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

Polarized spectroscopic properties of disordered Er3+:Gd2SrAl2O7 crystal

Open Access Open Access

Abstract

A multi-position disordered Er3+:Gd2SrAl2O7 single crystal was grown by the Czochralski method. Room temperature polarized spectral properties of the crystal were investigated in detail. The Judd–Ofelt theory was applied to analyze the polarized absorption spectra. The peak stimulated emission cross-sections of the 4I13/24I15/2 transition reach 0.79 ×10−20 and 1.25×10−20 cm2 for σ and π polarizations, respectively. The gain curves are flat with full widths at half the maximum larger than 60 nm. Fluorescence lifetimes of the 4I13/2 multiplet for Er3+ ions were fitted to be 4.93 ms and 3.48 ms from measured fluorescence decay curves for bulk and powder samples, respectively. The results show that the Er3+:Gd2SrAl2O7 crystal may be a potential gain medium for a 1.55 µm laser with low pumping threshold.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The eye-safe 1.55 µm laser is located in an atmospheric transmission window and the sensitive detection areas of Ge and InGaAs detectors operating at room temperature. For now, 1.55 µm laser operation via the 4I13/24I15/2 transition in Er3+-doped glasses and crystals have been widely used in the fields of optical communication, lidar, and remote sensing measurement [1,2]. When Er3+ ions are doped as an activator for 1.55 µm laser, Yb3+ ions are usually highly co-doped as a sensitizer after considering the weak absorption of Er3+ [3,4]. In materials with phonon energy about 1100 cm−1, such as phosphate glass [5] and Lu2Si2O7 [6] crystal, the doped Er3+ ions have a strong multiphonon non-radiative transition from 4I11/2 to 4I13/2, therefore, an efficient 1.55 µm laser operation may be realized in above materials co-doped with Er3+ and Yb3+ ions. But hosts with higher phonon energy such as RAl3(BO3)4 (R = Y, Gd, Lu) crystals [710], which are the most efficient gain media for the 1.55 µm laser after co-doped with Er3+ and Yb3+ ions, the shorter fluorescence lifetime of the upper laser multiplet 4I13/2, which is caused by the stronger non-radiative transition, decreases their energy storage capacities and leads to a high pumping threshold and heavy thermal load for laser operation. Er3+ ions in hosts with lower phonon energy have higher fluorescence quantum efficiency and energy storage capacity because of the weaker non-radiative transition. However, the longer fluorescence lifetimes of the pumping multiplet 4I11/2 makes reverse energy transfer and up-conversion process more serious, in turn leads to a lower efficiency of laser operation. By co-doped with Ce3+, the transition from 4I11/2 to 4I13/2 for Er3+ ions can be increased through the energy transfer of 4I11/2(Er3+) + 2F5/2(Ce3+) →4I13/2(Er3+) + 2F7/2(Ce3+) and has little effect on the fluorescence lifetime of the 4I13/2 multiplet [11]. The laser operations based on Er3+/Yb3+/Ce3+ tri-doped Ca2Al2SiO7 [12] and NaGd(WO4)2 [13] crystals have been achieved. Diode lasers at wavelength around 1.5 µm can also be used to pump Er3+ ions to the upper laser level 4I13/2 directly (so called resonant pump) in Er3+ single-doped crystals with lower phonon energy such as Er:YAG and Er:YVO4 [14]. The remarkable laser performance around 1.6 µm proves the research potential in this field.

Gd2SrAl2O7 (GSAO) crystal with phonon energy about 800 cm−1, which belongs to the tetragonal system with space group I4/mmm, has attracted research interest recently [15]. Its unit cell parameters are a = 3.705 Å, c = 19.781 Å, Z = 2 [16], It owns a layered perovskite-like structure [17], which is formed by perovskite (P) and rock salt (RS) interpenetrating in the sequence of PP/RS/PP/RS……, similar situation also occurs in Ln2SrAl2O7(Ln = La-Ho) [16]. In GSAO, Gd3+ and Sr2+ generally co-occupy site A (C4V) with the nine-coordinated in the RS layer and site B (D4h) with the twelve-coordinated in the PP layer, but Gd3+ prefers to occupy the former (site A) [18], Er3+ ions doped into these Gd3+ sites would result in inhomogeneous broadening of spectral lines, and which would be benefit to the realization of ultrashort pulses and tunable lasers. This disordered crystal melts congruently at about 1780 °C and can be grown by the Czochralski method. At room temperature, thermal conductivities of the crystal are 4.98 and 5.24 Wm−1K−1 along the a- and c-axis, respectively [15]. It is regarded as a promising host for solid state laser. At present, laser operations based on Yb3+-doped, Nd3+-doped, and Tm3+-doped GSAO crystals have been achieved [15,19]. Especially, the laser operation based on the Yb3+-doped GSAO has the maximum output power of 7.345 W and slope efficiency of 53.7% [19]. However, no details about the spectral properties related to the 1.55 µm laser operation of the Er3+:GSAO crystal have been reported to our knowledge. So this work is devoted to analyze the spectral properties of the Er3+:GSAO crystal related to the 1.55 µm laser.

2. Experiment

A single crystal of Er3+:Gd2SrAl2O7 was grown by the Czochralski method in a 2 kHz mid-frequency induction furnace (DJL-400). Sr2CO3 and Al2O3 with 99% purity and Er2O3 and Gd2O3 with 99.995% purity were used as starting materials. The raw materials were weighed according to the following reaction equation:

xEr2O3+(1-x)Gd2O3+SrCO3+Al2O3=Er2xGd(1-2x)SrAl2O7+CO2

After being mixed and ground fully, the materials had been sintered at 1200 °C for 10 h, and then 1500 °C for 10 h. The sintered polycrystal was placed in an iridium crucible and a crystal was grown at a pulling rate of 0.5–1.0 mm/h and a rotating rate of 10–15 rpm. By controlling the temperature, the crystal growth passes through four steps: necking, shoulder enlargement which has a cooling rate of about 2°C/h, equal diameter, and end, the grown crystal was detached slowly from the melt and cooled down to room temperature at a cooling rate of 10–30 °C/h. A Φ20 × 20 mm3 pale-yellow crystal was obtained finally (see Fig. 1). The yellow color of the crystal may be caused by defects in it. Similar phenomena have also been found in other Er3+-doped aluminate crystals [20]. The Er3+ concentration in the grown crystal was measured to be 0.78 at% (1.14×1020 cm−3) by the inductively coupled plasma atomic emission spectrometry (ICP-AES) method. Segregation coefficient of Er3+ in the crystal was calculated to be 0.78. The pattern of X-ray powder diffraction (XRD) of the grown Er3+:GSAO crystal was obtained by a diffractometer (Miniflex-600). Obviously, the diffraction peak positions of the Er3+:GSAO crystal are basically aligned with those of the standard card (see Fig. 2), which reveals that the low concentration of Er3+ doping hardly affects the crystal structure.

 figure: Fig. 1.

Fig. 1. The as-grown Er3+:GSAO crystal and sample after cut, annealed and polished.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. XRD patterns of the Er3+:GSAO and GSAO (JCPDS No.76-0095) crystals.

Download Full Size | PDF

After the optical indicatrix axes of the crystal was determined with a polarization microscope, a sample with dimensions of 5.66 × 5.57 × 2.25 mm3 was cut from the grown crystal. The 5.57 mm side of the crystal sample is parallel to the c-axis. In order to eliminate defects and thermal stress, the crystal sample was annealed at 1000 °C for 48 h in air and cooled to room temperature at a rate of 15–50 °C/h. After polished, the sample was used for spectral experiments (see Fig. 1). Room-temperature (RT) polarized absorption spectra were recorded with a Pekin-Elmer UV-VIS-NIR spectrophotometer (Lambda 950) in a range from 300 to 1700nm. RT polarized fluorescence spectra from 1400 to 1670 nm under excitation at 520 nm were recorded using a spectrometer (FLS1000, Edinburgh). The spectral resolutions were 1.0 nm in both the absorption and emission measurements. RT fluorescence decay curves at wavelengths of 980 and 1530 nm, corresponding to the transitions of 4I11/24I15/2 and 4I13/24I15/2 were recorded using the spectrometer (FLS1000, Edinburgh) when a microsecond flash xenon lamp (lF900 Edinburgh) was used as the exciting source and both the exciting wavelengths were set at 520 nm, corresponding to the transitions of 4I15/22H11/2.

3. Result and discussion

3.1. Absorption spectrum and Judd-Ofelt analysis

The RT polarized absorption spectra of the Er3+:GSAO crystal are shown in Fig. 3. They display strong anisotropy for different polarizations, and the absorptions for π polarization are generally stronger than those for σ polarization. It is worth mentioning that the absorption cross section around 1.5 µm in Er3+:GSAO is close to that in Er3+:YAG [21]. The absorption bands centered at 523 and 379 nm, which correspond to the 4I15/22H11/2 and 4I15/24G11/2 transitions, respectively, are usually considered as the hypersensitive transitions [22,23].

 figure: Fig. 3.

Fig. 3. RT polarized absorption spectra of the Er:GSAO crystal.

Download Full Size | PDF

The Judd–Ofelt theory [24,25] has been applied to calculate spectroscopic parameters of rare-earth ions in crystals and glasses [26]. Due to the lack of the values of refractive index for the GSAO crystal, the same refractive index (n≈1.94) [15] is adopted for both polarizations in calculation process. Referring to the Sellmeier equation of no and ne in CaGdAlO4 [27] with the same structure, the error caused by the possible error of the refractive index is generally less than 15% for the spontaneous transition probability. For simplicity, detailed calculating process following Ref. [26] is not described repeatedly here.

The measured oscillator strength fexp and the calculated one fcal are listed in Table 1. The J-O intensity parameters Ωt (t = 2, 4, 6) of some Er3+-doped crystals are listed in Table 2. The effective J-O parameters of the uniaxial Er3+:GSAO crystal, which are defined as Ωt,eff = (2Ωt,σ+Ωt,π)/3 [28,29], were calculated as: Ω2,eff = 3.87×10−20 cm2, Ω4,eff = 2.13×10−20 cm2, Ω6,eff = 1.77×10−20 cm2. It is generally thought that Ω2 is sensitive to the structure surrounding rare earth ions, and is associated with the asymmetry and the covalency of the lanthanide sites [30]; therefore, Ω2 is closely related to the hypersensitive transition [22,23]. The distortion index D of the rare earth ion coordination polyhedron in the crystal can be calculated through the formula [31]:

$$D = \frac{1}{n}\sum\nolimits_{i = 1}^n {\frac{{|{{l_i} - {l_{av}}} |}}{{{l_{av}}}}}$$
where n is the coordination number, li is the distance from the central atom to the ith coordinating atom, and lav is the average bond length. The crystal structure of GSAO and its coordination polyhedral is shown in Fig. 4. The values of sites A and B in the GSAO are 0.04768 and 0.01636, respectively. The distortion index of site A is larger than those in CaGdAlO4 (0.032) and CaYAlO4 (0.031) crystals which belong to the tetragonal system with space group of I4/mmm as well. So the absorption spectra especially for the bands correspond to hypersensitive transitions 4I15/22H11/2 and 4I15/24G11/2 of the Er3+:GSAO show stronger anisotropy for different polarizations than those of the Er3+:CaGdAlO4 [32] and Er3+:CaYAlO4 [33]. Moreover, the absorption bands of Er3+:GSAO are flatter and broader than those of the Er3+:CaGdAlO4 and Er3+:CaYAlO4 because of its multi-position and disordered structure. However, the average distance between Er3+ and its coordination ions of site A and B in Er3+:GSAO are 2.521 Å and 2.686 Å, respectively, both of them are longer than those in the CaGdAlO4 (2.518 Å) and CaYAlO4 (2.509 Å). The longer distance leads to a weaker crystal field, and thus a smaller Ω2,eff [34]. Under the combined effect of distortion index and crystal field, the value of Ω2,eff for the Er3+:GSAO is close to those for the Er3+:CaYAlO4 and Er3+:CaGdAlO4

 figure: Fig. 4.

Fig. 4. Crystal structure of GSAO and coordination polyhedral Gd(Sr)O9 and Sr(Gd)O12.

Download Full Size | PDF

Tables Icon

Table 1. Mean wavelengths and polarized experimental and calculated oscillator strengths of Er3+ ions in the GSAO crystal at room-temperature.

Tables Icon

Table 2. Comparison of the J-O parameters of Er3+ ions in the GSAO and some other crystals (in unit of 10−20cm2).

Based on the J-O intensity parameters, the spontaneous transition probability Aq, fluorescence branching ratio β, and radiative lifetime τr for different fluorescence multiplets of Er3+ in the GSAO can be calculated, the results of some transitions are listed in Table 3.

Tables Icon

Table 3. Spectral parameters of the Er3+:GSAO crystal calculated by the J–O theory.

3.2. Analysis of infrared emission properties

The Fuchtbauer–Ladenburg (F–L) formula can be used to calculate stimulated emission cross-section [38].

$${\mathrm{\sigma }_{em,q}}(\mathrm{\lambda } )= \frac{{{\mathrm{\lambda }^5}{A_q}({J \to J^{\prime}} ){I_q}(\mathrm{\lambda } )}}{{8\pi cn_q^2\int \mathrm{\lambda } {I_q}(\mathrm{\lambda } )d\mathrm{\lambda }}}$$
where q indicates the polarization of the fluorescence spectrum, ${I_q}(\lambda )$ is the fluorescence intensity at wavelength λ, $\; $c is the velocity of light, and nq is the refractive index of the crystal. It can be found from Fig. 4 that the emission cross-sections for π polarization are larger than those for σ polarization, and the peak emission cross-sections for the 4I13/24I15/2 transition are 0.79 ×10−20 cm2 at 1549 nm and 1.25×10−20 cm2 at 1546 nm for σ and π polarizations, respectively. Comparing with those of other Er3+-doped materials, for example, 1.13×10−20 cm2 (σ) for YAl3(BO3)4 [39], 1.12×10−20 cm2 (E//x) for Lu2Si2O7 [6], 1.09×10−20cm2 (π) for CaGdAlO4 [32], and 0.6×10−20cm2 for YAG [40], it can be found that the peak emission cross-sections of the Er3+:GSAO around 1550 nm are close to those of the typical Er3+-doped crystals.

Generally, the emission cross-section for the 4I13/24I15/2 transition calculated from the measured fluorescence spectrum by the F–L formula may be influenced by the radiation trapping [41]. Alternately, a reciprocity method (RM) can also be used to calculate the emission cross-section [42].

$$ \sigma_{e m, q}(\lambda)=\sigma_{a b s, q}(\lambda) \frac{z_{l}}{z_{u}} \exp \left(\frac{E_{z p l}-h c \lambda^{-1}}{k_{B} T}\right) $$
where ${\sigma _{abs,q}}(\lambda )$ is the polarized absorption cross-section, zl and zu are the partition functions of the lower and upper multiplets, due to lack of values for crystal field level positions of the Er3+:GSAO, those of aluminate crystals with the same structure [4345] were adopted and the value of zl/zu was estimated as 0.8, Ezpl is the energy separation between the lowest crystal field level of the upper and lower multiplets, which was estimated as 6591 cm−1 by the same way, kB is the Boltzmann constant, h is the Planck constant, and T is the temperature.

The wavelength dependences of polarized stimulated emission cross-sections of the 4I13/24I15/2 transition calculated by the above two methods are shown in Fig. 5. It can be found that the emission cross-section calculated by the two methods are similar at short wavelength, but at long wavelength, the emission cross-sections calculated by the F–L formula are larger than those calculated by the reciprocity method, which is caused by the radiation trapping.

 figure: Fig. 5.

Fig. 5. Polarized stimulated emission cross-sections for the 4I13/24I15/2 transition of the Er3+:GSAO crystal calculated by the F–L formula and the reciprocity method.

Download Full Size | PDF

The gain curve can be used to predict the possible laser wavelength. The 1.55 µm laser via the 4I13/24I15/2 transition operates in a three-level scheme, the gain curve can be calculated by

$$ \sigma_{g a i n}(\lambda)=P \sigma_{e m}(\lambda)-(1-P) \sigma_{a b s}(\lambda) $$
where P is the ratio of the population of Er3+ ions in the upper laser multiplet 4I13/2 to the total number of Er3+ ions. The gain curves are shown in Fig. 5.

It can be found from Fig. 6 that the gain cross section for π-polarization is larger than that for σ-polarization in a certain P, it indicates that π-polarization has a lower threshold of laser oscillation [46]. Therefore, polarized 1.55 µm laser output may be realized in a c-cut Er3+:GSAO crystal. For π-polarization with P ≥ 0.5, a positive gain can be achieved in a range from 1516 to 1635 nm at least. The curves are quite flat with full widths at half the maximum of 66 and 62 nm for σ and π polarizations, respectively, when P = 0.5. They are larger than those of Er:Yb:Lu2Si2O7 (about 30 nm for E//Y) [6], Er:Yb:KGd(PO3)4 (48 nm for E//Y) [47], and Er:CaGdAlO4 (56 nm for π polarization) [32] with the same value of P, and which is due to the multi-position disordered structure of the crystal. The broad and flat gain curves reveal that tunable and ultra-short lasers may be realized in the Er3+:GSAO crystal.

 figure: Fig. 6.

Fig. 6. Gain curves of the 4I13/24I15/2 transition for the Er3+:GSAO crystal with different P (P = 0.1, 0.2, …, 0.6, 0.7).

Download Full Size | PDF

The radiation trapping enhanced by the total reflection in crystal will cause the measured fluorescence lifetime longer than the intrinsic one [48]. The fluorescence decay curves of the 4I13/24I15/2 transition were recorded from both the bulk and powder samples (see Fig. 7) to display this phenomenon. It is worth mentioning that, when the measurement was carried out, the powder sample was dispersed in bromobenzene, which is low-toxic and has a refractive index (n = 1.55) close to that of the GSAO (n = 1.9). The linear relationship in Fig. 6 shows a single exponential behavior of the fluorescence decays and the fluorescence lifetimes of the 4I13/2 multiplet were fitted to be 4.93 and 3.48 ms for the bulk and powder samples, respectively. As for the 4I11/2 multiplet, the fluorescence lifetimes were fitted to be 612 and 585 µs for the bulk and powder samples, respectively. The shorter fluorescence lifetimes for the powder sample confirms the existence of radiation trapping, but the fluorescence lifetimes of the 4I13/2 multiplet for powder is still longer than the radiative lifetime of the 4I13/2 multiplet calculated by J-O theory (see Table 3). This may be caused by the uncertainty of the J–O theory [49] and still existence of the radiation trapping in the powder sample. Meanwhile, it also shows that the multiplet has a high fluorescence quantum efficiency. The same phenomenon also occurs in other Er3+-doped crystals with lower phonon energy, such as: Er3+:CaGdAlO4 [32], Er3+:LiNbO3 [50], and Er3+:Ca9Y(VO4)7 [51].

 figure: Fig. 7.

Fig. 7. RT fluorescence decay curves of the Er:GSAO bulk and powders samples at 1535 nm. Exciting wavelength is 520 nm.

Download Full Size | PDF

In general, the product of the emission cross-section σem and the fluorescence lifetime τf is used to compare the laser oscillating thresholds of different gain media [52]. For the convenience of comparison, the fluorescence lifetimes τf used in the calculation were all measured from crystals. The value of the product for Er3+:GSAO is 6.2×10−23 cm2s and larger than the 4.67×10−24 cm2s for Er3+:YAl3(BO3)4, and slightly larger than the 5.4×10−23 cm2s for Er3+:YAG. Therefore, when other factors are similar, it can be expected that the Er:GSAO has a lower oscillating threshold for 1.55 µm laser.

4. Conclusions

An Er3+:Gd2SrAl2O7 single crystal was grown by the Czochralski method. RT polarized spectral properties of the crystal as a gain medium for 1.55 µm laser were investigated in detail. The absorption spectra exhibit strong anisotropy. The emission cross-sections were determined from the polarized emission spectra by the F–L formula and also from the polarized absorption spectra by the reciprocity method. The peak emission cross sections around 1.55 µm are 0.79 ×10−20 cm2 for σ-polarization and 1.25×10−20 cm2 for π-polarization. For π-polarization with P ≥ 0.5, a positive gain can be achieved in a range from 1516 to 1635 nm at least. Because of the multi-position disordered structure of the crystal, the gain curves with P = 0.5 are flat with full widths at half the maximum of 66 and 62 nm for σ and π polarizations, respectively. The large emission cross section, wide and flat gain curves, and long fluorescence lifetime for upper laser level reveal that tunable and ultra-short 1.55 µm lasers may be realize in the crystal with a low oscillating threshold.

Funding

Ministry of Science and Technology of the People's Republic of China (2016YFB0701002); Chinese Academy of Sciences Key Project (KFJ–STS–QYZX–069, XDB20000000); Natural Science Foundation of Fujian Province (2019J01127).

Disclosures

The authors declare no conflicts of interest.

References

1. B. Denker and E. Shklovsky, Handbook of Solid-State Lasers, (Woodhead, 2013), pp. 341–358.

2. S. D. Setzler, M. P. Francis, and Y. E. Young, “Resonantly pumped eyesafe Erbium lasers,” IEEE J. Quantum Electron. 11(3), 645–657 (2005). [CrossRef]  

3. T. Fan, “Heat generation in Nd: YAG and Yb: YAG,” IEEE J. Quantum Electron. 29(6), 1457–1459 (1993). [CrossRef]  

4. L. D. DeLoach, S. A. Payne, L. Chase, L. K. Smith, W. L. Kway, and W. F. Krupke, “Evaluation of absorption and emission properties of Yb3+ doped crystals for laser applications,” IEEE J. Quantum Electron. 29(4), 1179–1191 (1993). [CrossRef]  

5. P. Laporta, S. Taccheo, S. Longhi, O. Svelto, and C. Svelto, “Erbium-ytterbium microlasers: optical properties and lasing characteristics,” Opt. Mater. 11(2-3), 269–288 (1999). [CrossRef]  

6. J. Huang, Y. Chen, H. Wang, Y. Lin, X. Gong, Z. Luo, and Y. Huang, “Efficient 2017 nm continuous wave laser operation of Czochralski grown Er3+:Yb3+:Lu2Si2O7 crystal,” Opt. Express 25(20), 24001–24006 (2017). [CrossRef]  

7. Y. Chen, Y. Lin, X. Gong, Q. Tan, Z. Luo, and Y. Huang, “2.0 W diode-pumped Er:Yb:YAl3(BO3)4 laser at 1.5–1.6 µm,” Appl. Phys. Lett. 89(24), 241111 (2006). [CrossRef]  

8. Y. Chen, Y. Lin, X. Gong, Z. Luo, and Y. Huang, “1.1 W diode-pumped Er:Yb laser at 1520 nm,” Opt. Lett. 32(18), 2759–2761 (2007). [CrossRef]  

9. Y. Chen, Y. Lin, X. Gong, Z. Luo, and Y. Huang, “Spectroscopic properties and laser performance of Er3+ and Yb3+ co-doped GdAl3(BO3)4 crystal,” IEEE J. Quantum Electron. 43(10), 950–956 (2007). [CrossRef]  

10. Y. Chen, Y. Lin, J. Huang, X. Gong, Z. Luo, and Y. Huang, “Spectroscopic and laser properties of Er3+:Yb3+:LuAl3(BO3)4 crystal at 1.5-1.6 µm,” Opt. Express 18(13), 13700–13707 (2010). [CrossRef]  

11. W. Stambouli, H. Elhouichet, C. Barthou, and M. Ferid, “Energy transfer induced photoluminescence improvement in Er3+/Ce3+/Yb3+ tri-doped tellurite glass,” J. Alloys Compd. 580, 310–315 (2013). [CrossRef]  

12. B. Simondi-Teisseire, B. Viana, A. M. Lejus, J. M. Benitez, D. Vivien, C. Borel, R. Templier, and C. Wyon, “Room-temperature CW laser operation at∼1.55 µm (eye-safe range) of Yb:Er and Yb:Er:Ce:Ca2Al2SiO7 crystals,” IEEE J. Quantum Electron. 32(11), 2004–2009 (1996). [CrossRef]  

13. J. Huang, Y. Chen, Y. Lin, X. Gong, Z. Luo, and Y. Huang, “Enhanced efficiency of Er:Yb:Ce:NaGd(WO4)2 laser at 1.5–1.6 µm by the introduction of high-doping Ce3+ ions,” Opt. Lett. 33(21), 2548–2550 (2008). [CrossRef]  

14. S. D. Setzler, M. P. Francis, Y. E. Young, J. R. Konves, and E. P. Chicklis, “Resonantly pumped eye safe Erbium lasers,” IEEE J. Quantum Electron. 11(3), 645–657 (2005). [CrossRef]  

15. F. Yuan, W. Liao, Y. Huang, L. Zhang, S. Sun, Y. Wang, Z. Lin, G. Wang, and G. Zhang, “A new ∼1 µm laser crystal Nd:Gd2SrAl2O7: growth, thermal, spectral and lasing properties,” J. Phys. D: Appl. Phys. 51(12), 125307 (2018). [CrossRef]  

16. I. Zvereva, Y. Smirnov, V. Gusarov, V. Popova, and J. Choisnet, “Complex aluminates RE2SrAl2O7 (RE = La, Nd, Sm–Ho): Cation ordering and stability of the double perovskite slab–rocksalt layer P2/RS intergrowth,” Solid State Sci. 5(2), 343–349 (2003). [CrossRef]  

17. I. A. Zvereva, V. F. Popova, E. A. Tugova, N. S. Pylkina, and V. Gusarov, “Phase equilibria in the Gd2O3–SrAl2O4 system,” Glass Phys. Chem. 31(6), 808–811 (2005). [CrossRef]  

18. J. Feng, B. Xiao, R. Zhou, W. Pan, and D. R. Clarke, “Anisotropic elastic and thermal properties of the double perovskite slab–rock salt layer Ln2SrAl2O7 (Ln = La, Nd, Sm, Eu, Gd or Dy) natural superlattice structure,” Acta Mater. 60(8), 3380–3392 (2012). [CrossRef]  

19. F. Yuan, Y. Huang, L. Liu, L. Zhang, S. Sun, G. Wang, Z. Du, Z. Lin, and J. Xu, “Spectroscopic properties and continuous-wave laser performance of Yb:Gd2SrAl2O7 crystal,” J. Alloys Compd. 808, 151715 (2019). [CrossRef]  

20. D. Zhou, X. Xu, X. Chen, H. Zhu, D. Li, J. Di, C. Xia, F. Wu, and J. Xu, “Crystal growth and spectroscopic properties of Er3+-doped CaYAlO4,” Phys. Status Solidi A 209(4), 730–735 (2012). [CrossRef]  

21. M. Eichhorn, S. T. Fredrich-Thornton, E. Heumann, and G. Huber, “Spectroscopic properties of Er3+:YAG at 300-550 K and their effects on the 1.6 µm laser transitions,” Appl. Phys. B 91(2), 249–256 (2008). [CrossRef]  

22. C. K. Jørgensen and B. Judd, “Hypersensitive pseudoquadrupole transitions in lanthanides,” Mol. Phys. 8(3), 281–290 (1964). [CrossRef]  

23. S. F. Mason, R. D. Peacock, and B. Stewart, “Ligand-polarization contributions to the intensity of hypersensitive trivalent lanthanide transitions,” Mol. Phys. 30(6), 1829–1841 (1975). [CrossRef]  

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

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

26. M. P. Hehlen, M. G. Brik, and K. W. Krämer, “50th anniversary of the Judd–Ofelt theory: An experimentalist's view of the formalism and its application,” J. Lumin. 136, 221–239 (2013). [CrossRef]  

27. P. Loiko, P. Becker, L. Bohatý, C. Liebald, M. Peltz, S. Vernay, D. Rytz, J. M. Serres, X. Mateos, Y. Wang, X. Xu, J. Xu, A. Major, A. Baranov, U. Griebner, and V. Petrov, “Sellmeier equations, group velocity dispersion, and thermo-optic dispersion formulas for CaLnAlO4 (Ln = Y, Gd) laser host crystals,” Opt. Lett. 42(12), 2275–2278 (2017). [CrossRef]  

28. Z. Luo, X. Chen, and T. Zhao, “Judd-Ofelt parameter analysis of rare earth anisotropic crystals by three perpendicular unpolarized absorption measurements,” Opt. Commun 134(1-6), 415–422 (1997). [CrossRef]  

29. H. Dai and O. M. Stafsudd, “Polarized absorption spectrum and intensity analysis of trivalent neodymium in sodium β″ alumina,” J. Phys. Chem. Solids 52(2), 367–379 (1991). [CrossRef]  

30. R. D. Peacock, “The intensities of lanthanide f-f transitions,” Struct. Bonding 22, 83–122 (1975). [CrossRef]  

31. W. H. Baur, “The comparison of the crystal structure of magnesium sulfate pentahydrate with copper sulfate pentahydrate and magnesium chromate pentahydrate,” Acta Crystallogr. 30(5), 1195–1215 (1974). [CrossRef]  

32. J. H. Huang, X. H. Gong, Y. J. Chen, Y. F. Lin, Z. D. Luo, and Y. D. Huang, “Spectral properties of Er3+-doped CaGdAlO4 crystal for laser application around 1.55 µm,” J. Alloys Compd. 585, 163–167 (2014). [CrossRef]  

33. J. C. Souriau, C. Borel, C. Wyon, C. Li, and R. Moncorgé, “Spectroscopic properties and fluorescence dynamics of Er3+ and Yb3+ in CaYA1O4,” J. Lumin. 59(6), 349–359 (1994). [CrossRef]  

34. M. Beltaif, M. Dammak, M. Megdiche, and K. Guidara, “Synthesis, optical spectroscopy and Judd–Ofelt analysis of Eu3+ doped Li2BaP2O7 phosphors,” J. Lumin. 177, 373–379 (2016). [CrossRef]  

35. R. Q. Piao, Y. Wang, Q. Xu, Z. B. Zhang, and D. L. Zhang, “Cross section spectral distribution of Er3+ electronic transitions in SrGdGa3O7 single-crystal,” J. Lumin. 210, 348–351 (2019). [CrossRef]  

36. H. L. Xu, L. G. Zhou, Z. W. Dai, and Z. K. Jiang, “Decay properties of Er3+ ions in Er3+:YAG and Er3+:YAlO3,” Phys. B 324(1-4), 43–48 (2002). [CrossRef]  

37. Y. Ding, G. Zhao, J. Chen, Q. Dong, Y. Nakai, and T. Tsuboi, “Near-infrared emission bands of Er3+-doped YAP and LSO crystals,” J. Lumin. 131(8), 1577–1583 (2011). [CrossRef]  

38. 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]  

39. W. You, Y. Lin, Y. Chen, Z. Luo, and Y. Huang, “Polarized spectroscopy of Er3+ ions in YAl3(BO3)4 crystal,” Opt. Mater. 29(5), 488–493 (2007). [CrossRef]  

40. 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]  

41. F. Auzel, G. Baldacchini, L. Laversenne, and G. Boulon, “Radiation trapping and self-quenching analysis in Yb3+, Er3+, and Ho3+ doped Y2O3,” Opt. Mater. 24(1-2), 103–109 (2003). [CrossRef]  

42. D. E. McCumber, “Einstein relations connecting broadband emission and absorption spectra,” Phys. Rev. 136(4A), A954–A957 (1964). [CrossRef]  

43. J. Andrew, J. Hutchinson, R. Horacio, and D. Merkle, “Spectroscopic evaluation of CaYA1O4 doped with trivalent Er, Tm, Yb and Ho for eyesafe laser applications,” Opt. Mater. 3(4), 287–306 (1994). [CrossRef]  

44. J. P. R. Wells, M. Yamaga, R. W. Mosses, T. P. J. Han, H. G. Gallagher, and T. Yosida, “Polarized laser selective excitation and electron paramagnetic resonance of Er3+ centers in SrLaAlO4 crystals,” Phys. Rev. B 61(6), 3905–3914 (2000). [CrossRef]  

45. J. P. R. Wells, M. Yamaga, N. Kodama, and T. P. J. Han, “Polarized laser spectroscopy and crystal-field analysis of Er3+ doped CaGdAlO4,” J. Phys.: Condens. Matter 11(39), 7545–7555 (1999). [CrossRef]  

46. W. Koechner, Solid-state Laser Engineering, (Springer, 2006).

47. G. Gong, Y. Chen, Y. Lin, J. Huang, X. Gong, Z. Luo, and Y. Huang, “Growth and spectroscopic properties of Er:Yb:KGd(PO3)4 crystal as a promising 1.55 µm laser gain medium,” Opt. Mater. Express 6(11), 3518–3526 (2016). [CrossRef]  

48. D. S. Sumida and T. Y. Fan, “Effect of radiation trapping on fluorescence lifetime and emission cross-section measurements in solid-state laser media,” Opt. Lett. 19(17), 1343–1345 (1994). [CrossRef]  

49. D. H. Zhou, X. D. Xu, C. T. Xia, and J. Xu, “Spectroscopic analysis and lasing of Er:Lu1.5Y1.5Al5O12 crystals,” J. Opt. Soc. Am. B 28(10), 2543–2548 (2011). [CrossRef]  

50. J. Amin, B. Dussardier, T. Schweizer, and M. Hempstead, “Spectroscopic analysis of Er3+ transitions in lithium niobite,” J. Lumin. 69(1), 17–26 (1996). [CrossRef]  

51. F. Yuan, W. Zhao, S. Sun, L. Zhang, Y. Huang, Z. Lin, and G. Wang, “Polarized spectroscopic properties of Er3+:Ca9Y(VO4)7 crystal,” J. Lumin. 154, 241–245 (2014). [CrossRef]  

52. Z. Luo and Y. Huang, Physics of Solid-State Laser Materials, (Springer Series in Materials Science, 2020), p. 316.

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 (7)

Fig. 1.
Fig. 1. The as-grown Er3+:GSAO crystal and sample after cut, annealed and polished.
Fig. 2.
Fig. 2. XRD patterns of the Er3+:GSAO and GSAO (JCPDS No.76-0095) crystals.
Fig. 3.
Fig. 3. RT polarized absorption spectra of the Er:GSAO crystal.
Fig. 4.
Fig. 4. Crystal structure of GSAO and coordination polyhedral Gd(Sr)O9 and Sr(Gd)O12.
Fig. 5.
Fig. 5. Polarized stimulated emission cross-sections for the 4I13/24I15/2 transition of the Er3+:GSAO crystal calculated by the F–L formula and the reciprocity method.
Fig. 6.
Fig. 6. Gain curves of the 4I13/24I15/2 transition for the Er3+:GSAO crystal with different P (P = 0.1, 0.2, …, 0.6, 0.7).
Fig. 7.
Fig. 7. RT fluorescence decay curves of the Er:GSAO bulk and powders samples at 1535 nm. Exciting wavelength is 520 nm.

Tables (3)

Tables Icon

Table 1. Mean wavelengths and polarized experimental and calculated oscillator strengths of Er3+ ions in the GSAO crystal at room-temperature.

Tables Icon

Table 2. Comparison of the J-O parameters of Er3+ ions in the GSAO and some other crystals (in unit of 10−20cm2).

Tables Icon

Table 3. Spectral parameters of the Er3+:GSAO crystal calculated by the J–O theory.

Equations (4)

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

D = 1 n i = 1 n | l i l a v | l a v
σ e m , q ( λ ) = λ 5 A q ( J J ) I q ( λ ) 8 π c n q 2 λ I q ( λ ) d λ
σ e m , q ( λ ) = σ a b s , q ( λ ) z l z u exp ( E z p l h c λ 1 k B T )
σ g a i n ( λ ) = P σ e m ( λ ) ( 1 P ) σ a b s ( λ )
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.