Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Tunable laser operation of ytterbium in disordered single crystals of Yb:NaGd(WO4)2

Open Access Open Access

Abstract

Lasing of Yb3+ in a disordered single crystal host, NaGd(WO4)2, is reported. Pump efficiencies as high as 20% and slope efficiencies as high as 30% are achieved for both σ- and π-polarizations with Ti:sapphire laser pumping. The emission of Yb:NaGd(WO4)2 is centered near 1030 nm. Tunability between 1016 and 1049 nm is obtained with a Lyot filter.

©2004 Optical Society of America

1. Introduction

Solid-state lasers based on rare-earth doped crystals operating in the 1 µm spectral region are nowadays widely spread as compact and reliable sources for various applications due to the possibility for highly efficient diode pumping. The potential of the trivalent ytterbium ion, Yb3+, as a dopant was recognized in the very early years of diode pumping because in comparison with Nd3+ this ion possesses longer energy-storage lifetime and smaller quantum defect, and can be pumped by the optically more robust InGaAs laser diodes operating in the 900–1000 nm spectral range [1]. The relatively simple two-manifold structure of Yb3+ has the advantage that excited state absorption, up-conversion and cross-relaxation processes are absent but the operation in a quasi-three-level laser scheme poses a significant design and implementation challenge. An additional advantage of the Yb3+ over the Nd3+ ion is related to the broader absorption and emission linewidths which on the one hand reduces the requirements to the pump laser diodes and on the other hand holds a greater promise for the generation of shorter light pulses by mode-locking in the high-power regime.

The monoclinic potassium double tungstates KR(WO4)2 with R=Y, Bi, Ln (Ln=Sm…Lu) [2] occupy a special position among the laser crystals due to their extremely large absorption and emission cross sections which is related also to their strong anisotropy and biaxiality [3]. In addition, the separation of adjacent R3+ ions in these ordered hosts is relatively large and therefore less concentration quenching occurs for increased doping levels. Highly efficient laser operation has already been demonstrated for Yb:KY(WO4)2 and Yb:KGd(WO4)2 [4], and more recently for Yb:KLu(WO4)2 [5]. Continuous wave (cw) room-temperature laser operation was possible even with the 100% Yb-doped (stoichiometric) KYb(WO4)2 [6].

Sodium double tungstates NaR(WO4)2 and molybdates NaR(MoO4)2 where R=Y, La, Bi and Ln (Ln=Ce…Lu) are also related to the Ca2+ ion substitution in the CaWO4 tetragonal scheelite structure, space group C4h6 (I41/a), by a pair of Na+ and R3+ ions [7]. For these compounds only weak phase transformations are observed upon cooling from high temperature. As a consequence the tetragonal crystalline symmetry is preserved at room temperature although some distortions towards lower symmetry space groups may appear as observed in the Bi-containing compounds [8,9].

In the tetragonal double tungstates and molybdates an almost random distribution of Na+ and trivalent metal R3+ ions over two lattice sites occurs which holds also for the active laser dopant. As a consequence, a locally variable crystal field around the dopant ion is present and the linewidths of the electronic transitions for the rare earth elements are found to be broader than in ordered crystals (see Fig. 10 in ref.[10]). Such disordered crystals with their inhomogeneous line broadening [11] have an intermediate position with respect to their opto-mechanical and spectroscopic properties between ordered laser hosts [12] and glasses [13]. Moreover, the uniaxial sodium double tungstates and molybdates possess an important technological advantage over their monoclinic potassium analogues. Some of them (e.g. R=La, Nd…Er, Bi) melt congruently and can be pulled in large sizes by the Czochralski method.

The major part of the literature dealing with disordered sodium double tungstates or molybdates is devoted to their growth and their structural and spectroscopic properties. Pulsed laser emission at room temperature is known only for the Nd ion in NaR(WO4)2 and NaR(MoO4)2 with R=La, Bi, Gd and Y [11,14], and also in the isostructural disordered KLa(WO4)2 [15], KLa(MoO4)2 [11,14], LiLa(MoO4)2 [11,14] and LiGd(MoO4)2 [11,14]. Room temperature cw laser operation has been demonstrated, however, only in four disordered Nd-hosts of this type: NaLa(MoO4)2 [16], LiGd(MoO4)2 [17], and more recently in NaGd(WO4)2 [18] and NaY(WO4)2 [1921].

The present work is devoted to cw laser studies of Yb3+-doped disordered NaGd(WO4)2 or shortly Yb:NaGdW. In order to determine the optimum pump conditions for the two polarizations we used in this initial experiment a tunable Ti:sapphire laser for pumping and in order to assess the mode-locking potential of Yb:NaGdW tunable operation for the σ and p polarizations using a Lyot filter was studied.

2. Crystal growth and characterization of Yb:NaGdW

Single crystals of NaGd1-xYbx(WO4)2 with x=0.1 in the melt were grown in air by the Czochralski method using a pulling equipment supplied with a crystal diameter control system. The raw materials, Gd2O3 (99.99%), Yb2O3 (99.99%), Na2CO3 (99.9%) and WO3 (99.8%) from Alfa Aesar, were mixed in the appropriate ratios and the desired phase was synthesized in a two step process: firstly the mixture was heated up to 750°C at a rate of 50°C/h, held at this temperature for 18 h and then cooled down to room temperature. The product was ground and in a second step heated to 850°C for 24 h to complete the synthesis. The phase obtained was assessed by X-ray powder diffraction (XRPD) and then heated to 25°C above the melting temperature in a 75 cc Pt crucible. The melt was kept at this temperature for one hour for homogenization and then the optimum crystallization temperature was determined by monitoring the crucible weight when seeding with a (001)-oriented NaLa(WO4)2 seed. The rotation and pulling rates were 10 rpm and 2.3 mm/h, respectively. The grown crystals were cooled to room temperature at a rate of 10°C/h. The crystals obtained were colourless. The dimensions of the pulled boule were 20 mm in diameter and 50 mm in length. Some crystal cracking appeared likely upon cooling. The refractive index was measured for this composition only at 1047 nm: no=1.948 and ne=1.943 indicate that the NaGdW host is a negative crystal with relatively low birefringence.

The crystal phase obtained was further studied by XRPD and the Yb content determined from ICP emission measurements. A tetragonal crystal structure was found. The lattice parameters derived from a Rietveld refinement of the XRPD profile were a=5.2370(2) Å, c=11.3636(6) Å [V=311.66(3) Å3]. The molar Yb content in the crystal was 8% giving a distribution coefficient of 0.8. Hence the Yb density in the crystal is 5.1×1020cm-3. The laser sample prepared was a 3.273 mm thick plate with a clear aperture of ≈10×15 mm2 (Fig. 1) allowing polarization either parallel or perpendicular to the optical axis c to be used.

The resonant transition of Yb3+ is actually also magnetic-dipole allowed and three spectra have to be measured for a complete description in an uniaxial crystal: α (E⊥c, H⊥c), π (E//c) and σ (H//c). In the case of Yb:NaGdW, however, we recorded very similar optical densities for the α and σ polarizations (Fig. 1). The maximum absorption cross sections near 974 nm calculated for the π and σ polarizations are 1.51 and 1.21×10-20 cm2, respectively; very close to the estimations in [22]. For the Brewster angle geometry used in the laser the small signal absorption amounts to 94.4% and 89.4% for the π and σ polarizations, respectively.

 

Fig. 1. The Yb:NaGdW sample used in the present work (left) and the recorded optical density at room temperature (right) for the three polarizations E//c (π), H//c (σ), and E⊥c, H⊥c (α).

Download Full Size | PDF

The Yb segregation coefficient, crystal lattice parameters and optical absorption properties so far reported for our sample agree rather well with those previously reported for Yb:NaGdW crystals grown in N2/O2 reductive atmosphere [22–25], despite the fact that the latter are supposed to contain W and O lattice vacancies and exhibit coloration in the visible and infrared spectral range. This suggests that the interaction of such lattice defects with the active Yb3+ ions can be neglected at least for the doping level considered here, x<10 molar %. The maximum emission cross sections calculated near 996 nm amount to ≈2×10-20 cm2 for the π polarization and ≈1.5×10-20 cm2 for the σ polarization [22]. Nevertheless the reabsorption makes it difficult to predict which polarization will have better laser performance and calculations of the gain cross sections (substracted emission and absorption cross sections assuming some inversion level) provide reliable information basically on the lasing wavelength at threshold only [22].

Measurements of fine powders in an immersion liquid in [22] provided a value of 320 µs for the fluorescence decay time independent of the Yb-doping level in the 0.5 to 5% range. This value is close to the radiative lifetime of 397 µs calculated from the absorption spectra [22], and it is also comparable to the lifetimes estimated for the monoclinic potassium double tungstates [5]. The pump saturation intensity we calculated with it is equal to 42.3 kW/cm2 (π) and 52.7 kW/cm2 (σ), respectively. Note that both the emission and the absorption cross sections of Yb:NaGdW are almost an order of magnitude lower than the cross sections observed for the E//Nm polarization in the Yb-doped monoclinic double tungstates [46]. That is why it is normal to expect lower efficiencies and higher laser thresholds for Yb:NaGdW than those found with the ordered double tungstates.

3. Experimental results

The experimental set-up used in the present work is shown in Fig. 2. The astigmatically compensated three-mirror cavity consisted of a total reflector M1 for the laser radiation with 50 mm radius of curvature, a folding totally reflecting mirror M2 with 100 mm radius of curvature and a plane output coupler M3 whose transmission was varied between 3 and 10%. Both M1 and M2 were highly transmitting for the pump radiation which ensured single side pumping and helped to avoid any feedback to the pump laser. The antireflection coated pump lens L (focal length 62.8 mm) and M2 introduced only negligible reduction of the pump power so that the incident power Pin on the crystal could be conveniently measured outside the resonator. The total cavity length amounted to 73 cm.

 figure: Fig. 2.

Fig. 2. Scheme of the three-mirror laser cavity of the Yb:NaGdW laser.

Download Full Size | PDF

The pump and the laser spot diameters were almost equal in the position of the crystal and we estimated from the Brewster reflection a Gaussian pump waist of wpump≈22 µm. Thus the pump beam size is πwp2/2=7.6×10-6 cm2. The peak incident on-axis intensity for the maximum incident power of Pin=2 W used in the present experiment is 260 kW/cm2. Averaging across the cross section by taking a factor of 0.5 means that we are approximately 3 times (π) and 2.5 times (σ) above the saturation intensity. For both polarizations, however, our measurements showed that the absorption depletion effect is partially compensated at high pump levels by the recycling effect (Fig. 3). Nonetheless, the absorbed power is always higher for the π-polarization.

 figure: Fig. 3.

Fig. 3. Measured single pass absorption with lasing and with lasing interrupted in the arm containing the output coupler versus incident pump power Pin for TOC=10%.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Output power Pout versus absorbed pump power Pabs (symbols) of the cw Yb:NaGdW laser obtained for E//c (a) and H//c (b) at the optimum pump wavelength λp indicated. The linear fits give the calculated slope efficiencies η for three different output couplers (TOC).

Download Full Size | PDF

Both the thresholds (at Pabs≈0.5 W) and the slope efficiencies obtained with three different output couplers were essentially the same for the two polarizations (Fig. 4). In both cases the optimum pump wavelengths λP indicated in the figures are very close to the absorption maximum. Such a similar performance for E//c and H//c can be attributed not only to the different reabsorption effect but also to the comparable emission cross sections at the actual laser wavelengths (Fig. 4) which are far from the corresponding maxima. Thermal effects were not observed up to Pin=2 W although no special cooling was provided to the crystal.

 figure: Fig. 5.

Fig. 5. Output power Pout versus laser wavelength λL for a maximum incident pump power of Pin=2 W and TOC=5% near 1030 nm. The tuning performance by a Lyot filter for both the π (filled triangles) and σ (empty triangles) polarizations is shown.

Download Full Size | PDF

The tuning operation of the laser was investigated inserting a two-plate Lyot filter under Brewster angle in the M2-M3 arm (Fig. 2) close to the output coupler. Under optimum alignment the power reduction with the filter inside the cavity did not exceed 10%. The tuning curve obtained for the π-polarization is slightly broader than for the σ-polarization (Fig. 5). Both curves seem to be slightly truncated at shorter wavelengths by the reflection characteristics of M1 and M2. Tunability extending from 1016 to 1049 nm at the zero-level was obtained. The FWHM for the π-polarization corresponds to Δν=6.5×1012 Hz. For a sech2-shaped mode-locked pulse this means than sub-50-fs pulses can be expected from this laser in the mode-locking regime. This is about two times less than the shortest pulses typically achieved nowadays with femtosecond lasers based on the monoclinic double tungstates for similar TOC values [26].

4. Conclusion

In conclusion cw and tunable room temperature laser operation of Yb3+:NaGd(WO4)2 under Ti:sapphire laser pumping was demonstrated. The performance for the two possible polarizations was very similar achieving an output power of about 300 mW for 1.5 W of absorbed pump radiation. The optimum pump wavelength corresponds to the maximum absorption and is suitable for InGaAs laser diodes. Initial tuning experiments indicate interesting potential for mode-locking with this disordered Yb-host. We note that in comparison to the recent laser demonstration of Yb:CaF2 [27] the disorder and the inhomogeneous line broadening in the sodium double tungstates is rather due to occupation with some probability of two possible sites (each with multiple environments) than to lattice defects.

The mode-locking experiments planned for the near future will reveal the actual advantages of this new Yb host. Besides this we intend to investigate the spectroscopic and laser properties for higher doping levels which could be interesting for highly absorbing thin layer laser designs.

Acknowledgments

This work was supported within the EU Project DT-CRYS and also by MEC (Spain) Project MAT2002-4605-C05-05. Dr. M. Rico thanks the Universities and Education Secretary of State and the European Social Fund for financial support.

References

1. W. F. Krupke, “Ytterbium solid-state lasers — the first decade,” IEEE J. on Selected Topics in Quantum Electron. 6, 1287–1296 (2000). [CrossRef]  

2. P. V. Klevtsov, L. P. Kozeeva, and L. Yu. Kharchenko, “Study of the crystallization and polymorphism of double potassium and trivalent metal tungstates, KR(WO4)2,” Sov. Phys. Crystallogr.20, 732–735 [transl. from Kristallographiya 20, 1210–1215 (1975)].

3. G. Boulon, “Yb3+-doped oxide crystals for diode-pumped solid state lasers: crystal growth, optical spectroscopy, new criteria of evaluation and combinatorial approach,” Opt. Mat. 22, 85–87 (2003). [CrossRef]  

4. N. V. Kuleshov, A. A. Lagatsky, V. G. Shcherbitsky, V. P. Mikhailov, E. Heumann, T. Jensen, A. Diening, and G. Huber, “CW laser performance of Yb and Er, Yb doped tungstates,” Appl. Phys. B 64, 409–413 (1997). [CrossRef]  

5. X. Mateos, V. Petrov, M. Aguiló, R. M. Solé, J. Gavaldá, J. Massons, F. Díaz, and U. Griebner, “Continuous-wave laser oscillation of Yb3+ in monoclinic KLu(WO4)2,” IEEE J. Quantum Electron. 40, 1056–1059 (2004). [CrossRef]  

6. M. C. Pujol, M. A. Bursukova, F. Güell, X. Mateos, R. Solé, J. Gavaldá, M. Aguiló, J. Massons, F. Díaz, P. Klopp, U. Griebner, and V. Petrov, “Growth, optical characterization, and laser operation of a stoichiometric crystal KYb(WO4)2,” Phys. Rev. B 65, 165121 (2002). [CrossRef]  

7. V. K. Trunov, V. A. Efremov, and J. A. Velikodnyj, Kristallochimija i svojstva dvojnych molibdatov i vol’framatov, (Nauka, Leningrad, 1986), in russian.

8. J. Hanuza, A. Benzar, A. Haznar, M. Maczka, A. Pietraszko, and J. H. van der Maas, “Structure and vibrational dynamics of tetragonal NaBi(WO4)2 scheelite crystal,” Vibrational Spectr. 12, 25–36 (1996). [CrossRef]  

9. J. Hanuza, A. Haznar, M. Maczka, A. Pietraszko, A. Lemiec, J. H. van der Maas, and E. T. G. Lutz, “Structure and vibrational properties of tetragonal scheelite NaBi(MoO4)2,” J. Ram. Spectr. 28, 953–963 (1997). [CrossRef]  

10. A. Méndez-Blas, M. Rico, V. Volkov, C. Cascales, C. Zaldo, C. Coya, A. Kling, and L. C. Alves “Optical spectroscopy of Pr3+ in M+Bi(XO4)2, M+=Li or Na and X=W or Mo, locally disordered single crystals,” J. Phys: Condens. Matter 16, 2139–2160 (2004). [CrossRef]  

11. A. A. Kaminskii, “Modern developments in the physics of crystalline laser materials,” phys. stat. sol. (a) 200, 215–296 (2003). [CrossRef]  

12. L. D. DeLoach, S. A. Payne, L. L. Chase, L. K. Smith, W. L. Kway, and W. F. Krupke, “Evaluation of absorption and emission properties of Yb3+ doped crystals for laser applications,” IEEE J. Quantum Electron. 29, 1179–1191 (1993). [CrossRef]  

13. X. Zou and H. Toratani, “Evaluation of spectroscopic properties of Yb3+-doped glasses,” Phys. Rev. B 52, 15889–15897 (1995). [CrossRef]  

14. A. A. Kaminskii, Laser crystals: their physics and properties, 2-nd edition, Springer series in optical sciences, vol. 14, (Springer, Berlin, 1990).

15. X. M. Han, L. Z. Zhang, M. W. Qiu, and C. F. Wang, “1.06 µm laser characteristics of Nd3+:KLa(WO4)2 crystal,” Mat. Res. Innovations 8, 68–70 (2004).

16. A. A. Kaminskii, G. Ya. Kolodnyni, and N. I. Sergeeva, “Continuous LaNa(MoO4)2:Nd3+ laser operating at 300°K,” J. Appl. Spectrosc.9, 1275–1276 (1968) [transl. from Zhurnal Prikladnoi Spektroskopii 9, 884–886 (1968)].

17. A. A. Kaminskii, P. V. Klevtsov, Kh. S. Bagdasarov, A. A. Maier, A. A. Pavlyuk, A. G. Petrosyan, and M. V. Provotorv, “New cw crystal lasers,” JETP Lett.16, 387–389 (1972) [transl. from Pis’ma v Zhurnal Eksperimental’noi i Teoreticheskoi Fiziki 16, 548–551 (1972)].

18. N. Faure, C. Borel, M. Couchaud, G. Basset, R. Templier, and C. Wyon, “Optical properties and laser performance of neodymium doped scheelites CaWO4 and NaGd(WO4)2,” Appl. Phys. B 63, 593–598 (1996).

19. W.-L. Zhou, X. X. Zhang, and B. H. T. Chai, “Laser oscillation at 1059 nm of a new laser crystal: Nd3+ doped NaY(WO4)2,” in OSA Trends in Optics and Photonics Series vol. 10., Advanced Solid State Lasers, C. R. Pollock and W. R. Bosenberg, eds., (Optical Society of America, Washington, D.C., 1997), pp. 451–454.

20. K. Fu, Z.-P. Wang, Z.-X. Cheng, J.-H. Liu, R.-B. Song, H.-C. Chen, and Z.-S. Shao, “Effect of Nd3+ concentration on the laser performance of a new laser crystal: Nd:NY(WO4)2,” Optics Laser Techn. 33, 593–595 (2001). [CrossRef]  

21. Z. Cheng, S. Zhang, K. Fu, and H. Cheng, “Growth, thermal and laser properties of neodymium doped sodium yttrium double tungstate crystal,” Jpn. J. Appl. Phys. 40, 4038–4040 (2001). [CrossRef]  

22. Yu. K. Voron’ko, E. V. Zharikov, D. A. Lis, A. A. Sobol’, K. A. Subbotin, S. N. Ushakov, V. E. Shukshin, and S. Dröge, “Growth and luminescent properties of NaGd(WO4)2:Yb3+ crystals,” Inorg. Mat.39, 1308–1314 (2003) [transl. from Neorganicheskie Materialy 39, 1509–1516 (2003)]. [CrossRef]  

23. E. V. Zharikov, G. M. Kuz’micheva, D. A. Lis, Yu. M. Papin, V. B. Rybakov, B. A. Smirnov, and K. A. Subbotin, “X-ray diffraction study of the structure and defect system of Yb-activated (Na0.5Gd0.5)WO4 crystals,” Inorg. Mat.39, 151–160 (2003) [transl. from Neorganicheskie Materialy 39, 200–211 (2003)]. [CrossRef]  

24. K. A. Subbotin, E. V. Zharikov, and V. A. Smirnov, “Yb-and Er-doped single crystals of double tungstates NaGd(WO4)2, NaLa(WO4)2, and NaBi(WO4)2 as active media for lasers operating in the 1.0 and 1.5 µm ranges,” Optics and Spectr.92, 601–608 (2002) [transl. from Optika i Spektroskopiya 92, 657–664 (2002)]. [CrossRef]  

25. D. Lis, E. Zharikov, K. Subbotin, Y. Voron’ko, A. Sobol, S. Ushakov, and V. Shukshin, “Growth and spectroscopic study of Yb3+:NaGd(WO4)2 as potential laser material,” in Advanced Solid-State Photonics 2003, paper TuB9, OSA Technical Digest, (Optical Society of America, Washingron, D. C.), pp. 183–185.

26. P. Klopp, V. Petrov, U. Griebner, and G. Erbert, “Passively mode-locked Yb:KYW laser pumped by a tapered diode laser,” Opt. Exp. 10, 108–113 (2002). [CrossRef]  

27. V. Petit, J. L. Doualan, P. Camy, V. Menard, and R. Moncorge, “CW and tunable laser operation of Yb3+ doped CaF2,” Appl. Phys. B 78, 681–684 (2004) [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. The Yb:NaGdW sample used in the present work (left) and the recorded optical density at room temperature (right) for the three polarizations E//c (π), H//c (σ), and E⊥c, H⊥c (α).
Fig. 2.
Fig. 2. Scheme of the three-mirror laser cavity of the Yb:NaGdW laser.
Fig. 3.
Fig. 3. Measured single pass absorption with lasing and with lasing interrupted in the arm containing the output coupler versus incident pump power Pin for TOC=10%.
Fig. 4.
Fig. 4. Output power Pout versus absorbed pump power Pabs (symbols) of the cw Yb:NaGdW laser obtained for E//c (a) and H//c (b) at the optimum pump wavelength λp indicated. The linear fits give the calculated slope efficiencies η for three different output couplers (TOC).
Fig. 5.
Fig. 5. Output power Pout versus laser wavelength λL for a maximum incident pump power of Pin=2 W and TOC=5% near 1030 nm. The tuning performance by a Lyot filter for both the π (filled triangles) and σ (empty triangles) polarizations is shown.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.