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

Growth, structure, and polarized spectroscopy of monoclinic Er3+:MgWO4 crystal

Open Access Open Access

Abstract

We report on the growth, structure, and polarized spectroscopy of a novel promising laser crystal, erbium-doped magnesium monotungstate, Er3+:MgWO4. 1.01 t.% Er3+:MgWO4 was grown by the Top-Seeded Solution Growth method using Na2WO4 as a solvent. The crystal structure was refined by the Rietveld method. Er3+:MgWO4 belongs to the monoclinic class (sp. gr. P2/c, wolframite-type structure, lattice parameters: a = 4.6939(6) Å, b = 5.6747(4) Å, c = 4.9316(6) Å and β = 90.7858(4) Å. The transition intensities for Er3+ ions were determined using the Judd-Ofelt theory accounting for an intermediate configuration interaction (ICI). Er3+ ions in MgWO4 exhibit intense, strongly polarized and broad absorption and emission bands owing to their accommodation in distorted low-symmetry sites (C2). The stimulated-emission cross-section for the 4I13/24I15/2 transition is 0.31×10−20 cm2 at 1637 nm (light polarization: E || b). The radiative lifetime of the 4I13/2 state is 4.85 ± 0.05 ms. The multiphonon non-radiative relaxation for Er3+ excited multiplets is quantified. Er3+ ions in MgWO4 feature large Stark splitting of the ground-state, ΔE(4I15/2) = 435 cm-1. Er3+:MgWO4 is attractive for low-threshold lasers at ∼1.64 µm.

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

1. Introduction

The Erbium ion (Er3+) is well known for its emission in the eye-safe spectral range of ∼1.5 - 1.6 µm originating from the 4I13/24I15/2 electronic transition. The applications of eye-safe erbium lasers are in the fields of range-finding, environmental sensing, aerial navigation and telecom. So far, Er3+,Yb3+-codoped phosphate glasses represent the state-of-the-art erbium laser media [1]. Here, the Yb3+ ion acts as a sensitizer to enhance the pump absorption at ∼0.96 - 0.98 µm, the spectral range addressed by commercial InGaAs laser diodes [2]. Such glasses feature attractive spectroscopic behavior, i.e., broadband Yb3+ absorption, efficient Yb3+ → Er3+ energy transfer and long Er3+ upper laser lifetime. Er,Yb:phosphate glass lasers can generate mJ-level nanosecond pulses making them ideal for portable range-finders [3]. At the same time, glasses suffer from poor thermo-mechanical properties limiting the power scaling capabilities.

As an alternative to glassy gain media, single crystals can be used. However, so far, only a few materials were found to be suitable for efficient lasers based on the Yb3+,Er3+ codoping scheme. These include vanadates (YVO4 [4]), borates (REAl3(BO3)4 [5,6], RECa4O(BO3)3 [7]), and silicates (Y2SiO5 [8]). Tolstik et al. demonstrated an Er,Yb:YAl3(BO3)4 (Er,Yb:YAB) laser delivering an output power of 1 W at 1555 nm with a slope efficiency of 35% [5]. Still, further power scaling of such lasers is limited by severe thermo-optic effects originating from upconversion losses. Moreover, the growth technology of REAl3(BO3)4 borate crystals is far from being mature.

An alternative approach is the use of singly Er3+-doped crystals which can be pumped at ∼1.54 µm. This corresponds to excitation directly to the upper laser level (in-band or resonant pumping). Owing to the relatively high absorption cross-sections for the 4I15/24I13/2 Er3+ transition, high pump efficiencies can be achieved even for low Er3+ doping levels (∼1 at.%). This leads to weak energy-transfer upconversion, which, together with the low quantum effect inherent to the resonant pumping scheme, determines weak heat loading and high laser slope efficiencies. Contrary to the Er3+, Yb3+ codoping scheme for which only a few particular host matrices are suitable, single Er3+ doping is more advantageous from the material point of view: efficient laser operation can be achieved in many singly Er3+-doped laser materials (crystals) under in-band pumping. In-band pumped Er:Y3Al5O12 (Er:YAG) lasers have delivered multi-watt output at 1617 nm and 1645 nm [9,10]. However, one disadvantage of cubic Er:YAG crystals are the depolarization losses appearing at high pump powers in lasers containing polarization-selective elements. Thus, it is still important to search for novel singly Er3+-doped materials offering intrinsic optical anisotropy, broad and intense spectral bands for polarized light, and good thermo-mechanical properties.

Magnesium monotungstate (MgWO4) has recently emerged as a promising host crystal for doping with laser-active rare-earth ions [1113]. This crystal is monoclinic and optically biaxial offering strong optical anisotropy. It was found to exhibit good thermo-mechanical properties, i.e., high thermal conductivity of ∼8.7 Wm-1K-1 [14] and low anisotropy of thermal expansion [13]. High-power laser operation was demonstrated using such crystals: Loiko et al. developed a diode-pumped Yb3+:MgWO4 laser delivering 18.2 W at ∼1056 nm with a high slope efficiency of ∼89% and a linearly polarized output [12]. The low-symmetry distorted coordination of rare-earth ions in MgWO4 leads to broad emission bands [15]. Because of the broadband emission properties, such crystals have been implemented in femtosecond mode-locked lasers [16,17]. Although Er3+:MgWO4 has never been studied so far, other monoclinic crystals doped with Er3+ ions appear promising for the development of eye-safe lasers: Serres et al. reported on a diode-pumped 1 at.% Er:KLu(WO4)2 laser generating 268 mW at 1610 nm with a slope efficiency of 30% [18].

In the present work, we report on the growth, structure refinement and a polarization-resolved spectroscopic study of an Er3+-doped MgWO4 crystal, for the first time, to the best of our knowledge.

2. Crystal growth and structure

2.1 Crystal growth

Er3+:MgWO4 was grown by the top-seeded solution growth (TSSG) method [11,19] from the flux with a composition of MgWO4: Na2WO4 = 5:7 mol (sodium tungstate, Na2WO4, was used as a solvent) in a vertical tubular furnace. The starting materials were Na2CO3, MgO, WO3 (purity: analytical grade) and Er2O3 (dopant, purity: 99.99%). They were weighed according to the above-mentioned composition with the initial Er3+ concentration of 10 at.% (with respect to Ca2+). The weighed materials were mixed, ground and then put into a Pt crucible with dimensions of Ø55 × 60 mm3. The crucible was placed into a resistive furnace equipped with a programmable temperature controller, nickel-chrome heating wires and a Pt–Rh/Pt thermocouple. A [010]-oriented seed from an undoped MgWO4 crystal was used. The solution was kept at 980°C for 2 days to ensure that the starting materials melted completely and homogeneously. The saturation temperature was determined to be 953°C by repeated seeding trials. The crystal was grown at a cooling rate of 0.6 - 1°C/day and a rotation speed of 10 rpm in the temperature range of 953 - 930°C. Once the growth was completed, the crystal was slowly pulled out of the solution and cooled down to room temperature (RT, 293 K) at a rate of 10 K/h. Figure 1 shows an as-grown Er3+:MgWO4 crystal with dimensions of 20 × 8 × 6 mm3. It has a rose coloration due to erbium doping.

 figure: Fig. 1.

Fig. 1. A photograph of the as-grown 1.01 at.% Er3+:MgWO4 crystal, the growth direction is along the [010] axis.

Download Full Size | PDF

The actual Er3+ doping level was determined by inductively coupled plasma atomic emission spectrometry (ICP-AES, Ultima2, Jobin-Yvon) to be 1.01 at.% (ion density: NEr = 1.419 × 1020 at/cm3). No significant variation of the Er3+ doping level across the crystal boule was observed. The segregation coefficient of Er3+ KEr was then 0.1. Such a low value is due to a significant difference of ionic radii of Mg2+ and Er3+ (see below). Despite the low KEr value for MgWO4, it is relatively easy to access the desired doping level typical for singly Er3+-doped crystals (∼1 at.%). Higher doping levels can be accessed by increasing the initial content of Er3+ ions in the growth charge or probably by using other charge compensators (e.g., Li+ cations). However, in in-band-pumped Er lasers, the doping levels above 1 at.% are usually not used due to the enhanced energy-transfer upconversion.

2.2 Structure refinement and electronic structure

The X-ray powder diffraction (XRD) data were collected using a Rigaku MiniFlex 600 X-ray diffractometer with CuKα radiation (λ = 1.5418 Å) in an angular range of 2θ = 10-80° with a scan step of 0.02° and a scan speed of 5°/min, as shown in Fig. 2(a). The measured XRD pattern was well assigned using a standard diffraction pattern of undoped MgWO4 (JCPDS card #96-101-0643) and no other phases were found.

 figure: Fig. 2.

Fig. 2. Structural study of Er3+:MgWO4: (a) X-ray powder diffraction (XRD) analysis: observed (black), calculated (red) and residual (blue) patterns, dashes – Bragg positions, (hkl) – Miller’s indices; (b,c) projection of the crystal structure on (b) the a-c plane and (c) the b-c plane, black lines – unit-cell.

Download Full Size | PDF

The crystal structure was refined by the Rietveld method (Match3 software was used). The crystallographic data of undoped MgWO4 [20] were used as a starting model. Er3+:MgWO4 crystallizes in the monoclinic system with space group C42h - P2/c, No. 13, and centrosymmetric point group 2/m. The lattice parameters are a = 4.6939(6) Å, b = 5.6747(4) Å, c = 4.9316(6) Å, monoclinic angle of β = 90.7858(4)° (number of formula units per unit-cell Z = 2), unit-cell volume of V = 131.351 Å3 and calculated density ρcalc = 6.078 g/cm3. The reliability factors are Rp = 10.5%, Rwp = 14.5%, Rexp = 10.29% and χ2 = (Rwp/Rexp)2 = 1.98 indicating good convergence of the fit.

The determined fractional atomic coordinates, site occupancy factors (O.F.) and isotropic displacement parameters Biso are listed in Table 1. Figures 2(b) and (c) present a fragment of the crystal structure calculated according to the determined atomic positions. Er3+:MgWO4 exhibits a wolframite [(Fe,Mn)WO4] type structure (MgWO4 is called huanzalaite in mineral form). The Mg2+|Er3+ cations occupy the 2f Wyckoff positions with VI-fold O2- coordination; in the distorted [Mg|ErO6] octahedra, there are two shorter [1.9392(0) Å], two intermediate [2.1732(0) Å] and two longer [2.2199(4) Å] Mg-O distances. The W6+ cations (2e Wyckoff positions) are also located in distorted octahedra with W-O distances in the range 1.7747(7) - 2.1673(4) Å. The network of Er3+:MgWO4 is made up of alternating zig-zag chains of edge-sharing [Mg|ErO6] and [WO6] polyhedra along the c-axis. The shortest distance Mg|Er - Mg|Er is 3.3323(4) Å is observed along the vector [u v w]:[0 - 0.3950 0.5].

Tables Icon

Table 1. Fractional Atomic Coordinates (x, y, z), Site Occupancy Factors (O.F.) and Isotropic Displacement Parameters Biso2) for Er3+:MgWO4

In MgWO4, erbium ions are expected to replace for the host-forming cations Mg2+ in a single type of sites (2f, site symmetry: C2). The local charge compensation is most probably ensured by univalent Na+ cations entering from the Na2WO4 solvent (Er3+ + Na+ ↔ 2Mg2+). The corresponding ionic radii for VI-fold oxygen coordination are: RMg = 0.72 Å, REr = 0.89 Å and RNa = 1.02 Å [21]. Large ionic radii of the dopant and charge compensation cations determine the expansion of the unit-cell [for MgWO4, a = 4.68892(2) Å, b = 5.67529(3) Å, c = 4.92891(2) Å and β = 90.726(1) Å] [20]. The difference in the charge and ionic radii of Mg2+, Er3+ and Na+ is expected to distort the crystal field around the Er3+ ions leading to additional spectral broadening.

The electronic structure of the host matrix, MgWO4, was analyzed by the density functional theory in which the generalized gradient approximation with the Perdew–Burke–Ernzerhof functional was used to study the exchange–correlation effects. The calculations were performed using the CASTEP code. The energy cutoff was set to 517 eV. The criterion for the self-consistent field was eigenenergy convergence within 2.0×10−6 eV per atom. K-space sampling was performed using a Monkhorst–Pack grid of 4 × 4 × 4 atoms with respect to k-points in the irreducible Brillouin zone.

The results are shown in Fig. 3(a). The top of the valence band (VB) and the bottom of the conduction band (CB) are located at different points in the Brillouin zone (B and Y, respectively), indicating an indirect bandgap of Eg,calc = 3.42 eV. For Er3+-doped MgWO4, Eg,calc is higher, 3.47 eV, in agreement with the optical bandgap derived from the absorption spectrum using the Tauc plot [22], Eg = 3.53 eV. The assignment of electronic bands was performed with the help of calculated total and partial densities of states, Fig. 3(b). The VB band in the range from the Fermi energy level around 0 eV to -6.26 eV is mainly derived from the p-states of O. The CB extends from 3.42 to 22 eV and it is mainly due to the p- and s-states of Mg.

 figure: Fig. 3.

Fig. 3. Electronic structure of MgWO4: (a) calculated band structure, Eg – bandgap energy; (b) calculated total and partial densities of states (TDOS and PDOS).

Download Full Size | PDF

3. Polarized optical spectroscopy

3.1 Experimental

MgWO4 is an optically biaxial crystal. For the polarization-resolved study, we have prepared a six-side polished rectangular sample oriented in the crystallographic frame by means of single-crystal XRD having the dimensions 4.39(a) × 4.15(b) × 4.35(c*) mm3. Here, c* is a direction being orthogonal to the a-axis and lying in the a-c plane (as the monoclinic angle of Er3+:MgWO4 is close to 90°, c* ≈ c below in the text).

The spectroscopic studies were performed at RT and low temperature (LT, 10 K). The absorption spectra were measured using a spectrophotometer (Lambda 1050, Perkin Elmer) and the luminescence spectra - by an optical spectrum analyzer (AQ6375B, Yokogawa). For polarization-resolved studies, a Glan-Taylor polarizer was implemented. The luminescence decay curves were measured using a nanosecond optical parametric oscillator (Horizon, Continuum), a 1/4 m monochomator (Oriel 77200), a PMT tube or an InGaAs detector and an 8 GHz digital oscilloscope (DSA70804B, Tektronix). For LT studies, the crystal was mounted on an APD DE-202 closed-cycle cryo-cooler equipped with an APD HC 2 Helium vacuum cryo-compressor and a Laceshore 330 temperature controller.

3.1 Absorption spectra and Judd-Ofelt analysis

The RT polarized absorption spectra of Er3+ in MgWO4 are shown in Fig. 4. The crystal exhibits a significant anisotropy of the absorption properties. For the 4I15/24I11/2 transition which can be addressed by commercial InGaAs diode lasers, the maximum absorption cross-section σabs is 0.77 × 10−20 cm2 at 983.7 nm and the corresponding absorption bandwidth (full width at half maximum, FWHM) is 4.1 nm (for E || b). Slightly lower σabs = 0.74 × 10−20 cm2 at 980.8 nm is observed for E || c at a much narrower bandwidth, 1.4 nm.

 figure: Fig. 4.

Fig. 4. (a-f) RT (293 K) absorption cross-section, σabs, spectra of Er3+:MgWO4 crystal, the light polarizations are E || a, b, c.

Download Full Size | PDF

Er3+:MgWO4 is also attractive for in-band pumping (directly into the 4I13/2 manifold): the maximum σabs is 1.70×10−20 cm2 at 1482.9 nm and the absorption bandwidth is as broad as ∼10 nm (for E || b) and at longer wavelengths addressed by Er fiber lasers, σabs = 1.51×10−20 cm2 at 1528.0 nm with still broad absorption bandwidth of ∼11 nm (for E || c).

The 4fn transition intensities of Er3+ were analyzed using the Judd-Ofelt (J-O) formalism based on the measured absorption spectra. Both the standard J-O theory [23,24] and its modification accounting for an intermediate configuration interaction (ICI) with an excited configuration of the opposite parity 4fn-15d1 [25,26] were implemented to determine the electric dipole (ED) contributions. The set of squared reduced matrix elements U(k) (k = 2, 4, 6) for Er3+ was calculated in the present work based on the free-ion parameters reported in [27]. The magnetic dipole (MD) contributions to transition intensities (for transitions following the selection rule ΔJ = 0, ±1, except of 0 ↔ 0’) were calculated here within the Russell–Saunders approximation on wave functions of Er3+ under an assumption of a free-ion. The refractive index data from [13] were used. More details can be found elsewhere [28].

Table 2 presents the experimental <fexp> and calculated fcalc absorption oscillator strengths of Er3+ in MgWO4. The obtained intensity parameters are Ω2 = 11.111, Ω4 = 3.394, Ω6 = 0.598 [10−20 cm2] for the J-O theory and Ω2 = 11.480, Ω4 = 3.782, Ω6 = 0.703 [10−20 cm2], R2 = -0.154, R4 = 0.518, R6 = 0.064 [10−4 cm] for the ICI approximation. Note that these intensity parameters are derived from polarization-averaged absorption oscillator strengths. The ICI model provided lower root mean square (r.m.s.) deviation between <fexp> and fcalc values, as well as better agreement between the radiative and measured lifetimes of the lowest-lying excited-state, 4I13/2. Thus, it was selected for further calculations.

Tables Icon

Table 2. Judd-Ofelt Analysisa of Transitions in Absorption for Er3+ in MgWO4

The probabilities of spontaneous radiative transitions (AΣJJ’, where Σ indicates both ED + MD contributions), the mean luminescence wavelength <λ> calculated from the barycenter energies of the corresponding multiplets (cf. Table 2), the luminescence branching ratios BJJ’ and the radiative lifetimes τrad are listed in Table 3 (all the values were calculated within the ICI approximation). For the upper laser level 4I13/2, τrad = 4.78 ms.

Tables Icon

Table 3. Probabilitiesa of Spontaneous Radiative Transitions of Er3+ in MgWO4

3.2 Emission (spectra and lifetimes)

The stimulated-emission (SE) cross-sections for the 4I13/24I15/2 transition of Er3+ were calculated using two complementary methods: the Füchtbauer–Ladenburg (F-L) equation [29] and the reciprocity method (RM) [30]. The combined SE cross-section spectra are shown in Fig. 5(a). The maximum σSE is 1.42×10−20 cm2 at 1533.9 nm (for light polarization E || c) corresponding to the zero-phonon line (ZPL) transition at RT (see below). At the wavelengths exceeding ZPL corresponding to the expected Er3+ laser emission (see the gain spectra), σSE is lower, namely 0.26×10−20 cm2 at 1640 nm (E || a), 0.31×10−20 cm2 at 1637 nm (E || b) and 0.26×10−20 cm2 at 1633 nm (E || c).

 figure: Fig. 5.

Fig. 5. Stimulated-emission (SE) cross-sections, σSE, for the 4I13/24I15/2 Er3+ transition in MgWO4: (a) combined spectra for E || a, b, c; (b) a comparison of spectra calculated by the Füchtbauer – Ladenburg (F-L) equation and the reciprocity method (RM), for E || c.

Download Full Size | PDF

A relatively good agreement between the σSE values calculated by both methods is achieved for a radiative lifetime of the 4I13/2 state τrad = 4.85 ± 0.05 ms, Fig. 5(b). This value is in line with the J-O analysis. Lower SE cross-sections achieved by the F-L method at shorter wavelengths are due to the reabsorption affecting the measured luminescence spectra.

Erbium ions represent a quasi-three-level laser scheme with intrinsic reabsorption (the 4I13/24I15/2 transition). Gain cross-sections, σgain = βσSE – (1 – β)σabs, where β = N2(4I13/2)/NEr is the inversion ratio, are thus calculated to conclude about the expected laser wavelength and polarization. The calculated gain profiles for light polarizations E || a and E || c are shown in Fig. 6. For the high-gain polarization E || a, a local peak at 1640 nm dominates in the spectra. The gain bandwidth (FWHM) is ∼17 nm.

 figure: Fig. 6.

Fig. 6. RT gain cross-section, σgain, profiles for the 4I13/24I15/2 transition of Er3+ in MgWO4: the light polarization is (a) E || a and (b) E || c, β – inversion ratio.

Download Full Size | PDF

RT luminescence decay curves from the 4I13/2 and 4I11/2 Er3 + multiplets are shown in Fig. 7. They were measured using a powdered crystal sample to reduce the effect of radiation trapping (reabsorption). The luminescence decay is single exponential in agreement with a single type of sites for Er3+ ions in MgWO4 (C2 symmetry). The luminescence lifetimes τlum are 4.93 ms and 50.4 µs, respectively. The luminescence lifetime of the 4I13/2 metastable level is close to the radiative one obtained from the J-O calculations (4.78 ms) and from the evaluation of SE cross-sections (4.85 ms), indicating a luminescence quantum efficiency close to unity. The slightly longer value of τlum is probably due to residual reabsorption effect.

 figure: Fig. 7.

Fig. 7. RT luminescence decay curves from the 4I13/2 (a) and 4I11/2 (b) multiplets of Er3+ in MgWO4: circles – experimental data, lines – single-exponential fits.

Download Full Size | PDF

The relatively long lifetime of the upper laser level (for oxide crystals) makes Er3+:MgWO4 attractive for passively Q-switched lasers.

The luminescence lifetimes of five Er3+ excited states from 4I13/2 to 4S3/2 were measured to determine the rates of multiphonon non-radiative relaxation WNR = (1/τlum) – (1/τrad), cf. Table 4. The values of WNR were plotted vs. the energy gap between the emitting multiplet and the lower-lying manifold, ΔE, as shown in Fig. 8. The experimental points were fitted using the equation WNR = Ce-αΔE, where C and α are constants characteristic of the material [31,32]. C has the meaning of a rate constant at the limit of zero energy gap (ΔE → 0), and α = –ln(ε)/(ph), where ε is the ratio between the probabilities of m-phonon and m – 1-phonon relaxation and ph is the dominant (maximum) phonon energy of the host matrix. The experimental points in Fig. 8 are well fitted with the above-mentioned equation yielding the values of C= 6.5 ± 0.3×109 s-1 and α = 3.44 ± 0.1×10−3 cm. The multiphonon relaxation in MgWO4 is slightly stronger than that in the monoclinic KLu(WO4)2 crystal [31] owing to the high maximum phonon energy of the former material (ph = 916 cm-1 [11]).

 figure: Fig. 8.

Fig. 8. The rate of multiphonon non-radiative relaxation WNR vs. the energy gap to the lower-lying manifold ΔE for Er3+ in MgWO4: circles – data obtained from luminescence lifetime measurements, red solid line – their fit, blue dashed line – fit for KLu(WO4)2 [31].

Download Full Size | PDF

Tables Icon

Table 4. Evaluationa of Non-Radiative Relaxation Rates for Er3+ in MgWO4

3.3 Low-temperature spectroscopy

LT absorption and luminescence spectra were measured to determine the Stark splitting of the upper (4I13/2) and lower (4I15/2) laser multiplets of Er3+, cf. Fig. 9(a,b). The assignment of the electronic transitions followed previous work on Er3+-doped KLu(WO4)2 crystal [33]. For C2 symmetry sites, each 2S+1LJ multiplet with non-integer J is split into J + ½ Stark sub-levels. All the sub-levels are identified from the measured spectra leading to the energy-level scheme shown in Fig. 9(c). The partition functions for the lower and upper laser manifolds are Zl = 4.534 and Zu = 4.343, respectively, and their ratio Zl/Zu = 1.044 (these data were used for the calculation of SE cross-sections via RM).

 figure: Fig. 9.

Fig. 9. (a,b) LT (10 K) (a) absorption and (b) luminescence spectra of Er3+ in the MgWO4 crystal corresponding to the 4I15/24I13/2 transition, “+” - assigned electronic transitions, dashes – electronic transitions for Er3+ in KLu(WO4)2 crystal [33]; (c) experimental Stark splitting of the 4I15/2 and 4I13/2 Er3+ multiplets, numbers indicate the sub-level energies in cm-1.

Download Full Size | PDF

The energy of the ZPL transition occurring between the lowest Stark sub-levels of both multiplets EZPL is 6520 cm-1 (1533.7 nm). The total Stark splitting of the lower laser level (4I15/2) is 435 cm-1 which is larger than that for Er3+ in KLu(WO4)2, 361 cm-1 [33]. This is attributed to the stronger crystal field for distorted C2 sites in MgWO4. Large total Stark splitting of the ground-state explains the relatively long emission wavelengths (cf. Fig. 5) and is beneficial for low threshold laser operation.

4. Conclusion

To conclude, Er3+-doped MgWO4 is a promising gain material for eye-safe lasers emitting at ∼1.6 µm. It exhibits (i) relatively intense and broad absorption bands around ∼1.54 µm which is attractive for in-band pumping, (ii) high stimulated-emission cross-sections for polarized light potentially leading to naturally polarized laser emission, (iii) broad emission bands making it feasible to access laser wavelengths at ∼1.62 - 1.64 µm similarly to Er:YAG and (iv) relatively long upper laser level lifetime and high luminescence quantum efficiency which is attractive for low-threshold operation, as well as generation of giant pulses in Q-switched lasers. In addition, as shown in previous work, MgWO4 features good thermal properties. The observed broad, intense and polarized spectral bands for Er3+:MgWO4 are assigned to low-symmetry (C2) coordination of Er3+ ions replacing for the host-forming Mg2+ cations which is distorted by the heteroallene doping mechanism involving univalent Na+ cations, i.e., the difference in the charge and the ionic radii of Mg2+, Er3+ and Na+.

Funding

National Key Research and Development Program of China (2021YFB3601504); National Natural Science Foundation of China (61975208, 61905247, 61875199, U21A20508, 61850410533); Sino-German Scientist Cooperation and Exchanges Mobility Program (M-0040); Key-Area Research and Development Program of Guangdong Province (2020B090922006); Grant PID2019-108543RB-I00 funded by MCIN/AEI.

Acknowledgment

Xavier Mateos acknowledges the Serra Húnter program.

Disclosures

The authors declare no conflicts of interest.

Data availability

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

References

1. G. Karlsson, F. Laurell, J. Tellefsen, B. Denker, B. Galagan, V. Osiko, and S. Sverchkov, “Development and characterization of Yb-Er laser glass for high average power laser diode pumping,” Appl. Phys. B 75(1), 41–46 (2002). [CrossRef]  

2. G. Karlsson, V. Pasiskevicius, F. Laurell, J. A. Tellefsen, B. Denker, B. I. Galagan, V. V. Osiko, and S. Sverchkov, “Diode-pumped Er–Yb:glass laser passively Q-switched by use of Co2+:MgAl2O4 as a saturable absorber,” Appl. Opt. 39(33), 6188–6192 (2000). [CrossRef]  

3. V. Vitkin, P. Loiko, O. Dymshits, A. Zhilin, I. Alekseeva, D. Sabitova, A. Polishchuk, A. Malyarevich, X. Mateos, and K. Yumashev, “Passive Q-switching of an Er, Yb:glass laser with Co:Mg(Al,Ga)2O4-based glass-ceramics,” Appl. Opt. 56(8), 2142–2149 (2017). [CrossRef]  

4. N. A. Tolstik, A. E. Troshin, S. V. Kurilchik, V. E. Kisel, N. V. Kuleshov, V. N. Matrosov, T. A. Matrosova, and M. I. Kupchenko, “Spectroscopy, continuous-wave and Q-switched diode-pumped laser operation of Er3+, Yb3+:YVO4 crystal,” Appl. Phys. B 86(2), 275–278 (2007). [CrossRef]  

5. N. A. Tolstik, S. V. Kurilchik, V. E. Kisel, N. V. Kuleshov, V. V. Maltsev, O. V. Pilipenko, E. V. Koporulina, and N. I. Leonyuk, “Efficient 1 W continuous-wave diode-pumped Er,Yb:YAl3(BO3)4 laser,” Opt. Lett. 32(22), 3233–3235 (2007). [CrossRef]  

6. K. N. Gorbachenya, V. E. Kisel, A. S. Yasukevich, V. V. Maltsev, N. I. Leonyuk, and N. V. Kuleshov, “Highly efficient continuous-wave diode-pumped Er, Yb:GdAl3(BO3)4 laser,” Opt. Lett. 38(14), 2446–2448 (2013). [CrossRef]  

7. P. Wang, J. M. Dawes, P. Burns, J. A. Piper, H. Zhang, L. Zhu, and X. Meng, “Diode-pumped cw tunable Er3+:Yb3+:YCOB laser at 1.5–1.6 µm,” Opt. Mater. 19(3), 383–387 (2002). [CrossRef]  

8. T. Schweizer, T. Jensen, E. Heumann, and G. Huber, “Spectroscopic properties and diode pumped 1.6 µm laser performance in Yb-codoped Er:Y3Al5O12 and Er:Y2SiO5,” Opt. Commun. 118(5-6), 557–561 (1995). [CrossRef]  

9. D. Y. Shen, J. K. Sahu, and W. A. Clarkson, “Highly efficient in-band pumped Er:YAG laser with 60 W of output at 1645 nm,” Opt. Lett. 31(6), 754–756 (2006). [CrossRef]  

10. J. W. Kim, D. Y. Shen, J. K. Sahu, and W. A. Clarkson, “High-power in-band pumped Er:YAG laser at 1617 nm,” Opt. Express 16(8), 5807–5812 (2008). [CrossRef]  

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

12. P. Loiko, M. Chen, J. M. Serres, M. Aguiló, F. Díaz, H. Lin, G. Zhang, L. Zhang, Z. Lin, P. Camy, S.-B. Dai, Z. Chen, Y. Zhao, L. Wang, W. Chen, U. Griebner, V. Petrov, and X. Mateos, “Spectroscopy and high-power laser operation of a monoclinic Yb3+:MgWO4 crystal,” Opt. Lett. 45(7), 1770–1773 (2020). [CrossRef]  

13. L. Zhang, P. Loiko, J.M. Serres, E. Kifle, H. Lin, G. Zhang, E. Vilejshikova, E. Dunina, A. Kornienko, L. Fomicheva, U. Griebner, V. Petrov, Z. Lin, W. Chen, K. Subbotin, M. Aguiló, F. Díaz, and X. Mateos, “Growth, spectroscopy and first laser operation of monoclinic Ho3+:MgWO4 crystal,” J. Lumin. 213, 316–325 (2019). [CrossRef]  

14. L. Zhang, Y. Huang, S. Sun, F. Yuan, Z. Lin, and G. Wang, “Thermal and spectral characterization of Cr3+:MgWO4—a promising tunable laser material,” J. Lumin. 169, 161–164 (2016). [CrossRef]  

15. P. Loiko, L. Zhang, J.M. Serres, Y. Wang, M. Aguiló, F. Díaz, Z. Lin, H. Lin, G. Zhang, E. Vilejshikova, E. Dunina, A. Kornienko, L. Fomicheva, V. Petrov, U. Griebner, W. Chen, and X. Mateos, “Monoclinic Tm:MgWO4 crystal: Crystal-field analysis, tunable and vibronic laser demonstration,” J. Alloys Compd. 763, 581–591 (2018). [CrossRef]  

16. Y. Wang, W. Chen, M. Mero, L. Zhang, H. Lin, Z. Lin, G. Zhang, F. Rotermund, Y. J. Cho, P. Loiko, X. Mateos, U. Griebner, and V. Petrov, “Sub-100 fs Tm:MgWO4 laser at 2017nm mode locked by a graphene saturable absorber,” Opt. Lett. 42(16), 3076–3079 (2017). [CrossRef]  

17. L. Wang, W. Chen, Y. Zhao, Y. Wang, Z. Pan, H. Lin, G. Zhang, L. Zhang, Z. Lin, J. E. Bae, T. G. Park, F. Rotermund, P. Loiko, X. Mateos, M. Mero, U. Griebner, and V. Petrov, “Single-walled carbon-nanotube saturable absorber assisted Kerr-lens mode-locked Tm:MgWO4 laser,” Opt. Lett. 45(22), 6142–6145 (2020). [CrossRef]  

18. J. M. Serres, P. Loiko, V. Jambunathan, X. Mateos, V. Vitkin, A. Lucianetti, T. Mocek, M. Aguiló, F. Díaz, U. Griebner, and V. Petrov, “Efficient diode-pumped Er:KLu(WO4)2 laser at ∼1.61 µm,” Opt. Lett. 43(2), 218–221 (2018). [CrossRef]  

19. L. Zhang, W. Chen, J. Lu, H. Lin, L. Li, G. Wang, G. Zhang, and Z. Lin, “Characterization of growth, optical properties, and laser performance of monoclinic Yb:MgWO4 crystal,” Opt. Mater. Express 6(5), 1627–1634 (2016). [CrossRef]  

20. V. B. Mikhailik, H. Kraus, V. Kapustyanyk, M. Panasyuk, Y. Prots, V. Tsybulskyi, and L. Vasylechko, “Structure, luminescence and scintillation properties of the MgWO4-MgMoO4 system,” J. Phys.: Condens. Matter 20(36), 365219 (2008). [CrossRef]  

21. R. D. Shannon, “Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides,” Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 32(5), 751–767 (1976). [CrossRef]  

22. J. Tauc, “Optical properties and electronic structure of amorphous Ge and Si,” Mater. Res. Bull. 3(1), 37–46 (1968). [CrossRef]  

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

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

25. A. A. Kornienko, A. A. Kaminskii, and E. B. Dunina, “Dependence of the line strength of f–f transitions on the manifold energy. II. Analysis of Pr3+ in KPrP4O12,” Phys. Status Solidi B 157(1), 267–273 (1990). [CrossRef]  

26. P. Loiko, A. Volokitina, X. Mateos, E. Dunina, A. Kornienko, E. Vilejshikova, M. Aguiló, and F. Díaz, “Spectroscopy of Tb3+ ions in monoclinic KLu(WO4)2 crystal: Application of an intermediate configuration interaction theory,” Opt. Mater. 78, 495–501 (2018). [CrossRef]  

27. J. B. Gruber, J. R. Quagliano, M. F. Reid, F. S. Richardson, M. E. Hills, M. D. Seltzer, S. B. Stevens, C. D. Morrison, and T. H. Allik, “Energy levels and correlation crystal-field effects in Er3+-doped garnets,” Phys. Rev. B 48(21), 15561–15573 (1993). [CrossRef]  

28. P. A. Loiko, E. V. Vilejshikova, N. M. Khaidukov, M. N. Brekhovskikh, X. Mateos, M. Aguiló, and K. V. Yumashev, “Judd–Ofelt modeling, stimulated-emission cross-sections and non-radiative relaxation in Er3+:K2YF5 crystals,” J. Lumin. 180, 103–110 (2016). [CrossRef]  

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

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

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

32. M. J. Weber, “Radiative and multiphonon relaxation of rare-earth ions in Y2O3,” Phys. Rev. 171(2), 283–291 (1968). [CrossRef]  

33. S. Bjurshagen, P. Brynolfsson, V. Pasiskevicius, I. Parreu, M. C. Pujol, A. Pena, M. Aguilo, and F. Diaz, “Crystal growth, spectroscopic characterization, and eye-safe laser operation of erbium-and ytterbium-codoped KLu(WO4)2,” Appl. Opt. 47(5), 656–665 (2008). [CrossRef]  

Data availability

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

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (9)

Fig. 1.
Fig. 1. A photograph of the as-grown 1.01 at.% Er3+:MgWO4 crystal, the growth direction is along the [010] axis.
Fig. 2.
Fig. 2. Structural study of Er3+:MgWO4: (a) X-ray powder diffraction (XRD) analysis: observed (black), calculated (red) and residual (blue) patterns, dashes – Bragg positions, (hkl) – Miller’s indices; (b,c) projection of the crystal structure on (b) the a-c plane and (c) the b-c plane, black lines – unit-cell.
Fig. 3.
Fig. 3. Electronic structure of MgWO4: (a) calculated band structure, Eg – bandgap energy; (b) calculated total and partial densities of states (TDOS and PDOS).
Fig. 4.
Fig. 4. (a-f) RT (293 K) absorption cross-section, σabs, spectra of Er3+:MgWO4 crystal, the light polarizations are E || a, b, c.
Fig. 5.
Fig. 5. Stimulated-emission (SE) cross-sections, σSE, for the 4I13/24I15/2 Er3+ transition in MgWO4: (a) combined spectra for E || a, b, c; (b) a comparison of spectra calculated by the Füchtbauer – Ladenburg (F-L) equation and the reciprocity method (RM), for E || c.
Fig. 6.
Fig. 6. RT gain cross-section, σgain, profiles for the 4I13/24I15/2 transition of Er3+ in MgWO4: the light polarization is (a) E || a and (b) E || c, β – inversion ratio.
Fig. 7.
Fig. 7. RT luminescence decay curves from the 4I13/2 (a) and 4I11/2 (b) multiplets of Er3+ in MgWO4: circles – experimental data, lines – single-exponential fits.
Fig. 8.
Fig. 8. The rate of multiphonon non-radiative relaxation WNR vs. the energy gap to the lower-lying manifold ΔE for Er3+ in MgWO4: circles – data obtained from luminescence lifetime measurements, red solid line – their fit, blue dashed line – fit for KLu(WO4)2 [31].
Fig. 9.
Fig. 9. (a,b) LT (10 K) (a) absorption and (b) luminescence spectra of Er3+ in the MgWO4 crystal corresponding to the 4I15/24I13/2 transition, “+” - assigned electronic transitions, dashes – electronic transitions for Er3+ in KLu(WO4)2 crystal [33]; (c) experimental Stark splitting of the 4I15/2 and 4I13/2 Er3+ multiplets, numbers indicate the sub-level energies in cm-1.

Tables (4)

Tables Icon

Table 1. Fractional Atomic Coordinates (x, y, z), Site Occupancy Factors (O.F.) and Isotropic Displacement Parameters Biso2) for Er3+:MgWO4

Tables Icon

Table 2. Judd-Ofelt Analysisa of Transitions in Absorption for Er3+ in MgWO4

Tables Icon

Table 3. Probabilitiesa of Spontaneous Radiative Transitions of Er3+ in MgWO4

Tables Icon

Table 4. Evaluationa of Non-Radiative Relaxation Rates for Er3+ in MgWO4

Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.