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

Laser operation of Yb3+in disordered Li0.75Gd0.75Ba0.5(MoO4)2 crystal with small quantum defect

Open Access Open Access

Abstract

A new strategy has been developed to enhance the optical bandwidths of rare earth dopants by partial substitution of the divalent cation (D) in the original DWO4 or DMoO4 crystal structures. For demonstration, the monoclinic (space group C2/c) Yb-doped Li0.75Gd0.75Ba0.50(MoO4)2 crystal was grown in a Li2Mo2O7 flux, and 300 K Yb3+laser operation is reported. The laser emission is characterized by rather short wavelengths related to the specific features of the absorption and emission spectra, which leads to very small quantum defect, i.e., as low as 0.7% for E//b-axis. Using a Ti:sapphire laser at 976.6 nm, up to 295 mW of cw power are obtained without special cooling while the tunability range extends over 33 nm around 1020 nm.

©2007 Optical Society of America

1. Introduction

Solid-state Yb-lasers presently receive much attention for high power applications in diode pumped thin disk [1], waveguide [2], and fiber [3] design configurations. The main reasons for this interest are: i) The availability of affordable, robust, and high power diode lasers emitting at 980 nm where Yb3+has rather high optical absorption. ii) The small quantum defect characteristic for the absorption and emission processes in Yb3+, which minimizes the heat transfer to the host. iii) The existence of only two 2S+1LJ multiplets (2F5/2 and 2F7/2) for the Yb3+energy level scheme, which excludes optical losses by up-conversion, and the long lifetime of the excited 2F5/2 state (typically ~1 ms), ensuring efficient energy storage. Moreover, due to the large number of electrons in the 4f13 configuration of Yb3+, these are less shielded from the host interaction than in other lanthanides, for instance in Nd3+, therefore Yb3+exhibits a significant broadening of its optical transitions, which makes it suitable for tunable lasers and for ultrashort pulse operation. This broadening can be further enhanced by the presence of local disorder for Yb3+ in a given host. The main drawback for room temperature laser operation of Yb3+remains the relatively large population of the terminal level of its quasi 3-level laser system. This promotes the search for new hosts with large splitting of the ground state multiplet.

Since the early days of solid state laser technology tungstate and molybdate crystals with the general formula DXO4 (D being a divalent cation and X=W6+or Mo6+) have attracted attention as laser hosts due to their inert character. The incorporation of laser active trivalent lanthanide ions was achieved by the full substitution of the divalent D2+cation in the above formula by a pair of monovalent M+and trivalent T3+cations [4]. This led to the so-called double tungstate (DT) and double molybdate (DMo) crystals. The two principal crystalline structures developed for laser applications are the monoclinic C2/c, represented by KTW (T=Y, Gd, and Lu) compounds [5], and the tetragonal I4̄ structure characteristic of crystals such as NaTW (T=Y [6], La [7], Gd [8] and Lu [9]) or LiGdMo [10]. In these formulae, for the sake of simplicity, W and Mo denote (WO4)2 and (MoO4)2, respectively. The properties of Yb3+and other lanthanide lasants in such crystals have been studied extensively in recent years, but further improvements can be still expected by the development of novel DT and DMo crystal compositions. In this work we demonstrate an alternative partial substitution of the divalent cation leading to a large number of new compounds. Moreover, we present for the first time to our knowledge experimental results on Yb-laser operation in one of such compounds, Li0.75Gd0.75Ba0.5(MoO4)2, hereafter LiGdBaMo, which is a monoclinic crystal with local disorder associated with the partial substitution of the divalent cation. The quantum defect for the laser operation of Yb3+in this host succeeds to be one of the smallest observed in a solid state laser.

2. Crystal growth and structural details

The full substitution of the divalent cation D2+mentioned above can be described by:

a) 2 [D2+XO4]→M+T3+(XO4)2

Partial substitution of D can be achieved in several ways keeping constant the total number of cations and the total electric charge of the original D2+ions. Two possibilities can be schematized as shown below:

b) 4[D2+XO4]→M+T3+D2(XO4)4=2[M0.5T0.5D(XO4)2]

c) 8[D2+XO4]→M3 +T3 3+D2(XO4)8=4[M0.75T0.75D0.5(XO4)2]

We followed the last approach, (c), in order to grow 10 at % Yb-doped crystals of Li0.75Gd0.75Ba0.5(MoO4)2. The Top Seeded Solution Growth (TSSG) slow cooling method was applied using a Li2Mo2O7 flux. The molar solute/flux ratio used was 2.66:97.34. The mixture was melted at 610°C in a Pt crucible and then held for 20 hours at ≈ 50°C above the melting temperature in order to achieve good homogenization. A Pt wire rotating at 12 rpm was used as the seed. For crystallization, the cooling interval was 11 0C and the cooling rate was 0.08°C/h. More details on the growth procedures can be found elsewhere [11].

The Yb:LiGdBaMo crystals obtained are monoclinic with the symmetry of the centrosymmetric space group C2/c. The single crystal X-ray diffraction refinement yielded the composition formula Li0.715Gd0.725Yb0.06Ba0.5(MoO4)2 [11]. The slight Li+deficiency is compensated by excess incorporation of trivalent Gd3+and Yb3+ions. This balance between monovalent and trivalent cations leads to extra positive electric charge, which is likely compensated by a small Mo6+deficiency, as reported in other DMo compounds [12]. The Yb3+content in the crystal is [Yb]=3.76×1020 at/cm3 [11]. In the refined structure Gd3+(Yb3+) and Li(1)+share a single 8f site, while Ba2+and Li(2)+fully occupy different 4e sites, and Mo1, Mo2 and eight types of O are found in 8f sites. This ion distribution corrects a previous structural determination [13]. Our present results lead to the conclusion that only one crystallographic site with disordered local environment exists for Yb3+.

3. Yb3+optical spectroscopy and laser results

The disordered crystalline environment of Yb is reflected in a broad zero-line absorption, 2F7/2(0)→2F5/2(0’), with a full width at half maximum of FWHM=19 cm-1. For comparison the FWHM of the same transition in ordered monoclinic KYW is ≈10 cm-1. The sequence of Yb3+energy levels has been established (Ref. 11) as 2F7/2=0, 223, 353, 460 cm-1, and 2F5/2=10248, 10413 and 10634 cm-1. This gives a ratio of the partition functions of the upper and lower Yb3+multiplets close to unity.

The absorption of Yb3+is anisotropic with peak values of the absorption cross section σABS equal to 1.92×10-20 cm2 for E//a, 2.58×10-20 cm2 for E//b, and 1.25×10-20 cm2 for E//c, at λ=976.6 nm in all cases. The corresponding peak emission cross sections, σEMI, calculated by the reciprocity principle were 1.01×10-20 cm2 for E//a, 2.81×10-20 cm2 for E//b, and 1.30×10-20 cm2 for E//c, at λ=1000 nm in all cases [11]. For the sake of completeness we have included in Fig. 1 the spectral dependence of the absorption and emission cross sections. It is worth to observe that the monoclinic Yb:LiGdBaMo exhibits much stronger anisotropy of the spectroscopic properties than the tetragonal DT or DMo crystals, while still retaining enhanced optical bandwidths.

Continuous wave (cw) laser operation of Yb:LiGdBaMo was studied in the ≈140 cm long astigmatically compensated four mirror (Z-shaped) optical cavity shown in Fig. 2, with single pass longitudinal pumping. We used a c-cut sample with a thickness of 2.4 mm. The uncoated sample was inserted between the two focusing mirrors M2 and M3 (radius of curvature equal to RC=-10 cm), under Brewster angle. Laser experiments were performed for the sample orientations with largest cross sections, i.e., polarization parallel to the a-axis (E//a) or to the b-axis (E//b), see Fig. 1(a)–(b). This was achieved by rotation of the sample while the pump polarization remained in both cases the same as that of the laser. The sample was not actively cooled. The pump source was a Ti:sapphire laser with a maximum power of ~1.5 W at the optimum pump wavelength, λP=976.6 nm. The pump beam was focused by a 6.28 cm ARcoated lens (FL). The transmission of the plane output coupler M1 was varied from 1% to 10%.

 figure: Fig. 1.

Fig. 1. 300 K polarized absorption (dots) and calculated emission (solid line) cross sections of Yb3+in LiGdBaMo laser.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Cavity used to study the cw and tunable laser operation of Yb:LiGdBaMo crystal

Download Full Size | PDF

The absorption measured at low incident pump power (0.5 W) was roughly 75%, both for E//a and E//b polarizations. Bleaching reduced this value to about 50% at the maximum incident power of 1.5 W applied, but when the laser was aligned, the absorption was enhanced by the recycling effect in the 3-level system of Yb3+and the same level of 75% was estimated almost independently of the incident pump power and the output coupler used.

The cw laser operation results obtained at room temperature are summarized in Fig. 3 and in Table 1. The output power is initially linearly proportional to the absorbed power, however, roll-off is observed at higher pump levels; this effect is more significant for the E//a polarization. Since the linear dependence was recovered using a chopper with a duty cycle of 25% for the pump beam, thermal effects should be responsible for this behaviour. The maximum output power obtained in the true cw regime for E//a polarization with an output coupler of TOC=10 % was 295 mW, Fig 3(a), at an absorbed pump power of 1.06 W. The maximum slope efficiency with respect to the absorbed power, obtained for TOC=10%, was η=47% (E//a) and η=43% (E//b). Note that although E//a polarization exhibited thermal degradation of the performance, it presented lower thresholds and higher slope efficiencies. The oscillation wavelength λL decreased with TOC, from 1026 nm (TOC=1%) to 1017–1019 nm (TOC=10%). This behaviour is in agreement with the predictions of the gain cross section profiles with increasing inversion rate β [11].

Tables Icon

Table 1. Slope efficiency η, laser wavelength λL, and pump power at laser threshold Pth, of the Yb:LiGdBaMo laser for several TOC.

 figure: Fig. 3.

Fig. 3. Output power versus absorbed pump power of the cw Yb:LiGdBaMo laser for E//a-and E//b-polarization and several output coupler transmissions TOC: 1% (circles), 3% (squares), 5% (triangles) and 10% (diamonds). The linear fits shown give the slope efficiencies obtained for each TOC value.

Download Full Size | PDF

The cw laser tunability has been studied with a two-plate birefringent filter inserted in the vicinity of the output coupler, see Fig. 2. Figure 4 shows the results obtained for the two polarizations considered. Continuous tuning was achieved from 1006 to 1039 nm for the E//apolarization with a FWHM of the tuning curve of 22 nm, and from 1003 to 1023 nm for the E//b-polarization with a FWHM of 9 nm. The shape of the tuning curves reflects the behaviour of the gain cross sections [11] or that of the emission cross sections shown in Fig. 1. The tuning ranges obtained are narrower than those typically observed with the tetragonal DT and DMo Yb-hosts of the NaTW or NaTMo-types, but they are characterized by rather short emission wavelengths, i.e., small quantum defects when pumped at the optimum absorption wavelength, 976.6 nm. Although in present experiments we have used an intracavity Lyot filter, the birefringence of Yb:LiGdBaMo would allow tuning by rotation of the current active laser crystal [14].

 figure: Fig. 4.

Fig. 4. Wavelength tunability of the Yb:LiGdBaMo laser for TOC=3% and the two possible polarizations. The incident pump power is ≈1.43 W.

Download Full Size | PDF

As already mentioned, a small quantum defect (equal to (λL-λP)/λP) is a prerequisite for power scaling. Using a 125 µm plate of KYbW water-cooled crystal, a quantum defect of 1.6% was reported at room temperature [15], and even lower quantum defect of 0.6% was observed with a 130 µm plate of 10 at% Yb-doped KYW crystal, using eight pump passes [16]. It is obvious from Fig. 1 that Yb:LiGdBaMo can be pumped at wavelengths longer than the one corresponding to maximum absorption, thus we have explored this opportunity by increasing the pump wavelength to the limits set by the pump focusing optics (in particular, the mirror M2 in Fig. 2). The results obtained with TOC=5% are shown in Fig. 5. As can be seen, for E//a polarization, the quantum defect achieved is about 2.8%. For E//b polarization, it was possible to obtain laser oscillation for pump wavelengths as long as λP=996.8 nm and the emission wavelength of λL=1003.8 nm gives then a quantum defect of only 0.7%. This result compares very well with the lowest values previously indicated for Yb-lasers, although no special cooling has been applied in the current case of Yb:LiGdBaMo. The very small quantum defect obtained in the present work with a bulk crystal of Yb:LiGdBaMo can be attributed to the specific anisotropy of the crystal host leading to favourable absorption and emission spectral profiles for polarization E//b, as shown in Fig. 1.

 figure: Fig. 5.

Fig. 5. Spectral record of the room temperature pump and emission wavelengths of the Yb:LiGdBaMo laser using a TOC=5%.

Download Full Size | PDF

4. Conclusions

The monoclinic (space group C2/c) Yb:LiGdBaMo crystal, successfully grown in the present work by the TSSG method, has a single 8f site occupied simultaneously by Gd (72.5%), Yb (6%) and Li (21.5%). This leads to locally disordered environment around the Yb3+dopant ions, expressed in inhomogeneous broadening of the spectral features. The spectroscopic properties of the crystal are anisotropic with the largest absorption and emission cross sections for E//b-polarization. Room temperature laser emission in the cw regime was obtained both for E//a-and E//b-polarizations using a c-cut sample. The highest output power and slope efficiency obtained with a 10% output coupler and E//a are 295 mW and 47%, respectively, with a threshold of 412 mW (absorbed pump power). Tunability of 33 nm at the zero level is obtained for this polarization using a Lyot filter. However, the other polarization, E//b, has allowed operating the laser at extremely low quantum defect, 0.7%, despite the non-optimized cooling conditions. The Yb:LiGdBaMo crystal is a good candidate for scaling the output power in a thin disk laser configuration. Thus, the partial substitution of the divalent cation in the general host formula DXO4 leads to new compounds with extraordinary optical properties for Yb3+.

Acknowledgments

This work was supported by the project MAT2005-6354-C03-01, Spain. A.G-C additionally acknowledges his FPU2003-3018 Spanish grant.

References and links

1. K. Petermann, D. Fagundes-Peters, J. Johannsen, M. Mond, V. Peters, J.J. Romero, S. Kutovoi, J. Speiser, and A. Giesen, “Highly Yb-doped oxides for thin-disc lasers,” J. Cryst. Growth 275, 135–140 (2005). [CrossRef]  

2. D. P. Shepherd, S. J. Hettrick, C. Li, J. I. Mackenzie, R. J. Beach, S. C. Mitchell, and H. E. Meissner, “High-power planar dielectric waveguide lasers,” J. Phys. D: Appl. Phys. 34, 2420–2432 (2001). [CrossRef]  

3. J. Kim, D. B. S. Soh, J. Nilsson, D. J. Richardson, and J. K. Sahu, “Fiber design for high-power low-cost Yb:Al-doped fiber laser operating at 980 nm.” IEEE J. Sel. Top. Quantum Electron. 13, 588–597 (2007). [CrossRef]  

4. L. G. van Uitert and R. R. Soden, “Single crystal tungstates for resonance and emission studies,” J. Appl. Phys. 31, 328–330 (1960). [CrossRef]  

5. V. Petrov, M. C. Pujol, X. Mateos, O. Silvestre, S. Rivier, M. Aguiló, R. M. Solé, J. Liu, U. Griebner, and F. Díaz, “Growth and properties of KLu(WO4)2, and novel ytterbium and thulium lasers based on this monoclinic crystalline host,” Laser & Photon. Rev. 1, 179–212 (2007). [CrossRef]  

6. A. García-Cortés, J. M. Cano-Torres, M. D. Serrano, C. Cascales, C. Zaldo, S. Rivier, X. Mateos, U. Griebner, and V. Petrov, “Spectroscopy and lasing of Yb-doped NaY(WO4)2: Tunable and femtosecond mode-locked laser operation,” IEEE J. Quantum Electron. 43, 758–764 (2007). [CrossRef]  

7. J. Liu, J. M. Cano-Torres, F. Esteban-Betegón, M. D. Serrano, C. Cascales, C. Zaldo, M. Rico, U. Griebner, and V. Petrov, “Continuous-wave diode-pumped operation of an Yb:NaLa(WO4)2 laser at room temperature,” Opt. Laser Technol. 39, 558–561 (2007). [CrossRef]  

8. C. Cascales, M. D. Serrano, F. Esteban-Betegón, C. Zaldo, R. Peters, K. Petermann, G. Huber, L. Ackermann, D. Rytz, C. Dupré, M. Rico, J. Liu, U. Griebner, and V. Petrov, “Structural, spectroscopic, and tunable laser properties of Yb3+-doped NaGd(WO4)2,” Phys. Rev. B 74, 174114:1–15 (2007).

9. A. García-Cortés, J. M. Cano-Torres, X. Han, C. Cascales, C. Zaldo, X. Mateos, S. Rivier, U. Griebner, V. Petrov, and F. J. Valle, “Tunable continuous wave and femtosecond mode-locked Yb3+laser operation in NaLu(WO4)2,” J. Appl. Phys. 101, 063110:1–7 (2007). [CrossRef]  

10. M. Rico, U. Griebner, V. Petrov, P. Ortega, X. Han, C. Cascales, and C. Zaldo, “Growth, spectroscopy, and tunable laser operation of the disordered crystal LiGd(MoO4)2 doped with ytterbium,” J. Opt. Soc. Am. B 23, 1083–1090 (2006). [CrossRef]  

11. A. García-Cortés, C. Cascales, and C. Zaldo, “Growth, structural and spectroscopic properties of Yb3+-doped Li0.75Gd0.75Ba0.5(MoO4)2 crystals,” Mat. Sc. Eng. B (2007), doi:10.1016/j.mseb.2007.07.049.

12. V. Volkov, M. Rico, A. Méndez-Blas, and C. Zaldo, “Preparation and properties of disordered NaBi(XO4)2, X=W or Mo, crystals doped with rare-earths,” J. Phys. Chem. Sol. 63, 95–105 (2002). [CrossRef]  

13. R. F. Klevtsova, A. D. Vasil’ev, L. A. Glinskaya, A. I. Kruglik, N. M. Kozhevnikova, and V. P. Korsun, “Crystal structure investigation of ternary molybdates Li3Ba2Ln3(MoO4)8 (Ln=Gd, Tm),” J. Struct. Chem.33, 443–447 (1992) [Translated from Zhurnal Strukt. Khimii, 33, 126–130 (1992)]. [CrossRef]  

14. I. Iparraguirre, I. Aramburu, J. Azkargorta, M. Illarramendi, J. Fernández, and R. Balda, “Wavelength tuning of Titanium Sapphire Laser by its own crystal birefringence,” Opt. Express 13, 1254–1259 (2005). [CrossRef]   [PubMed]  

15. P. Klopp, V. Petrov, and U. Griebner, “Potassium ytterbium tungstate provides the smallest laser quantum defect,” Jpn. J. Appl. Phys. 42, L246–L248 (2003). [CrossRef]  

16. M. Hildebrandt, U. Bünting, U. Kosch, D. Haussmann, T. Levy, M. Krause, O. Müller, U. Bartuch, and W. Viöl, “Diode-pumped Yb:KYW thin-disk laser operation with wavelength tuning to small quantum defects,” Opt. Commun. 259, 796–798 (2006). [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. 300 K polarized absorption (dots) and calculated emission (solid line) cross sections of Yb3+in LiGdBaMo laser.
Fig. 2.
Fig. 2. Cavity used to study the cw and tunable laser operation of Yb:LiGdBaMo crystal
Fig. 3.
Fig. 3. Output power versus absorbed pump power of the cw Yb:LiGdBaMo laser for E//a-and E//b-polarization and several output coupler transmissions TOC : 1% (circles), 3% (squares), 5% (triangles) and 10% (diamonds). The linear fits shown give the slope efficiencies obtained for each TOC value.
Fig. 4.
Fig. 4. Wavelength tunability of the Yb:LiGdBaMo laser for TOC =3% and the two possible polarizations. The incident pump power is ≈1.43 W.
Fig. 5.
Fig. 5. Spectral record of the room temperature pump and emission wavelengths of the Yb:LiGdBaMo laser using a TOC =5%.

Tables (1)

Tables Icon

Table 1. Slope efficiency η, laser wavelength λL , and pump power at laser threshold Pth , of the Yb:LiGdBaMo laser for several TOC .

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.