Spectral properties of Pr:CNGG crystals grown by micro-pulling-down method
Xue Yan-Yan1, Li Na1, Hu Dong-Hua1, Song Qing-Song2, Xu Xiao-Dong2, †, Wang Dong-Hai1, Wang Qing-Guo1, Li Dong-Zhen2, Wang Zhan-Shan1, Xu Jun1, ‡
School of Physics Science and Engineering, Institute for Advanced Study, Tongji University, Shanghai 200092, China
Jiangsu Key Laboratory of Advanced Laser Materials and Devices, School of Physics and Electronic Engineering, Jiangsu Normal University, Xuzhou 221116, China

 

† Corresponding author. E-mail: xdxu79@jsnu.edu.cn xujun@mail.shcnc.ac.cn

Abstract

Pr3+-doped calcium niobium gallium garnet (Pr:CNGG) single crystals with different Pr3+concentrations are successfully grown by the micro-pulling-down (μ-PD) method. The crystal structure, room-temperature absorption spectra, and fluorescence spectra of Pr:CNGG crystals are measured and discussed. The fluorescence results indicate their large dependence on the doping concentration. The fluorescence lifetime of the energy level is also determined. The results indicate that Pr:CNGG crystal could be a potential solid-state laser gain medium.

1. Introduction

The trivalent praseodymium ion (Pr3+) has been widely studied in the past few years due to the rich emission spectrum in ultraviolet, visible, and infrared regions.[13] The first Pr-based laser output was achieved in CaWO4 crystal, operating at .[4] After that, blue light emission was realized in Pr:LiYF4 crystal by using a pulsed dye laser.[5] With the development of GaN/InGaN laser diodes (LDs), the research process of the Pr3+-doped materials has been promoted. So far, Pr3+-doped hosts, such as crystals[68] and glasses,[9] have been widely investigated.

The CNGG crystal belongs to the cubic phase (Ia3d) garnet-type structure, which is the representative disordered crystal. It is characterized by low melting point (about 1460 °C) in comparison with other garnet crystals, and excellent mechanical and thermal properties in comparison with glasses.[10] In addition, the random distribution of Nb5+, Ga3+ and cationic vacancies in the host lattices cause the spectra of the doped rare-earth ions to significantly inhomogeneously broaden,[11] which makes it possible to produce ultra-short pulses. Up to now, a large number of research focuses on rare ion-doped CNGG crystals.[1215] Nevertheless, only one paper reported the crystal growth and structural properties of Pr:CNGG crystal,[16] but the spectral properties were not mentioned in detail, to the best of our knowledge.

The μ-PD method allows the direct growing of fiber-shaped single crystals with a diameter as low as , and it has many advantages over the conventional crystal growth methods, such as fast growth rate, low cost, and controllable crystal shape. So the μ-PD method is a promising method of growing gain media for miniature lasers.

In this paper, we grow the Pr:CNGG single crystals by the μ-PD method. The crystal structure, room-temperature spectral properties and lifetimes of Pr:CNGG crystals are measured and discussed in detail.

2. Experimental details
2.1. Crystal growth

The CNGG single crystals doped with 0.5, 1.0, and 2.0 at.% Pr3+ were successfully grown by the μ-PD method. The CaCO3, Nb2O5, Ga2O3, and Pr6O11 powders with a purity of 99.999% were prepared as starting materials and weighted according to the formula Ca3Nb1.6875Ga3.1875O12. The Pr6O11 powders were doped together with additional Ga2O3 in a ratio of 3:5 (mol%) to form Pr3Ga5O12.[17] Pr-doping as a garnet phase is important to maintain garnet composition in melt. The powders were well mixed and pressed into bulks, then sintered at 1300 ° C for 12 h in a muffle furnace with air atmosphere in order to react completely. The polycrystalline material was melted within a platinum crucible, and then passed through the micro-nozzle at the bottom of the crucible with a pulling rate of 0.5 mm/min. The growth direction was along the orientation. The as-grown Pr:CNGG single crystals with an average diameter of about 2 mm is shown in Fig. 1. The crystals were transparent and had a little yellow color, changing from light to deep with the increase of Pr3+ concentration.

Fig. 1. As-grown Pr:CNGG single crystals.
2.2. Characterizations

To identify the structure of the grown Pr:CNGG crystals, the x-ray diffraction (XRD) analysis was carried out in the 2θ ranging from 10° to 90° using an automated Ultima IV diffractometer (Cu target, , Rigaku, Japan) with a scan width of 0.014° at room temperature. Samples with a thickness of 1.07 mm were cut perpendicularly to the growth orientation and their faces were optically polished for the spectroscopic measurements. The absorption spectra with a resolution of 1 nm were measured from 400 nm to 2500 nm by a spectrophotometer (Cary5000, UV-VIS-IR). The fluorescence spectra and decay curves of the crystals were obtained by an FLS980 fluorescence spectrometer (Edinburgh Company, England). The exciting source was a xenon lamp. All the measurements were carried out at room temperature.

3. Results and discussion
3.1. XRD pattern

The samples are cut from the as-grown crystals and grounded into powder for XRD measurement. Figure 2 shows the XRD patterns of Pr:CNGG crystals. The diffraction peak positions and relative intensity of samples are in good agreement with those of the undoped CNGG crystal and ICSD Card

Fig. 2. XRD patterns of Pr:CNGG crystals.

No. 261821,[18] and no other phase is found in the lattice structure after Pr3+ ions have been doped into CNGG crystals, which implies that the Pr:CNGG crystals have the same crystallographic structure as pure CNGG crystal and belong to the cubic phase (Ia3d) garnet-type structure.

3.2. Spectral properties

Figure 3 shows the absorption spectra of the Pr:CNGG crystals with different dopant concentrations in the range of 420 nm–620 nm and 1200 nm–2500 nm. The absorption spectra consist of seven prominent absorption bands, which correspond to the transitions from ground state to excited states , , , , , , and , respectively. The absorption coefficients of 0.5, 1.0, and 2.0 at.% Pr:CNGG crystals at 449 nm are 1.26 cm−1, 2.88 cm−1, and 4.80 cm−1, respectively. The full widths at half-maximum (FWHMs) of the absorption peak at 449 nm for different concentrations are similar, which is about 7 nm. The full width at half-maximum (FWHM) of Pr:CNGG crystal is larger than that of Pr:LiLuF4 (1.7 nm),[19] Pr:LiYF4 (1.8 nm),[19] and Pr:LaF3 (5 nm),[20] which makes Pr:CNGG crystal more suitable to be pumped by the GaN/InGaN LD.

Fig. 3. Absorption spectra of Pr:CNGG crystals.

The fluorescence spectra of the Pr:CNGG crystals in the visible range under excitation at 449 nm and in the near-infrared (NIR) range under excitation at 606 nm are presented in Fig. 4. The emission bands in the visible range mainly correspond to the transition from and energy level to lower levels. In the NIR range, there are two intensive fluorescence bands centered at 824 nm and 1082 nm, corresponding to the transitions from excited state to and , respectively.

Fig. 4. Fluorescence spectra of the Pr:CNGG crystals in the visible and NIR range.

Obvious differences could be found when the doping concentration of Pr3+ increases from 0.5 at.% to 2.0 at.%. Firstly, 0.5-at.% Pr:CNGG crystal shows a larger fluorescence ratio of and transition centered at 606 nm and 632 nm than other Pr:CNGG crystals with higher Pr3+ concentration. There are two kinds of cross relaxation processes between Pr3+ ions[21] in Pr:CNGG crystal as shown in Fig. 5. The populations accumulate on the energy level by the multi-phonon relaxation from energy level and the cross relaxation between and (CR1). As the Pr3+ concentration increases, the cross relaxation between and (CR2) increases, resulting in the fluorescence quenching of energy level. Thus, 0.5-at.% Pr:CNGG crystal shows relatively large fluorescence ratio of emission bands centered 606 nm and 632 nm, derived from the transition of energy level. The result of multi-phonon relaxation and CR1 is that the fluorescence intensity centered at 617 nm, corresponding to the transition from to , is very weak. The most intensive fluorescence band of 2.0-at.% Pr:CNGG crystal is centered at 654 nm in the red spectral region.A similar phenomenon has been reported in Pr3+-doped phosphate glasses.[21] Secondly, 1.0-at.% and 2.0-at.% Pr:CNGG crystals show similar fluorescence ratios in the NIR range. The FWHMs of the fluorescence bands centered at 606 nm and 654 nm are both about 7 nm, which are larger than that of Pr:LiLuF4, Pr:LiYF4,[19] and Pr:YAlO3.[22] The FWHM of the emission band centered at 1082 nm is about 41 nm, indicating that a tunable and ultrafast laser around 1082 nm may be realized in Pr:CNGG crystal.

Fig. 5. Simple diagram of energy configuration of Pr3+ in CNGG crystal and cross relaxation processes.

The fluorescence decay curves at 606 nm corresponding to the energy level, under excitation at 449 nm are shown in Fig. 6. The fluorescence lifetime of the energy level in the 0.5, 1.0, and 2.0-at.% Pr:CNGG crystals are estimated at , , and , which are smaller than that of 0.5% Pr:YAG ( ),[23] 1% Pr:GGG ( ),[24] but larger than that of 1% Pr:YAP ( ).[25] The fluorescence lifetime decreases from to when the doping concentration of Pr3+ increases from 0.5 at.% to 2.0 at.%, which is mainly due to the depopulation of the level, caused by CR2 as shown in Fig. 5 and concentration quenching effect.[26,27]

Fig. 6. Intensity decaying with time of energy level of Pr:CNGG crystals.
4. Conclusions

The CNGG crystals doped with 0.5-, 1.0-, and 2.0-at.% Pr3+ ions are successfully grown by the μ-PD method. The doping of Pr3+ does not influence the lattice structure of CNGG. The spectral properties are studied at room temperature. The Pr:CNGG crystals have higher absorption coefficient and broader absorption band width at 449 nm, which are suitable for LD pumping. For the 0.5-at.% Pr:CNGG crystal, the orange emission band at 606 nm is more marked while the most intensive fluorescence band of 2.0-at.% Pr:CNGG crystal is centered at 654 nm in the red spectral region. In the NIR range, 1.0-at.% and 2.0-at.% Pr:CNGG crystals exhibit similar fluorescence ratios. The room temperature fluorescence lifetimes of energy level are fitted to be , , and for 0.5-, 1.0-, and 2.0-at.% Pr:CNGG crystals, respectively. Results of this work demonstrate that Pr:CNGG crystal owes good spectral properties and may be a potential candidate for solid-state laser operation.

Reference
[1] Chivian J S Case W E Eden D D 1979 Appl. Phys. Lett. 35 124 https://doi.org/10.1063/1.91044
[2] Koch M E Kueny A W Case W E 1990 Appl. Phys. Lett. 56 1083 https://doi.org/10.1063/1.103328
[3] Kück S Sokólska I Henke M Scheffler T Osiac E 2005 Phys. Rev. B 71 165112 https://doi.org/10.1103/PhysRevB.71.165112
[4] Lv S Z Wang Y Zhu Z J You Z Y Li J F Gao S F Wang H Y Tu C Y 2014 Appl. Phys. B 116 83 https://doi.org/10.1007/s00340-013-5651-7
[5] Esterowitz L Allen R Kruer M Bartoli F Goldberg L S Jenssen H P Linz A Nicolai V O 1977 J. Appl. Phys. 48 650 https://doi.org/10.1063/1.323648
[6] Liu W Zhang Q 2017 Mater. Res. Innovat. 21 65 https://doi.org/10.1080/14328917.2016.1183334
[7] Gün T Metz P Huber G 2011 Opt. Lett. 36 1002 https://doi.org/10.1364/OL.36.001002
[8] Macalik L Ma̧czka M Hanuza J Godlewska P Solarz P Ryba-Romanowski W Kaminskii A A 2008 J. Alloys Compd. 451 232 https://doi.org/10.1016/j.jallcom.2007.04.248
[9] Deopa N Rao A S Mahamuda S Gupta M Jayasimhadri M Haranath D Vijaya Prakash G 2017 J. Alloys Compd. 708 911 https://doi.org/10.1016/j.jallcom.2017.03.020
[10] Naito K Yokotani A Sasaki T Okuyama T Yamanaka M Nakatsuka M Nakai S Fukuda T Timoshechkin M I 1993 Appl. Opt. 32 7387 https://doi.org/10.1364/AO.32.007387
[11] Voronko Y K Sobol A A Karasik A Y Eskov N A Rabochkina P A Ushakov S N 2002 Opt. Mater. 20 197 https://doi.org/10.1016/S0925-3467(02)00063-0
[12] Voronko Y K Gessen S B Es’ kov N A Osiko V V Sobol A A Timoshechkin M I Ushakov S N Tsymbal L I 1992 Sov. J. Quantum Elect. 22 581 https://doi.org/10.1070/QE1992v022n07ABEH003545
[13] Shimamura K Sugiyama K Uda S Fukuda T 1995 Jpn. J. Appl. Phys. 34 4894 https://doi.org/10.1143/JJAP.34.4894
[14] Tsuboi T Tanigawa M Shimamura K 2000 Opt. Commun. 186 127 https://doi.org/10.1016/S0030-4018(00)01059-2
[15] Serrano M D ÁlvarezP érez J O Zaldo C Sanz J Sobrados I Alonso J A Cascales C FernándezD íaz M T Jezowski A 2017 J. Mater. Chem. C 5 11481 https://doi.org/10.1039/C7TC02760E
[16] Gheorghe L Greculeasa M Voicu F Gheorghe C Hau S Vlaicu A M Belikov K N Yu Bryleva E Gaiduk O V 2018 Opt. Mater. 84 335 https://doi.org/10.1016/j.optmat.2018.07.029
[17] Brenier A Boulon G Shimamura K Fukuda T 1999 J. Cryst. Growth 204 145 https://doi.org/10.1016/S0022-0248(99)00151-7
[18] Kaminskii A A Belokoneva E L Butashin A V Kurbanov K Markosian A A Mill B V Nikolskaia O K Sarsikov S E 1986 Inorg. Mater 22 927 https://inis.iaea.org/search/search.aspx?orig_q=RN:19021726
[19] Cornacchia F Di Lieto A Tonelli M Richter A Heumann E Huber G 2008 Opt. Express 16 15932 https://doi.org/10.1364/OE.16.015932
[20] Reichert F Moglia F Marzahl D T Metz P Fechner M Hansen N O Huber G 2012 Opt. Express 20 20387 https://doi.org/10.1364/OE.20.020387
[21] Zhang L Xia Y Shen X Wei W 2019 Spectrochim Acta Part. A 206 454 https://doi.org/10.1016/j.saa.2018.08.024
[22] Liu B Shi J Wang Q Tang H Liu J Zhao H Li D Liu J Xu X Wang Z Xu J 2018 J. Lumin. 196 76 https://doi.org/10.1016/j.jlumin.2017.11.019
[23] Özen G Forte O Di Bartolo B 2005 J. Appl. Phys. 97 013510 https://doi.org/10.1063/1.1823577
[24] Wang Y Li J You Z Zhu Z Tu C 2010 J. Alloys Compd. 502 184 https://doi.org/10.1016/j.jallcom.2010.04.140
[25] Özen G Forte O Di Bartolo B Collins J M 2007 J. Lumin 125 223 https://doi.org/10.1016/j.jlumin.2006.08.033
[26] Kränkel C Marzahl D T Moglia F Huber G Metz P W 2016 Laser Photon. Rev. 10 548 https://doi.org/10.1002/lpor.201500290
[27] Yu D C Chen Q J Lin H H Wang Y Z Zhang Q Y 2016 Opt. Mater. Express 6 197 https://doi.org/10.1364/OME.6.000197