Effect of Hf4+ doping on structure and enhancement of upconversion luminescence in Yb:Tm:LiNbO3 crystals
Dai Li1, †, Liu Chunrui1, Han Xianbo1, Wang Luping2, Shao Yu1, Xu Yuheng3
College of Science, Harbin University of Science and Technology, Harbin 150080, China
School of Materials Science and Engineering, Harbin University of Science and Technology, Harbin 150040, China
Department of the Applied Chemistry, Harbin Institute of Technology, Harbin 150001, China

 

† Corresponding author. E-mail: daili198108@126.com

Project supported by Special Funds of Harbin Innovation Talents in Science and Technology Research, China (Grant No. 2015RQQXJ045) and Science Funds for the Young Innovative Talents of HUST, China.

Abstract

A series of Yb:Tm:LiNbO3 crystals doped with x mol% Hf4+ ions (x = 2, 4, and 6) were grown by the Czochralski method. The dopant occupancy and defect structure of Hf:Yb:Tm:LiNbO3 crystals were investigated by x-ray diffraction and infrared transmission spectra. The influence of Hf4+ ions concentration on UV–VIS–NIR absorption spectra of Hf:Yb:Tm:LiNbO3 crystals was discussed. The upconversion luminescence of Hf:Yb:Tm:LiNbO3 crystals was obtained under 980 nm excitation. Strong emissions were observed at 475 nm in the blue wavelength range and 651 nm in the red wavelength range. Remarkably, enhancement of the red and blue upconversion luminescence was achieved by tridoping Hf4+ ions.

1. Introduction

Recently, rare-earths (RE) ions doped optical materials with the characteristic of efficient upconversion luminescence from infrared (IR) to visible radiation[1,2] attract considerable attention due to their potential applications in solid-state lasers, optical communications, solar cell, biological fluorescence labels, and so on.[3,4] Upconversion emission is usually obtained in bulk materials with relatively low phonon energy,[5] which decreases multiphonon relaxation rate and increases the lifetimes of the intermediate metastable states.[6] LiNbO3 crystals with low phonon energy are promising matrices for upconversion luminescence due to their good mechanical performance, high stability, excellent electro-optic, acousto-optic, nonlinear optic, and piezoelectric performances.[79] To obtain efficient upconversion luminescence, RE ions such as Er3+, Tm3+, Nd3+, and Ho3+ doped LiNbO3 crystals have been studied for many years. Among them, Tm3+ ion is well known as an activator of blue (1G43H6 transition), red (1G43F4 and 3F33H6 transitions), and NIR (3H43H6 and 3F43H6 transitions) emissions.[1,10,11] It is noteworthy that Tm3+ ions can be applied to blue laser which exhibits applications in optical communication, color display, and so on. Although upconversion luminescence of Tm3+ ions can be obtained under 980 nm excitation, the absorption of Tm3+ ions at the pumping wavelength is weak. While, as a sensitizer, Yb3+ ion can strongly absorb the pumping light and transfer energy to Tm3+ ions, which increases the pumping light absorption efficiency and lowers the laser threshold.[10,12] The upconversion luminescence of Yb3+ and Tm3+ codoped LiNbO3 crystals has been reported by Xing et al.[13]

As we all know, the application of Yb:Tm:LiNbO3 crystal in frequency doubler and laser medium material is limited by the photorefractive damage and the problem can be overcome by doping anti-photorefractive ions such as Mg2+, Zn2+, Sc3+, In3+, Hf4+, and Zr4+ ions.[1419] It is worthy of notice that luminescence behavior can be improved by doping anti-photorefractive ions.[20] Recently, a few reports have researched the influence of anti-photorefractive ions on upconversion emission in RE3+ ions doped LiNbO3 crystals.[2022] Shur et al. found that the addition of ZrO2 enhanced the 470 nm emission of Tm:LiNbO3 in blue wavelength region.[21] Sun et al. reported that 1.5 mol% In2O3 enhanced the 1.5 μm emission but suppressed 550 nm green upconversion emission of Er:LiNbO3.[22] Hf4+ has been reported to be an effective ion to improve the anti-photorefractive property of LiNbO3, while the effect of Hf4+ on the optical properties of Yb:Tm:LiNbO3 crystals have not been reported.

In this paper, Hf:Yb:Tm:LiNbO3 crystals with varying Hf4+ concentrations (2 mol%, 4 mol%, and 6 mol%) were successfully grown by the Czochralski technique. The defect structure was analyzed by x-ray diffraction (XRD) and infrared (IR) transmission spectra. The UV–VIS–NIR absorption spectra of Hf:Yb:Tm:LiNbO3 crystals were discussed. The effect of Hf4+ ions on the upconversion luminescence of Hf:Yb:Tm:LiNbO3 crystals was investigated under 980 nm excitation. A new strategy to improve the properties of red and blue upconversion luminescence of Yb:Tm:LiNbO3 crystals is given in this paper.

2. Experimental details
2.1. Crystal growth

The congruent Hf:Yb:Tm:LiNbO3 crystals ([Li]/[Nb] = 0.946) were grown along the c axis using traditional Czochralski method in air. The doping concentrations of Yb3+ and Tm3+ ions were both 1 mol%. The doping concentrations of Hf4+ were 2 mol%, 4 mol%, and 6 mol%, respectively, which were referred to as Hf2, Hf4, and Hf6 for convenience. After precisely weighed, the raw materials with purity of 99.99% were thoroughly mixed for 24 h to grow high quality crystals. Subsequently, the raw materials were heated at 750 °C for 3 h to remove CO2 and then sintered at 1150 °C for 4 h to form poly-crystals in single crystal furnace. The optimum conditions of the crystals growth were selected as follows: the temperature gradient of 30 °C/cm−1, the rotating rate of 15–25 r/pm, and the pulling rate of 0.6–1.5 mm/h. After cooled down to room temperature with a speed of 50 °C/h, the crystals were polarized for 2 h at 1100 °C with the current density of 5 mA/cm2. The grown sample Hf2 is shown in Fig. 1. Finally, the samples were cut into Y-cut plates with the dimensions of 10 mm×20 mm×2 mm (x × z × y) from the central part of the crystal and polished to optical grade.

Fig. 1. (color online) The photo of sample Hf2: (a) top view, (b) front view.
2.2. Measurement

The lattice constant and crystal structure of the samples were characterized by a SHMADZU XRD-6000 diffractometer with Cu target. The angular range of the measurement is from 10° to 80° and the tube voltage/tube current is 40 kV/50 MA.

Since the O–H vibration is sensitive to the change of ionic environment, the position of O–H vibration absorption peak is used to analyze the defect structures of the doped crystals. The IR transmission spectra of the samples were measured at a resolution of 1 cm−1 by a Nicolet-710 FT-IR spectrometer with the effective range of 3400–3600 cm−1.

The UV–VIS–NIR absorption spectra of the samples were measured by a Varian Cary 5000 spectrometer. The experiment conditions are as follows: room temperature, the range of 300–3000 nm, the measurement step distance of 1 nm, and the scanning rate of 600 nm/min.

The upconversion luminescence spectra of Hf:Yb:Tm:LiNbO3 crystals were obtained under 980 nm excitation. The experimental system is shown in Fig. 2. A pump source which is obtained by 980 nm diode laser with excitation power of 600 mW excites the sample after focusing. Through the upconversion performance of the sample, the 980 nm laser converts into visible light. Furthermore, the upconversion luminescence spectra were recorded by Zolix-SBP300 grating spectrometer which is equipped with CR131 photomultiplier tube. The light signal is converted into an electrical signal and the intensity of the signal is amplified. Finally, the data is collected and the spectrum is displayed in the computer.

Fig. 2. (color online) The schematic diagram of upconversion luminescence experimental setup.
3. Results and discussion

XRD patterns of the Hf:Yb:Tm:LiNbO3 crystals and the grown pure LiNbO3 crystals are displayed in Figs. 3 and 4. It is obvious that there is no new diffraction peak compared with the pure LiNbO3 structure. Hence, the diffraction patterns of the samples are indexed as hexagonal phase of LiNbO3 crystal, which reflects that all grown crystals are single phase. According to the data of XRD, the lattice parameters were calculated by the least square method and listed in Table 1. The unit cell volumes were computed by the formula V = a2c cos30°. The lattice constants of the pure congruent LiNbO3 crystal is shown in Table 1 for comparison.

Fig. 3. (color online) XRD patterns of Hf:Yb:Tm:LiNbO3 crystals.
Fig. 4. (color online) XRD pattern of the pure congruent LiNbO3 crystal.
Table 1.

Lattice constants of Hf:Yb:Tm:LiNbO3 crystals.

.

It is clear that the unit cell volumes of the LiNbO3 crystals change with the concentration of Hf4+ ions. Since the concentration of Li+ is less than that of Nb5+ in pure congruent LiNbO3 crystal, the anti-site Nb ( ) and Li vacancy ( ) exist. As is known to us, three factors affect the unit cell volume: the concentration of , the polarization ability and ionic radius of the doping ions. When doping 2 mol% Hf4+, 1 mol% Yb3+, and 1 mol% Tm3+ ions, the unit cell volumes of LiNbO3 crystals decrease. The reasons are as follows. Hf4+, Yb3+, and Tm3+ ions substitute and in the form of , , and . Based on charge balance, , , , and defects need defects to form , , , and defects. The doping ions reduce the concentration of and increase the concentration of , , and , which leads to the decrease of concentration followed by the decrease of unit cell volume. Another reason is that the polarization ability of is higher than Nb5+ and . When the doping concentration of Hf4+ increases up to 4 mol%, Hf4+ ions replace the normal Li and Nb sites to constitute its own charge balance in the form of . Theoretically, the unit cell volume of LiNbO3 crystal should also decrease. However, the unit cell volume increases gradually in the experiment. The reason may be that ionic radius becomes the main factor affecting the unit cell volume with the doping concentration of Hf4+ increased and the ionic radius of Hf4+ (0.81 Å[23]) ions is larger than that of Li (0.69 Å[24]) and Nb (0.68 Å[24]) ions.

IR transmission spectra of Hf:Yb:Tm:LiNbO3 crystals with various Hf4+ concentrations are shown in Fig. 5, in which the position of OH absorption peak shifts to the lager wavenumber with the increase of Hf4+ concentration. The OH absorption peak of sample Hf2 is located at 3484 cm−1. While the OH absorption peaks of samples Hf4 and Hf6 are shifted to 3495 cm−1 and 3498 cm−1, respectively. It is known that there are many intrinsic defects including anti-site Nb ( ) and Li vacancy ( ) in congruent LiNbO3 crystals and the OH absorption peak of congruent LiNbO3 crystals is located at 3482 cm−1 owing to the OH stretching vibration of complex defect ( –OH).[25] Based on a large number of reports, the OH absorption peak of doped LiNbO3 crystals will obvious shifts from about 3482 cm−1 of pure congruent LiNbO3 crystals to lager wavenumber when the doping concentration exceed its thresholds.[2527] The OH absorption peak at 3495 cm−1 indicates that the 4 mol% Hf4+ doping in Yb:Tm:LiNbO3 crystal is up to the threshold concentration. This is similar to the previously reported threshold concentration in Hf:LiNbO3 crystals.[26] In sample Hf2, the Hf4+, Yb3+, and Tm3+ ions enter into LiNbO3 crystal by replacing defects to form , , and defects, which repel H+ ions. As a result, the OH absorption peak of sample Hf2 still reflects the OH vibration around defects and is located at 3484 cm−1. The slight shift is caused by the reducing of and defects. When the doping level of Hf4+ exceeds its threshold in samples Hf4 and Hf6, defects are replaced completely and partial Hf4+ ions begin enter the normal Li and Nb sites in the form of and defects. Since defects have a stronger force to attract H+ than , H+ ions aggregate around to form –OH defects. Meanwhile more energy is required for OH stretching vibration. So the absorption peak that reflects the OH vibration around shows blue shift.

Fig. 5. (color online) IR transmission spectra of Hf:Yb:Tm:LiNbO3 crystals.

The UV–VIS–NIR absorption spectra of Hf:Yb:Tm:LiNbO3 crystals are shown in Fig. 6. Besides the absorption band centered at 980 nm caused by 2F7/22F5/2 transition of Yb3+ ions, the other absorption bands owe to the electronic transition of Tm3+ ions. The absorption bands centered at 703 nm, 795 nm, and 1215 nm are attributed to the transitions from the ground state 3H6 to the excited states 3F2, 3, 3H4, and 3H5. And the absorption bands centered at 1666 nm and 1757 nm both correspond to the transition from 3H6 to 3F4 due to the stark sublevel absorption caused by the crystal-field splitting in Hf:Yb:Tm:LiNbO3 crystals. In addition, the absorption band centered at 2352 nm is attributed to the 3H53H4 transition of Tm3+ ions. Among these absorption bands, 980 nm which is well adapted for emission wavelength of commercial diode laser can be selected as the pump wavelength. By comparison of the absorption spectra, the following characteristics are identified. Firstly, the absorption spectra of samples Hf2, Hf4, and Hf6 are similar in the shape and band position, except disappearance of absorption band centered at 2352 nm in sample Hf6. Since the Tm3+ doped crystal has a potential application for 2–2.5 μm laser, the high doping Hf4+ ions will eliminate the absorption at this wavelength and be benefit for the application. Secondly, the intensities of the absorption bands first increase and then decrease with the increase of Hf4+ ions concentration.

Fig. 6. (color online) The UV–VIS–NIR absorption spectra of the Hf:Yb:Tm:LiNbO3 crystals.

Upconversion luminescence spectra of Hf:Yb:Tm:LiNbO3 crystals in the visible wavelength range excited by 980 nm laser diode are shown in Fig. 7. The strongest intensity peak is centered at 475 nm in the blue wavelength range. This peak is attributed to the 1G43H6 transition of Tm3+ ions. Another peak is centered at 651 nm in the red wavelength range and attributed to the 1G43F4 transition. Compared with sample Hf2, the intensity of blue and red emissions increases obviously in samples Hf4 and Hf6. Especially, in sample Hf6, the blue emission is about 4.7 times and the red emission is about 1.9 times more than that in sample Hf2. It is obvious that the increase of Hf4+ doping concentration is beneficial for upconversion luminescence of Hf:Yb:Tm:LiNbO3 crystals.

Fig. 7. (color online) Room-temperature upconversion luminescence spectra of Hf:Yb:Tm:LiNbO3 crystals: (a) 400–550 nm, (b) 550–700 nm.

As is well known, the intensity of upconversion luminescence is closely related to the local environment, the doping concentration, and the distribution of active ions in a host material.[20] In consideration of the fixed Tm3+ and Yb3+ ions concentration, the change of the intensity of upconversion luminescence is caused by the local environment of Tm3+ and Yb3+ ions adjusted by Hf4+ ions. One reason for the enhancement is the minimization of structural defects in crystals. The atomic radius difference of Li (1.54 Å) and Nb (1.48 Å) is more remarkable than Hf (1.59 Å). Hence, doping Hf4+ ions that reduce defects and replace defects to form will minimize the structural defects of the crystal. Another reason is explained as follows. As we have known, three and two photons are involved in the blue and red upconversion emissions, respectively. The energy transfer mechanism of upconversion process in Hf:Yb:Tm:LiNbO3 crystals is shown in Fig. 8. In the upconversion process, Yb3+ ions are excited from the ground state 2F7/2 to the 2F5/2 state by absorbing 980 nm NIR photon via ground state absorption (GSA). Then the energy transfers from Yb3+ to Tm3+ ions followed by populating 3H5, 3F2, and 1G4 intermediate states of Tm3+ ions. The 3H53F4 and 3F23H4 transitions of Tm3+ ions are achieved by nonradiative relaxation. Finally, blue and red upconversion emissions are obtained though the 1G43H6 and 1G43F4 transitions respectively in radiative relaxation progress. However, the existence of cross relaxation (CR) processes (3H4+3H63F4+3F4) will suppress the population of 3H4.[28] Because the excited state absorption (ESA) which is an inherent result of a given pump choice is hard to change and CR processes can be suppressed by decrease of Tm3+ cluster sites.[2931] The enhancement of upconversion emission indicates the suppression of CR processes and reduction of Tm3+ cluster sites, which is caused by high doping Hf4+ ions. In conclusion, Hf4+ ions can improve the luminescence environment and enhance the intensity of blue and red emissions in Hf:Yb:Tm:LiNbO3 crystals.

Fig. 8. (color online) The energy levels of Yb3+ and Tm3+ ions as well as the upconversion luminescence mechanism of Hf:Yb:Tm:LiNbO3 crystals.
4. Conclusion

In this work, Hf:Yb:Tm:LiNbO3 crystals with variable Hf4+ concentrations have been successfully grown by the Czochralski method. The IR transmission spectrum indicates that defects are replaced completely and Hf4+ ions begin enter the normal Li and Nb sites when the Hf4+ doping concentration is up to 4 mol%. The UV–VIS–NIR absorption spectra first increase and then decrease with the increase of Hf4+ ions concentration and display strong absorption peaks at 980 nm. Under excitation at this wavelength, the blue (475 nm) and red (651 nm) upconversion luminescences are achieved though the 1G43H6 and 1G43F4 transitions of the Tm ions. The blue emissions intensity of 6 mol% Hf4+ tridoped Yb:Tm:LiNbO3 crystal is about 4.7 times larger than that of 2 mol% Hf4+ tridoped Yb:Tm:LiNbO3 crystal and the red emission intensity enhances by 1.9 times. The upconversion luminescence spectra remarkably increase with the increase of Hf4+ doping concentration. The generation and enhancement of the blue and red upconversion luminescence in Hf:Yb:Tm:LiNbO3 crystals are interesting for solid-state lasers and color display.

Reference
[1] Mokoena T P Linganiso E C Kumar V Swart H C Cho S H Ntwaeaborwa O M 2017 J. Colloid Interface Sci. 496 87
[2] Dai L Liu C R Tan C Yan Z H Xu Y H 2017 Chin. Phys. 26 044207
[3] Fischer S Goldschmidt J C Loper P Bauer G H Bruggemann R Kramer K Biner D Hermle M Glunz S W 2010 J. Appl. Phys. 108 044912
[4] Lu H C Yi G S Zhao S Y Chen D P Guo L H Cheng J 2004 J. Mater. Chem. 14 1336
[5] Mialon G Türkcan S Dantelle G Collins D P Hadjipanayi M Taylor R A Gacoin T Alexandrou A Boilott J P 2010 J. Phys. Chem. 114 22449
[6] Joubert M F 1999 Opt. Mater. 11 181
[7] Tumuluri A Bharati M S S Rao S V Raju K C J 2017 Mater. Res. Bull. 94 342
[8] Wu L Ling F R Zuo Z G Liu J S Yao J Q 2012 Chin. Phys. 21 017802
[9] Zhang D L Qiu C X Wong W H Pun E Y B 2014 Mater. Res. Bull. 60 771
[10] Wang Z J Li X Z Wei Q Long X F 2008 Mater. Res. Innov. 12 174
[11] Mahalingam V Naccache R Vetrone F Capobianco J A 2012 Opt. Express 20 111
[12] Xing L L Yang W Q Lin J C Huang M Xue Y Q 2017 Sci. Rep. 7 14725
[13] Xing L L Xu Y L Wang R Xu W Zhang Z G 2014 Opt. Lett. 39 454
[14] Xu Z P Bi W H Jiao B L Ben Y Z Lu Q Z Tang H H 2010 Cryst. Res. Technol. 45 85
[15] Sidorov N V Palatnikov M N Yanichev A A Titov R A Makarova O V 2017 Tech. Phys. 62 417
[16] Zhen X H Li Q Li H T Xu Y H Zhao L C 2005 Cryst. Res. Technol. 40 649
[17] Zhang T Wang B Fang S Q Ma D C Zhao Y Q Xu Y H 2004 Mater. Lett. 58 3074
[18] Sun X D Shi H X Luo S H Meng Q X Jiang Y Y 2010 Chin. Phys. 19 094209
[19] Dai L Jiao S S Xu C Li D Y Lin J Q Chen C T 2014 Mater. Res. Bull. 53 132
[20] Qian Y N Zhang Z Y Tang X Z Ivanov M Zhang H Y Wu Q B 2017 Opt. Mater. 70 116
[21] Shur J W Choi K H Yoon D H 2011 J. Cryst. Growth 318 653
[22] Sun L Yang C H Li A H Xu Y H Zhao L C 2009 J. Appl. Phys. 105 043512
[23] Lorenzo A Jaffrezic H Roux B 1995 Appl. Phys. Lett. 67 3735
[24] Mignotte C 2001 Appl. Surf. Sci. 185 11
[25] Xu Z P Xu S W Xu Y H Wang R 2005 Mod. Phys. Lett. 19 1277
[26] Li S Q Liu S G Kong Y F Deng D L Gao G Y Li Y B Gao H C Zhang L Hang Z H Chen S L Xu J J 2006 J. Phys.: Condens. Matter 18 3527
[27] Zhang T Wang X Geng T Tong C G Kang C 2015 J. Alloys Compd. 629 255
[28] Jackson S D Mossman S 2003 Appl. Opt. 42 2702
[29] Qian Y N Wang R Xing L L Xu Y L Yang C H Liu X R 2011 Cryst. Res. Technol. 46 1137
[30] Renard S Camy P Doualan A B J L Moncorge R 2008 J. Alloys Compd. 451 71
[31] Jackson S D 2004 Opt. Commun. 230 197