Next Article in Journal
Investigation on Mechanical and Microstructure Properties of Silt Improved by Titanium Gypsum-Based Stabilizer
Previous Article in Journal
Effect of Laser Micro-Texturing on Laser Joining of Carbon Fiber Reinforced Thermosetting Composites to TC4 Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RE3+-Doped Ca3(Nb,Ga)5O12 and Ca3(Li,Nb,Ga)5O12 Crystals (RE = Sm, Dy, and Pr): A Review of Current Achievements

1
National Institute for Laser, Plasma and Radiation Physics, 077125 Magurele, Romania
2
Doctoral School of Physics, Faculty of Physics, University of Bucharest, 077125 Magurele, Romania
*
Author to whom correspondence should be addressed.
Materials 2023, 16(1), 269; https://doi.org/10.3390/ma16010269
Submission received: 5 December 2022 / Revised: 21 December 2022 / Accepted: 22 December 2022 / Published: 27 December 2022

Abstract

:
Spectroscopic characteristics of RE3+ ions (RE = Sm, Dy, and Pr) doped in partially disordered Ca3Nb1.6875Ga3.1875O12-CNGG and Ca3Li0.275Nb1.775Ga2.95O12-CLNGG crystals are reviewed in detail to assess their prospects as laser crystals with emission in the visible spectral domain. All investigated crystals were grown using the Czochralski crystal growth technique. High-resolution absorption and emission measurements at different temperatures, as well as emission dynamics measurements, were performed on the grown crystals. The spectroscopic and laser emission characteristics of the obtained crystals were determined based on the Judd-Ofelt parameters. The obtained results indicate that CNGG:RE3+ and CLNGG:RE3+ (RE = Sm, Dy, and Pr) crystals can be promising materials for lasers in the visible range.

1. Introduction

A natural limitation of solid-state lasers due to the nature of the quantum processes underlying their operation is the existence of well-specified laser emission wavelengths determined by both the laser active ion and the host material. This situation requires further identification of new laser materials or emission schemes. One of the most viable solutions is to use host crystals with partially disordered structures determined by the mixed occupation of the host cationic sites with different ions. One of the most viable solutions is the use of host crystals with a partially disordered structure caused by the simultaneous occupation of some of the host cationic sites with different ions. Partial compositional disorder preserves the global crystallographic structure and symmetry of the crystal, but influences properties dependent on the local composition, such as the crystal field in the position of the laser active ion. Significant in this respect is the possibility of controlling the host disordering by producing manageable effects on the dopant ion spectra (especially the absorption and emission linewidths) through suitable control of the crystal field distribution in a compositionally disordered crystal. In the case of intrinsically disordered crystals in which certain crystallographic sites are filled with different types of cations with varying valences, the charge compensation induced by doping is achieved by changing the amounts of host cations. Due to the disorder in the close vicinity of the dopant ion, a multicenter structure and an inhomogeneous widening of the absorption and emission lines occur. Controlling the compositional disordering of the host laser material and a proper selection of the materials with an intrinsic disorder, are essential for identifying new laser materials or emission regimes.
Many families of disordered crystals such as aluminates, niobates, borates, double tungstates, etc., are currently known [1,2,3,4,5,6,7], but most of them are of low symmetry. Over recent years, trivalent rare earth ions (RE3+) doped in intrinsically disordered calcium-niobium-gallium-garnet (CNGG) and calcium-lithium-niobium-gallium-garnet (CLNGG) crystals have regained much interest as laser materials. Their main advantages are the low melting temperature (~1480 °C) [8,9] which allows the use of cheaper platinum (Pt) crucibles than iridium (Ir) ones, a moderate thermal conduction of 3–4.7 Wm−1 K−1, and the broad transition lines [9,10]. The stoichiometric composition of CNGG (Ca3Nb1.5Ga3.5O12) is different from the congruent one Ca3Nb1.6875Ga3.1875?0.125O12 which has a deficiency of Ga3+ and an excess of Nb5+ compensated by cationic vacancies (?) [8,9,11,12,13], which can be harmful, especially in the case of high-average power lasers. By adding Li+ cations into the CNGG crystal, the vacancies are eliminated and a new Ca3LiyNb(1.5+y)Ga(3.5–2y)O12 (CLNGG) crystal is obtained. It was found that good quality CLNGG crystals can be obtained for the compositional parameter y comprised between 0.25 and 0.275, and the best results in terms of transparency of the grown crystal, the starting material purity, and the purity of the solidified melt remaining after crystal growth can be attained for y = 0.275 [9,10,12,14,15,16]. Regarding the cationic occupation in the CNGG structure, Ca2+ ions occupy dodecahedral c-sites, Nb5+, Ga3+, and vacancies (?) share the octahedral a-sites, and Ga3+ ions are situated in the tetrahedral d-sites. A small percentage (2% to a maximum of 5%) of the tetrahedral d-sites can be occupied by Nb5+ ions. In the CLNGG structure, Li+ ions are mainly located in the octahedral sites and less likely positioned in the tetrahedral sites [15]. In the CNGG and CLNGG structure, RE3+ ions replace Ca2+ cations in the dodecahedral positions of the garnet lattice, thus imposing the necessity of charge compensation by compositional modification of the host, mainly of the octahedral sublattices, according to each specific doping concentration. This determines the presence of cations of varying valence and/or ionic sizes in the first cation spheres of RE3+ ions, which further leads to the formation of a multicenter structure and an inhomogeneous broadening of the absorption and emission lines of RE3+ ions. The spectroscopic properties of different RE3+ (RE = Pr, Nd, Sm, Tb, Dy, Ho, Er, Tm, Yb) ions doped in CNGG or CLNGG crystals were reported [15,17,18,19,20,21,22,23,24,25]. For example, when doping with Nd3+ ions, five nearly similar nonequivalent centers were highlighted by low-temperature spectroscopic measurements on CNGG:Nd and CLNGG:Nd crystals [15]. Laser emission of Nd3+- or Yb3+-doped CNGG or CLNGG in various operating regimes and pumping conditions were reported [26,27,28,29,30,31,32].
The main criteria for the selection of Pr3+, Sm3+, and Dy3+ as doping ions are the electronic configuration and transition probabilities. Due to the strong absorption and emission transitions of the grouped manifolds, 3P0, 3P1, 3P2, and 1I6, respectively, the Pr3+ ion is a very attractive laser-active ion. The most important laser transitions of Pr3+ ions are in the red, yellow, orange, green, and blue spectral domains and arise from the 3P0 level [33,34,35,36,37]. Sm3+ ions exhibit strong orange-red luminescence assigned to the 4G5/26HJ (J = 5/2, 7/2, 9/2) electronic transitions and a long lifetime (ms) of the emission level, which make Sm3+ ion a perfect choice for phosphor materials. In recent years, due to the increased demand for different lasers and light sources in the visible (VIS) range, the research on Sm3+ ions for the development of yellow, red, or orange lasers under InGaN laser diode pumping has become more and more significant [21,38,39,40,41]. Similarly, Dy3+ ion is also a very good choice for investigating VIS emission. Yellow laser emission [42,43], white-light emitting phosphors [44,45], or blue emission for temperature sensing [46] were recently reported.
Optical properties of CNGG:RE3+ and CLNGG:RE3+ (RE = Sm, Dy, and Pr) crystals are systematically examined in this work to evaluate their prospects as laser crystals in the VIS domain. The oscillator strengths and transition probabilities of the 4f2, 4f5, and 4f9 electronic configurations of the trivalent Pr3+, Sm3+, and Dy3+ ions, respectively, were obtained. High-resolution spectroscopy at low (10 K) and room temperature was employed to determine the optical properties of RE3+ doping ions. The stimulated emission cross-sections were determined using the Füchtbauer-Ladenburg formula. The emission kinetics of the 1D2 (Pr3+), 4G5/2 (Sm3+), and 4F9/2 (Dy3+) excited levels were measured, and the Inokuti-Hirayama model was used to analyze the energy transfer processes and determine the transfer microparameters.

2. Materials and Methods

The polycrystalline host compounds Ca3Nb1.6875Ga3.1875O12 (CNGG) and Ca3Li0.275Nb1.775Ga2.95O12 (CLNGG) were synthesized by the solid-state reaction method. For doping of CNGG and CNLGG hosts with Sm3+, Dy3+, and Pr3+ ions, Sm3Ga5O12 (SmGG), Sm3Li0.275Nb0.275Ga4.45O12 (SmLNGG), Dy3Ga5O12 (DyGG), Dy3Li0.275Nb0.275Ga4.45O12 (DyLNGG), Pr3Ga5O12 (PrGG), and Pr3Li0.275Nb0.275Ga4.45O12 (PrLNGG) polycrystalline compounds, respectively, were also synthesized under the same conditions. The compositions of Li+-free compounds (SmGG, DyGG, PrGG) were selected according to the stoichiometric one, and the compositions of Li+-containing compounds (SmLNGG, DyLNGG, PrLNGG) were selected according to the optimal Li+ content of 0.275 in the CLNGG host. Powders with a purity of at least 4N of CaCO3, Nb2O5, Li2CO3, Ga2O3, Sm2O3, Dy2O3, and Pr6O11, were used as raw materials. In order to remove the absorbed water, CaCO3 and Li2CO3 powders were first heated at 400 °C for 10 h. Then, for each polycrystalline compound, all powders were weighed, mixed, pressed into cylindrical tablets, and heated for 12 h in an air atmosphere at 900 °C to decompose CaCO3 and Li2CO3. Finally, the tablets were further heated for 12 h at 1200 °C to complete the sintering. To achieve the 5 at.% doping level of the CNGG and CLNGG compounds with Sm3+, Dy3+, and Pr3+ ions, respectively, the polycrystalline compounds of the hosts and the corresponding dopants (SmGG, DyGG, PrGG in the case of CNGG, and SmLNGG, DyLNGG, PrLNGG for CLNGG) were mixed in the proper amounts, considering that the RE3+ (RE = Sm, Dy, and Pr) ions substitute Ca2+ in the CNGG and CLNGG hosts.
RE-doped CNGG (CNGG:RE) and RE-doped CLNGG (CLNGG:RE) crystals were grown using the Czochralski crystal growth technique from Pt crucibles with a height of 30 mm and a diameter of 30 mm in air atmosphere. The growth temperature was determined to lie between 1460 °C and 1480 °C. High-quality crystals of CNGG:RE and CLNGG:RE were grown along the <111> crystallographic direction using optimized rotation and pulling rates of 15–20 rpm and 1.3–1.5 mm/h, respectively. After growth, the crystals were thermally treated in an air atmosphere for 20 h at 1300 °C to release the thermal stress and stabilize the oxidation state of RE3+ doped ions. Detailed conditions about the growth parameters were presented in our earlier publication [47]. The chemical composition of the obtained crystals composition was determined by the inductively coupled plasma-atomic emission spectrometry (ICP−AES) method, and a Trace Scan Advantage spectrometer (Thermo Jarrell Ash Corp., Franklin, MA, USA) was used. The structural characterization of the sintered polycrystalline compounds and the grown crystals was done through X-ray powder diffraction (XRPD) using a Bruker AXS D8 ADVANCE X-ray diffractometer (Bruker, Brno, Czech Republic) with Cu Kα radiation (λ = 1.5406 Å). Rietveld analysis was employed to quantify any residual phases in the sintered compounds.
For low or room temperature spectroscopic measurements, a system containing a Jarrell Ash or Horiba Jobin Yvone monochromator (Jarrell Ash Division/Fisher Scientific Company, Waltham, MA, USA and Horiba Jobin Yvon GmbH, Bensheim, Germany, respectively) equipped with S20 or S1 type photomultipliers (Photek Ltd., St Leonards-on-Sea, East Sussex, UK) coupled to a lock-in amplifier coupled to a computer was employed. As excitation sources, a Melles Griot Argon laser (National Laser Comp., Salt Lake City, Utah, USA) or a tunable system containing an Oriel Cornerstone 260 monochromator (Newport Corporation, California, USA coupled with a monochromator illuminator Oriel APEX (Newport Corporation, California, USA) and a Xenon lamp (Sciencetech Inc., Ontario, Canada) were used. A closed cycle He refrigerator ARS-2HW (Advanced Research Systems, Inc., Macungie, PA, USA) was used for low-temperature measurements. To measure the luminescence decay, a Tektronix 2024B oscilloscope (Tektronix, Inc., Beaverton, OR, USA) was used for data recording, and a tunable OPOTEK RADIANT 355 LD laser (Opotek LLC., California, CA, USA) in the range of 410–2500 nm or the third harmonic at 355 nm of an Nd:YAG Quantel laser was utilized as an excitation source.

3. Results

3.1. Crystals Growth and Structural Characterization

XRPD spectra on the SmGG, DyGG, PrGG, SmLNGG, DyLNGG, and PrLNGG sintered compounds revealed a dominant garnet-type cubic phase (space group Ia3d) and some residual phases like Ga2O3, SmNbO4, LiGa5O8, DyNbO4, and PrGaO3. According to Rietveld analysis, the majority garnet phase was found to be in the range of 85–91 wt.%, while the total of minor secondary phases was quantified to be less than 15 wt.% in each sintered compound [47]. The relationships between interplanar spacings and Miller indices, as well as the lattice constants for the garnet phase in each sintered compound, are presented in Reference [47].
Figure 1 shows the CNGG:RE and CLNGG:RE grown crystals. As can be observed, they have very good transparency, are free of macroscopic defects, and are of high optical quality. The grown crystal sizes are approximately 12 and 25–30 mm in diameter and length, respectively. The grown crystals present good mechanical properties, being hard enough for cutting and polishing.
The XRPD patterns on the CNGG:RE and CLNGG:RE grown crystals are shown in Figure 2. The patterns are very well indexed by the garnet-type structure (space group Ia3d) and no residual phases could be found. The lattice parameters were determined and are given in Table 1 together with those of the undoped host. The determined lattice parameters match very well with the ionic radii of Ca2+ (1.12 Å), Pr3+ (1.126 Å), Sm3+ (1.079 Å), and Dy3+ (1.027 Å) ions in 8-fold oxygen coordination, thus proving the insertion of RE3+ ions into the dodecahedral c-sites of the obtained crystals. Moreover, the elemental compositions of the crystals with a measurement error of ± 0.2% are given in Table 1. The effective segregation coefficients (keff) of RE3+ ions in the CNGG host crystal were evaluated to be keff (Sm) = 0.69, keff (Dy) = 0.84, and keff (Pr) = 0.36, being similar to those obtained in the case of the CLNGG host crystal. Thus, the concentration of RE3+ dopant ions in CNGG:RE and CLNGG:RE crystals was determined as being 3.4 at.%, 4.2 at.%, and 1.8 at.% for Sm3+, Dy3+, and Pr3+, respectively. The cation densities of RE3+ ions were calculated to be NA = 4.236 × 1020 ions/cm3 (CNGG:Sm), NA = 5.14 × 1020 ions/cm3 (CNGG:Dy), and NA = 2.209 × 1020 ions/cm3 (CNGG:Pr) in the case of CNGG:RE crystals, and NA = 4.232 × 1020 ions/cm3 (CLNGG:Sm), NA = 5.15 × 1020 ions/cm3 (CLNGG:Dy), and NA = 2.204 × 1020 ions/cm3 (CLNGG:Pr) for CLNGG:RE crystals.

3.2. Spectroscopic Investigations

The room temperature absorption spectra of CNGG:RE and CLNGG:RE crystals (RE3+ = Pr3+, Sm3+, Dy3+) were registered and analyzed within the Judd-Ofelt (JO) theory [48,49] to calculate the oscillator strengths and transition probabilities of the 4f2, 4f5, and 4f9 electronic configurations of the Pr3+, Sm3+, and Dy3+ ions, respectively. The JO theory is described in detail in References [21,23,24], and the most important formulas are given below. The electric dipole line strength (Smeas) of a transition can be determined from the absorption measurements as:
S m e a s J J = 3 c h 2 J + 1 n N A 3 π 3 λ ¯ 9 ( n 2 + 2 ) 2 k λ d λ
where c is the speed of light, h is the Planck constant, n is the bulk index of refraction, NA is the RE3+ ion concentration, λ ¯ is the mean wavelength of the absorption band that corresponds to the J→J’ transition, k λ d λ is the integrated absorption coefficient, and k(λ) is the absorption coefficient which depends on the wavelength. Based on several values of the refractive indices at various wavelengths of undoped CNGG [50,51] and CLNGG [17] crystals and a least-squares fitting program for the Sellmeier equations, dispersion curves of refractive indices were determinate and further used to calculate Smeas for each transition. The theoretical (Stheor) [52] and measured (Smeas) oscillator strengths of the absorption transitions were obtained and used to determine the Ωt (t = 2, 4, 6) intensity parameters. The root-mean-square (rms) deviation   Δ S r m s = ( q p ) 1 ( Δ S ) 2 1 / 2 , where q is the number of the analyzed transitions, p is the number of parameters, and ∆S = Stheor-Smeas, represents the matching error.
Other spectroscopic parameters such as spontaneous emission probabilities (AJJ), branching ratios (β), and radiative lifetimes (τr), were determined based on JO intensity parameters. The total spontaneous electric dipole emission transition probabilities from the excited state J to the lower state J’ are given by the formula:
A J J ' e d = 64 π 4 e 3 3 h 2 J + 1 λ ¯ 3 n ( n 2 + 2 ) 2 9 t = 2 , 4 , 6 Ω t S , L J U t S , L J 2
The radiative lifetime τr for an excited state J and the luminescence branching ratios β(J→J’) for the various emission transitions from this state can be then calculated as:
τ r = 1 A J J ,   and   β J J = A J J J A J J
From room temperature emission spectra originating from 1D2 (Pr3+), 4G5/2 (Sm3+), and 4F9/2 (Dy3+) manifolds, stimulated emission cross-sections em) were calculated by using the Fuchtbauer-Ladenburg (FL) equation [53]:
σ λ = λ 5 A J J ' 8 π n 2 c I λ λ I λ d λ
where A (J→J’) is the spontaneous emission probability from the excited state J to the terminal state J’, I(λ) is the emission intensity at wavelength λ, n is the refraction index, c is the speed of light, and λ I λ d λ is the integrated emission intensity.
The energy transfer (ET) processes induced by the static interactions between the dopant ions strongly influence the excitation flow between the energy levels of the active ion. The interionic process represents the direct transfer of excitation between two ions without the absorption or emission of photons. The ions involved (donors D and acceptors) are connected through the multipolar, exchange, or super-exchange interactions explained by the theory developed by Förster [54] and Dexter [55]. The ET from the donor to the acceptor, in addition to radiative and non-radiative de-excitation, represents a process of de-excitation of the donor. This transfer modifies the excited level lifetime of the donor. Depending on the dopant concentration, the kinetics of the emission level can be described by an exponential or non-exponential function. At very low dopant concentration, the measured lifetime is likely to be the radiative lifetime if there is no non-radiative contribution to the decay curve. For higher dopant concentration, the non-exponential luminescence decays can be evaluated using the formula τav = 0 t I t d t / 0 I t d t and extracting an average lifetime for the emitting level. At high doping concentrations, the type of interaction between RE3+ ions, such as dipole-dipole (DD), dipole-quadrupole (DQ), or quadrupole-quadrupole (QQ) interaction, can be determined from the non-exponential profile of the luminescence decay. In this case, the Inokuti-Hirayama (IH) energy transfer model [56] was employed to analyze the emission decay by using the following equation for the luminescence intensity, Φ(t):
Φ t = A × exp t τ 0 4 π 3 Γ 1 3 s N A R 0 3 t τ 0 3 s
where A is the amplitude, τ0 is the lifetime of isolated RE3+ ions, Γ is Eulers’ function (Γ is 1.77, 1.43, and 1.3 for s = 6, 8, and 10, respectively), s is the mechanism of multipolar interaction (6 for DD, 8 for DQ, 10 for QQ), NA is the concentration of RE3+ ions, and R0 is the critical transfer distance between two neighboring RE3+ ions. When non-radiative losses through cross-relaxation processes between two neighboring dopant ions are present, the microparameter of donor-acceptor interaction (CDA) and the transfer rates (WDA) can be calculated by the following equations, CDA = R 0 s τ 0 1 and WDA = CDA/ R 0 s , respectively. The energy transfer rate (WET) through cross relaxation [57], as well as quantum efficiency ɳ, can be also calculated by the formulas W E T = 1 τ 1 τ r a d and ɳ (ɳ = τ/τrad), respectively.
  • Sm3+ ions
The absorption spectra of 3.4 at.% Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals were measured at 300 K in the spectral range of 350–3000 nm and analyzed within the JO theory [48,49]. The obtained spectra are shown in Figure 3. Twelve absorption lines of Sm3+ ions were identified and analyzed to determine the JO parameters [21].
Based on measured and calculated line strengths [21], the Ωt (t = 2, 4, 6) parameters of CNGG:Sm and CLNGG:Sm crystals were obtained and are listed in Table 2. The order Ω4 > Ω2 > Ω6 of the parameters is in trend with that obtained for different Sm3+-doped crystals with similar structures [57,58,59,60]. The significance of each JO parameter was studied [61,62,63] and it was established that Ω2 is an intensity parameter very sensitive to the crystal field asymmetry in the RE3+ ion site, the covalency of the RE3+ ion, as well as to any modification in the energy gap between the 4fn and 4fn−15d states of the RE3+ ion. Ω6 is a parameter that reacts more to any variation in the electron density of the 4f and 5d configurations. Any alteration of the Ω2 and Ω6 parameters have an impact on the Ω4 parameter, which frequently complicates the establishment of the real factors that influence its modification. The values of spontaneous emission probabilities (AJJ’), branching ratios JJ’), and radiative lifetimes (τr) for the 4G5/2 excited level were determined based on the Ωt (t = 2, 4, 6) parameters [21], and are given in Table 3. The radiative lifetime values of the 4G5/2 level were found to be 1.58 ms and 1.5 ms, for CNGG:Sm and CLNGG:Sm crystals, respectively.
Figure 4a shows the excitation spectra of CNGG and CNLGG host crystals obtained by observing the blue emission line at 450 nm (Nb5+). The spectra exhibit a charge transfer (CT) band in the UV range positioned at 272 nm and 280 nm for CLNGG and CNGG, respectively, assigned to the Nb5+-O2− interactions into the [NbO4] tetrahedrons of the host lattices [17]. Compared to CNGG, the shift to the higher energy of the CT band peak for CLNGG is due to the insertion of Li+ ions into the CNGG structure which practically removes the vacancies producing a slight distortion of the [NbO4] tetrahedrons. The emission spectra under 280 nm and 272 nm excitation wavelengths of the CNGG and CLNGG hosts, respectively, are shown in Figure 4b. As can be seen, both crystals display a wide emission band in the VIS range with a peak located at about 450 nm. Figure 4c shows the excitation spectra of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals by observing the emission line of Sm3+ ions in orange at 615 nm attributed to the 4G5/26H7/2 transition. The spectra present the CT bands from 200 to 300 nm assigned to the Nb5+-O2− interactions into the [NbO4] tetrahedrons of the host lattices, and a group of narrow lines in the UV-VIS region assigned to the 4f5 electronic configuration of Sm3+ ions.
The emission spectra of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals under 270 nm and 405 nm excitations were recorded at room temperature (Figure 5). Sm3+ ion has an intense absorption band around 405 nm attributed to the 6H5/26P3/2, 6P5/2 transitions, which is very appropriate for the efficient pumping with InGaN laser diodes at ~405 nm. The values of the absorption cross-sections for the peaks at 404.9 nm and 405.6 nm were determined to be 2.1 × 10−20 cm2 and 2.3 × 10−20 cm2, respectively, for CNGG:Sm, and 2.2 × 10−20 cm2 and 2.8 × 10−20 cm2, respectively, for the CLNGG:Sm crystal [21]. Under both excitation wavelengths (270 nm and 405 nm), the emission spectra of Sm3+ ions exhibit three emission bands centered at 567 nm, 615 nm, and 662 nm corresponding to the 4G5/26H5/2, 4G5/26H7/2, and 4G5/26H9/2 transitions, respectively. Under 270 nm excitation, a low-intensity emission band around 450 nm attributed to the host emission can be observed (Figure 5a). The emission line at 615 nm assigned to the 4G5/26H7/2 transition is the most intense line observed for both CNGG:Sm and CLNGG:Sm crystals, under both 270 nm and 405 nm excitation wavelengths. The emission cross sections were calculated for each emission band corresponding to the transitions from the 4G5/2 level to 6H5/2, 6H7/2, and 6H9/2 lower levels of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals. The highest values of the emission cross-sections were obtained for the CLNGG:Sm crystal at 615 nm (4G5/26H7/2) and 662 nm (4G5/26H9/2). All obtained values are presented and compared with other similar materials in Table 4. The values of Sm3+ emission cross-sections at 615 nm suggest that CNGG:Sm and CLNGG:Sm are potential laser crystals with orange emission at 615 nm.
The partial energy levels and the multiplet barycenter (Bc) of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals were identified based on absorption and emission spectra at 10 K [21]. The energy levels involved in the main Sm3+ emission transitions are given and compared with those for the YAG:Sm crystal [57] in Table 5, to highlight the effect of partial structural disorder on the energy levels of Sm3+ in CNGG:Sm and CLNGG:Sm compared to the structurally ordered YAG:Sm crystal. The main crystal field effects are given by the perturbations of the crystal field induced by the disorder through the mixed occupation of the cationic sublattices, charge difference, and size mismatch effects. All these facts lead to the formation of multiple Sm3+ centers, a low local crystal field, as well as changes in the spectral properties of the Sm3+ ions by modifying the emission wavelengths.
The luminescence kinetics of the 4G5/2 level of Sm3+ ions in CNGG:Sm and CLNGG:Sm at room temperature were recorded by observing the emission line at 615 nm under 405 nm excitation. To find out the lifetime of isolated Sm3+ ions, CNGG:Sm and CLNGG:Sm sintered ceramics doped with 0.1 at.% Sm3+ ions were also measured. The normalized decay curves are shown in Figure 6a,b. For 0.1 at. % Sm, the decay curves are nearly exponential, and the measured lifetimes were determined to be τ0 = 1.483 ms for CNGG:Sm and τ0 = 1.44 ms for CLNGG:Sm. The values obtained for isolated Sm3+ ions (τ0) are near to the values determined by the JO method, as being τr = 1.58 ms for the CNGG:Sm crystal and τr = 1.5 ms for the CLNGG:Sm crystal. This fact indicates that the values of τ0 can be considered as the radiative lifetimes, since that the energy gap between the 4G5/2 level and the next lower level is about 7200 cm−1 and the maximum phonon energy in the hosts is about 750 cm−1 [12], thus resulting in a negligible non-radiative contribution to the luminescence decay.
For 3.4 at.% Sm, the decay curves present a non-exponential shape. The average luminescence lifetimes were estimated to be τav=1 ms and τav = 0.95 ms for CNGG:Sm and CLNGG:Sm crystals, respectively. Employing the Inokuti-Hirayama (IH) model, the non-exponential decay curves were evaluated, and the experimental transfer functions were determined (Figure 6c). From the fitting of the 4G5/2 decay curve, the critical distance (R0), microparameter of donor-acceptor interaction (CDA), transfer rates (WDA), energy transfer rate (WET), and quantum efficiency (ɳ) were determined (Table 6).
The emission quenching of the 4G5/2 level with Sm concentration could be mostly due to the energy transfer (ET) by cross-relaxation [57]. The ET rate (WET) is 325 s−1 and 358 s−1 for the CNGG:Sm and CLNGG:Sm crystals, respectively. The quantum efficiency ɳ of the 4G5/2 level, which is defined as the ratio of the number of photons emitted to the number of photons absorbed, was estimated to be ~66 % for both crystals. This shows that the multiphonon relaxations and the ET processes of the 4G5/2 level are negligible, which is advantageous for lasers and photonic devices.
  • Dy3+ ions
The absorption spectra of 4.2 at.% Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals were recorded in the extended spectral domain of 330–2000 nm at room temperature (Figure 7a) [23]. The absorption spectra were analyzed within the JO theory [48,49] to evaluate the intensity parameters Ωt (t = 2, 4, 6) contributing to the determination of the electric and magnetic dipole spontaneous emission probabilities (AJJ’), spectroscopic quality factor (χ = Ω46), branching ratios JJ’), and the radiative lifetime r) of the 4F9/2 Dy3+ emitting manifold [23]. Table 7 summarizes all the results obtained.
Figure 7b shows the excitation spectra of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals [23] obtained by observing the yellow line at 579 nm assigned to the 4F9/26H13/2 transition. For both crystals, the spectra present a CT band in the UV range of 200–300 nm assigned to the Nb5+-O2− interactions into the [NbO4] tetrahedrons of the host lattices, and a group of narrow lines in the UV-VIS domain assigned to the 4f9 electronic configuration of Dy3+ ions. Being positioned at wavelengths shorter than 200 nm according to Reference [68], the CT band due to the Dy3+−O2− interaction could not be highlighted.
The emission spectra of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals (Figure 8a,b) were recorded at 300 K under different excitation wavelengths of λex = 272, 280, and 352 nm. The obtained spectra show emission bands in the blue, yellow, and red spectral regions, assigned to the 4F9/26H15/2, 4F9/26H13/2, and 4F9/26H11/2 transitions, respectively, but the most intense emission line is in the yellow domain at 579 nm for both CNGG:Dy and CLNGG:Dy crystals. Under the host excitations of λex = 272 and 280 nm (Figure 8), it can be seen that the emission lines of Dy3+ ions have low intensity suggesting that the energy transfer from the host to Dy3+ ions is weak in both investigated crystals. The value of the emission cross-section for yellow emission at 579 nm was calculated to be σem = 0.33×10−20 cm2, being similar in both CNGG:Dy and CLNGG:Dy crystals. Moreover, other laser parameters such as branching ratios JJ’) and Y/B (yellow/blue) ratio, required for efficient lasing in the yellow domain, were determined to have similar or even higher values compared to other crystals (Table 8). Therefore, CNGG:Dy and CLNGG:Dy crystals are promising yellow gain media.
The partial energy levels and the multiplet barycenter (Bc) of Dy3+ ions in the CNGG:Dy and CLNGG:Dy crystals were determined by using the absorption and emission spectra at 10 K [23]. The energy levels involved in the yellow emission transition of Dy3+ compared with those of the Dy:YAG crystal [73] are given in Table 9.
The luminescence decays of the 4F9/2 level of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals at room temperature were measured by observing the yellow line at 579 nm under the excitation wavelength of 355 nm. To find out the lifetime of isolated Dy3+ ions, sintered ceramic samples of CNGG:Dy and CLNGG:Dy doped with 0.1 at.% Dy were also measured. The normalized luminescence decays are shown in Figure 9. The decay curve is exponential for the low concentration of 0.1 at.% Dy in the CNGG:Dy ceramic, with a measured lifetime of τ0 ~ 543 μs (Figure 9a), while in the case of the CLNGG:Dy ceramic, the decay has a non-exponential profile (Figure 9b) indicating the presence of ET processes at the doping concentration of 0.1 at.% Dy. In this case, the average lifetime was determined to be τav = 405 μs.
For the 4.2 at.% Dy concentration, the decay curves present a non-exponential shape (Figure 9), and the average luminescence lifetimes were estimated to be τav=229 μs and τav = 143 μs for the CNGG:Dy and CLNGG:Dy crystals, respectively. Using the IH model, the non-exponential decay curves were evaluated, and the experimental transfer functions were determined. From the fitting of the 4F9/2 luminescence decay, the critical distance (R0), microparameter of donor-acceptor interaction (CDA), transfer rates (WDA), energy transfer rate (WET), and quantum efficiency (ɳ) were determined (Table 10) and compared with other Dy3+-doped crystals in Table 10 [57,58,59]. The analysis of the decay curves highlighted the efficient ET between Dy3+ ions in the CNGG:Dy and CLNGG:Dy crystals. The high values of the quantum efficiency of the 4F9/2 level, ɳ ~ 73 %, coupled with the favorable emission cross-sections at 579 nm, σem = 0.33 × 10−20 cm2, indicate that Dy3+ ions doped in CNGG:Dy and CLNGG:Dy are potential laser crystals in yellow at 579 nm.
  • Pr3+ ions
The absorption spectra of the 1.8 at.% Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals (Figure 10) were measured in the 400–2750 nm domain at room temperature [24] and analyzed within the JO theory [48,49]. For the estimation of JO parameters, seven absorption bands of Pr3+ ions were considered. Generally, the traditional Judd-Ofelt theory supposes that the 4f5d configurations are situated at higher energies than the 4fn configuration. For the Pr3+ ions, the difference between the 4f2 configuration and 4f15d1 level is only about ΔE~15000 cm−1, implying the existence of significant mixing between these two configurations. Therefore, a large difference between the calculated and theoretical results is usually obtained, particularly when the 3H43P2 transition is considered in the JO parameters calculation [74]. These incompatibilities are generated by the theoretical characteristics of the standard JO model and can be eliminated by removing the 3H43P2 transition or by employing a modified JO analysis.
The modified OJ theory proposed by Kornienko et al. [74] assumes that the transition probabilities from the ground level to the 4f15d1 state are higher than those calculated by standard JO theory, and therefore the energy difference between the 4f2 and 4f15d1 configurations is taken into account, and the absorption oscillator strength is expressed as:
S mod S standard × 1 + E i 2 E f 0 E 5 d E f 0
where Sstandard is the standard oscillator strength and Ei, E5d, and E f 0 are the energies of the final state, the 4f15d1 state, and the mean energy of the 4f2 configuration, respectively. For our crystals, the mean energy of the 4f2 configuration was determined to be E f 0 = 10578 cm−1 for the CNGG:Pr crystal and E f 0 = 10660 cm−1 for the CLNGG:Pr crystal. Based on the absorption and excitation spectra, the energy of the 4f15d1 lower level of the 4f5d configuration was found to be E5d = 31348 cm−1 (319 nm) in both CNGG:Pr and CLNGG:Pr crystals. Consequently, the modified JO analysis excluding the 3H43P2 + 1I6 transitions was applied, and the obtained Ωt mod (t = 2, 4, 6) intensity parameters are given in Table 11. The values of spontaneous emission probabilities (AJJ’), branching ratios JJ’), and radiative lifetimes (τr) for the excited levels of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals [24] were determined based on the values of Ωt mod parameters, and are also shown in Table 11.
The excitation spectra of the 1.8 at.% Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals were recorded by observing the emission line at λem = 606 nm assigned to the 1D23H4 transition (Figure 11). The obtained spectra present a wide UV band in the 200–400 nm spectral range and a group of narrow lines in the 420–490 nm spectral domain assigned to the 4f2 electronic configuration of Pr3+ ions. The wide bands in the UV domain contain two types lines: the first one is associated with the CT bands assigned to the Nb5+−O2− interactions into the [NbO4] tetrahedrons of the host lattices (peaks at 272 nm for CLNGG and 280 nm for CNGG), while the second type are the peaks placed at λ = 319 nm and attributed to the 4f15d1 lower level of the 4f5d configuration of Pr3+ ions.
However, when the Pr3+ and Nb5+ ions are both contained in a material, a photo-induced redox process Pr3++Nb5+→Pr4++ Nb4+ takes place, leading to the formation of an intervalence charge transfer (IVCT) band. If the IVCT band is located at low energies, it can interact with the 3P0,1,2 manifolds of the Pr3+ ions leading to the emission quenching of these manifolds. The quenching of emission from the 3P0 level in materials containing M4+ or M5+ metal ions (M = Ti, V, Nb, and Ta) was intensively studied [24,75,76,77,78]. When the IVCT band is positioned at higher energies, there is no influence on the 3P0 level emissions and intense emissions from the 3P0 level can be obtained, especially a strong emission in the blue domain attributed to the 3P03H4 transition [24,77]. Therefore, the UV line situated around 319 nm can contain both the 4f15d1 lower level of 4f5d configuration and the IVCT band.
The emission spectra of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals in the VIS (Figure 12a,b) and near-infrared spectral ranges (Figure 12c,d) were registered under UV (275–375 nm) and 450 nm excitation wavelengths at room temperature [24]. Under UV excitation, the spectra exhibit very low intensities for the emission lines arising from the 3PJ (J=0,1,2,) manifolds, while the emission lines arising from the 1D2 manifold are intense and well-structured in both spectral domains. Under UV excitation, the electrons migrate from the (4f15d1 level + IVCT) band passing through the 3PJ (J = 0,1,2,) manifolds to the 1D2 multiplet, thus leading to obtaining a dominating 1D23H4 emission [79]. A total quenching of the 3P0 level emission was observed for the NaYTiO4:Pr phosphor [80], where the energy gap between the 3P0 level and the IVCT band is less than 7400 cm−1. For the CNGG:Pr and CLNGG:Pr crystals [24], this energy gap is around 11827 cm−1, thus explaining the partial quenching of emission from the 3P0 level. Under direct excitation at 450 nm in the 3P2 level, the obtained spectra show more intense emission lines from both 3PJ (J = 0,1,2,) and 1D2 manifolds compared with those obtained under 275–375 nm UV excitation.
Table 12 shows the stimulated emission cross-sections corresponding to the 3P03H4, 1D23H4, and 3P03F2 transitions obtained for the investigated crystals in comparison with other similar crystals. The high values of the emission cross-sections indicate that the CNGG:Pr and CLNGG:Pr are potential laser crystals in the blue, red, and orange domains.
The partial energy level scheme and the manifold barycenter (Bc) of Pr3+ ions are presented in detail in Reference [24]. The levels involved in the blue, orange, and red emissions assigned to the 3P03H4, 1D23H4, and 3P03F2 transitions, respectively, of the Pr3+ ions in CNGG:Pr and CLNGG:Pr are given and compared with those of the Pr: YAG crystal [84] in Table 13.
The emission kinetics of 1D2 level of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals were measured at room temperature by observing the emission line at 606 nm under direct excitation at 590 nm [24]. To obtain the lifetime of isolated Pr3+ ions, ceramic samples of CNGG:Pr and CLNGG:Pr doped with 0.01 at.% Pr were also sintered and measured. The normalized luminescence decays are presented in Figure 13. The luminescence decay is exponential for the low concentration of 0.01 at.% Pr in the CNGG:Pr ceramic, with a measured lifetime of τ0 ~ 128 μs (Figure 13a), while in the case of the CLNGG:Pr ceramic, the decay has a non-exponential profile (Figure 13b) indicating the presence of ET processes at the doping concentration of 0.01 at.% Pr. In this case, the average lifetime was determined to be τav = 110 μs. For the doping concentration of 1.8 at.% Pr, the decay curves have a non-exponential shape, and the measured lifetime decrease drastically to τav = 18 μs and τav = 15 μs, respectively. The quenching emission of the 1D2 level at room temperature can increase due to the increase of the Pr3+-Pr3+ non-radiative interactions if the resonance between levels is fulfilled or the ET process is phonon-assisted. Due to the presence of Pr3+ multicenters, the application of the IH model does not seem appropriate since the analysis of the 1D2 decay profiles may lead to inaccurate results.

4. Conclusions

The growth and spectroscopic characteristics of RE3+ ions (RE = Sm, Dy, Pr) doped in partially disordered Ca3Nb1.6875Ga3.1875O12 – CNGG and Ca3Li0.275Nb1.775Ga2.95O12 – CLNGG crystals were reviewed. High-quality CNGG:RE and CLNGG:RE crystals were grown using the Czochralski crystal growth technique. The JO model was employed to determine the Ωt (t = 2, 4, 6) intensity parameters and to evaluate the spectroscopic properties and laser emission features of the Pr3+, Sm3+, and Dy3+ ions, respectively, doped in the grown crystals. The partial energy levels of each dopant ion were determined. The luminescence decays of the 4G5/2 (Sm3+) and 4F9/2 (Dy3+) levels were evaluated within the IH theory, and the ET parameters were determined. For the Pr3+ ion, the emission decays of the 1D2 level show a drastic decrease of the lifetime values at high doping concentrations due to the cross-relaxation processes, and the application of the IH model does not seem appropriate because may lead to inaccurate results. The high values obtained for the stimulated emission cross-sections indicate that CNGG and CLNGG crystals doped with Sm3+, Dy3+, and Pr3+ ions could be promising materials to achieve laser emission in the orange (Sm3+), yellow (Dy3+), and blue, orange, and red (Pr3+) domains.

Author Contributions

Conceptualization, C.G.; writing-original draft preparation, C.G.; methodology, C.G, L.G. and S.H.; formal analysis, C.G, L.G. and S.H.; investigation, F.V, M.G, A.B. and G.S.; data curation, C.G., L.G., S.H., F.V., M.G., A.B. and G.S.; writing-review and editing, C.G., S.H. and L.G.; supervision, C.G. and L.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Romanian Ministry of Research, Innovation and Digitization, through grant agreement PCE 49/2021, project number PN-III-P4-ID-PCE-2020–2203.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ramírez, M.; Romero, J.; Molina, P.; Bausá, L.E. Near infrared and visible tunability from a diode pumped Nd3+ activated strontium barium niobate laser crystal. Appl. Phys. B 2005, 81, 827–830. [Google Scholar] [CrossRef]
  2. Bettinelli, M.; Speghini, A.; Ródenas, A.; Molina, P.; Ramírez, M.; Capote, B.; Jaque, D.; Bausá, L.E.; García Solé, J. Luminescence of lanthanide ions in strontium barium niobate. J. Lumin. 2007, 122–123, 307–310. [Google Scholar] [CrossRef]
  3. Xie, G.Q.; Tang, D.Y.; Tan, W.D.; Luo, H.; Zhang, H.J.; Yu, H.H.; Wang, J.Y. Subpicosecond pulse generation from a Nd:CLNGG disordered crystal laser. Opt. Lett. 2009, 34, 103–105. [Google Scholar] [CrossRef] [PubMed]
  4. Gheorghe, C.; Gheorghe, L.; Loiseau, P.; Aka, G. Spectroscopic features, and laser performance at 1.06 µm of Nd3+ doped Gd1−xLuxCa4O(BO3)3 single crystal. J. Appl. Phys. 2012, 111, 13102. [Google Scholar] [CrossRef]
  5. Marzahl, D.T.; Reichert, F.; Metz, P.W.; Fechner, M.; Hansen, N.O.; Huber, G. Efficient laser operation of diode-pumped Pr3+,Mg2+:SrAl12O19. Appl. Phys. B 2014, 116, 109–113. [Google Scholar] [CrossRef]
  6. Sani, E.; Brugioni, A.; Mercatelli, L.; Parisi, D.; Zharikov, E.V.; Lis, D.A.; Subbotin, K.A. Yb-doped double tungstates for down-conversion applications. Opt. Mater. 2019, 94, 415–422. [Google Scholar] [CrossRef]
  7. Broasca, A.; Greculeasa, M.; Voicu, F.; Hau, S.; Croitoru, G.; Gheorghe, C.; Pavel, N.; Gheorghe, L. Efficient near-infrared laser emission and nonlinear optical properties of a newly developed Yb:LYSB laser crystal. J. Alloy. Compd. 2020, 844, 156143. [Google Scholar] [CrossRef]
  8. Kaminskii, A.A.; Belokoneva, E.L.; Butashin, A.V.; Markosian, A.A.; Mill, B.V.; Nikolskaia, O.; Sarkisov, S.E. Crystal structure and spectral luminescence properties of the cation-deficient garnet Ca3(Nb, Ga)2Ga3O12-Nd3+. Inorg. Mater. 1986, 22, 927–936. [Google Scholar]
  9. Voron’ko, Y.K.; Kudryavtsev, A.B.; Es’kov, N.A.; Osiko, V.V.; Sobol’, A.A.; Sorokin, E.V.; Spiridonov, F.M. Raman scattering of light in crystals and melt of calcium-niobium gallium garnet. Sov. Phys. Dokl. 1988, 32, 70–73. [Google Scholar]
  10. Shi, Z.B.; Zhang, H.J.; Wang, J.Y.; Yu, Y.G.; Wang, Z.P.; Yu, H.H.; Sun, S.Q.; Xia, H.R.; Jiang, M.H. Growth and characterization of Nd:CLNGG crystal. J. Cryst Growth 2009, 311, 3792. [Google Scholar] [CrossRef]
  11. Voronko, Y.K.; Gessen, S.B.; Es’Kov, N.A.; Sobol, A.A.; Ushakov, S.N.; Tsymbal, L.I. Efficient active media based on Nd3+-activated calcium niobium gallium garnets. Sov. J. Quantum Electron. 1990, 20, 246. [Google Scholar] [CrossRef]
  12. Yu, Y.M.; Chani, V.I.; Shimamura, K.; Fukuda, T. Growth of Ca3(Li,Nb,Ga)5O12 garnet crystals from stoichiometric melts. J. Cryst. Growth 1997, 171, 463. [Google Scholar] [CrossRef]
  13. Brenier, A.; Boulon, G.; Shimamura, K.; Fukuda, T. Growth by the μ-pulling-down method and spectroscopic investigation of Nd3+-doped calcium niobium gallium garnet. J. Cryst. Growth 1999, 204, 145. [Google Scholar] [CrossRef]
  14. Agnesi, A.; Dell’Acqua, S.; Guandalini, A.; Reali, G.; Cornacchia, F.; Toncelli, A.; Toncelli, M.; Shimamura, K.; Fukuda, T. Optical spectroscopy and diode-pumped laser performance of Nd3+ in the CNGG crystal. IEEE J. Quant. Electron. 2001, 37, 304–313. [Google Scholar] [CrossRef]
  15. Lupei, A.; Lupei, V.; Gheorghe, L.; Rogobete, L.; Petraru, A. The nature of nonequivalent Nd3+centers in CNGG and CLNGG. Opt. Mater. 2001, 16, 403–411. [Google Scholar] [CrossRef]
  16. Voron’Ko, Y.K.; Sobol’, A.A.; Ushakov, S.N.; Shukshin, V.E. Spectroscopy of Disordered Laser Crystals. Inorg. Mater. 2002, 38, 390–396. [Google Scholar] [CrossRef]
  17. Voronko, Y.; Sobol, A.; Karasik, A.; Eskov, N.; Rabochkina, P.; Ushakov, S. Calcium niobium gallium and calcium lithium niobium gallium garnets doped with rare earth ions–effective laser media. Opt. Mater. 2002, 20, 197–209. [Google Scholar] [CrossRef]
  18. Gao, W.L.; Xie, G.Q.; Ma, J.; Liu, M.N.; Yuan, P.; Qian, L.J.; Yu, H.H.; Zhang, H.J.; Wang, J.Y.; Zhang, J. Spectroscopic characteristics and efficient laser operation of Tm: CLNGG disordered crystal. Laser Phys. Lett. 2013, 10, 055809. [Google Scholar] [CrossRef]
  19. Toma, O.; Georgescu, S. Judd-Ofelt analysis of Er3+ions in calcium lithium niobium gallium garnet. J. Lumin. 2014, 147, 259–264. [Google Scholar] [CrossRef]
  20. Xiao, K.; Lin, B.; Zhang, Q.-L.; Zhang, D.-X.; Feng, B.-H.; He, J.-L.; Zhang, H.-J.; Wang, J.-Y. Q-switched 1329 nm Nd: CNGG laser. Appl. Opt. 2001, 54, 7071–7075. [Google Scholar] [CrossRef]
  21. Gheorghe, C.; Hau, S.; Gheorghe, L.; Voicu, F.; Greculeasa, M.; Achim, A.; Enculescu, M. Optical properties of Sm3+ doped Ca3(Nb,Ga)5O12 and Ca3(Li,Nb,Ga)5O12 single crystals. J. Lumin. 2017, 186, 175–182. [Google Scholar] [CrossRef]
  22. Ryabochkina, P.A.; Chabushkin, A.N.; Zakharov, N.G.; Vorontsov, K.V.; Khrushchalina, S.A. Tunable 2-μm lasing in calcium—niobium—gallium garnet crystals doped with Ho3+ ions. Quantum Electron. 2017, 47, 607–609. [Google Scholar] [CrossRef]
  23. Gheorghe, C.; Hau, S.; Gheorghe, L.; Voicu, F.; Greculeasa, M.; Enculescu, M.; Belikov, K.N.; Bryleva, Y.E.; Gaiduk, O.V. Yellow laser potential of cubic Ca3(Nb,Ga)5O12:Dy3+ and Ca3(Li,Nb,Ga)5O12:Dy3+ single crystals. J. Alloys Compd. 2018, 739, 806–816. [Google Scholar] [CrossRef]
  24. Hau, S.; Gheorghe, C.; Gheorghe, L.; Voicu, F.; Greculeasa, M.; Stanciu, G.; Broasca, A.; Enculescu, M. Spectroscopic investigations of Pr3+ ions doped CNGG and CLNGG single crystals. J. Alloys Compd. 2019, 799, 288–301. [Google Scholar] [CrossRef]
  25. Hau, S.; Stanciu, G.; Avram, D.; Gheorghe, L.; Gheorghe, C. Energy transfer and luminescent properties of Tb3+ and Tb3+, Yb3+ doped CNGG phosphors. J. Rare Earths 2022, 40, 1445–1453. [Google Scholar] [CrossRef]
  26. Yu, H.; Zhang, H.; Wang, Z.; Wang, J.; Yu, Y.; Shi, Z.; Zhang, X.; Jiang, M. High-power dual-wavelength laser with disordered Nd:CNGG crystals. Opt. Lett. 2009, 34, 151–153. [Google Scholar] [CrossRef]
  27. Yu, H.; Zhang, H.; Wang, Z.; Wang, J.; Yu, Y.; Shi, Z.; Zhang, X.; Jiang, M. Continuous-wave and passively Q-switched laser performance with a disordered Nd:CLNGG crystal. Opt. Express 2009, 17, 19015–19020. [Google Scholar] [CrossRef]
  28. Li, Y.L.; Jiang, H.L.; Ni, T.Y.; Zhang, T.Y.; Tao, Z.H.; Zeng, Y.H. Efficient 1061 and 1329 nm laser emission of Nd:CNGG lasers under 885 nm diode pumping into the emitting level. Laser Phys. 2011, 21, 485–488. [Google Scholar] [CrossRef]
  29. Liu, C.-X.; Shen, X.-L.; Li, W.-N.; Wei, W. Visible and near-infrared optical properties of Nd: CLNGG crystal waveguides formed by proton implantation. Chin. Phys. B 2017, 26, 034207. [Google Scholar] [CrossRef]
  30. Schmidt, A.; Griebner, U.; Zhang, H.; Wang, J.; Jiang, M.; Liu, J.; Petrov, V. Passive mode-locking of the Yb:CNGG laser. Opt. Commun. 2010, 283, 567. [Google Scholar] [CrossRef]
  31. Ma, J.; Pan, Z.; Wang, J.; Yuan, H.; Cai, H.; Xie, G.; Qian, L.; Shen, D.; Tang, D. Generation of sub-50fs soliton pulses from a mode-locked Yb,Na: CNGG disordered crystal laser. Opt Express 2017, 25, 14968–14973. [Google Scholar] [CrossRef] [PubMed]
  32. Si, W.; Ma, Y.-J.; Wang, L.-S.; Yuan, H.-L.; Kong, W.-J.; Liu, J.-H. Acousto-Optically Q-Switched Operation of Yb: CNGG Disordered Crystal Laser. Chin. Phys. Lett 2017, 34, 124201. [Google Scholar] [CrossRef]
  33. Merkle, L.D.; Zandi, B. Spectroscopy and laser operation of Pr, Mg:SrAl12O19. J. Appl. Phys. 1996, 79, 1849. [Google Scholar] [CrossRef]
  34. Fibrich, M.; Jelínková, H.; Šulc, J.; Nejezchleb, K.; Škoda, V. Visible cw laser emission of GaN-diode pumped Pr:YAlO3 crystal. Appl. Phys. B 2009, 97, 363. [Google Scholar] [CrossRef]
  35. Reichert, F.; Moglia, F.; Marzahl, D.-T.; Metz, P.; Fechner, M.; Hansen, N.-O.; Huber, G. Diode pumped laser operation and spectroscopy of Pr3+:LaF3. Opt. Express 2012, 20, 20387–20395. [Google Scholar] [CrossRef]
  36. Luo, S.; Cai, Z.; Xu, H.; Shen, Z.; Chen, H.; Li, L.; Cao, Y. Direct oscillation at 640-nm in single longitudinal mode with a diode-pumped Pr:YLF solid-state laser. Opt. Laser Techn. 2019, 116, 112–116. [Google Scholar] [CrossRef]
  37. Damiano, E.; Cittadino, G.; Di Lieto, A.; Tonelli, M. Deep Red Tunability and Output Power Performances of Visible Laser Emission in a Pr:Ba(Y1−xLux)2F8 Single Crystal. Materials 2020, 13, 4655. [Google Scholar] [CrossRef]
  38. Wang, G.Q.; Lin, Y.F.; Gong, X.H.; Chen, Y.J.; Huang, J.H.; Luo, Z.D.; Huang, Y.D. Polarized spectral properties of Sm3+:LiYF4 crystal. J. Lumin. 2014, 147, 23–26. [Google Scholar] [CrossRef]
  39. Demesh, M.P.; Dernovich, O.P.; Gusakova, N.V.; Yasukevich, A.S.; Kornienko, A.A.; Dunina, E.B.; Fomicheva, L.A.; Pavlyuk, A.A.; Kuleshov, N.V. Growth and spectroscopic properties of Sm3+:KY(WO4)2 crystal. Opt. Mater. 2018, 75, 821–826. [Google Scholar] [CrossRef]
  40. Huang, J.; Huang, J.; Lin, Y.; Gong, X.; Chen, Y.; Luo, Z.; Huang, Y. Spectroscopic properties of Sm3+-doped NaGd(MoO4)2 crystal for visible laser application. J. Lumin. 2017, 187, 235–239. [Google Scholar] [CrossRef]
  41. Zhu, K.; Zhou, H.; Qiu, J.; Wang, L.G.; Ye, L. Optical temperature sensing characteristics of Sm3+ doped YAG single crystal fiber based on luminescence emission. J. Alloys Compd. 2022, 890, 161844. [Google Scholar] [CrossRef]
  42. Fujimoto, Y.; Ishii, O.; Yamazaki, M. 575 nm laser oscillation in Dy3+-doped waterproof fluoro-aluminate glass fiber pumped by violet GaN laser diodes. In Solid State Lasers XX: Technology and Devices; SPIE: Washington, DC, USA, 2011; Volume 7912. [Google Scholar] [CrossRef]
  43. Haro-González, P.; Martín, L.L.; Martín, I.R.; Dominiak-Dzik, G.; Ryba-Romanowski, W. Pump and probe measurements of optical amplification at 584nm in dysprosium doped lithium niobate crystal. Opt. Mater. 2010, 33, 196–199. [Google Scholar] [CrossRef]
  44. Ganesh Kumar, K.; Balaji Bhargav, P.; Aravinth, K.; Arumugam, R.; Ramasamy, P. Dysprosium activated strontium aluminate phosphor: A potential candidate for WLED applications. J. Lumin. 2020, 223, 117126. [Google Scholar] [CrossRef]
  45. Tang, W.; Guo, Q.; Su, K.; Liu, H.; Zhang, Y.; Mei, L.; Liao, L. Structure and Photoluminescence Properties of Dy3+ Doped Phosphor with Whitlockite Structure. Materials 2022, 15, 2177. [Google Scholar] [CrossRef]
  46. Boruc, Z.; Kaczkan, M.; Fetlinski, B.; Turczynski, S.; Malinowski, M. Blue emissions in Dy3+ doped Y4Al2O9 crystals for temperature sensing. Opt. Lett. 2012, 37, 5214–5216. [Google Scholar] [CrossRef]
  47. Gheorghe, L.; Greculeasa, M.; Voicu, F.; Gheorghe, C.; Hau, S.; Vlaicu, A.M.; Belikov, K.N.; Bryleva, Y.E.; Gaiduk, O.V. Crystal growth and structural characterization of Sm3+, Pr3+ and Dy3+-doped CNGG and CLNGG single crystals. Opt. Mater. 2018, 84, 335–338. [Google Scholar] [CrossRef]
  48. Judd, B.R. Optical Absorption Intensities of Rare-Earth Ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  49. Ofelt, G.S. Intensities of Crystal Spectra of Rare-Earth Ions. J. Chem. Phys. 1962, 37, 511–520. [Google Scholar] [CrossRef]
  50. Zhang, H.; Liu, J.H.; Wang, J.Y.; Fan, J.D.; Tao, X.T.; Mateos, X.; Petrov, V.; Jiang, M.H. Spectroscopic properties and continuous-wave laser operation of a new disordered crystal: Yb-doped CNGG. Opt. Express 2007, 15, 9464. [Google Scholar] [CrossRef]
  51. Basiev, T.T.; Es’kov, N.A.; Karasik, A.Y.; Osiko, V.V.; Sobol, A.A.; Ushakov, S.N.; Helbig, M. Disordered garnets Ca3(Nb, Ga)5O12:Nd3+-prospective crystals for powerful ultrashort-pulse generation. Opt. Lett. 1992, 17, 201. [Google Scholar] [CrossRef]
  52. Carnall, W.T.; Fields, P.R.; Rajnak, K. Electronic Energy Levels in the Trivalent Lanthanide Aquo Ions. I. Pr3+, Nd3+, Pm3+, Sm3+, Dy3+, Ho3+, Er3+, and Tm3+. J. Chem. Phys. 1968, 49, 4424–4442. [Google Scholar] [CrossRef]
  53. Krupke, W.F. Induced-emission cross sections in neodymium laser glasses. IEEE J. Quant. Electron. 1974, 10, 450–457. [Google Scholar] [CrossRef]
  54. Förster, T. Experimental and theoretical investigation of intermolecular transfer of electronic excitation energy. Z. Naturforsch. 1949, 4a, 321. [Google Scholar] [CrossRef]
  55. Dexter, D.L. Theory of the optical properties of imperfections in nonmetals. Solid State Phys. 1958, 6, 355. [Google Scholar] [CrossRef]
  56. Inokuti, M.; Hirayama, F. Influence of energy transfer by the exchange mechanism on donor luminescence. J. Chem. Phys. 1965, 43, 1978–1989. [Google Scholar] [CrossRef]
  57. Malinowski, M.; Wolski, R.; Frukacz, Z.; Lukasiewicz, T.; Luczyrlski, Z. Spectroscopic studies of YAG:Sm3+ crystals. J Appl Spectrosc 1995, 62, 840–843. [Google Scholar] [CrossRef]
  58. Di, J.; Xu, X.; Xia, C.; Zeng, H.; Cheng, Y.; Li, D.; Zhou, D.; Wu, F.; Cheng, J.; Xu, J. Crystal growth and optical properties of Sm:CaNb2O6 single crystal. J. Alloys Compd. 2012, 536, 20. [Google Scholar] [CrossRef]
  59. Dominiak-Dzik, G.; Ryba-Romanowski, W.; Lisiecki, R.; Solarz, P.; Berkowski, M. Dy-doped Lu2SiO5 single crystal: Spectroscopic characteristics and luminescence dynamics. Appl. Phys. B 2010, 99, 285–297. [Google Scholar] [CrossRef]
  60. Wang, Y.; Wang, B.; Wu, D.; Lin, S.; Zhou, Z.; Zhu, Y. Optical properties of Sm3+ doped Mg:LiNbO3 and Zn:LiNbO3 single crystals. Opt. Mater. 2012, 34, 845. [Google Scholar] [CrossRef]
  61. Krupke, W.F. Optical Absorption and Fluorescence Intensities in Several Rare-Earth-Doped Single Crystals. Phys. Rev. 1966, 145, 325. [Google Scholar] [CrossRef]
  62. Ravi Kanthumar, V.V.; Bhatnnagar, A.K. Effect of modifier ions on the covalency of Nd3+ ions in cadmium borate glasses. Opt. Mater. 1998, 11, 41. [Google Scholar] [CrossRef]
  63. Voronko, Y.K.; Popov, A.V.; Sobol, A.A.; Ushakov, S.N. Yb3+ ion in calcium niobium gallium garnet crystals: Nearest neighbor environment and optical spectra. Inorg. Mater. 2006, 42, 1133–1137. [Google Scholar] [CrossRef]
  64. He, X.; Zhang, L.; Chen, G.; Hang, Y. Crystal growth and spectral properties of Sm:GdVO4. J. Alloys Compd. 2012, 536, 366. [Google Scholar] [CrossRef]
  65. Malinowski, M.; Kaczkan, M.; Turczynski, S. Energy transfer and upconversion of Sm3+ ions in YAlO3. Opt. Mater. 2017, 63, 128–133. [Google Scholar] [CrossRef]
  66. Solarz, P.; Sobczyk, M. Spectroscopic properties of Sm3+ in KZnLa(PO4)2 in IR–VUV region. Opt. Mater. 2012, 34, 1826. [Google Scholar] [CrossRef]
  67. Strzep, A.; Lisiecki, R.; Solarz, P.; Ryba-Romanowski, W.; Berkowski, M. Spectroscopic characterization of Sm3+ doped (Lu0.4Gd0.6)2SiO5 single crystals. Opt. Mater. 2014, 36, 740. [Google Scholar] [CrossRef]
  68. Nagli, L.; Bunimovich, D.; Katzir, A.; Gorodetsky, O.; Molev, V. The luminescence properties of Dy-doped high silicate glass. J. Non-Cryst. Solids 1997, 217, 208–214. [Google Scholar] [CrossRef]
  69. Bowman, S.R.; O’Connor, S.; Condon, N.J. Diode pumped yellow dysprosium lasers. Opt. Express 2012, 20, 12906–12911. [Google Scholar] [CrossRef]
  70. Lisiecki, R.; Dominiak-Dzik, G.; Solarz, P.; Ryba-Romanowski, W.; Berkowski, M.; Głowacki, M. Optical spectra and luminescence dynamics of the Dy-doped Gd2SiO5 single crystal. Appl. Phys. B 2010, 98, 337–346. [Google Scholar] [CrossRef]
  71. Ning, K.; He, X.; Zhang, L.; Liu, Y.; Yin, J.; Zhang, P.; Chen, G.; Wang, X.; Chen, Z.; Shi, C.; et al. Spectroscopic characteristics of GdVO4: Dy3+ crystal. Opt. Mater. 2014, 37, 745–749. [Google Scholar] [CrossRef]
  72. Wang, Y.; You, Z.; Li, J.; Zhu, Z.; Tu, C. Optical properties of Dy3+ ion in GGG laser crystal. J. Phys. D Appl. Phys. 2010, 43, 075402. [Google Scholar] [CrossRef]
  73. Lupei, A.; Lupei, V.; Gheorghe, C.; Ikesue, A.; Enculescu, M. Spectroscopic characteristics of Dy3+ doped Y3Al5O12 transparent ceramics. J. Appl. Phys. 2011, 110, 083120. [Google Scholar] [CrossRef]
  74. Kornienko, A.A.; Kaminskii, A.A.; Dunina, E.B. Dependence of the Line Strength of f-f Transitions on the Manifold Energy, Phys. Stat. Sol. 1990, 157, 261–266. [Google Scholar] [CrossRef]
  75. Geng, J.; Chen, Y.; Gu, G.; Tian, L. Tunable white-light-emitting Sr2-xCaxNb2O7:Pr3+ phosphor by adjusting the concentration of Ca2+ ion. Opt. Mater. 2014, 36, 1093–1096. [Google Scholar] [CrossRef]
  76. Noto, L.; Yagoub, M.; Ntwaeaborwa, O.; Swart, H. Persistent photoluminescence emission from SrTa2O6:Pr3+ phosphor prepared at different temperatures. Ceram. Int. 2015, 41, 8828–8836. [Google Scholar] [CrossRef]
  77. Liu, C.; Pan, F.; Peng, Q.; Zhou, W.; Shi, R.; Zhou, L.; Zhang, J.; Chen, J.; Liang, H. Excitation wavelength dependent luminescence of LuNbO4:Pr3+-influences of intervalence charge transfer and host sensitization. J. Phys. Chem. C 2016, 120, 26044–26053. [Google Scholar] [CrossRef]
  78. Lei, R.; Luo, X.; Yuan, Z.; Wang, H.; Huang, F.; Deng, D.; Xu, S. Ultrahigh-sensitive optical temperature sensing in Pr3+:Y2Ti2O7 based on diverse thermal response from trap emission and Pr3+ red luminescence. J. Lumin. 2019, 205, 440–445. [Google Scholar] [CrossRef]
  79. Boutinaud, P.; Cavalli, E.; Bettinelli, M. Emission quenching induced by intervalence charge transfer in Pr3+- or Tb3+-doped YNbO4 and CaNb2O6. J. Phys. Condens. Matter. 2007, 19, 386230. [Google Scholar] [CrossRef]
  80. Boutinaud, P.; Pinel, E.; Oubaha, M.; Mahiou, R.; Cavalli, E.; Bettinelli, M. Making red emitting phosphors with Pr3+. Opt. Mater. 2006, 28, 9–13. [Google Scholar] [CrossRef]
  81. Pinelli, S.; Bigotta, S.; Toncelli, A.; Tonelli, M.; Cavalli, E.; Bovero, E. Study of the visible spectra of Ca3Sc2Ge3O12 garnet crystals doped with Ce3+ or Pr3+. Opt. Mater. 2004, 25, 91–99. [Google Scholar] [CrossRef]
  82. Malinowski, M.; Pracka, I.; Surma, B.; Lukasiewies, T.; Wolinski, W.; Wolski, R. Spectroscopic and laser properties of SrLaGa3O7: Pr3+ crystals. Opt. Mater. 1996, 6, 305–312. [Google Scholar] [CrossRef]
  83. Cai, G.X.; Zhou, M.; Liu, Z.; Luo, Z.Q.; Bu, Y.K.; Xu, H.Y.; Cai, Z.P.; Ye, C.C. Spectroscopic analysis of Pr3+: Gd3Ga5O12 crystal as visible laser material. Opt. Matter. 2010, 33, 191–195. [Google Scholar] [CrossRef]
  84. Gruber, J.B.; Hills, M.E.; Macfarlane, R.M.; Morrison, C.A.; Turner, G.A. Symmetry, selection rules, and energy levels of Pr3+: Y3Al5O12. Chem. Phys. 1989, 134, 241–257. [Google Scholar] [CrossRef]
Figure 1. CNGG:RE and CLNGG:RE (RE = Sm, Dy, and Pr) obtained crystals. (Reprinted with permission from Ref. [47], copyright 2018, Elsevier).
Figure 1. CNGG:RE and CLNGG:RE (RE = Sm, Dy, and Pr) obtained crystals. (Reprinted with permission from Ref. [47], copyright 2018, Elsevier).
Materials 16 00269 g001
Figure 2. XRPD patterns of CNGG:RE and CLNGG:RE (RE = Sm, Dy, Pr) grown crystals. (Reprinted with permission from Ref. [47], copyright 2018, Elsevier).
Figure 2. XRPD patterns of CNGG:RE and CLNGG:RE (RE = Sm, Dy, Pr) grown crystals. (Reprinted with permission from Ref. [47], copyright 2018, Elsevier).
Materials 16 00269 g002
Figure 3. Absorption spectra at room temperature of Sm3+ ions in CNGG:Sm (a) and CLNGG:Sm (b) crystals.
Figure 3. Absorption spectra at room temperature of Sm3+ ions in CNGG:Sm (a) and CLNGG:Sm (b) crystals.
Materials 16 00269 g003
Figure 4. Excitation (a) and emission (b) spectra of the hosts, and excitation spectra of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals (c).
Figure 4. Excitation (a) and emission (b) spectra of the hosts, and excitation spectra of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals (c).
Materials 16 00269 g004
Figure 5. Emission spectra of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals under 270 nm (a) and 405 nm (b) excitation wavelengths.
Figure 5. Emission spectra of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals under 270 nm (a) and 405 nm (b) excitation wavelengths.
Materials 16 00269 g005
Figure 6. Emission kinetics of the 4G5/2 level of Sm3+ ions in CNGG:Sm (a) and CLNGG:Sm (b), and the experimental transfer function (c).
Figure 6. Emission kinetics of the 4G5/2 level of Sm3+ ions in CNGG:Sm (a) and CLNGG:Sm (b), and the experimental transfer function (c).
Materials 16 00269 g006
Figure 7. Absorption (a) and excitation (b) spectra of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals. (Reprinted with permission from Ref. [23], copyright 2018, Elsevier).
Figure 7. Absorption (a) and excitation (b) spectra of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals. (Reprinted with permission from Ref. [23], copyright 2018, Elsevier).
Materials 16 00269 g007
Figure 8. Emission spectra of Dy3+ ions in CNGG:Dy (a) and CLNGG:Dy (b) crystals at room temperature. (Reprinted with permission from Ref. [23], copyright 2018, Elsevier).
Figure 8. Emission spectra of Dy3+ ions in CNGG:Dy (a) and CLNGG:Dy (b) crystals at room temperature. (Reprinted with permission from Ref. [23], copyright 2018, Elsevier).
Materials 16 00269 g008
Figure 9. Emission kinetics of the 4F9/2 level of Dy3+ ions in CNGG:Dy (a) and CLNGG:Dy (b) samples. (Reprinted with permission from Ref. [23], copyright 2018, Elsevier).
Figure 9. Emission kinetics of the 4F9/2 level of Dy3+ ions in CNGG:Dy (a) and CLNGG:Dy (b) samples. (Reprinted with permission from Ref. [23], copyright 2018, Elsevier).
Materials 16 00269 g009
Figure 10. Absorption spectra of Pr3+ ions in CNGG:Pr (a) and CLNGG:Pr (b) crystals. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Figure 10. Absorption spectra of Pr3+ ions in CNGG:Pr (a) and CLNGG:Pr (b) crystals. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Materials 16 00269 g010
Figure 11. Excitation spectra of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Figure 11. Excitation spectra of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Materials 16 00269 g011
Figure 12. Emission spectra of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Figure 12. Emission spectra of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Materials 16 00269 g012
Figure 13. Emission kinetics of the 1D2 level of Pr3+ ions in CNGG:Pr (a) and CLNGG:Pr (b) samples. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Figure 13. Emission kinetics of the 1D2 level of Pr3+ ions in CNGG:Pr (a) and CLNGG:Pr (b) samples. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Materials 16 00269 g013
Table 1. Elemental compositions and lattice parameters of CNGG:RE and CLNGG:RE (RE = Sm, Dy, Pr) grown crystals. (Reprinted with permission from Ref. [47], copyright 2018, Elsevier).
Table 1. Elemental compositions and lattice parameters of CNGG:RE and CLNGG:RE (RE = Sm, Dy, Pr) grown crystals. (Reprinted with permission from Ref. [47], copyright 2018, Elsevier).
CrystalElemental CompositionLattice Parameters (Å)
CNGGCa3Nb1.6875Ga3.1875O1212.508(2) [17]
CNGG:Sm(Ca0.966Sm0.034)3Nb1.688Ga3.187O1212.502(5)
CNGG:Dy(Ca0.958Dy0.042)3Nb1.686Ga3.189O1212.493(1)
CNGG:Pr(Ca0.982Pr0.018)3Nb1.684Ga3.191O1212.525(1)
CLNGGCa3Li0.275Nb1.775Ga2.95O1212.511(0) [17]
CLNGG:Sm(Ca0.966Sm0.034)3Li0.279Nb1.777Ga2.944O1212.506(4)
CLNGG:Dy(Ca0.958Dy0.042)3Li0.277Nb1.776Ga2.947O1212.501(0)
CLNGG:Pr(Ca0.982Pr0.018)3Li0.276Nb1.775Ga2.949O1212.517(1)
Table 2. JO intensity parameters for Sm3+ in different oxide crystals with partially disordered structure (Reprinted with permission from Ref. [21], copyright 2017, Elsevier).
Table 2. JO intensity parameters for Sm3+ in different oxide crystals with partially disordered structure (Reprinted with permission from Ref. [21], copyright 2017, Elsevier).
Crystal2 × 10−20 (cm2)4 × 10−20 (cm2)6 × 10−20 (cm2)
YAG:Sm [57]1.22.081.72
CNGG:Sm3.063.892.6
CLNGG:Sm4.194.42.49
CaNb2O6:Sm [58]6.336.493.72
LiNbO3:Sm [59]2.114.51.45
LiNbO3:Mg:Sm [60]1.862.821.19
LiNbO3:Zn:Sm [60]1.682.721.12
Table 3. Spontaneous emission probabilities (AJJ’), branching ratios (βJJ’), and radiative lifetimes (τr) of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals. (Reprinted with permission from Ref. [21], copyright 2017, Elsevier).
Table 3. Spontaneous emission probabilities (AJJ’), branching ratios (βJJ’), and radiative lifetimes (τr) of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals. (Reprinted with permission from Ref. [21], copyright 2017, Elsevier).
Transition from
4G5/2
CNGG:SmCLNGG:Sm
λ ¯ (nm) A E D
(S−1)
A M D
(S−1)
βJJ’
(%)
τr(ms) λ ¯
(nm)
A E D
(S−1)
A M D
(S−1)
βJJ’(%)τr(ms)
6F11/213590.58 0.0009 13610.56 0.0007
6F9/211463.21 0.005 11474.22 0.006
6F7/210165.242.150.01 10165.922.140.011
6F5/293623.66.80.04 93625.846.840.053
6H15/28800.86 0.02 8790.8 0.02
6F3/28784.4510.760.0004 8785.910.70.001
6F1/28694.05 0.006 8695.4 0.007
6H13/278210.7 0.016 77111 0.015
6H11/270852.2 0.08 70856.2 0.08
6H9/2647205 0.32 647215.8 0.32
6H7/260023325.20.41 600240.8250.37
6H5/256614.330.20.071.5856617.529.970.061.5
Table 4. Emission cross-sections of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals compared to other similar crystals. (Reprinted with permission from Ref. [21], copyright 2017, Elsevier).
Table 4. Emission cross-sections of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals compared to other similar crystals. (Reprinted with permission from Ref. [21], copyright 2017, Elsevier).
CrystalsTransition from
4G5/2
λ ¯ σem ×10−21 cm2
CNGG:Sm6H5/25670.187
6H7/26150.87
6H9/26620.63
CLNGG:Sm6H5/25670.184
6H7/26151
6H9/26630.79
CaNb2O6:Sm [58]6H7/26102.42
LiNbO3:Sm [59]6H7/26170.82
GdVO4:Sm [64]6H7/26040.9
Table 5. Partial energy levels of Sm3+ ions involved in the main emission lines in CNGG:Sm and CLNGG:Sm crystals.
Table 5. Partial energy levels of Sm3+ ions involved in the main emission lines in CNGG:Sm and CLNGG:Sm crystals.
MultipletsCNGG:Sm
[21]
Bc (cm−1)CLNGG:Sm
[21]
Bc (cm−1)YAG:Sm
[57]
Bc (cm−1)
4G5/2176301785217633178531760117784
179081790917867
188461885017883
6H9/2254324042555241926182461
249125252568
240824172466
233923522401
223822442254
6H7/2136212141368121914191261
128912991371
119011911242
101610191012
6H5/2228109233113251130
98106141
000
Table 6. Emission kinetic parameters of the 4G5/2 level of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals.
Table 6. Emission kinetic parameters of the 4G5/2 level of Sm3+ ions in CNGG:Sm and CLNGG:Sm crystals.
IH FittingCNGG:Sm
[21]
CLNGG:Sm
[21]
YAP:Sm
[65]
KZnLa(PO)4:Sm
[66]
LGSO:Sm
[67]
S66666
R0 (Å)6.596.867.2210.967.9
CDA (m6s−1)5.31 × 10−537.01 × 10−535.9 × 10−534.04 × 10−523.64 × 10−52
WDA (s−1)525691416.5233.081497.4
WET (s−1)325358510865450
τr (ms)1.581.52.44.31.89
τ0 (ms)(0.1 at.%)1.4831.44
τav (ms)10.952.140.9111.74
ɳ (%)6766892192
Table 7. Intensity parameters Ωt (t = 2, 4, 6) and other spectroscopic parameters of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals.
Table 7. Intensity parameters Ωt (t = 2, 4, 6) and other spectroscopic parameters of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals.
CNGG:Dy [23]CLNGG:Dy [23]
Transition fromLevel λ ¯ (nm) n A E D
(S−1)
A M D
(S−1)
βJJ’(%) λ ¯ (nm) n A E D
(S−1)
A M D
(S−1)
βJJ’(%)
4F9/26F1/213351.90390.14 013151.89970.15 0
6F3/21280.51.90540.1064 012801.90070.1054 0
6F5/21152.81.910111.2 0.006111601.905010.78 0.006
6F7/2999.81.91836.89.840.009999.21.91366.759.780.0092
6H5/2930.91.92355.3 0.0029931.71.91875.27 0.0029
6F9/2849.41.931512.49.660.012849.61.926712.339.580.012
6H7/2829.61.933812.65.990.01829.11.933712.646.0080.0104
6F11/2765.11.942732.390.80.06764.61.938132.1589.140.0678
6H9/2747.21.945726.45.520.01747.21.941026.185.490.0177
6H11/2661.51.9637107.821.80.07660.91.9591107.221.730.072
6H13/2574.91.99211131.8 0.628574.81.98741122.7 0.627
6H15/2480.82.0468311.2 0.17480.82.0410308.57 0.172
AT(4F9/2) = 1801.6 s−1, τr = 0.55 msAT(4F9/2) = 1786.5 s−1, τr = 0.559 ms
t2= 5.04 × 10−20 cm2
4= 1.81 × 10−20 cm2
6 = 1.53 × 10−20 cm2
2= 5.29 × 10−20 cm2
4= 1.49 × 10−20 cm2
6 = 1.37 × 10−20 cm2
χ1.181.08
Table 8. Emission cross sections (σem), branching ratios JJ’), and Y/B (yellow/blue) ratio of the yellow transition of Dy3+ ions in CNGG:Dy and CLNGG:Dy compared with other materials. (Reprinted with permission from Ref. [23], copyright 2018, Elsevier).
Table 8. Emission cross sections (σem), branching ratios JJ’), and Y/B (yellow/blue) ratio of the yellow transition of Dy3+ ions in CNGG:Dy and CLNGG:Dy compared with other materials. (Reprinted with permission from Ref. [23], copyright 2018, Elsevier).
Crystals4F9/26H13/2
σem (×10−20 cm2)
Y/Bßcalc(%)τrad(ms)τexp(ms)
CNGG:Dy [23]0.3372.9562.80.550.438
CLNGG:Dy [23]0.338362.70.5590.332
YAG:Dy [69]0.31.350.960.90.669
Gd2SiO5:Dy [70]0.681.6654.350.5420.497
GdVO4:Dy [71]0.9468.30.2640.084
GGG:Dy [72]0.261.32491.10.79
Table 9. Partial energy levels of Dy3+ ions involved in 4F9/26H13/2 yellow laser transition in the CNGG:Dy and CLNGG:Dy crystals.
Table 9. Partial energy levels of Dy3+ ions involved in 4F9/26H13/2 yellow laser transition in the CNGG:Dy and CLNGG:Dy crystals.
Multiple’sCNGG:Dy
[23]
Bc (cm−1)CLNGG:Dy
[23]
Bc (cm−1)YAG:Dy
[73]
Bc (cm−1)
4F9/221,54721,16921,55421,17221,55121,096
21,21021,21021,119
21,13621,14021,056
21,00121,00120,896
20,95020,95420,859
6H13/2396037763960377739483731
393639383821
384538473776
377437763720
369736953699
364336473590
357535773564
Table 10. Emission kinetic parameters of the 4F9/2 manifold of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals.
Table 10. Emission kinetic parameters of the 4F9/2 manifold of Dy3+ ions in CNGG:Dy and CLNGG:Dy crystals.
IH FittingCNGG:Dy
[23]
CLNGG:Dy
[23]
Bi4Si3O12:Dy
[57]
LSO:Dy
[58]
Lu2SiO5:Dy
[59]
s66666
R0 (Å)11.812.4139.158.908.9
CDA (m6s−1)0.512 × 10−520.908 × 10−520.65 × 10−517.8 × 10−525.3 × 10−52
WDA (s−1)184324811107.61569.41066
WET (s−1)681681856338390
τr (ms)0.5500.5591.6840.6350.635
τ0 (ms)(0.1 at.%)0.5430.405
τav (ms)0.2290.1430.6900.2020.509
ɳ (%)7372413280
Table 11. Spontaneous emission probabilities (AJJ’), branching ratios (βJJ’), and radiative lifetimes (τr) of the 3P0 and 1D2 excited levels of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals.
Table 11. Spontaneous emission probabilities (AJJ’), branching ratios (βJJ’), and radiative lifetimes (τr) of the 3P0 and 1D2 excited levels of Pr3+ ions in CNGG:Pr and CLNGG:Pr crystals.
MultipletsCNGG:Pr [24]CLNGG:Pr [24]
ν (cm−1)AEDAMDβν (cm−1)AEDAMDβ
3P01D237219.31 0.0002371611.63 0.0002
1G410,680909.7 0.0210,5721086 0.02
3F413,4753503 0.0713,4064457 0.08
3F313,9700 013,9620 0
3F215,22914,644 0.3215,22518,298 0.33
3H616,0461495 0.03216,1421320 0.02
3H517,9990 018,0240 0
3H420,29823,725 0.5220,30329,254 0.53
Ʃ A = 44,286 s−1, τr = 22 μsƩ A = 54,426.63 s−1, τr = 18.3 μs
1D21G46959429 0.096856501.6 0.09
3F497541998.4 0.45969024,44.6 0.43
3F310,249187.711.20.0410,246233140.04
3F211,508574.65.40.1211,509709.36.670.13
3H612,325514 0.1112,426644 0.11
3H514,27826 0.0514,308319 0.05
3H416,577714 0.1516,587751 0.13
Ʃ A= 4460 s−1, τr = 224 μsƩ A= 5623.2 s−1, τr = 177.8 μs
t mod2 = 3.607 × 10−20 cm2
4 = 3.580 × 10−20 cm2
6 = 1.264 × 10−20 cm2
rms = 0.086 × 10−20 cm2
2 = 4.563 × 10−20 cm2
4 = 4.451 × 10−20 cm2
6 = 1.105 × 10−20 cm2
rms = 0.12 × 10−20 cm2
Table 12. Calculated emission cross-sections (σem) of Pr3+ ions in CNGG:Pr and CLNGG:Pr in comparison with other Pr-doped crystals. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Table 12. Calculated emission cross-sections (σem) of Pr3+ ions in CNGG:Pr and CLNGG:Pr in comparison with other Pr-doped crystals. (Reprinted with permission from Ref. [24], copyright 2019, Elsevier).
Crystal3P03H4
σem (cm2)
1D23H4
σem (cm2)
3P03F2
σem (cm2)
CNGG:Pr5.72 × 10−200.48 × 10−201.9 × 10−19
CLNGG:Pr7.48 × 10−200.51 × 10−202.96 × 10−19
CaSGG:Pr [81]7.3 × 10−20-1.02 × 10−19
SrLaGa3O7:Pr [82]3.43 × 10−20-0.44 × 10−19
GGG:Pr [83]--1.24 × 10−19
Table 13. Energy levels of Pr3+ ions involved in the 3P03H4, 1D23H4, and 3P03F2 emission transitions in CNGG:Pr and CLNGG:Pr crystals.
Table 13. Energy levels of Pr3+ ions involved in the 3P03H4, 1D23H4, and 3P03F2 emission transitions in CNGG:Pr and CLNGG:Pr crystals.
Pr3+ MultipletsCNGG:Pr [24]Bc (cm−1)CLNGG:Pr
[24]
Bc(cm−1)YAG:Pr
[84]
Bc(cm−1)
3P020,60320,60320,60120,60120,5340
1D217,235
17,105
16,974
16,590
16,512
16,88317,227
17,111
16,978
16,575
16,534
16,88517,210
17,088
16,881
16,409
16,400
16,798
3F25464
5409
5368
5335
5303
53755479
5410
5352
5332
5311
53766994
6973
6943
6876
6857
6929
3H4-
623
554
517
-
-
66
21
0
--
625
555
513
-
-
74
22
0
--
742
576
533
-
-
50
19
0
-
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gheorghe, C.; Hau, S.; Gheorghe, L.; Voicu, F.; Greculeasa, M.; Broasca, A.; Stanciu, G. RE3+-Doped Ca3(Nb,Ga)5O12 and Ca3(Li,Nb,Ga)5O12 Crystals (RE = Sm, Dy, and Pr): A Review of Current Achievements. Materials 2023, 16, 269. https://doi.org/10.3390/ma16010269

AMA Style

Gheorghe C, Hau S, Gheorghe L, Voicu F, Greculeasa M, Broasca A, Stanciu G. RE3+-Doped Ca3(Nb,Ga)5O12 and Ca3(Li,Nb,Ga)5O12 Crystals (RE = Sm, Dy, and Pr): A Review of Current Achievements. Materials. 2023; 16(1):269. https://doi.org/10.3390/ma16010269

Chicago/Turabian Style

Gheorghe, Cristina, Stefania Hau, Lucian Gheorghe, Flavius Voicu, Madalin Greculeasa, Alin Broasca, and George Stanciu. 2023. "RE3+-Doped Ca3(Nb,Ga)5O12 and Ca3(Li,Nb,Ga)5O12 Crystals (RE = Sm, Dy, and Pr): A Review of Current Achievements" Materials 16, no. 1: 269. https://doi.org/10.3390/ma16010269

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop