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

Evidence of AlOHC responsible for the radiation-induced darkening in Yb doped fiber

Open Access Open Access

Abstract

Using a combination of experimental techniques such as optical absorption, Raman scattering, continuous wave and pulse Electron Spin Resonance (ESR), we characterize a set of γ-irradiated Yb3+ doped silica glass preforms with different contents of phosphorous and aluminum. We demonstrate that when P is introduced in excess compared to Al, nearly no radiodarkening is induced by γ-rays. On the other hand, when Al>P, a large absorption band is induced by radiation. Thermal annealing experiments reveal the correlation between the decrease of the optical absorption band and the decrease of the Al-Oxygen Hole Center (AlOHC) ESR signal, demonstrating the main role of AlOHC defects in the fiber darkening. HYSCORE (HYperfine Sublevel CORElation) pulse-ESR experiments show a high Al-P nuclear spin coupling when P>Al and no coupling when Al>P. This result suggests that both AlOHC and POHC creation is inhibited by Al-O-P linkages. Confronting our data with previous works, we show that the well-known photodarkening process, meaning losses induced by the IR pump, can also be explained in this framework.

©2013 Optical Society of America

1. Introduction

Ytterbium-doped high power laser fibers, operating near 1 micron, are currently used for industrial material processing and biomedical devices. Such active fibers are also promising candidates to achieve the inertial confinement fusion [1,2] or to operate in space for inter-satellites optical communication [3,4]. However, radiative environments imply the formation of absorbing defects, and the resulting absorption features located in the UV and visible range have long absorption tails that overlap with the pump and emission wavelengths of Yb3+. As a consequence, the radiation-induced darkening can lead to a drastic decrease of the laser performance. Several techniques can be used to improve the radiation hardening like Cerium codoping, O2 or H2-treatment [48].

Nevertheless, the role of the codopants (such as Al or P which are incorporated to reduce rare-earths clustering [9,10]) in the radiation response of the fibers remains unclear. Previous works suggest that these dopants are responsible for a strong part of the output losses [1114], but the absorbing defects, namely in the visible-near IR region, have not been yet clearly identified. Even so, this step remains crucial to understand the mechanisms of radiation-induced absorption and to improve efficiently the fibers performance in harsh environment.

In this paper, using optical absorption and electronic spin resonance (ESR) spectroscopy, it is experimentally found that the Aluminium-Oxygen Hole Center (AlOHC) paramagnetic defect is mainly responsible of the radiodarkening (gamma-rays induced absorption) in Yb-doped silica-based fiber preforms. It is also demonstrated that phosphorous codoping allows to prevent the formation of such absorbing defects. A structural study of the pristine samples, combining Raman and pulse ESR spectroscopy, allows a better understanding of the role of phosphorous and aluminium codopants in the vicinity of ytterbium. Confronting this set of experimental data with previous works, we show that the well-known attenuation induced by high-power IR laser diode pumping, called photodarkening [1518], can also be explained in this framework.

2. Experimental

2.1 Samples synthesis and composition

Ytterbium doped aluminophosphosilicate fiber preforms were fabricated by mean of the modified chemical vapour deposition (MCVD) technique. The set of samples has been designed to investigate the interaction between ytterbium and Al/P dopants, and the respective role of Al and P in the darkening of such fibers under irradiation. Thereby, the different cores present roughly the same content of Yb, Si and Ge, and a variable amount of Al and P. The average composition of the cores has been determined by quantitative chemical analyses using Castaing’s microprobe spectroscopy. Each composition is displayed in the Table 1.

Tables Icon

Table 1. Chemical core composition of the fiber preforms determined using Castaing’s microprobe. The Al/P ratio is also provided. Concentrations are in at%.

2.2 Gamma irradiation experiment

All samples have been exposed to 60Co γ-radiation at room temperature using an armored cell at CEA Saclay (DSM, SIS2M). The irradiation was performed at a dose-rate of 6.0 Gy.min−1 during 5 hours, implying a cumulated dose of 1800 Gy, which corresponds to the typical accumulated dose for a 10-20 years mission in a space environment [3,4]. The pictures of some of the samples before and after irradiation are presented Fig. 1. It is shown that the core samples exhibit a strong darkening after exposure to γ-rays when Al>P, and almost no coloration appears when P>Al.

 figure: Fig. 1

Fig. 1 Samples A, B, G, and F before (left) and after (right) gamma irradiation. The samples F and G, which contains an excess of aluminum compared to phosphorous, are darkened by gamma rays.

Download Full Size | PDF

3. Defects creation and thermal relaxation in post-irradiated samples

3.1 Optical absorption and continuous-wave ESR

After γ-irradiation, each preform was cut into different adapted geometry. The samples used for absorption spectroscopy were double-face polished (up to an optical grade) thin disks of preform of roughly 2.5 mm diameter and 1 mm thickness. Absorption spectra, displayed Fig. 2(a), were recorded in the 280-1340 nm wavelengths range. Only the core of the preforms was probed, using a hole of the dimension of the core, i.e 1 mm diameter. Figure 2(a) clearly reveals the different behaviors of the fibers under gamma-irradiation and the importance of Al/P ratio. The two post-irradiated preforms A and B, which contain an excess of phosphorous compared to aluminum, present no absorption band in the visible and near IR range (the sharp band around 980 nm corresponds to 2F7/2-2F5/2 absorption band of Yb3+). The other samples display a large absorption feature in the visible region with a tail up to the near IR, and the intensity of this absorption increases with the excess of Al compared to P. In order to identify the colored centers generated by γ-irradiation and responsible of the induced absorption, continuous-wave ESR measurements were conducted at room using an X-band EPR-Brüker spectrometer operating at 9.868 GHz, in the 340-360 mT magnetic field range, with a power of 2mW. The post-irradiated samples used for this experiment were cylinders of 2.5 mm diameter and 3.8 mm height.

 figure: Fig. 2

Fig. 2 (a) Optical absorption spectra of the seven irradiated preforms. Samples A and B (P>Al) exhibit low absorption band in the visible range. The sharp peak around 980 nm is the absorption band of Yb3+. The spectra have been cut between 850 nm and 930 nm due to discontinuities induced by the change of the lamp during the experiment; (b) cw-ESR spectra of the seven irradiated samples. The inset displays an enlargement of the weak signal of AlOHC defects around 348 mT. All the spectra have been recorded at ambient temperature.

Download Full Size | PDF

The cw-ESR spectra Fig. 2(b) present a composite and intense feature between 352 and 356 mT, characteristic of the presence of Ge-related paramagnetic point defects [19,20]. E’-Ge, Ge(1) and Ge(2) defects are induced by γ-irradiation, but could not be responsible of the darkening, due to their absorption located respectively at 6.3, 4.4 and 5.8 eV, i.e in the UV range [19,20]. The inset of Fig. 2(b) focuses on the weak signal at 348 mT (g = 2.025) which is the signature of AlOHC point defect [21,22]. To better visualize this AlOHC signal, ESR spectroscopy has been performed on the sample G at different powers, in the range 0.1mW-50mW (Fig. 3). Figure 3(a) diplays the evolution of the ESR signal and Fig. 3(b) the intensity of the AlOHC signal in function of the power. Because saturation is observed above P = 2mW, this power value has been choosing as the working power.

 figure: Fig. 3

Fig. 3 (a) cw-ESR spectra of the preform G at different power between 0.1mW and 50 mW around the AlOHC defect resonance; (b) Intensity of the AlOHC signal in function of the power square root. A linear behaviour is observed up to P = 2mW, the working power.

Download Full Size | PDF

The AlOHC ESR band is nearly missing in the A and B samples, whereas it appears for the others and grows with the excess of Al compared to P. Comparing optical absorption and cw-ESR measurements, a correlation seems to exist between the generation of AlOHC point defects and the large absorption band in the visible-near IR region in Al-rich preforms (we note however a poor correlation concerning the sample C which displays no AlOHC ESR signal and visible absorption, probably due to experimental errors concerning absorption measurements) In order to confirm this assumption, thermal annealing measurements follow by absorption and cw-ESR characterization have been performed.

3.2 Thermal annealing experiment

The sample G, which contains the higher Al excess, has been chosen to investigate the effect of thermal annealing on the AlOHC point defects using ESR and optical absorption. The irradiated G preform was annealed during 10 min at temperatures ranging from 20°C to 650 °C. The Fig. 2 displays the optical absorption recorded after heating at each temperature. A gradual decrease of the large absorption band from 300 nm up to the IR region is observed when the annealing temperature increases.

Figure 4 shows the intensity of the absorption at 550 nm and the intensity of the AlOHC ESR signal in function of the annealing temperature. The visible absorption remains stable from Tamb to 150-200 °C. Beyond this temperature, the absorption band decreases rapidly up and almost entirely vanishes above 450 °C. Taking into account the experimental errors, the evolution of the AlOHC defect probed by ESR follows the evolution of the optical absorption, putting in evidence the correlation between the visible-near IR range absorption and the AlOHC defect (Fig. 5).

 figure: Fig. 4

Fig. 4 Optical absorption spectra of the preform G recorded during thermal annealing experiment up to 650 °C. A gradual decrease of the visible absorption induced by irradiation is observed.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 (a) Evolution of the intensity of the visible absorption band; (b) the intensity of AlOHC ESR signal during the thermal annealing experiment. Both characterizations display a stable behaviour up to 200 °C, and a decrease of the signal between 200 °C and 400 °C. Above 450 °C, the visible absorption band and the AlOHC ESR signal have completely disappeared.

Download Full Size | PDF

This result unambiguously reveals that the AlOHC point defect is the main colored center responsible of the darkening in post-irradiated fibers.

4. Glass structure and Yb3+ environment in pre-irradiated samples

In order to better understand the formation of AlOHC colored centers induced by γ-irradiation, a structural study of the pristine samples has been performed using Raman scattering and pulse-ESR spectroscopy. Raman spectroscopy allows to probe the global glass structure, namely the insertion of Al and P in the silica network, and the pulse-ESR spectroscopy gives direct information concerning the local environment of the Yb3+ paramagnetic ions.

4.1 Raman spectroscopy: Al and P codopants insertion

Raman analyses were performed using a Renishaw spectrometer in backscattering configuration, with the 532 nm excitation line provided from a Nd3+-YAG laser. Figure 6 shows the Raman spectra acquired on the A, C, D and F preforms as well as the pure a-SiO2 Raman spectrum. The Raman spectra display a strong broad band around 440 cm−1 and weaker features at 490, 600 and 800 cm−1. All these vibrational modes are presents in pure SiO2 glass Raman spectrum and so provides from SiO2 network vibrations. The 440 cm−1 main band, assigned to Si-O-Si bending mode, has been used to normalize the spectra. The two sharp bands around 500 and 600 cm−1 come from the breathing modes of 4 and 3-member rings respectively, and the weak high-frequency bands at 1060 and 1190 cm−1 arise from TO and LO antisymmetric stretching vibrations of Si-O-Si linkages respectively [23]. This high-frequency range is largely modified by Al and/or P addition.

 figure: Fig. 6

Fig. 6 Normalized Raman spectra of preform fibers A, C, D, F (straight lines) and of pure silica glass (dotted line). Above 1000 cm−1, Al-O-P linkages are revealed in the doped samples. The P = O stretching band at 1330 cm−1 well appears when P>Al (sample A).

Download Full Size | PDF

Raman spectra of samples C, D and F present similar features at high frequency: a broad band centred at 1135 cm−1 and a shoulder at 1240 cm−1. A sharp band at 1330 cm−1, well attributed to P = O stretching vibration in SiO2-P2O5 glasses [24,25], is also observed in C sample. On the other hand, significant differences are observed in the Raman spectrum of F preform. First, the 1330 cm−1 vibration is much more intense. Secondly, the shape of the broad high frequency band is rather different, and its maximum is located at 1145 cm−1, revealing P-O-P linkages [25]. These linkages are confirmed by the presence of the 730 cm−1 P-O-P stretching mode [25].

A precise quantification of the vibrations is rather complicated because many vibrational bands due to different types of linkages occur in the 1000-1300 cm−1 wave numbers region. Moreover, the Raman bands frequency and width depend of bond angle distribution, and the Raman cross-section, linked to the bonds polarizability, depend also of the linkages type. All these reasons rule out the Raman spectra deconvolution. However, some characteristics can be extracted from high frequency Raman spectra. For Al/P>1, the shape of the 1000-1300 cm−1 broad band looks like AlPO4 molecular sieves spectra, indicating the presence of AlPO4-like units [26]. For Al/P<1, AlPO4 units vibrations contribute to the 1000-1300 cm−1 broad band, as well as P-O-P vibrations. Concerning the 1330 cm−1 band, a transition can be observed at Al/P = 1, which clearly indicate the presence of P = O linkages for P≥Al glasses.

No detail about ytterbium connectivity can be extracted from Raman spectroscopy. The next section, which displays pulse-ESR measurements, will bring such information.

4.2 Pulse-ESR spectroscopy: Yb3+ local environment

In order to compare the local environment of Yb3+ as a function of Al and P contents, two-dimensional hyperfine sublevel correlation (2D-HYSCORE) ESR spectra have been performed recorded at 4K, using the pulse sequence π/2-τ- π/2-t1- π-t2- π/2-τ-echo with a time delay τ = 136 ns and at a static magnetic field of 350 mT where all the samples present significant spin echo absorption (the two-pulsed echo filed swept spectra of these samples can be found in [10]). This type of experiment, based on the coupling between electronic spin and nuclear spin (hyperfine interaction) allows to probe the local environment of the Yb3+ paramagnetic centers via the spin echo modulation at Larmor frequency of the neighbour nuclei. Oxygen atoms that fully occupy the first coordination shell of Yb3+ are not observed (I = 0). The different 2D spectra are displayed Fig. 7.

 figure: Fig. 7

Fig. 7 2D-HYSCORE pulse ESR spectra of the pristine samples A, B, D, E, F, and G recorded at liquid He temperature and at a magnetic field of 350 mT. A and B samples reveal off-diagonal peaks showing the proximity of Al and P nuclei around Yb3+ when phosphorous is introduced in excess compared to aluminium.

Download Full Size | PDF

The three diagonal peaks at 3.0 MHz, 3.9 MHz and 6.0 MHz correspond to the Larmor frequencies of 29Si, 27Al and 31P nucleus respectively. We note that when P>Al, the intense pattern at 6.0 MHz indicates that phosphorous formed a cage (or solvation shell) surrounding Yb3+ [10], also demonstrated in the case of Er3+ [27,28] that Additional off-diagonal cross-peaks at (3.9-6.0) MHz and (6.0-3.9) MHz are observed for samples A and B, i.e when P>Al. These cross-peaks revealed the Al-P coupling, demonstrating the proximity of Al and P nuclei in the vicinity of the Yb3+ paramagnetic centers when phosphorous is introduced in excess compared to aluminium. On the other hand, no Al-P correlation is observed for samples D, E, F and G where Al>P.

5. Discussion

5.1 AlOHC defect and glass structure

From the results of thermal bleaching combining optical absorption and cw-ESR, it was demonstrated that AlOHC point defects are mainly responsible of the darkening in post-irradiated preforms (see section 3). However, it is known that P-related point defects, namely Phosphorous-Oxygen Hole Center (POHC), can be induced by γ irradiation [12,29]. The P = O link, observed by Raman spectroscopy in samples where Al/P<1 (Fig. 6), is yet a precursor of the POHC paramagnetic defect [29]. Nevertheless, no POHC species have been detected by cw-ESR in the whole set of samples, in particular in the preform A where P is largely introduced in excess compared to Al. The absence (or the weak concentration) of POHC is also confirmed by the lack of absorption band in the visible range of this post-irradiated preform (Fig. 2).

A surprising result is that the level of AlOHC induced by radiation does not depend linearly on the absolute concentration of Al incorporated in the glass matrix, but rather on the Al/P ratio. Indeed, the post-irradiated samples A and B which present non negligible Al concentration, exhibit rather no darkening phenomenon and no AlOHC formation. This means that AlOHC are created only when phosphorous is not introduced in excess compared to Al. Raman spectroscopy measurements reveal that Al and P formed preferentially AlPO4 units. When P>Al, the excess of phosphorous formed P-O-P and P = O linkages. In this case, the absence of AlOHC after irradiation shows that AlPO4 structures inhibit such defects creation. On the other hand, when Al>P, the Al excess, which is not located into AlPO4 linkages, is ionized by γ-rays.

Among the different trapped-electron centers which could be created with AlOHC during the pair generation process (for a complete review, see [30]), one can namely mention Si-E’, Ge-E’, Al-E’ or P-related electron-trapped paramagnetic defects. Unfortunately, the high magnitude of Ge-related defects in the post-irradiated ESR spectra (Fig. 2) does not allow to correlate the AlOHC creation with a particular electron-trapped defect. Another potential candidate is the Yb2+ specie, which can be formed by Yb3+ reduction mechanism. This possibility seems highly probable considering the proximity between Al and Yb3+, revealed by several recent theoretical and experimental studies [10,3133], and confirmed by our pulse-ESR measurements (Fig. 7). Indeed, the HYSCORE spectra of D, E, F and G samples show important Al pattern without Al-P coupling. This reveals that the Al atoms in the vicinity of Yb3+ are not bounded to P atoms and do not form AlPO4 units, allowing AlOHC and Yb2+ pair generation. We note that ytterbium doped aluminosilicate have a charge transfer band at 230 nm (at the origin of Yb2+ and colored centers formation under high energy irradiation), which could be at the origin of the photodarkening process [17,34].

5.2 AlOHC and photodarkening

The complex mechanism of photodarkening process, meaning the IR pump induced losses in Yb3+ doped fibers, is a subject of a long and controversy debate [1518,3445]. In our present study, the purpose is not to bring new insights concerning the energy transfers and cooperative effects leading to UV emission at the origin of colored centers formation. However, a comparison of previous photodarkening studies with the present results seems to indicate that AlOHC defects are mainly responsible of the coloration in both cases: different recent studies show the dependence of the output losses in function of the Al and P codopants content [35,42], and reveal that the addition of phosphorous in aluminosilicate matrix reduced the photodarkening effects [35]. This result is coherent with our measurements showing the formation of AlPO4 units and the decrease of AlOHC formation with P adding. Moreover, results of thermal bleaching in a photodarkened Al-rich glass with a composition close to the preform G [41] display also similar behaviour than the thermal treatment of post γ-irradiated G sample (Fig. 4). This reinforced the hypothesis of AlOHC defects creation/annihilation during the photodarkening/thermal bleaching process.

However, one can note that in Al-rich preforms, meaning when Al>P, gamma irradiation induced more darkening as Al excess increases (see Fig. 2) while for similar preforms, the photodarkening is reduced as Al excess increases [35,46]. This apparent contradiction probably finds its origin in the different radiation ways (between radiodarkening and photodarkening) leading to the defects creation. In the case of radiodarkening experiments, all the samples have received the same gamma energy in the whole material. More complex is the photodarkening process, for which the probability of the emission of a high energy photon (able to create defects) highly depends on the matrix composition and its atomic structure.

6. Conclusion

A set of gamma-irradiated Yb3+ doped aluminophosphosilicate glass preforms have been extensively investigated using several kinds of spectroscopic measurements. The results show a high dependence of the Al/P ratio in the darkening process. For Al/P<1, a very slight coloration is observed whereas Al/P>1 samples revealed an important darkening after irradiation. The combination of optical absorption and cw-ESR measurements at different annealing temperatures demonstrates that the creation of AlOHC colored centers is the main process at the origin of the darkening in Al-rich glasses, while these defects are not favored in P-rich glass in which Al atoms are in AlPO4 units as it has been shown by Raman spectroscopy. Pulse-ESR spectra reveal that Al-O-P linkages are in the vicinity of Yb3+ ions when P>Al, but not when P<Al.

Moreover, several similarities between radiodarkening and photodarkening have been highlighted, suggesting that AlOHC defects are at the origin of the darkening in both gamma and high-power IR irradiations. However, the rate of AlOHC created after irradiation highly depends on the kind of irradiation sources (γ-rays or IR light).

As perspectives, several measurements are indispensable to clarify the different mechanisms involved in situ during the irradiation process, namely the dynamics of transient defects [4, 47], or even annihilation via thermalbleaching [41] or photobleaching [37,48]. Additionally, the way to reach, considering photodarkening, high energy photons that create the defects will also be a key point. Other points remain obscured: does Yb2+ is the electron-trapped specie conjugated to AlOHC, or other species are involved? The contribution of spectroscopy to probe the atomic organization to correlate with optical properties is, in our opinion, a key issue to better understand radiation induced darkening in optical waveguides.

Acknowledgments

This project was funded by the CEA transversal programm “matériaux avancés”. The authors gratefully acknowledge O. Cavani (Prysmian group) for fiber preforms synthesis and S. Esnouf (CEA-Saclay) for gamma irradiations.

References and links

1. E. Moses, “Multi-megajoule NIF: ushering in a new era in high density science,” Proc. SPIE 7005, 70050F, 70050F-11 (2008). [CrossRef]  

2. S. Girard, J. Baggio, J. L. Leray, J. P. Meunier, A. Boukenter, and Y. Ouerdane, “Laser megajoule CEA vulnerability analysis of optical fibers for laser megajoule facility: preliminary studies,” IEEE Trans. Nucl. Sci. 52(5), 1497–1503 (2005). [CrossRef]  

3. J. Ma, M. Li, L. Tan, Y. Zhou, S. Yu, and Q. Ran, “Experimental investigation of radiation effect on erbium-ytterbium co-doped fiber amplifier for space optical communication in low-dose radiation environment,” Opt. Express 17(18), 15571–15577 (2009). [CrossRef]   [PubMed]  

4. Y. Ouerdane, S. Girard, B. Tortech, T. Robin, C. Marcandella, A. Boukenter, B. Cadier, J. P. Meunier, and P. Crochet, “Vulnerability of rare-earth-doped fibers for space missions: origins of radiation-induced attenuation,” Proc. SPIE 7316(731617), 731617, 731617-9 (2009). [CrossRef]  

5. W. Li and M. Lu, “The effect of added O2 on the transmittance and radiation resistance of radiation resistant glasses,” Opt. Express 18(25), 26307–26312 (2010). [CrossRef]   [PubMed]  

6. M. Vivona, S. Girard, C. Marcandella, T. Robin, B. Cadier, M. Cannas, A. Boukenter, and Y. Ouerdane, “Influence of Ce codoping and H2 pre-loading on Er/Yb-doped fiber: Radiation response characterized by confocal micro-luminescence,” J. Non-Cryst. Solids 357(8–9), 1963–1965 (2011). [CrossRef]  

7. S. Girard, M. Vivona, A. Laurent, B. Cadier, C. Marcandella, T. Robin, E. Pinsard, A. Boukenter, and Y. Ouerdane, “Radiation hardening techniques for Er/Yb doped optical fibers and amplifiers for space application,” Opt. Express 20(8), 8457–8465 (2012). [CrossRef]   [PubMed]  

8. Y. Sheng, L. Yang, H. Luan, Z. Liu, Y. Yu, J. Li, and N. Dai, “Improvement of radiation resistance by introducing CeO2 in Yb-doped silicate glasses,” J. Nucl. Mater. 427(1–3), 58–61 (2012). [CrossRef]  

9. K. Arai, H. Namikawa, K. Kumata, T. Honda, Y. Ishii, and T. Handa, “Aluminium or phosphorous co-doping effects on the fluorescence and structural properties of neodymium-doped silica glass,” J. Appl. Phys. 59(10), 3430–3436 (1986). [CrossRef]  

10. T. Deschamps, N. Ollier, H. Vezin, and C. Gonnet, “Clusters dissolution of Yb3+ in codoped SiO2-Al2O3-P2O5 glass fiber and its relevance to photodarkening,” J. Chem. Phys. 136(1), 014503 (2012). [CrossRef]   [PubMed]  

11. S. Girard, J. Baggio, and J. Bisutti, “14-MeV, γ-ray and pulse X-ray radiation-induced effects on multimode silica-based optical fibers,” IEEE Trans. Nucl. Sci. 53(6), 3750–3757 (2006). [CrossRef]  

12. G. Origlio, F. Messina, S. Girard, M. Cannas, A. Boukenter, and Y. Ouerdane, “Spectroscopic studies of the origin of radiation-induced degradation in phosphorous-doped optical fibers and preforms,” J. Appl. Phys. 108(12), 123103 (2010). [CrossRef]  

13. M. Lezius, K. Predehl, W. Stower, A. Turler, M. Greiter, C. Hoeschen, P. Thirolf, W. Assmann, D. Habs, A. Prokofiev, C. Ekstrom, T. W. Hansch, and R. Holzwarth, “Radiation induced absorption in rare earth doped optical fibers,” IEEE Trans. Nucl. Sci. 59(2), 425–433 (2012). [CrossRef]  

14. areB. P. Fox, Z. V. Schneider, K. Simmons-Potter, W. J. Thomes Jr, D. C. Meister, R. P. Bambha, D. A. V. Kliner, and M. J. Söderlund, “Gamma radiation effects in Yb-doped optical fiber,” Proc. SPIE 6453(645328), 645328, 645328-8 (2007). [CrossRef]  

15. J. J. Koponen, M. J. Söderlund, H. J. Hoffman, and S. K. T. Tammela, “Measuring photodarkening from single-mode ytterbium doped silica fibers,” Opt. Express 14(24), 11539–11544 (2006). [CrossRef]   [PubMed]  

16. S. Jetschke, S. Unger, U. Röpke, and J. Kirchhof, “Photodarkening in Yb doped fibers: experimental evidence of equilibrium states depending on the pump power,” Opt. Express 15(22), 14838–14843 (2007). [CrossRef]   [PubMed]  

17. M. Engholm and L. Norin, “Preventing photodarkening in ytterbium-doped high power fiber lasers; correlation to the UV-transparency of the core glass,” Opt. Express 16(2), 1260–1268 (2008). [CrossRef]   [PubMed]  

18. R. Peretti, C. Gonnet, and A. M. Jurdyc, “Revisiting literature observations on photodarkening in Yb3+doped fiber considering the possible presence of Tm impurities,” J. Appl. Phys. 112(9), 093511 (2012). [CrossRef]  

19. K. Médjahdi, A. Boukenter, Y. Ouerdane, F. Messina, and M. Cannas, “Ultraviolet-induced paramagnetic centers and absorption changes in singlemode Ge-doped optical fibers,” Opt. Express 14(13), 5885–5894 (2006). [CrossRef]   [PubMed]  

20. D. L. Griscom, “On the nature of radiation-induced point defects in GeO2-SiO2 glasses: reevaluation of a 26-year-old ESR and optical data set,” Opt. Mater. Express 1(3), 400–412 (2011). [CrossRef]  

21. K. L. Brower, “Electron paramagnetic resonance of Al E1’centers in vitreous silica,” Phys. Rev. B 20(5), 1799–1811 (1979). [CrossRef]  

22. K. Chah, B. Boizot, B. Reynard, D. Ghaleb, and G. Petite, “Micro-Raman and EPR studies of β-radiation damages in aluminosilicate glasses,” Nucl. Inst. and Meth. in Phys. Res. B 191(1–4), 337–341 (2002).

23. F. L. Galeener and A. E. Geissberger, “Vibrational dynamics in 30Si-substituted vitreous SiO2,” Phys. Rev. B 27(10), 6199–6204 (1983). [CrossRef]  

24. A. Alessi, S. Girard, M. Cannas, A. Boukenter, and Y. Ouerdane, “Phosphorous doping and drawing effects on the Raman spectroscopic properties of O = P bond in silica-based fiber and preform,” Opt. Mater. Express 2(10), 1391–1396 (2012). [CrossRef]  

25. V. G. Plotnichenko, V. O. Sokolov, V. V. Koltashev, and E. M. Dianov, “On the structure of phosphosilicate glasses,” J. Non-Cryst. Solids 306(3), 209–226 (2002). [CrossRef]  

26. B. G. Aitken, R. E. Youngman, R. R. Deshpande, and H. Eckert, “Structure−property relations in mixed-network glasses: multinuclear solid state NMR investigations of the system xAl2O3:(30 − x)P2O5:70SiO2,” J. Phys. Chem. C 113(8), 3322–3331 (2009). [CrossRef]  

27. R. Peretti, A. M. Jurdyc, B. Jacquier, W. Blanc, and B. Dussardier, “Spectroscopic signature of phosphate crystallization in Erbium-doped optical fibre preforms,” Opt. Mater. 33(6), 835–838 (2011). [CrossRef]  

28. A. Saitoh, S. Matsuishi, C. Se-Weon, J. Nishii, M. Oto, M. Hirano, and H. Hosono, “Elucidation of codoping effects on the solubility enhancement of Er3+ in SiO2 glass: striking difference between Al and P codoping,” J. Phys. Chem. B 110(15), 7617–7620 (2006). [CrossRef]   [PubMed]  

29. D. L. Griscom, E. J. Friebele, K. J. Long, and J. W. Fleming, “Fundamental defect centers in glass: Electron spin resonance and optical absorption studies of irradiated phosphorus doped silica glass and optical fibers,” J. Appl. Phys. 54(7), 3743–3762 (1983). [CrossRef]  

30. D. L. Griscom, “Trapped-electron centers in pure and doped glassy silica: A review and synthesis,” J. Non-Cryst. Solids 357(8–9), 1945–1962 (2011). [CrossRef]  

31. A. Saitoh, S. Matsuishi, C. Se-Weon, J. Nishii, M. Oto, M. Hirano, and H. Hosono, “Elucidation of codoping effects on the solubility enhancement of Er3+ in SiO2 glass: striking difference between Al and P codoping,” J. Phys. Chem. B 110(15), 7617–7620 (2006). [CrossRef]   [PubMed]  

32. A. Monteil, S. Chaussedent, G. Alombert-Goget, N. Gaumer, J. Obriot, S. J. L. Ribeiro, Y. Messaddeq, A. Chiasera, and M. Ferrari, “Clustering of rare earth in glasses, aluminum effect: experiments and modeling,” J. Non-Cryst. Solids 348, 44–50 (2004). [CrossRef]  

33. J. Du, L. Kokou, J. L. Rygel, Y. Chen, C. G. Pantano, R. Woodman, and J. Belcher, “Structure of cerium phosphate glasses: molecular dynamics simulation,” J. Am. Ceram. Soc. 94(8), 2393–2401 (2011). [CrossRef]  

34. M. Engholm, L. Norin, and D. Aberg, “Strong UV absorption and visible luminescence in ytterbium-doped aluminosilicate glass under UV excitation,” Opt. Lett. 32(22), 3352–3354 (2007). [CrossRef]   [PubMed]  

35. S. Jetschke, S. Unger, A. Schwuchow, M. Leich, and J. Kirchhof, “Efficient Yb laser fibers with low photodarkening by optimization of the core composition,” Opt. Express 16(20), 15540–15545 (2008). [CrossRef]   [PubMed]  

36. S. Yoo, C. Basu, A. J. Boyland, C. Sones, J. Nilsson, J. K. Sahu, and D. Payne, “Photodarkening in Yb-doped aluminosilicate fibers induced by 488 nm irradiation,” Opt. Lett. 32(12), 1626–1628 (2007). [CrossRef]   [PubMed]  

37. I. Manek-Hönninger, J. Boullet, T. Cardinal, F. Guillen, S. Ermeneux, M. Podgorski, R. Bello Doua, and F. Salin, “Photodarkening and photobleaching of an ytterbium-doped silica double-clad LMA fiber,” Opt. Express 15(4), 1606–1611 (2007). [CrossRef]   [PubMed]  

38. P. D. Dragic, C. G. Carlson, and A. Croteau, “Characterization of defect luminescence in Yb doped silica fibers: part I NBOHC,” Opt. Express 16(7), 4688–4697 (2008). [CrossRef]   [PubMed]  

39. A. D. Guzman Chávez, A. V. Kir’yanov, Y. O. Barmenkov, and N. N. Il’ichev, “Reversible photo-darkening and resonant photo-bleaching of Ytterbium-doped silica fiber at in-core 977-nm and 543-nm irradiation,” Laser Phys. Lett. 4(10), 734–739 (2007). [CrossRef]  

40. F. Mady, M. Benabdesselam, and W. Blanc, “Thermoluminescence characterization of traps involved in the photodarkening of ytterbium-doped silica fibers,” Opt. Lett. 35(21), 3541–3543 (2010). [CrossRef]   [PubMed]  

41. M. Leich, U. Röpke, S. Jetschke, S. Unger, V. Reichel, and J. Kirchhof, “Non-isothermal bleaching of photodarkened Yb-doped fibers,” Opt. Express 17(15), 12588–12593 (2009). [CrossRef]   [PubMed]  

42. K. E. Mattsson, “Photo darkening of rare earth doped silica,” Opt. Express 19(21), 19797–19812 (2011). [CrossRef]   [PubMed]  

43. H. Gebavi, S. Taccheo, D. Milanese, A. Monteville, O. Le Goffic, D. Landais, D. Mechin, D. Tregoat, B. Cadier, and T. Robin, “Temporal evolution and correlation between cooperative luminescence and photodarkening in ytterbium doped silica fibers,” Opt. Express 19(25), 25077–25083 (2011). [CrossRef]   [PubMed]  

44. R. Peretti, A. M. Jurdyc, B. Jacquier, C. Gonnet, A. Pastouret, E. Burov, and O. Cavani, “How do traces of thulium explain photodarkening in Yb doped fibers?” Opt. Express 18(19), 20455–20460 (2010). [CrossRef]   [PubMed]  

45. S. Jetschke, M. Leich, S. Unger, A. Schwuchow, and J. Kirchhof, “Influence of Tm- or Er-codoping on the photodarkening kinetics in Yb fibers,” Opt. Express 19(15), 14473–14478 (2011). [CrossRef]   [PubMed]  

46. S. Jetschke, S. Unger, M. Leich, and J. Kirchhof, “Photodarkening kinetics as a function of Yb concentration and the role of Al codoping,” Appl. Opt. 51(32), 7758–7764 (2012). [CrossRef]   [PubMed]  

47. R. Peretti, C. Gonnet, and A. M. Jurdyc, “A new vision of photodarkening in Yb3+-doped fibers,” Proc. SPIE 8257, 825705, 825705-7 (2012). [CrossRef]  

48. H. Gebavi, S. Taccheo, L. Lablonde, B. Cadier, T. Robin, D. Méchin, and D. Tregoat, “Mitigation of photodarkening phenomenon in fiber lasers by 633 nm light exposure,” Opt. Lett. 38(2), 196–198 (2013). [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 (7)

Fig. 1
Fig. 1 Samples A, B, G, and F before (left) and after (right) gamma irradiation. The samples F and G, which contains an excess of aluminum compared to phosphorous, are darkened by gamma rays.
Fig. 2
Fig. 2 (a) Optical absorption spectra of the seven irradiated preforms. Samples A and B (P>Al) exhibit low absorption band in the visible range. The sharp peak around 980 nm is the absorption band of Yb3+. The spectra have been cut between 850 nm and 930 nm due to discontinuities induced by the change of the lamp during the experiment; (b) cw-ESR spectra of the seven irradiated samples. The inset displays an enlargement of the weak signal of AlOHC defects around 348 mT. All the spectra have been recorded at ambient temperature.
Fig. 3
Fig. 3 (a) cw-ESR spectra of the preform G at different power between 0.1mW and 50 mW around the AlOHC defect resonance; (b) Intensity of the AlOHC signal in function of the power square root. A linear behaviour is observed up to P = 2mW, the working power.
Fig. 4
Fig. 4 Optical absorption spectra of the preform G recorded during thermal annealing experiment up to 650 °C. A gradual decrease of the visible absorption induced by irradiation is observed.
Fig. 5
Fig. 5 (a) Evolution of the intensity of the visible absorption band; (b) the intensity of AlOHC ESR signal during the thermal annealing experiment. Both characterizations display a stable behaviour up to 200 °C, and a decrease of the signal between 200 °C and 400 °C. Above 450 °C, the visible absorption band and the AlOHC ESR signal have completely disappeared.
Fig. 6
Fig. 6 Normalized Raman spectra of preform fibers A, C, D, F (straight lines) and of pure silica glass (dotted line). Above 1000 cm−1, Al-O-P linkages are revealed in the doped samples. The P = O stretching band at 1330 cm−1 well appears when P>Al (sample A).
Fig. 7
Fig. 7 2D-HYSCORE pulse ESR spectra of the pristine samples A, B, D, E, F, and G recorded at liquid He temperature and at a magnetic field of 350 mT. A and B samples reveal off-diagonal peaks showing the proximity of Al and P nuclei around Yb3+ when phosphorous is introduced in excess compared to aluminium.

Tables (1)

Tables Icon

Table 1 Chemical core composition of the fiber preforms determined using Castaing’s microprobe. The Al/P ratio is also provided. Concentrations are in at%.

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.