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

Synthesis, spectroscopy, and efficient laser operation of “mixed” sesquioxide Tm:(Lu,Sc)2O3 transparent ceramics

Open Access Open Access

Abstract

A novel transparent 4.76 at.% Tm:(Lu2/3Sc1/3)2O3 “mixed” sequioxide ceramic is synthesized by hot isostatic pressing (HIP) at 1800 °C / 195 MPa in an Ar atmosphere. Its structure is studied by scanning electron microscopy, X-ray diffraction, and Raman spectroscopy. The spectroscopic properties of the Tm3+ ion are described within the Judd-Ofelt theory, which resulted in intensity parameters of Ω2 = 2.429, Ω4 = 1.078 and Ω6 = 0.653 [10−20 cm2]. For the 3F43H6 transition, the maximum stimulated-emission cross-section σSE is 7.15 × 10−21 cm2 at 1951 nm. The radiative lifetime of the 3F4 state is 4.01 ms. Under diode-pumping at 802 nm, a microchip Tm:(Lu,Sc)2O3 ceramic laser generated ~1 W at 2100 nm with a slope efficiency of 24%. The spectroscopic and laser properties of the Tm:(Lu,Sc)2O3 ceramic are compared with those of a Tm:LuScO3 single crystal. The ceramic exhibits very broad and flat gain cross sections, which is promising for ultrashort pulse generation.

© 2017 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Cubic (C-type, bixbyite structure) rare-earth sesquioxides A2O3 (where A = Lu, Y, Sc, etc.) are well known host crystals for trivalent laser-active lanthanide ions (Ln3+) such as Yb3+, Tm3+ or Ho3+ [1]. They feature good thermal and thermo-optical properties (high thermal conductivity, κ = 12.8 W/mK for Lu2O3 which is weakly dependent on the Ln3+ concentration [2], and positive dn/dT coefficients [3]), broad transparency range (0.22-8 µm) and high refractive index (~1.91 for Lu2O3 at ~1 µm) [1], one of the lowest maximum phonon frequencies among the oxide matrices [1], as well as broad spectral bands for the Ln3+ dopant ions [1,4,5]. Efficient, wavelength-tunable and power-scalable near-IR Ln-doped sesquoxide lasers operating in the continuous-wave (CW) [1,6,7] and mode-locked (ML) regimes (at ~1 and ~2 µm) [8–10] have already been reported. These crystals have also been recognized to be very suitable for thin-disk oscillators [11,12].

Cubic A2O3 crystals are in particular attractive for thulium lasers. The Tm3+ ion (electronic configuration: [Xe]4f12) is known for its emission around 2 µm due to the 3F43H6 4f-4f electronic transition [13]. This eye-safe emission is used in range-finding, environmental sensing (LIDAR) and medicine. The Tm3+ ions can be excited by commercial AlGaAs laser diodes at ~0.8 µm (into the 3H4 state) while the possible cross-relaxation process, 3H4 + 3H63F4 + 3F4, increases the excitation quantum efficiency [14]. In cubic A2O3 crystals they feature a strong Stark splitting of the ground state (3H6), > 800 cm−1, leading to very broad emission and gain spectra extending up to ~2.1 µm [1,5,6] which is much longer than for many other Tm3+-doped oxide crystals. Moreover, this feature can be enhanced in “mixed” Tm:(Lu1-x,Scx)2O3 crystals exhibiting a compositional disorder. A 1 at.% Tm:LuScO3 sesquioxide crystal was previously studied in [15] generating a CW output power of 705 mW at ~2.1 µm with a slope efficiency of 55% (under Ti:Sapphire laser pumping). A 155 nm-broad (1960-2115 nm) tuning range of the laser emission was demonstrated [15]. In the ML regime, 105 fs pulses were achieved at 2010 nm using a semiconductor saturable absorber mirror (SESAM) [16]. This corresponded to the shortest pulse duration ever reported for any ML Tm3+:A2O3 oscillator, proving the potential of “mixed” crystals for fs ML lasers.

Due to the high melting point of the cubic A2O3 crystals (2450 °C for Lu2O3), their growth typically requires the use of expensive rhenium (Rh) crucibles [17]. Rh can be the source of crystal coloration. The growth of large-volume highly Tm3+-doped single crystals with high optical quality is complicated. The technology of Tm3+-doped sesquioxide transparent ceramics was proposed as an alternative allowing for an easier and size-scalable production. There are multiple reports on Tm:Lu2O3 transparent ceramics [18,19]. In [18], a 2 at.% Tm:Lu2O3 ceramic laser generated 26 W at 2066 nm with a slope efficiency of 42%. Bulk fs ML lasers [20] and thin-disk lasers [21] based on Tm:Lu2O3 ceramic are also known.

In the present work, we focused on a novel “mixed” Tm:(Lu2/3Sc1/3)2O3 transparent ceramic exhibiting promising potential for ultrafast lasers and amplifiers. We present structural and detailed spectroscopic characterization and compare this ceramic with a Tm:LuScO3 single-crystal. In addition to the ease of fabrication, we demonstrate that this new ceramic material can provide superior laser characteristics in comparison with the previously studied “mixed” Tm:LuScO3 ceramics [22] and single-crystals [15].

2. Fabrication of the Tm:(Lu,Sc)2O3 ceramics

The Tm:(Lu,Sc)2O3 ceramics were fabricated by the Hot Isostatic Pressing (HIP) sintering method using powders of Sc2O3, Lu2O3, and Tm2O3 (purity: 99.99%, Alfa Aesar) as raw materials. The raw materials, with a stoichiometric amount of 100 at.% Lu + Sc (taken in a proportion of Lu:Sc = 2:1) and 5 at.% Tm over it, were mixed uniformly by ball milling for 24 h, dried for 6 h at 70 °C, sieved, dry-pressed at 10 MPa, and cold isostactically pressed at 200 MPa. The green bodies of Tm:(Lu,Sc)2O3 ceramics were firstly pre-sintered at 1750 °C for 10 h under vacuum (pressure, P < 10−3 Pa) to densify the preforms. For further densification, the pre-sintered Tm:(Lu,Sc)2O3 ceramic samples were post-sintered by HIP at 1800 °C for 2 h in an Ar atmosphere (P = 195 MPa) to eliminate the closed pores around the grain boundaries. Finally, the ceramics were annealed at 1500 °C for 10 h in an O2 atmosphere to eliminate the oxygen vacancies and remove internal stresses. Samples of Tm:(Lu,Sc)2O3 ceramics with a diameter of 15 mm and a thickness of 5 mm were obtained. Some polished samples are shown in Fig. 1(a). For the spectroscopic experiments, 3 × 3 × 3 mm3 cubes were cut from the prepared samples and polished to laser-grade quality. The composition of the ceramics can be represented as 4.76 at.% Tm:(Lu2/3Sc1/3)2O3.

 figure: Fig. 1

Fig. 1 Some of the fabricated 4.76 at.% Tm:(Lu,Sc)2O3 ceramics (laser-grade-polished ceramic disks) (a); X-ray diffraction (XRD) pattern (in blue) compared to the standard pattern of ScLuO3 (in red) (b); Raman spectrum, excitation wavelength: 514 nm (c).

Download Full Size | PDF

The polycrystalline phase of the samples was analyzed at room-temperature (RT, 293 K) by X-ray diffraction (XRD) (D2 PHASER, Bruker Co. Ltd), using Cu Kα radiation. The experiment was carried out in the 2θ = 10-80° range with a step of 0.02° and a scan speed of 0.1°/min. As shown in Fig. 1(b), the measured diffraction pattern agrees with the JCPDS card No. 04-002-0541 of cubic LuScO3 (C-type structure, sp. gr. Ia3¯ - T7h, No. 206). Based on the position of the XRD peaks, the lattice constant of the Tm:(Lu,Sc)2O3 ceramic is a = 10.3683 Å (Z = 16, calculated density, ρcalc = 7.493 g/cm3), which is larger than that for the standard LuScO3 crystal (a = 10.105 Å). This is attributed to the larger fraction of Lu3+ with larger ionic radius (RLu = 0.861 Å for VI-fold O2- coordination) as compared to Sc3+ (RSc = 0.745 Å), as well as the Tm3+ doping (RTm = 0.88 Å).

The vibronic properties of the ceramics were studied with Raman spectroscopy, Fig. 1(c). We used a Renishaw inVia confocal Raman microscope with an excitation wavelength λexc = 514 nm. The Raman spectrum of the Tm:(Lu,Sc)2O3 ceramic is typical for cubic sesquioxides A2O3 [23] while significant broadening of the Raman bands is observed as compared to single crystals. The most intense Raman band is centered at 399 cm−1. It has an asymmetric shape with a full width at half maximum (FWHM) of 32 cm−1 and is composed of several poorly resolved components. A factor group analysis predicts the following irreducible representations at the Г-point: Гtotal = 4Ag + 4Eg + 14Fg + 5Au + 5Eu + 17Fu of which 22 modes (Ag, Eg and Fg) are Raman-active, 17 modes (Fu) are IR-active and the rest are silent [24]. We have resolved 16 Raman bands as indicated in Fig. 1(c). While the strongest band is typically ascribed to the Ag + Fg vibrations, the assignment of the modes is complicated [24].

The morphology of a polished surface and a fracture surface was characterized by field-emission scanning electron microscopy (SEM, Inspect F, FEI). The Tm:(Lu,Sc)2O3 sample is almost pore-free, see Fig. 2. The grain boundaries are clean and there is almost no secondary phase at the grain boundaries or inner grains. The average grain size in our samples is of 5-6 μm, as determined by the linear intercept method. The grain size distribution is uniform over the sample cross-section.

 figure: Fig. 2

Fig. 2 Field-emission scanning electron microscopy (SEM) images of a polished (a) and a fracture (b) surface of the 4.76 at.% Tm:(Lu,Sc)2O3 ceramic. The scale bar in (a,b) is 20 µm.

Download Full Size | PDF

3. Optical spectroscopy

3.1 Optical absorption

The absorption spectrum of the Tm:(Lu,Sc)2O3 ceramic was measured using a Varian CARY 5000 spectrophotometer (Agilent) with a resolution of 0.2 nm. In addition to RT studies, we measured the absorption spectra at 6 K using an Oxford Instruments Ltd. cryostat (Model SU 12) with He-gas close-cycle flow. From the absorption coefficient αabs measured at RT, we calculated the RT absorption cross-section, σabs = αabs/NTm, where NTm = 13.8 × 1020 cm−3 corresponds to 4.76 at.% Tm doping.

The transmission of the laser-grade polished 3 mm-thick Tm:(Lu,Sc)2O3 ceramic sample measured at 2.15 µm (out of the absorption band of Tm3 + ) was ~77% which is slightly lower than the value arising purely from the Fresnel losses expected for this material (81.1%, if assuming a refractive index n of 1.92). No noticeable distortion or scattering of the probe He-Ne laser beam passing through the sample were detected.

The RT absorption spectrum is shown in Fig. 3. The absorption band at 1.4-2.1 µm is associated with the 3H63F4 transition of Tm3 + . The maximum σabs for this transition is 4.26 × 10-21 cm2 at 1622 nm. This band is suitable for in-band pumping of Tm3 + ions, e.g., by Raman-shifted Er fiber lasers [25]. The band at 0.74-0.83 µm corresponds to the 3H63H4 transition and it renders the Tm:(Lu,Sc)2O3 ceramics suitable for pumping with AlGaAs laser diodes. Its maximum σabs is 2.80 × 10−21 cm2 at 793 nm with a FWHM of ~25 nm. At shorter wavelengths, the transitions to the 3F2,3 (0.64-0.71 µm), 1G4 (0.45-0.49 µm), 1D2 (0.35-0.37 µm) and 3P0-2 + 1I6 (0.27-0.30 µm) excited states were also observed. The last one overlaps with the UV absorption edge of the Tm:(Lu,Sc)2O3 ceramic at λg = 282 nm (Eg = 4.4 eV).

 figure: Fig. 3

Fig. 3 (a-f) Room-temperature (at 293 K, in red) and low-temperature (at 6 K, in blue, not in scale) absorption spectra of the 4.76 at.% Tm:(Lu,Sc)2O3 ceramic.

Download Full Size | PDF

The low-temperature absorption spectra of the Tm:(Lu,Sc)2O3 ceramic are also shown in Fig. 3 (in blue, not in scale with respect to the RT spectra). Due to the compositional disorder, the absorption spectra are very broad even at 6 K and the transitions to the individual Stark levels are not completely resolved.

The absorption bands of the Tm3+ ion in the (Lu,Sc)2O3 ceramic related to the transitions from the 3H6 ground-state to the excited states from 3F4 to 1D2 were analyzed using the standard Judd-Ofelt (J-O) theory [26,27]. The absorption oscillator strengths were determined from the measured absorption spectra:

fΣexp(JJ')=mec2πe2NTmλ2Γ(JJ'),
where me and e are the electron mass and charge, respectively, c is the speed of light, Г(JJ’) is the integrated absorption coefficient within the absorption band and 〈λ〉 is the “center of gravity” of the absorption band. The results are summarized in Table 1.

Tables Icon

Table 1. Experimental and Calculated Absorption Oscillator Strengths for 4.76 at.% Tm:(Lu,Sc)2O3 Ceramic

The values of fΣexp were used to determine the J-O (intensity) parameters, Ωk, k = 2, 4, 6: Ω2 = 2.429, Ω4 = 1.078 and Ω6 = 0.653 [10−20 cm2]. Using these parameters, the absorption oscillator strengths were calculated showing a good agreement with the experiment (as indicated by a moderate root-mean-square (rms) deviation, 0.35):

fΣcalc(JJ')=83h(2J'+1)λ(n2+2)29nSEDcalc(JJ')+fMDcalc(JJ'),
SEDcalc(JJ')=k=2,4,6U(k)Ωk,whereU(k)=(4fn)SLJ||Uk||(4fn)S'L'J'2.
Here, SEDcalc are the line strengths, h is the Planck constant, n is the refractive index of the crystal and U(k) are the squared reduced matrix elements [28]. The J-O theory describes the electric-dipole (ED) transitions. The contribution of the magnetic-dipole (MD) transition with JJ’ = 0, ± 1 (for the 3H63H5 transition), fMDcalc, was taken from [29].

3.2 Optical emission

The probabilities for spontaneous radiative transitions were calculated from the corresponding line strengths which were determined from the J-O parameters Ωk and the squared reduced matrix elements U(k), see Eq. (2b) [28]:

AΣcalc(JJ')=64π4e23h(2J'+1)λ3n(n2+23)2SEDcalc(JJ')+AMDcalc(JJ').
The MD contributions AMDcalc were taken from [29]. Using the values of AΣcalc for the separate emission channels JJ’, the total probability, Atot, the corresponding radiative lifetime of the excited state, τrad, and the luminescence branching ratios for the emission channels, B(JJ’) were determined:
τrad=1AtotandB(JJ')=AΣcalc(JJ')Atot,whereAtot=J'AΣcalc(JJ').
The results are shown in Table 2. The calculated radiative lifetime of the lowest excited state is τrad(3F4) = 4.01 ms. Previously, luminescence decay studies of a 1 at.% Tm:LuScO3 single crystal [30] and a 2 at.% Tm:LuScO3 ceramic [22] revealed a lifetime τlum(3F4) of 3.84 and 3.2 ms, respectively. The J-O modelling also predicts the radiative lifetime of the 3H4 pump level, τrad(3H4) = 0.72 ms. In [30] for a 1 at.% Tm:LuScO3 crystal, τlum(3H4) was measured to be 55 µs while no values for the LuScO3 crystals with lower doping are available (e.g., for 0.2 at.% Tm:Lu2O3 crystal, τlum(3H4) = 300 µs).

Tables Icon

Table 2. Calculated Emission Probabilities for Tm3+ in (Lu,Sc)2O3 Ceramic

The stimulated emission (SE) cross-sections, σSE, for the 3F43H6 transition of Tm3+ in (Lu,Sc)2O3 ceramic were calculated from the measured absorption spectrum, see Fig. 3(a), using the modified reciprocity method [31]:

σSE(λ)=18πn2τradcσabs(λ)ehc/(kTλ)λ4σabs(λ)ehc/(kTλ)dλ.
Here, k is the Boltzmann constant, T is the crystal temperature (RT), τrad is the radiative lifetime of the emitting state (3F4). To reduce the error in σSE at long wavelengths (> 2 µm) originating from the exponential term in Eq. (5), we additionally used the Füchtbauer–Ladenburg (F-L) equation [32] which is adequate because of the negligible reabsorption in this spectral range, Fig. 3(a). The σSE spectrum is shown in Fig. 4(a). The maximum σSE amounts to 7.15 × 10−21 cm2 at 1951 nm. Above 2 µm where laser operation is expected for low-loss cavities, a local maximum of σSE = 2.38 × 10−21 cm2 is found at 2090 nm.

 figure: Fig. 4

Fig. 4 Spectroscopy of Tm3+-doped (Lu,Sc)2O3 ceramic: (a) absorption, σabs, and stimulated-emission, σSE, cross-sections for the 3H63F4 transition; (b) gain cross-sections, σg = βσSE – (1 – β)σabs, for the 3F43H6 transition. β = N(3F4)/NTm is the inversion ratio.

Download Full Size | PDF

The gain cross-section, σg = βσSE – (1 – β)σabs, where β = N(3F4)/NTm is the inversion ratio, is typically calculated to estimate the expected laser wavelength, see Fig. 4(b). For low β < 0.06, broad and smooth gain spectra are observed centered at ~2.10 µm. For higher β, a peak at 1.97 µm and further one at 1.95 µm dominate the spectrum.

4. Continuous-wave laser operation

4.1 Laser set-up

The laser experiments were performed in a compact microchip-type cavity, Fig. 5(a). The 2.91 mm-thick polished and uncoated Tm:(Lu,Sc)2O3 ceramic active element (AE) with an aperture of 3.01 × 3.06 mm2 was mounted in a water cooled (12 °C) Cu-holder. Indium foil was used to provide a better thermal contact from all 4 lateral sides of the AE. The plano-plano cavity consisted of a flat pump mirror (PM) antireflection (AR) coated for 0.77-1.05 μm and high reflection (HR) coated for 1.80-2.12 μm, and a flat output coupler (OC) with transmission of TOC = 1.5%, 3% or 5% at 1.84-2.12 μm. Both PM and OC were placed close to the AE with minimum air gaps, so that the geometrical cavity length was ~3 mm. The AE was pumped through the PM by a fiber-coupled AlGaAs laser diode (LuOcean model Lu0808D200, Lumics, fiber core diameter: 200 μm, NA: 0.22) emitting randomly polarized output at 802 nm (the wavelength was stabilized by water-cooling of the diode). We used a lens assembly (focal length: 30 mm, 1:1 imaging ratio) to focus the pump light into the AE. The radius of the pump beam in the crystal was 100 μm (confocal parameter of the pump beam 2zR = 1.9 mm). The OCs provided a partial reflection (R ≈40%) at the pump wavelength. The total pump absorption under lasing conditions was 48%.

 figure: Fig. 5

Fig. 5 Tm:(Lu,Sc)2O3 ceramic microchip laser: (a) scheme of the laser, LD – laser diode, PM – pump mirror, OC – output coupler; (b) laser beam profile captured at Pabs = 5.05 W, corresponding to an output power of 1.01 W for TOC = 3%.

Download Full Size | PDF

The laser emission spectra were measured with a compact spectrometer (WaveScan, sensitivity: 1.0-2.6 µm, APE GmbH). The beam profile was captured in the far-field using a FIND-R-SCOPE near-IR camera (model 85726, sensitivity: 0.4-2.2 µm), see Fig. 5(b).

4.2 Microchip laser operation

Microchip laser operation has been achieved with all OCs. This indicates a positive thermal lens (and positive dn/dT coefficient) for Tm:(Lu,Sc)2O3 ceramics. The laser output was randomly polarized. The input-output dependences are shown in Fig. 6(a). The maximum output power reached 1.01 W at 2095-2102 nm with a slope efficiency η = 24% (with respect to the absorbed pump power, Pabs) for TOC = 3%. The laser threshold was at Pabs = 0.86 W and the optical-to-optical efficiency ηopt was 10% (with respect to the incident power). For TOC = 1.5%, slightly lower η = 19% was observed. For both OCs, the output dependence was linear up to Pabs ~5 W where thermal fracture was observed. For TOC = 5%, the laser output deteriorated (96 mW with η = 8%) which is attributed to stronger upconversion losses and the associated heating. Typical laser emission spectra are presented in Fig. 6(b). The laser emission occurred at ~2.09 µm in agreement with the gain spectra of Tm3 + for low inversion ratios β < 0.06, Fig. 4(b). A small β is expected for the utilized low output coupling. The multi-peak spectral behavior is due to etalon effects.

 figure: Fig. 6

Fig. 6 Tm:(Lu,Sc)2O3 ceramic microchip laser: (a) input-output dependences, η – slope efficiency; (b) typical laser emission spectra measured at Pabs = 5.05 W (for TOC = 1.5% and 3%) and 2.40 W (TOC = 5%).

Download Full Size | PDF

The Tm:(Lu,Sc)2O3 ceramic microchip laser generated a nearly-circular output beam with a measured M2x,y < 1.2 (at Pabs = 5.05 W, for TOC = 3%), see Fig. 5(b).

As compared with a previous report on a 2 at.% Tm:LuScO3 ceramic emitting at 1982 nm [22], the laser characteristics in the present work are considerably improved. In [22], under diode-pumping at 790 nm, the maximum output power was limited to 211 mW (at Pabs ~3.5 W) by thermal roll-over with a slope efficiency of only 8%. Here, despite the higher Tm doping, the ceramic quality allowed to work at higher pump levels to achieve an output power as high as ~1 W at 3 times higher slope efficiency.

5. Comparison of ceramics with a single crystal

For the comparative spectroscopic and laser studies, we used a 1.0 at.% Tm:LuScO3 crystal grown by the heat-exchanger method (HEM) [30].

At first, we compare the spectroscopic properties of the single crystal and the ceramic. The absorption cross-section spectra for the 3H63H4 transition are shown in Fig. 7(a). The spectra are similar in shape while less structured for the ceramics. The maximum σabs of 2.60 × 10−21 cm2 for the Tm:LuScO3 single crystal is observed at 793 nm. It is known that for the Tm3+ ion in a “mixed” LuScO3 crystal, σabs is lower compared to Tm:Lu2O3 and Tm:Sc2O3 [30] due to a compositional disorder. A similar effect is expected for “mixed” ceramics.

 figure: Fig. 7

Fig. 7 Spectroscopy of Tm3+ in a 4.76 at.% Tm:(Lu2/3Sc1/3)2O3 ceramic and in a 1 at.% Tm:LuScO3 single-crystal: (a) σabs spectra for the 3H63H4 transition, (b) σSE spectra for the 3F43H6 transition. The noise of the black curve in (b) is due to water absorption affecting the measured luminescence spectrum.

Download Full Size | PDF

The SE cross-section spectra for the 3F43H6 transition for the single crystal and the ceramic are compared in Fig. 7(b). A slight reduction of σSE for the ceramic is observed: for the Tm:LuScO3 single crystal, the maximum σSE is 8.0 × 10−21 cm2 at 1956 nm. The σSE value for the single crystal determined with the F-L method can be overestimated as compared to that obtained by the reciprocity method in analogy to Tm:Lu2O3 [30]. The physical reason is the shorter measured luminescence decay time τlum for Tm:A2O3 crystals than the radiative one τrad. The peak in the SE cross-section spectra at >2 µm is shifted to a longer wavelength for the Tm:LuScO3 single-crystal (2098 nm) due to the larger fraction of Sc3+ ions as compared to the ceramic.

For the laser experiments with the Tm:LuScO3 single-crystal, we cut a 2.91 mm-thick cubic active element with an aperture of 2.82 × 3.13 mm2. It was polished to laser-grade quality and remained uncoated. The experiments were performed in the set-up depicted in Fig. 5(a). Microchip laser operation has been achieved and the results are presented in Fig. 8. Due to the observed degradation of the Tm:LuScO3 laser output for TOC > 3%, we additionally studied a high-reflective OC (TOC = 0.2%). The laser emission was randomly polarized. The output power reached 221 mW at 2092-2095 nm with η = 21% for TOC = 1.5%. The laser threshold was at Pabs = 0.38 W and ηopt was 7%.

 figure: Fig. 8

Fig. 8 Tm:LuScO3 single crystal microchip laser: (a) input-output dependences, η – slope efficiency; (b) typical laser emission spectra measured at Pabs = 1.41 W.

Download Full Size | PDF

Previously, a similar crystal with a thickness of 3 mm was studied under Ti:Sapphire pumping at 796 nm [15,30]. The output power was 324 mW at 1978-1992 nm with η = 40% (for TOC = 1.5%). This is comparable to our result considering the better quality of the pump beam from the Ti:Sapphire laser. Using a 7.7 mm-thick sample, the same authors extracted 705 mW at 2090-2115 nm with η = 55% (at a maximum Pabs = 1.4 W) [30].

6. Conclusion

“Mixed” sesquioxide Tm:(Lu,Sc)2O3 transparent ceramics produced by the HIP method are promising materials for broadly tunable CW and ultrashort (femtosecond) ML oscillators at 1.95-2.1 µm. They allow for high Tm3+ doping levels by keeping the optical quality high, size-scalable synthesis, and similar spectroscopic properties as compared to those for HEM-grown Tm:LuScO3 single crystals. In the present work, we report on the first transparent ceramics with composition (Lu2/3Sc1/3)2O3 doped with 4.76 at.% Tm. Their morphology, structure and vibronic properties are studied. The transition probabilities of Tm3+ ions are determined within the J-O theory. The ceramic provides broad and smooth gain spectra at 1.9-2.15 µm, as well as relatively high stimulated-emission cross-section at >2 µm, namely 2.38 × 10−21 cm2 at 2090 nm. Using the Tm:(Lu,Sc)2O3 ceramic in a compact diode-end-pumped microchip laser, watt-level output at ~2.1 µm has been achieved with a slope efficiency of 24%, representing the highest output power demonstrated for any Tm3+-doped (Lu,Sc)2O3 crystalline or ceramic laser material.

Funding

Spanish Government (MAT2016-75716-C2-1-R, (AEI/FEDER,UE); MAT2013-47395-C4-4-R, TEC 2014-55948-R); Generalitat de Catalunya (2014SGR1358).

Acknowledgments

W. J. acknowledges financial support from the Key Laboratory of Science and Technology on High Energy Laser, CAEP. F.D. acknowledges additional support through the ICREA academia award 2010ICREA-02 for excellence in research. P. L. acknowledges financial support from the Government of the Russian Federation (Grant No. 074-U01) through ITMO Post-Doctoral Fellowship scheme. P. L. expresses his gratitude to Dr. Christian Kränkel (Institut für Laser-Physik, Universität Hamburg) for providing a Tm:LuScO3 single-crystal and the information about its spectroscopic properties, as well as for the fruitful discussion.

References and links

1. C. Kränkel, “Rare-earth-doped sesquioixides for diode-pumped high-power lasers in the 1-, 2-, and 3-µm spectral range,” IEEE J. Sel. Top. Quantum Electron. 21(1), 1602013 (2015).

2. R. Peters, C. Kränkel, S. T. Fredrich-Thornton, K. Beil, K. Petermann, G. Huber, O. H. Heckl, C. R. E. Baer, C. J. Saraceno, T. Südmeyer, and U. Keller, “Thermal analysis and efficient high power continuous-wave and mode-locked thin disk laser operation of Yb-doped sesquioxides,” Appl. Phys. B 102(3), 509–514 (2011).

3. P. A. Loiko, K. V. Yumashev, R. Schödel, M. Peltz, C. Liebald, X. Mateos, B. Deppe, and C. Kränkel, “Thermo-optic properties of Yb:Lu2O3 single crystals,” Appl. Phys. B 120(4), 601–607 (2015).

4. K. Petermann, G. Huber, L. Fornasiero, S. Kuch, E. Mix, V. Peters, and S. A. Basun, “Rare-earth-doped sesquioxides,” J. Lumin. 87–89, 973–975 (2000).

5. P. Koopmann, R. Peters, K. Petermann, and G. Huber, “Crystal growth, spectroscopy and highly efficient laser operation of thulium-doped Lu2O3 around 2 μm,” Appl. Phys. B 102(1), 19–24 (2011).

6. P. Koopmann, S. Lamrini, K. Scholle, P. Fuhrberg, K. Petermann, and G. Huber, “Efficient diode-pumped laser operation of Tm:Lu2O3 around 2 μm,” Opt. Lett. 36(6), 948–950 (2011). [PubMed]  

7. P. Koopmann, S. Lamrini, K. Scholle, M. Schäfer, P. Fuhrberg, and G. Huber, “Holmium-doped Lu2O3, Y2O3, and Sc2O3 for lasers above 2.1 μm,” Opt. Express 21(3), 3926–3931 (2013). [PubMed]  

8. A. Schmidt, P. Koopmann, G. Huber, P. Fuhrberg, S. Y. Choi, D.-I. Yeom, F. Rotermund, V. Petrov, and U. Griebner, “175 fs Tm:Lu2O3 laser at 2.07 µm mode-locked using single-walled carbon nanotubes,” Opt. Express 20(5), 5313–5318 (2012). [PubMed]  

9. M. Tokurakawa, A. Shirakawa, K. Ueda, H. Yagi, M. Noriyuki, T. Yanagitani, and A. A. Kaminskii, “Diode-pumped ultrashort-pulse generation based on Yb3+:Sc2O3 and Yb3+:Y2O3 ceramic multi-gain-media oscillator,” Opt. Express 17(5), 3353–3361 (2009). [PubMed]  

10. A. A. Lagatsky, P. Koopmann, P. Fuhrberg, G. Huber, C. T. A. Brown, and W. Sibbett, “Passively mode locked femtosecond Tm:Sc2O3 laser at 2.1 μm,” Opt. Lett. 37(3), 437–439 (2012). [PubMed]  

11. R. Peters, C. Kränkel, K. Petermann, and G. Huber, “Broadly tunable high-power Yb:Lu2O3 thin disk laser with 80% slope efficiency,” Opt. Express 15(11), 7075–7082 (2007). [PubMed]  

12. C. R. E. Baer, C. Kränkel, C. J. Saraceno, O. H. Heckl, M. Golling, R. Peters, K. Petermann, T. Südmeyer, G. Huber, and U. Keller, “Femtosecond thin-disk laser with 141 W of average power,” Opt. Lett. 35(13), 2302–2304 (2010). [PubMed]  

13. R. C. Stoneman and L. Esterowitz, “Efficient, broadly tunable, laser-pumped Tm:YAG and Tm:YSGG cw lasers,” Opt. Lett. 15(9), 486–488 (1990). [PubMed]  

14. P. Loiko and M. Pollnau, “Stochastic model of energy-transfer processes among rare-earth ions. Example of Al2O3:Tm3+,” J. Phys. Chem. C 120(46), 26480–26489 (2016).

15. P. Koopmann, S. Lamrini, K. Scholle, P. Fuhrberg, K. Petermann, and G. Huber, “Laser operation and spectroscopic investigations of Tm:LuScO3,” in CLEO/Europe - EQEC 2011 (Optical Society of America, 2011), P. CA1_4.

16. A. A. Lagatsky, P. Koopmann, O. L. Antipov, C. T. A. Brown, G. Huber, and W. Sibbett, “Femtosecond pulse generation with Tm-doped sesquioxides,” in CLEO/Europe - EQEC 2011 (Optical Society of America, 2013), P. CA_6_3.

17. R. Peters, C. Krankel, K. Petermann, and G. Huber, “Crystal growth by the heat exchanger method, spectroscopic characterization and laser operation of high-purity Yb:Lu2O3,” J. Cryst. Growth 310(7–9), 1934–1938 (2008).

18. O. L. Antipov, A. A. Novikov, N. G. Zakharov, and A. P. Zinov’ev, “Optical properties and efficient laser oscillation at 2066 nm of novel Tm:Lu2O3 ceramics,” Opt. Mater. Express 2(2), 183–189 (2012).

19. O. L. Antipov, A. A. Novikov, N. G. Zakharov, A. P. Zinoviev, H. Yagi, N. V. Sakharov, M. V. Kruglova, M. O. Marychev, O. N. Gorshkov, and A. A. Lagatskii, “Efficient 2.1-μm lasers based on Tm3+:Lu2O3 ceramics pumped by 800-nm laser diodes,” Phys. Status Solidi 10(6), 969–973 (2013).

20. A. A. Lagatsky, O. L. Antipov, and W. Sibbett, “Broadly tunable femtosecond Tm:Lu2O3 ceramic laser operating around 2070 nm,” Opt. Express 20(17), 19349–19354 (2012). [PubMed]  

21. E. J. Saarinen, E. Vasileva, O. Antipov, J.-P. Penttinen, M. Tavast, T. Leinonen, and O. G. Okhotnikov, “2-µm Tm:Lu2O3 ceramic disk laser intracavity-pumped by a semiconductor disk laser,” Opt. Express 21(20), 23844–23850 (2013). [PubMed]  

22. X. Xu, Z. Hu, D. Li, P. Liu, J. Zhang, B. Xu, and J. Xu, “First laser oscillation of diode-pumped Tm3+-doped LuScO3 mixed sesquioxide ceramic,” Opt. Express 25(13), 15322–15329 (2017). [PubMed]  

23. L. Laversenne, Y. Guyot, C. Goutaudier, M. T. Cohen-Adad, and G. Boulon, “Optimization of spectroscopic properties of Yb3+-doped refractory sesquioxides: cubic Y2O3, Lu2O3 and monoclinic Gd2O3,” Opt. Mater. 16(4), 475–483 (2001).

24. N. D. Todorov, M. V. Abrashev, V. Marinova, M. Kadiyski, L. Dimowa, and E. Faulques, “Raman spectroscopy and lattice dynamical calculations of Sc2O3 single crystals,” Phys. Rev. B 87(10), 104301 (2013).

25. O. Antipov, A. Novikov, S. Larin, and I. Obronov, “Highly efficient 2 μm CW and Q-switched Tm3+:Lu2O3 ceramics lasers in-band pumped by a Raman-shifted erbium fiber laser at 1670 nm,” Opt. Lett. 41(10), 2298–2301 (2016). [PubMed]  

26. B. R. Judd, “Optical absorption intensities of rare-earth ions,” Phys. Rev. 127(3), 750–761 (1962).

27. G. S. Ofelt, “Intensities of crystal spectra of rare-earth ions,” J. Chem. Phys. 37(3), 511–520 (1962).

28. L. Zhang, H. Lin, G. Zhang, X. Mateos, J. M. Serres, M. Aguiló, F. Díaz, U. Griebner, V. Petrov, Y. Wang, P. Loiko, E. Vilejshikova, K. Yumashev, Z. Lin, and W. Chen, “Crystal growth, optical spectroscopy and laser action of Tm3+-doped monoclinic magnesium tungstate,” Opt. Express 25(4), 3682–3693 (2017). [PubMed]  

29. C. M. Dodson and R. Zia, “Magnetic dipole and electric quadrupole transitions in the trivalent lanthanide series: Calculated emission rates and oscillator strengths,” Phys. Rev. B 86(12), 125102 (2012).

30. P. Koopmann, “Thulium- and holmium-doped sesquioxides for 2 μm lasers,” Ph.D. dissertation, Dept. of Physics, University of Hamburg, Hamburg, Germany, 2012.

31. A. S. Yasyukevich, V. G. Shcherbitskii, V. E. Kisel’, A. V. Mandrik, and N. V. Kuleshov, “Integral method of reciprocity in the spectroscopy of laser crystals with impurity centers,” J. Appl. Spectrosc. 71(2), 202–208 (2004).

32. B. F. Aull and H. P. Jenssen, “Vibronic interactions in Nd:YAG resulting in nonreciprocity of absorption and stimulated emission cross sections,” IEEE J. Quantum Electron. 18(5), 925–930 (1982).

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 (8)

Fig. 1
Fig. 1 Some of the fabricated 4.76 at.% Tm:(Lu,Sc)2O3 ceramics (laser-grade-polished ceramic disks) (a); X-ray diffraction (XRD) pattern (in blue) compared to the standard pattern of ScLuO3 (in red) (b); Raman spectrum, excitation wavelength: 514 nm (c).
Fig. 2
Fig. 2 Field-emission scanning electron microscopy (SEM) images of a polished (a) and a fracture (b) surface of the 4.76 at.% Tm:(Lu,Sc)2O3 ceramic. The scale bar in (a,b) is 20 µm.
Fig. 3
Fig. 3 (a-f) Room-temperature (at 293 K, in red) and low-temperature (at 6 K, in blue, not in scale) absorption spectra of the 4.76 at.% Tm:(Lu,Sc)2O3 ceramic.
Fig. 4
Fig. 4 Spectroscopy of Tm3+-doped (Lu,Sc)2O3 ceramic: (a) absorption, σabs, and stimulated-emission, σSE, cross-sections for the 3H63F4 transition; (b) gain cross-sections, σg = βσSE – (1 – β)σabs, for the 3F43H6 transition. β = N(3F4)/NTm is the inversion ratio.
Fig. 5
Fig. 5 Tm:(Lu,Sc)2O3 ceramic microchip laser: (a) scheme of the laser, LD – laser diode, PM – pump mirror, OC – output coupler; (b) laser beam profile captured at Pabs = 5.05 W, corresponding to an output power of 1.01 W for TOC = 3%.
Fig. 6
Fig. 6 Tm:(Lu,Sc)2O3 ceramic microchip laser: (a) input-output dependences, η – slope efficiency; (b) typical laser emission spectra measured at Pabs = 5.05 W (for TOC = 1.5% and 3%) and 2.40 W (TOC = 5%).
Fig. 7
Fig. 7 Spectroscopy of Tm3+ in a 4.76 at.% Tm:(Lu2/3Sc1/3)2O3 ceramic and in a 1 at.% Tm:LuScO3 single-crystal: (a) σabs spectra for the 3H63H4 transition, (b) σSE spectra for the 3F43H6 transition. The noise of the black curve in (b) is due to water absorption affecting the measured luminescence spectrum.
Fig. 8
Fig. 8 Tm:LuScO3 single crystal microchip laser: (a) input-output dependences, η – slope efficiency; (b) typical laser emission spectra measured at Pabs = 1.41 W.

Tables (2)

Tables Icon

Table 1 Experimental and Calculated Absorption Oscillator Strengths for 4.76 at.% Tm:(Lu,Sc)2O3 Ceramic

Tables Icon

Table 2 Calculated Emission Probabilities for Tm3+ in (Lu,Sc)2O3 Ceramic

Equations (6)

Equations on this page are rendered with MathJax. Learn more.

f Σ exp (JJ')= m e c 2 π e 2 N Tm λ 2 Γ(JJ'),
f Σ calc (JJ')= 8 3h(2J'+1)λ ( n 2 +2) 2 9n S ED calc (JJ')+ f MD calc (JJ'),
S ED calc (JJ')= k=2,4,6 U (k) Ω k , where U (k) = (4 f n )SLJ|| U k ||(4 f n )S'L'J' 2 .
A Σ calc (JJ')= 64 π 4 e 2 3h(2J'+1) λ 3 n ( n 2 +2 3 ) 2 S ED calc (JJ')+ A MD calc (JJ').
τ rad = 1 A tot and B(JJ')= A Σ calc (JJ') A tot , where A tot = J' A Σ calc (JJ') .
σ SE (λ)= 1 8π n 2 τ rad c σ abs (λ) e hc/(kTλ) λ 4 σ abs (λ) e hc/(kTλ) dλ .
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.