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

Spectroscopy of solid-solution transparent sesquioxide laser ceramic Tm:LuYO3

Open Access Open Access

Abstract

We report on a detailed spectroscopic study of a Tm3+-doped transparent sesquioxide ceramic based on a solid-solution (lutetia – yttria, LuYO3) composition. The ceramic was fabricated using commercial oxide powders by hot isostatic pressing at 1600°C for 3 h at 190 MPa argon pressure. The most intense Raman peak in Tm:LuYO3 at 385.4 cm-1 takes an intermediate position between those for the parent compounds and is notably broadened (linewidth: 12.8 cm-1). The transition intensities of Tm3+ ions were calculated using the Judd-Ofelt theory; the intensity parameters are Ω2 = 2.537, Ω4 = 1.156 and Ω6 = 0.939 [1020 cm2]. For the 3F43H6 transition, the stimulated-emission cross-section amounts to 0.27 × 10−20 cm2 at 2059nm and the reabsorption-free luminescence lifetime is 3.47 ms (the 3F4 radiative lifetime is 3.85 ± 0.1 ms). The Tm3+ ions in the ceramic exhibit long-wave multiphonon-assisted emission extending up to at least 2.35 µm; a phonon sideband at 2.23 µm is observed and explained by coupling between electronic transitions and the dominant Raman mode of the sesquioxides. Low temperature (12 K) spectroscopy reveals a significant inhomogeneous spectral broadening confirming formation of a substitutional solid-solution. The mixed ceramic is promising for ultrashort pulse generation at >2 µm.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Rare-earth sesquioxides R2O3 (where R is a lanthanide, Y or Sc) represent an important class of optical materials. The most interesting form of R2O3 compounds is the cubic one (also called C-type, sp. gr. Ia$\bar{3}$, bixbyite structure). Cubic rare-earth sesquioxides such as Y2O3 (yttria), Sc2O3 (scandia) or Lu2O3 (lutetia) are known as excellent host media for doping with trivalent rare-earth ions (RE3+) [1]. As host matrices, they feature good thermo-mechanical and thermo-optical properties [2,3] (i.e., high thermal conductivity with weak dependence on the doping level, weak isotropic thermal expansion, small and positive dn/dT coefficient), low phonon energies (for oxide materials), and wide transparency [4]. The RE3+ dopant ions substitute for the host-forming R3+ metal cations in two types of sites (C2 and C3i symmetry) with VI-fold oxygen coordination [5,6]. However, the spectroscopic properties of RE3+-doped R2O3 compounds are mainly determined by C2 species (about 3/4 of ions) since for the centrosymmetric C3i ones (about 1/4 of ions), the electric dipole transitions are not enabled [1]. The RE3+ ions in cubic sesquioxides experience strong crystal fields leading to large Stark splitting of their multiplets and, consequently, broad emission bands. The crystal-field strength increases in the series R3+ = Y3+ → Lu3+ → Sc3+ according to the variation of the ionic radius, making it possible to alter the spectral properties by changing the host composition [7].

The main drawback of R2O3 crystals are their extremely high melting points (e.g., 2425°C for Y2O3) complicating the crystal growth. So far different techniques have been used for growing R2O3 crystals, such as heat exchanger growth method (HEM), micro-pulling-down (µ-PD), Czochralski (Cz), etc. [8,9]. Still, single-crystals suffer from coloration, Rhenium (Re) impurities (when using Re crucibles), and a gradient of RE3+ dopant concentration (especially for Sc2O3). During the past decades, the transparent ceramic technology emerged as a competitive approach to the single-crystal growth [10]. It offers: (i) much lower synthesis temperatures (<1800°C); (ii) easier RE3+ doping in terms of higher available concentrations and more uniform ion distribution; (iii) well-preserved spectroscopic and thermal properties; (iv) possibility to fabricate mixed compositions (R1,R2)2O3 with well-controlled R1/R2 ratio; and (v) size-scalability. Laser-quality RE3+-doped yttria [11,12], lutetia [13,14] and scandia [15] ceramics have been developed. The main challenge for ceramics is reaching high transparency close to the theoretical limit (i.e., weak light scattering owing to residual pores, possible secondary phases at the grain boundaries, etc.).

Substitutional solid solutions A1-xBx (0 < x < 1) also called “mixed” materials are attracting attention for tailoring the spectroscopic properties of the dopant RE3+ ions [1619] as, under the condition of different crystal-field strengths associated with the parent compounds A and B, the absorption / emission lines of the dopants may experience strong inhomogeneous broadening. For cubic sesquioxides, isostructural solid-solutions with a general composition (Y1-x-yLuxScy)2O3 exist in the full range of x and y. The same is true also for mixing with some active RE3+ ions, as long as their stoichiometric compositions (RE2O3) possess the same cubic symmetry, e.g. Yb, Tm, Ho or Er. Although the growth of “mixed” sesquioxide crystals has been reported [19,20], it is much easier to fabricate such compounds via a transparent ceramic technology. Recently, laser ceramics in the lutetia-scandia and lutetia-yttria binary systems were developed [16,18,2125].

Thulium (Tm3+) doped sesquioxides attract attention because of their suitability for efficient lasing around 2 µm according to the 3F43H6 electronic transition [26]. Tm3+ ions experience strong crystal-fields in R2O3 compounds. The resulting substantial Stark splitting of their multiplets determines broad emission spectra naturally extending above 2 µm [7]. The latter property is of practical importance for broadly tunable and especially mode-locked lasers [27,28] as this helps to avoid the structured water vapor absorption in the atmosphere spectrally located at < 2 µm which is detrimental for achieving femtosecond pulses. Recently, femtosecond mode-locked Tm ceramic lasers based on “mixed” sesquioxides such as Tm:(Lu,Sc)2O3 and Tm:(Lu,Y)2O3 were reported [2830].

The present work is devoted to the spectroscopic properties of Tm3+ ions in a “mixed” (lutetia – yttria) sesquioxide laser ceramic, with the goal of revealing the effect of compositional disorder on the inhomogeneous broadening of absorption and emission lines, in order to better understand the potential of such materials for generation of few-optical-cycle pulses.

2. Synthesis of transparent ceramics

Commercial powders of rare-earth oxides, Lu2O3 (purity: 4N), Y2O3 (5N) and Tm2O3 (4N), from Jiahua Advanced Material Resources, China, were used as raw materials. They were weighed according to the composition of 3.0 at.% Tm:(Lu0.5Y0.5)2O3. The mixed powders were ball-milled in ethanol for 24 h with 1.0 at % monoclinic ZrO2 powder (Shandong Sinocera Functional Material, China) serving as a sintering aid. The milled slurries were dried in an oven at 70°C for 24 h and then sieved through a 100-mesh screen. After that, the powders were dry-pressed into pellets by a 12 mm-diameter mold and cold isostatically pressed at 200 MPa for 5 min. All the green bodies were calcined at 850°C for 4 h to remove the residue organics. After that, the samples were first pre-sintered at 1650°C for 4 h in vacuum under a pressure lower than 1.0 × 10−3 Pa, equipped with a tungsten mesh as the heating element. Then, the pre-sintered ceramics were treated by hot isostatic pressing (HIPing) at 1600°C for 3 h in 190 MPa argon pressure to eliminate residual pores. Finally, the HIP-treated samples were annealed at 1200°C for 24 h in a muffle furnace in air to compensate the oxygen loss during the vacuum pre-sintering and the following HIP treatment. Then, the ceramic disks were polished on both sides to laser quality level, Fig. 1. The obtained ceramic disks were transparent and slightly yellow colored due to the Tm3+ doping. The calculated Tm3+ ion density in the “mixed” ceramic was NTm = 8.34 × 1020 at/cm3.

 figure: Fig. 1.

Fig. 1. A photograph of annealed and polished 3.0 at.% Tm:LuYO3 ceramic disk.

Download Full Size | PDF

To observe the microstructure of the ceramic, its polished surface was thermally etched at 1400°C for 2 h in a muffle furnace in air. The thermally etched surfaces of the 3.0 at.% Tm:LuYO3 ceramics before and after HIPing were characterized using a Scanning Electron Microscope (SEM, TM3000, Hitachi, Japan). The pre-sintered ceramic has an average grain size of 1 µm and it contains sub-µm sized pores localized at the grain boundaries, Fig. 2(a). After HIPing, the mean grain size increases to 2.4 µm. The ceramic exhibits a close-packed microstructure with clean grain boundaries, Fig. 2(b).

 figure: Fig. 2.

Fig. 2. SEM images of the thermally etched surface of the 3.0 at.% Tm:LuYO3 ceramics: (a) a pre-sintered sample; (b) a sample after HIPing.

Download Full Size | PDF

For comparison, two parent ceramics, Tm:Lu2O3 and Tm:Y2O3, doped with 3 at.% Tm3+ (NTm = 8.61 × 1020 cm-3 and 7.85 × 1020 cm-3, respectively), were fabricated.

3. Results and discussion

3.1 Raman spectra

The room temperature (RT) Raman spectra were measured using a confocal laser microscope (Renishaw inVia) equipped with a×50 Leica objective and an Ar+ ion laser (514nm). For cubic sesquioxides (sp. gr. Ia$\bar{3}$) possessing a body-centered structure, the factor group analysis predicts the following irreducible representations for the optical and acoustical modes at the center of the Brillouin zone (k=0): Γop=4Ag+4Eg+14Fg+5A2u+5Eu+16Fu (of which 22 modes (Ag, Eg, and Fg) are Raman-active, 16 modes (Fu) are IR-active, and the rest are silent) and Γac=Fu [31,32].

Figure 3(a) shows the Raman spectra of the Tm3+-doped LuYO3, Y2O3 and Lu2O3 ceramics. They are typical for cubic (C-type) sesquioxides. The vibrational spectra exhibit two distinct frequency ranges: the relative position and intensities of the modes above 300cm-1 are rather similar for different ceramic compositions indicating that they are most probably related to oxygen motions and deformations of the [AO6] octahedrons [31]. The most intense peak is assigned to Fg+Ag vibrations [31], Fig. 3(b). For the yttria and lutetia ceramics, it is centered at 377.8 and 391.4cm-1, respectively, and its linewidth is nearly the same (6.7 and 7.4cm-1, respectively). For the mixed” ceramic, this peak takes an intermediate position (385.4cm-1) and it is notably broadened (linewidth: 12.8cm-1) confirming the formation of a substitutional solid-solution. A similar tendency is observed for other well assigned modes in the high-frequency range of ∼300 – 600cm-1. The peak corresponding to the maximum phonon energy (the Fg+Ag vibrations) is observed at 593cm-1 (Tm:Y2O3), 603cm-1 (Tm:LuYO3) and 612cm-1 (Tm:Lu2O3).

 figure: Fig. 3.

Fig. 3. Unpolarized Raman spectra of Tm3+-doped Lu2O3, Y2O3 and LuYO3 ceramics: (a) overview spectra; (b) a close look at the most intense mode (Fg + Ag), λexc = 514 nm. Numbers indicate the Raman frequencies in cm-1.

Download Full Size | PDF

3.2 Absorption spectra and Judd-Ofelt analysis

The transmission / absorption spectra were measured using a spectrophotometer (Lambda 1050, Perkin Elmer).

The Tm:LuYO3 ceramic exhibited a relatively high linear transmission of 81.5% at 2.2µm (out of the Tm3+ absorption bands), close to the theoretical limit, T0=2n/(n2+1)=82.0% (a formula accounting for multiple light reflections, n =1.918 is the estimated refractive index of LuYO3 [33]). The absorption spectrum of Tm3+ ions in the mixed” ceramic is shown in Fig. 4. Bands related to transitions from the ground-state (3H6) to the excited-states ranging from 3F4 up to 3P0-2 are observed. The UV absorption edge is observed at ∼250nm (for undoped Lu2O3, the optical bandgap Eg is 5.6eV [34] or ∼221nm).

 figure: Fig. 4.

Fig. 4. RT absorption spectra of the 3 at.% Tm:LuYO3 ceramic in the spectral range of (a) 245 – 320 nm, (b) 340 – 500 nm, (c) 640 – 840 nm, (d) 1050 – 2050 nm.

Download Full Size | PDF

The absorption cross-sections σabs for the 3H63H4 transition of Tm3+ in the mixed” LuYO3 ceramic, and in Y2O3 and Lu2O3, are shown in Fig. 5. This absorption band is suitable for pumping Tm-lasers using commercially available AlGaAs diode lasers emitting around 0.8µm. For the mixed” ceramic, the absorption spectrum is broadened as compared to both parent compounds; the maximum σabs is 0.33×1020cm2 at 796.2nm corresponding to an absorption bandwidth Δλabs of ∼21nm (combining several peaks), compared with σabs=0.37×1020cm2 at 796.7nm with Δλabs of ∼7nm for the Tm:Y2O3 ceramic.

 figure: Fig. 5.

Fig. 5. Absorption cross-sections, σabs, for the 3H63H4 transition of Tm3+ ions in the LuYO3, Y2O3 and Lu2O3 ceramics.

Download Full Size | PDF

The measured absorption spectrum was analyzed using the standard Judd-Ofelt (J-O) theory [35,36]. Eight Tm3+ transitions were considered. The set of reduced squared matrix elements U(k) was taken from [37]. The magnetic-dipole (MD) contributions to transition intensities (for transitions with ΔJ = J – J’ = 0, ±1) were calculated within the Russel-Saunders approximations using the wave functions of the free Tm3+ ion. The refractive index of the “mixed” ceramic LuYO3 was calculated using the dispersion curves of the parent compounds [34]. More details about the J - O analysis can be found elsewhere [38].

For calculating the absorption oscillator strengths, we have used the full Tm3+ ions density (NTm), although some authors suggest to account only for ions located in C2 sites [∼(3/4)NTm] [39]. However, for a “mixed” ceramic, the actual distribution of dopant ions over the C2 and C3i sites may significantly differ from that for the parent material.

Table 1 contains the experimental (fΣexp) and calculated (fΣcalc) absorption oscillator strengths. Here, the superscript “Σ” indicates a total value (ED + MD). The root mean square (r.m.s.) deviation between the fΣexp and fΣcalc values is δrms = 1.202, mainly due to the transitions to thermally coupled levels 3F2 + 3F3 and 1I6 + 3P0 + 3P1. For the lowest-lying excited-state (3F4), a relatively good agreement is observed. The corresponding J - O parameters are Ω2 = 2.537, Ω4 = 1.156 and Ω6 = 0.939 [1020 cm2]. These values agree well with those reported recently for another Tm3+-doped “mixed” sesquioxide ceramic with a composition (Lu,Sc)2O3, Ω2 = 2.429, Ω4 = 1.078 and Ω6 = 0.653 [1020 cm2] [16].

Tables Icon

Table 1. Experimental and Calculated Absorption Oscillator Strengthsa for Tm3+ Ions in LuYO3a

Using the determined J - O parameters, the probabilities (ED + MD) of spontaneous radiative transitions for particular emission channels J → J’ AΣcalc(JJ’), the total probabilities of radiative transitions from excited-states Atot = ΣJ'AΣcalc(JJ’), the luminescence branching ratios B(JJ’) = AΣcalc(JJ’)/Atot and the radiative lifetimes τrad = 1/Atot were calculated, cf. Table 2. The mean emission wavelengths 〈λem〉 were estimated using the barycenter energies of Tm3+ multiplets 〈E〉 from Table 1. For the 3F4 and 3H4 states, τrad amounts to 4.18 ms and 0.61 ms, respectively.

Tables Icon

Table 2. Calculated Emission Probabilities for Tm3+ Ions in LuYO3a

3.3 Emission (spectra and lifetimes)

The Tm3+ luminescence was excited at ∼796 nm using a CW Ti:Sapphire laser and the spectra were measured using an optical spectrum analyzer (OSA, AQ6376, Yokogawa) and a ZrF4 fiber. The spectral sensitivity of the set-up was calibrated using a 20 W quartz iodine lamp.

The normalized RT emission spectra of Tm3+ ions in the LuYO3, Y2O3 and Lu2O3 ceramics are shown in Fig. 6(a). The observed emission is related to the 3F43H6 transition. The spectra are very broad spanning from 1.6 up to 2.35 µm. For the “mixed” ceramic, the spectrum exhibits a notable inhomogeneous broadening as compared to those of the parent compounds. The spectral maximum is found at ∼1936nm. For the quasi-three-level 3F43H6 Tm3+ laser, emission is expected at the long-wave wing of the luminescence spectrum. For Tm3+-doped cubic sesquioxides, a broad peak above 2 µm is observed where maximum laser gain will occur. For the Tm:LuYO3 ceramic, it is centered at 2058 nm, in between the positions for Tm:Y2O3 (2049 nm) and Tm:Lu2O3 (2063 nm). The emission bandwidth for this peak also exceeds those for the parent compounds supporting our assumption about the formation of a solid-solution.

 figure: Fig. 6.

Fig. 6. (a,b) RT luminescence spectra of Tm:LuYO3, Tm:Y2O3 and Tm:Lu2O3 ceramics around 2 µm: (a) overview of the spectra, (b) a close look at the 2100 – 2375 nm range, semi-log scale, grey curve – emission spectrum for the 3H43H5 Tm3+ transition in a low-doped (<0.1 at.%) Tm:Lu2O3 shown for comparison. λexc = 796 nm.

Download Full Size | PDF

A careful examination of the long-wave part of the luminescence spectra of all the studied ceramics indicates that their emission extends up to at least 2.35 µm (further measurement was limited by the sensitivity of our set-up). It is associated with the 3F43H6 transition of Tm3+ according to the luminescence decay studies: the luminescence lifetime remains nearly constant when changing the detection wavelength between 1.8 – 2.3 µm. One may argue that such long-wave emission may also partially originate from the 3H43H5 electronic transition [40]. However, no characteristic spectral features of this transition are found in the spectra. To prove this, we have used a low-doped (<0.1 at.%) Tm:Lu2O3 crystal exhibiting almost no self-quenching of the 3H4 lifetime to measure the luminescence spectrum of the 3H43H5 transition, Fig. 6(b). It reveals several well-resolved emission peaks centered at 2282, 2308, 2324 and 2356 nm which are not found in the spectra of the ceramics. In contrast, the long-wave part of the emission spectrum of the ceramics is almost structureless and follows an exponential law at >2.25 µm, as seen from Fig. 6(b) plotted in a semi-log scale.

The long-wave limit for purely electronic transitions 3F43H6 is determined by the crystal-field splitting of the involved multiplets and, in particular, by the energy gap between the lowest sub-level of the 3F4 state (5643 cm-1, Y1) and the highest sub-level of the 3H6 ground-state (810 cm-1, Z13, see below) [41]. This yields a value of 2069 nm. All the emissions above this wavelength are multiphonon-assisted (or vibronic) as they are related to the coupling between electrons and host vibrations (phonons) [42].

The very broad peak centered at 2.23 µm (for Tm:LuYO3) is interpreted as a phonon sideband, see [42,43] for this term. Note that it takes an intermediate position between those for Tm:Y2O3 (2.21 µm) and Tm:Lu2O3 (2.25 µm) which agrees with the difference in the crystal-field strengths. This indicates that the position of this peak could be linked to those of electronic transitions. Indeed, for the “mixed” ceramic, the energy gap between the prominent electronic emission band centered at 2058nm and the above-described sideband (2236 nm) is ∼387 cm-1. This well matches the most intense Raman peak of this material corresponding to vibration with an energy ph = 385.4 cm-1, Fig. 3(b). Thus, one can describe the appearance of the phonon sideband in the Tm3+ emission spectrum as transitions to a virtual energy level located slightly above the highest electronic sub-levels of the 3H6 multiplet [44], i.e., having an energy E(Z13) + ph.

The stimulated-emission (SE) cross-sections, σSE, for the 3F43H6 transition of Tm3+ ions in the LuYO3 ceramic were calculated using two methods, namely, (i) the reciprocity method [45], and (ii) the Füchtbauer – Ladenburg (F-L) formula [46]. The σSE spectra obtained by both methods were in good agreement with each other, considering the effect of reabsorption on the measured emission spectrum. In the F-L formula, we have used a radiative lifetime of the 3F4 state τrad = 3.85 ± 0.1 ms to fit the two methods which reasonably agrees with that determined using the J - O theory (4.18 ms). In Fig. 7(a), the combined SE cross-section spectrum is shown. The maximum σSE is 0.75 × 10−20 cm2 at 1937 nm and at longer wavelengths where the laser operation is expected, σSE = 0.27 × 10−20 cm2 at 2059 nm.

 figure: Fig. 7.

Fig. 7. The 3H63F4 transition of Tm3+ ions in the LuYO3 ceramic: (a) absorption, σabs, and stimulated-emission (SE), σSE, cross-sections; (b) the gain cross-sections, σgain= βσSE (1 – β)σabs, for different inversion ratios β = N2(3F4)/NTm.

Download Full Size | PDF

The 3F43H6 transition of Tm3+ ions represents a quasi-three-level laser scheme with reabsorption at the laser wavelength. Thus, the gain cross-sections, σgain= βσSE (1 – β)σabs, are calculated, where β = N2(3F4)/NTm is the population inversion ratio. The gain profiles of the “mixed” ceramic are shown in Fig. 7(b). The spectra are smooth and broad extending until 2.35 µm. For small inversion ratios (β < 0.10), two local maxima appear in the spectra, centered at ∼2085 and 2059nm. For β = 0.04, the gain bandwidth (FWHM) is as broad as 75 nm. For higher β > 0.10, the gain maxima experience a blue-shift to ∼1960 and 1938nm. The observed broadband gain properties indicate the high suitability of this ceramic for generation of sub-100 fs pulses.

The existence of gain at long wavelengths well above 2.1 µm due to the multiphonon-assisted transitions is an important prerequisite for generation of ultrashort pulses from mode-locked Tm sesquioxide lasers. Indeed, the emission spectra of such lasers delivering pulses in the sub-100 fs time domain contained spectral components extending up to 2.3 µm [30,47].

The luminescence dynamics was studied using a ns optical parametric oscillator (Horizon, Continuum), a 1/4 m monochomator (Oriel 77200), an InGaAs detector and an 8 GHz digital oscilloscope (DSA70804B, Tektronix). To avoid the effect of reabsorption (radiation trapping) on the measured lifetimes, finely powdered samples were used.

For the 3F4 Tm3+ state, the decay curves are well described by a single-exponential law, Fig. 8(a), yielding τlum = 3.470 ms for the Tm:LuYO3 ceramic. This value is slightly longer compared to the parent compounds, 3.224 ms (Tm:Lu2O3) and 2.919 ms (Tm:Y2O3). Note that the measured luminescence lifetimes of the 3F4 state for the studied sesquioxide ceramics are close to those obtained for single-crystals with low Tm3+ doping levels (<0.3 at.%), i.e., 3.38 ms (Tm:Lu2O3) and 3.54 ms (Tm:Y2O3) [7], indicating a relatively weak concentration quenching.

 figure: Fig. 8.

Fig. 8. RT luminescence decay curves for Tm3+ ions in the LuYO3, Y2O3, and Lu2O3 ceramics: (a) decay from the 3F4 state, λexc = 1644 nm, λlum = 2000nm; (b) decay from the 3H4 state, λexc = 785 nm, λlum = 821 nm. Powdered samples.

Download Full Size | PDF

For the 3H4 pump level, the decay is clearly not single-exponential, Fig. 8(b), owing to the efficient cross-relaxation (CR) process, 3H4 + 3H63F4 + 3F4. For the Tm:LuYO3 ceramics, the mean luminescence lifetime of the 3H4 state <τlum> is only 24 µs. It is much shorter than the so-called intrinsic lifetime (measured at a very low Tm3+ doping level, i.e., unaffected by the CR process), τlum,0 = 350 µs for Tm:Lu2O3 [7]. Thus, the estimated CR rate, WCR = (1/τlum) – (1/τlum,0), is about 3.88 × 104 s-1 (for 3 at.% Tm3+ doping). The CR rate is quadratically proportional to the doping concentration, WCR = CCR(NTm)2 [48], where CCR = 0.56 × 10−37 cm6s-1 is the concentration-independent CR parameter.

Table 3 summarizes the measured luminescence lifetimes.

Tables Icon

Table 3. Measured Luminescence Lifetimes of the 3F4 and 3H4 Tm3+ States in the LuYO3, Y2O3 and Lu2O3 Ceramics

3.4 Low-temperature spectroscopy

For low temperature (LT, 12 K) studies, the samples were mounted in an APD DE-202 closed-cycle cryo-cooler equipped with an APD HC 2 Helium vacuum cryo-compressor and a Laceshore 330 temperature controller.

The 3H63F4 (absorption) and 3F43H6 (luminescence) Tm3+ transitions were considered giving access to the crystal-field splitting of both multiplets. In Fig. 9, the obtained LT absorption spectra are plotted vs. the photon energy and the LT emission spectra – vs. (EZPL – photon energy). The considered transitions are of pure ED nature and thus no signatures of C3i sites could be observed. For the RE3+ ions in C2 sites, each multiplet 2S+1LJ with an integer J is split into a total of 2J + 1 sub-levels. A careful examination of the LT spectra of the parent compounds allowed us to determine the full set of Stark sub-levels for both mutiplets, cf. Table 4. Here, we use the empirical notations for the electronic levels proposed by Lupei et al. [49]: 3H6 = Zi (i = 1…13) and 3F4 = Yj (j = 1…9). The total Stark splitting of the ground-state, ΔE(3H6), increased in the R = Y → Lu series, from 788 to 828 cm-1. The peaks corresponding to electronic transitions were relatively narrow and experienced a notable shift between Y2O3 and Lu2O3 according to the different crystal-field strengths in these compounds. For the most intense electronic line in absorption, designated as Z0 → Y7, the peak positions / widths are 6112.8 / 6.1 cm-1 (Y2O3) and 6134.9 / 5.5 cm-1 (Lu2O3), respectively.

 figure: Fig. 9.

Fig. 9. Low-temperature (LT, 12 K) spectroscopy of Tm3+ ions in the LuYO3, Y2O3, Lu2O3 ceramics: (a) absorption spectra, the 3H63F4 transition; (b) luminescence spectra, the 3F43H6 transition, ZPL – zero-phonon-lines, “+” indicate the assigned electronic transitions.

Download Full Size | PDF

Tables Icon

Table 4. Experimental Crystal-Field Splitting of the 3H6 and 3F4 Tm3+ Multiplets in LuYO3, Y2O3, and Lu2O3 Ceramics

For the “mixed” ceramic, the LT spectroscopy does not reveal any signatures of two distinct Tm3+ species (e.g., the presence of two sets of electronic transitions or any dependence of the luminescence spectrum on the excitation wavelength). Instead, the electronic lines in both the LT absorption and emission spectra are notably broadened and their peak positions take an intermediate place between those for the parent compounds. E.g., the above-mentioned Z0 → Y7 Tm3+ absorption line shifts to 6123.9 cm-1 and its width increases to 34.8 cm-1 (∼5 times that of Tm:Lu2O3). This confirms the formation of a substitutional solid-solution (Lu1-xYx)2O3 with a mixture of the host-forming cations at the atomic level. The compositional disorder originates from the second coordination sphere of Tm3+ ions formed by different sets of Lu3+ and Y3+ cations with different ionic radii. Thus, strictly speaking, one cannot speak about a predominant substitution of the host-forming Y3+ or Lu3+ cations by the dopant Tm3+ ones. Table 3 presents an attempt to assign the Stark sub-levels of the Tm3+ ion in the LuYO3 ceramic.

A similar analysis for Yb3+ ions in a mixed (Lu,Sc)2O3 sesquioxide crystal was performed recently in [50,51] however not going down to temperatures of about 10 K. The authors have shown a tendency for increasing the crystal-field strengths in R2O3 sesquioxides with decreasing the R3+ ionic radius (which agrees with our analysis, as R(Y3+) = 0.90 Å and R(Lu3+) = 0.861 Å for a VI-fold oxygen coordination). They also observed a monotonous shift of the energies of Yb3+ Stark sub-levels with the Lu/Sc ratio. One can expect a monotonouis variation of the barycenter multiplet energies of Tm3+ ions in (Lu1-xYx)2O3 solid-solutions with changing the Lu/Y ratio.

4. Conclusion

In the present work, we tried to reveal the effect of a “mixed” host composition on the spectroscopic properties of the dopant Tm3+ ions using the cubic sesquioxide system (R2O3) and analyzing a lutetia – yttria (LuYO3) transparent ceramic fabricated by HIPing. Our study evidences the formation of a substitutional sesquioxide solid-solution with a mixture of cations at the atomic level according to the following findings: (i) the dominant peak in the Raman spectrum characteristic to C-type bixbyite structure takes an intermediate position between those for the parent compounds and is notably broadened; (ii) the absorption and emission spectra of the Tm3+ ion exhibit significant inhomogeneous broadening; (iii) at 12 K, the absorption / emission peaks corresponding to electronic transitions are notably broadened, their positions follow the variation of the crystal-field strength in the R = Y → Lu series while no evidence of two distinct Tm3+ species in C2 sites with a second coordination sphere predominantly formed by Lu3+ or Y3+ is observed.

The Tm:LuYO3 ceramic benefits from inhomogeneously broadened emission related to the 3F43H6 transition naturally extending above 2 µm (the limit set by the total Stark splitting of the Tm3+ ground-state, Δ(3H6) = 810 cm-1). In addition, it exhibits a long-wave emission observed up to at least 2.35 µm. The analysis of its spectral shape reveals a phonon sideband at 2.23 µm and a part with a nearly exponential dependence at longer wavelengths. The phonon sideband is associated to the coupling of electronic transitions with the most intense Raman mode of the C-type bixbyite structure (centered at 385.4 cm-1 for LuYO3) and the exponential part – with multiphonon-assisted (vibronic) processes. All this leads to smooth (structureless) and broad gain spectra of Tm3+ ions supporting the generation of ultrashort (sub-100 fs) pulses. The utilization of multiphonon-assisted emission sidebands of Tm3+ ions in sesquioxides may be a viable way for further pulse shortening in mode-locked lasers emitting above 2 µm.

Funding

Agence Nationale de la Recherche ANR SPLENDID2 (ANR-19-CE08-0028); “RELANCE” Chair of Excellence project funded by the Normandy Region; The research at IChHPS RAS was funded by the Russian Science Foundation (21-13-00397); National Natural Science Foundation of China (61975208, 61875199, 61905247, 52032009, 61850410533, 62075090, U21A20508); Sino-German Scientist Cooperation and Exchanges Mobility Program (M-0040).

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. C. Kränkel, “Rare-earth-doped sesquioxides for diode-pumped high-power lasers in the 1-, 2-, and 3-µm spectral range,” IEEE J. Sel. Top. Quantum Electron. 21(1), 250–262 (2015). [CrossRef]  

2. Z. Liu, A. Ikesue, and J. Li, “Research progress and prospects of rare-earth doped sesquioxide laser ceramics,” J. Eur. Ceram. Soc. 41(7), 3895–3910 (2021). [CrossRef]  

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). [CrossRef]  

4. 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). [CrossRef]  

5. R. P. Leavitt, J. B. Gruber, N. C. Chang, and C. A. Morrison, “Optical spectra, energy levels, and crystal-field analysis of tripositive rare-earth ions in Y2O3. II. Non-Kramers ions in C2 sites,” J. Chem. Phys. 76(10), 4775–4788 (1982). [CrossRef]  

6. J. B. Gruber, R. P. Leavitt, C. A. Morrison, and N. C. Chang, “Optical spectra, energy levels, and crystal-field analysis of tripositive rare-earth ions in Y2O3. IV. C3i sites,” J. Chem. Phys. 82(12), 5373–5378 (1985). [CrossRef]  

7. P. Loiko, P. Koopmann, X. Mateos, J. M. Serres, V. Jambunathan, A. Lucianetti, T. Mocek, M. Aguiló, F. Díaz, U. Griebner, V. Petrov, and C. Kränkel, “Highly-efficient, compact Tm3+:RE2O3 (RE = Y, Lu, Sc) sesquioxide lasers based on thermal guiding,” IEEE J. Sel. Top. Quantum Electron. 24(5), 1–13 (2018). [CrossRef]  

8. L. Fornasiero, E. Mix, V. Peters, K. Petermann, and G. Huber, “Czochralski growth and laser parameters of RE3+-doped Y2O3 and Sc2O3,” Ceram. Int. 26(6), 589–592 (2000). [CrossRef]  

9. C. Goutaudier, F. S. Ermeneux, M. T. Cohen-Adad, and R. Moncorge, “Growth of pure and RE3+-doped Y2O3 single crystals by LHPG technique,” J. Cryst. Growth 210(4), 693–698 (2000). [CrossRef]  

10. A. Ikesue and Y. L. Aung, “Ceramic laser materials,” Nat. Photonics 2(12), 721–727 (2008). [CrossRef]  

11. J. Lu, K. Takaichi, T. Uematsu, A. Shirakawa, M. Musha, K. I. Ueda, H. Yagi, T. Yanagitani, and A. A. Kaminskii, “Yb3+:Y2O3 ceramics–a novel solid-state laser material,” Jpn. J. Appl. Phys. 41(Part 2, No. 12A), L1373–L1375 (2002). [CrossRef]  

12. P. A. Ryabochkina, A. N. Chabushkin, Y. L. Kopylov, V. V. Balashov, and K. V. Lopukhin, “Two-micron lasing in diode-pumped ceramics,” Quantum Electron. 46(7), 597–600 (2016). [CrossRef]  

13. K. Takaichi, H. Yagi, A. Shirakawa, K. Ueda, S. Hosokawa, T. Yanagitani, and A. A. Kaminskii, “Lu2O3:Yb3+ ceramics–a novel gain material for high-power solid-state lasers,” Phys. Status Solidi A 202(1), R1–R3 (2005). [CrossRef]  

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

15. J. Lu, J. F. Bisson, K. Takaichi, T. Uematsu, A. Shirakawa, M. Musha, K. Ueda, H. Yagi, T. Yanagitani, and A. A. Kaminskii, “Yb3+:Sc2O3 ceramic laser,” Appl. Phys. Lett. 83(6), 1101–1103 (2003). [CrossRef]  

16. W. Jing, P. Loiko, J. Maria Serres, Y. Wang, E. Vilejshikova, M. Aguiló, F. Díaz, U. Griebner, H. Huang, V. Petrov, and X. Mateos, “Synthesis, spectroscopy, and efficient laser operation of “mixed” sesquioxide Tm:(Lu,Sc)2O3 transparent ceramics,” Opt. Mater. Express 7(11), 4192–4202 (2017). [CrossRef]  

17. J. Saikawa, Y. Sato, T. Taira, and A. Ikesue, “Passive mode locking of a mixed garnet Yb:Y3ScAl4O12 ceramic laser,” Appl. Phys. Lett. 85(24), 5845–5847 (2004). [CrossRef]  

18. A. Pirri, B. Patrizi, R. N. Maksimov, V. A. Shitov, V. V. Osipov, M. Vannini, and G. Toci, “Spectroscopic investigation and laser behaviour of Yb-doped laser ceramics based on mixed crystalline structure (ScxY1-x)2O3,” Ceram. Int. 47(20), 29483–29489 (2021). [CrossRef]  

19. A. Schmidt, V. Petrov, U. Griebner, R. Peters, K. Petermann, G. Huber, C. Fiebig, K. Paschke, and G. Erbert, “Diode-pumped mode-locked Yb:LuScO3 single crystal laser with 74 fs pulse duration,” Opt. Lett. 35(4), 511–513 (2010). [CrossRef]  

20. C. Kränkel, A. Uvarova, É Haurat, L. Hülshoff, M. Brützam, C. Guguschev, S. Kalusniak, and D. Klimm, “Czochralski growth of mixed cubic sesquioxide crystals in the ternary system Lu2O3–Sc2O3–Y2O3,” Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 77(4), 550–558 (2021). [CrossRef]  

21. W. Jing, P. Loiko, J. M. Serres, Y. Wang, E. Kifle, E. Vilejshikova, M. Aguiló, F. Díaz, U. Griebner, H. Huang, and V. Petrov, “Synthesis, spectroscopic characterization and laser operation of Ho3+ in “mixed” (Lu,Sc)2O3 ceramics,” J. Lumin. 203, 145–151 (2018). [CrossRef]  

22. L. Basyrova, P. Loiko, W. Jing, Y. Wang, H. Huang, E. Dunina, A. Kornienko, L. Fomicheva, B. Viana, U. Griebner, V. Petrov, M. Aguiló, F. Díaz, X. Mateos, and P. Camy, “Spectroscopy and efficient laser operation around 2.8 µm of Er:(Lu,Sc)2O3 sesquioxide ceramics,” J. Lumin. 240, 118373 (2021). [CrossRef]  

23. 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). [CrossRef]  

24. Z. Hao, L. Zhang, Y. Wang, H. Wu, G. H. Pan, H. Wu, X. Zhang, D. Zhao, and J. Zhang, “11 W continuous-wave laser operation at 2.09 µm in Tm:Lu1.6Sc0.4O3 mixed sesquioxide ceramics pumped by a 796 nm laser diode,” Opt. Mater. Express 8(11), 3615–3621 (2018). [CrossRef]  

25. W. Jing, P. Loiko, L. Basyrova, Y. Wang, H. Huang, P. Camy, U. Griebner, V. Petrov, J. M. Serres, R. M. Solé, M. Aguiló, F. Díazd, and X. Mateos, “Spectroscopy and laser operation of highly-doped 10 at.% Yb:(Lu,Sc)2O3 ceramics,” Opt. Mater. 117, 111128 (2021). [CrossRef]  

26. 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). [CrossRef]  

27. A. Suzuki, C. Kränkel, and M. Tokurakawa, “High quality-factor Kerr-lens mode-locked Tm:Sc2O3 single crystal laser with anomalous spectral broadening,” Appl. Phys. Express 13(5), 052007 (2020). [CrossRef]  

28. Y. Wang, W. Jing, P. Loiko, Y. Zhao, H. Huang, X. Mateos, S. Suomalainen, A. Härkönen, M. Guina, U. Griebner, and V. Petrov, “Sub-10 optical-cycle passively mode-locked Tm:(Lu2/3Sc1/3)2O3 ceramic laser at 2 µm,” Opt. Express 26(8), 10299–10304 (2018). [CrossRef]  

29. Y. Zhao, L. Wang, W. Chen, Z. Pan, Y. Wang, P. Liu, X. Xu, Y. Liu, D. Shen, J. Zhang, M. Guina, X. Mateos, P. Loiko, Z. Wang, X. Xu, J. Xu, M. Mero, U. Griebner, and V. Petrov, “SESAM mode-locked Tm:LuYO3 ceramic laser generating 54-fs pulses at 2048 nm,” Appl. Opt. 59(33), 10493–10497 (2020). [CrossRef]  

30. Y. Zhao, L. Wang, Y. Wang, J. Zhang, P. Liu, X. Xu, Y. Liu, D. Shen, J. E. Bae, T. G. Park, F. Rotermund, X. Mateos, P. Loiko, Z. Wang, X. Xu, J. Xu, M. Mero, U. Griebner, V. Petrov, and W. Chen, “SWCNT-SA mode-locked Tm:LuYO3 ceramic laser delivering 8-optical-cycle pulses at 2.05 µm,” Opt. Lett. 45(2), 459–462 (2020). [CrossRef]  

31. Y. Repelin, C. Proust, E. Husson, and J. M. Beny, “Vibrational spectroscopy of the C-form of yttrium sesquioxide,” J. Solid State Chem. 118(1), 163–169 (1995). [CrossRef]  

32. 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). [CrossRef]  

33. D. E. Zelmon, J. M. Northridge, N. D. Haynes, D. Perlov, and K. Petermann, “Temperature-dependent Sellmeier equations for rare-earth sesquioxides,” Appl. Opt. 52(16), 3824–3828 (2013). [CrossRef]  

34. D. Zhang, W. Lin, Z. Lin, L. Jia, W. Zheng, and F. Huang, “Lu2O3: A promising ultrawide bandgap semiconductor for deep UV photodetector,” Appl. Phys. Lett. 118(21), 211906–4 (2021). [CrossRef]  

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

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

37. A. A. Kornienko and E. B. Dunina, “Determination of intensity parameters from the fine details of the Stark structure in the energy spectrum of Tm3+ ions in Y3Al5O12,” Opt. Spectrosc. 97(1), 68–75 (2004). [CrossRef]  

38. L. Zhang, H. Lin, G. Zhang, X. Mateos, J. Maria 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). [CrossRef]  

39. R. Moncorgé, Y. Guyot, C. Kränkel, K. Lebbou, and A. Yoshikawa, “Mid-infrared emission properties of the Tm3+-doped sesquioxide crystals Y2O3, Lu2O3, Sc2O3 and mixed compounds (Y,Lu,Sc)2O3 around 1.5-, 2-and 2.3-µm,” J. Lumin. 241, 118537 (2022). [CrossRef]  

40. L. Guillemot, P. Loiko, R. Soulard, A. Braud, J. L. Doualan, A. Hideur, and P. Camy, “Close look on cubic Tm:KY3F10 crystal for highly efficient lasing on the 3H43H5 transition,” Opt. Express 28(3), 3451–3463 (2020). [CrossRef]  

41. P. Loiko, E. Kifle, L. Guillemot, J. L. Doualan, F. Starecki, A. Braud, M. Aguiló, F. Díaz, V. Petrov, X. Mateos, and P. Camy, “Highly efficient 2.3 µm thulium lasers based on a high-phonon-energy crystal: evidence of vibronic-assisted emissions,” J. Opt. Soc. Am. B 38(2), 482–495 (2021). [CrossRef]  

42. S. Tanabe, K. Hirao, and N. Soga, “Local structure of rare-earth ions in fluorophosphate glasses by phonon sideband and Mössbauer spectroscopy,” J. Non-Cryst. Solids 142, 148–154 (1992). [CrossRef]  

43. F. Auzel, “Multiphonon-assisted anti-Stokes and Stokes fluorescence of triply ionized rare-earth ions,” Phys. Rev. B 13(7), 2809–2817 (1976). [CrossRef]  

44. P. Loiko, X. Mateos, S. Y. Choi, F. Rotermund, J. M. Serres, M. Aguiló, F. Díaz, K. Yumashev, U. Griebner, and V. Petrov, “Vibronic thulium laser at 2131 nm Q-switched by single-walled carbon nanotubes,” J. Opt. Soc. Am. B 33(11), D19–D27 (2016). [CrossRef]  

45. S. A. Payne, L. L. Chase, L. K. Smith, W. L. Kway, and W. F. Krupke, “Infrared cross-section measurements for crystals doped with Er3+, Tm3+ and Ho3+,” IEEE J. Quantum Electron. 28(11), 2619–2630 (1992). [CrossRef]  

46. 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). [CrossRef]  

47. Y. Zhao, L. Wang, W. Chen, P. Loiko, Y. Wang, Z. Pan, H. Yang, W. Jing, H. Huang, J. Liu, X. Mateos, Z. Wang, X. Xu, U. Griebner, and V. Petrov, “Kerr-lens mode-locked Tm-doped sesquioxide ceramic laser,” Opt. Lett. 46(14), 3428–3431 (2021). [CrossRef]  

48. 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). [CrossRef]  

49. A. Lupei, V. Lupei, S. Grecu, C. Tiseanu, and G. Boulon, “Crystal-field levels of Tm3+ in gadolinium gallium garnet,” J. Appl. Phys. 75(9), 4652–4657 (1994). [CrossRef]  

50. W. Liu, D. Lu, S. Pan, M. Xu, Y. Hang, H. Yu, H. Zhang, and J. Wang, “Ligand engineering for broadening infrared luminescence of Kramers ytterbium ions in disordered sesquioxides,” Cryst. Growth Des. 19(7), 3704–3713 (2019). [CrossRef]  

51. R. Guo, D. Huang, D. Lu, F. Liang, Q. Zhang, H. Yu, and H. Zhang, “Spectral broadening mechanism of Yb3+-doped cubic LuxSc2-xO3 sesquioxide crystals for ultrafast lasers,” Opt. Mater. Express 12(5), 1963–1976 (2022). [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 (9)

Fig. 1.
Fig. 1. A photograph of annealed and polished 3.0 at.% Tm:LuYO3 ceramic disk.
Fig. 2.
Fig. 2. SEM images of the thermally etched surface of the 3.0 at.% Tm:LuYO3 ceramics: (a) a pre-sintered sample; (b) a sample after HIPing.
Fig. 3.
Fig. 3. Unpolarized Raman spectra of Tm3+-doped Lu2O3, Y2O3 and LuYO3 ceramics: (a) overview spectra; (b) a close look at the most intense mode (Fg + Ag), λexc = 514 nm. Numbers indicate the Raman frequencies in cm-1.
Fig. 4.
Fig. 4. RT absorption spectra of the 3 at.% Tm:LuYO3 ceramic in the spectral range of (a) 245 – 320 nm, (b) 340 – 500 nm, (c) 640 – 840 nm, (d) 1050 – 2050 nm.
Fig. 5.
Fig. 5. Absorption cross-sections, σabs, for the 3H63H4 transition of Tm3+ ions in the LuYO3, Y2O3 and Lu2O3 ceramics.
Fig. 6.
Fig. 6. (a,b) RT luminescence spectra of Tm:LuYO3, Tm:Y2O3 and Tm:Lu2O3 ceramics around 2 µm: (a) overview of the spectra, (b) a close look at the 2100 – 2375 nm range, semi-log scale, grey curve – emission spectrum for the 3H43H5 Tm3+ transition in a low-doped (<0.1 at.%) Tm:Lu2O3 shown for comparison. λexc = 796 nm.
Fig. 7.
Fig. 7. The 3H63F4 transition of Tm3+ ions in the LuYO3 ceramic: (a) absorption, σabs, and stimulated-emission (SE), σSE, cross-sections; (b) the gain cross-sections, σgain= βσSE (1 – β)σabs, for different inversion ratios β = N2(3F4)/NTm.
Fig. 8.
Fig. 8. RT luminescence decay curves for Tm3+ ions in the LuYO3, Y2O3, and Lu2O3 ceramics: (a) decay from the 3F4 state, λexc = 1644 nm, λlum = 2000nm; (b) decay from the 3H4 state, λexc = 785 nm, λlum = 821 nm. Powdered samples.
Fig. 9.
Fig. 9. Low-temperature (LT, 12 K) spectroscopy of Tm3+ ions in the LuYO3, Y2O3, Lu2O3 ceramics: (a) absorption spectra, the 3H63F4 transition; (b) luminescence spectra, the 3F43H6 transition, ZPL – zero-phonon-lines, “+” indicate the assigned electronic transitions.

Tables (4)

Tables Icon

Table 1. Experimental and Calculated Absorption Oscillator Strengthsa for Tm3+ Ions in LuYO3a

Tables Icon

Table 2. Calculated Emission Probabilities for Tm3+ Ions in LuYO3a

Tables Icon

Table 3. Measured Luminescence Lifetimes of the 3F4 and 3H4 Tm3+ States in the LuYO3, Y2O3 and Lu2O3 Ceramics

Tables Icon

Table 4. Experimental Crystal-Field Splitting of the 3H6 and 3F4 Tm3+ Multiplets in LuYO3, Y2O3, and Lu2O3 Ceramics

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.