Next Article in Journal
Influence of the Fly Ash Content on the Fresh and Hardened Properties of Alkali-Activated Slag Pastes with Admixtures
Next Article in Special Issue
Effect of Se on Structure and Electrical Properties of Ge-As-Te Glass
Previous Article in Journal
The Effect of the Substrate on the Microstructure and Tribological Properties of Cold Sprayed (Cr3C2-25(Ni20Cr))-(Ni-graphite) Cermet Coatings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Temperature Dependence of Absorption and Energy Transfer Efficiency of Er3+/Yb3+/P5+ Co-Doped Silica Fiber Core Glasses

1
Key Laboratory of High Power Laser Materials, Shanghai Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Shanghai 201800, China
2
Hangzhou Institute for Advanced Study, University of Chinese Academy of Sciences, Hangzhou 310024, China
3
Department of Physics, Shanghai University, Shanghai 200444, China
*
Authors to whom correspondence should be addressed.
Submission received: 24 December 2021 / Revised: 22 January 2022 / Accepted: 23 January 2022 / Published: 27 January 2022
(This article belongs to the Special Issue Advanced Functional Glass: Preparation, Properties, and Applications)

Abstract

:
A high phosphorus Er3+/Yb3+ co-doped silica (EYPS) fiber core glass was prepared using the sol-gel method combined with high-temperature sintering. The absorption spectra, emission spectra, and fluorescence decay curves were measured and compared in temperatures ranging from 300 to 480 K. Compared to 915 and 97x nm, the absorption cross-section at ~940 nm (~0.173 pm2) demonstrates a weaker temperature dependence. Hence, the 940 nm pump mechanism is favorable for achieving a high-power laser output at 1.5 μm. Additionally, the double-exponential fluorescence decay of Yb3+ ions and the emission intensity ratio of I1018nm/I1534nm were measured to evaluate the energy transfer efficiency from Yb3+ ions to Er3+ ions. Through the external heating and active quantum defect heating methods, the emission intensity ratios of I1018nm/I1534nm increase by 30.6% and 709.1%, respectively, from ~300 to ~480 K. The results indicate that the temperature rises significantly reduce the efficiency of the energy transfer from the Yb3+ to the Er3+ ions.

1. Introduction

In recent years, eye-safe 1.5 μm high-power fiber lasers have been extensively used in free-space communication, LIDAR, and range finding, which has attracted significant attention [1,2,3,4,5]. The 1.5 μm laser is mainly based on the 4I13/2 to 4I15/2 transition of the Er3+ ions. In silica fiber, the Er3+ ions have small absorption cross-sections at ~980 nm, and a serious concentration quenching effect [6,7]. The current output power record of the single-mode Er3+-doped silica fiber amplifier is only 107 W [8]. To increase the output power further, doping the Yb3+ ions can effectively improve the pump absorption coefficient [1,9]. Additionally, doping high-concentration phosphorus is also necessary for Er3+/Yb3+ co-doped silica fiber (EYDF) [10,11]. This has two advantages, namely, an improved dispersion effect of the Er3+ and Yb3+ ions [12], and a weak back energy transfer (ET) from the Er3+ to Yb3+ ions in the high phonon energy environment [13,14]. The current power record of the single-mode EYDF is 302 W, which has an optical efficiency of 56%. The gain fiber (Nufern-LMA-EYDF-25P/300-HE) was doped with ~11mol% P2O5 [10]. At present, two primary factors limit the increase in power of the EYDF: first, the significant thermal effect caused by the quantum defect (QD) [10,15,16], which reduces the pump absorption and damages the fiber coating [17,18,19]; second, the amplified spontaneous emission (ASE) and parasitic oscillations of the Yb3+ ions at ~1 μm, reducing the efficiency of the ET from the Yb3+ to the Er3+ ions [11]. Both of the two effects directly affect the 1.5 μm laser performance.
Generally, a ~915 nm, ~940 nm, ~97x nm LD pump, or a 1018 nm fiber laser, can be used as the pump source of the EYDF [20]. The absorption coefficients of the EYDF at ~915, ~940, ~97x, and ~1018 nm are significantly different, and are affected by the composition of the fiber core glass. The low pump absorption can achieve a smoother population inversion distribution along the EYDF [20], thereby suppressing the significant thermal effect and leading to a Yb-ASE gain [6,10,21,22]. For the high phosphorus-doped EYDF, the absorption spectrum at 940 nm is flatter than 915 nm and 97x nm [22,23]. In summary, the ~940 nm pump is a more practical scheme for achieving a high-power EYDF amplifier.
Several studies have been conducted on the doping concentration and doping ratio of the Yb3+ and Er3+ ions in silica glass [24]. The results show that when the doping concentration of the Yb3+ ions reaches 3–5 wt%, and the molar concentration ratio of Yb3+/Er3+ ions is 8–15, the EYDF exhibits a better laser performance [10,24,25,26]. However, the strong pump absorption of the Yb3+ ions inevitably causes the temperature of the glass to rise. Due to the limitation of fiber polymer coatings, the operating temperature of fiber laser is required to be lower than ~200 °C [27]. Only a few studies have compared the change of primary pump absorption cross-sections and the efficiency of the ET from the Yb3+ to the Er3+ ions with the temperature rise of the EYDF amplifier [28]. In this work, the Er3+/Yb3+ co-doped with ~15 mol% P2O5 silica glass was prepared using the sol-gel method combined with high-temperature sintering for the first time. The temperature dependence (from 300 to 480 K) of the primary absorption cross-sections (at 915, 940, 974, and 1018 nm) was calculated and analyzed. Additionally, the fluorescence decay of the Yb3+ ions and the emission intensity ratio of I1018nm/I1534nm were measured to evaluate the efficiency of the ET from the Yb3+ to the Er3+ ions under two rising temperature conditions. The results provide sufficient data supporting the experimental design and modeling of the EYDF amplifier.

2. Experiment Details

2.1. Sample Preparation

The chemical composition of the high phosphorus-doped silica core glass is presented in Table 1. Both the EYPS and Yb3+/P5+-doped silica (YPS) bulk glasses were prepared using the sol-gel method, combined with high-temperature sintering, with an excellent refractive index uniformity [29]. TEOS, ErCl3·6H2O, YbCl3·6H2O, H3PO4, and C2H5OH were used as precursors. All precursors were mixed in pure water and stirred at ~30 °C to form a homogeneous rare earth-doped sol. After heating to 1000 °C, a gel powder with decomposed hydroxyl and organic matter was obtained. Transparent bulk glasses were prepared after the dry powder was melted at 1650 °C for 2 h in a vacuum state. To simulate the fiber drawing process, the bulk glasses were heat-treated in a hydrogen–oxygen flame at ~2000 °C and then rapidly cooled in air. The detailed preparation process is described in Refs. [30,31]. The thickness of the bulk glasses was polished to 1 mm. The EYPS and YPS bulk glasses contain the same concentrations of P2O5 and Yb2O3, which are 15.0mol% and 1.0 mol%, respectively. The actual concentrations of Yb2O3, Er2O3, and P2O5 were measured using a Thermo iCAP 6300 radial view inductively coupled plasma optical emission spectrometry (Thermo Fisher Scientific, Waltham, MA, USA). The decrease in phosphorus content is due to the volatilization during high-temperature sintering.

2.2. Characterization of the Bulk Glass

The absorption spectrum was measured using a Lambda UV-VIS-NIR spectrophotometer with a scanning step of 0.1 nm. The temperature values were set to 300, 340, 380, 420, and 480 K using the temperature controller shown in Figure 1a. As shown in Figure 1b, a high-resolution Edinburgh FLS 920 Spectrofluorometer (Edinburgh Instruments, Wales, UK). was used to measure the near-infrared emission spectrum under 915 nm pulse excitation at 300–480 K. Figure 1b shows two methods for temperature control in the vacuum chamber: ① external heating of the EYPS bulk glass using the temperature controller; ② active quantum defect heating of the EYPS bulk glass by increasing the 915 nm LD pump power, with the temperature controller turning off. The temperature was detected in real time using a Fluke Ti400 (Fluke Corporation, Everett, WA, USA) infrared thermal imager. The fluorescence decay curves of Er3+:4I13/2 and Yb3+:2F5/2 energy levels were measured at 1018 and 1534 nm under 980 nm pulse excitation, respectively. Additionally, the fluorescence decay with time evolution in the wavelength range of 1000–1700 nm under 980 nm pulse excitation was measured using an FLS920 Spectrofluorometer at ~300 K.

3. Results and Discussion

3.1. Efficient Energy Transfer of EYPS Bulk Glass

The ~15mol% P2O5 was introduced in the EYPS bulk glass, and the composition was close to that of the commercial erbium–ytterbium co-doped silica fiber (Nufern-LMA-EYDF-25P/300-HE) [10]. The high doping concentration of rare earth ions and the efficiency of the ET from the Yb3+ to the Er3+ ions were guaranteed. The fluorescence decay of the Yb3+ ions and emission intensity ratio of I~1μm/I1534nm in the EYPS glass can be used to evaluate the efficiency of the ET from the Yb3+ to the Er3+ ions [2,32]. Figure 2a shows the normalized emission spectrum of the EYPS under 915 nm pump excitation. The weak emission intensity in the range of 1–1.1 μm indicates the high efficiency of the ET from the Yb3+ to the Er3+ ions. Figure 2b shows the fluorescence decay curves of the Er3+ and Yb3+ ions in the EYPS at ~300 K. The fluorescence lifetime of the Er3+ ions is 8.99 ms at 1534 nm. However, the decay of the Yb3+ is double exponential at 1018 nm. The emission intensity at 1018 nm decreases to 1/e of the initial intensity at 8.14 μs. A short fluorescence lifetime, τ1 (8.14 μs), and a long fluorescence lifetime, τ2 (~1.3 ms), were obtained. This is attributed to the existence of two types of Yb3+ ions in the EYPS: those closely coupled to the Er3+ ions and those uncoupled from the Er3+ ions [11,33]. The distance between the Yb3+ and Er3+ ions determines whether the two can be coupled [24], and a detailed discussion is presented in Section 3.3. In summary, the study of spectral properties based on the EYPS bulk glass has a high reference value for high-power EYDF amplifiers.

3.2. Temperature Dependence of the Primary Absorption Cross-Sections of the EYPS

The high-power operation of the EYDF amplifier results in a high heat load, and the fiber core temperature may reach 500 K under air-cooling conditions [10,15,16,34], which limits the power increase. Figure 3a shows the absorption cross-section curves of the Yb3+ ions in the EYPS bulk glass at 300, 340, 380, 420, and 480 K. The calculation formula of the absorption cross-sections were described in detail in Ref. [17]. The ratios of the primary absorption cross-section at different temperatures to that at 300 K in the EYPS bulk glass are shown in Figure 3b. As the temperature increases from 300 to 480 K, the absorption cross-sections of the EYPS bulk glass decrease by 27.2% (0.173 pm2→0.126 pm2) at 915 nm and 37.9% (0.943 pm2→0.586 pm2) at 974 nm. The absorption cross-section at 1018 nm increases by 106.9% (0.029 pm2→0.060 pm2). As we all know, the 1018 nm pump wavelength has a high quantum conversion efficiency for the 1.5 μm laser. The 106.9% increase in the absorption cross-section significantly impacts the amplification using the 1018 nm co-band pump. The higher reabsorption of the fiber core glass at 1–1.1 μm can also suppress the Yb-ASE gain. Additionally, the absorption cross-section at 940 nm is 0.173 pm2, which has little change with temperature. Compared with 915 and 974 nm in Figure 3b, the absorption cross-section at ~940 nm demonstrates a weak temperature dependence simultaneously, which is conducive to the stable output of the high-power EYDF amplifier.
The reduction in the absorption cross-section is caused by the temperature-dependent electron–phonon interaction at the sub-level of the rare earth ions [17,35]. The particles in the ground state are redistributed with the temperature rise, and this process conforms to the Boltzmann distribution. In Equation (1), b1N represents the fraction of the population of the N-th Stark sub-energy level within the ground-state manifold (where N = 1 or 2; see Figure 4) relevant to the absorption transition [35], kb is the Boltzmann constant, T is the temperature, and E1NE11 is the difference in energy relative to the lowest Stark energy level [17]:
b 1 N ( T ) = exp [ ( E 1 N E 11 ) / k b T ] i = 1 4 exp [ ( E 1 i E 11 ) / k b T ]
Figure 4a,b show the absorption curves of the Yb3+ ions in EYPS bulk glass after the multiple peaks fitting at 300 and 480 K, respectively. For the convenience of discussion, the ground state 2F7/2 and the excited state 2F5/2 of the Yb3+ ions are simplified into two sub-levels and three sub-levels, respectively, as shown in the inset of Figure 4a. The dotted lines represent the decomposition peaks corresponding to different Stark transitions. The absorption transitions at ~915 and 974 nm correspond to 2F7/22F5/2 (b11→b23) and 2F7/22F5/2 (b11→b21), respectively. The absorption transitions at ~940 nm correspond to 2F7/22F5/2 (b11, b12→b22), and the transitions at ~1018 nm correspond to 2F7/22F5/2 (b12→b21). The proportions of the integrated area of the fitted peaks corresponding to b12→b22 and b12→b21 increase, while that of the integrated area involving b11 decreases. According to Equation (1), as the temperature rises from 300 to 480 K, the value of b11 decreases by 24.0%, and the value of b12 increases by 31.4%. These results are consistent with the changing trend of the above-mentioned integrated area of fitted peaks. Therefore, as the temperature increases, the absorption cross-sections at ~915 and 974 nm, which are determined by the value of b11, significantly decrease. The absorption cross-sections at ~940 nm, which are affected by both the value of b11 and b12, are maintained. The absorption cross-sections at ~1018 nm which are determined by the value of b12, significantly increase.

3.3. Analysis of the Temperature Dependence of the Efficiency of the Energy Transfer from Yb3+ to Er3+ Ions in EYPS

The efficiency of the ET from the Yb3+ to the Er3+ ions in the EYDF directly affects the output power, the ~1 μm ASE, and the ytterbium parasitic lasing [11]. The efficiency of the ET from the Yb3+ to the Er3+ ions can be evaluated using the fluorescence decay of the Yb3+ ions [32]. However, the decay of the Yb3+ ions is double exponential in Figure 2b. Therefore, the mechanism of the double-exponential decay was first probed. Figure 5a shows the fluorescence decay curves of the EYPS bulk glass, with the time evolution in the wavelength range of 1000–1700 nm under 980 nm pulse excitation at 300 K. During the pulse duration (~28 μs), the Er3+ ions have exhibited a strong emission intensity at 1534 nm, due to the high efficiency of the ET from the Yb3+ to the Er3+ ions. After the pulse is stopped, the emission intensity of the Yb3+ ions at ~1 μm rapidly decreases to 1/e of the initial intensity within ~10 μs, and the peak emission intensity of the Er3+ ions at 1534 nm continues to increase. Combined with Figure 2b, τ1 (8.14 μs) corresponds to the ET process from the Yb3+ ions to the Er3+ ions, and τ2 (~1.3 ms) corresponds to the spontaneous emission (SE) process of the Yb3+ ions [11,32,33]. The fluorescence decay of the Yb3+ ions is significantly affected by the distance between the Yb3+ and Er3+ ions in the matrix [33,36,37]. Most of the Yb3+ and Er3+ ions in the EYPS are concentrated around phosphorus [38]. Therefore, two types of Yb3+ ions are assumed: those coupled and those uncoupled to Er3+. Figure 5b shows the simplified model of the coupled and isolated Yb3+ ions in the EYPS bulk glass. The Yb3+ ions closely coupled with the Er3+ ions show ET and SE1 processes simultaneously, and the emission of the Yb3+ ions which are not close to the Er3+ ions decays during the SE2 processes.
ET : Er 3 + : F 15 / 2 4 + Yb 3 + : F 5 / 2 2 Er 3 + : F 13 / 2 4 + Yb 3 + : F 7 / 2 2 + Phonon
SE : Yb 3 + : F 5 / 2 2 Yb 3 + : F 7 / 2 2 + Photon
To further study the ET and SE processes in the high-power EYDF amplifier, the fluorescence decay of the Yb3+ and Er3+ ions in the EYDF was measured under a 980 nm excitation wavelength in 300 to 480 K in Figure 6a,b. The results show that the fluorescence lifetime of the Er3+ ions gradually decreases because of the increased probability of the non-radiative transition under temperature rise conditions. Table 2 is obtained by fitting the fluorescence decay of the Yb3+ ions according to Equation (2):
I ( t ) = I 0 + A 1 e t / τ 1 + A 2 e t / τ 2
In Equation (2), τ1 and τ2 correspond to the fluorescence lifetimes of the ET and SE2 processes of the Yb3+ ions, respectively. A1 and A2 are the relative proportions of the population participating in the ET and SE2, respectively. As the temperature increases, the probability of the non-radiative transition of the excited electron in the Yb3+ ions increases, and the population participating in the ET and SE2 processes decreases. Table 2 shows that the values of A1 and A2 are ~86% and ~14%, respectively. As the temperature increases, the value of A1 decreases, while that of A2 increases, indicating that the proportion of the SE2 process increases. The value of τ1 decreases, owing to the increase in the ET rate, which is assisted by multi-phonons after the temperature increases. The decrease in τ2 and τYb is due to the increase in the probability of non-radiative transition. In summary, as the temperature rises, the relative proportion (A2) of the Yb3+ ions involved in the SE2 process increases. The increase in A2 indicates that the relative emission intensity at ~1 μm increases, and the efficiency of the ET from the Yb3+ to the Er3+ ions decreases.
Subsequently, the emission intensity ratio of I1018nm/I1534nm was also measured to evaluate the effect of temperature on the efficiency of the ET from the Yb3+ to the Er3+ ions. Figure 6c shows that the fluorescence spectra of the EYPS bulk glass were measured under a 50 mW 915 nm LD pump excitation at 300–480 K. With the increase in temperature caused by the external heating, the peak emission intensity of the Yb3+ and Er3+ ions in the EYPS bulk glass decreases significantly. Figure 6d shows the normalized fluorescence spectra, and the inset in Figure 6d shows an enlarged view at ~1 μm. The emission peaks broaden as the temperature rises from 300 to 480 K [1,17,35]. The emission intensity ratio of I1018nm/I1534nm increases by 30.5%, and the result is confirmed by the increase in the A2 value in Table 2. The efficiency of the ET from the Yb3+ to the Er3+ ions significantly decreases as the temperature rises.
Considering the practical application of the EYDF, without changing the heat dissipation conditions, the temperature rise of the fiber core is due to the thermal effect associated with the increase in the LD pump power. The active quantum defect heating in the vacuum chamber, caused by increasing the 915 nm LD pump power, is shown in Figure 1b, with the temperature controller turning off. Similar to the above discussion, the emission intensity ratio of I1018nm/I1534nm was measured to evaluate the effect of temperature on efficiency of the ET from the Yb3+ to the Er3+ ions. Figure 7a shows the changes of the emission intensity (I1534nm and I1018nm) and the EYPS surface temperature with the increasing pump power. As the 915 nm LD pump power increases from 50 to 3000 mW, the surface temperature of the EYPS bulk glass increases from 305.9 to 481.1 K, caused by the significant thermal effects of LD pumping. The emission intensity I1018nm continuously increases with increased pump power. However, the emission intensity I1534nm reaches a maximum when the LD pump power increases to 2000 mW. The results indicate that the ET process from the Yb3+ to the Er3+ ions reaches a bottleneck. The increase in temperature and the population inversion of the Yb3+ and Er3+ ions [23,33] cooperatively limit the ET process. Figure 7b shows the normalized emission spectrum of the EYPS bulk glass under the excitation of the 915 nm LD pump. As the EYPS surface temperature rises from 305.9 to 481.1 K, the emission intensity ratio of I1018nm/I1534nm increases by 709.1% from 0.045 to 0.3641, and the efficiency of the ET from the Yb3+ to the Er3+ ions is reduced significantly. In summary, the emission intensity ratios of I1018nm/I1534nm in Figure 6d and Figure 7b increase by 30.6% and 709.1%, respectively. Compared with the external heating of the EYPS, the quantum defect heating not only causes a temperature increase, but results in a more significant reduction in the efficiency of the ET of the Yb3+ to the Er3+ ions. Under a high enough pump power, the ET process will be saturated. The increase in the population inversion of Yb3+ and Er3+ ions [23,33] results in a stronger ~1 μm intensity, adversely affecting the 1.5 μm laser output.

4. Conclusions

The Er3+/Yb3+ ion co-doped silica glasses with 15mol% P2O5 were prepared using the sol-gel combined high-temperature sintering method. The high efficiency of the ET from the Yb3+ to the Er3+ ions was guaranteed by high phosphorus doping. In this study, the temperature dependence (the range of 300 to 480 K) of the primary absorption (at 915, 940, 974, and 1018 nm) and the efficiency of the ET from the Yb3+ to the Er3+ ions in the EYPS fiber core glass were investigated. The results indicated that the absorption cross-section decreased by 27.2% (0.173 pm2→0.126 pm2) at 915 nm and 37.9% (0.943 pm2→0.586 pm2) at 974 nm, and increased by 106.9% (0.029 pm2→0.060 pm2) at 1018 nm. However, the pump absorption cross-section at ~940 nm was maintained at 0.173 pm2 from 300 to 480 K. The weak temperature dependence of the pump absorption at ~940 nm is conducive to the stable high-power output of EYDF amplifiers.
Additionally, the double-exponential fluorescence decay of the Yb3+ ions was observed, corresponding to the ET and SE2 processes. This is attributed to the existence of two types of Yb3+ ions in the EYPS glass: those coupled and those uncoupled to the Er3+ ions. As the temperature increased from 300 to 480 K because of the external heating, the population of isolated Yb3+ ions participating in the SE2 process increased, and the emission intensity ratio of I1018nm/I1534nm increased by 30.5%. The reduction in the efficiency of the ET from the Yb3+ to the Er3+ ions was confirmed as the temperature rose. Considering the practical application of the EYDF, the quantum defect heating scheme was adopted by increasing the LD pump power. The surface temperature of the EYPS bulk glass increased from ~306 to ~481 K. The emission intensity ratio of I1018nm/I1534nm increased by 709.1%, which is primarily attributed to the increase in the population inversion of the Yb3+ and Er3+ ions. The saturated ET process resulted in a strong ~1 μm emission intensity, and the efficiency of the ET from the Yb3+ to the Er3+ ions decreased significantly.

Author Contributions

Conceptualization, Y.C. and C.Y.; methodology, Y.C. and C.Y.; resources, H.D. and S.W.; validation, Y.C., H.D. and C.Y.; formal analysis, L.H., Q.Y., Y.J., S.W., C.S. and Y.D.; writing—review and editing, Y.C., Q.Y., Y.J., S.W., Y.D. and C.Y.; funding acquisition, C.Y., C.S. and L.H. All authors have read and agreed to the published version of the manuscript.

Funding

The National Natural Science Foundation of China (Grant Nos. 61775224, 61875216, 62005297).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Eichhorn, M.; Pollnau, M. Spectroscopic Foundations of Lasers: Spontaneous Emission Into a Resonator Mode. IEEE J. Sel. Top. Quantum Electron. 2015, 21, 486–501. [Google Scholar] [CrossRef]
  2. Shao, C.; Wang, F.; Jiao, Y.; Wang, S.; Wang, X.; Yu, C.; Hu, L. Relationship between glass structure and spectroscopic properties in Er3+/Yb3+/Al3+/P5+-doped silica glasses. Opt. Mater. Express 2020, 10, 1169–1181. [Google Scholar] [CrossRef]
  3. Sobon, G.; Sliwinska, D.; Abramski, K.M.; Kaczmarek, P. 10 W single-mode Er/Yb co-doped all-fiber amplifier with suppressed Yb-ASE. Laser Phys. Lett. 2014, 11, 025103. [Google Scholar] [CrossRef]
  4. Zhu, Z.W.; Liu, Y.; Luo, D.P.; Gu, C.L.; Zhou, L.; Xie, G.H.; Deng, Z.J.; Li, W.X. Tunable optical frequency comb from a compact and robust Er:fiber laser. High Power Laser Sci. Eng. 2020, 8, e17. [Google Scholar] [CrossRef]
  5. Zha, S.; Chen, Y.; Li, B.; Lin, Y.; Liao, W.; Zou, Y.; Huang, C.; Lin, Z.; Zhang, G. High-repetition-rate 1.5 µm passively Q-switched Er:Yb:YAl3(BO3)4 microchip laser. Chin. Opt. Lett. 2021, 19, 071402. [Google Scholar] [CrossRef]
  6. Yu, W.; Xiao, Q.; Wang, L.; Zhao, Y.; Qi, T.; Yan, P.; Gong, M. 219.6 W large-mode-area Er:Yb codoped fiber amplifier operating at 1600 nm pumped by 1018 nm fiber lasers. Opt. Lett. 2021, 46, 2192–2195. [Google Scholar] [CrossRef] [PubMed]
  7. Yang, Q.; Jiao, Y.; Yu, C.; Shao, C.; Lou, F.; Wang, S.; Zhang, L.; Yang, Q.; Hu, L. Gain and laser performance of heavily Er-doped silica fiber fabricated by MCVD combined with the sol-gel method. Chin. Opt. Lett. 2021, 19, 110603. [Google Scholar] [CrossRef]
  8. Michaud, L.C.; Veilleux, C.; Bilodeau, G.; Gilbert-Paquet, O.; Lebel-Cormier, M.A.; Lemieux-Tanguay, M.; Pelletier-Ouellet, S.; Paradis, P.; Bellec, M.; Gregoire, N.; et al. 100-W-level single-mode ytterbium-free erbium fiber laser. Opt. Lett. 2021, 46, 2553–2556. [Google Scholar] [CrossRef]
  9. Kiritchenko, N.V.; Kotov, L.V.; Melkumov, M.A.; Likhachev, M.E.; Bubnov, M.M.; Yashkov, M.V.; Laptev, A.Y.; Guryanov, A.N. Effect of ytterbium co-doping on erbium clustering in silica-doped glass. Laser Phys. 2015, 25, 025102. [Google Scholar] [CrossRef]
  10. Matniyaz, T.; Kong, F.; Kalichevsky-Dong, M.T.; Dong, L. 302 W single-mode power from an Er/Yb fiber MOPA. Opt. Lett. 2020, 45, 2910–2913. [Google Scholar] [CrossRef]
  11. Jeong, Y.; Harker, A.; Lovelady, M.; Piper, A.; Yoo, S.; Codemard, C.A.; Nilsson, J.; Sahu, J.K.; Payne, D.N.; Horley, R.; et al. Erbium:Ytterbium Codoped Large-Core Fiber Laser With 297-W Continuous-Wave Output Power. IEEE J. Sel. Top. Quantum Electron. 2007, 13, 573–579. [Google Scholar] [CrossRef] [Green Version]
  12. Kobayashi, Y.; Sekiya, E.H.; Banno, M.; Nishimura, R.; Okazaki, T.; Hashimoto, Y.; Araki, T.; Ichii, K.; Saito, K. Effect of P-to-Rare Earth Atomic Ratio on Energy Transfer in Er-Yb-Doped Optical Fiber. J. Lightwave Technol. 2020, 38, 4504–4512. [Google Scholar] [CrossRef]
  13. Wang, S.; Lou, F.; Yu, C.; Zhou, Q.; Wang, M.; Feng, S.; Chen, D.; Hu, L.; Chen, W.; Guzik, M.; et al. Influence of Al3+ and P5+ ion contents on the valence state of Yb3+ ions and the dispersion effect of Al3+ and P5+ ions on Yb3+ ions in silica glass. J. Mater. Chem. C 2014, 2, 4406–4414. [Google Scholar] [CrossRef]
  14. Townsend, J.E.; Barnes, W.L.; Crubb, S.G. Yb3+ Sensitised Er3+ Doped Silica Optical Fibre with Ultra High Transfer Efficiency and Gain. MRS Online Proc. Libr. 1991, 244, 143. [Google Scholar] [CrossRef] [Green Version]
  15. Canat, G.; Mollier, J.-C.; Jaouën, Y.; Dussardier, B. Evidence of thermal effects in a high-power Er3+–Yb3+ fiber laser. Opt. Lett. 2005, 30, 3030–3032. [Google Scholar] [CrossRef]
  16. Hendow, S.T.; Chen, W.; Sha, J.; Wang, Y.; Shen, D. Investigation of temperature influence on output performances of high-power cladding-pumped Er,Yb co-doped fiber laser. Int. Soc. Opt. Photonics 2013, 8601, 86012A. [Google Scholar] [CrossRef]
  17. Cheng, Y.; Yang, Q.; Yu, C.; Guo, M.; Jiao, Y.; Dai, Y.; Wang, S.; Hu, L. Temperature dependence of the spectral properties of Yb3+/P5+/Al3+ co-doped silica fiber core glasses. Opt. Mater. Express 2021, 11, 2459–2467. [Google Scholar] [CrossRef]
  18. He, Q.; Wang, F.; Lin, Z.; Shao, C.; Wang, M.; Wang, S.; Yu, C.; Hu, L. Temperature dependence of spectral and laser properties of Er3+/Al3+ co-doped aluminosilicate fiber. Chin. Opt. Lett. 2019, 17, 101401. [Google Scholar] [CrossRef]
  19. Mescia, L.; Bia, P.; Girard, S.; Ladaci, A.; Chiapperino, M.A.; Robin, T.; Laurent, A.; Cadier, B.; Boutillier, M.; Ouerdane, Y.; et al. Temperature-Dependent Modeling of Cladding-Pumped Er3+/Yb3+-Codoped Fiber Amplifiers for Space Applications. J. Lightwave Technol. 2018, 36, 3594–3602. [Google Scholar] [CrossRef]
  20. Yu, W.; Yan, P.; Xiao, Q.; Qi, T.; Li, D.; Gong, M. Power scalability of a continuous-wave high-power Er-Yb co-doped fiber amplifier pumped by Yb-doped fiber lasers. Appl. Opt. 2021, 60, 2046–2055. [Google Scholar] [CrossRef]
  21. Guan, X.; Zhao, Q.; Lin, W.; Tan, T.; Yang, C.; Ma, P.; Yang, Z.; Xu, S. High-efficiency and high-power single-frequency fiber laser at 1.6  μm based on cascaded energy-transfer pumping. Photonics Res. 2020, 8, 2327–9125. [Google Scholar] [CrossRef]
  22. Ballato, J.; Creeden, D.; Pretorius, H.; Limongelli, J.; Setzler, S.D. Single frequency 1560nm Er:Yb fiber amplifier with 207W output power and 50.5% slope efficiency. Int. Soc. Opt. Photonics 2016, 9728, 97282L. [Google Scholar] [CrossRef]
  23. De Varona, O.; Fittkau, W.; Booker, P.; Theeg, T.; Steinke, M.; Kracht, D.; Neumann, J.; Wessels, P. Single-frequency fiber amplifier at 1.5 μm with 100 W in the linearly-polarized TEM00 mode for next-generation gravitational wave detectors. Opt. Express 2017, 25, 24880–24892. [Google Scholar] [CrossRef] [PubMed]
  24. Melkumov, M.A.; Laptev, A.Y.; Yashkov, M.V.; Vechkanov, N.N.; Guryanov, A.N.; Bufetov, I.A. Effects of Yb3+ and Er3+ concentrations and doping procedure on excitation transfer efficiency in Er-Yb doped phosphosilicate fibers. Inorg. Mater. 2010, 46, 299–303. [Google Scholar] [CrossRef]
  25. Rybaltovsky, A.A.; Lipatov, D.S.; Lobanov, A.S.; Abramov, A.N.; Umnikov, A.A.; Bazakutsa, A.P.; Bobkov, K.K.; Butov, O.V.; Gur’yanov, A.N. Photosensitive highly Er/Yb co-doped phosphosilicate optical fibers for continuous-wave single-frequency fiber laser applications. J. Opt. Soc. Am. B 2020, 37, 3077–3083. [Google Scholar] [CrossRef]
  26. Khudyakov, M.M.; Lobanov, A.S.; Lipatov, D.S.; Abramov, A.N.; Vechkanov, N.N.; Guryanov, A.N.; Melkumov, M.A.; Bobkov, K.K.; Aleshkina, S.S.; Kochergina, T.A.; et al. Single-mode large-mode-area Er–Yb fibers with core based on phosphorosilicate glass highly doped with fluorine. Laser Phys. Lett. 2019, 16, 025105. [Google Scholar] [CrossRef]
  27. Moore, S.W.; Barnett, T.; Reichardt, T.A.; Farrow, R.L. Optical properties of Yb3+-doped fibers and fiber lasers at high temperature. Opt. Commun. 2011, 284, 5774–5780. [Google Scholar] [CrossRef]
  28. Aubry, M.; Morana, A.; Mescia, L.; Mekki, J.; Ranger, C.; Laurent, A.; Robin, T.; Marin, E.; Ouerdane, Y.; Boukenter, A.; et al. Combined Temperature and Radiation Effects on the Gain of Er- and Er–Yb-Doped Fiber Amplifiers. IEEE Trans. Nucl. Sci. 2021, 68, 793–800. [Google Scholar] [CrossRef]
  29. Wang, S.; Xu, W.; Wang, F.; Lou, F.; Zhang, L.; Zhou, Q.; Chen, D.; Feng, S.; Wang, M.; Yu, C.; et al. Yb3+-doped silica glass rod with high optical quality and low optical attenuation prepared by modified sol-gel technology for large mode area fiber. Opt. Mater. Express 2017, 7, 2012–2022. [Google Scholar] [CrossRef]
  30. Guo, M.; Shao, C.; Shi, F.; Wang, F.; Wang, S.; Yu, C.; Zhang, L.; Ren, J.; Hu, L. Effect of thermal annealing on structures and properties of Yb3+/Al3+/P5+ co-doped silica glasses. J. Non-Cryst. Solids 2019, 522, 119563. [Google Scholar] [CrossRef]
  31. Dong, H.; Wang, Z.; Shao, C.; Wang, S.; Lou, F.; Yu, C.; Chen, Y.; Hu, L. Effect of co-dopants on the spectral property of Yb3+ doped silica glasses at 1018 nm. Opt. Mater. 2021, 122, 111716. [Google Scholar] [CrossRef]
  32. Laroche, M.; Girard, S.; Sahu, J.K.; Clarkson, W.A.; Nilsson, J. Accurate efficiency evaluation of energy-transfer processes in phosphosilicate Er3+-Yb3+-codoped fibers. J. Opt. Soc. Am. B 2006, 23, 195–202. [Google Scholar] [CrossRef]
  33. Dong, L.; Matniyaz, T.; Kalichevsky-Dong, M.T.; Nilsson, J.; Jeong, Y. Modeling Er/Yb fiber lasers at high powers. Opt. Express 2020, 28, 16244–16255. [Google Scholar] [CrossRef] [PubMed]
  34. Jeong, Y.; Baek, S.; Dupriez, P.; Maran, J.N.; Sahu, J.K.; Nilsson, J.; Lee, B. Thermal characteristics of an end-pumped high-power ytterbium-sensitized erbium-doped fiber laser under natural convection. Opt. Express 2008, 16, 19865–19871. [Google Scholar] [CrossRef]
  35. Yong, Y.S.; Aravazhi, S.; Vazquez-Cordova, S.A.; Carjaval, J.J.; Diaz, F.; Herek, J.L.; Garcia-Blanco, S.M.; Pollnau, M. Temperature-dependent absorption and emission of potassium double tungstates with high ytterbium content. Opt. Express 2016, 24, 26825–26837. [Google Scholar] [CrossRef] [Green Version]
  36. Dexter, D.L. A Theory of Sensitized Luminescence in Solids. J. Chem. Phys. 1953, 21, 836–850. [Google Scholar] [CrossRef]
  37. Federighi, M.; Pasquale, F.D.J.I.P.T.L. The effect of pair-induced energy transfer on the performance of silica waveguide amplifiers with high Er3+/Yb3+ concentrations. IEEE Photonics Technol. Lett. 1995, 7, 303–305. [Google Scholar] [CrossRef]
  38. Deschamps, T.; Vezin, H.; Gonnet, C.; Ollier, N. Evidence of AlOHC responsible for the radiation-induced darkening in Yb doped fiber. Opt. Express 2013, 21, 8382–8392. [Google Scholar] [CrossRef]
Figure 1. Front view of (a) the temperature-controllable absorption test device; front view of (b) the fluorescence test device, using two methods for temperature control: ① external heating of the EYPS bulk glass using the temperature controller; ② active quantum defect heating of the EYPS bulk glass by increasing the 915 nm LD pump power, with the temperature controller turning off.
Figure 1. Front view of (a) the temperature-controllable absorption test device; front view of (b) the fluorescence test device, using two methods for temperature control: ① external heating of the EYPS bulk glass using the temperature controller; ② active quantum defect heating of the EYPS bulk glass by increasing the 915 nm LD pump power, with the temperature controller turning off.
Materials 15 00996 g001
Figure 2. (a) Fluorescence spectrum of the EYPS at ~300 K; (b) fluorescence decay curves of Yb3+ and Er3+ ions in the EYPS at ~300 K.
Figure 2. (a) Fluorescence spectrum of the EYPS at ~300 K; (b) fluorescence decay curves of Yb3+ and Er3+ ions in the EYPS at ~300 K.
Materials 15 00996 g002
Figure 3. (a) Absorption cross-section spectra of Yb3+ ion in EYPS bulk glass in the temperature range from 300 to 480 K; (b) ratios of the absorption cross-section at different temperatures to that at 300 K for the primary pump wavelength of the EYPS bulk glass (the curves are obtained by fitting).
Figure 3. (a) Absorption cross-section spectra of Yb3+ ion in EYPS bulk glass in the temperature range from 300 to 480 K; (b) ratios of the absorption cross-section at different temperatures to that at 300 K for the primary pump wavelength of the EYPS bulk glass (the curves are obtained by fitting).
Materials 15 00996 g003
Figure 4. Absorption curves of Yb3+ ions in EYPS bulk glass are processed by Lorentz fitting at (a) 300 K and (b) 480 K. The dashed curves show the multiple decomposition peaks corresponding to different Stark transitions, and the inset is a simplified Yb3+ ion energy level.
Figure 4. Absorption curves of Yb3+ ions in EYPS bulk glass are processed by Lorentz fitting at (a) 300 K and (b) 480 K. The dashed curves show the multiple decomposition peaks corresponding to different Stark transitions, and the inset is a simplified Yb3+ ion energy level.
Materials 15 00996 g004
Figure 5. (a) Time−resolved emission spectra of the EYPS bulk glass under 980 nm pulse excitation at 300 K, and the inset shows the duration of the 980 nm pulse; (b) simplified modeling of the ET and SE processes of coupled and isolated Yb3+ ions in the EYPS bulk glass.
Figure 5. (a) Time−resolved emission spectra of the EYPS bulk glass under 980 nm pulse excitation at 300 K, and the inset shows the duration of the 980 nm pulse; (b) simplified modeling of the ET and SE processes of coupled and isolated Yb3+ ions in the EYPS bulk glass.
Materials 15 00996 g005
Figure 6. (a) Double−exponential fluorescence decay curves of Yb3+ ions and (b) fluorescence decay curves of Er3+ ions in the EYPS in 300−480 K; (c) fluorescence spectra of the EYPS and (d) its normalized spectra in 300−480 K. The inset is an enlarged view at ~1 μm, and the external heating caused the temperature rise.
Figure 6. (a) Double−exponential fluorescence decay curves of Yb3+ ions and (b) fluorescence decay curves of Er3+ ions in the EYPS in 300−480 K; (c) fluorescence spectra of the EYPS and (d) its normalized spectra in 300−480 K. The inset is an enlarged view at ~1 μm, and the external heating caused the temperature rise.
Materials 15 00996 g006
Figure 7. Under the 915 nm LD pump excitation in the power range of 50–3000 mW, (a) the change in the emission intensity (I1534nm and I1018nm) and surface temperature of the EYPS with increasing pump power. The inset shows the surface temperature of the EYPS in real time when the pump power is increased to 2000 mw; (b) the normalized emission spectra of the EYPS bulk glass. The active quantum defect heating caused the temperature rise.
Figure 7. Under the 915 nm LD pump excitation in the power range of 50–3000 mW, (a) the change in the emission intensity (I1534nm and I1018nm) and surface temperature of the EYPS with increasing pump power. The inset shows the surface temperature of the EYPS in real time when the pump power is increased to 2000 mw; (b) the normalized emission spectra of the EYPS bulk glass. The active quantum defect heating caused the temperature rise.
Materials 15 00996 g007
Table 1. Mean compositions of the EYPS and YPS bulk glasses (mol%).
Table 1. Mean compositions of the EYPS and YPS bulk glasses (mol%).
SamplesTheoretical CompositionActual Composition
Er2O3Yb2O3P2O5SiO2Er2O3Yb2O3P2O5SiO2
EYPS0.11.01583.90.090.8612.4586.60
YPS-1.01584.0-0.8311.0888.09
Table 2. Fitted results of fluorescence decay curves of Yb3+ and Er3+ ions in the EYPS and YPS bulk glasses from 300 to 480 K.
Table 2. Fitted results of fluorescence decay curves of Yb3+ and Er3+ ions in the EYPS and YPS bulk glasses from 300 to 480 K.
TemperatureEYPSYPS
A1τ1 (μs)A2τ2 (μs)τEr (ms)τYb (μs)
300 K0.8648.200.1361333.769.121793.46
340 K0.8628.140.1381312.628.921754.72
380 K0.8597.910.1411306.878.761736.54
420 K0.8597.900.1411299.428.761719.04
480 K0.8547.560.1461279.258.141686.01
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cheng, Y.; Dong, H.; Yu, C.; Yang, Q.; Jiao, Y.; Wang, S.; Shao, C.; Hu, L.; Dai, Y. Temperature Dependence of Absorption and Energy Transfer Efficiency of Er3+/Yb3+/P5+ Co-Doped Silica Fiber Core Glasses. Materials 2022, 15, 996. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15030996

AMA Style

Cheng Y, Dong H, Yu C, Yang Q, Jiao Y, Wang S, Shao C, Hu L, Dai Y. Temperature Dependence of Absorption and Energy Transfer Efficiency of Er3+/Yb3+/P5+ Co-Doped Silica Fiber Core Glasses. Materials. 2022; 15(3):996. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15030996

Chicago/Turabian Style

Cheng, Yue, Hehe Dong, Chunlei Yu, Qiubai Yang, Yan Jiao, Shikai Wang, Chongyun Shao, Lili Hu, and Ye Dai. 2022. "Temperature Dependence of Absorption and Energy Transfer Efficiency of Er3+/Yb3+/P5+ Co-Doped Silica Fiber Core Glasses" Materials 15, no. 3: 996. https://0-doi-org.brum.beds.ac.uk/10.3390/ma15030996

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop