig. 19 can be estimated to be 76%. Which indicate highly efficient energy transfer from Er3+
:
4S3/2 to the Cr3+
:
4T2g(4F). In rare-earth ion systems, energy transfer exchange interactions are generally negligible due to the small spatial extent of 4f orbitals, making multipolar mechanisms dominant. Specifically, in GdAlO3, efficient trapping of excitation energy by Cr3+ and rare earth ions like Er3+ is attributed to multipolar interactions when there is spectral overlap, while exchange interactions play a minor role and are only significant for rare earth ions lacking allowed absorption bands.52 Therefore, for Er3+–Cr3+ pairs in GdAlO3, multipolar interactions are expected to dominate the energy transfer process, provided there is suitable spectral overlap between their energy levels. The non-radiative energy transfer in Gd0.99Er0.01Al0.995Cr0.05O3 is taken place through the following cross-relaxations: (2H11/2 + 4A2(4F) → 4I15/2 + 4T2(4F)), (4S3/2 + 4A2(4F) → 4I15/2 + 4T2(4F)) transitions and via resonant phonon-assisted energy transfer (Fig. 19). This energy transfer process involving Er3+ ions explain the strong luminescence quenching of the emission bands at 542 nm, 521 nm and the appearance of an intense emission line at 726 nm when Cr3+ is incorporated. The decay of the emission at 697 nm under 377 nm excitation in Gd0.99Er0.01AlO3 is bi-exponential with lifetimes τ′ = 2.61 ms. and τ′′ = 1.18 ms. Biexponential decay indicates that there are two different mechanisms that affect the decay dynamics and energy transfer may be one of these mechanisms. The decay curve shows two lifetimes: a longer time 2.61 ms (no transfer), a shorter time 1.18 (efficient transfer). This fact indicates that there is energy transfer from intrinsic defects to Er3+ ions in Gd0.99Er0.01AlO3. However, the decay curve at 697 nm under 377 nm excitation in Gd0.99Er0.01Al0.995Cr0.05O3 is described by a double-exponential equation I(t) = I01
exp(−t/τ1) + I02
exp (−t/τ2), where I is the luminescence intensity; I01 = 1844 and I02 = 1258 are constants; t represents time, τ1 = 1.69 ms and τ2 = 10.77 ms are decay times for the respective exponential components. The absence of the bands at 565 nm and 413 nm in the PLE spectrum of Gd0.99Er0.01AlO3 monitored at 697 nm indicates that the 697 nm emission line arise not only from electron transition between intrinsic defects centers, but also from the transition 2T1(2G)→4A2(4F) of Cr3+. This assignment is justified since the experimental value 697 nm (14
347 cm−1) is well reproduced by theoretical value of the transition in Table 2. Moreover, the highest-value lifetime τ2 = 10.77 ms is characteristic of the spin-forbidden transition. τ1 = 1.69 ms is fluorescence lifetime of 697 nm emission coming from the electron transition between two defects centers. The co-doping with chromium leads to a reduction of fluorescence lifetime of 697 nm emission from 2.61 ms to 1.69 ms, indicating that there is an increase in the transfer rate KT constant. A dramatic decrease in the 697 nm emission line intensity and the appearance of an intense emission line at 726 nm support the energy transfer from oxygen-vacancies to Cr3+ ions via resonant phonon-assisted energy transfer from oxygen defect to Cr3+
:
4T2(4F) level and cross-relaxation processes (Fig. 19). The value of the constants I01 and τ1 in the PL decay at 697 nm under 377 nm excitation in Gd0.99Er0.01Al0.995Cr0.05O3 indicate that a very low part I01
exp(−t/τ1) of the emission at 697 nm is originating from electron transition between two intrinsic defects centers coupled to the B3g (4) and B1g(7) vibrational modes. The highest part I02
exp (−t/τ2) of the emission at 697 nm inGd0.99Er0.01Al0.995Cr0.05O3 is originating from the transition 2T1 (2G) →4A2 (4F) of Cr3+. The energy transfer from Er3+, defects centers to Cr3+ and the high multiphonon relaxation from 2T1 (2G) to the 2E (2G) level, induces the intense line at 726 nm through the transition 2E (2G) →4A2 (4F) of Cr3+ (Fig. 19). Based on the experimental results and the theoretical optical considerations, the energy transfer process between Er3+, Cr3+ and oxygen-vacancies under λex = 377 nm is presented in Fig. 19 and (Table 3).
 |
| Fig. 19 Energy level diagram and energy transfer mechanism in Gd0.99Er0.01Al0.995Cr0.05O3. | |
Table 3 Fluorescence lifetime of Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples, monitored at Er3+ emission wavelengths of λem = 542 nm and λem = 697 nm under λex = 377 nm
Samples |
Gd0.99Er0.01AlO3 |
Gd0.99Er0.01Al0.995Cr0.05O3 |
The recorded wavelength |
697 nm |
542 nm |
697 nm |
542 nm |
Double-exponential |
Mono-exponential |
Double-exponential |
Double-exponential |
 |
2.61–1.18 |
0.13 |
— |
— |
 |
— |
— |
1.69–10.77 |
0.13–0.03 |
4. Conclusion
Gd0.99Er0.01AlO3 and Gd0.99Er0.01Al0.995Cr0.05O3 samples were successfully synthesized using the solid-state reaction method. XRD analysis confirmed that both samples crystallize in an orthorhombic structure with a Pbnm space group. Using the Derivation of absorption spectrum fitting (DASF) and the first derivative of reflectance dR/dλ methods, the optical band gaps were determined to be 5.93 eV and 5.90 eV, respectively. Red emission peaks at 680, 697, 705, 717, and 758 nm were observed in both samples and are ascribed transition involving intrinsic defects coupled with the B3g (4) and B1g(7) vibrational modes. Co-doping with Cr3+ induced a significant decrease in Er3+ emission intensity, particularly the intensity of red lines at 680, 697, 705, 717 nm. The weak peak at 693 nm and the intense peak at 726 nm are assigned to the 2T1 (2G) → 4A2 (4F) and 2Eg (2G) → 4A2g (4F) transitions of Cr3+. PL spectra and decay curves at 697 nm and 542 nm under 377 nm excitation confirmed efficient non-radiative energy transfer from Er3+ and intrinsic defects to Cr3+ ions. The energy transfer occurs via resonant phonon-assisted processes from the Er3+ 2H11/2 and 4S3/2 levels to the Cr3+
:
4T2 (4F) level, followed by cross-relaxations (2H11/2+4A2(4F) → 4I15/2+4T2(4F)),(4S3/2+4A2(4F) → 4I15/2+4T2(4F)). Decay curves at 697 nm indicate that the lower part of this emission originates from transition involving intrinsic defects coupled with vibrational modes, while the highest part of the emission at 697 nm in Gd0.99Er0.01Al0.995Cr0.05O3 is due to the Cr3+ 2T1 (2G) →4A2 (4F) transition. Finally, energy transfer from Er3+, defect to Cr3+and the high relaxation from 2T1 (2G) to the 2E (2G) level induce the intense line at 726 nm due to the transition 2E (2G) →4A2 (4F) of Cr3+.
Author contributions
F. Mselmi: Conceptualization, methodology, writing – original draft, formal analysis. Abir Hadded: Writing – original draft, formal analysis, software. Hajer Souissi: Investigation, validation. Souha Kammoun: Investigation, validation, supervision. J. Pina: Software, investigation. B. F. O. Costa: Investigation, validation, supervision.
Conflicts of interest
This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.
Data availability
All the data for the manuscript is available in the manuscript.
Acknowledgements
This work was supported by the Tunisian National Ministry of Higher Education, Scientific Research and Technology.
References
- R. K. Tamrakar and K. Upadhyay, J. Disp. Technol., 2016, 12, 599–604 CAS.
- K. Dhahri, M. Bejar, E. Dhahri, M. J. Soares, M. F. P. Graça and M. A. Sousa, Mater. Lett., 2014, 128, 235–237 CrossRef CAS.
- C. J. Shilpa, A. K. Jayaram, N. Dhananjaya, H. Nagabhushana, S. C. Prashantha, D. V. Sunitha, S. C. Sharma, C. Shivakumara and B. M. Nagabhushana, Spectrochim. Acta, Part A, 2014, 133, 550–558 CrossRef CAS.
- G. R. Remya, S. Solomon, J. K. Thomas and A. John, Mater. Today: Proc., 2015, 2, 1012–1016 Search PubMed.
- S. Sasaki, C. T. Prewitt and R. C. Liebermann, Am. Mineral., 1983, 68, 1189–1198 CAS.
- N. L. Ross, J. Zhao, J. B. Burt and T. D. Chaplin, J. Phys.: Condens. Matter, 2004, 16, 5721 Search PubMed.
- M. Gao, J. Yu, S. Shi, J. Wang and L. Fu, Spectrochim. Acta, Part A, 2025, 326, 125194 CrossRef CAS PubMed.
- K. Li and D. Zhu, Mater. Res. Bull., 2025, 181, 113117 Search PubMed.
- M. Sekulić, V. Đorđević, Z. Ristić, M. Medić and M. D. Dramićanin, Adv. Opt. Mater., 2018, 6(17), 1800552 Search PubMed.
- J. Xu, S. Tanabe, A. D. Sontakke and J. Ueda, Appl. Phys. Lett., 2015, 107, 081903 CrossRef.
- T. Maldiney, A. Bessière, J. Seguin, E. Teston, S. K. Sharma, B. Viana, A. J. Bos, P. Dorenbos, M. Bessodes and D. Gourier, Nat. Mater., 2014, 13, 418–426 CrossRef.
- J. Li, C. Wang, J. Shi, P. Li, Z. Yu and H. Zhang, J. Lumin., 2018, 199, 363–371 CrossRef.
- J. Xu, D. Murata, Y. Katayama, J. Ueda and S. Tanabe, J. Mater. Chem. B, 2017, 5, 6385–6393 RSC.
- Z.-J. Wei, K. Long, C. Yin, X. Yuan, M. Sun, W. Wang and Z. Yuan, J. Mater. Chem. B, 2025, 13, 6508–6518 RSC.
- R. S. Yadav, A. Bahadur and S. B. Rai, Phys. Scr., 2023, 98, 105919 CrossRef.
- E. Rai, R. S. Yadav, D. Kumar, A. K. Singh, V. J. Fulari and S. B. Rai, RSC Adv., 2023, 13, 4182–4194 RSC.
- K. Dhahri, M. Bejar, E. Dhahri, M. J. Soares, M. Sousa and M. A. Valente, J. Alloys Compd., 2015, 640, 501–503 CrossRef.
- H. Luo and P. Dorenbos, J. Mater. Chem. C, 2018, 6, 4977–4984 RSC.
- R. A. Young, The Rietveld Method, International Union of Crystallography, 1993, vol. 5 Search PubMed.
- V. Lojpur, S. Ćulubrk, M. Medić and M. Dramicanin, J. Lumin., 2016, 170, 467–471 CrossRef.
- S. K. Mohamed, M. M. Abd El-Raheem, M. M. Wakkad, A. A. Hakeeam and H. F. Mohamed, Mem. - Mater. Devices Circuits Syst, 2023, 6, 100085 Search PubMed.
- N. S. Gonçalves, J. A. Carvalho, Z. M. Lima and J. M. Sasaki, Mater. Lett., 2012, 72, 36–38 CrossRef.
- A. Hadded, F. Mselmi, S. Kammoun and E. Dhahri, Opt Laser. Technol., 2025, 183, 112244 Search PubMed.
- H. M. Moghaddam and S. Nasirian, Nanosci. Methods, 2012, 1, 201–212 CrossRef.
- A. Bala and S. Rani, Opt. Quantum Electron., 2023, 55, 866 CrossRef.
- A. Nallathambi, A. Prakasam and R. A. Raj, Phys. E, 2020, 116, 113716 CrossRef.
- G. V. Priya, N. Murali, M. K. Raju, B. Krishan, D. Parajuli, P. Choppara, B. C. Sekhar, R. Verma, K. M. Batoo and P. V. L. Narayana, Appl. Phys. A, 2022, 128, 663 CrossRef.
- Y. Pepe, A. Karatay, Y. O. Donar, S. Bilge, E. A. Yildiz, M. Yuksek, A. Sınağ and A. Elmali, Mater. Chem. Phys., 2020, 255, 123596 CrossRef.
- A. N. Mallika, A. R. Reddy and K. V. Reddy, J. Adv. Ceram., 2015, 4, 123–129 CrossRef CAS.
- A. Chopelas, Phys. Chem. Miner., 2011, 38, 709–726 CrossRef CAS.
- Y. U. Jinqiu, C. U. I. Lei, H. E. Huaqiang, Y. A. N. Shihong, H. U. Yunsheng and W. U. Hao, J. Rare Earths, 2014, 32, 1–4 CrossRef.
- X. Chen, E. Ma and G. Liu, J. Phys. Chem. C, 2007, 111, 10404–10411 CrossRef CAS.
- A. Angnanon, S. Nualpralaksana, B. Damdee, N. Wongdamnern, N. Intachai, S. Kothan and J. Kaewkhao, Optik, 2024, 296, 171431 CrossRef CAS.
- D. Souri and Z. E. Tahan, Appl. Phys. B, 2015, 119, 273–279 CrossRef CAS.
- N. F. Mott and E. A. Davis, Electronic processes in non-crystalline materials, Oxford Clarendon, 1979, vol. 2, p. 660 Search PubMed.
- J. Tauc and A. Menth, J. Non-Cryst. Solids, 1972, 8, 569–585 CrossRef.
- I. Rhrissi, O. El Harafi, Y. Arba and R. Moubah, Mater. Sci. Eng., B, 2023, 297, 116784 CrossRef CAS.
- E. Mamani Flores, B. S. Vera Barrios, J. C. Huillca Huillca, J. A. Chacaltana García, C. A. Polo Bravo, H. E. Nina Mendoza, A. B. Quispe Cohaila, F. Gamarra Gómez, R. M. Tamayo Calderón and G. de L. Fora Quispe, Crystals, 2024, 14, 998 CrossRef CAS.
- B. Di. Bartolo, O. Forte, in Advances in Spectroscopy for Lasers and Sensing, Springer Netherlands, Dordrecht, 2006, pp. 403–433 Search PubMed.
- B. Wang, C. Gong, X. Xue, M. Li, Q. Zhu, X. Wang and J.-G. Li, Dalton Trans., 2023, 52, 16780–16790 RSC.
- M. Suganya, K. Ganesan, P. Vijayakumar, S. Jakathamani, A. S. Gill, O. Annalakshmi, S. K. Srivastava, R. M. Sarguna and S. Ganesamoorthy, Opt. Mater., 2020, 107, 110095 CrossRef CAS.
- B. Lou, J. Wen, L. Ning, M. Yin, C.-G. Ma and C.-K. Duan, Phys. Rev. B, 2021, 104, 115101 Search PubMed.
- B. R. Jovanic and J. P. Andreeta, J. Phys.:Condens. Matter, 1998, 10, 271 CrossRef CAS.
- W. T. Carnall, P. R. Fields and K. Rajnak, J. Chem. Phys., 1968, 49, 4424–4442 CrossRef CAS.
- F. Mselmi, I. Elhamdi, M. Bejar and E. Dhahri, Opt. Mater., 2023, 137, 113555 CrossRef.
- G. Wang, H. G. Gallagher, T. P. J. Han and B. Henderson, Radiat. Eff. Defects Solids, 1995, 136, 43–46 CrossRef.
- L. Reddy, Heliyon, 2024, 10(15), e34477 CrossRef.
- I. Elhamdi, F. Mselmi, S. Kammoun, E. Dhahri, A. J. Carvalho, P. Tavares and B. F. O. Costa, Dalton Trans., 2023, 52, 9301–9314 RSC.
- Z.-Y. Yang and Q. Wei, Phys. B, 2005, 370, 137–145 CrossRef.
- Y. Y. Yeung and C. Rudowicz, Comput. Chem., 1992, 16, 207–216 CrossRef.
- K. Sun, X. Yin, Z. Li, H. Lin, R. Hong, D. Zhang, Z. Zhang, G. Zheng and Y. Ding, Opt. Mater. Express, 2022, 12, 2942–2953 CrossRef.
- A. J. De Vries, W. J. J. Smeets and G. Blasse, Mater. Chem. Phys., 1987, 18, 81–92 CrossRef.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.