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

Spectroscopic properties and quenching mechanism of 2 μm emission in Ho3+ doped germanate glasses and fibers

Open Access Open Access

Abstract

The 2 μm emission spectra and lifetimes of Ho3+ ions in germanate glasses with different doping concentrations were investigated. The Judd-Ofelt parameters, radiative transition rates, and emission cross-section of Ho3+ were calculated based on the absorption and emission spectra. The energy transfer rate to hydroxyl groups and non-radiative rate of 5I7 level were calculated by fitting the variations of lifetimes vs. the doping concentrations. Besides, the unclad fibers with highly Ho3+ doped (6 × 1020 cm−3) were fabricated and pumped using a homemade 1.94 μm fiber laser, and the peak of emission spectra showed a redshift with the increasing fiber length.

© 2015 Optical Society of America

1. Introduction

As one of the atmospheric transparent windows, single frequency fiber laser operating at 2 μm has attracted many researchers because of its numerous potential applications including eye safe range-finding, LIDAR for heterodyne measurement of wind velocity and as sources for mid-infrared optical parametric oscillators [1–4]. Usually, Tm3+ -doped, Ho3+ -doped, and Raman fiber lasers are the most widespread way of getting 2-micron laser output. However, Raman fiber lasers requires high power pump sources in the spectral range, which is a shortcoming in the view of practical applications [5]. The Tm3+ doped sources are typically limited to efficient operation at <2.05 μm [6], although Tm3+ doped fiber lasers with high output power and slope efficiency have been demonstrated in silica [7], silicate [8] and germanate fibers [9]. The transition (5I65I7) of Ho3+ ions produces radiation in the range of 2.05 μm to 2.2 μm [10], which could totally match the applications which require good atmospheric propagation. Atmospheric transmission spectra provided by ModTran [6] show the advantage of operating at wavelength beyond 2.1μm in comparison to the windows accessible thulium sources.

In order to get a single frequency output, a short laser cavity with enough net gain is required [1]. Therefore, host materials must have high solubility of holmium ions, which are usually exhibited by heavy metal glasses [1] (e.g., germanate, telluride, and fluoride glasses). Another concern for choosing host materials is the phonon energy. High phonon energy results in fast multiphonon relaxation, thereby causing thermal damage and low quantum efficiency [5]. Unlike telluride and fluoride glasses, germanate glasses combine large rare earth solubility, moderate phonon energy (800 cm−1 to 900 cm−1) with strong mechanical strength and high damage threshold. In 2009, A single frequency fiber laser operating at 2.05 μm with over 50 mW output power has been obtained in a 2 cm long Ho3+ -doped germanate glass fiber. The maximum doping concentration of Ho2O3 in the fiber is 3wt% [9].

In this work, we prepared a series of Ho2O3 -doped lead germanate glasses with doping concentration varied from 0.1 wt% to 4 wt%. The optical and spectroscopic properties of the lead germanate glasses with different doping concentrations were presented and studied. The lifetime quenching caused by OH groups was also analyzed, and the energy transfer rates to OH groups under different doping concentrations were obtained. Unclad germanate fibers with 4 wt% Ho2O3 were fabricated, and pumped by using an in-house constructed 2 μm fiber laser.

2. Experiment methods

Glass samples of the lead-germanate, 50 mol%GeO2 + 5 mol%SiO2-20 mol% PbO-20 mol%CaO + 5 mol% K2O with x wt% of Ho2O3 (x = 0.1, 0.5, 2, 3, 4), denoted as G1 to G5, were prepared by using a conventional quenching method. The commercially available raw materials used were GeO2 and Ho2O3 with 99.999%, and the others were of reagent grade. Well-mixed 20 g batches were melted at 1200 °C for 40 min in a platinum crucible. Dried O2 was bubbled into the glass melt for 30 minutes to eliminate OH. The melts were then quenched on a stainless steel plate and annealed for 2 h at 500 °C in an muffle furnace. Another glass sample with the same components of G5, named G6, was prepared using the same process described above, except that the batch of G6 was first dried in a vacuum drying oven at 110 °C for 24 h before melting to remove the crystal water content in the raw materials, and the entire preparation process of G6 was conducted in a dried environment (relative humidity ≤5% at 18°C). The annealed samples were cut and polished to dimensions of 10 mm × 10 mm × 0.5 mm for optical and spectroscopic measurements.

A glass batch with components similar to G6, in quantity of 300 g of powder, was melted using the same process applied to G6. Considering the large mass of this glass batch, a longer time of each procedure is required. The well-mixed 300g batch was heated for more than 48 h in the vacuum drying oven before melting. Then, the mixture was melted in platinum crucibles for 3 h in an electrically heated furnace at ~1200 °C with a protective atmosphere with dried oxygen. Three hours later, dried oxygen was bubbled into the melt for 1 hour to further remove the OH- groups in the melt. The stir effect caused by bubbles could homogenize the melt and reduce the cluster of rare earth ions. Finally, the glass batch were quenched on a stainless steel plate and finely annealed at glass transition temperature for 100 h. The entire procedure was also conducted in a drying room. A glass rod was drilled from the bulk glass (BG) and finely polished, then drawn into a 125 μm diameter unclad fiber in a fiber drawing tower.

The refractive index and density of the samples were measured by the prim minimum deviation and the Archimedes method using distilled water as an immersion liquid. The absorption spectra were recorded by using a PerkinElmer Lambda 900 UV/VIS/NIR spectrophotometer. The emission spectra were measured with an Edinburg FLS920 type spectrometer. All measurements were conducted at room temperature.

The concentration of Ho ions was calculated using the measured density and the quoted weight fraction of Ho. The material density was measured as 4.98 g/cm3. For a given weight fraction (W), the Ho density (NHo), is calculated as 2Wρ/M, where ρ is the density and M is the mass of Ho2O3 (6.2767 × 10−22 g). For the 2 wt% sample (G2), the doping concentration of Ho3+ ions was calculated to be 3.12 × 1020 cm−3.

3.Results and discussion

3.1 Infrared (IR) transmission, absorption spectroscopic properties, and Judd-Ofelt analysis

The infrared transmittance spectrum of G4 is presented in Fig. 1. As is shown, the maximum transmittance is 84.6%, and the IR cutoff wavelength reaches approximately 5.8 μm. The refractive index of G4 at 2.5 μm was fitted by Cauchy dispersion equation and the value is 1.787. This value corresponds to 15.3% Fresnel reflection loss of the two surfaces of the used glass sheet, and such caused the low IR transmittance of the G4 glass. The wide absorption band around 3 μm was caused by the free OH groups [11]. The hydroxyl absorption coefficient was calculated using the equation [12]

 figure: Fig. 1

Fig. 1 IR transmittance spectrum of G4; the inset shows the absorption coefficients of all samples.

Download Full Size | PDF

αOH=1lln(T0T)

where l is the sample thickness, T is the transmittance at 3 μm, and T0 is the transmittance of the host matrix. The calculated results of G1 to G6 are presented in the inset of Fig. 1. The absorption coefficients of G1 to G5 are nearly kept at 2 cm−1, whereas the value of G6 remarkably decreases to 0.45 cm−1. It means that the preparation procedures applied to G6 could effectively eliminate the OH groups. According to reference [11], the content of the free OH groups was estimated from the measured maximum absorption coefficient. The free OH content NOH (ions/cm−3) was acquired as follows [11]

NOH=NAεlln(T0T)

where NA is Avogadro’s constant, and ε is the molar absorptivity of the free OH groups in the glass. Here we adopt the molar absorptivity in silicate glasses, 49.1 × 103 cm2/mol [13], to estimate the OH content in the germanate glass because of unavailable result for germanate glass by now. As a result, the typical contents of OH groups are about 2.45 × 1019 cm−3 in G1 to G5, and 0.5 × 1019 cm−3 in G6, 4 times smaller than that of G1 to G5. Therefore, it can be concluded that the dehydration method utilized to G6 could efficiently reduce the OH content.

Figure 2 shows the subtracted absorption spectrum of the 3 wt% Ho2O3-doped germanate glass from 300 nm to 2200 nm. Eight inhomogeneously broadened absorption peaks were observed from the absorption curve. The bands located at approximately 420 nm, 448 nm, 486 nm, 538 nm, 640 nm, 890 nm, 1148 nm, and 1936 nm were designed to the transition from the ground state 5I8 level to the 5G5, 5F1 + 5G6,5F2 + 3K8, 5F3,5F4 + 5S2,5F5, 5I5,5I6, and 5I7 levels, respectively. The energy level locations were very similar to those previously reported for telluride [14] and fluoride [15] glass hosts. Owing to the strong intrinsic bandgap absorption located around 370 nm, the absorption from Ho3+ cannot be clearly identified below 370 nm. We adopted a 640 nm laser diode to pump the Ho3+ -doped glass samples in this work because of the unavailable of 1150 nm pump sources.

 figure: Fig. 2

Fig. 2 Subtracted absorption spectrum of G4.

Download Full Size | PDF

Three Judd-Ofelt intensity parameters Ωt(t = 2, 4, 6) of Ho3+ ions in the germanate glasses, which dominate the radiative transition probabilities, are derived from the least square fit to the measured and experimental oscillator strengths. In the Judd-Ofelt theory, the line strength of the electric-dipole transition between an initial level aJ and a final level bJ'is given by [16, 17]

Sed(aJ,bJ)=t=2,4,6Ωt|aJU(t)bJ|2

The squared terms in Eq. (3) are the matrix elements of the doubly reduced unit tensor U(t) (t = 2, 4, and 6), which are known to be almost independent on host materials. Thus, the U(t) values calculated for Ho3+ in LaF3 by Weber et al. [18] were used.

The electric-dipole line strength is related to the integrated absorption coefficient over the band and was obtained by a equation described as [17]

bandκ(λ)dλ=8π3e2λ¯N3ch(2J+1)(n2+2)9nSed(aJ,bJ)
where κ(λ) is the absorption coefficient of the transition from a level aJ to a level bJ', λ¯ is the mean wavelength of the absorption band, N is the concentration of rare-earth ions, and n is the refractive index of the host material. A least squares methods for Eqs. (3) and (4) was used in order to determine the set of Ωt parameters. By adopting the absorption band labeled in Fig. 2 for the least square fitting excluding the absorption form the ground state 5I8 to 5I7, which has a magnetic dipole contribution (selection rules: ΔJ = 0, ± 1, ΔS = 0, ΔL = 0), the Judd-Ofelt parameters of Ho3+ ions in this lead germanate glass (refractive index n(633 nm) = 1.812) were determined as Ω2 = 4.7 × 10−20 cm2, Ω4 = 1.63 × 10−20cm2, and Ω6 = 0.81 × 10−20 cm2. Generally, Ω2 is closely related to the ligand symmetry of the host material and the covalence of the bond formed by rare earth ions, and it is hypersensitive to the compositional changes in the host materials. Ω6 is related to the overlap integral of the 4f and 5d orbits. The values for Ω4 and Ω6 also provide some information of the rigidity and viscosity of the host [19]. However, compared with Ω2, structural informations carried by Ω4 and Ω6 is marginal and sometimes inaccurate. The value of Ω2 in GeO2-BaO-Ga2O3 system [20] is larger than that in this lead-germanate glass, which indicates the higher local symmetry of Ho3+ ions in the lead-germanate glass. The higher symmetry of Ho3+ ions is originated from the large amount of non-bridging oxygen in the lead-germanate glass due to the lack of intermediate oxide of glass network, thus the of requirements of surrounding configuration of rare-earth ions could be easily satisfied.

Table 1 presents the calculated total radiative transition probability Ar, lifetimes τr, and branch ratios β of several energy levels. The results show that the value Ar of the 5I7 level of Ho3+ was 94.5s−1, which was larger than that of silicate glasses [21], but smaller than that of tellurite glasses [22]. Given that the radiative transition probability was proportional to n(n2 + 2)2/9, and the refractive index in germanate glass was usually intermediate between silicate and tellurite glasses, it was reasonable to obtain the presented result [14].

Tables Icon

Table 1. Radiative transition rates (Ar), branching ratios (β), and radiative lifetimes (τr) for different energy levels

3.2 Emission properties and lifetimes of the 5I75I8 transition with different Ho3+ concentrations

The emission spectra of Ho3+-doped germanate samples with different doping concentrations were measured at room temperature. A 640 nm laser diode was used to excite Ho3+ for fluorescence properties characterization. Considering the large phonon energy of 822 cm−1 of this glass determined from the phonon side spectrum, we supposed that most of the excited Ho3+ ions in 5F5 level would finally relax to 5I6 level via multiphonon relaxation after the ions exciting from ground level to 5F5 level. All detailed process are indicated in Fig. 3.

 figure: Fig. 3

Fig. 3 Energy level diagram of Ho3+ in the lead-germanate glass; the dash lines represent the non-radiative processes, the solid lines stand for absorption (red line) and emission (A31 and A21) transition of Ho3+ ions, and the solid curve means the energy transfer from Ho3+ to OH groups.

Download Full Size | PDF

Figure 4(a) shows the normalized fluorescence spectra of G1 to G6. Four major emission peaks were observed: one was at 1890 nm, and three were located at 1942, 2010 and 2050 nm, respectively. The multiple peak profiles of the emission spectrum mainly comes from the rich Stark splits of two transition levels of 5I7 and 5I8. The long wavelength end of the emission spectrum extends well beyond 2100 nm. Besides, no broadening effect is observed from the normalized fluorescence spectra, with, which are usually observed in the glasses with high doping concentrations. In addition, the strongest emission is always located at 2010 nm. The stimulated emission cross-section of Ho3+ in this germanate glass was calculated from the fluorescence spectra using Füchtbauer-Ladenburg equation [11]

σem=λ4Ar8πcn2λI(λ)λI(λ)dλ 
where n is the refractive index. λ is the wavelength, I(λ) is the emission spectrum and c is the light speed. The emission cross section spectrum is performed in Fig. 4(b), and the maximum cross section was 4.46 × 10−21 cm2 at 2010 nm.

 figure: Fig. 4

Fig. 4 (a) Normalized emission spectra of G1 to G6; (b) overlaps of absorption and emission cross-section spectra of the 2 μm band; (c) (τm1τr1)(black squares) and integrated 2 μm emission intensity (red squares) as a function of Ho3+ ions; τmand τr are the measured and calculated lifetimes of 5I7 level, respectively; (d) fluorescence decay curves of Ho3+ doped germanate glasses;.

Download Full Size | PDF

Variations in the integrated emission intensity (red cubic) of the 5I75I8 transition with different Ho3+ concentrations are presented in Fig. 4(c). The integrated emission intensity increased with increasing Ho3+ concentration, although there were some approaches in which could decrease the emission intensity instantaneously such as energy transfer to OH or other quenching centers. Especially, the integrated emission intensity of G6 was twice as strong as that of G5, indicating that removal of OH groups could efficiently improve the 2 μm emission intensity.

As Fig. 4(d) shows, the fluorescence lifetime of 5I75I8 transition of Ho3+ ions near 2-micron decreased from 2.96 ms to 1.85 ms when the doping concentration of Ho2O3 increased from 2 wt% to 4 wt%, whereas the lifetime of G6 increased to 4.85 ms, which was higher than that of G5 at the same doping concentration. We did not acquire accurate decay curves for the samples with low doping concentration (≤0.5 wt%) because of the weak fluorescence intensity under short pulse pumping condition. The lifetime is defined as the time for the fluorescence intensity decrease to 1/e of the initial value. The fluorescence decay curves are strait lines respect to a log scale of Y-axis, which indicate there is no other significant nonlinear energy transfer between Ho3+ ions involved. Therefore, a possible mechanism for the lifetime decreasing is the energy transfer to the OH groups [23]. Therefore, considering multiphonon decay, the measure lifetime τm of Ho3+-excited state is finally given by [23]

τm1=τr1+W21+WOH
where W21 is the multiphonon decay rate from 5I7 to 5I8 level taken as a constant here and WOH the energy transfer rate between Ho3+ and OH groups. Here [22]
WOH=92πNHo3+(αNOH)τrN02
where NOHand NHo3+ are the concentrations of OH groups and Ho3+ ions, respectively. αrepresents the proportion of Ho3+ ions coupled to OH groups. N0 is the critical concentration defined as N0=(4πR03/3)1, R0 is the critical distance at which the energy transfer for an isolated donor-acceptor pair separated by R0occurs with the same rate as the spontaneous deactivation in the donor itself. R0 is given by [11]
R06=6cτr16π4n2glowDgupDσemiD(λ)σabsA(λ)dλ
where glow and gup are the degeneracies of donor (D) states, respectively, from the lower and upper levels involved in the process. σemiD(λ) and σabsA(λ) are emission (donor) and absorption (acceptor) cross section spectra. In this case, the donor and the acceptor are all Ho3+ ions, the overlaps of the absorption and emission spectra are presented in Fig. 4(b). By using Eq. (6), R0 is calculated to be 18.9 Å, and the corresponding critical concentration N0 is 3.54 × 1019 cm−3.Combining Eqs. (6) and (7), we can deduced that

τm-1-τr-1=W21+92πNHo3+(αNOH)τrN02

Taking the values of NOH, N0, and τr into Eq. (9), and then fitting the Eq. (9) to the data shown in Fig. 4, we obtained the values of W21 and α as 32.9 ± 8.5 s−1, 24.4%, respectively. Thus, the value of WOH for different doping concentration of Ho2O3 also can be calculated. The maximum quantum efficiency (ηm) of the 5I75I8 transition of Ho3+ ions in this lead-germanate glass expressed as A21/(A21 + W21), was 74.1%. The quantum efficiencies, η, of G3~G6 were defined to be A21/(A21 + W21 + WOH). Based on the deduced parameters, we also could obtain WOH, η and τm of G1 and G2 samples. All the results are listed in Table 2. The quantum efficiency of G6 is 47.1%, while 17.5% in the silicate glass [21]. A higher η could be reached by further removing the OH groups or other quenching centers.

Tables Icon

Table 2. Doping concentrations (NHo), energy transfer rate to OH groups (WOH), quantum efficiency (η), measured lifetimes (τm) at 5I7 level, and the product of τm × NHo of Ho3+ in G1 to G6 glasses.

Moreover, using Eq. (7), the WOH of G6 is calculated as 82.3 s−1, which was close to the value (74.6 s−1) calculated by combining the measured lifetime and W21. Therefore, it can be concluded that the calculations we did match the experimental value well.

3.3 Gain properties and emission spectra in the unclad fibers

The parameter of material gain is an important index for the host glass. From a steady rate equation, the material gain is simply given by [19]

G=exp[σgσaNτml]
where σgis the gain cross-section, σais the pump absorption cross-section for the pumping wavelength, N is the concentration of doping ions, τmis the ions excited state lifetime at the corresponding concentration N, and l is the gain length. Due to the negligible influence to σgand σa for different doping concentrations, the gain is mainly determined by the product of N × τm in the same gain length. A larger product predicts a higher gain. The products for all samples are listed in Table 2, and G6 exhibits the largest value. Therefore, G6 was chosen to fabricate the glass fibers. The fibers were fabricated by the way descripted in Section 2. The loss of the unclad fibers was measured by cut back method using a pig-tailed single mode laser operating at 1310 nm. The measured loss is ~1.7 dB/m.

Figure 5 compares the mid-infrared emission spectra of 4 wt% of Ho2O3-doped BG glass fibers as a function of fiber length, using the pump excitation at 1.94 μm with a pump power of 500 mW. The spectrum line width in the bulk glass is 154 nm, which is much larger than in the fiber (59 nm). The line width in different fiber length exhibits no obvious change. Thus, the large overlaps between the absorption and emission spectra are proposed to causing the linewidth narrowing. With increasing the fiber length, the center wavelength shifts to the red side, from 2127 nm to 2142 nm, which can be ascribed to the growing radiation trapping along the fiber.

 figure: Fig. 5

Fig. 5 Normalized infrared emission spectra of 4 wt% Ho2O3 doped germanate unclad fibers with different fiber length under the excitation of 1940 nm fiber laser.

Download Full Size | PDF

4. Conclusion

In summary, the mid-infrared luminescence properties of Ho-doped germanate glasses and fibers at 2 μm were investigated. Spectroscopic parameters for characterizing Ho3+ doped germanate glasses were analyzed and calculated based on Judd-Ofelt theory. The lifetime quenching mechanism of 5I7 level of Ho3+ ion was also presented, the quenching rate to OH groups decreased dramatically from 412.2 s−1 to 74.6 s−1 under the same doping concentration (~6 × 1020 ions/cm3) by reducing the OH groups. The index of gain medium, τm × NHo, was also improved to 30.26 × 1020 ms/cm3 with a quantum efficiency of 47.1%, which predict a high gain and a low pumping threshold.

The emission spectra of 2 μm in bulk glasses and fibers were compared and analyzed. The linewidth was 154 nm in bulk glasses, while 59 nm in the fibers. The large overlap between emission and absorption spectra was proposed for the spectra narrowing. The central wavelength of the emission band shows a red shift with the increment of fiber length and locates around 2120 nm, which can be ascribed to the increasing radiation trapping with the fiber length. The above results indicate that this germanate glasses were attractive and promising laser materials for 2.1 μm single frequency laser output.

Acknowledgments

This research is financially supported by the Chinese National Natural Science Foundation (Grant No.60937003) and Natural Science Foundation of Shanghai, China (Grant No. 12ZR1451600).

References and links

1. Q. Fang, W. Shi, K. Kieu, E. Petersen, A. Chavez-Pirson, and N. Peyghambarian, “High power and high energy monolithic single frequency 2 μm nanosecond pulsed fiber laser by using large core Tm-doped germanate fibers: experiment and modeling,” Opt. Express 20(15), 16410–16420 (2012). [CrossRef]  

2. G. Li, Y. Gu, B. Yao, L. Shan, and Y. Wang, “High-power high-brightness 2 μm continuous wave laser with a double-end dif fusion-bonded Tm, Ho: YVO4 crystal,” Chin. Opt. Lett. 11(9), 091404 (2013). [CrossRef]  

3. J. Yang, Y. Tang, and J. Xu, “Development and applications of gain-switched fiber lasers [Invited],” Photonics Research 1(1), 52–57 (2013). [CrossRef]  

4. X. Wang, P. Zhou, X. Wang, H. Xiao, and L. Si, “51.5 W monolithic single frequency 1.97 m Tm-doped fiber amplifier,” High Power Laser Science and Engineering 1(3-4), 123–125 (2013). [CrossRef]  

5. A. S. Kurkov, V. V. Dvoyrin, and A. V. Marakulin, “All-fiber 10 W holmium lasers pumped at λ=1.15 microm,” Opt. Lett. 35(4), 490–492 (2010). [CrossRef]   [PubMed]  

6. A. Hemming, N. Simakov, J. Haub, and A. Carter, “A review of recent progress in holmium-doped silica fibre sources,” Opt. Fiber Technol. 20(6), 621–630 (2014). [CrossRef]  

7. S. D. Jackson and T. King, “High-power diode-cladding-pumped Tm-doped silica fiber laser,” Opt. Lett. 23(18), 1462–1464 (1998). [CrossRef]   [PubMed]  

8. J. Geng, Q. Wang, T. Luo, S. Jiang, and F. Amzajerdian, “Single-frequency narrow-linewidth Tm-doped fiber laser using silicate glass fiber,” Opt. Lett. 34(22), 3493–3495 (2009). [CrossRef]   [PubMed]  

9. J. Wu, Z. Yao, J. Zong, and S. Jiang, “Highly efficient high-power thulium-doped germanate glass fiber laser,” Opt. Lett. 32(6), 638–640 (2007). [CrossRef]   [PubMed]  

10. X. Yang, H. Huang, D. Shen, H. Zhu, and D. Tang, “2.1 μm Ho:LuAG ceramic laser intracavity pumped by a diode-pumped Tm:YAG laser,” Chin. Opt. Lett. 12(12), 121405 (2014). [CrossRef]  

11. X. Fan, K. Li, X. Li, P. Kuan, X. Wang, and L. Hu, “Spectroscopic properties of 2.7 μm emission in Er3+ doped tellurite glasses and fibers,” J. Alloys Compd. 615, 475–481 (2014). [CrossRef]  

12. F. Huang, X. Liu, W. Li, L. Hu, and D. Chen, “Energy transfer mechanism in Er3+ doped fluoride glass sensitized by Tm3+ or Ho3+ for 2.7 μm emission,” Chin. Opt. Lett. 12(5), 051601 (2014). [CrossRef]  

13. L. Nemec and J. Gotz, “Infrared absorption of OH− in E glass,” J. Am. Ceram. Soc. 53(9), 526 (1970). [CrossRef]  

14. G. X. Chen, Q. Y. Zhang, G. F. Yang, and Z. H. Jiang, “Mid-Infrared Emission Characteristic and Energy Transfer of Ho3+-Doped Tellurite Glass Sensitized by Tm 3+.,” J. Fluoresc. 17(3), 301–307 (2007). [CrossRef]   [PubMed]  

15. B. Peng and T. Izumitani, “Optical properties, fluorescence mechanisms and energy transfer in Tm3+, Ho3+ and Tm3+-Ho3+ doped near-infrared laser glasses, sensitized by Yb3+,” Opt. Mater. 4(6), 797–810 (1995). [CrossRef]  

16. X. Wang, L. Hu, K. Li, Y. Tian, and S. Fan, “Spectroscopic properties of thulium ions in bismuth silicate glass,” Chin. Opt. Lett. 10(10), 101601 (2012). [CrossRef]  

17. W. F. Krupke, “Optical Absorption and Fluorescence Intensities in Several Rare-Earth-Doped Y2O3 and LaF3 Single Crystals,” Phys. Rev. 145(1), 325–337 (1966). [CrossRef]  

18. M. J. Weber, B. H. Matsinger, V. L. Donlan, and G. T. Surratt, “Optical Transition Probabilities for Trivalent Holmium in LaF3 and YAlO3,” J. Chem. Phys. 57(1), 562–567 (1972). [CrossRef]  

19. S. Tanabe, T. Ohyagi, N. Soga, and T. Hanada, “Compositional dependence of Judd-Ofelt parameters of Er3+ ions in alkali-metal borate glasses,” Phys. Rev. B Condens. Matter 46(6), 3305–3310 (1992). [CrossRef]   [PubMed]  

20. R. Xu, M. Wang, Y. Tian, L. Hu, and J. Zhang, “2.05 µm emission properties and energy transfer mechanism of germanate glass doped with Ho3+, Tm3+, and Er3+,” J. Appl. Phys. 109(5), 053503 (2011). [CrossRef]  

21. X. Wang, X. Fan, S. Gao, K. Li, and L. Hu, “Spectroscopic properties of Tm3+–Ho3+ codoped SiO2–Al2O3–CaO–SrO glasses,” Ceram. Int. 40(7), 9751–9756 (2014). [CrossRef]  

22. Y. Yan, A. J. Faber, and H. de Waal, “Luminescence quenching by OH groups in highly Er-doped phosphate glasses,” J. Non-Cryst. Solids 181(3), 283–290 (1995). [CrossRef]  

23. X. Feng, S. Tanabe, and T. Hanada, “Hydroxyl groups in erbium-doped germanotellurite glasses,” J. Non-Cryst. Solids 281(1-3), 48–54 (2001). [CrossRef]  

24. 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]  

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 (5)

Fig. 1
Fig. 1 IR transmittance spectrum of G4; the inset shows the absorption coefficients of all samples.
Fig. 2
Fig. 2 Subtracted absorption spectrum of G4.
Fig. 3
Fig. 3 Energy level diagram of Ho3+ in the lead-germanate glass; the dash lines represent the non-radiative processes, the solid lines stand for absorption (red line) and emission (A31 and A21) transition of Ho3+ ions, and the solid curve means the energy transfer from Ho3+ to OH groups.
Fig. 4
Fig. 4 (a) Normalized emission spectra of G1 to G6; (b) overlaps of absorption and emission cross-section spectra of the 2 μm band; (c) ( τ m 1 τ r 1 ) (black squares) and integrated 2 μm emission intensity (red squares) as a function of Ho3+ ions; τ m and τ r are the measured and calculated lifetimes of 5I7 level, respectively; (d) fluorescence decay curves of Ho3+ doped germanate glasses;.
Fig. 5
Fig. 5 Normalized infrared emission spectra of 4 wt% Ho2O3 doped germanate unclad fibers with different fiber length under the excitation of 1940 nm fiber laser.

Tables (2)

Tables Icon

Table 1 Radiative transition rates (Ar), branching ratios (β), and radiative lifetimes (τr) for different energy levels

Tables Icon

Table 2 Doping concentrations (NHo), energy transfer rate to OH groups (WOH), quantum efficiency (η), measured lifetimes (τm) at 5I7 level, and the product of τ m × NHo of Ho3+ in G1 to G6 glasses.

Equations (10)

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

α OH = 1 l ln( T 0 T )
N OH = N A εl ln( T 0 T )
S ed (aJ,b J )= t=2,4,6 Ω t | aJ U (t) b J | 2
band κ( λ ) dλ= 8 π 3 e 2 λ ¯ N 3ch(2J+1) ( n 2 +2) 9n S ed (aJ,b J )
σ em = λ 4 A r 8πc n 2 λI(λ) λI( λ )dλ  
τ m 1 = τ r 1 + W 21 + W OH
W OH = 9 2π N H o 3+ (α N OH ) τ r N 0 2
R 0 6 = 6c τ r 16 π 4 n 2 g low D g up D σ emi D ( λ ) σ abs A ( λ )dλ
τ m -1 - τ r -1 = W 21 + 9 2π N H o 3+ (α N OH ) τ r N 0 2
G=exp[ σ g σ a N τ m l]
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.