Next Article in Journal
Elimination of Composition Segregation in 33Al–45Cu–22Fe (at.%) Powder by Two-Stage High-Energy Mechanical Alloying
Previous Article in Journal
Review on Preformed Crowns in Pediatric Dentistry—The Composition and Application
Previous Article in Special Issue
In Situ Investigation of Strain Localization in Sintered, Porous Segmented Alumina
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Achievements and Future Perspectives of the Trivalent Thulium-Ion-Doped Mixed-Sesquioxide Ceramics for Laser Applications

1
Istituto di Fisica Applicata “N. Carrara”, Consiglio Nazionale delle Ricerche, 50019 Sesto Fiorentino, FI, Italy
2
Institute of Electrophysics UrB RAS, 620016 Ekaterinburg, Russia
3
Ural Federal University Named after the First President of Russia B.N. Yeltsin, 620002 Ekaterinburg, Russia
4
Key Laboratory of Transparent Opto-Functional Inorganic Materials, Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 201899, China
5
Center of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
6
Istituto Nazionale di Ottica, Consiglio Nazionale delle Ricerche, INO-CNR, 50019 Sesto Fiorentino, FI, Italy
*
Author to whom correspondence should be addressed.
Submission received: 9 February 2022 / Revised: 5 March 2022 / Accepted: 7 March 2022 / Published: 11 March 2022
(This article belongs to the Special Issue Multiple Applications for Ceramic Materials)

Abstract

:
This paper is devoted to reviewing the latest results achieved in solid-state lasers based on thulium-doped mixed-sesquioxide ceramics, i.e., (Lu,Sc,Y)2O3. The near- and mid-infrared regions are of interest for many applications, from medicine to remote sensing, as they match molecular fingerprints and cover several atmospheric transparency windows. These matrices are characterized by a strong electron–phonon interaction—which results in a large splitting of the ground state—and by a spectral broadening of the optical transition suitable for developing tunable and short-pulse lasers. In particular, the manuscript reports on the trivalent thulium laser transitions at 1.5, 1.9, and 2.3 µm, along with the thermal and optical characteristics of the (Lu,Sc,Y)2O3 ceramics, including the fabrication techniques, spectroscopic and optical properties, and laser performances achieved in different pumping regimes, such as continuous-wave (CW), quasi-CW, and pulsed modes. A comparison of the performance obtained with these mixed-sesquioxide ceramics and with the corresponding crystals is reported.

1. Introduction

Over the last several years, a large number of scientific papers focused on trivalent thulium ions (Tm3+) as activators in laser-grade materials have been published, because Tm3+, together with Ho3+ [1], is an excellent candidate for the development of laser sources emitting in the near- and mid-infrared, from 1.5 to 2.3 µm [2]. As a matter of fact, the most common bulk matrices (i.e., crystals and transparent polycrystalline ceramics [3])—such as YAG [3,4,5,6,7,8,9,10,11], Lu2O3 [12,13,14,15,16], Sc2O3 [17,18,19,20,21], Y2O3 [22,23,24,25], YLF [26,27], CaF2 [28,29,30], LuAG [31,32,33]—were developed and studied achieving good performance in terms of laser output power, short pulse duration, and broad range of tunability (from 1.84 to 2.07 µm).
What makes trivalent thulium ions even more interesting is the possibility to excite the Tm3+ laser transitions at 2 µm [3] and 2.3 µm [34] by means of high-efficiency, commercially available semiconductor-based pump sources, exploiting the strong absorption band located at 790–800 nm [35]. Moreover, as the corresponding full width at half-maximum (FWHM) is in the range of a few tens of nanometers (i.e., 25–40 nm), tight control of the wavelength peak of laser diodes (LDs) is not mandatory. In the past, complex and more expensive pumping laser systems such as Ti:sapphire were used.
As many applications require laser emission at longer wavelengths—as will be clarified later on—some crystals [36,37,38,39,40,41,42] and mixed ceramics have attracted the interest of the scientific community because they are characterized by slightly broader and smoother gain spectra, which can be exploited in developing tunable and sub-100 fs laser systems [43,44,45].
Among outstanding hosts, many research groups have focused their attention on three cubic sesquioxide ceramics [46]—i.e., Lu2O3 [14,15,16], Sc2O3 [21], and Y2O3 [22]—because they can form solid solutions—i.e., (Lu,Sc,Y)2O3—maintaining excellent thermomechanical properties and optical quality. It is well known that sesquioxide ceramics show high thermal conductivity, a high refractive index (i.e., 1.94 at 800 nm and 1.92 at 2066 nm in 2at.% Tm:Lu2O3 [16]), a broad transparency range (0.22–8 µm), a positive dn/dt thermo optical coefficient (i.e., (2.0 ± 0.5) × 10−5 K−1 in 2at.% Tm:Lu2O3 [47]), low-energy phonons, and a high nonlinear refractive index (at 2070 nm n2 = 3.3 × 10−16 cm2/W [15] and 8.6 × 10−16 cm2/W at 1064 nm [48]). Moreover, they support high concentrations of dopants, preserving their optical quality and good values of thermal conductivity in comparison with the undoped compositions. In particular, the thermal conductivity of Lu3+-based matrices is almost insensitive to the concentration of dopants, as Tm3+ and Lu3+ have similar ionic radii (i.e., rion(Tm3+) = 0.0869 nm and rion(Lu3+) = 0.0861 nm for VI-fold oxygen coordination [49]) and mass [50]. Furthermore, sesquioxides are characterized by a strong Stark splitting of manifolds, which results in an emission wavelength longer than 2050 nm. They have been tested in CW, Q-switching (100 ns at 2066 nm in Tm:Lu2O3 [16]), and mode-locking regimes (180 fs at 2070 nm 2at.% Tm:Lu2O3 [15]). Tm3+-doped sesquioxides are still a lively field of investigation.
Thanks to many efforts made to improve the manufacturing techniques, in recent years Tm3+-doped mixed-sesquioxide ceramics such as (Lu,Y)2O3 [51,52], (Sc,Y)2O3 [53], and (Lu,Sc)2O3 [54,55,56], as well as garnets (i.e., YSAG [57,58]), have been fabricated and their potentiality tested, finding excellent results. For instance, pulses as short as 54 fs were generated in Tm:(Lu2/3Sc1/3)2O3 [52], and a CW tuning range of 200 nm was measured in Tm:LuYO3 [59].
Regarding their applications, infrared laser sources are currently used in many technological applications [60], such as material processing [61], remote sensing [62,63], gas monitoring [60], medicine [64,65], and optical communication systems. All are triggered by two important properties of the infrared spectral range: first, many roto-vibrational absorption lines of various molecules are located in this range, known as the molecular fingerprint region; secondly, it contains several atmospheric transparency windows [66]. Moreover, 2 µm range radiation, known as the “eye-safe region”, plays a crucial role in all applications where eye safety is strategic; this is because the vitreous body of the human eye absorbs the radiation in this spectral range, preventing damage to the retina.
For example, radiation below 2.3 µm matches the absorption lines of methane and carbon dioxide [67], while in medicine it allows noninvasive measurements of glucose levels [68]; its strong absorption in water is attractive for surgical applications, since its depth of penetration into biological tissues is a few hundred micrometers, enabling microsurgical approaches (i.e., at 2.09 µm and 2.02 µm the penetration depth in water is 300 µm and 180 µm, respectively). Two-micrometer high-power laser systems can serve as pump sources for mid-infrared optical parametric oscillators (OPOs) [69] which, in turn, allow generation of laser wavelengths from 3 to 12 µm [70]. An interesting example is ZnGeP2-based OPOs, which are pumped at 2.1 µm and emit in the range from 3 to 5 µm [71]. Radiation at 1.5 μm is a useful wavelength for optical communication systems because it belongs to the S-band, i.e., 1.47–1.52 μm; it can be used as a laser pump for erbium-doped fiber amplifiers, as well as for Raman fiber amplifiers.
The main purpose of this paper is to review the important goals achieved with (Lu,Y)2O3, (Lu,Sc)2O3, and (Sc,Y)2O3 mixed-sesquioxide ceramics doped with Tm3+ ions. The full potential of these matrices is still to be explored. The knowledge of the state of the art can be useful to focus on future goals. This article is organized as follows: First, the main conditions for exciting a selected laser transition of Tm3+ are discussed. Secondly, for each mixed-sesquioxide ceramic, we report and discuss their spectroscopic and optical properties, as well as their laser behavior in CW and pulsed regimes. The results are then compared with those obtained in pure crystals or ceramics (whenever possible), because this paves the way for understanding how the induced disorder in lattices, by substituting some ions of the ordered host, influences the laser behavior of the mixed matrices.

2. Tm3+ Ion Laser Transitions: General Considerations

Thulium is the least abundant chemical element among those belonging to the rare-earth family (REEs); it is never found in pure form, and in 1911 it took Charles James around 15,000 recrystallizations of Tm(BrO3)3 to obtain a spectroscopically pure sample [72]. Thulium has many isotopes, of which only one is stable: natural Thulium is composed entirely by 169Tm.
The trivalent ion, Tm3+, with an electron configuration [Xe]4f12, is employed in solid-state laser physics because of its three active optical transitions in near- and mid-infrared—i.e., 3H4 → 3H5, 3F4 → 3H6, and 3H4 → 3F4—with emissions centered at around 2.3 µm, 1.9 µm, and 1.5 µm, respectively. As the energy diagram of Tm3+ is extremely complex [73,74], a specific oscillation wavelength can be reached only if some conditions are fulfilled, since Tm3+ metastable states are involved in energy transfer processes such as non-radiative decay, excited-state absorption (ESA), and upconversion processes that take place. In Figure 1, the energy level scheme of Tm3+-doped Lu2O3 is reported.
Secondly, the Stark splitting of the Tm3+ manifolds is tightly connected to the host crystal field. In the most commonly used garnets, such as YAG or LuAG, the value of the splitting is lower than in mixed sesquioxides; for this reason, with the garnets, the range of tunability spans from 1.84 to 2 µm, while with the sesquioxides it can exceed 2.3 µm.
The concentration of Tm3+ ions is a crucial factor because, at high levels of doping cross-relaxation (CR) processes can take place between two neighboring Tm3+ ions [4,75], triggering the 2 µm transition, as is explained below (see Figure 2).
In the following sections we will discuss the properties of the main laser transitions and the conditions needed to activate them.

2.1. Tm3+ 3H4 → 3F4 Near-Infrared Emission at 1.5 µm

The existence of the 3H4 → 3F4 laser emission at around 1.5 µm [76,77,78,79] was demonstrated for the first time in 1983 in (Yb,Tm):BaYb2F8 and (Tm,Yb):LiYbF4 crystals pumped with a 1054 nm neodymium laser [80]. Five years later, laser oscillations in pure Tm3+-doped multimode fluoride fibers at around 1.48 µm, pumping with a krypton ion laser operating at 676 nm, were demonstrated [81]. Later on, other demonstrations followed in fibers [79,82]. Concerning its usage as a bulk material, few studies have been reported in the literature; the reason can be found in the nature of the transition, as it is a self-terminating four-level laser transition. As the lifetime of the lower laser level is longer than that of the upper state, laser oscillation occurs for a limited time, because the population accumulates in the lower laser level, so that eventually the population inversion cannot be maintained. This drawback is overcome by co-doping the matrices in order to either lengthen the upper level or shorten the lower level lifetime of the laser transition. It has been demonstrated that if hosts are co-doped with Tb3+ or Ho3+ ions and pumped at 0.6 or 0.8 µm, the lifetime of the 3H4 level can be shortened. Another possibility to depopulate the 3F4 is to exploit its excited level absorption toward the 3F2 level. This scheme was implemented by Komukai et al. [82] in ZBLAN glass fibers, using a Nd:YAG laser as a pump source. The excitation of Tm3+ to the 3H4 level was possible thanks to the weak absorption of Tm3+ in the ground state, in combination with the long absorption length allowed by the fibers. This effectively resulted in a two-photon pumping scheme. A similar scheme was proposed by Antipenko et al. [80] for crystals; in this case, the excitation of Tm was provided by absorption and multistep excitation transfer from Yb3+ ions used as sensitizers and pumped on the 2F7/2 → 2F5/2 transition. Concerning the nature of the host, essentially, when a sample doped with Tm3+ is pumped at 0.8 µm (i.e., 3H6 → 3H4), all radiative decay paths are activated. Fortunately, the mechanisms underpinning the laser’s action at different laser wavelengths are triggered differently. The usage of low-energy phonons together with low-doping materials is a good compromise; however, there is no shortage of exceptions. In CaF2, which is in principle an excellent large host with high thermal conductivity compared to other fluoride crystals and glasses, special attention is required. Due to the charge compensation [83] necessary to allow the neutrality of the host, even for low concentrations of dopant, aggregates of thulium ions can be formed, enhancing the probability of CR process triggering the transition at 2 µm. Fortunately, it has been demonstrated that in matrices co-doped with Y2+ ions it is possible to reduce the thulium aggregation. Thus far, no laser emission at this wavelength has been achieved in any laser ceramics.

2.2. Tm3+ 3F4 → 3H6 Near-Infrared Emission at 2 µm

The optical transition Tm3+ 3F4 → 3H6 [4] is widely used in laser systems with emission at around 1.9 µm. As it involves one of the thermally populated Stark levels of 3H6, emission from 1.8 to 2.2 µm can be obtained. In principle, the 3F4 → 3H6 optical transition can be excited by two different pumping schemes, which are characterized by different efficiencies. The least efficient scheme is based on direct pumping of the lower manifold 3H6 by a pump source at around 1.7–1.8 µm to the upper manifold 3F4 [5]; the most efficient is triggered by a non-radiative mechanism such as the cross-relaxation process (3H6 + 3H4 → 3F4 +3F4) in a matrix activated with a sufficient doping concentration [4,75], because the dipole–dipole interaction depends on the ion spacing, and shows a square dependence on the ion excitation density. If a pump wavelength centered at around 780 nm is used, Tm3+ ions are excited from the 3H6 to the 3H4 level. Thanks to the CR process, an electron of one ion relaxes from 3H4 to 3F4 (first excited ion), while an electron of the other ions is excited to 3F4 (second excited ion). Both ions contribute to populating the upper laser level 3F4 so that two excited ions are generated for each absorbed pump photon. This is referred to as a “two-for-one cross-relaxation mechanism”. As at room temperature a small percentage of the population (i.e., 2at.% in Tm:YAG) is placed on the high-energy states of the ground manifold, the 3F43H6 laser transition can be considered to be a quasi-three-level system whose overall quantum efficiency approaches 2 (beyond Stokes limit) [84], compensating for the high quantum defect. This transition can be activated in matrices with a level of doping higher than 2at.%, and moderates phonon energy. Historically, the first laser based on this transition with emission at 2 µm (Tm:YAG) was built in 1965 [85], while the pulsed regime was achieved 10 years later in a co-doped (Cr,Tm):YAG [34].

2.3. 3H4 → 3H5 Transition at 2.3 µm

The emission of Tm3+ 3H4 → 3H5 optical transition is centered at around 2.3 µm [34,86], and can be excited at 0.8 µm; it can be managed as a quasi-four-level laser transition, as non-radiative multiphonon relaxations depopulate the 3H5 level. However, its activation is complex, because the 3H4 level is a metastable state involved in several decay mechanisms. First, the 3H4 level is the upper level of the laser transition centered at 1.5 µm (i.e., 3H4 → 3F4); secondly, quenching mechanisms such as non-radiative relaxation, CR near Tm3+ ions, and energy transfer to impurities can depopulate the level. This oscillation was demonstrated in low-doping fluoride crystals [87,88,89], where non-radiative relaxation processes scarcely take place [90]. In YLF matrices, a continuous tunability was measured from 2.20 to 2.46 µm [91]. Lately, crystal oxides such as Tm-doped YAlO3 [92] and YAG [93] have been successfully tested. The dopant concentration underlying the CR mechanism between the 3H4 and 3F4 levels can be minimized, as several studies have already demonstrated, doping the host with a level of activators lower than 2at.%.
Unfortunately, to date, no laser oscillations at this wavelength have been demonstrated in ceramics.

2.4. Tm3+ Visible Laser Emissions

Although so far, to the best of our knowledge, no visible laser oscillations have been achieved and reported in the literature in mixed ceramics, it is interesting to observe that a Tm:YLF laser emitting in the blue, resonantly pumped by a XeF laser, was already presented in 1981 by Baer et al. [94]. Spectroscopic investigations focused on fluorophosphate glass hosts have been performed since 1984 and have demonstrated that the Tm3+ 1D2 → 3H4 transition can be used to obtain a laser oscillation in the visible region, at around 450 nm [95,96]. In fact, pumping of a 1at.% Tm:YLF crystal with two laser dyes at 780.78 nm and 648.77 nm oscillations in the blue (1D2 → 3F4 transition) and green (1D2 → 3H5 transition) regions was achieved; upconversion laser operation was at 77 K [97]. In 1992, blue upconversion laser emission at 450.2 nm (1D2 → 3F4 transition) and at 483.0 nm (1G4 → 3H6 transition) was also achieved in a Tm:YLiF4 crystal. The first transition was excited by sequential two-photon absorption with a Ti:sapphire laser source at 784.5 nm and 648 nm. In this experiment, the operation temperature was also up to 70 K. The laser emission at 483 nm was obtained by pumping with a single red dye laser whose wavelength was resonant with the absorption from a metastable intermediate state. The laser emission was observed up to 160 K [98]. In a (Tm,Yb):BaY2F8 crystal, where ytterbium plays the role of sensitizer, upconversion laser operation has been demonstrated in the visible region—in particular at 456 nm, 482 nm, 512 nm (Tm3+ 1D2 → 3H4 transition), and 649 nm. These measurements were performed at room temperature by pumping the sample with a CW 960 nm Ti:sapphire laser as well as a diode laser [99].

3. Mixed Ceramic Matrices

Before focusing on mixed-sesquioxide ceramics, some general consideration of ceramics may be useful for readers.
It has largely been demonstrated that transparent polycrystalline ceramics can overcome some drawbacks related to the melt growth of cubic sesquioxide crystals, which are characterized by high melting points (2490 °C in Lu2O3, 2485 °C in Sc2O3, 2425 °C in Y2O3) and by phase transition points lower than the melting points of the lattice [100].
Generally speaking, in ceramics it is possible to achieve high levels of doping and uniformity in dopant distribution, as well as excellent thermomechanical and optical properties comparable to their crystalline counterparts. Laser-grade ceramics can be fabricated thanks to the constant improvement in their manufacturing technique, because the mixture of powder fabrication methods and sintering processes plays a fundamental role in achieving the greatest optical transparency of ceramics. Today, ceramics with optical transparency near Fresnel’s theoretical limit are manufactured [101,102,103].
Nanocrystalline powders can be fabricated via several processes, such as combustion synthesis [104,105], the hydrothermal method [106], emulsion synthesis, sol–gel processing [107], co-precipitation in aqueous media [108,109,110], or laser ablation [111,112]. As far as the sintering process is concerned, the techniques currently being applied include vacuum sintering [113], vacuum sintering followed by hot isostatic pressing [114,115], hot pressing (HP) [116], hot pressing and hot isostatic pressing (HP–HIP) [117], spark plasma sintering (SPS) [118,119], and pressureless sintering under a flowing H2 atmosphere [120]. Of course, each production technique has strengths and weaknesses, imposing severe limitations on the dimensions and shape of the sample, its uniformity, and the fabrication time. For example, ceramics made by SPS can be fabricated in a short time; however, they have a small size. Moreover, the vacuum sintering temperatures of each material reported in the literature to date show a fairly narrow interval of variation, from 1700 °C to 1850 °C.
The mixed-sesquioxide ceramics mentioned in this review represent the solid solutions of the most common cubic sesquioxides, such as Lu2O3, Sc2O3, and Y2O3; they are obtained via a partial substitution of cations in the lattice by different rare-earth ions. For instance, (ScxY1−x)2O3 [53], (LuxY1−x)2O3 [51,52], and (LuxSc1−x)2O3 [54,55,56] matrices can be obtained by replacing an Y3+ ion with Sc3+ or Lu3+ in the first two hosts, and a Sc3+ ion with Lu3+ in the last matrix.
A major challenge in the process of lasing quality mixed-sesquioxide ceramics is the attainment of a pore-free microstructure with homogeneous distribution of chemical components at the nanoscale level, so as to avoid refractive index modulation near grain boundaries. In this regard, two main fabrication methods can be distinguished: The first method consists of a sintering plus HIP approach, with the utilization of commercially available powders or nanoparticles synthesized by co-precipitation and added ZrO2 as a grain growth inhibitor [54,121,122]. The second method involves conventional vacuum sintering of weakly agglomerated nanoparticles produced by laser ablation [53,112].
The main effect of substitutions, due to the different masses and radii of the cations (i.e., rion(Y3+) = 0.9 Å, rion(Sc3+) = 0.745 Å and rion(Lu3+) = 0.861 Å [49]), is a change in the lattice parameters because the size of the unit cell increases and, in turn, modifies the crystal field strength; the latter affects the position and the linewidth of the Tm3+ levels. The large splitting of the levels results in a broad range of tunability—broader than in pure matrices—while the emission spectra assume a smoother shape, better suited for a mode-locking regime. It was clearly demonstrated in Tm3+-doped (ScxY1−x)2O3 that the decrease in the cation radius increases the crystal field strength and, thus, the Stark splitting of manifolds involved in such laser transitions as, for instance, 3F4, 3H6, and so on [123]. At the same time, the separation between the lowest Stark electronic sublevels of the two manifolds 3F4 and 3H6 remains nearly identical when moving from Tm:Y2O3 to Tm:Sc2O3.
Concerning the crystallographic properties, yttrium, lutetium, and scandium sesquioxides are isostructural, and are capable of forming continuous series of solid solutions (LuxY1−x)2O3, (ScxY1−x)2O3, and (LuxSc1−x)2O3, with a bixbyite-type cubic structure (space group Ia3–Th7). However, when the difference in the ionic radii of the substituting and substituted cations is significant, crystallization in a different, completely ordered lattice (i.e., perovskite structure) can be observed. For example, at room temperature and pressure the smallest rare-earth ion Sc3+ and several lanthanides (Ln, from La3+ to Ho3+) form Ln-scandates—orthorhombic perovskites with the space group Pbnm [124]. In addition, pure perovskite-type YScO3 phases have previously been synthesized via the flux method using a PbO/PbF2 flux [125], by a combination of the polymerized complex method and solid-state reaction [126] and by heating up a corresponding mixture of oxides to 1000 °C at 20 kbar for 1 h [127]. Conversely, only a bixbyite-type cubic structure was detected in polycrystalline samples based on mixed (Lu,Y,Sc)2O3 matrices fabricated using ceramic technology.

3.1. Thermal Conductivity of Mixed Ceramics

As already mentioned, the thermal conductivity (K) of laser gain matrices depends on two parameters: the nature of the host, and the concentration of the dopant. Garnets, for instance, have a higher undoped thermal conductivity than fluorides, but lower than that of the sesquioxides, which are preferred for the development of high-average-power laser systems [128]. In all matrices, a decrease in thermal conductivity is observed with increased dopant concentration [50]. This phenomenon can be attributed to the different weight of the substituted cation and active RE ion, because the phonons—which transport the heating in insulator materials—are scattered at the doping ions, also known as mass defects, with a consequent decrease in K [129]. Accordingly, only Lu3+-based hosts do not show this trend, because the mass of Tm3+ is similar to that of Lu3+.
In undoped mixed matrices, the thermal conductivity is expected to be lower than in pure matrices due to the intrinsic disorder; moreover, it decreases when increasing the concentration of dopants. This has been very well demonstrated in mixed crystals such as LuScO3 [130]. We have no theoretical reasons to suppose that this should be different in mixed matrices; on the contrary, this could explain the saturation measured, for instance, in CW Tm3+:LuYO3 ceramic emissions [51].
As the data on thermal conductivity of mixed ceramics and its temperature dependence are scarce or nonexistent, only the values for several RE3+-doped disordered sesquioxide crystals can be reported. For instance, Liu et al. measured 5.488 W/(m K) and 5.019 W/(m K) for 1at.% Yb:Lu0.99Y1.01O3 and 1at.% Yb:LuScO3 crystals, respectively [131]. Krankel et al. reported Czochralski growth and temperature-dependent thermal conductivity of (Er0.07Sc0.50Y0.43)2O3 crystals [132]. At room temperature the thermal conductivity reached 4.1 W/(m K). Nevertheless, these values are comparable with those of other disordered matrices, such as RE3+-doped CaGdAlO4 and Ca3Nb1.5Ga3.5O12 [133,134], exhibiting broad gain bandwidth suitable for ultrashort-pulse laser operation.

3.2. Description and Comparison of the Fabrication Techniques Used to Develop the Mixed Sesquioxides Ceramics

Recently, an alternative method of synthesizing sesquioxide transparent ceramics has been proposed—to reduce the sintering temperature by forming a solid solution [46,135]. It was found that Y2O3, Sc2O3, and Lu2O3 can be mixed in arbitrary ratios to form a complete solid solution series with the formula of (LuxYySc1−x−y)2O3 (0 ≤ x < 1, 0 ≤ y < 1, and 0 < x + y ≤ 1), and the melting points of the binary and ternary systems are generally lower than that of any constituent component [136,137].
It is essential for laser applications to fabricate highly transparent sesquioxide ceramics with clean grain boundaries and narrowly distributed grain size. The availability of powders with desired properties, such as good dispersion and high reactivity, is crucial to obtaining better quality ceramics, because large particle size and hard aggregation may necessitate high sintering temperatures and, thus, may cause exaggerated grain growth [46].
As for the sintering methods, the solid-state reaction method is one of the conventional ways to sinter powders for polycrystalline transparent ceramics. This method is often used for the fabrication of complex oxides (such as YAG [7] starting from Y2O3 and Al2O3 powders, or LuAG from Lu2O3 and Al2O3 powders [31]), where the formation of the final complex oxide occurs during the sintering phase. Another conventional method is non-reactive sintering of nanopowders with high sinterability. In this kind of process, the starting powders already have their final chemical composition, and in the sintering phase the ceramic is formed by growth and densification of the microcrystals. The starting powders are usually obtained via wet chemical reactions, including homogeneous precipitation [138], the sol–gel method [139], co-precipitation [55,121,140,141,142] methods, etc. Other investigations have addressed the use of spark plasma sintering (SPS) for the production of Sc2O3, Y2O3, and Lu2O3 transparent ceramics [118,119], but not yet for the production of mixed sesquioxides.

3.3. Tm3+: (Lu1−xScx)2O3 with 0 < x < 1

In 2017, for the first time, a Tm3+:LuScO3 ceramic was fabricated by using a solid-state reactive sintering method with an average grain size of 1.65 µm [55]. The laser behavior of the sample was tested by CW pumping the sample at 790 with an AlGaAs laser diode, obtaining an output power of 211 mW with slope efficiency of ~8.2%. The mode-locking regime was demonstrated using single-walled carbon nanotubes (SWCNTs) as a saturable absorber. The shortest pulse width was 0.59 µs, with a repetition rate of 34.72 kHz. The maximum average output power was 32 mW. Concerning the spectroscopic data, and from comparison of spectra obtained for crystals and ceramics, no relevant differences in peak location were observed. The room temperature absorption spectrum of ceramic samples from 300 to 2100 nm shows six bands associated with Tm3+ transitions from the 3H6 ground state to 1D2, 1G4, 3F2,3, 3H4, 3H5, and 3F4 excited states, respectively. The peak absorption cross-section and its full width at half-maximum (FWHM) for the Tm3+ 3H6 → 3H4 transition at 793 nm were calculated to be 3.5 × 10−21 cm2 and 32 nm, respectively. The absorption cross-section is comparable to that of Tm3+:Lu2O3 ceramic and crystal (3.8 × 10−21 cm2 at 796 nm). The emission spectrum covers the wavelength range from 1550 to 2200 nm and corresponds to the transition 3F4 → 3H6 of the Tm3+ ion. The emission band centered at 1970 nm measures an FWHM of 75 nm and stays between the peaks measured in Sc2O3 at 1940 nm and in Lu2O3 at 1980 nm [143,144]. The multipeak structure measured in pure matrices disappears in mixed ceramics and crystals due to the spectral broadening generated by the disorder induced by the substitution of Lu3+ ions with Sc3+ ions. These ions are characterized by different masses and radii, as explained in the previous section; additionally, the latter exhibits smoothing of the whole spectrum, which could be helpful in obtaining ultrashort-pulse lasers via the mode-locking technique. Concerning the fluorescence lifetime of the 3F4 level, it was measured to be 3.2 ms, which is comparable with that of Tm3+:Lu2O3 ceramic (i.e., 3.7 ms [16]) and Tm3+:Lu2O3 crystal (i.e., 3.8 ms [143,144]). The large emission band and long lifetime clearly indicate that Tm3+:LuScO3 ceramic is a very promising material in mode-locking regimes.
The same year, an excellent result was obtained when a 4.76at.% Tm3+:(Lu2/3Sc1/3)2O3 mixed-sesquioxide ceramic [121] was synthesized via hot isostatic pressing (HIP), improving upon the results reported in [55]. According to the absorption spectrum data acquired at room temperature, the absorption cross-sections were calculated for the transitions 3H6 → 3F4 and 3H6 → 3H4, finding σabs = 4.26 × 10−21 cm2 at 1622 nm and σabs = 2.80 × 10−21 cm2 at 793 nm (the FWHM was 25 nm). The broadening of the emission peaks with respect to the pure matrices (i.e., Lu2O3, Sc2O3) was confirmed at 6 K as well. Spontaneous radiative transitions and radiative lifetimes were assessed by means of the Judd–Ofelt theory. The probabilities for the maximum spontaneous radiative transitions, σSE, were σSE = 7.15 × 10−21 cm2 at 1951 nm and σSE = 2.38 × 10−21 cm2 at 2090 nm. Concerning the radiative lifetime of the lowest excited state, it was found that τrad(3F4) = 4.01 ms, which is comparable with the values of the luminescence time measured in 1at.% Tm:LuScO3 single crystal and 2at.% Tm:LuScO3 ceramic, i.e., 3.84 and 3.2 ms, respectively. The laser behavior in the CW regime was tested by a microchip-type cavity. The maximum output power was 1.01 W at 2095 and 2102 nm, with a slope efficiency of η = 24% for TOC = 3%. The laser threshold was Pabs = 0.86 W. In the mode-locking regime obtained via a near-surface design GaSb-based SESAM, it was demonstrated that Tm3+:(Lu2/3Sc1/3)2O3 mixed ceramics can generate nearly Fourier-limited pulses of 63 fs pulse duration. The sample was pumped with a high-beam-quality Ti:sapphire laser source.
Mixed-sesquioxide ceramics with different Lu/Sc balances were fabricated via the solid-state reactive sintering method (pre-sintered in vacuum and then post-sintered by HIP) using ZrO2 (99.5%) powder as a sintering aid, i.e., 1at.% and 1.5at.% Tm:Lu1.6Sc0.4O3 with a thickness of 2.6 mm [54]. The high optical quality of these ceramics, which have an average grain size of 1.54 µm, was demonstrated by transmittance measurements, i.e., 82.21% at the lasing wavelength of 2.09 µm, which is near to the Fresnel theoretical value of 82.28% (n = 1.906 at 2 µm). The samples were excited by a fiber-coupled AlGaAs diode laser at 796 nm in order to measure the emission spectra at room temperature and 77 K, and these results were compared in order to calculate the Stark splitting of the levels involved in laser action: a Stark splitting of 920 cm−1 in the ground state 3H4 and 472 cm−1 in the 3F4 of Tm3+ was found. Thanks to these large Stark splittings, the laser can emit at around 2.09 µm due to transition from the lowest Stark level 3F4 to the highest Stark level of 3H6. The stimulated emission cross -section at 2.09 µm is σem = 1.1 × 10−21 cm2. From the absorption spectra, the absorption cross-section at 796 nm responsible for the 3H6 → 3H4 transition was calculated, finding it to be σabs = 3.1 × 10−21 cm2. A plano-concave cavity was used to test the samples; for the 1at.% Tm3+ ceramic, the maximum output power was 9.8 W, with a slope efficiency of 40%; the optical-to-optical conversion efficiency was measured at 25.8%. Comparable results were found with the 1.5at.% Tm3+ laser, i.e., 11 W with 39% slope efficiency and an optical-to-optical efficiency of 28.9%. With both samples, the emission wavelength was 2090 nm. By comparing the theoretical data (38%) for the conversion from one 796 nm photon to one 2090 nm photon, it was clear that cross-relaxation between two closed Tm3+ ions took place.
Over the years, various manufacturing techniques have been developed to improve the optical qualities of ceramics. Recently, a 2at.% Tm3+:(Lu0.8Sc0.2)2O3 was obtained via the gel-casting of well-dispersed nanopowders obtained by co-precipitation method using an alcohol–water solvent, which was expected to yield more homogeneous and transparent ceramics [122]. The excellent optical quality of samples was carried out from the optical in-line transmittance of a 12 mm long ceramic rod, with results as high as 80.3% at 2090 nm, which is comparable with the theoretical transmittance of ~81.4% at 2090 nm. The average grain sizes spanned from 2 to 5 µm. The attenuation coefficient at 2090 nm was 0.006 cm−1. The absorption and emission spectra of the sample were acquired at room temperature. Concerning the absorption bands, four bands centered at 686, 796, 1207, and 1625 nm were attributed to the Tm3+ transitions from its ground state of 3H6 to the excited states of 3F2,3, 3H4, 3H5, and 4F4, respectively. Focusing on the whole 3H63H4 absorption band that peaked at 796 nm, its FWHM was around 38 nm, which is broader than that of Tm3+-doped Lu2O3 ceramics. Accordingly, in the CW laser test, the sample, having been placed in a plano-concave cavity, was pumped using a commercial laser diode with emission at 796 nm. A maximum output power of 1.88 W was measured, with a slope efficiency of 24.6% (with respect to the input pump power) and an optical-to-optical efficiency of 14.1%. The laser threshold of the absorbed pump power was 3.2 W.
Recently, a 58 fs laser pulse was generated at 2081 nm by Kerr-lens mode locking in a 2.8at.% Tm:(Lu2/3Sc1/3)2O3 transparent ceramic [56] fabricated by hot isostatic pressing (HIP) of commercial powders at 1800 °C and 195 MPa in an Ar atmosphere [121]. In CW operation mode, this delivers a maximum output power of 490 mW at around 2088 nm (TOC = 0.5%), with an efficiency conversion of 26.8%. Short pulses were obtained in a Kerr-lens mode-locking regime by exploiting the high nonlinear refractive index (n2) of the ceramic. We should note that in Tm:Lu2O3 ceramics this was measured at n2 = 3.3 × 10−16 cm2/W at 2070 nm [15] and n2 = 8.6 × 10−16 cm2/W at 1064 nm, and it is expected that in mixed-sesquioxide ceramics this value should be almost preserved. The average output power was 220 mW at a pulse repetition rate of 84.8 MHz. The peak on-axis laser intensity in the sample was ∼510 GW/cm2. It was very well demonstrated that the emitted spectra at wavelengths longer than 2.2 µm are attributable to vibronic transitions of the Tm3+ and played a fundamental role in achieving laser pulses as short as 58 fs.
For the sake of clarity and completeness, we should recall that in 2011 the first 1at.% Tm3+-doped LuScO3 crystal was grown by Koopmann et al. [143]. In CW, an output power of 705 mW with a slope of 55% at around 2.1 µm was measured in a nearly concentric resonator; the laser threshold was only 0.38 mW. Concerning the range of tunability, it was demonstrated from 1960 to 2115 nm (i.e., 155 nm). Two years later, a Ti:sapphire-pumped mode-locked Tm:LuScO3 crystal laser was demonstrated [145]. In 2018, a new high-optical-quality mixed-sesquioxide crystal—i.e., 4at.% Tm:LuScO3—was grown via the heat exchanger method in a rhenium crucible and tested in a four-mirror z-fold cavity pumping for the first time by a laser diode at 793 nm [146]. In a CW regime, the laser delivered 660 mW at 2102 nm (TOC = 2%), with a slope efficiency of up to 33%, while the laser threshold was 194 mW. The tunability of laser was measured from 1973 nm to 2141 nm (i.e., 148 nm), with an FWHM of 75 nm. The femtosecond mode-locking operation was achieved via an ion-implanted InGaAsSb quantum-well-based SESAM, which enabled the generation of near-transform-limited pulses of 170 fs at 2093 nm, with an average output power of 113 mW and a pulse repetition frequency of 115.2 MHz.

3.4. Tm3+: (Lu1−xYx)2O3 with 0 < x < 1

The year 2017 was a profitable one, as the first 3at.% Tm-doped (Lu0.5Y0.5)2O3 was also fabricated by solid-state reactive sintering (Tsintering = 1700 °C) and tested in CW and Q-switching regimes [51]. The sample was pumped with a fiber-coupled AlGaAs laser diode at 790 nm. In CW, it delivered a maximum output power of 1.55 W, with a slope efficiency of 19.9% at 2050 nm. Passive Q-switching was obtained using a Cr:ZnSe saturable absorber. At 2047 nm, a maximum average output power of 0.54 W was found, along with a pulse width of 120.3 ns and a pulse energy of 20.5 µJ, while the pulse peak power was 170.6 W. Spectroscopic measurements have shown that the location of the main absorption peaks at 775.5, 796.5, and 811.5 nm matched those of Tm-doped Lu2O3 [16] and Y2O3 ceramics [25], and the calculated absorption coefficient at 796.5 nm and 790 nm was 2.6 cm−1 and 1.6 cm−1, respectively.
Two years later, the first mode-locking regime was demonstrated with a 3at.% Tm:LuYO3 mixed ceramic synthesized via a solid-state reactive sintering method [147]. The mode-locking laser operation was achieved by inserting an SESAM with a modulation depth of 1.2% and a relaxation time of 10 ps into the cavity. Pulses with a duration of 41 ps and a repetition rate of 139.3 MHz at 2061 nm were obtained. Concerning the spectroscopic measurements, the sample was pumped with a fiber-coupled laser diode at 796 nm. The absorption cross-section at the pump wavelength was calculated to be σabs = 3.8 × 10−21 cm2, with an FWHM of 25 nm. This is comparable with results obtained in Tm:Lu2O3, and higher than in Tm:LuScO3, where σabs = 3.5 × 10−21 cm2 was found [55]. Conversely, emission cross-sections calculated at the two main peaks of the fluorescence spectra—i.e., at 1937 nm and at 2055 nm—were found to be σem = 6.0 × 10−21 cm2 and 3.0 × 10−21 cm2, respectively; these results are lower than in 2at.% Tm3+:Lu2O3 (i.e., 8.2 × 10−21 cm2 [16]) and Y2O3 (i.e., 9.9 × 10−21 cm2 [148]). The fluorescence time of the 3F4 → 3H6 laser transition was 2.6 ms. CW laser behavior was tested by closing the cavity using output couplers with different transmissions (TOC). The maximum output power of 1.20 W, with a slope efficiency of 25.1% and threshold of 0.53 W, was obtained with a TOC = 3%. The highest slope efficiency of 33.1% was achieved with a TOC = 10%, with a maximum output power of 0.88 W and threshold of 1.5 W.
Passive mode-locking of a 3at.% Tm:LuYO3 ceramic was also demonstrated by using a single-walled carbon nanotube saturable absorber [59]. When pumping at 795 nm with a Ti:sapphire laser, pulses as short as 57 fs were measured at 2045 nm, with an average power of 63 mW and a repetition rate of 72.6 MHz. This is an excellent achievement compared with those obtained in a Tm:Lu2O3 ceramic operating in a mode-locking regime with a saturable absorber, i.e., 180 fs [15]; it is similar to the pulse duration of 63 fs obtained with Lu2/3Sc1/3O3 [54]. The tunability of the sample in the CW regime was preliminarily tested by inserting a 3.2 mm thick birefringent quartz plate into the cavity. The curve of tunability ranged from 1909 to 2109 nm—i.e., 200 nm—with the main peak located at 2074.2 nm and a corresponding maximum output power of 440 mW (TOC = 1.5%). Under the same experimental conditions, in 3at.% Tm:Lu2/3Sc1/3O3 ceramic, a tuning range as broad as 130 nm was found [54].
Recently, the same group has measured the shortest pulses [52,149] ever reported in literature for any bulk solid-state lasers based on Tm3+. The gain material was a 3at.% Tm:LuYO3 ceramic mode-locked with a GaSb-based SESAM. At a repetition rate of ∼78 MHz, a pulse of 54 fs was measured at 2049 nm, with an average output power of 51 mW.
The laser performances of a ceramic fabricated by hot isostatic pressing with a sintering temperature of 1700 °C were tested in CW, pumping the ceramic with a continuous-wave narrow-line-bandwidth Ti:sapphire laser at 795.3 nm. By closing the four-mirror bow-tie cavity using an output coupler with a transmission of TOC = 3%, an output power of 603 mW with a slope efficiency of 33.2% and a laser threshold of 200 mW was measured.
Concerning Tm:LuYO3 crystals, a sample was grown via the optical floating zone method for the first time in 2020 [150]; unfortunately, no data on laser behavior have been reported in the literature. Spectroscopic characterization at room temperature was performed, finding an absorption cross-section at 796 nm of σabs = 7.3 × 10−20 cm2 with the FWHM of 48 nm; lower values were measured at 2057 nm (σabs = 0.2 × 10−20 cm2) and 1936 nm (σabs = 1.4 × 10−20 cm2). The emission cross-section at 2056 nm was σem = 1.95 × 10−20 cm2, with an FWHM of nearly 300 nm. The measured fluorescence lifetime of 3F4 was 0.86 ms.

3.5. Tm3+: (Y1−xScx)2O3 with 0 < x < 1

(Y,Sc)2O3 is very attractive because the large difference in the ionic radii of Y3+ and Sc3+ [49] is expected to enhance the disorder of the lattice and, in turn, a large Stark splitting of Tm3+ levels should be measured. However, it has been extremely difficult to grow large, single crystals, due to a phase transition from hexagonal to cubic structure occurring at a temperature near to the melting point of yttrium oxide, above 2400 °C. However, thanks to the continued efforts of the scientific community, small matrices doped with Nd3+, Er3+, and Yb3+ have been grown and investigated. The first Nd3+:YScO3 crystal with a cubic bixbyite-type structure was grown in 1976 [151]. An investigation of the excitation spectra and fluorescence dynamics of Er3+:YScO3 single-crystal fiber grown via the laser-heated pedestal growth (LHPG) method was reported in 1991 [152]. The orthorhombic perovskite single-crystal Y0.96ScO2.94 grown from mixtures of sesquioxides with the flux reagents PbF2 and PbO was spectroscopically investigated, and nuclear magnetic resonance of a sample doped with 1at.% Yb3+ was carried out [125]. X-ray diffraction analysis and micro-Raman characterization were carried out in a 0.1at.% Nd3+:YScO3 crystal fiber grown via the LHPG method [153].
In 2013, for the first time, Yb3+-doped (ScxY1−x)2O3 ceramics with x = 0.1 and 0.9, developed using high-purity nanopowders and pressureless sintering in a H2 atmosphere, were fabricated [154,155]. The first laser oscillation was demonstrated only recently in 6at.% Yb-doped (ScxY1−x)2O3 ceramics with different Y3+/Sc3+ balances (x = 0, 0.273, 0.508, and 0.742), fabricated via solid-state vacuum sintering of laser-ablated mixed-sesquioxide nanoparticles [112].
To the best of our knowledge, only one study focused on Tm3+:(Y1−xScx)2O3 with 0 < x < 1 ceramic has been reported in the literature [53,156]. This study demonstrated the first laser action of a 5at.% Tm3+-doped (Sc0.252Y0.698)2O3 ceramic made via solid-state vacuum sintering of laser-ablated mixed-sesquioxide nanoparticles. Pumped at 793 nm in quasi-CW mode, this ceramic delivers 1.24 W with a slope efficiency of 9.45% at 2076 nm. Absorption spectra were acquired at several temperatures, from 93 K to 293 K: three broad bands located at 670–700 nm, 760–820 nm, and 1600–1700 nm were observed, associated with Tm3+ transitions from 3H6. The variation of the temperature affects the absorption coefficients (α) in different ways. In the case of the main peaks (657, 685, 764, and 774 nm), the α-coefficient decreases when increasing the temperature; conversely, the peaks located at 697 nm as well as in the range of 779–840 nm show an enhancement of α. This can be attributed to the change in the thermal population distribution of the sublevels in the 3H6 manifold. When increasing the temperature, the population in the upper sublevels increases, enhancing the absorption from the transitions starting from the upper sublevels, and the population in the lowest sublevel decreases, reducing the absorption coefficient of the transition starting from this sublevel. It is interesting to observe that when compared with Tm:Y2O3 the well-defined peak structure observable in the Tm:Y2O3 spectra is almost lost in the disordered sample—a clear signature of the disorder of the lattice. The individual sublevels of the 3H6 ground manifold are inhomogeneously broadened by the variation in the Stark splitting occurring at different lattice sites due to the random partial substitution of Y3+ cations with smaller Sc3+ ions. In mixed (Sc0.252Y0.698)2O3 the main peaks are less intense and broader; moreover, they are shifted to shorter wavelengths. In particular, the absorption coefficient (α) at 685.3 nm decreases from α = 140.6 cm−1 at 93 K to α = 81.9 cm−1 at room temperature. In Tm:Y2O3, the measured values were α = 79.8 cm−1 at 93 K and α = 66.1 cm−1 at RT. Regarding the emission spectrum, (Sc0.252Y0.698)2O3 shows two peaks located at 1946 nm and 2098 nm. If compared with Tm:Y2O3 [53], it can be observed that the emission cross-section at 1946 nm decreases from σem = 10.4 × 10−21 cm2 in Tm:Y2O3 to σem = 7.6 × 10−21 cm2 in the disordered matrix (Sc0.252Y0.698)2O3. However, the peak at a longer wavelength is shifted from 2060 nm (Tm:Y2O3) to 2098 nm (Sc0.252Y0.698)2O3. The emission cross-section at 2060 nm in Tm:Y2O3 was σem = 5.1 × 10−21 cm2, while σem = 3.9 × 10−21 cm2 at 2098 nm was found in (Sc0.252Y0.698)2O3. The CW tuning range was investigated in a coupled cavity configuration with an external grating, with 600 lines/mm placed in the Littrow configuration—a continuous range of tunability from 1927.5 nm to 2108.5 nm, i.e., 181 nm was measured.

4. Discussion

The continuous improvement of ceramic fabrication techniques has made it possible to obtain samples with high optical quality as well as materials that are difficult to produce in crystalline form, such as lutetia, for laser applications. Thanks to these ceramics, various research groups have developed lasers with output powers of hundreds of watts and short pulses. Without any doubt, today, the laser performance achieved by transparent polycrystalline ceramics is almost comparable with the corresponding crystals. Ceramic matrices as YAG, LuAG, LuYAG, sesquioxides, etc., doped with Tm3+ ions, have been spectroscopically tested, and laser oscillations have been achieved in the so-called eye-safe region (above 1.4 µm).
In the last five years, alongside the interest in the most common matrices [157,158], the idea of manufacturing mixed matrices (see Table 1 and Table 2) doped with various rare-earth ions [159,160] has emerged because, due to their crystalline structure distorted by the replacement of some cations of the matrix with other rare-earth ions, they could pave the way for the development of lasers with pulses in the order of a few tens of femtoseconds. The first results published in the literature are encouraging, as 50 fs laser pulses have been demonstrated (see Table 3). However, the knowledge we have on the spectroscopic behavior of this mixed matrix is limited to some hosts that have already been fabricated, and the physical and thermomechanical parameters are not yet known. Accordingly, a great effort will have to be made in terms of improving both the quality of the mixed matrices and the knowledge of the physical processes that determine the activation of some specific laser transitions. For instance, we believe that the development of lasers with visible emission operating at room temperature could be one of the most important goals.
The spectroscopic properties and laser behavior (i.e., laser emission wavelength, tuning range, and available pulse durations) of sesquioxide solid solutions are directly dependent on the composition of the mixed matrix, and the selection of a particular balance between cations should be based on the specific field in which the active medium and corresponding laser are planned to be applied. For instance, compared to (LuxY1−x)2O3 materials, mixed compositions with Sc—such as (LuxScx)2O3 and (YxScx)2O3—possess broader absorption and emission spectra of Tm3+ ions, supporting an extended tuning range and shorter pulse durations at the expense of a somewhat lower thermal conductivity, due to the larger difference in ionic radii and mass between Lu3+ (Y3+) and Sc3+. For enhanced Tm3+-doping concentrations and high-average-power laser action, disordered ceramics and crystals based on Lu2O3 seem more preferable, owing to the small discrepancy between Lu3+ and Tm3+, inducing a minor level of disorder and better preserving the thermophysical characteristics.
Considering possible mass production of mixed-sesquioxide gain media, the cost of raw materials should be taken into account. High-purity Y2O3 micro- and nanopowders are currently available on the market at a reasonable cost, while proper lutetium and scandium sesquioxide powders are much harder to find, and they are several times more expensive than yttria. Even though the modern fabrication methods yield disordered ceramic matrices with lasing quality, considerable efforts will be required for obtaining large-sized and large-aperture active components, so as to give an impulse to further development of the advanced 2 μm solid-state lasers required in various fields of science and technology. In this regard, the approach suggested by Wu et al., consisting of gel-casting of well-dispersed nanopowders followed by HIP treatment [122] for the synthesis of 12 mm long 2at.% Tm:(Lu0.8Sc0.2)2O3 ceramic rods, with an attenuation coefficient of ~0.006 cm−1 at 2090 nm, is a major step towards large, high-optical-quality samples.
Tm3+-doped mixed-sesquioxide materials could be elaborated in the direction of additional broadening of gain profiles by the utilization of Ho3+ as a co-dopant. The point is that the gain profiles of Tm3+ and Ho3+ in co-doped hosts have overlapping regions, and it is possible to combine stimulated emissions of both active ions for ultrashort-pulse lasing. For instance, Wang et al. [44] demonstrated 46 fs pulses at 2033 nm from an SESAM mode-locked Tm,Ho:Ca(Gd,Lu)AlO4 laser, which is the shortest duration ever obtained in a Tm- and/or Ho-based gain medium. The aspect of optimal ratio between Tm and Ho ions in sesquioxides—and especially in mixed sesquioxides—needs detailed investigation.

5. Conclusions

In this paper, we report a summary of fabrication techniques, spectroscopy, and laser performance of trivalent thulium-doped mixed-sesquioxide ceramics, i.e., (Lu,Sc,Y)2O3. In particular, a careful description of the properties and mechanisms of the thulium laser transition emission at around 1.5, 1.9, and 2.3 µm is given. The results obtained with Tm3+-doped ceramic hosts were compared to performance achieved with the corresponding crystals in all pumping regimes. The emission wavelengths ranging from 1.5 µm up to 2.3 µm make the trivalent thulium ions very useful, as Tm3+ can be used in a consistent number of applications, from medicine to LIDAR (light detection and ranging) tests. Although important milestones have already been achieved by thulium-based lasers, their features are still unexplored in fields such as, for instance, laser inertial fusion energy.

Author Contributions

Conceptualization, A.P.; data curation, A.P. and R.N.M.; writing—original draft preparation, A.P. and R.N.M.; visualization, G.T.; validation, G.T., J.L., R.N.M., M.V.; writing—review and editing, A.P., R.N.M., J.L., M.V. and G.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shen, Y.J.; Yao, B.Q.; Duan, X.M.; Zhu, G.; Wang, L.W.; Ju, Y.L.; Wang, Y.Z. 103 W in-band dual-end-pumped Ho:YAG laser. Opt. Lett. 2012, 37, 3558–3560. [Google Scholar] [CrossRef] [PubMed]
  2. Walsh, B.M. Review of Tm and Ho materials; spectroscopy and lasers. Laser Phys. 2009, 19, 855–866. [Google Scholar] [CrossRef]
  3. Ikesue, A.; Aung, Y.L. Ceramic laser materials. Nat. Photonics 2018, 2, 721–727. [Google Scholar] [CrossRef]
  4. Becker, T.; Clausen, R.; Huber, G.; Duczynski, E.W.; Mitzscherlich, P. Spectroscopic and Laser Properties of Tm-Doped YAG at 2 µm. In Advanced Solid State Lasers; Shand, M., Jenssen, H., Eds.; OSA Proceedings Series; Optical Society of America: Washington, DC, USA, 1989; Volume 5, p. DD1. [Google Scholar] [CrossRef]
  5. Peterson, P.; Gavrielides, A.; Sharma, M.P. Diode-pumped Tm: YAG solid-state lasers with indirect and direct manifold pumping. Appl. Phys. B 1995, 61, 195–200. [Google Scholar] [CrossRef]
  6. Honea, E.C.; Beach, R.J.; Sutton, S.B.; Speth, J.A.; Mitchell, S.C.; Skidmore, J.A.; Emanuel, M.A.; Payne, S.A. 115-W Tm:YAG diode-pumped solid-state laser. IEEE J. Quantum Electron. 1997, 33, 1592–1600. [Google Scholar] [CrossRef]
  7. Zhang, W.; Pan, Y.-B.; Zhou, J.; Liu, W.; Li, J.; Jiang, B.; Cheng, X.; Xu, J. Diode-pumped Tm:YAG ceramic laser. J. Am. Ceram. Soc. 2009, 92, 2434–2437. [Google Scholar] [CrossRef]
  8. Ma, Q.L.; Bo, Y.; Zong, N.; Pan, Y.-B.; Peng, Q.J.; Cui, D.F.; Xu, Z.Y. Light scattering and 2 µm laser performance of Tm:YAG ceramic. Opt. Commun. 2011, 284, 1645–1647. [Google Scholar] [CrossRef]
  9. Gao, W.L.; Ma, J.; Xie, G.Q.; Zhang, J.; Luo, D.W.; Yang, H.; Tang, D.Y.; Ma, J.; Yuan, P.; Qian, L.J. Highly efficient 2 µm Tm:YAG ceramic laser. Opt. Lett. 2012, 37, 1076–1078. [Google Scholar] [CrossRef]
  10. Zhang, S.; Wang, X.; Kong, W.; Yang, Q.; Xu, J.; Jiang, B.; Pan, Y. Efficient Q-switched Tm:YAG ceramic slab laser pumped by a 792 nm fiber laser. Opt. Commun. 2013, 286, 288–290. [Google Scholar] [CrossRef]
  11. Zou, Y.; Wei, Z.; Wang, Q.; Zhan, M.; Li, D.; Zhang, Z.; Zhang, J.; Tang, D. High-efficiency diode-pumped Tm:YAG ceramic laser. Opt. Mater. 2013, 35, 804–806. [Google Scholar] [CrossRef]
  12. Koopmann, P.; Peters, R.; Petermann, K.; Huber, G. Highly Efficient, Broadly Tunable Tm:Lu2O3 Laser at 2 µm. In Proceedings of the CLEO/Europe and EQEC 2009 Conference Digest, Munich, Germany, 14–19 June 2009; Optical Society of America: Washington, DC, USA, 2009; p. CA10_3. [Google Scholar] [CrossRef]
  13. Zhang, W.; Li, L.; Zhou, S.; Gao, Q. High-Efficiency Diode-Resonantly-Pumped Tm:Lu2O3 Ceramic Laser with Near Diffraction-Limited Beam Quality. J. Russ. Laser Res. 2021, 42, 358–362. [Google Scholar] [CrossRef]
  14. Antipov, O.L.; Getmanovskiy, Y.A.; Balabanov, S.S.; Larin, S.V.; Sharkov, V.V. 1940 nm, 1966 nm and 2066 nm multi-wavelength CW and passively-Q-switched operation of L-shaped Tm3+:Lu2O3 ceramic laser in-band fiber-laser pumped at 1670 nm. Laser Phys. Lett. 2021, 18, 055001. [Google Scholar] [CrossRef]
  15. Lagatsky, A.A.; Antipov, O.L.; Sibbett, W. Broadly tunable femtosecond Tm:Lu2O3 ceramic laser operating around 2070 nm. Opt. Express 2012, 20, 19349–19354. [Google Scholar] [CrossRef] [Green Version]
  16. Antipov, O.L.; Novikov, A.A.; Zakharov, N.G.; Zinoviev, A.P. Optical properties and efficient laser oscillation at 2066 nm of novel Tm:Lu2O3 ceramics. Opt. Mater. Express 2012, 2, 183–189. [Google Scholar] [CrossRef]
  17. Koopmann, P.; Lamrini, S.; Scholle, K.; Fuhrberg, P.; Petermann, K.; Huber, G. Long wavelength laser operation of Tm:Sc2O3 at 2116 nm and beyond. In Advanced Solid-State Photonics; Optical Society of America: Washington, DC, USA, 2011; p. ATuA5. [Google Scholar] [CrossRef] [Green Version]
  18. Lagatsky, A.A.; Koopmann, P.; Fuhrberg, P.; Huber, G.; Brown, C.T.A.; Sibbett, W. Passively mode locked femtosecond Tm:Sc2O3 laser at 2.1 μm. Opt. Lett. 2012, 37, 437–439. [Google Scholar] [CrossRef] [Green Version]
  19. Suzuki, A.; Krankel, C.; Tokurakawa, M. High quality-factor Kerr-lens mode-locked Tm:Sc2O3 single crystal laser with anomalous spectral broadening. Appl. Phys. Express 2020, 13, 052007. [Google Scholar] [CrossRef]
  20. Tokurakawa, M.; Fujita, E.; Suzuki, A.; Krankel, C. Sub-120 fs Kerr-lens mode-locked Tm:Sc2O3 laser at 2.1 µm wavelength range. In Proceedings of the International Conference Laser Optics (ICLO), St. Petersburg, Russia, 4–8 June 2018. [Google Scholar] [CrossRef]
  21. Lupei, V.; Lupei, A.; Gheorghe, C.; Ikesue, A. Spectroscopic characteristics of Tm3+ in Tm and Tm, Nd, Yb:Sc2O3 ceramic. J. Lumin. 2008, 128, 901–904. [Google Scholar] [CrossRef]
  22. Yue, F.; Jambunathan, V.; David, S.P.; Mateos, X.; Aguiló, M.; Díaz, F.; Šulc, J.; Lucianetti, A.; Mocek, T. Spectroscopy and diode-pumped continuous-wave laser operation of Tm:Y2O3 transparent ceramic at cryogenic temperatures. Appl. Phys. B 2020, 126, 44. [Google Scholar] [CrossRef] [Green Version]
  23. Ermeneux, F.S.; Sun, Y.; Cone, R.L.; Equall, R.W.; Hutcheson, R.L.; Moncorge, R. Efficient CW 2 µm Tm3+:Y2O3 Laser. In Advanced Solid State Lasers; Fejer, M., Injeyan, H., Keller, U., Eds.; OSA Trends in Optics and Photonics; Optical Society of America: Washington, DC, USA, 1999; Volume 26, p. TuB8. [Google Scholar] [CrossRef]
  24. Szela, J.W.; Sloyan, K.A.; Parsonage, T.L.; Mackenzie, J.I.; Eason, R.W. Laser operation of a Tm:Y2O3 planar waveguide. Opt. Express 2013, 21, 12460–12468. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, H.; Huang, H.; Liu, P.; Jin, L.; Shen, D.; Zhang, J.; Tang, D. Diode-pumped continuous-wave and Q-switched Tm: Y2O3 ceramics laser around 2050 nm. Opt. Mater. Express 2017, 7, 296–303. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Cai, Y.; Xu, B.; Yang, Y.; Hang, Y. Extending the wavelength tunability from 2.01 to 2.1 μm and simultaneous dual-wavelength operation at 2.05 and 2.3 µm in diode-pumped Tm:YLF lasers. J. Lumin. 2020, 218, 116873. [Google Scholar] [CrossRef]
  27. Schellhorn, M. High-power diode-pumped Tm:YLF laser. Appl. Phys. B 2008, 91, 71–74. [Google Scholar] [CrossRef]
  28. Camy, P.; Doualan, J.L.; Renard, S.; Braud, A.; Menard, V.; Moncorgé, R. Tm3+:CaF2 for 1.9 μm laser operation. Opt. Commun. 2004, 4–6, 395–402. [Google Scholar] [CrossRef]
  29. Ryabochkina, P.A.; Lyapin, A.A.; Osiko, V.V.; Fedorov, P.P.; Ushakov, S.N.; Kruglova, M.V.; Sakharov, N.V.; Garibin, E.A.; Gusev, P.E.; Krutov, M.A. Structural, spectral-luminescent, and lasing properties of nanostructured Tm:CaF2 ceramics. Quantum Electron. 2012, 42, 853–857. [Google Scholar] [CrossRef]
  30. Lyapin, A.A.; Fedorov, P.P.; Garibin, E.A.; Malov, A.V.; Osiko, V.V.; Ryabochkina, P.A.; Ushakov, S.N. Spectroscopic, luminescent and laser properties of nanostructured CaF2:Tm materials. Opt. Mater. 2013, 35, 1859–1864. [Google Scholar] [CrossRef]
  31. Wang, Y.; Lan, R.; Mateos, X.; Li, J.; Li, C.; Suomalainen, S.; Härkönen, A.; Guina, M.; Petrov, V.; Griebner, U. Thulium doped LuAG ceramics for passively mode locked lasers. Opt. Express 2017, 25, 7084–7091. [Google Scholar] [CrossRef]
  32. Barnes, N.P.; Jani, M.G.; Hutcheson, R.L. Diode-pumped, room-temperature Tm:LuAG laser. Appl. Opt. 1995, 34, 4290–4294. [Google Scholar] [CrossRef]
  33. Zhou, Z.; Huang, X.; Guan, X.; Lan, J.; Xu, B.; Xu, H.; Cai, Z.; Liu, P.; Yan, D.; Xu, X.; et al. Continuous-wave and passively Q-switched Tm3+-doped LuAG ceramic lasers. Opt. Mater. Express 2017, 7, 3441–3447. [Google Scholar] [CrossRef]
  34. Caird, J.A.; DeShazer, L.G.; Nella, J. Characteristics of room-temperature 2.3-µm laser emission from Tm3+ in YAG and YALO3. IEEE J. Quantum Electron. 1975, 11, 874–881. [Google Scholar] [CrossRef]
  35. Huber, G.; Duczynski, E.W.; Petermann, K. Laser pumping of Ho-, Tm-, Er-doped garnet lasers at room temperature. IEEE J. Quantum Electron. 1988, 24, 920–923. [Google Scholar] [CrossRef]
  36. Braud, A.; Tigreat, P.Y.; Doualan, J.L.; Moncorgé, R. Spectroscopy and cw operation of a 1.85 µm Tm:KY3F10 laser. Appl. Phys. B 2001, 72, 909–912. [Google Scholar] [CrossRef]
  37. Lisiecki, R.; Solarz, P.; Dominiak-Dzik, G.; Ryba-Romanowski, W.; Lukasiewicz, T. Effect of temperature on spectroscopic features relevant to laser performance of YVO4:Tm3+, GdVO4:Tm3+, and LuVO4:Tm3+ crystals. Opt. Lett. 2010, 35, 3940–3942. [Google Scholar] [CrossRef]
  38. Yu, H.; Pan, Z.; Zhang, H.; Wang, Z.; Wang, J.; Jiang, M. Efficient Tm:LuVO4 laser at 1.9 µm. Opt. Lett. 2011, 36, 2402–2404. [Google Scholar] [CrossRef] [PubMed]
  39. Wang, Y.; Lan, J.L.; Zhou, Z.Y.; Guan, X.F.; Xu, B.; Xu, H.Y.; Cai, Z.P.; Wang, Y.; Tu, C.Y. Continuous-wave laser operation of diode-pumped Tm-doped Gd3Ga5O12 crystal. Opt. Mater. 2017, 66, 185–188. [Google Scholar] [CrossRef]
  40. Qin, Z.P.; Liu, J.G.; Xie, G.Q.; Ma, J.; Gao, W.L.; Qian, L.J.; Yuan, P.; Xu, X.D.; Xu, J.; Zhou, D.H. Spectroscopic characteristics and laser performance of Tm:CaYAlO4 crystal. Laser Phys. 2013, 23, 105806. [Google Scholar] [CrossRef]
  41. Lan, J.L.; Zhang, X.Y.; Zhou, Z.Y.; Xu, B.; Xu, H.Y.; Cai, Z.P.; Chen, N.; Wang, J.; Xu, X.D.; Soulard, R.; et al. Passively Q-Switched Tm:CaYAlO4 Laser Using a MoS2 Saturable Absorber. IEEE Photonics Technol. Lett. 2017, 29, 515–518. [Google Scholar] [CrossRef]
  42. Moncorgé, R.; Guyot, Y.; Kränkel, C.; Lebbou, K.; Yoshikawa, A. 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. 2022, 241, 118537. [Google Scholar] [CrossRef]
  43. Wang, Y.; Zhao, Y.; Pan, Z.; Bae, J.E.; Choi, S.Y.; Rotermund, F.; Loiko, P.; Serres, J.M.; Mateos, X.; Yu, H.; et al. 78 fs SWCNT-SA mode-locked Tm:CLNGG disordered garnet crystal laser at 2017 nm. Opt. Lett. 2018, 43, 4268–4271. [Google Scholar] [CrossRef]
  44. Wang, L.; Chen, W.; Zhao, Y.; Loiko, P.; Mateos, X.; Guina, M.; Pan, Z.; Mero, M.; Griebner, U.; Petrov, V. Sub-50 fs pulse generation from a SESAM mode-locked Tm,Ho-codoped calcium aluminate laser. Opt. Lett. 2021, 46, 2642–2645. [Google Scholar] [CrossRef]
  45. Wang, L.; Chen, W.; Pan, Z.; Loiko, P.; Bae, J.E.; Rotermund, F.; Mateos, X.; Griebner, U.; Petrov, V. Sub-100 fs mode-locked Tm:CLTGG laser. Opt. Express 2021, 29, 31137–31144. [Google Scholar] [CrossRef]
  46. Liu, Z.; Ikesue, A.; Li, J. Research progress and prospects of rare-earth doped sesquioxide laser ceramics. J. Eur. Ceram. Soc. 2021, 41, 3895–3910. [Google Scholar] [CrossRef]
  47. Ivakin, E.V.; Kisialiou, I.G.; Antipov, O.L. Laser ceramics Tm:Lu2O3. Thermal, thermo-optical, and spectroscopic properties. Opt. Mater. 2013, 35, 499–503. [Google Scholar] [CrossRef]
  48. Senatsky, Y.; Shirakawa, A.; Sato, Y.; Hagiwara, J.; Lu, J.; Ueda, K.; Yagi, H.; Yanagitani, T. Nonlinear refractive index of ceramic laser media and perspectives of their usage in a high-power laser-driver. Laser Phys. Lett. 2004, 1, 500–506. [Google Scholar] [CrossRef]
  49. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  50. Gaumé, R.; Viana, B.; Vivien, D.; Roger, J.-P.; Furnier, D. A simple model for the prediction of thermal conductivity in pure and doped insulating crystals. Appl. Phys. Lett. 2003, 83, 1355–1357. [Google Scholar] [CrossRef]
  51. Zhou, Z.Y.; Guan, X.F.; Huang, X.X.; Xu, B.; Xu, H.Y.; Cai, Z.P.; Xu, X.D.; Liu, P.; Li, D.Z.; Zhang, J.; et al. Tm3+-doped LuYO3 mixed sesquioxide ceramic laser: Effective 2.05 µm source operating in continuous-wave and passive Q-switching regimes. Opt. Lett. 2017, 42, 3781–3784. [Google Scholar] [CrossRef]
  52. Zhao, Y.; Wang, L.; Chen, W.; Pan, Z.; Wang, Y.; Liu, P.; Xu, X.; Liu, Y.; Shen, D.; Zhang, J.; et al. SESAM mode-locked Tm:LuYO3 ceramic laser generating 54-fs pulses at 2048 nm. Appl. Opt. 2020, 59, 10493–10497. [Google Scholar] [CrossRef]
  53. Pirri, A.; Maksimov, R.N.; Shitov, V.A.; Osipov, V.V.; Sani, E.; Patrizi, B.; Vannini, M.; Toci, G. Continuously tuned (Tm0.05Sc0.2522Y0.698)2O3 ceramic laser with emission peak at 2076 nm. J. Alloys Compd. 2021, 889, 161585. [Google Scholar] [CrossRef]
  54. Wang, Y.; Jing, W.; Loiko, P.; Zhao, Y.; Huang, H.; Mateos, X.; Suomalainen, S.; Härkönen, A.; Guina, M.; Griebner, U.; et al. Sub-10 optical-cycle passively mode-locked Tm:(Lu2/3Sc1/3)2O3 ceramic laser at 2 µm. Opt. Express 2018, 26, 10299–10304. [Google Scholar] [CrossRef]
  55. Xu, X.; Hu, Z.; Li, D.; Liu, P.; Zhang, J.; Xu, B.; Xu, J. First laser oscillation of diode-pumped Tm3+-doped LuScO3 mixed sesquioxide ceramic. Opt. Express 2017, 25, 15322–15329. [Google Scholar] [CrossRef]
  56. Zhao, Y.; Wang, L.; Chen, W.; Loiko, P.; Wang, Y.; Pan, Z.; Yang, H.; Jing, W.; Huang, H.; Liu, J.; et al. Kerr-lens mode-locked Tm-doped sesquioxide ceramic laser. Opt. Lett. 2021, 46, 3428–3431. [Google Scholar] [CrossRef] [PubMed]
  57. Gruber, J.B.; Seltzer, M.D.; Hills, M.E.; Stevens, S.B.; Morrison, C.A. Energy levels and upconversion fluorescence in trivalent thulium-doped yttrium scandium aluminum garnet. J. Appl. Phys. 1993, 73, 1929–1935. [Google Scholar] [CrossRef]
  58. Feng, Y.; Toci, G.; Patrizi, B.; Pirri, A.; Hu, Z.; Chen, X.; Wei, J.; Pan, H.; Li, X.; Zhang, X.; et al. Fabrication, microstructure, and optical properties of Tm:Y3ScAl4O12 laser ceramics. J. Am. Ceram. Soc. 2020, 103, 1819–1830. [Google Scholar] [CrossRef]
  59. Zhao, Y.; Wang, L.; Wang, Y.; Zhang, J.; Liu, P.; Xu, X.; Liu, Y.; Shen, D.; Eun Bae, J.; Gwan Park, T.; et al. SWCNT-SA mode-locked Tm:LuYO3 ceramic laser delivering 8-optical-cycle pulses at 2.05 µm. Opt. Lett. 2020, 45, 459–462. [Google Scholar] [CrossRef]
  60. Scholle, K.; Lamrini, S.; Koopmann, P.; Fuhrberg, P. 2 µm Laser sources and their possible applications. In Frontiers in Guided Wave Optics and Optoelectronics; Pal, B., Ed.; BoD-Books on Demand; IntechOpen: London, UK, 2010; pp. 471–500. [Google Scholar] [CrossRef] [Green Version]
  61. Gattass, R.R.; Mazur, E. Femtosecond laser micromachining in transparent materials. Nat. Photonics 2008, 2, 219–225. [Google Scholar] [CrossRef]
  62. Singh, U.N.; Walsh, B.M.; Yu, J.; Petros, M.; Kavaya, M.J.; Refaat, T.F.; Barnes, N.P. Twenty years of Tm:Ho:YLF and LuLiF laser development for global wind and carbon dioxide active remote sensing. Opt. Mater. Express 2015, 5, 827–837. [Google Scholar] [CrossRef]
  63. Walsh, B.M.; Lee, H.R.; Barnes, L.B. Mid infrared laser for remote sensing applications. J. Lumin. 2016, 169, 400–405. [Google Scholar] [CrossRef]
  64. Cihelka, J.; Matulková, I.; Civiš, S. Laser diode photoacoustic and FTIR laser spectroscopy of formaldehyde in the 2.3 µm and 3.5 µm spectral range. J. Mol. Spectrosc. 2009, 256, 68–74. [Google Scholar] [CrossRef]
  65. Belyaev, A.N.; Chabushkin, A.N.; Khrushchalina, S.A.; Kuznetsova, O.A.; Lyapin, A.A.; Romanov, K.N.; Ryabochkina, P.A. Investigation of endovenous laser ablation of varicose veins in vitro using 1.885-µm laser radiation. Lasers Med. Sci. 2016, 31, 503–510. [Google Scholar] [CrossRef]
  66. Hale, G.M.; Querry, M.R. Optical Constants of Water in the 200-nm to 200-μm Wavelength Region. Appl. Opt. 1973, 12, 555–563. [Google Scholar] [CrossRef]
  67. Luo, P.L.; Kuo, C.C.; Lee, C.C.; Shy, J.T. Frequency stabilization of a single-frequency volume Bragg grating-based short-cavity Tm:Ho:YLF laser to a CO2 line at 2.06 μm. Appl. Phys. B 2012, 109, 327–331. [Google Scholar] [CrossRef]
  68. Fard, S.T.; Hofmann, W.; Fard, P.T.; Bohm, G.; Ortsiefer, M.; Kwok, E.; Amann, M.C.; Chrostowski, L. Optical absorption glucose measurements using 2.3 µm vertical-cavity semiconductor lasers. IEEE Photon. Technol. Lett. 2008, 20, 930–932. [Google Scholar] [CrossRef]
  69. Budni, P.A.; Pomeranz, L.A.; Lemons, M.L.; Miller, C.A.; Mosto, J.R.; Chicklis, E.P. Efficient midinfrared laser using 1.9-µm-pumped Ho:YAG and ZnGeP2 optical parametric oscillators. J. Opt. Soc. Am. B 2000, 17, 723–728. [Google Scholar] [CrossRef]
  70. Ell, R.; Morgner, U.; Kärtner, F.X.; Fujimoto, J.G.; Ippen, E.P.; Scheuer, V.; Angelow, G.; Tschudi, T.; Lederer, M.J.; Boiko, A.; et al. Generation of 5-fs pulses and octave-spanning spectra directly from a Ti:sapphire laser. Opt. Lett. 2001, 26, 373–375. [Google Scholar] [CrossRef]
  71. Hemming, A.; Richards, J.; Davidson, A.; Carmody, N.; Bennetts, S.; Simakov, N.; Haub, J. 99 W mid-IR operation of a ZGP OPO at 25% duty cycle. Opt. Express 2013, 21, 10062–10069. [Google Scholar] [CrossRef]
  72. James, C. Thulium I. Am. Chem. Soc. 1911, 33, 1332–1344. [Google Scholar] [CrossRef] [Green Version]
  73. Gruber, J.B.; Hills, M.E.; Macfarlane, R.M.; Morrison, C.A.; Turner, G.A.; Quarles, G.J.; Kintz, G.J.; Esterowits, L. Spectra and energy levels of Tm3+:Y3Al5O12. Phys. Rev. B 1989, 40, 9464–9478. [Google Scholar] [CrossRef]
  74. Loiko, P.; Doualan, J.L.; Guillemot, L.; Moncorgé, R.; Starecki, F.; Benayad, A.; Dunina, E.; Kornienko, A.; Fomicheva, L.; Braud, A.; et al. Emission properties of Tm3+-doped CaF2, KY3F10, LiYF4, LiLuF4 and BaY2F8 crystals at 1.5 µm and 2.3 µm. J. Lumin. 2020, 225, 117279. [Google Scholar] [CrossRef]
  75. French, V.A.; Petrin, R.R.; Powell, R.C.; Kotka, M. Energy-transfer processes in Y3Al5O12:Tm,Ho. Phys. Rev. B 1992, 46, 8018–8026. [Google Scholar] [CrossRef]
  76. Stoneman, R.C.; Esterowitz, L. Continuous-wave 1.50 µm thulium cascade laser. Opt. Lett. 1991, 16, 232–234. [Google Scholar] [CrossRef]
  77. Diening, A.; Mobert, P.E.A.; Huber, G. Diode-pumped continuous-wave, quasi-continuous-wave and Q-switched laser operation of Yb3+,Tm3+:YLiF4 at 1.5 and 2.3 µm. J. Appl. Phys. 1998, 84, 5900–5904. [Google Scholar] [CrossRef]
  78. Braud, A.; Girard, S.; Doualan, J.L.; Thuau, M.; Moncorgé, R. Energy-transfer processes in Yb:Tm-doped KY3F10, LiYF4, and BaY2F8 single crystals for laser operation at 1.5 and 2.3 µm. Phys. Rev. B 2000, 61, 5280–5292. [Google Scholar] [CrossRef]
  79. Androz, G.; Bernier, M.; Faucher, D.; Vallée, R. 2.3 W single transverse mode thulium-doped ZBLAN fiber laser at 1480 nm. Opt. Express 2008, 16, 16019–16031. [Google Scholar] [CrossRef]
  80. Antipenko, B.M.; Mak, A.A.; Raba, O.B.; Seirenyan, K.B.; Uvarova, T.V. New lasing transition in the Tm3+ ion. Sov. J. Quantum Electron. 1983, 13, 558–560. [Google Scholar] [CrossRef]
  81. Allain, J.Y.; Monerie, M.; Poignant, H. Tunable cw lasing around 0.82, 1.48, 1.88 and 2.35 μm in thulium-doped fluorozirconate fiber. Electron. Lett. 1989, 25, 1660–1662. [Google Scholar] [CrossRef]
  82. Komukai, T.; Yamamoto, T.; Sugawa, T.; Miyajima, Y. Upconversion Pumped Thulium-Doped Fluoride Fiber Amplifier and Laser Operating at 1.47 µm. IEEE J. Quantum Electron. 1995, 31, 1880–1889. [Google Scholar] [CrossRef]
  83. Renard, S.; Camy, P.; Braud, A.; Doualan, J.L.; Moncorgé, R. CaF2 doped with Tm3+: A cluster model. J. Alloys Compd. 2008, 451, 71–73. [Google Scholar] [CrossRef]
  84. Kintz, G.J.; Allen, R.; Esterowitz, L. Two for one photon conversion observed in alexandrite pumped Tm3+, Ho3+:YAG at room temperature. In Proceedings of the Conference on Lasers and Electro-Optics, Baltimore, MD, USA, 26 April–1 May 1987; Optical Society of America: Washington, DC, USA, 1987; p. ThU4. [Google Scholar]
  85. Johnson, L.F.; Geusic, J.E.; Uitert, L.G.V. Coherent oscillations from Tm3+, Ho3+,Yb3+ and Er3+ ions in yttrium aluminum garnet. Appl. Phys. Lett. 1965, 7, 127–129. [Google Scholar] [CrossRef]
  86. Allen, R.; Esterowitz, L. CW diode pumped 2.3 µm fiber laser. Appl. Phys. Lett. 1989, 55, 721–722. [Google Scholar] [CrossRef]
  87. Canbaz, F.; Yorulmaz, I.; Sennaroglu, A. Kerr-lens mode-locked 2.3-µm Tm3+:YLF laser as a source of femtosecond pulses in the mid-infrared. Opt. Lett. 2017, 42, 3964–3967. [Google Scholar] [CrossRef]
  88. Soulard, R.; Tyazhev, A.; Doualan, J.L.; Braud, A.; Hideur, A.; Laroche, M.; Xu, B.; Camy, P. 2.3 µm Tm3+:YLF mode-locked laser. In Laser Congress 2017 (ASSL, LAC); Optical Society of America: Washington, DC, USA, 2017; p. ATh3A.3. [Google Scholar] [CrossRef]
  89. Muti, A.; Tonelli, M.; Petrov, V.; Sennaroglu, A. Continuous-wave mid-infrared laser operation of Tm3+:KY3F10 at 2.3 µm. Opt. Lett. 2019, 44, 3242–3245. [Google Scholar] [CrossRef]
  90. Walsh, B.M.; Barnes, N.P.; Petros, M.; Yu, J.; Singh, U.N. Spectroscopy and modeling of solid state lanthanide lasers: Application to trivalent Tm3+ and Ho3+ in YLiF4 and LuLiF4. J. Appl. Phys. 2004, 95, 3255–3271. [Google Scholar] [CrossRef]
  91. Pinto, J.F.; Esterowitz, L.; Rosenblatt, G.H. Tm3+:YLF laser continuously tunable between 2.20 and 2.46 µm. Opt. Lett. 1994, 19, 883–885. [Google Scholar] [CrossRef]
  92. Guillemot, L.; Loiko, P.; Braud, A.; Doualan, J.L.; Hideur, A.; Koselja, M.; Moncorge, R.; Camy, P. Continuous-wave Tm:YAlO3 laser at ~2.3 µm. Opt. Lett. 2019, 44, 5077–5080. [Google Scholar] [CrossRef]
  93. Guillemot, L.; Loiko, P.; Kifle, E.; Doualan, J.L.; Braud, A.; Starecki, F.; Georges, T.; Rouvillain, J.; Hideur, A.; Camy, P. Watt-level mid-infrared continuous-wave Tm:YAG laser operating on the 3H4->3H5 transition. Opt. Mater. 2020, 101, 109745. [Google Scholar] [CrossRef]
  94. Baer, J.E.; Knights, M.G.; McCarthy, J.C.; Chicklis, E.P.; Jenssen, H.P. 450-nm operation of XeF-pumped Tm: YLF. In Conference on Lasers and Electro-Optics; Abbott, F., Ready, J., Li, T., Sincerbox, G., Eds.; OSA Technical Digest; Optica Publishing Group: Washington, DC, USA, 1981; p. WA5. [Google Scholar] [CrossRef]
  95. Tanabe, S.; Yoshii, S.; Hirao, K.; Soga, N. Upconversion properties, multiphonon relaxation, and local environment of rare-earth ions in fluorophosphate glasses. Phys. Rev. B 1992, 45, 4620–4625. [Google Scholar] [CrossRef]
  96. Kermaoui, A.; Barthou, C.; Denis, J.P.; Blanzat, B. Excited state absorption mechanisms of red to UV and blue conversion luminescence in Tm3+ doped fluorophosphate glass. J. Lumin. 1984, 29, 85–96. [Google Scholar] [CrossRef]
  97. Nguyen, D.C.; Faulkner, G.E.; Dulick, M. Blue-green (450-nm) upconversion Tm3+:YLF laser. Appl. Opt. 1989, 28, 3553–3555. [Google Scholar] [CrossRef]
  98. Hebert, T.; Wannemacher, R.; Macfarlane, R.M.; Length, W. Blue continuously pumped upconversion lasing in Tm:YLiF4. Appl. Phys. Lett. 1992, 60, 2592–2594. [Google Scholar] [CrossRef]
  99. Thrash, R.J.; Johnson, L.F. Upconversion laser emission from Yb3+-sensitized Tm3+ in BaY2F8. J. Opt. Soc. Am. B 1994, 11, 881–885. [Google Scholar] [CrossRef]
  100. Zinkevich, M. Thermodynamics of rare earth sesquioxides. Prog. Mater. Sci. 2007, 52, 597–647. [Google Scholar] [CrossRef]
  101. Ikesue, A.; Furusato, I.; Kamata, K. Fabrication of Polycrystal line, Transparent YAG Ceramics by a Solid-State Reaction Method. J. Am. Ceram. Soc. 1995, 78, 225–228. [Google Scholar] [CrossRef]
  102. Ikesue, A.; Aung, Y.L.; Taira, T.; Kamimura, T.; Yoshida, K.; Messing, G.L. Progress in ceramic lasers. Annu. Rev. Mater. Res. 2006, 36, 397–429. [Google Scholar] [CrossRef]
  103. Pirri, A.; Toci, G.; Patrizi, B.; Vannini, M. An Overview on Yb-Doped Transparent Polycrystalline Sesquioxide Laser Ceramics. IEEE J. Sel. Top. Quantum Electron. 2018, 24, 1602108. [Google Scholar] [CrossRef]
  104. Kim, W.-J.; Park, J.Y.; Oh, S.J.; Kim, Y.S.; Hong, G.W.; Kuk, I.-H. Characteristics and sintering behavior of Yttria powders synthesized by the combustion process. J. Mater. Sci. Lett. 1999, 18, 411–413. [Google Scholar] [CrossRef]
  105. Roy, S.; Sigmund, W.; Aldinger, F. Nanostructured Yttria powders via gel combustion. J. Mater. Res. 1999, 14, 1524–1531. [Google Scholar] [CrossRef]
  106. Sharma, P.K.; Jilavi, M.H.; Naû, R.; Schmidt, H. Seeding effect in hydrothermal synthesis of nanosize Yttria. J. Mater. Sci. Lett. 1998, 17, 823–825. [Google Scholar] [CrossRef]
  107. Ramanathan, S.; Shankar, P.; Subramanian, K.; Ramakrishnan, S.S.; Angelo, P.C.; Venkataraman, H. Synthesis of nanocrystalline yttria by sol-gel method. Mater. Lett. 2001, 48, 342–346. [Google Scholar] [CrossRef]
  108. Ikegami, T.; Mori, T.; Yajima, Y.; Takenouchi, S.; Misawa, T.; Moriyoshi, Y. Fabrication of transparent Yttria ceramics through the synthesis of Yttrium hydroxide at low temperature and doping by sulfate ions. J. Ceram. Soc. Jpn. 1999, 107, 297–299. [Google Scholar] [CrossRef] [Green Version]
  109. Noriko Saito, N.; Matsuda, S.-I.; Ikegami, T. Fabrication of transparent Yttria ceramics at low temperature using carbonate-derived powder. J. Am. Ceram. Soc. 1998, 81, 2023–2028. [Google Scholar] [CrossRef]
  110. Tool, C.J.J.; Cordfunke, E.H.P. Influence of precipitation on the microstructure and sinterability of yttria. Solid State Ion. 1989, 32–33, 691–697. [Google Scholar] [CrossRef]
  111. Osipov, V.V.; Shitov, V.A.; Maksimov, R.N.; Solomonov, V.I. Properties of transparent Re3+:Y2O3 ceramics doped with tetravalent additives. Opt. Mater. 2015, 50, 65–70. [Google Scholar] [CrossRef]
  112. Pirri, A.; Patrizi, B.; Maksimov, R.N.; Shitov, V.A.; Osipov, V.V.; Vannini, M.; Toci, G. Spectroscopic investigation and laser behaviour of Yb-doped laser ceramics based on mixed crystalline structure (ScxY1−x)2O3. Ceram. Int. 2021, 47, 29483–29489. [Google Scholar] [CrossRef]
  113. Jiang, B.; Hu, C.; Li, J.; Kou, H.; Shi, Y.; Liuy, W.; Pan, Y. Synthesis and properties of Yb:Sc2O3 transparent ceramics. J. Rare Earths 2011, 29, 951–953. [Google Scholar] [CrossRef]
  114. Wang, J.; Ma, J.; Zhang, J.; Liu, P.; Luo, D.; Yin, D.; Tang, D.; Kong, L.B. Yb:Y2O3 transparent ceramics processed with hot isostatic pressing. Opt. Mater. 2017, 71, 117–120. [Google Scholar] [CrossRef]
  115. Seeley, Z.M.; Kuntz, J.D.; Cherepy, N.J.; Payne, S.A. Transparent Lu2O3: Eu ceramics by sinter and HIP optimization. Opt. Mater. 2011, 33, 1721–1726. [Google Scholar] [CrossRef]
  116. Gazza, G.E.; Roderick, D.; Levine, B. Transparent Sc2O3 by hot-pressing. J. Mater. Sci. 1971, 6, 1137–1139. [Google Scholar] [CrossRef]
  117. Sanghera, J.; Kim, W.; Baker, C.; Villalobos, G.; Frantz, J.; Shaw, B.; Lutz, A.; Sadowski, B.; Miklos, R.; Hunt, M.; et al. Laser oscillation in hot pressed 10% Yb3+:Lu2O3 ceramic. Opt. Mater. 2011, 33, 670–674. [Google Scholar] [CrossRef]
  118. An, L.; Ito, A.; Goto, T. Fabrication of transparent lutetium oxide by spark plasma sintering. J. Am. Ceram. Soc. 2011, 94, 695–698. [Google Scholar] [CrossRef]
  119. Futami, Y.; Yanagida, T.; Fujimoto, Y.; Pejchal, J.; Sugiyama, M.; Kurosawa, S.; Yokota, Y.; Ito, A.; Yoshikawa, A.; Goto, T. Optical and scintillation properties of Sc2O3, Y2O3 and Lu2O3 transparent ceramics synthesized by SPS method. Radiat. Meas. 2013, 55, 136–140. [Google Scholar] [CrossRef]
  120. Zhou, D.; Shi, Y.; Xie, J.; Ren, Y.; Yun, P. Fabrication and luminescent properties of Nd3+-Doped Lu2O3 transparent ceramics by pressureless sintering. J. Am. Ceram. Soc. 2009, 92, 2182–2187. [Google Scholar] [CrossRef]
  121. Jing, W.; Loiko, P.; Serres, J.M.; Wang, Y.; Vilejshikova, E.; Aguiló, M.; Díaz, F.; Griebner, U.; Huang, H.; Petrov, V.; et al. Synthesis, spectroscopy, and efficient laser operation of "mixed" sesquioxide Tm:(Lu,Sc) 2O3 transparent ceramics. Opt. Mater. Express 2017, 7, 4192–4202. [Google Scholar] [CrossRef]
  122. Wu, H.; Pan, G.H.; Hao, Z.; Zhang, L.; Zhang, X.; Zhang, L.; Zhao, H.; Zhang, J. Laser-quality Tm:(Lu0.8Sc0.2)2O3 mixed sesquioxide ceramics shaped by gelcasting of well-dispersed nanopowders. J. Am. Ceram. Soc. 2019, 102, 4919–4928. [Google Scholar] [CrossRef]
  123. Loiko, P.; Koopmann, P.; Mateos, X.; Serres, J.; Jambunathan, V.; Lucianetti, A.; Mocek, T.; Aguilo, M.; Diaz, F.; Griebner, U.; et al. Highly efficient, compact Tm3+:RE2O3 (RE = Y, Lu, Sc) sesquioxide lasers based on thermal guiding. IEEE J. Sel. Top. Quantum Electron. 2019, 24, 1–13. [Google Scholar] [CrossRef]
  124. Liferovich, R.P.; Mitchell, R.H. A structural study of ternary lanthanide orthoscandate perovskites. J. Solid State Chem. 2004, 177, 2188–2197. [Google Scholar] [CrossRef]
  125. Sarkarainadar, B.; Rodewald, U.C.; Thomas, H.; van Wüllen, L.; Mohr, D.; Heinz, D.; Eckert, H.; Pöttgen, R. PbO/PbF2 Flux Growth of YScO3 and LaScO3 Single Crystals—Structure and Solid-State NMR Spectroscopy. Z. Nat. B 2010, 65, 1199–1205. [Google Scholar] [CrossRef] [Green Version]
  126. Shimizu, Y.; Ueda, K. Phase formation and UV luminescence of Gd3+ doped perovskite-type YScO3. J. Solid State Chem. 2016, 242, 170–174. [Google Scholar] [CrossRef]
  127. Clark, J.B.; Richter, P.W.; DuToit, L. 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. 1978, 23, 129–134. [Google Scholar] [CrossRef]
  128. Aggarwal, R.L.; Ripin, D.J.; Ochoa, J.R.; Fan, T.Y. Measurement of thermo-optic properties of Y3Al5O12, Lu3Al5O12, YAlO3, LiYF4, LiLuF4, BaY2F8, KGd(WO4)2, and KY(WO4)2 laser crystals in the 80–300 K temperature range. J. Appl. Phys. 2005, 98, 1035141–10351414. [Google Scholar] [CrossRef]
  129. Klemens, P.G. Thermal resistance due to point defects at high temperatures. Phys. Rev. 1960, 119, 507–509. [Google Scholar] [CrossRef]
  130. Peters, R.; Kränkel, C.; Fredrich-Thornton, S.T.; Beil, K.; Petermann, K.; Huber, G.; Heckl, O.H.; Baer, C.R.E.; Saraceno, C.J.; Südmeyer, T.; et al. Thermal analysis and efficient high power continuous-wave and mode-locked thin disk laser operation of Yb-doped sesquioxides. Appl. Phys. B 2011, 102, 509–514. [Google Scholar] [CrossRef]
  131. Liu, W.; Lu, D.; Guo, R.; Wu, K.; Pan, S.; Hang, Y.; Sun, D.; Yu, H.; Yu, H. Broadening of the Fluorescence Spectra of Sesquioxide Crystals for Ultrafast Lasers. Cryst. Growth Des. 2020, 20, 4678–4685. [Google Scholar] [CrossRef]
  132. Kränkel, C.; Uvarova, A.; Haurat, É.; Hülshoff, L.; Brützam, M.; Guguschev, C.; Kalusniak, S.; Klimm, D. Czochralski growth of mixed cubic sesquioxide crystals in the ternary system Lu2O3-Sc2O3-Y2O3. Acta Cryst. 2021, B77, 550–558. [Google Scholar] [CrossRef]
  133. Hu, Q.; Jia, Z.; Tang, C.; Lin, N.; Zhang, J.; Jia, N.; Wang, S.; Zhao, X.; Tao, X. The origin of coloration of CaGdAlO4 crystals and its effect on their physical properties. CrystEngComm 2017, 19, 537–545. [Google Scholar] [CrossRef]
  134. Pan, Z.; Loiko, P.; Wang, Y.; Zhao, Y.; Yuan, H.; Tang, K.; Dai, X.; Cai, H.; Serres, J.M.; Slimi, S.; et al. Disordered Tm3+,Ho3+-codoped CNGG garnet crystal: Towards efficient laser materials for ultrashort pulse generation at 2 µm. J. Alloys Compd. 2021, 853, 157100. [Google Scholar] [CrossRef]
  135. Pirri, A.; Toci, G.; Patrizi, B.; Maksimov, R.N.; Osipov, V.V.; Shitov, V.A.; Vannini, M. Yb3+:(LuxY1−x)2O3 mixed sesquioxide ceramics for laser applications. Part I: Fabrication, microstructure and spectroscopy. J. Alloys Compd. 2021, 869, 159227. [Google Scholar] [CrossRef]
  136. Badie, J.M.; Foex, M. Determination experimentale, calcul et prevision de certains diagrammes Sc2O3-Ln2O3. J. Sol. State Chem. 1978, 26, 311–319. [Google Scholar] [CrossRef]
  137. Badie, J.M. High-temperature phases and phase-transitions in the systems Sc2O3-Ln2O3 (Ln= lanthanide and yttriun). Rev. Int. Hautes Temp. Refract. 1978, 15, 183–189. Available online: https://0-scholar-google-com.brum.beds.ac.uk/scholar?q=High-temperature%20phases%20and%20phase-transitions%20in%20the%20systems%20Sc2O3Ln2O3 (accessed on 8 February 2022).
  138. Li, J.G.; Ikegami, T.; Mori, T.; Yajima, Y. Monodispersed Sc2O3 precursor particles via homogeneous precipitation: Synthesis, thermal decomposition, and the effects of supporting anions on powder properties. J. Mater. Res. 2003, 18, 1149–1156. [Google Scholar] [CrossRef]
  139. Barrera, E.W.; Pujol, M.C.; Cascales, C.; Zaldo, C.; Park, K.H.; Choi, S.B.; Rotermund, F.; Carvajal, J.J.; Mateos, X.; Aguilo, M.; et al. Spectroscopic characterization of sol-gel synthesized Tm:Lu2O3 nanocrystals. Appl. Phys. B 2012, 106, 409–417. [Google Scholar] [CrossRef]
  140. Liu, Q.; Dai, Z.F.; Hreniak, D.; Li, S.S.; Liu, W.B.; Wang, W.; Luo, W.; Li, C.; Dai, J.; Chen, H.; et al. Fabrication of Yb:Sc2O3 laser ceramics by vacuum sintering co-precipitated nano-powders. Opt. Mater. 2017, 72, 482–490. [Google Scholar] [CrossRef]
  141. Liu, W.; Kou, H.; Li, J.; Jiang, B.; Pan, Y. Transparent Yb:(LuxSc1−x)2O3 ceramics sintered from carbonate co-precipitated powders. Ceram. Int. 2015, 41, 6335–6339. [Google Scholar] [CrossRef]
  142. Jing, W.; Loiko, P.; Serres, J.M.; Wang, Y.; Kifle, E.; Vilejshikova, E.; Aguiló, M.; Díaz, F.; Griebner, U.; Huang, H.; et al. Synthesis, spectroscopic characterization and laser operation of Ho3+ in "mixed" (Lu,Sc)2O3 ceramics. J. Lumin. 2018, 203, 145–151. [Google Scholar] [CrossRef]
  143. Koopmann, P.; Lamrini, S.; Scholle, K.; Fuhrberg, P.; Petermann, K.; Huber, G. Laser Operation and Spectroscopic Investigation of Tm:LuScO3 CLEO/Europe and EQEC 2011 Conference Digest; OSA Technical Digest (CD); Optical Society of America: Washington DC, USA, 2011; p. CA1_4. [Google Scholar] [CrossRef]
  144. Koopmann, P.; Peters, P.; Petermann, K.; Huber, G. Crystal growth, spectroscopy, and highly efficient laser operation of thulium-doped Lu2O3 around 2 µm. Appl. Phys. B 2011, 102, 19–24. [Google Scholar] [CrossRef]
  145. Lagatsky, A.A.; Koopmann, P.; Antipov, L.; Brown, C.T.A.; Huber, G.; Sibbett, W. Femtosecond pulse generation with Tm-doped sesquioxides. In Proceedings of the Conference on Lasers and Electro-Optics—International Quantum Electronics Conference, Munich, Germany, 12–16 May 2013; Optical Society of America: Washington, DC, USA, 2013; p. CA_6_3. [Google Scholar]
  146. Stevenson, N.K.; Brown, C.T.A.; Hopkins, J.M.; Dawson, M.D. Diode-pumped femtosecond Tm3+-doped LuScO3 laser near 2.1 µm. Opt. Lett. 2017, 43, 1287–1290. [Google Scholar] [CrossRef] [Green Version]
  147. Li, D.; Kong, L.; Xu, X.; Liu, P.; Xie, G.; Zhang, J.; Xu, J. Spectroscopy and mode-locking laser operation of Tm:LuYO3 mixed sesquioxide ceramic. Opt. Express 2019, 27, 24416–24425. [Google Scholar] [CrossRef]
  148. Ryabochkina, P.A.; Chabushkin, A.N.; Kopylov, Y.L.; Balashov, V.V.; Lopukhin, K.V. Two-micron lasing in diode-pumped Tm:Y2O3 ceramics. Quantum Electron. 2016, 46, 597–600. [Google Scholar] [CrossRef]
  149. Zhao, Y.; Wang, L.; Pan, Z.; Wang, Y.; Zhang, J.; Liu, P.; Xu, X.; Shen, D.; Xu, J.; Suomalainen, S.; et al. Sub-60 fs SESAM mode-locked Tm:LuYO3 ceramic laser. In Proceedings of the 2019 Conference on Lasers and Electro-Optics Europe and European Quantum Electronics Conference, Munich, Germany, 23–27 June 2019; OSA Technical Digest; Optical Society of America: Washington, DC, USA, 2019; p. ca_6_2. [Google Scholar]
  150. Chen, G.; Li, S.; Zhang, L.; Tan, X.; Deng, W.; He, M.; Xu, M.; Yang, Y.; Zhang, S.; Hang, Y. Growth and spectra of Tm3+ doped LuYO3 single crystal for 2 µm lasers. Infrared Phys. Technol. 2020, 109, 103431. [Google Scholar] [CrossRef]
  151. Bagdasarov, K.S.; Kaminskii, A.A.; Kevorkov, A.M.; Li, L.; Prokhorov, A.M.; Tevosyan, T.A.; Sarkisov, S.E. Investigation of the stimulated emission of cubic crystals of YScO3 with Nd3+ ions. Sov. Phys. Dokl. 1976, 20, 681–683. [Google Scholar]
  152. Tissue, B.M.; Cockroft, N.J.; Lu, L.; Nguyen, D.C.; Yen, W.M. Comparison of the spectra and dynamics of Er3+:Y2−xScxO3 (x= 0, 1, 2). J. Lumin. 1991, 48, 477–480. [Google Scholar] [CrossRef]
  153. Alimov, O.; Dobretsova, E.; Guryev, D.; Kashin, V.; Kiriukhina, G.; Kutovoi, S.; Rusanov, S.; Simonov, S.; Tsvetkov, V.; Vlasov, V.; et al. Growth and characterization of neodymium-doped yttrium scandate crystal fiber with a bixbyite-type crystal structure. Cryst. Growth Des. 2020, 20, 4593–4599. [Google Scholar] [CrossRef]
  154. Lu, S.; Yang, Q.; Shi, Z. The effect of Sc2O3 on Yb3+-doped Y2O3 transparent ceramics. Adv. Mater. Res. 2011, 299, 629–632. [Google Scholar] [CrossRef]
  155. Lu, S.; Yang, Q.; Zhang, H.; Wang, Y.; Huang, D. Fabrication and spectral properties of Yb:(Sc0.9Y0.1)O3 transparent ceramics. Opt. Mater. 2013, 35, 793–797. [Google Scholar] [CrossRef]
  156. Pirri, A.; Maksimov, R.N.; Shitov, V.A.; Osipov, V.V.; Tikhonov, E.V.; Toci, G.; Patrizi, B.; Vannini, M. Novel Tm:(Y,Sc) 2O3 Transparent Ceramics for Laser Applications. In Proceedings of the Conference on Lasers and Electro-Optics Europe & European Quantum Electronics Conference (CLEO/Europe-EQEC), Munich, Germany, 21–25 June 2021; p. ce_10_3. [Google Scholar] [CrossRef]
  157. Liu, Z.; Toci, G.; Pirri, A.; Patrizi, B.; Li, J.; Hu, Z.; Wei, J.; Pan, H.; Xie, T.; Vannini, M.; et al. Fabrication and laser operation of Yb:Lu2O3 transparent ceramics from co-precipitated nano-powders. J. Am. Ceram. Soc. 2019, 102, 7491–7499. [Google Scholar] [CrossRef]
  158. Dai, Z.; Liu, Q.; Toci, G.; Vannini, M.; Pirri, A.; Babin, V.; Nikl, M.; Wang, W.; Chen, H.; Li, J. Fabrication and laser oscillation of Yb:Sc2O3 transparent ceramics from co-precipitated nano-powders. J. Eur. Ceram. Soc. 2018, 38, 1632–1638. [Google Scholar] [CrossRef]
  159. Toci, G.; Pirri, A.; Patrizi, B.; Maksimov, R.N.; Osipov, V.V.; Shitov, V.A.; Yurovskikh, A.S.; Vannini, M. High efficiency emission of a laser based on Yb-doped (Lu,Y)2O3 ceramic. Opt. Mater. 2018, 83, 182–186. [Google Scholar] [CrossRef]
  160. Toci, G.; Pirri, A.; Li, J.; Xie, T.; Pan, Y.; Babin, V.; Beitlerova, A.; Nikl, M.; Vannini, M. First laser emission of Yb0.15:(Lu0.5Y0.5)3Al5O12 ceramics. Opt. Express 2016, 24, 9611–9616. [Google Scholar] [CrossRef]
Figure 1. Energy level scheme of Tm3+ in Lu2O3 (adapted from [47]).
Figure 1. Energy level scheme of Tm3+ in Lu2O3 (adapted from [47]).
Materials 15 02084 g001
Figure 2. Scheme of the cross-relaxation (CR) process.
Figure 2. Scheme of the cross-relaxation (CR) process.
Materials 15 02084 g002
Table 1. Spectroscopic data on the mixed ceramics reported in the literature; λabs and λem are the wavelengths at which the absorption and emission cross-sections were calculated, respectively.
Table 1. Spectroscopic data on the mixed ceramics reported in the literature; λabs and λem are the wavelengths at which the absorption and emission cross-sections were calculated, respectively.
SampleDoping
at.%
σabs
(×10−21 cm−2)
λabs
(nm)
σem
(×10−21 cm−2)
λem
(nm)
Grain Size
(µm)
Lattice Const.
(Å)
Ref.
LuScO323.5793- 1.65 [55]
(Lu2/3Sc1/3)2O34.762.87937.1519514–510.3683[121]
(Lu2/3Sc1/3)2O34.764.216222.382090 [121]
Lu1.6Sc0.4O31.53.17961.1120901.54 [54]
Lu0.8Sc0.2O32- - 2.5 [122]
(Lu2/3Sc1/3)2O32.8- 7.01950--[56]
LuYO333.87966.019371.65-[51]
LuYO333.02055- 1.65-[147]
LuScO3 crystal12.67938.01956-10.105[143]
(Sc1/4Y3/4)2O35 3.9209828.210.401[53]
Table 2. CW and quasi-CW laser performance of the mixed ceramics reported in the literature; Pout: laser output power; λL: laser wavelength emission; η: slope efficiency; Pth: laser threshold.
Table 2. CW and quasi-CW laser performance of the mixed ceramics reported in the literature; Pout: laser output power; λL: laser wavelength emission; η: slope efficiency; Pth: laser threshold.
SampleDoping
at.%
Pout
(W)
λL
(nm)
η
(%)
Pth
(W)
Ref.
LuScO320.21119828.20.840[55]
(Lu2/3Sc1/3)2O34.7612100240.860[121]
Lu1.6Sc0.4O319.82090402.8[54]
Lu1.6Sc0.4O31.5112090395.0[54]
Lu0.8Sc0.2O321.88209024.63.2[122]
(Lu2/3Sc1/3)2O32.80.490208826.8-[56]
LuYO331.55205019.91.1[51]
LuYO331.20206725.10.530[147]
LuYO330.4402074-0.140[59]
YO330.603206033.20.250[52]
LuYO330.600207611.50.250[52]
(Sc1/4Y3/4)O351.2420779.453.49[53]
LuScO3 crystal10.2501982550.038[143]
LuScO3 crystal10.7052100550.038[143]
Table 3. Data obtained in the pulsed regime of the mixed ceramics reported in the literature; λL: laser wavelength emission; Pout: laser output power; τL: pulse duration; f: repetition rate.
Table 3. Data obtained in the pulsed regime of the mixed ceramics reported in the literature; λL: laser wavelength emission; Pout: laser output power; τL: pulse duration; f: repetition rate.
SampleλL
(nm)
Pout
(mW)
τL
(fs)
f
(MHz)
Ref.
LuScO3197532590 ps34.72[55]
Lu2/3Sc1/3O32057306378.9[54]
Lu2/3Sc1/3O320812205884.8[56]
LuYO32048515478[52]
LuYO32045635772.6[59]
LuYO32061121410139.3[147]
LuYO32047540120.3 ns26.31[51]
LuScO3c2093113170115.2[146]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pirri, A.; Maksimov, R.N.; Li, J.; Vannini, M.; Toci, G. Achievements and Future Perspectives of the Trivalent Thulium-Ion-Doped Mixed-Sesquioxide Ceramics for Laser Applications. Materials 2022, 15, 2084. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15062084

AMA Style

Pirri A, Maksimov RN, Li J, Vannini M, Toci G. Achievements and Future Perspectives of the Trivalent Thulium-Ion-Doped Mixed-Sesquioxide Ceramics for Laser Applications. Materials. 2022; 15(6):2084. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15062084

Chicago/Turabian Style

Pirri, Angela, Roman N. Maksimov, Jiang Li, Matteo Vannini, and Guido Toci. 2022. "Achievements and Future Perspectives of the Trivalent Thulium-Ion-Doped Mixed-Sesquioxide Ceramics for Laser Applications" Materials 15, no. 6: 2084. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15062084

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop