Next Article in Journal
Room Temperature Solid State Synthesis, Characterization, and Application of a Zinc Complex with Pyromellitic Acid
Next Article in Special Issue
Enhancing Light Extraction of Inorganic Scintillators Using Photonic Crystals
Previous Article in Journal / Special Issue
Super-Lattice Structure and Phase Evolution of Pb(Lu0.5Nb0.5)O3-PbTiO3 Single Crystal with Low PbTiO3
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Rare-Earth Tantalates and Niobates Single Crystals: Promising Scintillators and Laser Materials

1
The Key Laboratory of Photonic Devices and Materials, Anhui Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Hefei 230031, China
2
Science Island Branch of Graduate School, University of Science and Technology of China, Hefei 230026, China
*
Author to whom correspondence should be addressed.
Crystals 2018, 8(2), 55; https://doi.org/10.3390/cryst8020055
Submission received: 11 December 2017 / Revised: 26 December 2017 / Accepted: 28 December 2017 / Published: 24 January 2018
(This article belongs to the Special Issue Crystal Growth for Optoelectronic and Piezoelectric Applications)

Abstract

:
Rare-earth tantalates, with high density and monoclinic structure, and niobates with monoclinic structure have been paid great attention as potential optical materials. In the last decade, we focused on the crystal growth technology of rare-earth tantalates and niobates and studied their luminescence and physical properties. A series of rare-earth tantalates and niobates crystals have been grown by the Czochralski method successfully. In this work, we summarize the research results on the crystal growth, scintillation, and laser properties of them, including the absorption and emission spectra, spectral parameters, energy levels structure, and so on. Most of the tantalates and niobates exhibit excellent luminescent properties, rich physical properties, and good chemical stability, indicating that they are potential outstanding scintillators and laser materials.

1. Introduction

In recent decades, rare-earth tantalates and niobates have been paid great attention due to their rich physical properties and good chemical stability [1,2,3,4]. Commonly, chemical forms of the rare-earth tantalates and niobates are written as REBO4, RE3BO7, REB3O9, REB5O14, and REB7O19, where RE means rare earth ions, and B are Ta or Nb elements [5,6]. Orthotantalates and orthoniobates REBO4 belong to the fergusonite structure and have been reported extensively [7,8,9]. By comparison and analysis, REBO4 exhibited excellent luminescent properties [10]. Moreover, RE ions can be substituted easily by other trivalent ions to realize characteristic fluorescence emission, from the ultraviolet to the infrared region [11]. With the change of RE ions radius, there are three different structures. With the larger ionic radii (La-Pr), they crystallize in structure P21/c ( C 2 h 5 , #14, Z = 4), while with intermediate ionic radii (Nd-Tb) they exhibite the fergusonite M-type structure I2/a ( C 2 h 6 , #15, Z = 4). Furthermore, with the smaller ionic radii (Dy-Lu) they present the fergusonite M’-type structure P2/a ( C 2 h 4 , #13, Z = 2) [11,12,13].
At present, with an increasing demand in dense, fast, and bright scintillators for high-energy physics, medical diagnostics, and security screening devices, the scintillation characteristics of the lanthanide orthotantalates and orthoniobates REBO4 have been widely investigated for the development of new scintillator materials [9,14,15]. Among REBO4 materials, RETaO4 are favorable for the registration of high-energy particles due to their extremely high density ranging from 7.8 g/cm3 for LaTaO4 to 9.75 g/cm3 for LuTaO4, which are partly higher than the density of Bi4Ge3O12 (BGO 7.1 g/cm3) and PbWO4 (PWO 8.28 g/cm3). Normally, the rare-earth tantalates are studied systematically with polycrystalline samples produced by solid-state reaction [9]. Although the optical absorption and luminescence of RETaO4 crystals are described, the crystal size is too small, only 7 mm × 7 mm × 1 mm [10]. It is urgent to develop the growth technology of large RETaO4 crystals for the device application. Recently, our group has made great progress in the Cz growth of large-size tantalite crystals and obtained big and fine GdTaO4 single crystals, which would promote the growth and practical application of RETaO4 [16].
In addition, REBO4 have been considered appropriately as host matrices and have also attracted wide interests as potential rare earth-doped laser hosts [17]. They belong to monoclinic system and have large crystal field energy. The active ions occupy the site with C2 symmetry (RE sites), which is advantageous to remove the parity-forbidden rule of f-f transition. Therefore, it is beneficial for the realization of new emission and tunable wavelength. Additionally, their mechanical and thermal properties are also sufficient for laser applications.
Recently, on the one hand, our group have finished a lot of research works on scintillation characteristics of RETaO4, such as GdTaO4, Tb:GdTaO4, LuTaO4, Nd:LuTaO4, and so on. The luminescence results indicate that GdTaO4 and Nd:LuTaO4 are the promising scintillator candidates. On the other hand, the emission spectra of active ions (Yb, Nd, Er, Ho) doped GdBO4, YBO4, and mixed GdYBO4 crystals show that they can be used as the new materials for visible, near, and middle infrared lasers. Meanwhile, near infrared lasers are realized on the Yb and Nd-doped ReBO4 successfully.

2. Experiments

2.1. Polycrystalline Synthesis

In this paper, LuTaO4 and Nd-doped LuTaO4 series samples were prepared by solid-state reaction method using Lu2O3 (99.99%), Ta2O5 (99.99%), and Nd2O3 (99.999%) as starting materials. First, the raw materials were weighed accurately according to the appropriate stoichiometric ratio and then mixed and ground in a mortar. The mixed powders were then heated to 1500 °C at a heat rate of approximately 2.3 °C/min and calcined at 1500 °C for 24 h. Finally, the calcined samples were slowly cooled to room temperature and then carefully ground for the measurements.

2.2. Single Crystal Growth

All of the crystals, mentioned in the text, were grown by the Czochralski (Cz) method. Raw materials, with high purity (>4 N), were weighed stoichiometrically, mixed thoroughly, and then pressed into disks. The disks were put into an iridium crucible and melted in a JGD-60 furnace with an automatic diameter controlled system. Using a seed, the crystals were grown in nitrogen atmosphere with a rotation speed of 3.0–10.0 rpm and pulling rate of 0.35–1 mm/h. The growth process of all crystals is similar. The as-grown crystals and some samples are shown in Figure 1.

2.3. Crystal Structure and Chemical Etching

Generally, crystal structure can be determined by X-ray diffraction (XRD). The XRD patterns of as-grown crystals are measured with a X’pert PRO X-ray diffractometer (PANalytical, Almelo, The Netherlands) employing Cu Kα radiation (λ = 1.540598 Å). The diffraction data are recorded in the 2θ range of 10°–90° with a scan step of 0.033°. The structural parameters of above crystals are obtained by fitting the XRD data with the Rietveld refinement method. The structural parameters of pure GdTaO4 or GdNbO4 are taken as the initial values, the background function, lattice parameters, atomic coordinates, and isotropic temperature factors are refined with the software GSAS. The refinement results of cell parameters are shown in Table 1. Since the ionic radius of doped ions is smaller than that of the replaced ions, there is a slight decrease in the cell volume and density.
It is very meaningful to research the crystal defects in the as-grown crystal. The crystal with defects may destroy the mechanical, optical, and laser properties which will restrict the use of the crystals. Generally, the chemical etching method is an important and direct technique to investigate the defect structure of a single crystal. As we all know, acid and alkali are often used as the etchant to dissolve the dislocation sites.
The crystal of Tm,Ho:GdYTaO4 and Nd:GdYNbO4 are etched with KOH. Their dislocation etching pit patterns on the (100), (010), and (001) crystallographic faces are shown in Figure 2 and Figure 3. Figure 2a,c exhibit the similar triangular prism, different from Figure 2b: rhombohedral. Similarly, the shape of the etching pits of Nd:GdYNbO4 on (100) and (001) crystallographic faces are both present in strip shape, but not completely equivalent, as shown in Figure 3. Dislocations are typical defects introduced by the lattice distortion due to the incorporation of impurities and usually found in the molecules or atoms with the weak chemical bond. In the process of chemical etching, the weakest and unstable bonds are broken and form the specific etching pits. Therefore, dislocation etching pits have close relations with the atomic arrangement and symmetry.

2.4. Measurements

All the absorption spectra were recorded by a VIS-NIR-IR spectrophotometer PE lambda 950 (PerkinElmer, Waltham, MA, USA). The excitation and emission spectra were measured by a fluorescence spectrometer FLSP920 (Edinburgh, UK), with an excitation source of Xenon lamp or diode laser. An Opolette 355I (Carlsbad, CA, USA) was used to measure the fluorescence decay curves. Thermal expansion behavior of crystals along a, b, and c axes were measured in the temperature range of 300–893 K using a thermal dilatometer DIL-402C (Netzsch, Selbe, Germany) with a heating rate of 5 K/min. The samples coated with graphite were used for the measurements of specific heat and thermal diffusivity along the a, b, and c axes by a laser flash apparatus LFA457 (Netzsch, Selbe, Germany)

2.5. Judd–Ofelt Calculation

The Judd–Ofelt (J–O) theory [18,19] has been successfully applied to various systems doped with rare-earth ions. The great advantage of the J–O theory is the ability to express the probability of any transition between two fn states with only three intensity parameters (Ω2, Ω4, Ω6) which can be obtained from absorption spectra. Moreover, spectral parameters, such as line strength (s), oscillator strength (f), transition probability (A), and radiative lifetime (τ), can be calculated with absorption spectrum. The experimental oscillator strength from the initial state | S L J to the final state | S L J was calculated as follows:
f = 4 m ε 0 c 2 N e 2 λ ¯ 2 α ( λ ) d λ
where α (λ) is absorption coefficient, N is the total number of active ions per unit volume, m is the mass of the electron, ε0 is permittivity of vacuum, c is the velocity of light in vacuum, e is the electron charge, and λ is given as Equation (2):
λ ¯ = λ α ( λ ) d λ α ( λ ) d λ
However, under certain simplifying assumptions, the oscillator strength (f) can be expressed as the sum of electric-dipole (fed) and magnetic-dipole (fmd) oscillator strength:
f = f e d + f m d
Additionally, the relationships of oscillator strength and line strength are shown as follow:
S e d ( S L J S L J ) = 9 n ( n 2 + 2 ) 2 3 h ( 2 J + 1 ) λ ¯ 8 π 2 m c f e d
S m d ( S L J S L J ) = 3 h ( 2 J + 1 ) λ ¯ 8 π 2 m c n f m d
where J is the total angular momentum, n is the refractive index, and h is Planck constant. Sed and Smd can be calculated easily with Equation (6):
S e d = t = 2 , 4 , 6 Ω t | S L J | | U t | | S L J | 2
S m d = ( h 4 π m c ) 2 | S L J | | L + 2 S | | S L J | 2
In the above formulas, | S L J | | U t | | S L J | is the doubly reduced matrix elements of unitary tensor operator Ut with t = 2, 4, 6 between the state | S L J and | S L J , L + 2S is the magnetic dipole operator.
In addition, spontaneous transition probability (A), and radiative lifetime (τ) can be calculated with line strength:
A ( S L J S L J ) = [ 16 π 3 e 2 3 h ε 0 ( J + 1 ) λ ¯ 3 ] × [ n ( n 2 + 2 ) 2 9 S e d + n 3 S m d ]
τ = 1 J A ( S L J S L J )

2.6. Emission Cross-Section Calculation

In this paper, all the emission cross-sections σem are calculated with the F–L formula:
σ e m ( λ ) = λ 5 β I ( λ ) 8 π n 2 c τ λ I ( λ ) d λ
where c is the speed of light, τ is the radiative lifetime of the upper energy level, n is the refractive index, and β is the branching ratio.

3. Scintillator Materials

3.1. GdTaO4 and Nd:GdTaO4 Crystals

In recent years there has been renewed interest in developing new scintillator materials characterized by high light yield, fast response, and high density [20,21,22,23]. Scintillators with high density and high atomic number are mostly desirable, because high stopping power can reduce the needed amount of scintillator materials and, thus, reduce the volume of the detector. GdTaO4 is an attractive host and exhibits a high density (8.94 g/cm3). Attenuation length for GdTaO4 is calculated to be 1 cm, only second to PWO (0.89 cm) [23,24]. Previously, GdTaO4 crystal has been reported in a few papers [10,25,26], where the grown crystals are either with inclusions and twins [10] or with small size [26].
In our lab, nearly ten years have been spent on the Cz growth of GdTaO4 single crystals. At first, an iridium wire is used to pull a crystal from the melt for obtaining the seed. With the seed, GdTaO4 and Tb:GdTaO4 bulk single crystals are successfully grown by the Cz method [27]. However, there are cracks, inclusions, and twins within these two crystals. After a long period of technological optimization, finally, a crack-free GdTaO4 crystal with dimensions of Φ 23 mm × 30 mm was grown successfully, which is the largest size so far [16].
The luminescence and scintillation properties of GdTaO4 crystal have been studied in detail. Under the excitation of 273 nm, the GdTaO4 crystal shows a strong emission band from 400 to 700 nm, with a highly asymmetrical shape [16]. To analyze the asymmetrical emission band further, a series of temperature-dependent luminescent spectra of GdTaO4 crystal are measured. All of the emission spectra consist of two bands, 2.2 eV and 2.7 eV bands from 8 K to 300 K [28]. The emission intensity increases slightly as the temperature rising from 8 K to 80 K, and then decrease with increasing temperature. It is quenched rapidly above 150 K with the intensity decreasing by two orders of magnitude. The intensity variation of these two bands indicates the existence of thermal activation process. The activation energy of 2.2 eV and 2.7 eV bands is determined to be 156 meV and 175 meV, respectively. These two bands originate from different luminescent centers, which are tentatively assigned to self-trapped excitons (STE) localized at TaO43− groups (2.7 eV) and to relaxed excitons related to lattice imperfections (2.2 eV) [28]. The photoluminescence decay shows two components, including a fast one of 30 ns with 53% and a relatively slow one of 452 ns with 47%. The scintillation decay consists of a fast component of 72.6 ns (9.5%) and a slow component of 1236.2 ns (90.5%). Meanwhile, the scintillation efficiency of GdTaO4 is about four times as much as PWO by integrating the area of the radioluminescence spectra. The relative light yield of GdTaO4 is calibrated as 19 p.e./mes as that of PWO. Although, the scintillation decay of GTO is inferior to PWO, the latter presents a dominant decay of a dozen nanoseconds [24], the light yield of GdTaO4 is higher than PWO.
In addition, the Nd:GdTaO4 single crystal with density of 8.83 g/cm3 has been grown by Fang Peng et al. [29]. Its photoluminescence decay time of 417 nm from the 4f-4f transition of Nd3+ is 463 ns, and scintillation decay constants consist of 46.4 ns (48%) and 1199.7 ns (52%) under the excitation of 354 nm. The scintillation decay of Nd:GdTaO4 is much faster than that of GdTaO4, and the faster component percent is remarkably increased. This result may be caused by defects induced by Nd3+ doping. Moreover, because the concentration of Nd3+ is 0.67%, a rather low value, the light yield of Nd:GdTaO4 can be estimated to be equal to that of GdTaO4. However, the faster decay of Nd:GdTaO4 is encouraging and makes it more effective in detecting high energy rays or particles than GdTaO4.

3.2. LuTaO4

Lutetium tantalate (LuTaO4) is an efficient luminescent host material, especially excited by occurs by ionizing radiation [30,31]. LuTaO4 also exhibits extremely high density (9.81 g/cm3), which is the highest among the present luminescent host materials. Therefore, LuTaO4 may be an excellent heavy scintillator when it is doped with appropriate active ions, such as Nd3+. M’-type Lu1−xNdxTaO4 (x, 0.01–0.1) polycrystalline powders were synthesized by solid reaction method [32].
The emission at 418.5 nm corresponding to the 4D3/24I13/2 transition of Nd3+ is strongest, and the fluorescence lifetime of 418.5 nm emission is measured to be approximately 263.2 ns, which is faster than that of BGO (300 ns) [32]. Meanwhile, LuTaO4 can also emit directly. Therefore, LuTaO4 can be expected as a very promising heavy substrate and heavy scintillator, which has potential applications in nuclear medicine and high energy detection. Thus, it is significant to explore the growth of LuTaO4 and Nd:LuTaO4 crystals.
Liu et al. [31] have spent long time on the LuTaO4 crystal growth. However, unfortunately, it is hard to obtain sing crystal. According to the XRD analysis, there are Lu3TaO7, M’-LuTaO4, M-LuTaO4 phases in the crystal pulled from LuTaO4 melt, which indicates that the phase transition of the system Lu2O3–Ta2O5 is different from Gd2O3–Ta2O5. For designing or improving the single-crystal growth or ceramic preparation technique of LuTaO4, Xing et al. [33] investigated the detailed phase relations of the Lu2O3–Ta2O5 system. The compounds containing 25–60 mol% Ta2O5 are prepared by solid-state reaction at sintering temperature from 1350 °C to 2058 °C. The sintered compound phases are studied by XRD in details. Cubic Lu3TaO7, M’-LuTaO4, M-LuTaO4, O-Ta2O5, and T-Ta2O5 are observed. With the temperature increases, there is an irreversible phase transition from M’ to M-LuTaO4 near 1770 °C in the composition of 30–52 mol% Ta2O5, and another phase transition from T-Ta2O5 to O-Ta2O5 at about 1685 °C when the ratio of Ta2O5 is >52 and ≤60 mol%. Finally, a phase diagram of the Lu2O3–Ta2O5 system in the range 0–100 mol% Ta2O5 is constructed, as shown in Figure 4. These results are helpful to explain the phase transition of Lu2O3–Ta2O5 system and to design the preparation technique of LuTaO4 single crystals or ceramic scintillators.

4. Laser Materials

In the following parts, we will briefly describe the spectral and laser properties of some active ion-doped othotantalates and orthoniobates.

4.1. Er:GdTaO4 Crystal

Trivalent erbium ion (Er3+) has attracted wide attention due to its rich laser emission bands, such as, green laser (2H11/2,4S3/2)→4I15/2; red laser 4F9/24I15/2; 1.5–1.6 μm laser 4I13/24I15/2; 2.6–3 μm laser (4I11/24I13/2). These lasers can be applied in many different fields, including atmospheric monitoring, eye-safe laser, and medical treatment. Previously, 30 at% and 1 at% Er:GdTaO4 crystals are grown by the Cz method [34,35]. According to the absorption spectra, the transition intensity parameters Ωt (t = 2, 4, 6) are calculated by J–O theory and compared with other Er3+-doped crystals, as shown in Table 2. Ω2 is sensitive to the symmetry between the rare-earth ions and the ligand field. The spectroscopic quality factor Ω46 of 1 at% Er:GdTaO4 crystal is found to be 1.42, which is used to estimate the potential of active materials for laser operation when is linked to the luminescence branching ratios. The value of 1 at% Er:GdTaO4 is comparable with those in other Er3+-doped systems and larger than that of 30 at% Er:GdTaO4 crystal. High Er3+ doping concentration is proposed to overcome the self-terminating “bottleneck” effect by inducing upconversion (UC) and cross-relaxation (CR) processes [36]. However, the fluorescence intensity of 2H11/2, 4F9/2, and 4S3/2 states is quenched at high Er3+ concentration. The UC process depopulates the pumping level and the upper level of 1.6 μm laser and reabsorption in this wavelength increase with the increasing Er-concentration. Emission cross-section is an important parameters for evaluating the laser property of the materials, and larger cross-sections means easier laser realization. The largest emission cross-section of 30 at% Er:GdTaO4 crystal is 0.655 × 10−20 cm2 at 2.631 μm, which indicates that 30 at% Er:GdTaO4 crystal can be a promising laser medium around 2.6 μm [34]. The largest emission cross-section of 1 at% Er:GdTaO4 crystal is 1.022 × 10−20 cm2 around 1.6 μm, which make it e great potential for near infrared laser generation. Furthermore, 1 at% Er:GdTaO4 crystal also can be used as the green and red laser materials under some special environment such as at low temperature and this will be an important issue for our future research.

4.2. Nd:GdTaO4 and Nd:GdYTaO4 Crystals

Nd:GdTaO4 can be used not only as a heavy scintillator, but also as a new laser material. Nd:GdTaO4 and Nd:Gd0.69Y0.3TaO4 (Nd:GdYTaO4) (1 at%) single crystals with high optical quality are grown successfully [40,41] and their luminescence and laser properties in near infrared wavelength are studied. The absorption cross-section of the Nd:GdTaO4 crystal at 808 nm is 5.437 × 10−20 cm2, and the full width at half maximum (FWHM) of this absorption band is about 6 nm. The stimulated emission cross-section at 1066 nm is 3.9 × 10−19 cm2 and the measured lifetime of 4F3/2 level is 178.4 μs. A diode end-pumped Nd:GdTaO4 laser at 1066 nm with the maximum output power of 2.5 W is achieved in the continuous-wave mode. The optical-to-optical conversion efficiency and slope efficiency are 34.6% and 36%, respectively. In addition, the fluorescence branching ratio of 4F3/24I9/2 transition reaches 43%, indicating that Nd:GdTaO4 may be an efficient laser medium at 920 nm.
The maximum absorption cross-section of Nd:GdYTaO4 at 809 nm and the stimulated emission cross-section at 1066.6 nm are 6.886 × 10−20 cm2 and 22 × 10−20 cm2, respectively. The fluorescence lifetime is 182.4 μs. An 808 nm laser diode end-pumped continuous wave (CW) laser at 1066.5 nm is realized. The maximum output power of 2.37 W is obtained, corresponding to an optical conversion efficiency of 36.5% and a slope efficiency of 38%. Compared with the slope efficiency 36% of Nd:GdTaO4, Nd:GdYTaO4 shows an enhancement of the CW laser performance. Spectroscopic properties of Nd:GdTaO4 and Nd:GdYTaO4 are compared with other Nd-doped laser crystals, which are listed in Table 3. By comparison, Nd:GdYTaO4 is a better novel laser crystal with low symmetry and has great potential in low to moderate level lasers.

4.3. Yb:GdTaO4 Crystal

With only two multiplets ground state 2F7/2 and excited state 2F5/2, the activator Yb3+ ion has attracted great attention, 1 at% and 5 at% Yb:GdTaO4 are grown by our group. The main absorption peaks of 5 at% Yb:GdTaO4 crystal locate at 930 nm, 956 nm, and 974 nm, and the corresponding absorption cross-sections are calculated to be 0.53 × 10−20 cm2, 0.84 × 10−20 cm2, and 0.65 × 10−20 cm2, respectively. However, 1 at% Yb:GdTaO4 exhibited the bigger absorption cross-section, namely 0.81 × 10−20 cm2, 0.91 × 10−20 cm2, and 1.2 × 10−20 cm2 at 930 nm, 957 nm, and 974 nm, respectively. With the increase of Yb3+ ion concentration, the distance between Yb3+ ions becomes smaller, which causes the cross relaxation. Therefore, the absorption abilities of ground state ions are weaker. Oppositely, the emission cross-sections of 1 at% Yb:GdTaO4 (2.2 × 10−20 cm2 at 1016 nm; 1.36 × 10−20 cm2 at 1035 nm) are bigger than those of 5 at% Yb:GdTaO4 (0.54 × 10−20 cm2 at 1018 nm; 0.5 × 10−20 cm2 at 1036 nm). According to the spectra results, 1 at% Yb:GdTaO4 may be more suitable for laser putout.

4.4. Tm,Ho:GdYTaO4 and Yb,Ho:GdYTaO4 Crystal

The new wavelengths around 2.9 μm have attracted widely interests, due to their strong absorption in water, biological tissues, and vapor [42], which can be applied in medical, biological, and remote sensing. Moreover, laser wavelengths around 2.9 μm are also suitable pump sources for infrared optical parametric oscillation (OPO) or optical parametric generation (OPG) [43]. One possibility of generating 2.9 μm radiation is the 5I65I7 transition of Ho3+, which possesses rich energy levels [44]. Usually, Tm3+ are used to sensitize Ho3+ solid state lasers. Tm,Ho:GdYTaO4 and Yb,Ho:GdYTaO4 single crystals are grown and the spectral properties were studies in detail.
Their absorption spectra along a, b, and c axes are measured and the absorption coefficients are compared in Table 4. Their respective absorption coefficients along the c axis are larger than those along the other two directions. It indicates that the c-axis samples may be more beneficial for the laser performance by improving pumping efficiency.
The 2.9 μm emission spectra of Tm,Ho:GdYTaO4 and Yb,Ho:GdYTaO4 crystals are measured, excited by a 783 nm and 940 nm LD, respectively. Their emission cross-sections are calculated with F–L formula and compared in Figure 5. The main emission peaks of Tm,Ho:GdYTaO4 are located at 2895, 2915 and 2932 nm. Similarly, strong peaks of Yb,Ho:GdYTaO4 are located at 2865 and 2911 nm. The maximum emission cross-section of Tm,Ho:GdYTaO4 at 2933 nm is 37.2 × 10−20 cm2, which is larger than that of Yb,Ho:GdYTaO4 (2911 nm, 17.6 × 10−20 cm2). It indicates that Tm,Ho:GdYTaO4 may be easier to realize laser output than Yb,Ho:GdYTaO4. In addition, the emission spectrum also indicates that the energy transfer between Yb3+–Ho3+ and Tm3+–Ho3+ ions can be realized successfully.
The fluorescence decay times are obtained. Their lifetimes of upper level and low level are shown in Table 5. Compared with other hosts, the Tm,Ho:GdYTaO4 and Yb,Ho:GdYTaO4 exhibit shorter lifetime of 5I7 and longer lifetime of 5I6, which are in favor of the population inversion and laser output. Moreover, the laser threshold of Yb,Ho:GdYTaO4 may be lower than that of Tm,Ho:GdYTaO4.

4.5. Yb:GdNbO4 and Yb:YNbO4

The intensive absorption broadband of 5 at% Yb:GdNbO4 and 5 at% Yb:YNbO4 between 900–1000 nm correspond to the typical transitions from 2F7/2 to the sublevels of 2F5/2 of Yb3+. There are three obvious absorption peaks of 5 at% Yb: GdNbO4, located at 936, 955, and 975 nm, respectively. However, the absorption peaks of 5 at% Yb:YNbO4 are 933, 955, and 974 nm. Their different locations probably are due to the small difference for the radius of Gd3+ (0.938 Å) and Y3+ (0.9 Å). Therefore, the environment of the crystal field around Yb3+ is a little different. The absorption cross-sections of 5 at% Yb:GdNbO4 were calculated to 0.77 × 10−20 cm2, 0.85 × 10−20 cm2, 0.64 × 10−20 cm2, respectively. The absorption cross-section of 5 at% Yb:YNbO4 were calculated to be 0.73 × 10−20 cm2, 1.85 × 10−20 cm2, 0.86 × 10−20 cm2, respectively.
Their refractive indices n were fitted with the following Sellmeier equation by the least square method:
n 2 ( λ , T ) = A ( T ) + B ( T ) λ 2 C ( T ) + D ( T ) λ 2
where T is crystal transmission. And the fitted results of 5 at% Yb:GdNbO4 are A = 5.30671, B = 7711.03444 nm2, C = 255,262.4962 nm2, and D = −9.4997 × 10−8 (nm2)−1. Similarly, A = 4.5440, B = 467,665.596 nm2, C = −77,583.488 nm2, and D = −6.352 × 10−8 (nm2)−1 of 5 at% Yb:YNbO4.
The emission cross-sections of Yb:GdNbO4 and Yb:YNbO4 crystal are calculated, and compared with other hosts, as shown in Table 6. The emission cross-sections of Yb:YNbO4 are larger that of Yb:GdNbO4, and comparable to Yb:YAG. Therefore, it will be a promising near-infrared laser material.
In addition, the preliminary laser experiment of Yb:GdNbO4 is achieved. The pump source is an InGaAs LD with a maximum output powder of 25 W at around 976 nm in continuous mode. With a 3.54% transmission output coupler, the maximum laser power of 270 mW corresponding to the threshold of 6 W is obtained. The slope efficiency is 7.5%. High threshold pump power and low efficiency are mainly due to the crystal reabsorption, poor quality of the as-grown crystal, and crude plane-plane cavity. Good quality crystal and advanced cavity will improve the laser efficiency.

4.6. Nd:GdNbO4, Nd:YNbO4, Nd:GdYNbO4, and Nd:GdLaNbO4 Crystals

With the absorption spectra, the J–O intensity parameters Ω2,4,6, fluorescence branching ratios β(J,J′), radiative times τrad of Nd:GdNbO4 (Nd:GNO), Nd:YNbO4 (Nd:YNO), Nd:GdYNbO4 (Nd:GYNO) and Nd:GdLaNbO4 (Nd:GLNO) crystals are calculated by J–O theory [50,51,52,53]. The comparisons of β(J,J′) are listed in Table 7. As we can see, Nd:GLNO crystal has a relatively large value of β11/2 than other Nd-doped niobate crystals, which indicates that this crystal may be more easily to generate lasers at around 1.06 μm. The comparisons of Ω2,4,6 are shown in Figure 6. The Ω2 value of Nd:GLNO is higher than that of Nd:GNO crystal, point to the presence in Nd:GLNO crystal of Nd optical centers with lower environment symmetries determined by the doping of La3+ ions in GNO host, which induce a larger disorder around Nd3+ ions. Moreover, Nd:YNO crystal possesses the largest value of Ω2. The reason is that although these four crystals all belong to the monoclinic system, the angle of β in YNO is the largest, which indicates a lowest macroscopic symmetry in the Nd:YNO host. However, despite the Nd3+ ions possess a lowest environment symmetries in YNO host, which can supply a relatively strong crystal field for Nd3+ ions, the growth of Nd:YNO crystal is much harder than that of Nd:GNO crystal.
The room-temperature emission spectra of the as-grown Nd-doped niobate laser crystals at around 1.06 μm are shown in Figure 7. The strongest emission wavelength of Nd:GLNO, Nd:GNO, Nd:YNO, and Nd:GYNO are located at 1065.0 nm, 1065.7 nm, 1066 nm, and 1065.9 nm, respectively. Owing to the inhomogeneous broadening in the mixed crystal, Nd:GLNO and Nd:GYNO have a broader emission band at around 1.06 μm. In addition, stimulated emission cross-section can be estimated from the emission spectra using the F–L formula. Therefore, the stimulated emission cross-section values of Nd:GLNO Nd:GNO, Nd:YNO, and Nd:GYNO crystals at around 1.06 μm are estimated to be 18, 18.3, 22, and 20.5 × 10−20 cm2, respectively. Moreover, the fluorescence lifetimes of 4F3/24I11/2 transition are fitted to be 152, 178, 156, and 176.4 μs for Nd:YNO, Nd:GNO, Nd:GYNO, and Nd:GLNO, respectively. The small emission cross-section and long fluorescence lifetime indicates that the Nd:GLNO crystal possess good energy storage capacity, which is advantageous to its application in Q-switched laser.
In addition, laser performance of four crystals are operated based on a plano-plano resonator and the laser output power curves are shown in Figure 8. The transmission of output mirror in all of the laser experiment is 5.4% at 1.06 μm. The slope efficiency for Nd:YNO along b-orientation is 24.0%. The slope efficiency for Nd:GNO along three crystallographic axes (a-, b-, and c-) are 35.3%, 33.7%, and 28%, respectively. Additionally, the slope efficiency for Nd:GYNO along three crystallographic axes (a-, b- and c-) are 30.4%, 29.4% and 29.8%, respectively. Lastly, the slope efficiency for Nd:GLNO along c-orientation is 34.2%. Based on the above comparison of slope efficiency, Nd:GLNO is better than others.

5. Crystal Field Calculation

5.1. Calculations of Energy Levels

Using the relativistic model of ab initio self-consistent DV- method [54,55] and effective Hamiltonian model [56], the crystal-field and spin-orbit parameters of Nd3+ in GdTaO4 and LuTaO4 have been calculated. The parameters of Nd3+ in GdTaO4 and LuTaO4 have been shown in Table 8.
The experimental energy level values of Nd3+ are obtained from the previous spectroscopic analysis. Then, the crystal-field and spin-orbit parameters from DV-Xα method and effective Hamiltonian method and other free-ion parameters from Ref [57] were used as the initial values to fit the experimental energy levels of Nd3+ in GdTaO4, and LuTaO4 with the f-shell fitting program, which is developed by M. F. Ried with the Fortran language [58]. The calculated results, experimental results and the difference between the calculated values and experimental values are shown in Table 9.
From Table 8, most of the calculated levels are quite consistent with experimental energy levels. In Nd3+:LuTaO4, among the 152 experimental energy levels, there are four Stark levels (4354, 13,321, 14,767, and 17,194 cm−1) with poor fitting quality with deviations of −29.78, −28.10, −28.59, and 25.69 cm−1, respectively. Additionally, in Nd3+:GdTaO4, there are only two Stark levels (12,731 and 15,601 cm−1) with poor fitting quality with deviations of −24.05 and −22.02 cm−1, respectively. The deviation of calculated energy levels and experimental energy levels are less than 30 cm–1, which indicates that the fitted energy levels results are satisfactory.
It was concluded that the J-mixing effect does not play any significant role in the low-lying energy levels from the fitting results, especially for 4I9/2, 4I11/2, 4I13/2, 4I15/2, and 4F3/2 multiplets. However, at the higher-energy side, the J-mixing effect is clearly appreciable because the Stark components are overlapping with adjacent multiplets. On the other hand, the M-mixing effect is very clear for all energy levels.

5.2. Crystal Field Analysis

The experimental energy levels of Nd3+ give a basic set for a reliable energy level simulation. A model with 30 parameters including 16 free-ion and 14 crystal-field parameters is used for Nd3+ in rare earth tantalates, with the root mean square deviation (σ) of 12.66 and 14.6 cm−1. In the simulation process, it could be stated that the relative positions of some experimental energy levels depend on α, β, γ, and Ti. Such as the energy levels of 4S3/2 and 4F7/2, 4G9/2 and 2K13/2, 4G11/2, and 2K15/2 [59].
The free-ions and crystal-field parameters of Nd3+ in GdTaO4, LuTaO4, and YAlO3 hosts are shown in Table 10.
The values of free-ions parameters of Nd3+ in rare earth tantalates are similar to those in YAlO3 hosts. However, the electrostatic interaction between two 4f electrons of Nd3+ doped in rare earth tantalates is stronger than that of Nd3+ in YAlO3, because of the more delocalized electrons and decreased Slater integrals. With the Slater integrals decreasing, the individual 2S+1LJ state will be lowered, so the energy positions of 2P1/2 are 23,960 and 23,164 cm−1 for Nd3+ in GdTaO4 and YAlO3 hosts, respectively. The spin-orbit interaction varies slightly for these two hosts whereas the other free-ion parameters show no clear trend. This is because Nd3+ ion occupies the 4e sites of Gd(Lu) with C2 point symmetry, and each of the ions in rare earth tantalates is octahedrally coordinated with six O forming a distorted square antiprism, whereas Nd3+ ion occupies the 4c sites of Y with CS point symmetry, and each of the ions is dodecahedrally coordinated to eight O in YAlO3.
The values of crystal-field parameters are affected by the distances, bonding angles, and the nature of ligands. The crystal-field parameters B6 are only related to the ions of the nearest neighbor rather than the ions of next nearest neighbor and further, while B2 and B4 are not only related to the ions of the nearest neighbor but also the ions of next nearest neighbor and further. In Table 8, B2 and B4 of Nd3+ ions have relatively large differences among these rare earth tantalates and YAlO3, but the crystal-field parameters B6 are close to each other. This is because the interaction between 4f electrons of Nd3+ and Ta5+, Gd3+, and Lu3+ surrounding Nd3+ are strong in rare earth tantalates. It indicates that the charge interpenetration and coulomb exchange of Nd3+ in rare earth tantalates are stronger than those of Nd3+ in YAlO3 while covalency and overlap of Nd3+ in rare earth tantalates are weaker than those of Nd3+ in YAlO3. In fact, the overlap and coulomb exchange are responsible for about half the total parameters values, and the other half comes mostly from covalency. When the charge penetration contributions are taken into account, the magnitudes of the net electrostatic contributions are only of the order of 10% of the combined contributions from covalency and exclusion. Thus, in the analysis of the lanthanide crystal fields, approximations which neglect electrostatic contributions are likely to be far more realistic than those which neglect overlap and covalency contributions.
The Nν values are calculated from the two sets of crystal-field parameters, which are defined in terms of the length rather than the rotational invariants of the crystal-field parameters vector in the spherical harmonics space, allowing a direct comparison of the crystal-field strength in any type of symmetry[59]. From Table 9, we see that the contribution of rank six crystal-field parameters to the total crystal-field strength is larger in rare earth tantalates than that in YAlO3. Taking into consideration the electronegativity values of Gd(Lu) and Y ions, it might be concluded that in rare earth tantalates crystals, the neighbor ions of impurities are slightly nearer than those in the YAlO3 crystals. Thus, the contribution to the total crystal field of these neighbors is greater in rare earth tantalates than that in the YAlO3 hosts, as it was calculated. The Nν of Nd3+ in rare earth tantalates is larger than that in YAlO3, therefore, crystal-field interaction of rare earth tantalates is stronger than that of YAlO3.

6. Conclusions

  • Rare-earth orthotantalates have high density deserve to be studied as scintillators. Large volume and high-quality single crystals are grown and their properties are researched systematically by our group. The scintillation decay time of Nd:GdTaO4 single crystal is faster than GdTaO4 single crystal. LuTaO4 has the highest density 9.8 g/cm3 among the present luminescent material hosts, and the fluorescence lifetime of Nd:LuTaO4 is about 263.2 ns and expected as the most promising heavy scintillator.
  • Er:GdTaO4 and Tm,Ho:GdYTaO4 single crystals have good visible and mid-infrared fluorescence properties and can be the potential laser materials. Nd3+, Yb3+-doped othotantalates and orthoniobates have been realized laser output, which prove that they can be used as laser matrix hosts successfully.
  • The ab initio self-consistent DV-Xα method has been used in its relativistic model to investigate the crystal-field and spin-orbit parameters of Nd3+ doped in rare earth tantalates. The deviations of the calculated energy levels and experimental energy levels are less than 30 cm−1, which indicates that the energy levels fitting results are satisfactory. Through the calculation of the crystal field parameters it is shown that the crystal field is strong and is beneficial to the widen of the ion absorption band, which indicated that rare-earth orthotantalates will be promising laser crystal materials.

Acknowledgments

This work was financially supported by the National Natural Science Foundation of China (Grants Nos. 51502292, 51702322, and 61405206), National Key Research and Development Program of China (Grant No. 2016YFB0402101).

Author Contributions

Thank all the co-authors’ contributions for this paper. Renqin Dou researched the Yb-doped niobates and Tm/Yb,Ho-doped tantalates and wrote the paper; Yuanzhi Chen studied the Er-doped tantalates; Shoujun Ding analyzed the Nd-doped niobates; Fang Peng performed the Nd-doped tantalates; Wenpeng Liu contributed crystal growth; Dunlu Sun revised the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abe, R.; Higashi, M.; Sayama, K.; Abe, Y.; Sugihara, H. Photocatalytic activity of R3MO7 and R2Ti2O7 (R = Y, Gd, La; M = Nb, Ta) for water splitting into H2 and O2. J. Phys. Chem. B 2006, 110, 2219–2226. [Google Scholar] [CrossRef] [PubMed]
  2. Haugsrud, R.; Norby, T. Proton conduction in rare-earth ortho-niobates and ortho-tantalates. Nat. Mater. 2006, 5, 193–196. [Google Scholar] [CrossRef]
  3. Zhang, H.; Wang, Y.; Xie, L. Luminescent properties of Tb3+ activated GdTaO4 with M and M′ type structure under UV-VUV excitation. J. Lumin. 2010, 130, 2089–2092. [Google Scholar] [CrossRef]
  4. Forbes, T.Z.; Nyman, M.; Rodriguez, M.A.; Navrotskya, A. The energetics of lanthanum tantalate materials. J. Solid State Chem. 2010, 183, 2516–2521. [Google Scholar] [CrossRef]
  5. Sych, A.M.; Golub, A.M. Niobates and Tantalates of Trivalent Elements. Uspekhi Khimii 1977, 46, 417–444. [Google Scholar]
  6. Gundobin, N.V.; Petrov, K.I.; Plotkin, S.S. Vibration-Spectra and Structure Ln3MO7-Type Niobates and Tantalates. Zhurnal Neorganicheskoi Khimii 1977, 22, 2973–2977. [Google Scholar]
  7. Siqueira, K.P.F.; Carmo, A.P.; Bell, M.J.V.; Dias, A. Optical properties of undoped NdTaO4, ErTaO4 and YbTaO4 ceramics. J. Lumin. 2016, 179, 146–153. [Google Scholar] [CrossRef]
  8. Blasse, G.; Bril, A. Luminescence phenomena in compounds with fergusonite structure. J. Lumin. 1970, 3, 109–131. [Google Scholar] [CrossRef]
  9. Voloshyna, O.; Neicheva, S.V.; Starzhinskiy, N.G.; Zenyaa, I.M.; Gridina, S.S.; Baumerb, V.N.; Sidletskiya, O.T. Luminescent and scintillation properties of orthotantalates with common formulae RETaO4 (RE = Y, Sc, La, Lu and Gd). Mater. Sci. Eng. B 2013, 178, 1491–1496. [Google Scholar] [CrossRef]
  10. Kazakova, L.I.; Bykov, I.S.; Dubovsky, A.B. The luminescence of rare-earth tantalate single crystals. J. Lumin. 1997, 72–74, 211–212. [Google Scholar] [CrossRef]
  11. Brixner, L. On the structural and luminescent properties of the M′ LnTaO4 rare earth tantalates. J. Electrochem. Soc. 1983, 130, 2435–2443. [Google Scholar] [CrossRef]
  12. Stubičan, V. High-Temperature Transitions in Rare-Earth Niobates and TantaIates. J. Am. Ceram. Soc. 1964, 47, 55–58. [Google Scholar] [CrossRef]
  13. Siqueira, K.P.F.; Carvalho, G.B.; Dias, A. Influence of the processing conditions and chemical environment on the crystal structures and phonon modes of lanthanide orthotantalates. Dalton Trans. 2011, 40, 9454–9460. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, B.; Han, K.; Liu, X.; Gu, M.; Huang, S.; Ni, C.; Qi, Z.; Zhang, G. Luminescent properties of GdTaO4 and GdTaO4:Eu3+ under VUV-UV excitation. Solid State Commun. 2007, 144, 484–487. [Google Scholar] [CrossRef]
  15. Liu, X.; Chen, S.; Gu, M.; Wu, M.; Qiu, Z.; Liu, B.; Ni, C.; Huang, S. A novel M′-type LuTaO4:Ln(3+) (Ln = Eu, Tb) transparent scintillator films. Opt. Mater. Express 2014, 4, 172–178. [Google Scholar] [CrossRef]
  16. Yang, H.; Peng, F.; Zhang, Q.; Guo, C.; Shi, C.; Liu, W.; Sun, G.; Zhao, Y.; Zhang, D.; Sun, D.; et al. A promising high-density scintillator of GdTaO4 single crystal. CrystEngComm 2014, 16, 2480–2485. [Google Scholar] [CrossRef]
  17. Ning, K.; Zhang, Q.; Zhang, D.; Fan, J.; Sun, D.; Wang, X.; Hang, Y. Crystal growth, characterization of NdTaO4: A new promising stoichiometric neodymium laser material. J. Cryst. Growth 2014, 388, 83–86. [Google Scholar] [CrossRef]
  18. Ofelt, G.S. Intensities of Crystal Spectra of Rare-Earth Ions. J. Chem. Phys. 1962, 37, 511–520. [Google Scholar] [CrossRef]
  19. Judd, B.R. Optical Absorption Intensities of Rare-Earth Ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  20. Nikl, M.; Bruza, P.; Panek, D.; Vrbova, M.; Mihokova, E.; Mares, J.A.; Beitlerova, A.; Kawaguchi, N.; Fukuda, K.; Yoshikawa, A. Scintillation characteristics of LiCaAlF6-based single crystals under X-ray excitation. Appl. Phys. Lett. 2013, 102, 161907. [Google Scholar] [CrossRef]
  21. Rooh, G.; Kim H, J.; Park, H.; Kim, S. Investigation of scintillation and luminescence properties of cerium doped Rb2LiGdCl6 single crystals. J. Cryst. Growth 2013, 377, 28–31. [Google Scholar] [CrossRef]
  22. Nikl, M.; Solovieva, N.; Pejchal, J.; Shim, J.B.; Yoshikawa, A.; Fukuda, T.; Vedda, A.; Martini, M.; Yoon, D.H. Very fast YbxY1−xAlO3 single-crystal scintillators. Appl. Phys. Lett. 2004, 84, 882–884. [Google Scholar] [CrossRef]
  23. Moses, W.W.; Weber, M.J.; Derenzo, S.E.; Perry, D.; Berdahl, P.; Boatner, L.A. Prospects for dense, infrared emitting scintillators. IEEE Trans. Nucl. Sci. 1998, 45, 462–466. [Google Scholar] [CrossRef]
  24. Lecoq, P.; Dafinei, I.; Auffray, E.; Schneegans, M.; Korzhik, M.V.; Missevitch, O.V.; Pavlenko, V.B.; Fedorov, A.A.; Annenkov, A.N.; Kostylev, V.L.; et al. Lead tungstate (PbWO4) scintillators for LHC EM calorimetry. Nucl. Instrum. Methods Phys. Res. Sect. A 1995, 365, 291–298. [Google Scholar] [CrossRef]
  25. Almeida Silva, R.; Tirao, G.; Cusatis, C.; Andreeta, J.P. Growth and X-ray characterization of GdxYb1−xTaO4 (0 ≤ x ≤ 1) single crystals with large lattice spacing gradient. J. Cryst. Growth 2005, 277, 308–313. [Google Scholar] [CrossRef]
  26. Almeida Silva, R.; Tirao, G.; Cusatis, C.; Andreeta, J.P. Growth and structural characterization of M-type GdTaO4 single crystal fiber. J. Cryst. Growth 2005, 274, 512–517. [Google Scholar] [CrossRef]
  27. Liu, W.; Zhang, Q.; Zhou, W.; Gu, C.; Yin, S. Growth and Luminescence of M-Type GdTaO4 and Tb: GdTaO4 Scintillation Single Crystals. IEEE Trans. Nucl. Sci. 2010, 57, 1287–1290. [Google Scholar] [CrossRef]
  28. Yang, H.; Peng, F.; Zhang, Q.; Shi, C.; Guo, C.; Wei, X.; Sun, D.; Wang, X.; Dou, R.; Xing, X.; et al. Temperature-dependent luminescence of GdTaO4 single crystal. J. Lumin. 2015, 160, 90–94. [Google Scholar] [CrossRef]
  29. Peng, F.; Liu, W.; Zhang, Q.; Yang, H.; Shi, C.; Mao, R.; Sun, D.; Luo, J.; Sun, G. Crystal growth, optical and scintillation properties of Nd3+ doped GdTaO4 single crystal. J. Cryst. Growth 2014, 406, 31–35. [Google Scholar] [CrossRef]
  30. Blasse, G.; Dirksen, G.J.; Brixner, L.H.; Crawford, M.K. Luminescence of Materials Based On LuTaO4. J. Alloy. Compd. 1994, 209, 1–6. [Google Scholar] [CrossRef]
  31. Liu, W.; Zhang, Q.; Ding, L.; Sun, D.; Luo, J.; Yin, S. Photoluminescence properties of LuTaO4:RE3+ (RE3+ = Eu3+, Tb3+) with M′-type structure. J. Alloy. Compd. 2009, 474, 226–228. [Google Scholar] [CrossRef]
  32. Zhao, Y.; Zhang, Q.; Guo, C.; Shi, C.; Peng, F.; Yang, H.; Sun, D.; Luo, J.; Liu, W. The luminescence properties of the high-density phosphor Lu1−xNdxTaO4. J. Lumin. 2014, 155, 165–169. [Google Scholar] [CrossRef]
  33. Xing, X.; Wang, X.; Zhang, Q.; Sun, G.; Shi, C.; Liu, W.; Sun, D.; Yin, S. High-Temperature Phase Relations in the Lu2O3-Ta2O5 System. J. Am. Ceram. Soc. 2016, 99, 1042–1046. [Google Scholar] [CrossRef]
  34. Yang, H.; Zhang, Q.; Zhou, P.; Ning, K.; Gao, J.; He, Y.; Luo, J.; Sun, D.; Yin, S. Czochralski growth and optical investigations of Er3+:GdTaO4 laser crystal. In Proceedings of the SPIE 8206, Pacific Rim Laser Damage 2011: Optical Materials for High Power Lasers, Shanghai, China, 12 January 2012. [Google Scholar]
  35. Chen, Y.; Peng, F.; Zhang, Q.; Liu, W.; Dou, R.; Ding, S.; Luo, J.; Sun, D.; Sun, G.; Wang, X. Growth, structure and spectroscopic properties of 1 at.% Er3+: GdTaO4 laser crystal. J. Lumin. 2017, 192, 555–561. [Google Scholar] [CrossRef]
  36. Lupei, V.; Georgescu, S.; Florea, V. On the dynamics of population inversion for 3 μm Er3+ lasers. IEEE J. Quantum Electron. 1993, 29, 426–434. [Google Scholar] [CrossRef]
  37. Su, J.; Yang, C.; Li, Q.; Zhang, Q.; Luo, J. Optical spectroscopy of Er:YSGG laser crystal. J. Lumin. 2010, 130, 1546–1550. [Google Scholar] [CrossRef]
  38. Xu, X.; Zhao, G.; Wu, F.; Xu, W.; Zong, Y.; Wang, X.; Zhao, Z.; Zhou, G.; Xu, J. Growth and spectral properties of Er:Gd2SiO5 crystal. J. Cryst. Growth 2008, 310, 156–159. [Google Scholar] [CrossRef]
  39. Ding, Y.; Zhao, G.; Xu, X. Crystal growth and spectroscopic properties of erbium doped Lu2SiO5. J. Cryst. Growth 2010, 312, 2103–2106. [Google Scholar] [CrossRef]
  40. Peng, F.; Yang, H.; Zhang, Q.; Luo, J.; Liu, W.; Sun, D.; Dou, R.; Sun, G. Spectroscopic properties and laser performance at 1066 nm of a new laser crystal Nd:GdTaO4. Appl. Phys. B 2015, 118, 549–554. [Google Scholar] [CrossRef]
  41. Peng, F.; Yang, H.; Zhang, Q.; Luo, J.; Sun, D.; Liu, W.; Sun, G.; Dou, R.; Wang, X.; Xing, X. Growth, thermal properties, and LD-pumped 1066 nm laser performance of Nd3+ doped Gd/YTaO4 mixed single crystal. Opt. Mater. Express 2015, 5, 2536–2544. [Google Scholar] [CrossRef]
  42. Tempus, M.; Luthy, W.; Weber, H.P.; Ostroumov, V.G.; Shcherbakov, I.A. 2.79 μm YSGG:Cr:Er Laser pumped at 790 nm. IEEE J. Quantum Electron. 1994, 30, 2608–2611. [Google Scholar] [CrossRef]
  43. Vodopyanov, K.L.; Ganikhanov, F.; Maffetone, J.P.; Zwieback, I.; Ruderman, W. ZnGeP2 optical parametric oscillator with 3.8–12.4 μm tunability. Opt. Lett. 2000, 25, 841–843. [Google Scholar] [CrossRef] [PubMed]
  44. Diening, A.; Kuck, S. Spectroscopy and diode-pumped laser oscillation of Yb3+, Ho3+-doped yttrium scandium gallium garnet. J. Appl. Phys. 2000, 87, 4063–4068. [Google Scholar] [CrossRef]
  45. Dou, R.; Zhang, Q.; Liu, W.; Luo, J.; Wang, X.; Ding, S.; Sun, D. Growth, structure, chemical etching, and spectroscopic properties of a 2.9 μm Tm,Ho:GdYTaO4 laser crystal. Opt. Mater. 2015, 48, 80–85. [Google Scholar] [CrossRef]
  46. Dou, R.; Zhang, Q.; Sun, D.; Luo, J.; Yang, H.; Liu, W.; Sun, G. Growth, thermal, and spectroscopic properties of a 2.911 μm Yb,Ho:GdYTaO4 laser crystal. CrystEngComm 2014, 16, 11007–11012. [Google Scholar] [CrossRef]
  47. Chen, D.W.; Fincher, C.L.; Rose, T.S.; Vernon, F.L.; Fields, R.A. Diode-pumped 1-W continuous-wave Er:YAG 3 μm laser. Opt. Lett. 1999, 24, 385–387. [Google Scholar] [CrossRef] [PubMed]
  48. Deloach, L.D.; Payne, S.A.; Smith, L.K.; Kway, W.L.; Krupke, W.F. Laser And Spectroscopic Properties of Sr5(PO4)(3)F:Yb. J. Opt. Soc. Am. B 1994, 11, 269–276. [Google Scholar] [CrossRef]
  49. Boulon, G.; Collombet, A.; Brenier, A.; Cohen-Adad, M.-T.; Yoshikawa, A.; Lebbou, K.; Lee, J.H.; Fukuda, T. Structural and spectroscopic characterization of nominal Yb3+:Ca8La2(PO4)(6)O2 oxyapatite single crystal fibers grown by the micro-pulling-down method. Adv. Funct. Mater. 2001, 11, 263–270. [Google Scholar] [CrossRef]
  50. Ding, S.; Zhang, Q.; Luo, J.; Liu, W.; Lu, W.; Xu, J.; Sun, G.; Sun, D. Crystal growth, structure, defects, mechanical and spectral properties of Nd0.01:Gd0.89La0.1NbO4 mixed crystal. Appl. Phys. A 2017, 123, 636. [Google Scholar] [CrossRef]
  51. Ding, S.; Peng, F.; Zhang, Q.; Luo, J.; Liu, W.; Sun, D.; Dou, R.; Sun, G. Structure, spectroscopic properties and laser performance of Nd:YNbO4 at 1066 nm. Opt. Mater. 2016, 62, 7–11. [Google Scholar] [CrossRef]
  52. Ding, S.; Peng, F.; Zhang, Q.; Luo, J.; Liu, W.; Sun, D.; Dou, R.; Gao, J.; Sun, G.; Cheng, M. Crystal growth, spectral properties, and continuous wave laser operation of Nd:GdNbO4. J. Alloy. Compd. 2017, 693, 339–343. [Google Scholar] [CrossRef]
  53. Ding, S.; Zhang, Q.; Peng, F.; Liu, W.; Luo, J.; Dou, R.; Sun, G.; Wang, X.; Sun, D. Crystal growth, spectral properties and continuous wave laser operation of new mixed Nd:GdYNbO4 laser crystal. J. Alloy. Compd. 2017, 698, 59–63. [Google Scholar] [CrossRef]
  54. Rosen, A.; Ellis, D.E. Relativistic Molecular Calculations in Dirac-Slater Model. J. Chem. Phys. 1975, 62, 3039–3049. [Google Scholar] [CrossRef]
  55. Van Pieterson, L.; Reid, M.F.; Wegh, R.T.; Soverna, S.; Meijerink, A. 4fn → 4fn−15d transitions of the light lanthanides: Experiment and theory. Phys. Rev. B 2002, 65, 045113. [Google Scholar] [CrossRef]
  56. Reid, M.F.; Duan, C.-K.; Zhou, H. Crystal-field parameters from ab initio calculations. J. Alloy. Compd. 2009, 488, 591–594. [Google Scholar] [CrossRef]
  57. Molina, P.; Loro, H.; Álvarez-García, S.; Bausá, L.E.; Rodriguez, E.M.; Guillot-Noël, O.; Goldner, P.; Bettinelli, M.; Ghigna, P.; Solé, J.G. Site location and crystal field of Nd3+ ions in congruent strontium barium niobate. Phys. Rev. B 2009, 80, 054111. [Google Scholar] [CrossRef]
  58. Loro, H.; Vasquez, D.; Camarillo, G.E.; del Castillo, H.; Camarillo, I.; Muñoz, G.; Flores, J.C.; Hernández, A.J.; Murrieta, S.H. Experimental study and crystal-field modeling of Nd3+ 4f3 energy levels in Bi4Ge3O12 and Bi4Si3O12. Phys. Rev. B 2007, 75, 125405. [Google Scholar] [CrossRef]
  59. Dagama, A.A.S.; Desa, G.F.; Porcher, P.; Caro, P. Energy-Levels of Nd3+ in LiYF4. J. Chem. Phys. 1981, 75, 2583–2587. [Google Scholar] [CrossRef]
  60. Duan, C.-K.; Tanner, P.A.; Makhov, V.N.; Kirm, M. Vacuum ultraviolet spectra and crystal field analysis of YAlO3 doped with Nd3+ and Er3+. Phys. Rev. B 2007, 75, 195130. [Google Scholar] [CrossRef]
Figure 1. Photographs of the as-grown crystals: (a) GdTaO4; (b) Nd:GdTaO4; (c) Nd:GdYTaO4; (d) Er:GdTaO4; (e) Tm,Ho:GdYTaO4; (f) Yb,Ho:GdYTaO4; (g) Yb:GdNbO4; (h) Yb:YNbO4; (i) Nd:GdNbO4; (j) Nd:YNbO4; (k) Nd:GdYNbO4; and (l) Nd:GdLaNbO4.
Figure 1. Photographs of the as-grown crystals: (a) GdTaO4; (b) Nd:GdTaO4; (c) Nd:GdYTaO4; (d) Er:GdTaO4; (e) Tm,Ho:GdYTaO4; (f) Yb,Ho:GdYTaO4; (g) Yb:GdNbO4; (h) Yb:YNbO4; (i) Nd:GdNbO4; (j) Nd:YNbO4; (k) Nd:GdYNbO4; and (l) Nd:GdLaNbO4.
Crystals 08 00055 g001aCrystals 08 00055 g001b
Figure 2. Dislocation etching pit patterns of Tm,Ho:GdYTaO4 crystal at three different crystalline faces: (a) (100); (b) (010); and (c) (001).
Figure 2. Dislocation etching pit patterns of Tm,Ho:GdYTaO4 crystal at three different crystalline faces: (a) (100); (b) (010); and (c) (001).
Crystals 08 00055 g002
Figure 3. Dislocation etching pit patterns of Nd:GdYNbO4 crystal at three different crystalline faces: (a) (100); (b) (010); and (c) (001).
Figure 3. Dislocation etching pit patterns of Nd:GdYNbO4 crystal at three different crystalline faces: (a) (100); (b) (010); and (c) (001).
Crystals 08 00055 g003
Figure 4. Phase-relation diagram of the Lu2O3-Ta2O5 system [33]. Reproduced with permission from [Xue Xing et al.], [J. Am. Ceram. Soc.]; published by [John Wiley and Sons], [2015].
Figure 4. Phase-relation diagram of the Lu2O3-Ta2O5 system [33]. Reproduced with permission from [Xue Xing et al.], [J. Am. Ceram. Soc.]; published by [John Wiley and Sons], [2015].
Crystals 08 00055 g004
Figure 5. Emission cross-section spectra of Tm,Ho:GdYTaO4 and Yb,Ho:GdYTaO4 crystal at room temperature.
Figure 5. Emission cross-section spectra of Tm,Ho:GdYTaO4 and Yb,Ho:GdYTaO4 crystal at room temperature.
Crystals 08 00055 g005
Figure 6. Comparison of Ω2,4,6 between Nd:GNO, Nd:YNO, Nd:GYNO, and Nd:GLNO crystals.
Figure 6. Comparison of Ω2,4,6 between Nd:GNO, Nd:YNO, Nd:GYNO, and Nd:GLNO crystals.
Crystals 08 00055 g006
Figure 7. Emission spectra of Nd:GNO, Nd:YNO, Nd:GYNO, and Nd:GLNO crystals excited by 808 nm at room temperature.
Figure 7. Emission spectra of Nd:GNO, Nd:YNO, Nd:GYNO, and Nd:GLNO crystals excited by 808 nm at room temperature.
Crystals 08 00055 g007
Figure 8. Laser output power of Nd:YNO, Nd:GNO, Nd:GYNO, and Nd:GLNO crystals versus incident power for different directions.
Figure 8. Laser output power of Nd:YNO, Nd:GNO, Nd:GYNO, and Nd:GLNO crystals versus incident power for different directions.
Crystals 08 00055 g008
Table 1. Lattice parameters of as-grown crystals.
Table 1. Lattice parameters of as-grown crystals.
Crystala (Å)b (Å)c (Å)β (°)ρ3)V (g cm3)Rp%Rwp%
GdTaO45.4111.055.0795.588.83302.56--
Nd:GdTaO45.4011.065.0895.618.89302.094.36.24
Nd:GdYTaO45.3811.045.0895.588.42300.259.97.34
Er:GdTaO45.4111.065.0995.618.79302.705.207.20
Tm,Ho:GdYTaO45.3911.035.0895.618.595300.438.916.22
Yb,Ho:GdYTaO45.3911.045.0895.608.602300.682.921.11
Yb:GdNbO45.3611.075.1094.577.140301.4594.586.26
Nd:GdNbO45.3811.095.1194.566.875303.4104.145.44
Nd:YNbO47.04110.9525.30194.565.568299.6406.708.75
Nd:GdYNbO45.35011.0505.09094.566.525300.3124.045.28
Nd:GdLaNbO45.38111.1125.11194.566.791304.6945.864.20
Table 2. Comparison of Ωt values with other Er-doped hosts.
Table 2. Comparison of Ωt values with other Er-doped hosts.
CrystalsΩ2/10−20 cm2Ω4/10−20 cm2Ω6/10−20 cm2Ω46
30 at% Er:YSGG [37]0.230.860.372.32
0.6 at% Er:Gd2SiO5 [38]6.1681.8781.2551.50
0.5 at% Er:Lu2SiO5 [39]4.4511.6141.1581.39
30 at% Er:GdTaO4 [34]0.981.174.670.25
1 at% Er:GdTaO4 [35]7.5531.9991.4041.42
Table 3. Comparison of spectroscopic properties of Nd:GdTaO4 and Nd:GdYTaO4 with other Nd-doped laser crystals [41].
Table 3. Comparison of spectroscopic properties of Nd:GdTaO4 and Nd:GdYTaO4 with other Nd-doped laser crystals [41].
CrystalsFWHM (@808 nm)σα (10–20 cm2) (@808 nm)σem (10−20 cm2) (@1.06 μm)τem (μs)
Nd:GdYTaO46~126.922182
Nd:GdTaO465.139178
Nd:YAG1.58.334240
Nd:YVO44.44015699
Table 4. The absorption coefficient of a, b, and c directions at the pumping wavelength.
Table 4. The absorption coefficient of a, b, and c directions at the pumping wavelength.
Crystalsα/cm−1
(3H63H4)
α/cm−1
(2F7/22F5/2)
abcabc
Tm,Ho:GdYTaO42.752.793.88
Yb,Ho:GdYTaO45.085.609.67
Table 5. Comparison of the lifetimes of 5I7 and 5I6 in different Ho3+-doped crystals [45,46,47].
Table 5. Comparison of the lifetimes of 5I7 and 5I6 in different Ho3+-doped crystals [45,46,47].
Crystals5I6/μs5I7/ms5I7/5I6
Tm,Ho:GdYTaO41314.0931.2
Yb,Ho:GdYTaO44197.317.4
Yb,Ho:YSGG58510.217.4
Tm,Ho:LuAG2507.530.0
Tm,Ho:YAG4011.4285
Table 6. Absorption and emission cross-sections for Yb:YNbO4, Yb:YNbO4, and some other Yb3+-doped crystals [48,49] at room temperature. (σabs: absorption cross-section; σem: emission cross-section).
Table 6. Absorption and emission cross-sections for Yb:YNbO4, Yb:YNbO4, and some other Yb3+-doped crystals [48,49] at room temperature. (σabs: absorption cross-section; σem: emission cross-section).
Crystalσabs/10−20 cm2σem/10−20 cm2
Yb:YNbO40.73 (933 nm)1.81 (1005 nm)
1.85 (955 nm)1.11 (1021 nm)
0.86 (974 nm)0.57 (1030 nm)
0.44 (1003 nm)
Yb:GdNbO40.87 (936 nm)0.446 (1003)
0.97 (955 nm)0.487 (1018)
0.85 (975 nm)0.466 (1030)
0.34 (1000 nm)
Yb:Y2O31.16 (978 nm)0.95 (1032 nm)
Yb:YAG0.86 (935 nm)2.50 (1030 nm)
Yb:Lu2O31.23 (978 nm)0.95 (1034.5 nm)
Yb:GSO0.69 (925 nm)0.66 (1030 nm)
0.38 (1048 nm)
0.41 (1088 nm)
Yb:FAP10.0 (905 nm)5.9 (942 nm)
Table 7. Comparisons of τrad and β(J,J′) between Nd:GNO, Nd:YNO, Nd:GYNO, and Nd:GLNO crystals.
Table 7. Comparisons of τrad and β(J,J′) between Nd:GNO, Nd:YNO, Nd:GYNO, and Nd:GLNO crystals.
Crystalsβ9/2β11/2β13/2β15/2τrad
1 at% Nd:GLNO31.2255.1013.040.64176
1 at% Nd:GYNO33.0453.8812.470.61156
2 at% Nd:GNO31.5354.9012.950.62178
1 at% Nd:YNO34.3952.9812.050.58152
Table 8. Parameters of Nd3+ in GdTaO4 and LuTaO4.
Table 8. Parameters of Nd3+ in GdTaO4 and LuTaO4.
ParametersNd3+: GdTaO4Nd3+: LuTaO4
B 0 2 −1891778
B 2 2 −299−321
B 0 4 702−682
B 2 4 601 + 267i−1017 + 795i
B 4 4 −302 + 149i105 + 648i
B 0 6 −696−421
B 2 6 248 + 620i−302 − 118i
B 4 6 579 − 12i−651 − 848i
B 6 6 77 − 168i359 + 741i
ξ998902
Table 9. Energy levels of Nd3+ in rare earth tantalates.
Table 9. Energy levels of Nd3+ in rare earth tantalates.
2S+1LJNd3+ Energy Levels of GdTaO4Nd3+ Energy Levels of LuTaO4
E (calc)E (exp)ΔE (cm−1)E (calc)E (exp)ΔE (cm−1)
4I9/2−4.8104.81−21.81021.81
122.09117−5.09105.211159.79
242.10238−4.10162.71151−11.71
372.17367−5.17289.78275−14.78
489.454988.55640.44631−9.44
4I11/22319.311978.3419834.66
2369.081992.01200512.99
2450.842058.632039−19.63
2516.722123.432123−0.43
2592.422292.47230411.53
2656.792354.4023616.60
4I13/24619.13463212.873924.453921−3.45
4672.954668−4.953933.5539395.45
4735.5947459.413991.053980−11.05
4829.13484212.874055.6040571.40
4873.804250.4242609.58
4936.6749447.334297.58431012.42
5044.435036−8.434383.784354−29.78
4I15/26874.396872−2.395830.42
7000.32701312.685880.84
7071.00708413.005929.69
7134.39714813.615989.77
7226.49723912.516349.72
7322.9273241.086405.49
7462.4874707.526552.21
7621.336630.59
4F3/211,617.9811,615−2.9811,368.5911,365−3.59
11,878.5611,877−1.5611,451.7011,450−1.70
4F5/2 + 2H9/2(2)12,755.0512,731−24.0512,336.6412,3414.36
12,874.0912,862−12.0912,390.7212,40918.28
12,907.3012,898−9.3012,510.2512,509−1.25
12,954.0612,935−19.0612,551.6012,550−1.60
13,026.2213,014−12.2212,599.5912,590−9.59
13,185.5613,171−14.5612,695.5112,677−18.51
4F7/2 + 4S3/213,926.8113,9369.1913,349.1013,321−28.10
14,094.8214,082−12.8213,400.9413,382−18.94
14,151.6914,16311.3113,487.0413,4979.96
14,184.8113,499.6713,51717.33
13,514.6413,53217.36
13,611.0313,6.2311.97
4F9/215,451.3115,46412.6914,583.4414,573−10.44
15,623.0215,601−22.0214,795.5914,767−28.59
15,675.8515,664−11.8514,868.5214,8734.48
15,740.0815,842.0215,8441.98
15,814.3215,82510.6815,929.3115,927−2.31
2H11/2(2)16,824.2716,8338.7316,846.2516,844−2.25
16,881.0716,89614.9316,919.5316,9222.47
16,911.4917,058.5017,036−22.50
16,947.0916,96517.9117,168.3117,19425.69
4G5/2 + 2G7/217,629.6417,64616.3617,214.2317,2216.77
17,898.5617,884−14.5617,236.0917,25619.91
17,959.8817,940−19.8817,261.02
17,288.86
17,455.4717,4593.53
4G7/219,681.7419,69715.2618,707.7018,7113.30
19,834.9719,825−9.9718,811.8418,811−0.84
18,914.4818,9249.52
18,949.08
2K13/2 + 4G9/220,511.9520,502−9.9519,135.7819,15014.22
20,680.2220,6876.7819,292.5519,30714.45
19,378.4119,3790.59
2G9/2(1) + 4G11/2 + 2D3/2(1) + 2K15/221,672.1221,6785.8820,780.7920,7909.21
22,014.2420,844.5620,843−1.56
22,240.8822,230−10.8820,977.0220,973−4.02
22,847.4122,841−6.4121,022.9421,017−5.94
21,098.5321,088−10.53
21,183.8321,19410.17
21,519.9621,5233.04
21,604.8721,6061.13
21,762.0621,77411.94
2P1/223,973.7723,960−13.7723,085.1123,0959.89
2D(1)5/2 23,597.1023,585−12.10
23,698.3723,685−13.37
23,811.1523,787−24.15
2P3/2 25,952.6025,96512.40
26,074.5726,074−0.57
4D3/227,638.9727,628−10.9727,535.9827,5393.02
27,648.1727,636−12.17
4D5/2 27,840.1627,8487.84
28,029.6228,027−2.62
2I11/2 + 4D1/229,682.9829,676−6.9828,221.2428,218−3.24
28,328.9528,324−4.95
2I13/2 + 2L17/231,640.0131,6509.9928,928.9628,917−11.96
29,682.4329,677−5.43
4D7/2 + 2D(2)3/2 30,182.4330,177−5.43
32,893.1332,8951.87
2H(1)11/2 33,790.3433,80615.66
2F(2)5/235,808.8535,82213.1534,382.58
Table 10. Comparing the free-ions and crystal-field parameters.
Table 10. Comparing the free-ions and crystal-field parameters.
ParametersNd3+:GdTaO4Nd3+:LuTaO4Nd3+: YAlO3 [60]
Eavg25,16724,14124,119
F273,01873,29070,925
F452,78960,03050,794
F635,75744,22035,424
ξ1034883875
α21.3411.2223
β−593−436−691
γ1445−16941690
T2[298][298][458]
T3[35][35][38]
T4[59][59][75]
T6[−285][−285][−290]
T7[332][332][237]
T8[305][305][496]
M2.112.111.90
P192192206
B 0 2 −1560591−154
B 2 2 −415−153 − 70i578
B 0 4 626−753−541
B 2 4 422+315i−1428 + 611i967 + 24i
B 4 4 244 + 18i516 + 898i−309+608i
B 0 6 −613−522−671
B 2 6 −301 + 807i−147 − 43i512 − 18i
B 4 6 −100 − 103i172 − 820i1611 + 0i
B 6 6 −279 − 107i533 + 864i0 + 132i
σ12.6614.615.50
Nν295929272545

Share and Cite

MDPI and ACS Style

Dou, R.; Zhang, Q.; Gao, J.; Chen, Y.; Ding, S.; Peng, F.; Liu, W.; Sun, D. Rare-Earth Tantalates and Niobates Single Crystals: Promising Scintillators and Laser Materials. Crystals 2018, 8, 55. https://doi.org/10.3390/cryst8020055

AMA Style

Dou R, Zhang Q, Gao J, Chen Y, Ding S, Peng F, Liu W, Sun D. Rare-Earth Tantalates and Niobates Single Crystals: Promising Scintillators and Laser Materials. Crystals. 2018; 8(2):55. https://doi.org/10.3390/cryst8020055

Chicago/Turabian Style

Dou, Renqin, Qingli Zhang, Jinyun Gao, Yuanzhi Chen, Shoujun Ding, Fang Peng, Wenpeng Liu, and Dunlu Sun. 2018. "Rare-Earth Tantalates and Niobates Single Crystals: Promising Scintillators and Laser Materials" Crystals 8, no. 2: 55. https://doi.org/10.3390/cryst8020055

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