Next Article in Journal
A Portable Miniature Cryogenic Environment for In Situ Neutron Diffraction
Previous Article in Journal
Effect of Solid/Liquid and Eutectic Front Velocities on Microstructure Evolution in Al-20%Cu Alloys
Previous Article in Special Issue
Thermal Field Simulation and Optimization of PbF2 Single Crystal Growth by the Bridgman Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on Spectral Properties and Mid-Infrared Laser Performance of Er, La:CaF2 Crystals

by
Zhen Zhang
1,*,
Jingjing Liu
2,
Yunfei Wang
1,
Fengkai Ma
3,
Shaochen Liu
1,
Zhonghan Zhang
1,
Jie Liu
2 and
Liangbi Su
1,*
1
Synthetic Single Crystal Research Center, CAS Key Laboratory of Transparent and Opto-Functional Inorganic Materials, Shanghai Institute of Ceramics Chinese Academy of Sciences, Shanghai 201899, China
2
Shandong Provincial Key Laboratory of Optics and Photonic Device, School of Physics and Electronics, Shandong Normal University, Jinan 250014, China
3
Guangdong Provincial Engineering Research Center of Crystal and Laser Technology, Department of Optoelectronic Engineering, Jinan University, Guangzhou 510632, China
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(7), 639; https://doi.org/10.3390/cryst14070639
Submission received: 5 June 2024 / Revised: 1 July 2024 / Accepted: 3 July 2024 / Published: 11 July 2024
(This article belongs to the Special Issue Photoelectric Functional Crystals)

Abstract

:
Er3+-doped fluorite crystals, including CaF2 and SrF2, are considered as attractive laser gain materials in the mid-infrared (MIR) region with merits of high laser efficiency as well as low doping concentration. In this work, a series of Er, La:CaF2 crystals were grown and the modulation effect of co-doping La3+ ions on the spectral properties and mid-infrared laser performance was investigated. It was found that introducing La3+ ions can effectively manipulate the coordination environment of Er3+ ions embedded in CaF2 crystal, thus modulating the shape and intensity of absorption and emission bands. On the other hand, La3+ ions can partially substitute Er3+ sites in the clusters to form mixed clusters, which affects the energy transfer processes between Er3+ ions as well as ~3 μm laser performance, which is dominated by energy transfer up-conversion (ETU) processes between Er3+ ions. By co-doping La3+ ions into Er:CaF2 crystal at an appropriate concentration, the spectral parameter modulation can be achieved while maintaining a high MIR laser efficiency.

1. Introduction

A near 3 μm laser based on laser diode (LD)-pumped Er-doped gain materials has attracted increasing attention for significant applications in cosmetic medicine, soft tissue ablation, and gas sensing [1,2,3]. More importantly, it can be used directly as a pump source for Fe2+:ZnSe to achieve lasing at a wavelength range of 4~5 μm [4], which is critical for the advanced infrared countermeasure system. Compared to optical parameter oscillation (OPO) or optical parameter amplification (OPA) schemes, the Er laser-pumped Fe2+:ZnSe system possesses higher integration and improved robustness, and more importantly its output power is able to scale to the hundred-watts level with Master Oscillator Power-Amplifier (MOPA) technology [5]. Thus, there have been consistent and increasing efforts to realize a high-power or high-energy Er-based near 3 μm laser in the past decades.
It is well known that near 3 μm laser oscillation derived from Er3+: 4I11/24I13/2 transition is self-terminated because of the longer lifetime of the upper energy level 4I13/2 than that of the lower energy level 4I11/2 [6]. To overcome this unfavorable effect, high concentration doping is adopted for Er-doped gain materials in order to take advantage of the ETU1 process (4I13/2 + 4I13/24I9/2 + 4I15/2) which can accelerate depopulation of 4I13/2, and meanwhile recycle the excitation energy of 4I13/2 to multiply the population of 4I11/2 [7,8], as illustrated in Figure 1a. On the other hand, introduction of numerous dopants will degrade the thermal conductance of laser gain crystals [9,10] and result in severe thermal lens effect due to the intense absorption of the pump light, which hinders the power scaling of Er-doped crystals [11,12].
Er3+-doped fluorite crystals, including CaF2 and SrF2, stand out for the unique self-aggregation behavior or clustering effect [13] of the trivalent rare earth (RE3+) ions embedded in the crystals (shown in Figure 1b), which enables this kind of gain crystals to achieve high laser efficiency at a rather lower doping level. Benefiting from the ionic clustering-induced enhancement of the ETU1 process, Fan et al. and Liu et al. realized a highly efficient ~2.8 μm continuous-wave (CW) laser with a slope efficiency above 40% by 976 nm LD end-pumped Er:SrF2 and Er:CaSrF2 crystals [14,15]; this efficiency surpasses the Stokes limitation and still keeps the highest record for LD-pumped Er laser in the 3 μm wavelength range. In addition, a CW MIR laser with hundreds of milliwatts output power was also obtained in extremely light-doping (0.3 at.%) Er:CaF2 crystals [16], which breaks the long-standing consideration that a ~2.8 μm CW laser is unable to run in low concentration Er-doped materials. Based on the merit of low doping concentration, a 5.04 W CW laser at 2799 nm was obtained in a 2 at.% Er:CaF2 crystal [17]. The first semi-conductor saturable absorber mirror (SESAM) mode-locked laser operation of Er-doped bulk materials was realized in Er:CaF2-SrF2 mixed crystal [18].
The ~3 μm laser performance of Er-doped crystals is dominated by two different energy transfer up-conversion processes, that is, ETU1: 4I13/2 + 4I13/24I9/2 + 4I15/2 and ETU2: 4I11/2 + 4I11/24F7/2 + 4I15/2; the former is desired and the latter is adverse to the laser oscillation. These two processes both occur with lasing and their competitive strength varies with the Er3+ coordination environment [19]. The strength regulation of these two ETU processes is generally realized by changing the doping concentration of Er3+ ions in Er:LiYF4 [20], Er:YSGG [21], Er:BaY2F8 [22], and so on. In contrast, for Er3+-doped fluorite crystals, introducing inactive rare earth ions to manipulate the cluster structure should be another more effective way to achieve this regulation. In fact, co-doping optically inert RE3+ ions (such as Y3+, La3+, Gd3+, Lu3+ etc.) has been demonstrated to modulate the site structure and spectral properties of Nd3+ ions embedded in CaF2, SrF2, and their mixed crystals [23,24,25]. However, most works are focused on an Er3+ singly doped CaF2/SrF2 system at present, and the effect of co-doping site-modulated ions remains unknown.
In this work, a series of Er3+, La3+ co-doped CaF2 crystals were grown to investigate the modulation effect of co-doping La3+ ions on the spectral properties and MIR laser performance. It was found that introduction of La3+ ions into Er:CaF2 crystal could change the site structure of Er3+ ions within the ionic clusters, reshaping the absorption/emission band and enhancing the emission cross-section and lifetimes of ~3 μm laser energy levels. In particular, a moderate La3+ concentration such as 2 at.% can make improvements that do not impede ETU processes, resulting in a low laser threshold and high laser efficiency.

2. Materials and Methods

The 5 at.% Er, x at.% La:CaF2 (x = 0, 2, 5, 10) single crystals were successfully grown by temperature gradient method. Crystalline ErF3, LaF3, and CaF2 with a purity of 99.99% were ground and well mixed according to a stoichiometric ratio. An additional 1 wt.% PbF2 powder was added to serve as oxygen scavenger. The mixed powder was then loaded into a muti-hole graphite crucible with the caps dilled a Φ1 mm hole in the center. At first, the vacuum in the sealed furnace chamber was pumped to be lower than 1 × 10−3 Pa. Then, the crucible was heated to 200 °C at the rate of 30 °C per hour and the temperature was kept constant for 5 h to get rid of the absorbed water of the raw powder. The crucible was then heated to 800 °C and kept for 12 h to eliminate the oxygen in the chamber through the reaction of PbF2 + O2 → PbO + F2. The crucible was further heated to 1420 °C and maintained for 15 h to ensure the homogeneity of the melt. After that, the temperature declined at a cooling rate of 0.8 °C/h during the period of crystallization, and finally the crucible was cooled down to room temperature at a rate of 15 °C/h. Crystal chips with a thickness of 1 mm and blocks with a cross-section of 3 × 3 mm2 and a length of 10 mm were cut from the individual as-grown boules, then the end faces of the plates and blocks were finely polished. The thin chips were prepared for spectral measurements and the crystal blocks were used as gain elements in laser experiments.
The powder X-ray diffraction (PXRD) patterns were collected by a D8 ADVANCE high-resolution X-ray diffractometer (Bruker, Karlsruhe, Germany). The real doping concentrations of Er3+ and La3+ ions were determined by inductively coupled plasma atomic emission spectroscopy (ICP-AES, Agilent 725). Absorption spectra were measured using a Cary 5000 UV/VIS/IR spectrophotometer (Agilent, Santa Clara, CA, USA) with an integration time of 0.2 s and a step of 0.1 nm. Emission spectra and fluorescence decay measurements were carried out by an FLS 1000 fluorescence spectrometer (Edinburgh Instruments, Livingston, UK) where the excitation sources included a steady-state xenon lamp, a microsecond flash lamp, and a 980 nm semiconductor laser which can work in CW and pulsed mode. The detectors used involved visible and infrared photomultiplier tubes (PMTs) and a liquid nitrogen cooled InSb infrared photodiode. All measurements were carried out at room temperature.
The laser experiments were carried out with a linear plano-concave resonator as depicted in Figure 2. A fiber-coupled 976 nm LD was used as pump source. The core diameter of the fiber was 105 μm with a numerical aperture (NA) = 0.22. The pump beam was collimated and focused by a couple of convex lenses with the expansion ratio of 1:2. The input mirror (IM) was a concave lens with curvature radius of −50 mm and coated with high transmittance (HT) for the pump beam and high reflectivity (HR) at 2.7~2.95 μm. A series of plane lenses with different transmittance of 1%, 2%, and 3% for laser beam were utilized as the output mirror (OM). The laser beam was reflected to the power meter by a dichroic mirror with HR for the laser beam and HT for the pump beam so the laser output power could be measured.

3. Results and Discussion

Figure 3a shows the as-grown 5 at.% Er, 2 at.% La:CaF2 crystal ingot which was transparent and without any cracks or inclusions. For fluorite crystals, the cellular structure is apt to occur especially in the tail of the ingots; the chips and blocks shown in Figure 3c, d were selected from the homogeneous parts of the as-grown boules which are lacking the cellular structure. The actual concentrations of dopants were obtained by ICP-AES tests and the results are summarized in Table 1. There are slight deviations between the nominal and actual doping concentrations because of segregation during crystal growth. To check the phase purity of the crystals, PXRD measurements were performed and the results are presented in Figure 4a. All the diffraction peaks belong to the CaF2 phase, which demonstrates that the fluorite structure can be retained with even more than 15 at.% RE3+ dopants introduced. However, it was found that the series peaks shift to low angle with an increase in La3+ dopants, indicating that lattice expansion occurs when La3+ (1.30 Å) ions substitute original sites occupied by Ca2+ (1.26 Å) ions as well as when the equivalent amount of F ions enters the interstitial site. Many calculations and experimental results reported indicate that the introduced RE3+ ions and charge-compensated interstitial F in CaF2 will congregate spontaneously and form locally disordered ionic clusters under the combined effect of Coulomb force and lattice stress [26,27]. In Figure 4b, it can be seen that the variation of lattice constant on La concentration failed to follow Vegard’s law well, maybe due to the inconstant Er doping concentration.
The spectral characteristics were systematically investigated to analyze the modulation effect of co-doping La3+ ions. Figure 5 shows the absorption cross-sections (σabs) corresponding to Er3+: 4I15/24I11/3 and 4I13/2 transitions. The former transition can be directly excited by commercial ~980 nm LDs, thus it is paid more attention. For the singly doped crystal it can be seen that this absorption band consists of two peaks centered at 967.3 nm and 981.2 nm, and each peak has a shoulder which emerges at 970 nm and 979 nm, respectively. With La3+ dopants increasing, the 967.3 nm peak shows slight red-shift and the σabs is significantly reduced by 40%, from 2.28 × 10−21 cm2 to 1.44 × 10−21 cm2, which reduces to be lower than that of the shoulder for 10 at.% La3+ co-doped crystal. On the other hand, for the peak at 981.3 nm, a more obvious blue-shift occurs with La3+ concentration; the σabs shows a small increase at 2 at.% La3+ concentration, then decreases by about 20% (from 2.09 × 10−21 cm2 to 1.70 × 10−21 cm2), and the original shoulder becomes the new peak. Finally, the two peaks converge into one with the peak center at 975.9 nm. There are two factors that can account for the absorption band evolution: One is that the fraction of the active centers corresponding to the two original absorption peaks declines with La3+ concentration; in addition, La3+ and Er3+ ions doped in CaF2 crystal will form mixed clusters where the strength of crystal field applied to Er3+ ions becomes weaker than that in pure Er3+ clusters, leading to a reduction of Stark level splitting of Er3+: 4I11/2. The similar change also occurs to the absorption band at near 1550 nm: the σabs first slightly increases at 2 at.% La3+ concentration and then declines, accompanied with the gradual convergence of the original structured band.
The emission cross-sections (σem) near 2.8 μm were calculated according to the Judd–Ofelt theory and the Füchtbauer-Ladenburg formula [28], and are illustrated in Figure 6. The σem at 2729 nm also shows a rise for 2 at.% La3+ co-doped sample (from 7.3 × 10−21 cm2 to 7.8 × 10−21 cm2), then a decline with La3+ concentration increasing. The broad emission band in the long wavelength range overall decreased, which results in the width of this emission band narrowing.
We further measured the lifetimes of Er3+: 4I11/2 and 4I13/2 levels. The thin samples were excited by a 980 nm semiconductor laser and the pulse width was about 50 μs. To prevent the influence of self-absorption on the intrinsic fluorescence decay, the samples for lifetime measurements were processed into the thin chips with a thickness of 1 mm, and the signals were monitored at 1002 nm and 1570 nm, which is away from the corresponding absorption peaks. The decay curves of Er3+: 4I11/2 level are summarized in Figure 7a; they all can be well-fitted with a single exponential function, and the fitted lifetimes are 6.7 ms, 7.5 ms, 7.8 ms, and 7.9 ms, with increase of La3+ concentration. For the decay curves of 4I13/2, because of the population compensation by Er3+: 4I11/24I13/2 transition under 980 nm excitation, there is an obvious rising edge in the first 10 ms of decay curves, as shown in Figure 7b. These curves were fitted by the formula put forward in Ref [29] and the lifetimes were calculated to be 11.1 ms, 14.3 ms, 14.5 ms, and 15.3 ms for 0 at.%, 2 at.%, 5 at.%, and 10 at.% La3+-doped samples. The lifetimes were prolonged with 2 at.% La3+ introduced, and the rise in lifetime becomes smaller when further increasing La3+ concentration.
As mentioned before, introducing La3+ ions changes the composite of the original RE3+ clusters, which influences the coordinate structure of Er3+ ions residing in the clusters; on the other hand, the pure Er3+-clusters could be partially “broken up”, leading to the cooperative up-conversion between Er3+ ions being weakened. In order to investigate the “weakening effect” of La3+ dopants, the decay curves of 4S3/2 were measured. It has been demonstrated that the fluorescence originated from the 4S3/2 level is sensitive to Er3+ concentration and easy to quench by Er3+: 4S3/2 + 4I15/24I11/2 + 4I13/2 process [30], thus the lifetime of 4S3/2 can reflect the interaction strength between Er3+ ions. The results are shown in Figure 7c; all the decay curves fit well with the single exponential function and the fitted lifetimes were 16.8 μs, 17.4 μs, 21.4 μs, and 26.5 μs, as summarized in Figure 7d. With the La3+ concentration increasing, the green fluorescence of 4S3/2 with a radiation lifetime of several hundred microseconds [31] is seriously quenched for all crystal samples and the rise of the lifetime is inconspicuous with La3+ concentration. It is worth noting that for the 2 at.% La3+-doped sample, the lifetime increase is negligible, which is not consistent with that at ~1 μm or ~1.5 μm. This indicates that introducing La3+ into 5 at.% Er:CaF2 crystal only has a moderate influence on the interionic energy transfer processes; the spectral modulation, including reshaping the absorption/emission band and increasing emission cross-sections and lifetimes of laser energy levels, is mainly attributed to the change of the site structures surrounding the embedded Er3+ ions. In fact, it has been demonstrated that La3+ ions tend to aggregate with cubic coordination in CaF2 crystal, whereas Er3+ ions prefer to form clusters with tetragonal antiprism sublattice [32]. Therefore, when La3+ ions partially substitute Er3+ sites within clusters to form mixed ones, the extra lattice distortion with the structural characteristics of La3+ clusters will occur and achieve the local structure manipulation.
We also tested the laser performance of the Er3+, La3+ double doping crystals based on the setup previously mentioned, and the results are presented in Figure 8. For 5 at.% Er, 2 at.% La:CaF2 crystal, higher laser efficiencies were obtained when T = 2% and 3% OMs were utilized. The highest slope efficiency was 36.6% and the maximum output power was 0.92 W with a T = 3% OM. When using a 2% T OM, the slope efficiency was 35.6% with an output power of 0.9 W. The laser thresholds, increasing with the transmittance of the OM, were 31 mW, 98 mW, and 153 mW for the OM with T = 1%, 2%, and 3%. When La3+ concentration increased to 5 at.%, neither the slope efficiency or the maximum output power degraded; the highest slope efficiency was 29.4% with an output power of 0.435 W for T = 3% OM, and the laser threshold became higher, 146 mW for T = 1% OM. The degraded laser performance is partially due to the weakened ETU process, and the lower thermal conductivity with a higher doping concentration, resulting in more severe thermal effect that also impedes efficient laser operation. For the 5 at.% Er, 10 at.% La:CaF2 crystal, it is difficult to stabilize the laser oscillation because of the serious thermal effect. We also failed to perform the laser experiment on 5 at.% Er:CaF2 crystal due to the undesired crystal quality; however, the laser performance of 5 at.% Er:CaF2 crystal reported in Ref. [33] can serve as a reference. As shown in Table 2, the laser power and slope efficiency were comparable for 5 at.% Er, 2 at.% La:CaF2 and 5 at.% Er,:CaF2, which demonstrated that introduction of moderate La3+ dopants would not degrade laser performance for 5 at.% Er,:CaF2.

4. Conclusions

This work demonstrates a method to modulate the spectral parameters concerning ~3 μm MIR laser performance of Er3+-doped fluorite crystals. With La3+ concentration increasing, the absorption and emission bands are reshaped and the structured bands converge; meanwhile, the emission cross-sections and laser energy lifetimes are enhanced. By analysis of the fluorescence decay behavior of the quenching energy level 4S3/2, it is revealed that co-doping La3+ ions would not have a significant influence on the energy transfer process between Er3+ ions, and manipulation of the site structure of RE3+ ionic clusters plays the dominant role. An efficient CW MIR laser with a slope efficiency of 36.6% and an output power of 0.9 W is achieved in 5 at.% Er, 2 at.% La:CaF2 crystal. Further increasing La3+ concentration would degrade the laser performance due to a weakened ETU process and reduced thermal conductivity. Therefore, the spectral parameters of Er-doped fluorite crystals can be modulated through optimizing the type or concentration of the co-doped optically inert RE3+ ions.

Author Contributions

Conceptualization, L.S. and Z.Z. (Zhen Zhang); methodology, Z.Z. (Zhen Zhang), J.L. (Jingjing Liu), and F.M.; validation, Z.Z. (Zhen Zhang), J.L. (Jingjing Liu), and Y.W.; formal analysis, S.L. and Z.Z. (Zhonghan Zhang); investigation, J.L. (Jingjing Liu) and J.L. (Jie Liu); resources, L.S.; data curation, L.S.; writing—original draft preparation, Z.Z. (Zhen Zhang); writing—review and editing, F.M.; visualization, Y.W. and Z.Z. (Zhonghan Zhang); supervision, L.S. and J.L. (Jie Liu); project administration, Z.Z. (Zhonghan Zhang); funding acquisition, L.S. and Z.Z. (Zhen Zhang). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program of China (2023YFB3507402), the National Natural Science Foundation of China (61925508, 62205359), the CAS Project for Young Scientists in Basic Research under Grant (YSBR- 024), and the Open Fund of the Guangdong Provincial Key Laboratory of Optical Fiber Sensing and Communications (Jinan University) (2022GDSGXCG01).

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors would like to thanks Fei Tang and Xiaobo Qian for their help in crystal growth and Yuexiang Zhao for her help in crystal polishing.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kaufmann, R.; Beier, C. Laser skin ablation: An update on aesthetic and medical indications. Med. Laser Appl. 2004, 19, 212–222. [Google Scholar] [CrossRef]
  2. Parker, S. Surgical lasers and hard dental tissue. Br. Dent. J. 2007, 202, 445–454. [Google Scholar] [CrossRef]
  3. Walsh, B.M.; Lee, H.R.; Barnes, N.P. Mid infrared lasers for remote sensing applications. J. Lumin. 2016, 169, 400–405. [Google Scholar] [CrossRef]
  4. Myoung, N.; Martyshkin, D.V.; Fedorov, V.V.; Mirov, S.B. Energy scaling of 4.3 μm room temperature Fe:ZnSe laser. Opt. Lett. 2011, 36, 94–96. [Google Scholar] [CrossRef]
  5. Do, B.; Skolnik, L.; Skolnik, J.; Nguyen, T. Challenges in designing and building a high energy, pulsed diode-pumped MWIR Fe:ZnSe laser. In Proceedings of the Solid State Lasers XXXII: Technology and Devices, San Francisco, CA, USA, 31 January–1 February 2023; SPIE: Bellingham, WA, USA, 2023; Volume 12399, pp. 57–68. [Google Scholar]
  6. Zharikov, E.V.; Zhekov, V.I.; Kulevskii, L.A.; Murina, T.M.; Osiko, V.V.; Prokhorov, A.M.; Savel, A.D.; Smirnov, V.V.; Starikov, B.P.; Timoshechkin, M.I. Stimulated emission from Er3+ ions in yttrium aluminum garnet crystals at λ = 2.94 μ. Sov. J. Quantum Electron. 1975, 4, 1039. [Google Scholar] [CrossRef]
  7. Bagdasarov, K.S.; Zhekov, V.I.; Lobachev, V.A.; Murina, T.M.; Prokhorov, A.M. Steady-State Emission from a Y3Al5O12: Er3+ Laser (λ = 2.94 μ, T = 300 K). Sov. J. Quantum Electron. 1983, 13, 262. [Google Scholar] [CrossRef]
  8. Stoneman, R.C.; Lynn, J.G.; Esterowitz, L. Laser-pumped 2.8-µm Er3+:GSGG laser. In Proceedings of the Conference on Lasers and Electro-Optics, San Diego, CA, USA, 9–13 December 1991; Optical Society of America: Washington, DC, USA, 1991; p. CTuO6. [Google Scholar]
  9. Popov, P.A.; Fedorov, P.P.; Osiko, V.V. Thermal conductivity of single crystals with a fluorite structure: Cadmium fluoride. Phys. Solid State 2010, 52, 504–508. [Google Scholar] [CrossRef]
  10. Popov, P.A.; Fedorov, P.P.; Osiko, V.V. Thermal conductivity of single crystals of the Ca1−xYxF2+x solid solution. In Doklady Physics; Pleiades Publishing: Warrensburg, MI, USA, 2014; Volume 59. [Google Scholar]
  11. Pollnau, M. Analysis of heat generation and thermal lensing in erbium 3-μm lasers. IEEE J. Quantum Electron. 2003, 39, 350–357. [Google Scholar] [CrossRef]
  12. Dergachev, A.; Moulton, P.F. Tunable CW Er:YLF diode-pumped laser. In Proceedings of the Advanced Solid-State Photonics, San Antonio, TX, USA, 2–5 February 2003; Optical Society of America: Washington, DC, USA, 2003; p. 3. [Google Scholar]
  13. Orlovskii, Y.V.; Basiev, T.T.; Osiko, V.V.; Gross, H.; Heber, J. Fluorescence line narrowing (FLN) and site-selective fluorescence decay of Nd3+ centers in CaF2. J. Lumin. 1999, 82, 251–258. [Google Scholar] [CrossRef]
  14. Fan, M.; Li, T.; Zhao, J.; Zhao, S.; Li, G.; Yang, K.; Su, L.; Ma, H.; Kränkel, C. Continuous wave and ReS2 passively Q-switched Er:SrF2 laser at ~3 μm. Opt. Lett. 2018, 43, 1726–1729. [Google Scholar] [CrossRef]
  15. Liu, J.; Feng, X.; Fan, X.; Zhang, Z.; Zhang, B.; Liu, J.; Su, L. Efficient continuous-wave and passive Q-switched mode-locked Er3+:CaF2–SrF2 lasers in the mid-infrared region. Opt. Lett. 2018, 43, 2418–2421. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, J.; Feng, X.; Fan, X.; Zhang, Z.; Zhang, B.; Liu, J.; Su, L. Revisiting the self-termination effect of erbium based near 3 μm mid-infrared lasers. J. Lumin. 2022, 252, 119339. [Google Scholar]
  17. Yang, Z.; Zhang, Z.; Zhao, Z.; Su, L.; Liu, Z. High-Power Continuous-Wave and Tunable Er:CaF2 Lasers. Chin. J. Lasers 2024, 51, 1701006. [Google Scholar]
  18. Liu, J.; Feng, X.; Fan, X.; Zhang, Z.; Zhang, B.; Liu, J.; Su, L. Mid-infrared SESAM mode-locked Er:CaF2-SrF2 bulk laser at 2.73 μm. Chin. Opt. Lett. 2024, 22, 051406. [Google Scholar] [CrossRef]
  19. Chou, H.; Jenssen, H.P. Upconversion processes in Er-activated solid state laser materials. In Proceedings of the Advanced Solid State Lasers, Cape Cod, MA, USA, 19–20 January 1989; Optica Publishing Group: Washington, DC, USA, 1989; p. DD5. [Google Scholar]
  20. Jensen, T.; Diening, A.; Huber, G.; Chai BH, T. Investigation of diode-pumped 2.8 μm Er:LiYF4 lasers with various doping levels. Opt. Lett. 1996, 21, 585–587. [Google Scholar] [CrossRef] [PubMed]
  21. Jensen, T.; Ostroumov, V.G.; Huber, G. Upconversion processes in Er3+:YSGG and diode-pumped laser experiments at 2.8 μm. In Proceedings of the Advanced Solid State Lasers, Memphis, TE, USA, January 30–February 2 1995; Optical Society of America: Washington, DC, USA, 1995; p. IL4. [Google Scholar]
  22. Pollnau, M.; Lüthy, W.; Weber, H.P.; Jensen, T.; Huber, G.; Cassanho, A.; Jenssen, H.P.; McFarlane, R.A. Investigation of diode-pumped 2.8 μm laser performance in Er:BaY2F8. Opt. Lett. 1996, 21, 48–50. [Google Scholar] [CrossRef] [PubMed]
  23. Kaminskii, A.; Osico, V.; Prochorov, A.; Voronko, Y. Spectral investigation of the stimulated radiation of Nd3+ in CaF2-YF3. Phys. Lett. 1966, 22, 419–421. [Google Scholar] [CrossRef]
  24. Sulyanova, E.A.; Sobolev, B.P. The universal defect cluster architecture of fluorite-type nanostructured crystals. CrystEngComm 2022, 24, 3762–3769. [Google Scholar] [CrossRef]
  25. Ma, F.; Zhang, Z.; Jiang, D.; Zhang, Z.; Kou, H.; Strzep, A.; Tang, Q.; Zhou, H.; Zhang, M.; Zhang, P.; et al. Neodymium cluster evolution in fluorite laser crystal: A combined DFT and synchrotron X-ray absorption fine structure study. Cryst. Growth Des. 2022, 22, 4480–4493. [Google Scholar] [CrossRef]
  26. Corish, J.; Catlow CR, A.; Jacobs PW, M.; Ong, S.H. Defect aggregation in anion-excess fluorites. Dopant monomers and dimers. Phys. Rev. B 1982, 25, 6425–6438. [Google Scholar] [CrossRef]
  27. Bendall, P.J.; Catlow CR, A.; Corish, J.; Jacobs PW, M. Defect aggregation in anion-excess fluorites II. Clusters containing more than two impurity atoms. J. Solid State Chem. 1984, 51, 159–169. [Google Scholar] [CrossRef]
  28. Wang, Q.G.; Su, L.; Liu, J.F.; Liu, B.; Wu, F.; Luo, P.; Zhao, H.-Y.; Shi, J.-J.; Xue, Y.-Y.; Xu, X.-D.; et al. Intensities and spectral features of the–potential laser transition of Er3+ centers in CaF2–CeF3 disordered crystal. Chin. Phys. B 2017, 26, 114208. [Google Scholar] [CrossRef]
  29. Zhang, Z.; Ma, F.; Guo, X.; Wang, J.; Qian, X.; Liu, J.; Liu, J.; Su, L. Mid-infrared spectral properties and laser performance of Er3+ doped CaxSr1-xF2 single crystals. Opt. Mater. Express 2018, 8, 3820–3828. [Google Scholar] [CrossRef]
  30. Basiev, T.T.; Kharikov, E.V.; Zhekov, V.I.; Murina, T.M.; Osiko, V.V.; Prokhorov, A.M.; Starikov, B.P.; Timoshechkin, M.I.; Shcherbakov, I.A. Radiative and nonradiative transitions exhibited by Er3+ ions in mixed yttrium-erbium aluminum garnets. Sov. J. Quantum Electron. 1976, 6, 796. [Google Scholar] [CrossRef]
  31. Kumar, G.A.; Riman, R.; Chae, S.C.; Jang, Y.N.; Bae, I.K.; Moon, H.S. Synthesis and spectroscopic characterization of CaF2:Er3+ single crystal for highly efficient 1. 53 μm amplification. J. Appl. Phys. 2004, 95, 3243–3249. [Google Scholar] [CrossRef]
  32. Ma, F.; Su, F.; Zhou, R.; Ou, Y.; Xie, L.; Liu, C.; Jiang, D.; Zhang, Z.; Wu, Q.; Su, L.; et al. The defect aggregation of RE3+ (RE = Y, La~Lu) in MF2 (M = Ca, Sr, Ba) fluorites. Mater. Res. Bull. 2020, 125, 110788. [Google Scholar] [CrossRef]
  33. Normani, S.; Basyrova, L.; Loiko, P.; Benayad, A.; Braud, A.; Dunina, E.; Fomicheva, L.; Kornienko, A.; Hideur, A.; Camy, P. Erbium-Doped Fluorite-Type Crystals for 2.8 µm Lasers. In Proceedings of the Advanced Solid State Lasers, Barcelona, Spain, 11–15 December 2022; Optica Publishing Group: Washington, DC, USA, 2022; p. AM3A 2. [Google Scholar]
  34. Eichler, H.J.; Findeisen, J.; Liu, B.; Kaminskii, A.A.; Butachin, A.V.; Peuser, P. Highly efficient diode-pumped 3 μm Er3+:BaY2F8 laser. IEEE J. Sel. Top. Quantum Electron. 1997, 3, 90–94. [Google Scholar] [CrossRef]
  35. Li, T.; Beil, K.; Kränkel, C.; Huber, G. Efficient high-power continuous wave Er:Lu2O3 laser at 2.85 μm. Opt. Lett. 2012, 37, 2568–2570. [Google Scholar] [CrossRef] [PubMed]
  36. You, L.; Lu, D.; Pan, Z.; Yu, H.; Zhang, H.; Wang, J. High-efficiency 3 μm Er:YGG crystal lasers. Opt. Lett. 2018, 43, 5873–5876. [Google Scholar] [CrossRef]
  37. Hu, Q.; Nie, H.; Mu, W.; Yin, Y.; Zhang, J.; Zhang, B.; He, J.; Jia, Z.; Tao, X. Bulk growth and an efficient mid-IR laser of high-quality Er:YSGG crystals. CrystEngComm 2019, 21, 1928–1933. [Google Scholar] [CrossRef]
  38. Yao, W.; Uehara, H.; Kawase, H.; Chen, H.; Yasuhara, R. Highly efficient Er:YAP laser with 6.9 W of output power at 2920 nm. Opt. Express 2020, 28, 19000–19007. [Google Scholar] [CrossRef]
Figure 1. Illustration of (a) the ETU1 process between Er3+ and (b) the inequivalent substitution of RE3+ ions and their aggregating process in CaF2 crystal.
Figure 1. Illustration of (a) the ETU1 process between Er3+ and (b) the inequivalent substitution of RE3+ ions and their aggregating process in CaF2 crystal.
Crystals 14 00639 g001
Figure 2. Schematic diagram of the MIR laser setup.
Figure 2. Schematic diagram of the MIR laser setup.
Crystals 14 00639 g002
Figure 3. (a) As-grown Er,La:CaF2 crystal ingot, (b) illustration of fluorite structure of CaF2, (c) polished crystal chips, and (d) crystal blocks for spectral measurements and MIR laser experiments.
Figure 3. (a) As-grown Er,La:CaF2 crystal ingot, (b) illustration of fluorite structure of CaF2, (c) polished crystal chips, and (d) crystal blocks for spectral measurements and MIR laser experiments.
Crystals 14 00639 g003
Figure 4. (a) PXRD patterns of the crystals, (b) plot of lattice constant versus La concentration for Er,La:CaF2 crystals.
Figure 4. (a) PXRD patterns of the crystals, (b) plot of lattice constant versus La concentration for Er,La:CaF2 crystals.
Crystals 14 00639 g004
Figure 5. The absorption cross-sections of the crystal samples corresponding to (a) Er3+: 4I15/2 to 4I11/2 and (b) 4I15/2 to 4I13/2 transitions.
Figure 5. The absorption cross-sections of the crystal samples corresponding to (a) Er3+: 4I15/2 to 4I11/2 and (b) 4I15/2 to 4I13/2 transitions.
Crystals 14 00639 g005
Figure 6. The emission cross-sections of the crystal samples corresponding to Er: 4I11/2 to 4I13/2 transitions.
Figure 6. The emission cross-sections of the crystal samples corresponding to Er: 4I11/2 to 4I13/2 transitions.
Crystals 14 00639 g006
Figure 7. The measured fluorescence decay curves monitored at around (a) 1 μm (4I11/2), (b) 1.57 μm (4I13/2), and (c) 0.57 μm (4S3/2), and (d) the summary of the fitted lifetimes.
Figure 7. The measured fluorescence decay curves monitored at around (a) 1 μm (4I11/2), (b) 1.57 μm (4I13/2), and (c) 0.57 μm (4S3/2), and (d) the summary of the fitted lifetimes.
Crystals 14 00639 g007
Figure 8. Output power of 2.8 μm versus absorbed pump power at 976 nm for (a) 5 at.% Er, 2at.% La:CaF2 and (b) 5 at.% Er, 5 at.% La:CaF2.
Figure 8. Output power of 2.8 μm versus absorbed pump power at 976 nm for (a) 5 at.% Er, 2at.% La:CaF2 and (b) 5 at.% Er, 5 at.% La:CaF2.
Crystals 14 00639 g008
Table 1. Summary of the real doping concentrations of the crystals.
Table 1. Summary of the real doping concentrations of the crystals.
CrystalsEr3+La3+
at.%1021 ions/cm3at.%1021 ions/cm3
5 at.% Er:CaF24.741.16//
5 at.% Er, 2 at.% La:CaF24.671.112.120.449
5 at.% Er, 5 at.% La:CaF24.821.136.221.38
5 at.% Er, 10 at.% La:CaF25.251.2111.412.66
Table 2. Summary of ~3 μm CW laser performance of LD-pumped Er-doped crystals.
Table 2. Summary of ~3 μm CW laser performance of LD-pumped Er-doped crystals.
CrystalsEr concentration
(at.%)
Slope Efficiency
(%)
Output Power
(W)
Year
Er:LiYF415350.371996 [20]
Er:BaY2F810320.161997 [34]
Er:Lu2O3736/271.4/5.92012 [35]
Er:YGG1035.41.382018 [36]
Er:SrF23411.062018 [14]
Er:YSGG28.331.50.52019 [37]
Er:YAP530.66.92020 [38]
Er:CaF2537.90.72022 [33]
Er,La:CaF2536.60.92This Work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Z.; Liu, J.; Wang, Y.; Ma, F.; Liu, S.; Zhang, Z.; Liu, J.; Su, L. Study on Spectral Properties and Mid-Infrared Laser Performance of Er, La:CaF2 Crystals. Crystals 2024, 14, 639. https://doi.org/10.3390/cryst14070639

AMA Style

Zhang Z, Liu J, Wang Y, Ma F, Liu S, Zhang Z, Liu J, Su L. Study on Spectral Properties and Mid-Infrared Laser Performance of Er, La:CaF2 Crystals. Crystals. 2024; 14(7):639. https://doi.org/10.3390/cryst14070639

Chicago/Turabian Style

Zhang, Zhen, Jingjing Liu, Yunfei Wang, Fengkai Ma, Shaochen Liu, Zhonghan Zhang, Jie Liu, and Liangbi Su. 2024. "Study on Spectral Properties and Mid-Infrared Laser Performance of Er, La:CaF2 Crystals" Crystals 14, no. 7: 639. https://doi.org/10.3390/cryst14070639

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