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

Growth, spectroscopy and laser operation of Tm,Ho:GdScO3 perovskite crystal

Open Access Open Access

Abstract

We report on the growth, polarized spectroscopy and first laser operation of an orthorhombic (space group Pnma) Tm3+,Ho3+-codoped gadolinium orthoscandate (GdScO3) perovskite-type crystal. A single crystal of 3.76 at.% Tm, 0.35 at.% Ho:GdScO3 was grown by the Czochralski method. Its polarized absorption and fluorescence properties were studied revealing a broadband emission around 2 µm. The parameters of the Tm3+ ↔ Ho3+ energy transfer was quantified, P28 = 1.30 × 10−22 cm3µs-1, and P71 = 0.99 × 10−23 cm3µs-1, and the thermal equilibrium lifetime was measured to be 3.5 ms. The crystal-field splitting of Tm3+ and Ho3+ multiplets in Cs symmetry sites of the perovskite structure was determined by low-temperature spectroscopy and the mechanism of spectral line broadening is discussed. The continuous-wave Tm,Ho:GdScO3 laser generated 1.16 W at ∼2.1 µm with a slope efficiency of 50.5%, a laser threshold of 184 mW, a linear laser polarization (E || c) and a spatially single-mode output. The Tm,Ho:GdScO3 crystal is promising for broadly tunable and femtosecond mode-locked lasers emitting above 2 µm.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Perovskites are materials with a chemical formula of ABX3 where A and B are cations, often of rather different sizes, and X is an anion. This class of materials is named after the mineral perovskite, calcium titanium oxide CaTiO3. The perovskite structure is known for permitting diverse ionic substitutions at both cationic sites due to its high tolerance to structural distortions. Perovskites exhibit exciting compositionally driven ferro-, piezo-, and pyro-electrical properties as well as electro-optical effects, and they have been used as electric, magnetic, optical and catalytic materials in various applications [1,2].

Perovskite oxides, ABO3, are relevant as host materials for doping with trivalent rare-earth ions for luminescent and laser materials [3]. In ABO3 crystals, cations with a large ionic radius occupy A-sites and coordinate to XII oxygen atoms and those with a small ionic radius reside in VI-fold coordinated B-sites. The idealized form of ABO3 compounds is cubic ($Pm\bar{3}m$), e.g., SrTiO3 and BaTiO3. More often, the structure is modified by cation displacement or by tilting of octahedra lowering the crystal symmetry. The orthorhombic CaTiO3 phase (space group Pnma) is among the most common non-cubic variants. Orthorhombic yttrium aluminium perovskite, YAlO3 (shortly YAP) is a well-known example of laser host material.

Among the ABO3 perovskites, rare-earth (RE) orthoscandates AScO3 with A = La to Ho have attracted a lot of attention in the past years. Their structures have been known since 1957 when they were presented by Geller [4] and further described by other authors [58]. For predicting the stability of ABO3 perovskite structures, the Goldschmidt tolerance factor t is used [9]. The t factors for REScO3 crystals are lower compared to other REBO3 compounds suggesting that orthoscandates form highly distorted structures [7]. Gadolinium orthoscandate, GdScO3, adopts the GdFeO3 type structure (orthorhombic, sp. gr. Pnma, lattice constants: a = 5.745(1) Å, b = 7.929(2) Å, c = 5.481(1) Å [8]). Its Goldschmidt tolerance factor t is relatively low, 0.808, but well within the range of stable orthorhombic perovskites [7].

The structure of GdScO3 is composed of a framework of corner-connected [ScO6] octahedra and Gd3+ in interstices surrounded by eight [ScO6] octahedra, see Fig. 1. The orthorhombic structure is derived from the ideal ABO3 cubic (sp. gr. Pm$\bar{3}$m) aristotype by tilting about the axes of the cubic unit cell running parallel to the edges of the cubes formed by eight [ScO6] octahedra in an ab + a pattern (Glazer notation [10]) as a response to the occupancy of the A-site by a cation which is smaller than that required to form the ideal cubic structure [7]. The octahedron tilting results in a symmetry reduction. It is also known that for rare-earth orthoscandates, the RE3+ coordination polyhedron in the A-site is best regarded as [REO8] rather than [REO12], i.e., a four-fold anti-prism [7].

 figure: Fig. 1.

Fig. 1. The crystal structure of an orthorhombic GdScO3 crystal: (a) a fragment of the structure, black lines mark the unit-cell; (b) schematic of the [ScO6] (left) and [GdO8] (right) polyhedra showing the metal-to-oxygen (M – O) interatomic distances. The atomic coordinates reported by Veličkov et al. [8] are used.

Download Full Size | PDF

Very recently, GdScO3 crystals gained increasing attention for doping with laser-active RE3+ ions targeting ultrafast laser development in the near-infrared. GdScO3 melts congruently at a high temperature of ∼2100°C and can be grown by the standard Czochralski (Cz) method. As a host material, it features good thermo-mechanical properties, namely a relatively high thermal conductivity of about 5.5 Wm-1K-1 [11], and moderate average thermal expansion <α> of 10.9 × 10−6 K-1 [12] enabling efficient thermal management under high-power pumping. The low maximum phonon energy of 665 cm-1 [13] suppresses non-radiative relaxation from the excited state of RE3+ ions. GdScO3 is an optically biaxial crystal with a high refractive index (<n > = 2.07 at ∼1 µm [14]) and it features a notable polarization anisotropy of absorption and emission properties of the dopant RE3+ ions which leads to linearly polarized laser emission.

The spectroscopic properties of few RE3+ ions, including Nd3+ [14,15], Yb3+ [11], Tm3+ [16], Ho3+ [17], Er3+ [18], and Dy3+ [19], in GdScO3 have been studied. Laser operation of RE3+-doped GdScO3 crystals in the near infrared has also been described [2024]. Li et al. reported on the crystal growth, optical spectroscopy and Judd-Ofelt analysis of Tm:GdScO3 crystals [16]. Continuous-wave (CW) laser operation of this material was achieved with diode pumping yielding 4.53 W at 1988nm with a slope efficiency of 25% [21]. Liu et al. achieved a continuous wavelength tuning of a Tm:GdScO3 laser from 1845 to 2006nm [23]. Recently, Zhang et al. demonstrated femtosecond mode-locked (ML) Tm:GdScO3 laser operation: using a GaSb-based SEmiconductor Saturable Absorber Mirror, 44 fs pulses were generated at 2048nm with an average output power of 188 mW at a repetition rate of 77.6 MHz [24]. The growth, unpolarized spectroscopy and Judd-Ofelt analysis of Ho:GdScO3 crystals was reported by Hu et al. [17].

Holmium ions (Ho3+) are appealing because of their emission around 2.1 µm due to the 5I75I8 electronic transition. This emission does not overlap with the structured absorption lines of water vapor in the atmospheric air enabling generation of femtosecond pulses from ML lasers [25]. The Tm3+,Ho3+ co-doping scheme is often used for achieving Ho laser emission [26,27]. It relies on the strong absorption of Tm3+ ions at ∼0.8 µm (the 3H63H4 transition) followed by a cross-relaxation (CR) process, 3H4 + 3H63F4 + 3F4, populating the Tm3+ metastable state (3F4) in a “two-for-one” pump process, and a non-radiative energy-transfer (ET), 3F4 (Tm3+) → 5I8 (Ho3+), as illustrated in Fig. 2. The Tm3+,Ho3+ co-doping is also beneficial for combining the gain bandwidths of both ions supporting the generation of shorter pulses from ML lasers [28,29].

 figure: Fig. 2.

Fig. 2. A simplified energy level scheme of Tm3+ and Ho3+ ions showing processes relevant for laser operation around 2.1 µm: NR – multiphonon non-radiative relaxation, CR – cross-relaxation, ETU – energy-transfer upconversion, P28 and P71 – energy-transfer processes.

Download Full Size | PDF

In the present work, we report on the Cz growth, a detailed polarized room- and low-temperature spectroscopic study, and first laser operation of a Tm,Ho:GdScO3 crystal. The Tm,Ho:GdScO3 laser delivered a watt-level output at 2.1 µm with a high slope efficiency of about 50%.

2. Crystal growth

The Tm,Ho:GdScO3 crystal was grown by the Cz method in an induction-heating furnace. The raw materials were oxide powders of Gd2O3, Sc2O3, Tm2O3 and Ho2O3 with the purity of 99.99%. The raw materials were accurately weighed in the stoichiometric ratios with the doping concentrations of Tm3+ and Ho3+ ions of 5 at.% and 0.5 at.% (with respect to Gd3+) in GdScO3, respectively. The total weighed mass is 75% of the total mass that can be held in the crucible, considering that the growth charge expands as it melts. After thorough mixing, the raw materials were pressed into tablets with a diameter of 60 mm. Then, the tablets were sintered in air at 1500°C for 10 h. The crystal growth took place in an iridium crucible, with ZrO2 as an isolator in argon atmosphere. The power used for crystal growth was about 8 kW, which corresponds to the melting point of Tm,Ho:GdScO3 of about 2100 °C. The single crystal was grown using an [110] oriented seed from an undoped GdScO3 crystal. The complete process included the following steps: furnace loading, vacuumation, argon charging, heating, melting, crystal growth and slow cooling down to room temperature for 20 h. The pulling rate was 1.0 mm/h and the rotation speed was 10 r.p.m. (revolutions per minute). A photograph of an as-grown Tm,Ho:GdScO3 crystal is shown in Fig. 3(a). The boule had a cylindrical shape with the dimensions of Φ30 × 30 mm3. It was free of cracks, inclusions and bubbles and had a yellow-brown coloration due to color centers formed during the growth in an oxygen deficient atmosphere. This coloration was subsequently removed by annealing in air.

 figure: Fig. 3.

Fig. 3. Tm, Ho:GdScO3 crystal: (a) a photograph of the as-grown crystal, the growth direction is along the [110] orientation; (b) X-ray powder diffraction (XRD) pattern, numbers denote the Miller’s indices, (hkl), black lines – the reference pattern of GdScO3.

Download Full Size | PDF

The phase purity and the crystal structure of Tm,Ho:GdScO3 were confirmed by X-ray powder diffraction (XRD), see Fig. 3(b). All the reflections in the diffraction pattern were assigned to the orthorhombic phase isostructural to undoped GdScO3 (PDF card No. 79 - 0577). The characteristic GdFeO3 type structure has four distorted perovskite units in the crystallographic cell (space group Pnma – D16 2 h, No. 62). According to previous studies, it is expected that the dopant Tm3+ and Ho3+ ions replace for the host forming Gd3+ cations in Сs symmetry sites (Wyckoff: 4c) [3032].

The actual composition of the grown crystal was studied by Inductively Coupled Plasma Atomic Spectroscopy (ICP-AS). The RE doping levels were 3.76 at.% Tm, 0.35 at.% Ho (corresponding to ion densities of NTm = 6.25 × 1020 at/cm3 and NHo = 0.58 × 1020 at/cm3). The RE ion segregation coefficients, KRE = Ccrystal/Cmelt, are KTm = 0.75 and KHo = 0.70.

3. Results and discussion

3.1 Raman spectra

The room-temperature polarized Raman spectra of the Tm,Ho:GdScO3 crystal are presented in Fig. 4. Four polarization configurations, ${\boldsymbol a}({ij} )\bar{{\boldsymbol a}}$, where i, j = b and c (Porto’s notations) were used to identify vibration modes of different symmetry. For orthorhombic RE scandates (sp. gr. Pnma), according to the factor group analysis, four formula units consisting of 20 atoms per unit cell give rise to 60 vibrational modes at the center of the Brillouin zone (k=0): Γop=7Ag+5B1g+7B2g+5B3g+8Au+9B1u+7B2u+9B3u, of which 24 modes are Raman-active (Ag, B1g, B2g, and B3g), 25 modes are IR-active (B1u, B2u, and B3u) and 8 modes are silent (Au), and Γac=B1u+B2u+B3u [13]. The Raman spectra of Tm,Ho:GdScO3 are typical for orthorhombic perovskites and are strongly polarized. 22 active Raman modes were found and assigned out of 24 predicted ones. Note that to date, none of the perovskites has shown all the modes. The number of modes depends on the degree of distortion of the ideal perovskite structure. In the low-frequency range (below 200cm-1), the Raman modes are due to A-site (Gd3+) translations, while those at higher energies (above 200cm-1) are due to oxygen motions (bending, stretching, and tilting) in the [AO8] units. In orthorhombic symmetry, the B cation (Sc3+) occupies a site of inversion symmetry (Ci) and thus, the Raman spectra of such perovskites should not have modes which are due to B cation translations or vibrations. The dominant Raman peak appears at 320cm-1 (Ag) and the maximum phonon energy is 665cm-1 (B3g). The relatively low maximum phonon energy of GdScO3 is beneficial for reducing the multiphonon non-radiative relaxation rate from the 5I7 Ho3+ state.

 figure: Fig. 4.

Fig. 4. Polarized Raman spectra of an a-cut Tm,Ho:GdScO3 crystal, λexc = 514 nm, numbers - Raman frequencies in cm−1.

Download Full Size | PDF

3.2 Polarized absorption spectra

The orthorhombic GdScO3 crystal is optically biaxial, and its optical indicatrix frame aligns with the crystallographic axes. The spectroscopic properties of Tm3+,Ho3+-codoped GdScO3 were studied for the principal light polarizations E || a, b, c.

The room-temperature polarized absorption spectrum of Tm,Ho:GdScO3 is presented in Fig. 5(a-c) in terms of the absorption coefficient αabs. The assignment of electronic transitions of Tm3+ and Ho3+ ions follows that by Carnall et al. [33]. The broad absorption in the visible (320 – 500 nm) is assigned to color centers that are not entirely removed upon annealing in air [34]. Absorption bands related to transitions from the 3H6 Tm3+ and 5I8 Ho3+ ground states to the excited-states ranging from 3F4 Tm3+ and 5I7 Ho3+ up to 1D2 Tm3+ and 3L9 + 5G3 Ho3+ (wavelength range: 320 – 2150 nm) are observed. The UV absorption edge is observed at ∼320 nm (3.87 eV) whereas for undoped GdScO3, the optical bandgap Eg is 5.8 eV [35].

 figure: Fig. 5.

Fig. 5. Polarized absorption properties of the Tm,Ho:GdScO3 crystal: (a-c) absorption spectra; (d) absorption cross-sections, σabs, for the 3H63F4 Tm3+ transition, arrow indicates the pump wavelength.

Download Full Size | PDF

The absorption cross-sections, σabs = αabs/NTm, for the 3H63H4 Tm3+ transition around 0.8 µm are shown in Fig. 5(d). This band is suitable for pumping Tm3+,Ho3+-codoped crystals by high-brightness Ti:Sapphire lasers or commercial high-power spatially multimode AlGaAs laser diodes. The maximum σabs value reaches 1.52 × 10−20 cm2 at 794.2 nm and the absorption bandwidth (full width at half maximum, FWHM) is as large as 8.0 nm (for light polarization E || b). For the other two polarizations, the absorption is weaker, σabs = 0.93 × 10−20 cm2 at 792.4 nm (for E || a) and σabs = 0.75 × 10−20 cm2 at 801.5 nm (for E || c). The relatively broad absorption of Tm3+ ions around 0.8 µm is beneficial for suppressing the effect of the temperature drift of the pump diode wavelength.

Recently, resonant (in-band) Tm3+ pumping was considered as a viable way for power scaling of Tm,Ho lasers. It corresponds to the 3H63F4 Tm3+ transition. Raman-shifted Er-fiber lasers emitting at 1.68 µm can address this absorption band. For Tm3+ ions in GdScO3, the maximum σabs value is 0.98 × 10−20 cm2 at 1673.6 nm for light polarization E || b. For the other two polarizations, the absorption is weaker, σabs = 0.61 × 10−20 cm2 at 1670.7 nm (for E || a) and σabs = 0.67 × 10−20 cm2 at 1670.6 nm (for E || c).

3.3 Luminescence spectra and Tm3+ → Ho3+ energy transfer

The luminescence spectra of the Tm,Ho:GdScO3 crystal in the near-IR are shown in Fig. 6(a), measured for the three principal light polarizations E || a, b, c under excitation at 794 nm (into the 3H4 Tm3+ manifold). The spectra are strongly polarized and exhibit a notable spectral line broadening which is attributed to the strong electron-phonon coupling for this host crystal (note that homogeneous thermal line broadening is expected to be much stronger for Tm3+ ions than for Ho3+ ones in line with the measured spectra). The weak band at 1.37 – 1.53 µm is due to the 3H43F4 Tm3+ transition, and a broad band spanning from 1.55 to 2.3 µm is assigned to two spectrally overlapping transitions, namely 3F43H6 Tm3+ and 5I75I8 Ho3+. The latter emission is more intense despite the low Ho:Tm codoping ratio (1:10.7, according to the actual rare-earth ion densities), indicating an efficient Tm3+ → Ho3+ energy transfer. In the spectral range where laser operation is expected according to the quasi-three-level nature of the Ho3+ laser scheme with reabsorption from the ground-state (5I8), higher emission intensity corresponds to light polarization E || c and several local peaks are found in the luminescence spectra centered at 2022, 2064, 2098, and 2111 nm.

 figure: Fig. 6.

Fig. 6. Emission properties of Tm,Ho:GdScO3: (a) polarized luminescence spectra in the near-IR, light polarizations: E || a, b, c, arrows indicate the observed laser wavelengths (for light polarization E || c); (b) a comparison of unpolarized emission spectra of Tm3+,Ho3+-codoped and singly Tm3+-doped GdScO3 crystals around 2 µm, λexc = 794 nm.

Download Full Size | PDF

The luminescence dynamics from the lower-lying excited states of Tm3+ and Ho3+ ions in the Tm,Ho:GdScO3 crystal, which are responsible for laser emission around 2 µm, were studied under resonant Tm3+ excitation (at 1635 nm, to the 3F4 state). The luminescence was detected at two wavelengths corresponding to almost pure Tm3+ (1870nm) and Ho3+ (2040nm) emissions. The measured luminescence decay curves from the 3F4 Tm3+ and 5I7 Ho3+ manifolds plotted on a semi-log scale are shown in Fig. 7. In the initial decay stage, a rapid rise of Ho3+ emission intensity is observed within a few hundred microseconds. At the same time, a corresponding fast drop in the intensity of Tm3+ luminescence is observed. Before the thermal equilibrium is reached between both active ions, this phenomenon corresponds to a direct energy transfer from Tm3+ to Ho3+. The luminescence decay rates of the 3F4 Tm3+ and 5I7 Ho3+ manifolds become almost identical at a longer time scale, about 1 ms after the excitation pulse. During this phase, energy transfer occurs in both directions. However, it is masked by a quasi-equilibrium between the respective multiplets.

 figure: Fig. 7.

Fig. 7. Luminescence decay curves from the 3F4 Tm3+ and 5I7 Ho3+ manifolds measured under resonant Tm3+ excitation (λexc = 1635 nm), circles – experimental data, curves – their fits using Eq. (1), τ0 – thermal equilibrium decay time.

Download Full Size | PDF

The parameters of the bidirectional 3F4(Tm3+) ↔ 5I7(Ho3+) energy-transfer were estimated using the dynamical model developed by Walsh et al. [36]. The luminescence decay curves were fitted using the following equations:

$$\frac{{{n_2}(t)}}{{{n_2}(0)}} = \frac{\beta }{{\alpha + \beta }}\textrm{exp} ( - \frac{t}{\tau }) + \frac{\alpha }{{\alpha + \beta }}\textrm{exp} ( - (\alpha + \beta )t),$$
$$\frac{{{n_7}(t)}}{{{n_7}(0)}} = \frac{\beta }{{\alpha + \beta }}\textrm{exp} ( - \frac{t}{\tau }) - \frac{\alpha }{{\alpha + \beta }}\textrm{exp} ( - (\alpha + \beta )t),$$
where t is time after a short-pulse excitation, τ is the thermal equilibrium decay time, n2 and n7 are the electronic populations of the 3F4 Tm3+ and 5I7 Ho3+ manifolds, respectively, α = P28NHo and β = P71NTm are the transfer rates, where P28 is a parameter representing a direct nonradiative transfer of energy from Tm3+ to Ho3+ and P71 is a parameter representing back nonradiative energy transfer. The best-fit curves are shown in Fig. 7. The corresponding fit parameters are τ = 3.5 ± 0.5 ms, P28 = 1.30 ± 0.03 × 10−22 cm3µs−1 and P71 = 0.99 ± 0.03 × 10−23 cm3µs−1. The ratio of the energy-transfer parameters, Θ = P71/P28, known as the equilibrium constant, amounts to 0.076. It shows how the Tm3+ and Ho3+ ions share the excitation energy, and its value confirms the predominantly direct Tm3+ → Ho3+ energy-transfer. The Θ value for the Tm,Ho:GdScO3 crystal is close to that for Tm,Ho:LiYF4 (Θ = 0.076) [36], another well-known crystal for 2.1-µm lasers, and is lower than that for Tm,Ho:Y3Al5O12 (Θ = 0.120) [36]. A comparison of Tm3+ → Ho3+ energy-transfer parameters and thermal equilibrium decay times for various oxide and fluoride laser crystals is given in Table 1.

Tables Icon

Table 1. Thulium-Holmium Energy-Transfer Parameters for Various Laser Crystals

The equilibrium constant for a Tm3+,Ho3+-codoped crystal can be independently estimated from the crystal-field splitting of Tm3+ and Ho3+ multiplets involved in the energy-transfer [36]:

$$\Theta = \frac{{{Z_2}{Z_8}}}{{{Z_1}{Z_7}}}\textrm{exp} [\frac{{ - (E_{ZPL}^{Tm} - E_{ZPL}^{Ho})}}{{kT}}], $$
where Z1, Z2, Z7 and Z8 are the partition functions of the 3H6, 3F4 Tm3+ and 5I7, 5I8 Ho3+ states, respectively, EZPL are the energies of the zero-phonon-line (ZPL) transitions for Tm3+ and Ho3+ occurring between the lowest Stark sub-levels of their ground and excited manifolds, k is the Boltzmann constant and T is the crystal temperature (room temperature, RT). The partition functions were calculated using the experimental Stark splitting for Tm3+ and Ho3+ in GdScO3 determined in the present work (see below): Z1 = 4.967, Z2 = 5.213 (Tm3+) and Z7 = 7.249, Z8 = 7.530 (Ho3+). From Eq. (2), we arrive at Θ = 0.054 which is underestimated as compared with the value determined from the luminescence decay studies. Equation (2) considers only energy transfers between nearly resonant Stark sub-levels, while the Θ value obtained from the measured luminescence decay curves also considers phonon-assisted processes when phonons conserve the energy for non-resonant Stark levels.

3.4 Low-temperature spectroscopy

The dopant trivalent RE ions (Tm3+ and Ho3+, in our case) in GdScO3 predominantly substitute for the Gd3+ host-forming cations in Cs symmetry sites with an VIII-fold oxygen coordination (Wyckoff: 4c). The corresponding ionic radii RGd = 1.053 Å, RTm = 0.994 Å and RHo = 1.015 Å. The Sc3+ sites (Wyckoff: 4b) feature a VI-fold oxygen coordination (RSc = 0.745 Å). The 2S + 1LJ multiplets with integer total angular momentum J in the crystal-field of the symmetry Cs are split into a total of 2J + 1 Stark components. In the present work, we determined the crystal-field splitting of the two lowest multiplets of Tm3+ (3H6 and 3F4) and Ho3+ (5I8 and 5I7) in GdScO3 as they are relevant for describing the laser emission properties around 2 µm and energy-transfer between these two dopant ions.

The low-temperature (LT, 12 K) absorption and luminescence spectra corresponding to the 3H63F4 Tm3+ and 5I85I7 Ho3+ transitions were measured for singly Tm3+-doped and Tm3+,Ho3+-codoped GdScO3 crystals, respectively, using linearly polarized light, as shown in Fig. 8. The LT absorption spectra were used to resolve the crystal-field splitting of the RE3+ excited-states and they plotted vs. the photon energy (expressed in cm−1). The LT emission spectra were measured to determine the Stark sub-level energies of the RE3+ ground-states and they are plotted as a function of (EZPL – photon energy, expressed in cm−1). For labelling electronic transitions, we used the following phenomenological notations for Stark sub-levels: Yi, Zj, Yk, Zl for the 3F4 (I = 1 - 9), 3H6 (j= 1 - 13), 5I7 (k = 1 - 15), and 5I8 (l = 1 - 17) multiplets, respectively [43,44]. For the assignment of electronic transitions of Tm3+ and Ho3+ ions in GdScO3, we followed the previous work on the isostructural GdAlO3 crystal (sp. gr. Pnma) doped with Tm3+ and Ho3+ ions (Gd3+ site symmetry: Cs). The vertical dashes in Fig. 8 mark the electronic transitions for this reference crystals.

 figure: Fig. 8.

Fig. 8. Low-temperature (12 K) spectroscopy of Tm3+ and Ho3+ ions in the GdScO3 crystal: (a,c) absorption spectra: (a) the 3H63F4 Tm3+ transition; (c) the 5I85I7 Ho3+ transition; (b,d) luminescence spectra: (b) the 3F43H6 Tm3+ transition; (d) the 5I75I8 Ho3+ transition, “+” indicate the assigned electronic transitions. Vertical dashes – experimental Stark splitting for Tm3+ and Ho3+ ions in Cs sites of the GdAlO3 crystal (after [41,42]).

Download Full Size | PDF

Figure 9 presents the determined experimental crystal-field splitting of the (3H6 and 3F4) Tm3+ and (5I8 and 5I7) Ho3+ multiplets in GdScO3. The ZPL energies EZPL are 5712 cm−1 (Tm3+) and 5110 cm−1 (Ho3+). The total Stark splitting of the ground-state amounts to ΔE(3H6) = 681 cm−1 (Tm3+) and ΔE(5I8) = 382 cm−1 (Ho3+). Table 2 lists the Stark sub-level energies for the considered multiplets.

 figure: Fig. 9.

Fig. 9. Experimental crystal-field splitting of two lowest multiplets of Tm3+ and Ho3+ ions in the GdScO3 crystal. EZPL – zero-phonon line, Z1, Z2, Z7 and Z8 are the partition functions of the 3H6, 3F4 Tm3+ and 5I7, 5I8 Ho3+ multiplets, respectively.

Download Full Size | PDF

Tables Icon

Table 2. Experimental Crystal-Field Splitting of the 3H6, 3F4,Tm3+ and 5I8, 5I7 Ho3+ Multiplets in the GdScO3 crystal

Both Tm3+ and Ho3+ ions in the GdScO3 crystal exhibit broad and smooth absorption and emission spectra at room temperature for polarized light. The large spectral bandwidths may originate either from the homogeneous (thermal) broadening owing to a strong electron-phonon interaction with the host matrix or from the inhomogeneous broadening mechanisms.

Amanyan et al. studied Nd3+-doped GdScO3 crystals and indicated that for this compound, the spectral linewidths at room temperature were substantially larger than those for ordered orthoaluminate (YAlO3) and aluminium garnet (Y3Al5O12) crystals [14]. These authors also suggested that the relatively large Stark splitting of RE3+ multiplets in GdScO3 comes from the low-symmetry component in the crystal-field potential expansion indicating a strong low-symmetry distortion in the local symmetry of the activator center [14]. Guo et al. studied the LT fluorescence of Yb3+ ions in GdScO3 and indicated that the energy transitions of Yb3+ were almost independent of the excitation wavelength revealing a single type of Yb3+ species [30]. Zhang et al. indicated that although the shape of [Gd|REO8] polyhedra is the same through the GdScO3 unit-cell, their orientations are different which may contribute to additional inhomogeneous spectral line broadening [11]. However, no evidence of the latter effect was provided.

In the GdScO3 crystal, there are two possible sites for RE substitution, namely A-sites (Gd ones, Cs symmetry) and B-sites (Sc ones, Ci symmetry). The entering of RE3+ ions such as Tm3+ and Ho3+ into Sc sites looks very improbable due to the large difference in their ionic radii (RTm = 0.88 Å, RHo = 0.901 Å and RSc = 0.745 Å for VI-fold oxygen coordination), as pointed out by Piotrowski et al. [31]. Moreover, electric dipole (ED) transitions are forbidden for sites with a center of inversion, so that even if a small amount of RE3+ dopant ions are found in the Sc sites, they will not show themselves in the optical spectra (except of the transitions with a magnetic dipole component of the transition intensity). This situation is similar to the case of multi-site cubic sesquioxides (e.g., Y2O3) featuring a distribution of the dopant ions over C2 and C3i symmetry sites [45]. However, since both sites in the latter are occupied by the same host cations, 25% of the dopant ions remain optically inactive.

In the present work, no signs of multi-site behavior for eigher Tm3+ or Ho3+ ions were revealed by the LT spectroscopy and the measured LT emission spectra at 2 µm were almost independent of the excitation wavelength. To get further insight into this topic, we compared the LT (12 K) absorption spectra of Ho3+ ions (the 5I85I7 transition) for three crystals, namely ordered YAlO3 (Y site symmetry: Cs), GdScO3 (Gd site symmetry: Cs) and Y2O3 (Y site symmetry: C2 and C3i), as shown in Fig. 10. At 12 K, the temperature broadening of spectral lines is greatly suppressed. As seen from the figure, the spectral linewidths for all three crystals are comparable and they are much narrower than those for crystals with a structure and / or compositional disorder featuring a strong inhomogeneous spectral line broadening, such as Ca(Gd,Lu)AlO4 [38] crystal or solid-solution (“mixed”) sesquioxides (Lu,Sc)2O3 [45].

 figure: Fig. 10.

Fig. 10. LT (12 K) absorption spectra of Ho3+-doped GdScO3, Y2O3 and YAlO3 crystals, the 5I85I7 transition.

Download Full Size | PDF

From the above-mentioned considerations, we conclude that the main reasons for observation of broad and structureless spectra of RE ions in GdScO3 are (i) low-symmetry component in the crystal-field potential expansion leading to a large Stark splitting of RE3+ multiplets and (ii) a strong electron-phonon interaction leading to a large homogeneous thermal line widths at room temperature. The latter is in line with the fact that the spectral broadening is more significant for Tm3+ ions than for Ho3+ ions, as evidenced by the luminescence spectra at 2 µm, see Fig. 6(b). Indeed, it is known that the strength of the lattice-orbit interaction is not homogeneous across the lanthanide series and increases at its beginning and end being particularly strong for Tm3+ ions.

3.5 Laser operation

The layout of the laser setup is depicted in Fig. 11. The rectangular laser element was cut from the 3.76 at.% Tm, 0.35 at.% Ho:GdScO3 crystal for light propagation along the a-axis (a-cut) with a thickness of 3.1 mm and an aperture of 3.0(b) × 3.0(c) mm2. It was polished to laser quality on both sides, with good parallelism, and left uncoated. The laser element was mounted on a passively cooled copper holder using a thermally conductive silver paint for improved heat removal. The hemispherical laser cavity was formed by a flat pump mirror (PM) with high transmission (T = 80%) at 0.79 µm and high reflection at 1.86 – 2.31 µm and a series of concave output couplers (OCs) with a radius of curvature (RoC) of -100 mm and a transmission at the laser wavelength TOC ranging from 0.1% to 10%. The crystal was placed near the PM with a small air gap of less than 1 mm. The geometric cavity length was ∼99 mm. The pump source was a CW Ti:Sapphire laser (3900S, Spectra Physics) emitting up to 3.9 W at 791 nm (to address the 3H63H4 Tm3+ absorption band) with a nearly diffraction-limited beam quality (M2 ≈ 1). The pump power was adjusted by a half-wave plate with the polarization fixed by a polarizer to be E || b in the crystal. The pump radiation was focused into the crystal through the PM using an antireflection-coated aspherical lens with a focal length f of 75 mm. The diameter of the pump spot 2ωp was 70 ± 10 µm. The pumping was done in a double-pass configuration due to the non-negligible reflectance of the OCs at 0.79 µm. The pump absorption efficiency under lasing conditions was weakly dependent on the output coupling and amounted to ∼82%. A long-pass reflective filter was placed after the OC to filter out the residual pump. The laser spectra were measured using an optical spectrum analyzer (AQ6376, Yokogawa) and a ZrF4 fiber.

 figure: Fig. 11.

Fig. 11. Layout of the Tm,Ho:GdScO3 laser: P – Glan-Taylor polarizer; FL – aspherical focusing lens; PM – pump mirror; OC – output coupler; F – long pass filter.

Download Full Size | PDF

The Tm,Ho:GdScO3 laser generated a maximum output power of 1.16 W at ∼2067 and 2099nm (demonstrating dual-band emission) with a slope efficiency η of 50.5% (relative to the absorbed pump power) at the maximum absorbed pump power of 2.60 W for an intermediate TOC of 5%, as illustrated in Fig. 12(a). The optical efficiency was 29.1% (relative to the pump power incident on the crystal). Upon increasing the output coupling from 0.1% to 10%, there was a gradual rise in the laser threshold from 52 to 290 mW. The input-output dependencies were linear within the investigated range of pump powers, and no indications of unwanted thermal effects or thermal fracture of the crystal were detected.

 figure: Fig. 12.

Fig. 12. Tm,Ho:GdScO3 laser: (a) input-output dependences for various output coupling, a-cut crystal, η – slope efficiency, Pth – threshold pump power; (b) Findlay-Clay analysis for evaluating the round trip passive losses L; (c) typical spectra of laser emission measured well above the laser threshold, the laser polarization is E || c; (d) experimental crystal-field splitting of the 5I7 and 5I8 Ho3+ manifolds, arrows indicate the observed laser transitions.

Download Full Size | PDF

The round trip passive losses L in the Tm,Ho:GdScO3 laser were evaluated by the Findlay-Clay analysis, i.e., by plotting the laser threshold power Pth vs. ln(1/ROC) where ROC = 1 – TOC is the output-coupler reflectivity, Fig. 12(b). This yields L = 2.5 ± 0.5%, well below the Fresnel loss at the uncoated crystal surfaces owing to the etalon effect. The theoretical slope efficiency of the Ho laser is calculated as η = ηSt,L × ηq × ηOC × ηmode, where ηSt,L = λP/λL is the Stokes efficiency under lasing conditions, λP and λL are the pump and laser wavelengths, respectively, ηq is the quantum efficiency which reach 2 assuming a very efficient cross-relaxation for Tm3+ ions and efficient Tm3+ → Ho3+ energy transfer, ηOC = ln(1 - TOC)/ln[(1 - TOC) × (1 - L)] is the output-coupling efficiency and ηmode is the mode-matching efficiency. For TOC = 5%, the result is η = 51.2% which is just above the measured slope efficiency.

Typical laser emission spectra measured well above the laser threshold, are depicted in Fig. 12(c). For a very low TOC of 0.1%, the laser operated across a wide spectral range of 2097 – 2117 nm. For TOC between 0.5% and 2%, the emission occurred at ∼2.10 µm. Upon increasing TOC to 5% and 10%, the laser wavelength switched to ∼2.06 and 2.026 µm, respectively. For all studied OCs, the laser operated solely on the 5I75I8 Ho3+ transition without Tm3+ colasing confirming the properly selected Ho/Tm codoping ratio. The observed progressive blue-shift in the laser spectra can be attributed to the quasi-three-level nature of the 5I75I8 Ho3+ laser transition which involves reabsorption from the ground state. For all the studied OCs, the laser emission was linearly polarized (E || c), and the polarization state was naturally determined by the anisotropy of the gain, cf. Figure 6(a). Based on the determined crystal-field splitting of the 5I7 (Yk) and 5I8 (Zl) Ho3+ multiplets, we assigned the observed laser lines to the Y1 → Z17 (2115 nm), Z16 (2100 nm), Z13 (2067nm) and Z7 (2024nm) electronic transitions, as illustrated in Fig. 12(d).

The polarization behavior of the Tm,Ho:GdScO3 laser was studied in more detail in Fig. 13. By using a rotatory Glan-Taylor prism placed after the OC and the long-pass filter, we measured the polarization state of the output. The polar plot of the normalized laser power transmitted through the polarizer P vs. its orientation θ (θ = 0 for horizontal polarization, along the b-axis) is well fitted using the formula Ptrans(θ) = Pu/2 + Ppcos2(θφ), Fig. 13(a), where Pu and Pp are the unpolarized and polarized power fractions, respectively, and φ is the preferential orientation of the electric field vector E, for Pu < 0.001 and φ = 90° suggesting linearly polarized light along the c-axis. The polarization degree P = Pp/(Pp + Pu) > 99.9%. Then, by inserting another AR-coated half-wave plate before the focusing lens, the orientation of the pump polarization EP was changed between E || b and E || c. The output power of the Tm,Ho:GdScO3 laser was monitored as a function of EP at a fixed incident pump power, see Fig. 13(b). The strong absorption anisotropy of Tm3+ ions in the GdScO3 at the pump wavelength of 791 nm, as shown in Fig. 5(d), leads to a notable variation of the pump absorption efficiency when rotating the pump polarization and, consequently, a change in the output power of the laser. This analysis confirms the preferential pumping with the EP || b polarization.

 figure: Fig. 13.

Fig. 13. Polarization behavior of the Tm,Ho:GdScO3 laser based on an a-cut crystal: (a) study of the polarization state of laser emission (E || c); (b) dependence of output power on the pump polarization EP in the b-c plane, TOC = 5%, Pinc = 1.1 W.

Download Full Size | PDF

The Tm,Ho:GdScO3 laser generated a nearly circular output beam even at high pump powers. The corresponding 1D intensity distributions were well fitted by a Gaussian profile, as illustrated in Fig. 14(a,b). The laser was operating on the fundamental transverse mode (measured beam quality factors M2x,y < 1.05), leading to good mode-matching efficiency.

 figure: Fig. 14.

Fig. 14. Spatial and temporal emission characteristics of the Tm,Ho:GdScO3 laser: (a) a typical far-field beam profile; (b) the corresponding 1D intensity distributions (symbols) and their Gaussian fits (curves); (c) a typical switch-on oscilloscope trace, TOC = 5%, Pabs = 2.0 W.

Download Full Size | PDF

A typical oscilloscope trace of the output emission from the Tm,Ho:GdScO3 laser is shown in Fig. 14(c). It reveals a highly stable CW operation after the spiking following switching on of the pump, and no relaxation oscillations are observed.

The unpolarized spectra of visible and near-IR luminescence from the Tm,Ho:GdScO3 crystal were measured under lasing and non-lasing conditions, see Fig. 15. The spectra are dominated by intense Tm3+ emission from the pump manifold according to the 3H43H6 transition, observed at 0.8 µm. The upconversion emissions are solely attributed to Ho3+ ions indicating an efficient Tm3+ → Ho3+ energy transfer draining the population of the Tm3+ metastable state (3F4). In the visible, intense upconversion luminescence is observed in the green (the 5F4 + 5S25I8 transition, at 544 nm) and red (the 5F55I8 transition, at 654 nm). Under lasing conditions, the intensity of Ho3+ upconversion luminescence is notably decreased compared to the case without lasing due to the rapid depopulation of the metastable 5I7 Ho3+ state by the lasing process (5I75I8). In contrast, the intensity of Tm3+ emissions changes only marginally.

 figure: Fig. 15.

Fig. 15. Spectra of visible and near-IR luminescence (unpolarized emission) from the Tm,Ho:GdScO3 crystal under lasing and non-lasing conditions, λP = 791 nm.

Download Full Size | PDF

4. Conclusion

To conclude, Tm3+,Ho3+-codoped gadolinium orthoscandate (GdScO3) perovskite-type crystals are attractive for laser sources emitting around 2.1 µm due to their intense and broad absorption around 0.8 µm, broadband and polarized emission properties, predominantly unidirectional Tm3+ → Ho3+ energy transfer and a relatively large Stark splitting of the Tm3+ and Ho3+ ground manifolds assigned to the low-symmetry distortion in the local symmetry of the activator center. By employing low-temperature absorption and fluorescence spectroscopy, the crystal-field splitting of the two lowest manifolds of Tm3+ and Ho3+ ions were fully resolved. Our studies suggest that both rare-earth ions reside in a single type of sites (A-sites of the perovskite structure, Cs symmetry) and suggest that the spectral line broadening at room temperature mainly originates from the homogeneous (thermal) broadening owing to the electron-phonon interaction rather than some inhomogeneous mechanism. Highly-efficient, low-threshold and power scalable to the Watt-level continuous-wave laser operation of the Tm,Ho:GdScO3 crystal was achieved. Considering their attractive spectroscopic properties, the Tm,Ho:GdScO3 crystals are promising for sub-100 fs pulse generation from mode-locked lasers.

Funding

Natural Science Foundation of Shanghai (23ZR1471800, 23ZR1472000); National Natural Science Foundation of China (61975208, 62205354, U21A20508); Agence Nationale de la Recherche (ANR-19-CE08-0028); Région Normandie (“RELANCE” Chair of Excellence project); Sino-German Scientist Cooperation and Exchanges Mobility Program (M-0040); Ministerio de Ciencia e Innovación (PID2022-141499OB-I00).

Acknowledgment

Xavier Mateos acknowledges the Serra Húnter program.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. J. Hwang, R.R. Rao, L. Giordano, et al., “Perovskites in catalysis and electrocatalysis,” Science 358(6364), 751–756 (2017). [CrossRef]  

2. L. N. Quan, B. P. Rand, R. H. Friend, et al., “Perovskites for next-generation optical sources,” Chem. Rev. 119(12), 7444–7477 (2019). [CrossRef]  

3. H. Yamamoto, S. Okamoto, and H. Kobayashi, “Luminescence of rare-earth ions in perovskite-type oxides: from basic research to applications,” J. Lumin. 100(1-4), 325–332 (2002). [CrossRef]  

4. S. Geller, “Crystallographic studies of perovskite-like compounds. IV. Rare earth scandates, vanadites, galliates, orthochromites,” Acta Crystallogr. 10(4), 243–248 (1957). [CrossRef]  

5. J. B. Clark, P. W. Richter, and L. Du Toit, “High-pressure synthesis of YScO3, HoScO3, ErScO3, and TmScO3, and a reevaluation of the lattice constants of the rare earth scandates,” J. Solid State Chem. 23(1-2), 129–134 (1978). [CrossRef]  

6. S. N. Amanyan, E. V. Antipov, V. A. Antonov, et al., “Synthesis and structure of GdScO3,” Russ. J. Inorg. Chem. 32(9), 1225–1228 (1987).

7. R. P. Liferovich and R. H. Mitchell, “A structural study of ternary lanthanide orthoscandate perovskites,” J. Solid State Chem. 177(6), 2188–2197 (2004). [CrossRef]  

8. B. Veličkov, K. Volker, B. Rainer, et al., “Crystal chemistry of GdScO3, DyScO3, SmScO3 and NdScO3,” Z. Kristallogr. 222(9), 466–473 (2007). [CrossRef]  

9. V. M. Goldschmidt, “Die gesetze der krystallochemie,” Naturwissenschaften 14(21), 477–485 (1926). [CrossRef]  

10. A. M. Glazer, “The classification of tilted octahedra in perovskites,” Acta. Crystallogr. B. 28(11), 3384–3392 (1972). [CrossRef]  

11. Y. Zhang, S. Li, X. Du, et al., “Yb:GdScO3 crystal for efficient ultrashort pulse lasers,” Opt. Lett. 46(15), 3641–3644 (2021). [CrossRef]  

12. J. Hidde, C. Guguschev, S. Ganschow, et al., “Thermal conductivity of rare-earth scandates in comparison to other oxidic substrate crystals,” J. Alloys Compd. 738, 415–421 (2018). [CrossRef]  

13. O. Chaix-Pluchery and J. Kreisel, “Raman scattering of perovskite DyScO3 and GdScO3 single crystals,” J. Phys.: Condens. Matter 21(17), 175901 (2009). [CrossRef]  

14. S. N. Amanyan, P. A. Arsenev, K. S. Bagdasarov, et al., “Synthesis and examination of GdScO3 single crystals activated by Nd3+,” J. Appl. Spectrosc. 38(3), 343–348 (1983). [CrossRef]  

15. P. A. Arsenev, K. E. Bienert, and R. K. Sviridova, “Spectral properties of neodymium ions in the lattice of GdScO3 crystals,” Phys. Status Solidi A 9(2), K103–K104 (1972). [CrossRef]  

16. Q. Li, J. Dong, Q. Wang, et al., “Crystal growth, spectroscopic characteristics, and Judd-Ofelt analysis of Tm:GdScO3,” Opt. Mater. 109, 110298 (2020). [CrossRef]  

17. D. Hu, J. Dong, J. Tian, et al., “Crystal growth, spectral properties and Judd-Ofelt analysis of Ho:GdScO3 crystal,” J. Lumin. 238, 118243 (2021). [CrossRef]  

18. W. Hou, H. Zhao, Z. Qin, et al., “Spectroscopic and continuous-wave laser properties of Er:GdScO3 crystal at 2.7 µm,” Opt. Mater. Express 10(11), 2730–2737 (2020). [CrossRef]  

19. F. Peng, W. Liu, Q. Zhang, et al., “Growth, structure, and spectroscopic characteristics of a promising yellow laser crystal Dy:GdScO3,” J. Lumin. 201, 176–181 (2018). [CrossRef]  

20. Y. Zhang, C. Huang, M. Xu, et al., “Nd:GdScO3 crystal: Polarized spectroscopic, thermal properties, and laser performance at 1.08 µm,” Opt. Laser Technol. 167, 109709 (2023). [CrossRef]  

21. S. Li, Q. Fang, Y. Zhang, et al., “2 µm ultrabroad spectra and laser operation of Tm:GdScO3 crystal,” Opt. Laser Technol. 143, 107345 (2021). [CrossRef]  

22. W. Hou, Y. Xue, Z. Qin, et al., “Efficient continuous wave and passively Q switched Er:GdScO3 laser using Fe:ZnSe at 2.8 µm,” Opt. Lett. 48(8), 2118–2121 (2023). [CrossRef]  

23. J. Liu, S. Li, C. Qian, et al., “High-beam-quality 2 µm tunable Tm:GdScO3 laser pumped by a 793 nm laser diode,” Laser Phys. Lett. 19(11), 115801 (2022). [CrossRef]  

24. N. Zhang, Q. Song, J. Zhou, et al., “44-fs pulse generation at 2.05 µm from a SESAM mode-locked Tm:GdScO3 laser,” Opt. Lett. 48(2), 510–513 (2023). [CrossRef]  

25. Y. Wang, P. Loiko, Y. Zhao, et al., “Polarized spectroscopy and SESAM mode-locking of Tm,Ho:CALGO,” Opt. Express 30(5), 7883–7893 (2022). [CrossRef]  

26. P. Loiko, J. Maria Serres, X. Mateos, et al., “Microchip laser operation of Tm,Ho:KLu(WO4)2 crystal,” Opt. Express 22(23), 27976–27984 (2014). [CrossRef]  

27. G. L. Bourdet and G. Lescroart, “Theoretical modeling and design of a Tm, Ho:YLiF4 microchip laser,” Appl. Opt. 38(15), 3275–3281 (1999). [CrossRef]  

28. L. Wang, W. Chen, Y. Zhao, et al., “Sub-50 fs pulse generation from a SESAM mode-locked Tm,Ho-codoped calcium aluminate laser,” Opt. Lett. 46(11), 2642–2645 (2021). [CrossRef]  

29. Z. Pan, Y. Wang, Y. Zhao, et al., “Sub-80 fs mode-locked Tm,Ho-codoped disordered garnet crystal oscillator operating at 2081 nm,” Opt. Lett. 43(20), 5154–5157 (2018). [CrossRef]  

30. R. Guo, F. Wang, S. Wang, et al., “Exploration of the crystal growth and crystal-field effect of Yb3+ in orthorhombic GdScO3 and LaLuO3 crystals,” Cryst. Growth Des. 23(5), 3761–3768 (2023). [CrossRef]  

31. W. M. Piotrowski, K. Maciejewska, L. Dalipi, et al., “Cr3+ ions as an efficient antenna for the sensitization and brightness enhancement of Nd3+, Er3+-based ratiometric thermometer in GdScO3 perovskite lattice,” J. Alloys Compd. 923, 166343 (2022). [CrossRef]  

32. S. K. Gupta, V. Grover, R. Shukla, et al., “Exploring pure and RE co-doped (Eu3+, Tb3+ and Dy3+) gadolinium scandate: Luminescence behaviour and dynamics of energy transfer,” J. Chem. Eng. 283, 114–126 (2016). [CrossRef]  

33. W. T. Carnall, P. R. Fields, and K. Rajnak, “Electronic energy levels in the trivalent lanthanide aquo ions. I. Pr3+, Nd3+, Pm3+, Sm3+, Dy3+, Ho3+, Er3+, and Tm3+,” J. Chem. Phys. 49(10), 4424–4442 (1968). [CrossRef]  

34. J-H. Li, G-H. Sun, Q-L. Zhang, et al., “Effect of annealing atmosphere on the structure and spectral properties of GdScO3 and Yb: GdScO3 crystals,” Acta. Phys. Sin. 71(16), 164206 (2022). [CrossRef]  

35. C. Derks, K. Kuepper, M. Raekers, et al., “Band-gap variation in RScO3 (R = Pr, Nd, Sm, Eu, Gd, Tb, and Dy): X-ray absorption and O K-edge x-ray emission spectroscopies,” Phys. Rev. B 86(15), 155124 (2012). [CrossRef]  

36. B. M. Walsh, N. P. Barnes, and B. Di Bartolo, “The temperature dependence of energy transfer between the Tm 3F4 and Ho 5I7 manifolds of Tm-sensitized Ho luminescence in YAG and YLF,” J. Lum. 90(1-2), 39–48 (2000). [CrossRef]  

37. Z. Pan, P. Loiko, Y. Wang, et al., “Disordered Tm3+,Ho3+-codoped CNGG garnet crystal: Towards efficient laser materials for ultrashort pulse genera-tion at ∼2 µm,” J. Alloys Compd. 853, 157100 (2021). [CrossRef]  

38. Z. Pan, P. Loiko, J.M. Serres, et al., “Tm,Ho:Ca(Gd,Lu)AlO4 crystals: Polarized spectroscopy and laser operation,” J. Lumin. 257, 119638 (2023). [CrossRef]  

39. S. Kurilchik, N. Gusakova, M. Demesh, et al., “Energy transfer in Tm,Ho:KYW crystal and diode-pumped microchip laser operation,” Opt. Express 24(6), 6451–6458 (2016). [CrossRef]  

40. P. Loiko, R. Soulard, G. Brasse, et al., “Tm,Ho:LiYF4 planar waveguide laser at 2.05 µm,” Opt. Lett. 43(18), 4341–4344 (2018). [CrossRef]  

41. P. A. Arsenev and K. E. Bienert, “Absorption, luminescence, and stimulated emission spectra of Tm3+ in GdAlO3 crystals,” Phys. Status Solidi A 13(2), K125–K128 (1972). [CrossRef]  

42. P. A. Arsenev and K. E. Bienert, “Spectral properties of Ho3+ in GdScO3 crystals,” Phys. Status Solidi A 13(2), K129–K132 (1972). [CrossRef]  

43. J. B. Gruber, W. F. Krupke, and J. M. Poindexter, “Crystal-field splitting of trivalent thulium and erbium J levels in yttrium oxide,” J. Chem. Phys. 41(11), 3363–3377 (1964). [CrossRef]  

44. K. Rajnak and W. F. Krupke, “Energy levels of Ho3+ in LaCl3,” J. Chem. Phys. 46(9), 3532–3542 (1967). [CrossRef]  

45. K. Eremeev, P. Loiko, S. Balabanov, et al., “Spectroscopy of thulium ions in solid-solution sesquioxide laser ceramics: inhomogeneous spectral line broadening, crystal-field engineering and C3i sites,” Opt. Mater. 148, 114791 (2024). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (15)

Fig. 1.
Fig. 1. The crystal structure of an orthorhombic GdScO3 crystal: (a) a fragment of the structure, black lines mark the unit-cell; (b) schematic of the [ScO6] (left) and [GdO8] (right) polyhedra showing the metal-to-oxygen (M – O) interatomic distances. The atomic coordinates reported by Veličkov et al. [8] are used.
Fig. 2.
Fig. 2. A simplified energy level scheme of Tm3+ and Ho3+ ions showing processes relevant for laser operation around 2.1 µm: NR – multiphonon non-radiative relaxation, CR – cross-relaxation, ETU – energy-transfer upconversion, P28 and P71 – energy-transfer processes.
Fig. 3.
Fig. 3. Tm, Ho:GdScO3 crystal: (a) a photograph of the as-grown crystal, the growth direction is along the [110] orientation; (b) X-ray powder diffraction (XRD) pattern, numbers denote the Miller’s indices, (hkl), black lines – the reference pattern of GdScO3.
Fig. 4.
Fig. 4. Polarized Raman spectra of an a-cut Tm,Ho:GdScO3 crystal, λexc = 514 nm, numbers - Raman frequencies in cm−1.
Fig. 5.
Fig. 5. Polarized absorption properties of the Tm,Ho:GdScO3 crystal: (a-c) absorption spectra; (d) absorption cross-sections, σabs, for the 3H63F4 Tm3+ transition, arrow indicates the pump wavelength.
Fig. 6.
Fig. 6. Emission properties of Tm,Ho:GdScO3: (a) polarized luminescence spectra in the near-IR, light polarizations: E || a, b, c, arrows indicate the observed laser wavelengths (for light polarization E || c); (b) a comparison of unpolarized emission spectra of Tm3+,Ho3+-codoped and singly Tm3+-doped GdScO3 crystals around 2 µm, λexc = 794 nm.
Fig. 7.
Fig. 7. Luminescence decay curves from the 3F4 Tm3+ and 5I7 Ho3+ manifolds measured under resonant Tm3+ excitation (λexc = 1635 nm), circles – experimental data, curves – their fits using Eq. (1), τ0 – thermal equilibrium decay time.
Fig. 8.
Fig. 8. Low-temperature (12 K) spectroscopy of Tm3+ and Ho3+ ions in the GdScO3 crystal: (a,c) absorption spectra: (a) the 3H63F4 Tm3+ transition; (c) the 5I85I7 Ho3+ transition; (b,d) luminescence spectra: (b) the 3F43H6 Tm3+ transition; (d) the 5I75I8 Ho3+ transition, “+” indicate the assigned electronic transitions. Vertical dashes – experimental Stark splitting for Tm3+ and Ho3+ ions in Cs sites of the GdAlO3 crystal (after [41,42]).
Fig. 9.
Fig. 9. Experimental crystal-field splitting of two lowest multiplets of Tm3+ and Ho3+ ions in the GdScO3 crystal. EZPL – zero-phonon line, Z1, Z2, Z7 and Z8 are the partition functions of the 3H6, 3F4 Tm3+ and 5I7, 5I8 Ho3+ multiplets, respectively.
Fig. 10.
Fig. 10. LT (12 K) absorption spectra of Ho3+-doped GdScO3, Y2O3 and YAlO3 crystals, the 5I85I7 transition.
Fig. 11.
Fig. 11. Layout of the Tm,Ho:GdScO3 laser: P – Glan-Taylor polarizer; FL – aspherical focusing lens; PM – pump mirror; OC – output coupler; F – long pass filter.
Fig. 12.
Fig. 12. Tm,Ho:GdScO3 laser: (a) input-output dependences for various output coupling, a-cut crystal, η – slope efficiency, Pth – threshold pump power; (b) Findlay-Clay analysis for evaluating the round trip passive losses L; (c) typical spectra of laser emission measured well above the laser threshold, the laser polarization is E || c; (d) experimental crystal-field splitting of the 5I7 and 5I8 Ho3+ manifolds, arrows indicate the observed laser transitions.
Fig. 13.
Fig. 13. Polarization behavior of the Tm,Ho:GdScO3 laser based on an a-cut crystal: (a) study of the polarization state of laser emission (E || c); (b) dependence of output power on the pump polarization EP in the b-c plane, TOC = 5%, Pinc = 1.1 W.
Fig. 14.
Fig. 14. Spatial and temporal emission characteristics of the Tm,Ho:GdScO3 laser: (a) a typical far-field beam profile; (b) the corresponding 1D intensity distributions (symbols) and their Gaussian fits (curves); (c) a typical switch-on oscilloscope trace, TOC = 5%, Pabs = 2.0 W.
Fig. 15.
Fig. 15. Spectra of visible and near-IR luminescence (unpolarized emission) from the Tm,Ho:GdScO3 crystal under lasing and non-lasing conditions, λP = 791 nm.

Tables (2)

Tables Icon

Table 1. Thulium-Holmium Energy-Transfer Parameters for Various Laser Crystals

Tables Icon

Table 2. Experimental Crystal-Field Splitting of the 3H6, 3F4,Tm3+ and 5I8, 5I7 Ho3+ Multiplets in the GdScO3 crystal

Equations (3)

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

n 2 ( t ) n 2 ( 0 ) = β α + β exp ( t τ ) + α α + β exp ( ( α + β ) t ) ,
n 7 ( t ) n 7 ( 0 ) = β α + β exp ( t τ ) α α + β exp ( ( α + β ) t ) ,
Θ = Z 2 Z 8 Z 1 Z 7 exp [ ( E Z P L T m E Z P L H o ) k T ] ,
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.