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

Intense O + E + S-band emission from Pr3+-doped ZnF2-based glasses

Open Access Open Access

Abstract

Pr3+-doped ZnF2-based glasses were prepared by using a melt-quenching method in dry N2 atmosphere. Under the excitation of a 588 nm light emitting diode (LED), ultrabroadband emissions ranging from 1245 to 1640 nm were obtained from the Pr3+-doped ZnF2-based glasses, which originate from the transitions 1D21G4 (producing E + S-band emission) and 1G43H5 (producing O-band emission) of Pr3+. The shape of the emission spectra could be tailored by varying the concentration of Pr3+. Emission spectra with the maximum full width at half maximum (FWHM) of 215 nm (1289 nm-1504 nm, covering the O + E + S-band) was obtained in the ZnF2-based glass at a doping concentration of 5000 ppm. The effects of the phonon energy of the matrix on O + E + S-band emission were also investigated. Our results showed that Pr3+-doped ZnF2-based glasses with low phonon energy might be used for constructing O + E + S-band lasers and optical amplifiers.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Wideband optical fiber amplifiers are essential for increasing the scale and performance of communications systems [1,2]. Among optical fiber amplifiers, Pr3+-doped fluoride fibers are used as the gain media for realizing O-band optical amplifiers in which a 1017 nm laser diode is used as the pump source, since the transition 1G43H5 of Pr3+ can produce O-band emissions and low phonon energy of fluoride fibers is required for reducing the non-radiative transitions from the 1G4 level to the neighbored levels [3]. At present, Pr3+-doped fluoride fiber amplifiers have already been commercial and widely used in optical transmission systems [4,5]. Recently, much efforts have been devoted to further expand the bandwidth of Pr3+-doped optical amplifiers. In 2011, B. Zhou et al. reported broadband near-infrared emission ranging from 1250 to 1680 nm in Pr3+-doped bismuth gallate glasses excited at 488 nm or 595 nm, in which the transitions 1D21G4 and 1G43H5 of Pr3+ produce such broadband emissions [6]. In 2012, B. Zhou et al. demonstrated broadband emission extending from 1.3 to 1.67 µm in Pr3+-doped fluorotellurite glasses under 488 nm or 590 nm excitation [7]. In 2016, J. Pisarska et al. investigated influence of BaF2 and activator concentration on broadband near-infrared luminescence of Pr3+ ions in gallo-germanate glasses for achieving broadband near-infrared amplifiers [8]. Although broadband near-infrared emission can be obtained by using the transitions 1D21G4 and 1G43H5 of Pr3+, the efficiency of the transition 1G43H5 is largely affected by the phonon energy of the matrix since the energy difference between the 1G4 level to the neighbored state 3F4 is just 3000 cm-1 and the nonradiative transition between them can quench the O-band emission [1]. It indicates that efficient broadband near-infrared emission might be obtained in Pr3+-doped fluoride glasses with low phonon energy. Among fluoride glasses, ZnF2-based glasses have the relatively low phonon energy (∼ 420 cm-1) [9]. However, until now, broadband near-infrared emission properties in Pr3+-doped ZnF2-based glasses have not yet been investigated, and the effects of the phonon energy of the matrix on near-infrared emission of Pr3+ were also not clarified.

In this paper, we reported intense O + E + S-band emission in Pr3+-doped ZnF2-based glasses and the effects of the phonon energy of the matrix on the O + E + S-band emission of Pr3+. Pr3+ doped ZnF2-based glasses were prepared by using a melt-quenching method in a glove box filled with dry N2. With the excitation of a 588 nm LED, broadband emissions from 1245 to 1640 nm were observed in the glasses, which originate from the transitions 1D21G4 and 1G43H5 of Pr3+. With an increase of the doping concentration of Pr3+, the emissions in E + S-band were quenched by the cross relaxation process [1D2, 3H4]→[1G4,3F4,3], and the emission spectrum with a full width at half maximum (FWHM) of 215 nm (1289 nm-1504 nm) was obtained in the glass at a doping concentration of 5000 ppm. In addition, a comparison of emission spectra of Pr3+-doped ZnF2-based (the phonon energy: ∼ 420 cm-1), InF3-based (∼ 507 cm-1) and ZrF4-based (∼ 579 cm-1) glasses at a same doping concentration of 5000 ppm was given, the emissions in O-band increased with a decrease of the phonon energy of the glass due to the reduction of the nonradiative transition from 1G4 level to the neighbored state 3F4, which is beneficial for achieving broadband O + E + S-band emission. The above results indicated that Pr3+-doped ZnF2-based glasses might be used as the gain medium for constructing O + E + S-band lasers and optical amplifiers.

2. Experiments and results

In our experiments, Pr3+-doped ZnF2-based glasses with a composition of 60ZnF2-15BaF2-10SrF2-(15-x)YF3-xPrF3 (x = 0, 0.1, 0.3, 0.5, 0.7, 1, 2 and 3, respectively) were fabricated by using a conventional melt-quenching method in a glove box filled with dry N2. The detailed procedures could be seen in [10]. The starting materials were high-purity (99.99%) anhydrous fluorides. The mixed batches were melted in a platinum crucible at 950 °C for 2 h. Then the glass melts were injected into a copper mold and annealed near the glass transition temperature for 5 h. The glass samples were cut and polished to 10 × 5 × 0.43 mm for optical measurements. By using a Seiko TG/DTA6200 analyzer, the transition temperature (Tg) and the onset crystallization temperature (Tx) of the ZnF2-based glass were measured to be ∼315 and ∼369 °C, respectively.

Figure 1 shows the transmission spectrum of Pr3+-doped ZnF2-based glass at a doping concentration of 30000 ppm (or 3 mol%) measured with a Shimadzu UV3600 spectrometer (180-2500 nm) and Nicolet 6700 FTIR spectrophotometer (2500-14000 nm). ZnF2-based glass has a transmission window from 200 to 12000nm, which is wider than that of InF2-based, ZrF4-based and AlF3-based glasses, due to the relatively low phonon energy (∼ 420 cm-1) of ZnF2-based glass. The absorption peaks at 442, 467, 480, 588, 1015, 1442, 1535, 1942, 2250 and 4705 nm are caused by the transitions from the ground level 3H4 to highly excited levels 3P2, 1I6, 3P0, 1D2, 1G4, 3F4, 3F3, 3F2, 3H6 and 3H5 of Pr3+ ions, respectively [11]. In addition, no obvious hydroxyl (OH) absorption at ∼3 µm was observed in the transmission spectrum of the ZnF2-based glass, which showed a quite low content of residual OH absorption in the ZnF2-based glass. Such a relatively low OH content is good for obtaining efficient near-infrared emissions from Pr3+-doped ZnF2-based glasses since the existence of OH may quench near-infrared emissions of Pr3+ ions significantly.

 figure: Fig. 1.

Fig. 1. Optical transmittance spectrum of Pr3+-doped ZnF2-based glass (30000 ppm). Inset shows the detail of the transmittance bands from 320 to 5600 nm.

Download Full Size | PDF

Figure 2 shows the experimental setup used in the experiments for measuring the emission spectra from the Pr3+ doped ZnF2-based fluoride glasses. A 588 nm LED was used as the pump source. The pump light was launched to the surface of the Pr3+ doped ZnF2-based fluoride glasses through a couple of silica lens, and the emission spectra were recorded by using a grating spectrometer (SPEX 1000 M, Horiba) equipped with an InGaAS avalanche photodiode (APD, Hamamatsu). To avoid the effect of high order diffraction, a 1050 nm long pass filter was used for measuring the emission spectra in the range of 1100-1650 nm. Figures 3(a) and 3(b) show the measured emission spectra of Pr3+-doped ZnF2-based glasses with various doping concentration. The energy level diagram of Pr3+ ions is shown in Fig. 5. As the doping concentration of Pr3+ is ∼ 1000 ppm, several emission peaks at 800, 840, 1026, 1335 and 1464 nm were observed in Pr3+-doped ZnF2-based glass, which were attributed to the transitions 1D23H6, 1D23F2, 1D23F3,4, 1G43H5 and 1D21G4 (shown in Fig. 5), respectively [12]. Figure 3(c) shows the dependence of the intensity of these emission bands at 800, 840, 1026, 1335 and 1464 nm on the doping concentration of Pr3+ ions. With an increase of the doping concentration of Pr3+ ions, the intensity of the emission bands at 800, 840, 1026 and 1464 nm decreased monotonically owing to the occurrence of the cross relaxation process [1D2, 3H4]→[1G4,3F4,3], which depopulated the energy level 1D2 of Pr3+ ions and made the transitions from the energy level 1D2 to the other energy levels quenched significantly [7]. In addition, the concentration quenching and reabsorption effects in high concentration Pr3+ doped ZnF2-based fluoride glasses can also result in the decrease of the intensity for emissions at 800, 840, 1026 and 1464 nm. Interestingly, the intensity of the emission band at 1335 nm increased with an increase of the doping concentration of Pr3+ ions from 1000 to 5000 ppm due to the occurrence of the cross relaxation process [1D2, 3H4]→[1G4,3F4,3], which populated the energy level 1G4 of Pr3+ ions and further made the intensity of the emission band at 1335 nm (the transition 1G43H5) enhanced (shown in Fig. 5). However, as the doping concentration of Pr3+ ions was larger than 5000 ppm, the intensity of the emission band at 1335 nm decreased due to the occurrence of the cross relaxation process [1G4, 1G4]→[3H5,1D2] and other processes (such as the concentration quenching and reabsorption processes), which depopulated the energy level 1G4 of Pr3+ ions and further made the intensity of the emission band at 1335 nm decreased (shown in Fig. 5) [13]. The above results showed that the shape of the emission spectra in O + E + S-band could be tailored by varying the concentration of Pr3+ ions. Figure 3(d) shows the dependence of the FWHM of the emission bands at 1335 and 1464 nm on the doping concentration of Pr3+ ions. As the doping concentration of Pr3+ ions was ∼ 5000 ppm, emission spectra with the FWHM of 215 nm (1289 nm-1504 nm) was obtained in the Pr3+-doped ZnF2-based glass, which almost covered O + E + S-band. Those results indicated that broadband O + E + S-band emissions could be obtained in Pr3+-doped ZnF2-based glasses. Lifetimes of the 1D2 and 1G4 levels of Pr3+ doped in ZnF2-based glasses were also measured by monitoring the emissions at 1464 and 1335 nm, shown in Fig. 4. An optical parametric oscillator (OPO) nanosecond pulsed laser (Continuum SLIII-EX) was used as the pump source (588 and 980 nm were used). For the ZnF2-based glass with a Pr3+ doping concentration of 1000 ppm, the measured lifetime of the 1D2 and1G4 levels were about 354 and 183 µs, respectively, which are much longer than that obtained in Pr3+ doped gallo-germanate glasses (∼110 µs for the 1D2 level) [8] and Pr3+ doped ZrF4-based glasses (∼110 µs for the 1G4 level) [13]. With the increase of Pr3+ doping concentration to 30000 ppm, the lifetimes of the 1D2 and1G4 levels were dramatically decreased to about 4 and 3 µs, respectively.

 figure: Fig. 2.

Fig. 2. Schematic diagram of experimental setup for measuring the emission spectra of Pr3+ doped ZnF2-based fluoride glasses.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. Measured emission spectra in the range (a) 750-900 nm and (b) 900-1650 nm of Pr3+-doped ZnF2-based glasses with various doping concentration. (c) Dependence of the intensity of these emission bands at 800, 840, 1026, 1335 and 1464 nm on the doping concentration of Pr3+ ions. (d) Dependence of the FWHM of the emission bands at 1335 and 1464 nm on the doping concentration of Pr3+ ions.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Dependence of the measured lifetimes of the 1D2 and1G4 levels on the Pr3+ doping concentration in ZnF2-based glasses.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Energy level diagram of Pr3+ ions in the ZnF2-based glass.

Download Full Size | PDF

Furthermore, according to the Füchtbauer-Ladenburg (FL) theory and the emission spectra in O + E + S-band of Pr3+-doped ZnF2-based glass, the stimulated emission cross sections (σemi) are calculated by the following equation [14]:

$${\sigma _{emi}} = \frac{{{\lambda ^4}A}}{{8\pi c{n^2}}} \times \frac{{\lambda I(\lambda )}}{{\int {\lambda I(\lambda )d\lambda } }}$$
where λ is the wavelength, A is the radiative transition probability, I(λ) is the relative intensity of the emission spectrum, c and n are the speed of light in vacuum and the refractive index of the glass, respectively. The A can be calculated based on the absorption spectrum of Pr3+-doped ZnF2-based glass and the reduced matrix elements for the transitions of Pr3+ ions by using the Judd-Ofelt theory [15,16]. Table 1 lists out the calculated A and branching ratio (β) for the transitions from the energy levels 1D2 and 1G4 to lower levels of Pr3+ ions. The A for the transitions 1G43H5 and 1D21G4 of Pr3+ ions were calculated to be 201.4 and 179.7 s-1. Based on the parameters mentioned above, the stimulated emission cross section at 1335 and 1464 nm was calculated to be 3.18 × 10−21 and 3.63 × 10−21 cm2, respectively, as shown in Figs. 6(a) and 6(b), which is comparable with that of ZBLAN glasses (3.5 × 10−21 cm2 for the 1310 nm emission) [17]. Based on the stimulated emission cross section, the stimulated absorption cross section (σabs) can be obtained by using McCumber equation as follows [18]:
$${\sigma _{abs}}(\lambda ) = {\sigma _{emi}} \times \frac{{{Z_U}}}{{{Z_L}}}{\exp ^{\frac{{hc{\lambda ^{ - 1}} - \Delta E}}{{{k_B}T}}}}$$
where ΔE is the energy gap between the lowest Stark level of the upper and lower states, h is the Planck’s constant, kB is the Boltzmann constant, T is the temperature, ZL and ZU are the partition functions of the lower and upper manifolds, calculated by using the method proposed by Miniscalco and Quimby [19]. The calculated stimulated absorption cross section at 1284 and 1371 nm was calculated to be 1.81 × 10−21 and 2.24 × 10−21 cm2 (shown in Figs. 6(a) and 6(b)), respectively. By using the calculated σabs and σemi of Pr3+ ions, the gain cross sections can be calculated by the following equation [20]:
$$G(\lambda ) = P\mathrm{\ \times }{\sigma _{\textrm{em}i}}(\lambda ) - (1 - P)\mathrm{\ \times }{\sigma _{abs}}(\lambda )$$
where the population inversion P is assigned to the concentration ratio of Pr3+ ions in the 1D2 and 1G4 levels (or in the 1G4 and 3H5 levels). Figures 6(c) and 6(d) show the calculated gain cross sections for the 1G43H5 and 1D21G4 transitions of Pr3+ ions in ZnF2-based glasses, respectively. As P > 0.2, positive gain in O + E + S-band could be obtained by using the transitions 1D21G4 and 1G43H5 of Pr3+ ions in ZnF2-based glasses, which indicated that Pr3+-doped ZnF2-based glasses might be used as the gain medium for constructing O + E + S-band lasers and optical amplifiers.

Tables Icon

Table 1. Calculated radiative transition probability (A) and branching ratio (β) for the transitions from the energy levels 1D2 and 1G4 to lower levels of Pr3+ ions.

 figure: Fig. 6.

Fig. 6. Calculated absorption and emission cross sections of (a) Pr3+: 3H51G4 and (b) Pr3+: 1G41D2 transitions. (c) Calculated gain cross sections of Pr3+: 3H51G4 and (d) Pr3+: 1G41D2 transitions in ZnF2-based glass. From bottom to top, the population inversion ratios (P) are 0.0, 0.2, 0.4, 0.6, 0.8 and 1.0, respectively.

Download Full Size | PDF

In addition, to investigate the effects of the phonon energy of the glass on near-infrared emission of Pr3+ ions, Pr3+-doped InF3-based and ZrF4-based glasses with a composition of 15ZnF2-25.5InF3-11.5GaF3-18BaF2-8SrF2-5LiF-(2.5-0.5x)LaF3-(2.5-0.5x)YF3-12PbF2-xPrF3 and 50ZrF4-33BaF2-7AlF3-(10-x)YF3-xPrF3 (x = 0.5) were fabricated by using a conventional melt-quenching method in a glove box filled with dry N2. The glass samples were cut and polished to 10 × 10 × 2 mm for optical measurements. Figure 7(a) shows the measured Raman scattering spectra of undoped ZnF2-based, InF3-based and ZrF4-based glasses. The Raman scattering spectra of the fluoride glasses were measured with a confocal Raman microscope (LabRAM HR Evolution, HORIBA), and a 633 nm laser was selected as the pump source. The Raman shift for main Raman peaks of ZnF2-based, InF3-based and ZrF4-based glasses were 420, 507 and 579 cm-1, respectively, which correspond to the phonon energy of those three types of fluoride glasses [21]. Then we measured the near-infrared emission spectra of Pr3+ ions in ZnF2-based, InF3-based and ZrF4-based glasses with a same doping concentration of 5000 ppm under the excitation of a same 588 nm LED and the measured results were shown in Figs. 7(b) (normalized at 1464 nm). Two emission peaks at 1335 and 1464 nm were all observed in Pr3+-doped ZnF2-based, InF3-based and ZrF4-based glasses, which were attributed to the transitions 1G43H5 and 1D21G4 (shown in Fig. 5), respectively. Interestingly, with a decrease of the phonon energy of the glass, the relative intensity of the emission peak at 1335 nm (the transition 1G43H5) increased significantly since low phonon energy is beneficial for reducing the nonradiative transition from the 1G4 level to the neighbored state 3F4 and further make the intensity of the emission band at 1335 nm enhanced significantly [22,23]. In the future, we will systematically investigate the effect of cross relaxation, concentration quenching and pumping intensity on the emission intensities and intensity ratio between 1335 nm and 1464 nm. The above results showed that Pr3+-doped ZnF2-based glasses with low phonon energy are preferable for constructing O + E + S-band lasers and optical amplifiers.

 figure: Fig. 7.

Fig. 7. (a) Measured Raman scattering spectra of undoped ZnF2-based, InF3-based and ZrF4-based glasses. (b) Normalized near-infrared emission spectra of Pr3+ ions in ZnF2-based, InF3-based and ZrF4-based glasses with a same doping concentration of 5000 ppm under the excitation of a same 588 nm LED.

Download Full Size | PDF

3. Conclusions

In conclusion, we reported intense O + E + S-band emission in Pr3+-doped ZnF2-based glasses. With the excitation of a 588 nm LED, broadband emissions from 1245 to 1640 nm were observed in the glasses, which originated from the transitions 1D21G4 and 1G43H5 of Pr3+. The shape of the emission spectra could be tailored by varying the concentration of Pr3+. Emission spectra with the FWHM of 215 nm (1289 nm-1504 nm) was obtained in the ZnF2-based glass at a doping concentration of 5000 ppm. The effects of the phonon energy of the matrix on O + E + S-band emission were also investigated. Our results showed that Pr3+-doped ZnF2-based glasses with low phonon energy might be used as the gain medium for constructing O + E + S-band lasers and optical amplifiers.

Funding

National Key Research and Development Program of China (2020YFB1805800); National Natural Science Foundation of China (62090063, 62075082, U20A20210, 61827821, U22A2085, 62235014, 62205121); the Opened Fund of the State Key Laboratory of Integrated Optoelectronics.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

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. M. Yamada, “Overview of wideband optical fiber amplification technologies,” NTT Technical Review 2(12), 34–55 (2004).

2. Z. Li, A. M. Heidt, N. Simakov, et al., “Diode-pumped wideband thulium-doped fiber amplifiers for optical communications in the 1800–2050nm window,” Opt. Express 21(22), 26450–26455 (2013). [CrossRef]  

3. M. Yamada, T. Kanamori, Y. Ohishi, et al., “Pr3+-doped fluoride fiber amplifier module pumped by a fiber coupled master oscillator/power amplifier laser diode,” IEEE Photonics Technol. Lett. 9(3), 321–323 (1997). [CrossRef]  

4. https://www.fiberlabs.com/glossary/praseodymium-doped-fiber-amplifier/.

5. A. Ferrari, A. Napoli, J. K. Fischer, et al., “Assessment on the achievable throughput of multi-band ITU-T G.652.D fiber transmission systems,” J. Lightwave Technol. 38(16), 4279–4291 (2020). [CrossRef]  

6. B. Zhou and E. Y. Pun, “Superbroadband near-IR emission from praseodymium-doped bismuth gallate glasses,” Opt. Lett. 36(15), 2958–2960 (2011). [CrossRef]  

7. B. Zhou, L. Tao, Y. H. Tsang, et al., “Superbroadband near-IR photoluminescence from Pr3+-doped fluorotellurite glasses,” Opt. Express 20(4), 3803–3813 (2012). [CrossRef]  

8. J. Pisarska, M. Kowal, M. Kochanowicz, et al., “Influence of BaF2 and activator concentration on broadband near-infrared luminescence of Pr3+ ions in gallo-germanate glasses,” Opt. Express 24(3), 2427–2435 (2016). [CrossRef]  

9. S. Jia, C. Li, Z. Zhao, et al., “Er3+-doped ZnF2-BaF2-SrF2-YF3 fluoride glasses for 2.7, µm laser applications,” Mater. Lett. 227, 97–99 (2018). [CrossRef]  

10. Z. Zhao, S. Jia, H. He, et al., “Study on 2.86 µm emission properties of Ho3+-doped low-hydroxyl ZnF2 based fluoride glasses for mid-infrared laser applications,” Opt. Mater. Express 10(12), 3137–3144 (2020). [CrossRef]  

11. W. Seeber, E. A. Downing, L. Hesselink, et al., “Pr3+-doped fluoride glasses,” J. Non-Cryst. Solids 189(3), 218–226 (1995). [CrossRef]  

12. V. Naresh and B. S. Ham, “Influence of multiphonon and cross relaxations on 3P0 and 1D2 emission levels of Pr3+ doped borosilicate glasses for broad band signal amplification,” J. Alloys Compd. 664, 321–330 (2016). [CrossRef]  

13. Y. Ohishi, T. Kanamori, T. Nishi, et al., “Concentration effect on gain of Pr3+-doped fluoride fiber for 1.3 µm amplification,” IEEE Photonics Technol. Lett. 4(12), 1338–1341 (1992). [CrossRef]  

14. X. Wen, G. Tang, Q. Yang, et al., “Highly Tm3+ doped germanate glass and its single mode fiber for 2.0 µm laser,” Sci. Rep. 6(1), 20344 (2016). [CrossRef]  

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

16. G. S. Ofelt, “Intensities of Crystal Spectra of Rare-Earth Ions,” J. Chem. Phys. 37(3), 511–520 (1962). [CrossRef]  

17. Y. Miyajima, T. Sugawa, and Y. Fukasaku, “38.2 dB amplification at 1.31 µm and possibility of 0.98 µm pumping in Pr3+-doped fluoride fibre,” Electron. Lett. 27(19), 1706–1707 (1991). [CrossRef]  

18. D. E. McCumber, “Theory of phonon-terminated optical masers,” Phys. Rev. 134(2A), A299–A306 (1964). [CrossRef]  

19. W. J. Miniscalco and R. S. Quimby, “General procedure for the analysis of Er3+ cross sections,” Opt. Lett. 16(4), 258–260 (1991). [CrossRef]  

20. J. Wang, Z. Jia, C. Zhang, et al., “Thulium-doped fluorotellurite glass fibers for broadband S-band amplifiers,” Opt. Lett. 47(8), 1964–1967 (2022). [CrossRef]  

21. C. Zhang, K. Han, D. Zhou, et al., “High concentration Tm3+ doped TeO2-Al2O3-BaF2 glass for ∼ 2 µm fiber lasers,” J. Alloys Compd. 901, 163592 (2022). [CrossRef]  

22. Y. Nishida, T. Kanamori, Y. Ohishi, et al., “Efficient PDFA module using PbF2/InF3-based fluoride fiber,” IEEE Photonics Technol. Lett. 9(3), 318–320 (1997). [CrossRef]  

23. D. W. Hewak, R. S. Deol, J. Wang, et al., “Low phonon-energy glasses for efficient 1.3 µm optical fibre amplifiers,” Electron. Lett. 29(2), 237–239 (1993). [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 (7)

Fig. 1.
Fig. 1. Optical transmittance spectrum of Pr3+-doped ZnF2-based glass (30000 ppm). Inset shows the detail of the transmittance bands from 320 to 5600 nm.
Fig. 2.
Fig. 2. Schematic diagram of experimental setup for measuring the emission spectra of Pr3+ doped ZnF2-based fluoride glasses.
Fig. 3.
Fig. 3. Measured emission spectra in the range (a) 750-900 nm and (b) 900-1650 nm of Pr3+-doped ZnF2-based glasses with various doping concentration. (c) Dependence of the intensity of these emission bands at 800, 840, 1026, 1335 and 1464 nm on the doping concentration of Pr3+ ions. (d) Dependence of the FWHM of the emission bands at 1335 and 1464 nm on the doping concentration of Pr3+ ions.
Fig. 4.
Fig. 4. Dependence of the measured lifetimes of the 1D2 and1G4 levels on the Pr3+ doping concentration in ZnF2-based glasses.
Fig. 5.
Fig. 5. Energy level diagram of Pr3+ ions in the ZnF2-based glass.
Fig. 6.
Fig. 6. Calculated absorption and emission cross sections of (a) Pr3+: 3H51G4 and (b) Pr3+: 1G41D2 transitions. (c) Calculated gain cross sections of Pr3+: 3H51G4 and (d) Pr3+: 1G41D2 transitions in ZnF2-based glass. From bottom to top, the population inversion ratios (P) are 0.0, 0.2, 0.4, 0.6, 0.8 and 1.0, respectively.
Fig. 7.
Fig. 7. (a) Measured Raman scattering spectra of undoped ZnF2-based, InF3-based and ZrF4-based glasses. (b) Normalized near-infrared emission spectra of Pr3+ ions in ZnF2-based, InF3-based and ZrF4-based glasses with a same doping concentration of 5000 ppm under the excitation of a same 588 nm LED.

Tables (1)

Tables Icon

Table 1. Calculated radiative transition probability (A) and branching ratio (β) for the transitions from the energy levels 1D2 and 1G4 to lower levels of Pr3+ ions.

Equations (3)

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

σ e m i = λ 4 A 8 π c n 2 × λ I ( λ ) λ I ( λ ) d λ
σ a b s ( λ ) = σ e m i × Z U Z L exp h c λ 1 Δ E k B T
G ( λ ) = P   × σ em i ( 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.