Physico-chemical and optical properties of Er3+-doped and Er3+/Yb3+-co-doped Ge25Ga9.5Sb0.5S65 chalcogenide glass
-
Dianna Himics
, Lukas Strizik
, Jana Holubova , Ludvik Benes , Karel Palka , Bozena Frumarova , Jiri Oswald , Andrey S. Tverjanovich and Tomas Wagner
Abstract
We investigated the physicochemicаl properties, structure and optical properties of the Ge25Ga9.5Sb0.5S65: Er3+/Yb3+ glasses. The Judd-Ofelt theory was used to calculate the intensities of the intra-4f electronic transitions of Er3+ ions. We observed the upconversion photoluminescence (UCPL) at 530, 550, 660 and 810 nm under 980 nm excitation. In the Ge25Ga9.5Sb0.5S65: 0.1 at.% Er3+, we found that the Stokes photoluminescence (PL) at the green spectral region excited by the 490 and 532 nm laser is only ≈5 times higher than the UCPL emission under 810 or 980 nm excitation making these materials attractive for UCPL applications. The addition of 0.1–1 at.% of Yb3+ into Ge25Ga9.5Sb0.5S65: 0.1 at.% Er3+ glass reduces the UCPL as well as the Er3+ ≈1.5 μm emission intensity probably due to the reabsorption processes of the excitation light and concentration quenching. However, the observed Er3+: 4S3/2→4I13/2 (≈850 nm) emission in the Ge25Ga9.5Sb0.5S65: 0.1 at.% Er3+ sample populates the 4I13/2 level, which promises the using of this material for the 1.5 μm optical amplification.
Introduction
Rareearthdoped chalcogenide glasses exhibit many interesting properties therefore, they are promising materials in variety of applications. Chalcogenide glasses (ChGs) possess low phonon energy and high refractive index [1], [2], [3] which suppresses the multiphonon relaxation and promotes radiative recombination, respectively [4]. Moreover, ChGs are transparent from visible to mid-infrared spectral region and show large intra-4f cross sections when they are doped with rare-earth (RE) ions [5]. RE3+-doped ChGs can be applied such as optical amplifiers [6], waveguides [7], displays [8], sensors and detectors [9], [10], [11], lasers [12], [13].
The Er3+: 4I13/2→4I15/2 (λ≈1.5 μm) near-infrared emission has attracted the attention in telecommunication C-band [14], [15]. To improve the efficiency of the ≈1.5 μm emission, the Er3+ ions can be co-doped with the Yb3+ ions, where Yb3+ ions play role of the sensitizer at pumping wavelength of ≈980 nm and allow the energy transfer to neighboring Er3+ ions. It was demonstrated that the Er3+/Yb3+ co-doping has become an effective method for production of short amplifiers and efficient lasers in the long haul telecommunications [16], [17]. The efficient Yb3+→Er3+ energy transfer was first investigated by Snitzer and Woodcock in Na-K-Ba silicate glasses [17]. Recently, Er3+/Yb3+ -co-doped fiber lasers [17], [18] and planar waveguide amplifiers [14], [16], [19] have been demonstrated.
In this paper, we have chosen the Ga-Ge-Sb-S chalcogenide glasses as the promising host matrix for the RE3+ ions [20], [21]. The addition of Sb into Ge-Ga-S glasses improves the resistance against crystallization and resistance to moisture with still large solubility of the RE3+ ions due to the presence of Ga [20], [22]. Reference [21] reports the effect of compositional changes in Ge25Ga10−x SbxS65: 0.5 at.% Er3+ on the upconversion photoluminescence (UCPL). In the present study, we aim on the UCPL and ≈1.5 μm photoluminescence (PL) emission in (Ge0.25Ga0.095Sb0.005S0.65)99.9−x Er0.1Ybx, where x=0, 0.1, 0.5 and 1.0 at.% and (Ge0.25Ga0.095Sb0.005S0.65)99.85Er0.05Yb0.1 (further denoted as Ga-Ge-Sb-S: Er3+/Yb3+) chalcogenide glasses with respect to their optical properties and structure.
Experimental
The (Ge0.25Ga0.095Sb0.005S0.65)99.9−x Er0.1Ybx (x=0, 0.1, 0.5, 1) and (Ge0.25Ga0.095Sb0.005S0.65)99.85Er0.05Yb0.1 ChGs were synthesized from high-purity elements of Ge (5N), Ga (5N), Sb (5N), S (4.5N), Er (3N) and Yb (3N). Elements were weighted into silica ampoules, which were sealed at residual pressure of ~10−3 Pa. The total weight of batch was 10 g. The sealed ampoules were heated at 1270 K for 24 h in a rocking furnace. The melt of the glass was quenched into water and the glass was annealed at 20 K below the glass transition temperature Tg for 2 h to relax the mechanical strain. Finally, the prepared samples were cut into discs with diameter of ≈10 mm and thickness ≈2–4 mm and polished into optical quality.
The non-crystalline nature of prepared glasses has been confirmed by X-ray diffraction (XRD) analysis using the Bruker D8 advance powder XRD diffractometer. The XRD investigation with Cu Kα radiation was carried out on the as-prepared and annealed glasses in the 2θ range of 5–65° with a step of 0.02°. The compositions of prepared glasses were characterized by the energy dispersive X-ray (EDX) microanalyzer Aztec X-Max 20, Oxford Instruments at accelerating voltage of 20 kV. Differential scanning calorimeter DSC Diamond, Perkin-Elmer was employed to investigate thermal properties of the studied glasses in the range of 570–870 K and at heating rate of 10 K min−1. A piece of glass about ≈10 mg was sealed into an aluminum pan for the measurement. The glass transition temperature (Tg) was determined as half-height between extrapolated onset and endset, the crystallization temperature (Tc) as onset of crystallization peak.
The Raman spectra were measured under the Nd:YAG laser (λ=1064 nm) excitation at room temperature using the FT-IR spectrophotometer IFS 55 with the Raman FRA-106 accessory (Bruker) for back scattering geometry. Raman spectra were reduced by the Gammon-Shuker approximation [23] and decomposed into individual bands by the pseudo-Voigt functions.
Archimedes method was used for determination of the density of prepared glasses. Refractive index of glasses was determined by the variable angle spectroscopic ellipsometry (VASE, J.A. Woollam Co., Inc.) measured in the spectral region of 500–2300 nm with a spectral step of 20 nm and at angles of light incidence 65°, 70°, 75°. The ellipsometric data were parameterized by the Sellmeier model [24] in transparent spectral region of studied materials with obtained fit accuracy given by the mean square errors <1.5. Er3+ and Yb3+ absorption cross sections spectra were determined using the double-beam UV-Vis-NIR spectrophotometer (JASCO V-570) in the spectral region of 300–2500 nm with a spectral step of 2 nm. Intensities of the Er3+ intra-4f electronic transitions in the Ge25Ga9.5Sb0.5S65 host matrix were calculated by the Judd-Ofelt theory [25], [26], [27], [28] using the Er3+ absorption bands centered at 1538, 989, 815, 661, 545, and 522 nm and JOF program v. 2.3 [25], [26], [29]. The areas of absorption bands were fitted by Gausssians in Fityk program v. 0.9.8 [30]. The Er3+ 1.5 μm PL emission spectra of the Ga-Ge-Sb-S: Er3+/Yb3+ ChGs were measured in the spectral region of 1440–1650 nm with a 0.5 nm step and excited by the 980 nm diode laser (power density ≈80 W cm−2). PL signal was processed with a 1/2 m double grating monochromator and amplified by preamplifier and lock-in amplifier at chopping frequency of 30 Hz. The PL signal was detected under 90° by the two-step-cooled Ge detector. PL and UCPL emission spectra measured in the spectral region of 500–900 nm with a 0.5 nm step were acquired at pumping wavelengths of 490, 532, 810 and 980 nm using the Ti:sapphire laser (power density ≈3800 W cm−2) pumped with Nd:YVO4. The PL signal was processed through 1/8 m monochromator and detected by the GaAs photomultiplier tube. All photoluminescence measurements were carried out at room temperature.
Results and discussion
All measured XRD patterns presented in Fig. 1 confirmed the amorphous state of the prepared Ga-Ge-Sb-S: Er3+/Yb3+ ChGs.

XRD patterns for Ga-Ge-Sb-S: Er3+/Yb3+ chalcogenide glasses. (color online).
The chemical composition of (Ge0.25Ga0.095Sb0.005S0.65)99.9−x Er0.1Ybx (x=0, 0.1, 0.5, 1) and (Ge0.25Ga0.095Sb0.005 S0.65)99.85Er0.05Yb0.1 ChGs determined by the EDX spectroscopy is shown in Table 1. The observed results show a small deviation between experimental and theoretical chemical composition.
Chemical composition of Ge25Ga9.5Sb0.5S65 chalcogenide glasses doped with Er3+ or Er3+/Yb3+ ions. The error of EDX spectroscopy measurements is ±1 at.%.
| Sample | Chemical composition | |||||
|---|---|---|---|---|---|---|
| Ge (at.%) | Ga (at.%) | Sb (at.%) | S (at.%) | Er (at.%) | Yb (at.%) | |
| Er0.1 | 25.(0) | 7.(9) | 0.(5) | 66.(5) | – | – |
| Er0.1Yb0.1 | 23.(7) | 10.(2) | 0.(5) | 65.(6) | – | – |
| Er0.1Yb0.5 | 22.(3) | 11.(4) | 0.(6) | 65.(0) | – | 0.(7) |
| Er0.1Yb1 | 24.(1) | 9.(3) | 0.(5) | 65.(3) | – | 0.(9) |
| Er0.05Yb0.1 | 24.(2) | 9.(2) | 0.(5) | 66.(1) | – | – |
The temperature difference ∆T between the crystallization temperature Tc≈837 K and the glass transition temperature Tg≈713 K is >100 K suggesting the good thermal stability of synthesized glass. By the thermal stability is meant a broader temperature range at which the glass can be processed such as for fibers drawing.
Densities of prepared ChGs determined by the Archimedes method are shown in the Table 2. Densities were further used to calculate the concentration of Er3+ and Yb3+ ions per unit volume in Ge25Ga9.5Sb0.5S65 glass (see Table 2). The density of Ga-Ge-Sb-S ChGs increases with increasing Yb and Er content [31].
Physical properties of (Ge0.25Ga0.095Sb0.005S0.65)99.9−x Er0.1Ybx (x=0, 0.1, 0.5, 1) and (Ge0.25Ga0.095Sb0.005S0.65)99.85Er0.05Yb0.1 chalcogenide glasses.
| Sample | Er0.1 | Er0.1Yb0.1 | Er0.1Yb0.5 | Er0.1Yb1 | Er0.05Yb0.1 |
|---|---|---|---|---|---|
| Density (g cm−3) | 2.909±0.004 | 2.943±0.006 | 2.981±0.007 | 3.014±0.006 | 2.914±0.007 |
| Er3+ ions concentration (×1019 cm−3) | 3.822±0.006 | 3.837±0.008 | 3.822±0.009 | 3.836±0.008 | 1.912±0.004 |
| Yb3+ ions concentration (×1019 cm−3) | – | 3.842±0.008 | 19.222±0.004 | 38.391±0.008 | 3.822±0.008 |
| Refractive index n (at 663 nm) | 2.09±0.01 | 2.09±0.01 | 2.10±0.01 | 2.13±0.01 | 2.09±0.01 |
Moreover, the refractive index determined by the VASE slightly increases with increasing concentration of Yb3+ ions (Table 2).
The structure of the Ge25Ga9.5Sb0.5S65 glass doped with 0.1 at.% Er was investigated by Raman spectroscopy, as shown in Fig. 2. The spectrum was decomposed into several bands. The main and the most intense Raman band near 340 cm−1 can be assigned to a ν1 vibrational mode of the corner-sharing GeS4/2 and GaS4/2 tetrahedra. The small band at 265 cm−1 can be associated with vibrations of the metal–metal bonds in S3Ge(Ga)-Ge(Ga)S3 structural units. The band with maximum near of 375 cm−1 can be assigned to the ν1 mode of two edge shared tetrahedra Ge2S4S2/2 and Ga2S4S2/2. The last two bands located at 400 and 435 cm−1 correspond to the ν3 modes of the corner-sharing and edge- sharing GeS4/2 and GaS4/2 tetrahedra. Broad band in the region of 80–230 cm−1 does not show any fine structure and thus, it is difficult to decompose it into individual bands. However, the ν2 and the ν4 vibrational modes of the corner-sharing GeS4/2 and GaS4/2 originate at this spectral region. We did not find any band which can be associated with vibrational modes of Sb-based structural units, probably due to the low concentration of Sb (<1 at.%) in the studied glasses.

Reduced Raman spectrum of the (Ge0.25Ga0.095Sb0.005S0.65)99.9Er0.1 glass. (color online).
Absorption coefficients of the Er3+-doped and the Er3+/Yb3+-co-doped Ge25Ga9.5Sb0.5S65 glasses are shown in Fig. 3. We observed the ground state absorption (GSA) bands of Er3+ centered approximately at 1538, 979, 815, 661, 545, and 522 nm. One strong GSA band at 992 nm originates from Yb3+: 2F7/2→2F5/2 transitions and this band overlaps with the Er3+: 4I15/2→4I11/2 GSA band at 979 nm. In the Fig. 3 is evident a red shift of the absorption edge of the Ge25Ga9.5Sb0.5S65 host matrix with increasing Yb3+ content.

The absorption coefficient of the Ge25Ga9.5Sb0.5S65: Er3+/Yb3+ ChGs with labeled Er3+ and Yb3+ GSA transitions from respective levels 4I15/2 and 2F7/2 to the upper manifolds. (color online).
The Judd-Ofelt phenomenological parameters for the Er3+ ions embedded in the Ge25Ga9.5Sb0.5S65 glassy host were found to be Ω2=(12.58±1.10)×10−20 cm2, Ω4=(3.25±1.28)×10−20 cm2 and Ω6=(2.05±0.45)×10−20 cm2 which are comparable to similar chalcogenide glasses [21]. Root-mean-square (RMS) deviation between the theoretical and the experimental line strengths was 0.63×10−20 cm2. The large value of the Ω2 parameter can be related to a high degree of the covalent bonding around the Er3+ ions and in some extent to the asymmetry. The spectroscopic quality factor Ω4/Ω6=1.59 can be used as a prediction for the stimulated emission ability for such intra-4f transitions, where the first terms of doubly reduced matrix elements are negligible or zero. The calculated radiative transition probabilities, branching ratios and radiative lifetimes of selected Er3+ transitions are presented in Table 3.
Spontaneous electric dipole AEDand magnetic dipole AMDemission probabilities, branching ratios β and radiative lifetimes τof the (Ge0.25Ga0.095Sb0.005S0.65)99.9Er0.1 calculated by the Judd-Ofelt theory.
| Transition | λ (nm) | A ED (s−1) | A MD (s−1) | β (%) | τ (ms) |
|---|---|---|---|---|---|
| 2 H 11/2→4I15/2 | 533 | 49255.0 | 0 | 96 | 0.020 |
| 4 S 3/2→4I13/2 | 859 | 2055.7 | 0 | 26 | 0.404 |
| 4 S 3/2→4I15/2 | 556 | 5500.3 | 0 | 69 | 0.125 |
| 4 F 9/2→4I15/2 | 670 | 6926.6 | 0 | 91 | 0.131 |
| 4 I 9/2→4I15/2 | 823 | 771.8 | 0 | 79 | 1.018 |
| 4 I 11/2→4I15/2 | 1002 | 693.4 | 0 | 86 | 1.246 |
| 4 I 13/2→4I15/2 | 1580 | 448.2 | 85.6 | 100 | 1.873 |
The proposed UCPL mechanism in the Er3+/Yb3+-co-doped samples under 980 nm excitation wavelength is schematically drawn in the Fig. 4. The 980 nm wavelength excites the Er3+ and Yb3+ by the ground state absorption (GSA) into Er3+: 4I11/2 and Yb3+: 2F5/2 levels, respectively. Subsequently, the Er3+: 4F7/2 manifold can be populated: (1) by the excited state absorption (ESA) 4I11/2→4F7/2 within Er3+ ion, or (2) by the energy transfer (ETU1) between Er3+ and Yb3+ ions as Er3+: 4I11/2, Yb3+: 2F5/2→Er3+: 4F7/2, Yb3+: 2F7/2. The thermalized levels 2H11/2 and 4S3/2 responsible for the green UCPL will be populated from 4F7/2 in particular by the multiphonon relaxation due to their small energy difference. On the other hand, the energy transfer (ETU2) Er3+: 4I13/2, Yb3+: 2F5/2→Er3+: 4F9/2, Yb3+: 2F7/2 can be assumed as well promoting the Er3+ red UCPL 4F9/2→4I15/2. It should be noted that the ETU1 and ETU2 processes can originate between neighboring Er3+ ions (without Yb3+) due to similar energy positions of the Er3+: 4I11/2 and Yb3+: 2F5/2. Revealing the UCPL dynamics is a difficult task and among the ChGs was recently studied in Ge-Ga-S: Er3+ ChGs [32], [33]. Moreover, experimentally observed Er3+: 4S3/2→4I13/2 (≈850 nm) UCPL emission populating the 4I13/2 level is promising for the Er3+: 4I13/2→4I15/2 (≈1.5 μm) optical amplification.

Schematic energy level diagram for the Er3+/Yb3+ -co-doped Ge25Ga9.5Sb0.5S65 glass with observed emissions and proposed mechanisms standing behind the UCPL emissions. (color online).
The effect of Yb3+ addition to sensitize the UCPL emission was investigated for samples (Ge0.25Ga0.095 Sb0.005S0.65)99.9−x Er0.1Ybx (x=0, 0.1, 0.5, 1) and (Ge0.25Ga0.095Sb0. 005S0.65)99.85Er0.05Yb0.1 at pumping wavelength of 980 nm. The UCPL emission spectra are presented in the Fig. 5 from the visible to the near-infrared spectral region. The addition of Yb3+ ions lowers the total UCPL emission intensity and promotes the red-to-green UCPL intensity ratio. This is probably because of a red shift of the absorption edge of glassy host matrix with Yb addition (Fig. 3) which promotes the absorption of the excitation light (due to the ESA, ETU processes) by host matrix. Thus, the co-doping of the Ge25Ga9.5Sb0.5S65: 0.1 at.% Er3+ ChGs by the Yb3+ ions is inefficient to improve the UCPL emission intensity.

UCPL emission spectra of the (Ge0.25Ga0.095Sb0.005S0.65)99.9Er0.1 and the (Ge0.25Ga0.095Sb0.005S0.65)99.8Er0.1Yb0.1 glasses at pumping wavelength of 980 nm laser. (color online).
Stokes and anti-Stokes (UCPL) PL spectra of the (Ge0.25Ga0.095Sb0.005S0.65)99.9Er0.1 glass at pumping wavelengths of 490, 532, 810 and 980 nm are presented in Fig. 6. The observed emission bands at 529, 551, 660, 808, and 850 nm can be assigned to the Er3+: 2H11/2→4I15/2, 4S3/2→4I15/2, 4F9/2→4I15/2, 4I9/2→4I15/2 and 4S3/2→4I13/2 electronic transitions, respectively. The PL emission intensities at 550 nm originating from 4S3/2→4I15/2 transitions are similar for excitation wavelength of 490 and 532 nm which directly excite the upper 4F7/2 and 2H11/2 levels. From these manifolds the 4S3/2 level is populated by the multiphonon relaxation and thus, the emission characteristics are similar. On the other hand, the emissions at excitation wavelengths of 810 and 980 nm originate from the nonlinear UCPL processes and thus, they have lower intensity (approximately ≈5 times). However, the UCPL intensity is still sufficiently high, which makes these materials promising for the UCPL applications.

Emission spectra of the (Ge0.25Ga0.095Sb0.005S0.65)99.9Er0.1 glass measured at various excitation wavelengths λex. (color online).
In addition, we investigated the ≈1.5 μm PL emission originating from Er3+: 4I13/2→4I15/2 transitions at 980 nm pumping wavelength against Yb3+ content which is depicted in the Fig. 7. The PL emission intensity decreases with increasing Yb3+ content. This is probably due to the concentration quenching and/or due to UCPL processes depopulating the 4I13/2 energy level. Thus, the present study demonstrates that the Yb3+ addition into Ge25Ga9.5Sb0.5S65: Er3+ ChGs does not lead to appreciable sensitization of the UCPL as well as 1.5 μm PL emissions. However, we believe that this drawback can be overcome by using the host material with larger optical band gap energy. Such promising materials can be chalcohalide or oxychalcogenide glasses where the trade-off between intra-4f electronic intensities, solubility of Er3+/Yb3+ ions and optical band gap energy might be found.

(a) The PL Er3+: 4I13/2→4I15/2 emission spectra and (b) the same normalized spectra of the Ge25Ga9.5Sb0.5S65: 0.1 or 0.05 at.% Er3+ ChGs co-doped with various concentrations of Yb3+ ions under 980 nm excitation. (color online).
Conclusions
Er3+-doped and Er3+/Yb3+-co-doped Ge25Ga9.5Sb0.5S65 thermally stable chalcogenide glasses were synthesized by the melt-quenching technique. The Judd-Ofelt theory was used for calculation of the Er3+ intra-4f electronic transitions intensities. The Ge25Ga9.5Sb0.5S65 glass doped with 0.1 at.% Er3+ shows intense upconversion photoluminescence (UCPL) from green to near-infrared spectral region. The green UCPL emission is approximately 5 times lower than the green Stokes photoluminescence (PL) Stokes emission intensity, which makes this material attractive for the UCPL applications. However, it is presented that the Yb3+ ions sensitization is ineffective resulting in the decrease of the total UCPL emission intensity under 980 nm laser excitation. This is attributed to a red shift of the absorption edge of host matrix with Yb addition, which is merged with the upper Er3+ energy level manifolds 4F7/2, 2H11/2 and 4S3/2. Moreover, the ≈1.5 μm PL emission intensity decreases with the increasing Yb content probably due to the concentration quenching. Thus, the addition of Yb3+ ions was shown to be inefficient to improve the UCPL and ≈1.5 μm PL, which could be overcome by the using of chalcohalide or oxychalcogenide glasses as host matrices. Thereafter, the observed Er3+: 4S3/2→4I13/2 (≈850 nm) emission could be utilized for the optical amplification at ≈1.5 μm (4I13/2→4I15/2) via upconversion processes.
Article note
A collection of invited papers based on presentations at the 12th Conference on Solid State Chemistry (SSC-2016), Prague, Czech Republic, 18–23 September 2016.
Acknowledgments
Authors acknowledge the financial support from Grant Agency of Czech Science Foundation of the Czech Republic, project no. 15-07912S. This work was supported by the project CZ.1.05/4.1.00/11.0251 and by grant LM2015082 Center of Materials and Nanotechnologies from the Czech Ministry of Education, Youth and Sports of the Czech Republic.
References
[1] N. F. Mott, E. A. Davis. Electronic Processes in Non-Crystalline Materials, Clarendon, Oxford (1979).Search in Google Scholar
[2] B. Bureau, X. H. Zhang, F. Smektala, J.-L. Adam, J. Troles, H. Ma, C. Boussard- Plèdel, J. Lukcas, P. Lucas, D. Le Coq, M. R. Riley, J. H. Simmons. J. Non Cryst. Solid. 276, 345 (2004).10.1016/j.jnoncrysol.2004.08.096Search in Google Scholar
[3] T. Wagner, M. Frumar. “Optically-induced diffusion and dissolution of metals in amorphous chalcogenides”, in Photo-Induced Metastability in Amorphous Semiconductors, A. V. Kolobov (Ed.), pp. 160–177, Wiley-VCH Verlag, Berlin (2003).10.1002/9783527602544.ch10Search in Google Scholar
[4] V. G. Truong, B. S. Ham, A. M. Jurduc, B. Jacquier, J. Leperson, V. Nazabal, J. L. Adam. Phys. Rev. B. 74, 184103 (2006).10.1103/PhysRevB.74.184103Search in Google Scholar
[5] L. B. Shaw, B. Cole, P. A. Thielen, J. S. Sanghera, I. D. Aggarwal. IEEE J Quantum Elec.48, 1127 (2001).10.1109/3.945317Search in Google Scholar
[6] F. Prudenzano, L. Mescia, L. Allegretti, M. DeSario, F. Smektala, V. Mozain, V. Nazabal, J. Troles, J. L. Doualan, G. Canat, J. L. Adam, B. Boulard. Opt. Mater.31, 1292 (2009).10.1016/j.optmat.2008.10.004Search in Google Scholar
[7] A. V. Rode, A. Zakery, M. Samoc, R. B. Charters, E. G. Gamaly, B. Luther-Davies. Appl. Surf. Sci.481, 197 (2002).10.1016/S0169-4332(02)00375-6Search in Google Scholar
[8] N. G. Boetti, J. Lousteau, D. Negro, E. Mura, G. Scarpignato, S. Abrate, D. Milanese. Opt. Express. 20, 5409 (2012).10.1364/OE.20.005409Search in Google Scholar PubMed
[9] F. Charpentier, B. Bureau, J. Troles, C. Boussard-Plédel, K. Michel-LePierres, F. Smektala, J.-L. Adam. Opt. Mater.31, 496 (2009).10.1016/j.optmat.2007.10.014Search in Google Scholar
[10] W. Length, R. M. McFarlane. Opt. Photonics News. 3, 8 (1992).10.1364/OPN.3.3.000008Search in Google Scholar
[11] J. LePerson, F. Colas, C. Compere, M. Lehaitre, M.-L. Anne, C. Boussard-Plédel, B. Bureau, J.-L. Adam, S. Deputier, M. Guilloux-Viry. Sens. Actuators. B130, 771 (2008).10.1016/j.snb.2007.10.067Search in Google Scholar
[12] F. Prudenzano, L. Mescia, L. Allegretti, V. Mozain, V. Nazabal, F. Smektala. Opt. Mater.33, 241 (2010).10.1016/j.optmat.2010.10.046Search in Google Scholar
[13] F. Prudenzano, L. Mescia, L. A. Allegretti, M. De Sario, T. Palmisano, F. Smektala, V. Mozain, V. Nazabal, J. Troles. J. Non Cryst. Solids. 355, 1145 (2009).10.1016/j.jnoncrysol.2009.01.051Search in Google Scholar
[14] F. Rivera-Lopez, P. Babu, L. Jyothi, U. R. Rodriguez-Mendoza, I. R. Martin, C. K. Jayasankar, V. Lavin. Opt. Mater.34, 1235 (2012).10.1016/j.optmat.2012.01.017Search in Google Scholar
[15] J. D. B. Bradely, M. Pollanu. Lasers Photon.5, 368 (2011).10.1002/lpor.201000015Search in Google Scholar
[16] M. Federighi, F. Di Pasquale. IEEE Photonic Tech. L.7, 303 (1995).10.1109/68.372753Search in Google Scholar
[17] E. Snitzer, R. Woodcock. Appl. Phys. Lett.6, 45 (1965).10.1063/1.1754157Search in Google Scholar
[18] B. S. Reddy, H.-Y. Hwang, Y.-D. Jho, B. S. Ham, S. Sailaja, C. M. Reddy, B. V. Rao, S. J. Dhoble. Ceram. Int.41, 3684 (2015).10.1016/j.ceramint.2014.11.040Search in Google Scholar
[19] J.-M. P. Delavaux, S. Granlund, O. Mizuhara, L. D. Tzeng, D. Barbier, M. Rattay, F. Saint André, A. Kevorkian. IEEE Photonic Tech. L.9, 247 (1997).10.1109/68.553108Search in Google Scholar
[20] Y. Guimond, J.-L. Adam, A.-M. Jurdyc, H. L. Ma, J. Mugnier, B. Jacquier. J. Non Cryst. Solids. 256, 257, 378 (1999).10.1016/S0022-3093(99)00477-9Search in Google Scholar
[21] L. Strizik, J. Zhang, T. Wagner, J. Oswald, T. Kohoutek, B. M. Walsh, J. Prikryl, R. Svoboda, C. Liu, B. Frumarova, M. Frumar, M. Pavlista, W. J. Park, J. Heo. J. Lumin.147, 209 (2014).10.1016/j.jlumin.2013.11.021Search in Google Scholar
[22] J. Heo, J. M. Yoon, S. Y. Ryou. J. Non Cryst. Solids. 238, 115 (1998).10.1016/S0022-3093(98)00577-8Search in Google Scholar
[23] R. Shuker, R. W. Gammon. Phys. Rev. Lett. 25, 222 (1970).10.1103/PhysRevLett.25.222Search in Google Scholar
[24] W. Sellmeier. Ann. Phys. Chem.143, 271 (1871).Search in Google Scholar
[25] B. M. Walsh. “Judd-Ofelt Theory: principles and practices”, in Advances in Spectroscopy for Lasers and Sensing, B. Di Bartolo, O. Forte (Eds.), pp. 403–433, Springer, Netherlands (2006).10.1007/1-4020-4789-4_21Search in Google Scholar
[26] P. Goldner, F. Auzel. J. Appl. Phys.79, 7972 (1996).10.1063/1.362347Search in Google Scholar
[27] B. R. Judd. Phys. Rev.127, 750 (1962).10.1103/PhysRev.127.750Search in Google Scholar
[28] G. S. Ofelt. J. Chem. Phys.37, 511 (1962).10.1063/1.1701366Search in Google Scholar
[29] W. T. Carnall, P. R. Fields, B. G. Wybourne. J. Chem. Phys.42, 3797 (1965).10.1063/1.1695840Search in Google Scholar
[30] M. Wojdyr. J. Appl. Crystallogr.43, 1126 (2010).10.1107/S0021889810030499Search in Google Scholar
[31] S. W. Yung, H. J. Lin, Y. Y Lin, R. K. Brow, Y. S. Lai, J. S. Horng, T. Zhang. Mater. Chem. Phys.117, 29 (2009).10.1016/j.matchemphys.2008.11.060Search in Google Scholar
[32] L. Strizik, V. Prokop, J. Hrabovsky, T. Wagner, T. Aoki. J. Mater. Sci. Mater. El. (2017). https://link.springer.com/article/10.1007/s10854-016-6306-3.Search in Google Scholar
[33] T. Aoki, L. Strizik, J. Hrabovsky, T. Wagner. J. Mater. Sci-Mater. El. (2017). https://link.springer.com/article/10.1007/s10854-017-6363-2.Search in Google Scholar
©2017 IUPAC & De Gruyter. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. For more information, please visit: http://creativecommons.org/licenses/by-nc-nd/4.0/
Articles in the same Issue
- Frontmatter
- Graphical abstracts
- In this issue
- Preface
- 12th Conference on Solid State Chemistry (SSC-2016)
- Conference papers
- Structural modifications of metallic glasses followed by techniques of nuclear resonances
- Highly conductive barium iron vanadate glass containing different metal oxides
- Physico-chemical and optical properties of Er3+-doped and Er3+/Yb3+-co-doped Ge25Ga9.5Sb0.5S65 chalcogenide glass
- Spectroscopic ellipsometry characterization of spin-coated Ge25S75 chalcogenide thin films
- The challenge of methods of thermal analysis in solid state and materials chemistry
- Mössbauer spectroscopy: epoch-making biological and chemical applications
- Redistribution of iron ions in porous ferrisilicates during redox treatments
- Textural and morphology changes of mesoporous SBA-15 silica due to introduction of guest phase
- Carbon dioxide and methane adsorption over metal modified mesoporous SBA-15 silica
- Titania aerogels with tailored nano and microstructure: comparison of lyophilization and supercritical drying
- Solvent-free, improved synthesis of pure bixbyite phase of iron and manganese mixed oxides as low-cost, potential oxygen carrier for chemical looping with oxygen uncoupling
- Synthesis, structure and thermal expansion of the phosphates M0.5+x M′x Zr2−x (PO4)3 (M, M′–metals in oxidation state +2)
- Visible-light activated photocatalytic effect of glass and glass ceramic prepared by recycling waste slag with hematite
- Structure and properties of nanocrystalline nickel prepared by selective leaching at different temperatures
- Corrosion protection of zirconium surface based on Heusler alloy
- Toward the control of graphenic foams
Articles in the same Issue
- Frontmatter
- Graphical abstracts
- In this issue
- Preface
- 12th Conference on Solid State Chemistry (SSC-2016)
- Conference papers
- Structural modifications of metallic glasses followed by techniques of nuclear resonances
- Highly conductive barium iron vanadate glass containing different metal oxides
- Physico-chemical and optical properties of Er3+-doped and Er3+/Yb3+-co-doped Ge25Ga9.5Sb0.5S65 chalcogenide glass
- Spectroscopic ellipsometry characterization of spin-coated Ge25S75 chalcogenide thin films
- The challenge of methods of thermal analysis in solid state and materials chemistry
- Mössbauer spectroscopy: epoch-making biological and chemical applications
- Redistribution of iron ions in porous ferrisilicates during redox treatments
- Textural and morphology changes of mesoporous SBA-15 silica due to introduction of guest phase
- Carbon dioxide and methane adsorption over metal modified mesoporous SBA-15 silica
- Titania aerogels with tailored nano and microstructure: comparison of lyophilization and supercritical drying
- Solvent-free, improved synthesis of pure bixbyite phase of iron and manganese mixed oxides as low-cost, potential oxygen carrier for chemical looping with oxygen uncoupling
- Synthesis, structure and thermal expansion of the phosphates M0.5+x M′x Zr2−x (PO4)3 (M, M′–metals in oxidation state +2)
- Visible-light activated photocatalytic effect of glass and glass ceramic prepared by recycling waste slag with hematite
- Structure and properties of nanocrystalline nickel prepared by selective leaching at different temperatures
- Corrosion protection of zirconium surface based on Heusler alloy
- Toward the control of graphenic foams