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

Superbroad visible to NIR photoluminescence from Bi+ evidenced in Ba2B5O9Cl: Bi crystal

Open Access Open Access

Abstract

The nature of bismuth NIR luminescence is essential to develop the bismuth doped laser materials with high efficiency and desirable emission wavelength, and it, thereby, receives rising interests. Our previous work reported the Bi0 luminescence from Ba2B5O9Cl: Bi with a lifetime of ~30μs and the conversion of Bi2+ to Bi0. This work found indeed the conversion could be enabled in the compound by an in situ reduction technique and it, however, happens via an intermediate state of Bi+. Once the ion of Bi+ is stabilized and built into the compound, it can luminesce in a super broad spectral range from 600 to 1200nm with a lifetime longer than 1ms, due to the cascade transitions from 3P2 and 3P1 to 3P0. This is completely different from Bi0 and Bi2+ in the compound, and it has never been noticed before. We believe this work can help us better understand the complex nature of bismuth luminescence.

© 2016 Optical Society of America

1. Introduction

Infrared luminescent materials active in new spectral region of 1150-1500nm are of crucial importance to develop new fiber lasers and optical amplifiers, which could find potential applications such as in medicine, astrophysics, or next generation optical communication system [1–5 ]. Bismuth doped glasses were recently recognized as such a promising candidate which could luminesce in an extremely broad spectral range, that is, from 1000 to 1700nm, or even to 2000nm [6–12 ]. Shortly after this, the potency in fiber laser and amplifier was confirmed, and broadly tunable fiber laser was initially demonstrated from 1150nm to 1300nm, and soon extended to 1550nm [13–15 ]. These works have turned the researches on bismuth doped materials and devices into a very hot area in consequent years.

Nevertheless, one of open problems in the area is the nature of bismuth near infrared (NIR) luminescence. It remains unclear despite of the subsequent successes in bismuth fiber device. The problem comes from the multiple valence states (from positive to negative) and species (from single ion to cluster ion) of bismuth, and the easy conversion between them. It is very challenging to stabilize the element to a specific state and to identify the state with traditional techniques such as X-ray photoelectron spectrum, electron energy loss spectroscopy etc. For design of efficient bismuth doped glass in the future, it is essential to establish the origin of bismuth NIR emission. For this, many efforts have been made in experimental and theoretical aspects [16–27 ]. For instance, Sokolov et al had performed quantum chemical calculations on spectroscopic properties of Bi2 -, Bi2 2-, Bi+, Bi2 +, Bi+…VCl -, Bi…≡Si- Si≡ complex (formed by interstitial Bi atoms and glass intrinsic defects) etc [24,25 ]. Romanov et al had tried to prepare Bi+ doped single crystals of chlorides, such as CsCdCl3 [26,27 ].

Our previous work reported the Bi0 luminescence from Ba2B5O9Cl: Bi with a lifetime of ~30μs and the conversion of Bi2+ to Bi0. The luminescence peaks at 1030 and 1061nm due to substitution of Bi0 for two different types of barium sites in the compound, and it only appears after the doped sample was treated in a reducing atmosphere and it disappears once the sample was in turn sintered in an oxidizing atmosphere [21]. For the sample treated in the reducing atmosphere, will the valence of bismuth change directly from + 2 to 0 or via + 1 to 0? If the reduction reaction happened via + 1, we should be able to detect the luminescence from Bi+. Motivated by this, we doped trivalent bismuth into the sample first in air, and then reduced it in situ by hydrogen flow. We noticed that as the treatment time is prolonged gradually, Bi2+ and Bi0 luminescences reach maximum as the times are 0.5 and 2.5 hours, respectively [21]. In between, we observed a luminescence spanning from 600 to 1300nm with a lifetime longer than 1ms becomes strongest when the treating time is 1 hour. Comparison to the previous reports confirms it is from Bi+, and very different from either Bi2+ or Bi0 in luminescence wavelength and lifetime. We also observed the visible Bi+ emission at 660nm for the first time due to the transition from 3P2 to 3P0.

2. Experimental procedure

All samples of Ba2B5O9Cl: Bi were prepared by solid state reaction at high temperature. Analytic reagents BaCO3, H3BO3, BaCl2.2H2O and Bi2O3 were selected as raw materials. The samples were weighed according to Ba2(1- x%)B5O9Cl: 2x%Bi (x = 0, 0.1, 0.3, 0.5, 1.0, 1.5, 2.0, 2.5, 3.0, 5.0) and mixed, and sintered similar to reference [21]. The samples were sintered in air or in 95%N2/5%H2 for different times at 850°C to control the valence state of bismuth. Boiling temperature of Bi2O3 is 1890°C while the samples were synthesized at temperature not higher than 850°C. So, the bismuth loss due to volatility will be negligible, and the nominal content should be close to the bismuth content residual in the samples after sintered.

X-ray diffraction (XRD) patterns of the samples were recorded with a Rigaku D/max-IIIA X-ray diffractometer (40kV, 1.2° min−1, 40mA, Cu-Kα1, λ = 1.5405Å). Static excitation and emission spectra and dynamic emission decay spectra were measured with a high resolution spectrofluorometer Edinburgh Instruments FLS 920 equipped with a red sensitive single photon counting photomultiplier (Hamamatsu R928P) in Peltier air-cooled house for ultraviolet to visible range and a liquid nitrogen cooled photomultiplier (Hamamatsu R5509-72) for NIR range. A microsecond pulsed xenon flashlamp μF900 with an average power of 60W applicable to a lifetime of 1μs to 10s was used to measure the decay curves. A 450 W ozone-free xenon lamp was used for steady-state measurements. For the low temperature measurements, the sample was housed in a closed cycle helium cryostat. All the measurements were performed at room temperature except otherwise specified.

3. Results and discussion

3.1 Effect of bismuth doping on crystal structure

We examined the XRD patterns of all the samples which were prepared in air or N2/H2, and they are all pure phase of Ba2B5O9Cl. Rietveld refining results for the doped samples prepared in air reveal that the unit cell shrinks with the increase of bismuth content [21]. This is a sign that smaller Bi3+ ions have successfully substituted for Ba2+ ions. The samples were submitted for consequent treatment in the flow of nitrogen and hydrogen in different times to precipitate in situ different valences of bismuth species inside the compound.

3.2 Conversion of Bi3+ to Bi2+by in situ reduction process

The Bi doped samples prepared in air show only very weak blue luminescence from Bi3+ at room temperature. This is due to thermal quenching. As the temperature is lowered to 10K, the blue emission can be enhanced more than 6 times. Once the samples were annealed in hydrogen, red luminescence was detected immediately, and it was intensified as the annealing time increased when annealed at 850°C. It became strongest as the time prolonged to 0.5h, and the luminescence was weakened as the dwell time was longer than 0.5h. As the time was 2.5h, the intensity decreased to 60% of the sample which was annealed for 0.5h. The red luminescence is peaked at 655nm as Fig. 1(a) illustrates, and it depends tightly on the concentration of bismuth. The strongest luminescence was found at Bi% = 0.5%. As displayed in Fig. 1(b), the excitation spectrum of the emission at 655nm comprises three excitation bands located at 273, 440 and 624nm, respectively. Luminescence decay curves were collected and Fig. 1(c) illustrates one of examples, which was monitored at the emission at 655nm upon an excitation into 273nm. The decay follows well the single exponential decay equation and fitting to it produces a lifetime of 8.00μs. The lifetime shows clear dependence on temperature, and it reduces from 13.47μs @ 10K to 8.00μs @ 300K, as shown in Fig. 1(d). As we changed the excitation wavelengths to 440 or 624nm, the emission peak does not move and the lifetime is always around 10μs even at 10K. Surprisingly, the lifetime does not show the dependence on the Bi concentration. Comparing the results above to the reports on Bi2+ doped compounds confirms the red emission at 655nm is from Bi2+ ions in the compound [21]. This implies that the conversion takes place from Bi3+ to Bi2+ after in situ reduction process. The excitations at 273, 440 and 624 nm correspond to 2P1/22S1/2, 2P1/22P3/2(2), and 2P1/22P3/2(1) respectively and the emission at 655nm is due to 2P3/2(1) → 2P1/2.

 figure: Fig. 1

Fig. 1 (a) Emission spectra (λex = 273nm) of Ba2(1-x%)B5O9Cl: 2x%Bi (x = 0.1, 0.3, 0.5, 0.7, 1.0, 1.5, 2.0, 3.0, 5.0) treated in N2/H2 for 0.5h, and inset demonstrates in situ reduction; Orange balls are for barium atoms, yellow for boron, deep blue for oxygen, green for chlorine, and blue for H2 atmosphere; (b) excitation spectra (λem = 655nm) of Ba1.99B5O9Cl: 1%Bi treated in N2/H2 for 0.5h. Visible excitations were enlarged for clarity; (c) decay curves (λem = 655nm, λex = 273nm) of Ba1.99B5O9Cl: 1%Bi (0.5h treatment in N2/H2) at different temperatures as indicated and (d) the dependence of the emission lifetimes on temperature (green ball) and bismuth concentration (black ball).

Download Full Size | PDF

3.3 Conversion of Bi2+Bi+Bi0 by in situ reduction process

As the annealing time continues prolonging from 0.5h to 2.5h, Bi0 luminescence at 1055nm is intensified gradually at expense of Bi2+ emission and it reaches maximum at the dwell time of 2.5h. This reflects the valence change of bismuth from + 2 to 0 in the in situ reduction process. Beside Bi0 and Bi2+ luminescences, we found a very broad emission spectrum starting from 600 to 850nm upon 330nm excitation, as Fig. 2(a) shows, and it comprises two peaks at 660 and 790nm. Since the shape of the peak at 790nm implies there should be a tail in the longer wavelength, we detected the emission spectrum with another NIR photomultiplier.

 figure: Fig. 2

Fig. 2 (a) Emission spectra (λex = 330nm) of Ba1.99B5O9Cl: 1%Bi treated in N2/H2 for different time; (b) emission spectra (λex = 330nm) of Ba2(1-x%)B5O9Cl: 2x%Bi (x = 0.1, 0.5, 1.0, 1.5, 5.0) treated in N2/H2 for 1h detected by visible and NIR photomultipliers; The spectra were rescaled and combined; (c) excitation spectra (λem = 790nm, λem = 970nm) of Ba1.99B5O9Cl: 1%Bi treated in N2/H2 for 1h and 0h; (d) decay curves of Ba1.99B5O9Cl: 1%Bi treated in H2/N2 for 1h.

Download Full Size | PDF

Unexpectedly, a new emission peak was found centering at 970nm beside the 790nm emission, as shown in Fig. 2(b). The corresponding excitation spectrum of the emission at 790nm is depicted in Fig. 2(c), and it is made up of two new strong peaks at 330nm and 606nm along with a weak peak at 478nm from Bi0. The excitation spectrum of the 970nm emission comprises two strong peaks at ~320nm and ~670nm. This is different from the case of the 790nm emission. Perhaps, the excitation peak at 606nm has immerged into the tail of the excitation peak at 670nm since the 670nm peak is very broad. The excitation peaks of Bi0 become slightly stronger as compared to that of the emission at 790nm. This should be due to the more overlap between the 970nm emission and the Bi0 luminescence. Figure 2(d) illustrates the decay curves of the emissions at 660nm, 790nm and 970nm upon the excitation of 330nm. The decay of the 790nm emission complies well with the single exponential decay equation, and the lifetime is 1.151ms. The decays of the emissions at 660nm and 970nm are different from the 790nm emission. They experience a fast decay rightly followed by a slow decay, which should be due to the overlaps to Bi2+ and Bi0 emissions. The fast decay lies within 200μs and it is contributed mainly by Bi2+ and Bi0 species. The slow decays show the intrinsic lifetimes of the emissions at 660nm and 970nm to be 1.149 and 1.192ms, respectively. So the lifetimes of the emissions at 660nm, 790nm and 970nm are in same order of millisecond irrespective of upon which excitation wavelength it is, as displayed in Fig. 2(d), and they are very different from either Bi2+ or Bi0 species. So, what's the origin of these emissions?

The evolution of the emission along with the annealing time in hydrogen may give us the inspiration on it. Look back on Fig. 2(a), what we find is that the air sample does not show any luminescence in the spectral range from 600 to 850nm upon 330nm excitation while the annealed samples all exhibit the broad luminescence. As the annealing time increases, the emission at 660nm decreases all the way down, and the emission at 790nm increases first and it reaches the maximum as the time extends from 0.5h to 1.0h, afterwards, it is weakened rapidly. Since the time when the intensity maximum of the emission at 790nm is 1.0h, and it is right in between 0.5h for Bi2+ and 2.5h for Bi0, we believe the species of bismuth which is responsible for the long lived luminescence is most probably Bi+. This is because Bi+ is the only intermediate state between Bi2+ and Bi0, and also because it is very difficult to precipitate bismuth metallic colloid or polycations inside the compound due to the limit space of barium sites available to bismuth species in Ba2B5O9Cl.

In molten salts of AlCl3-NaCl, Bjerrum et al found five absorptions at 308, 333, 585, 658, 694 and 901 nm from Bi+, and they assigned 308 and 333 nm to the transition of 3P01D2, 585, 658 and 694 nm to 3P03P2, and 901nm to 3P03P1, respectively [28]. We did not observe all similar peaks. But as we compared our results to theirs, it can easily come to the conclusion that the excitation peak at 330 nm may be due to 3P01D2 and the peaks at 606 and 670nm are to 3P03P2, and the emissions at 660nm and 790nm may be to 3P23P0, and the emission at 970nm is to 3P13P0. The lifetime of the emissions at 660nm and 790nm are almost same, perhaps due to the same origin of transitions from 3P2 which is split by crystal field. It, however, is slightly shorter than the lifetime of the emission at 970nm as shown in Fig. 2(d). This is because the transition of 3P13P0 is slightly more forbidden. Romanov et al found the absorptions at 601nm and 660nm and an emission at 950nm with a lifetime of 400μs in KMgCl3: Bi+, and the absorptions at 596nm and 644nm and an emission at 980nm with a lifetime of 525μs in KAlCl4: Bi+ [26], and the absorptions at 610nm and 650nm and an emission at 980nm with a lifetime longer than 300μs in CsCdCl3: Bi+ [27]. These results are similar to this work and confirm the assignments. In view of radius match, Bi+ ions once formed will substitute barium sites, which are coordinated by seven oxygen and two chlorine atoms and located in the center of channels formed by [B5O9] chains and [BO3] units. They will experience stronger crystal field due to the enhanced interaction between Bi+, BO3 and BO4 units in the host as compared to the free state, and their absorptions and emission, therefore, lie at longer wavelengths than free Bi+ ions. The decay of Bi+ luminescence follows well the single exponential decay equation at different temperatures between 10 to 300K. The lifetime of the 790nm emission upon 330nm excitation depends tightly on temperature. It is 2.071, 2.041, 1.944, 1.754, 1.476ms at 10, 100, 150, 200, 250K, respectively.

4. Conclusions

In all, the valence conversion happens from Bi2+ via Bi+ to Bi0 in Ba2B5O9Cl: Bi by the in situ reduction process. The intermediate species shows extraordinarily broad luminescence from 600 to 1200nm with the lifetime longer than 1ms, due to the cascade transitions from 3P2 and 3P1 to 3P0, and it is rather different from Bi0 and Bi2+ in the compound. The lifetimes of Bi0, Bi+ and Bi2+ show similar dependence on temperature in the range from 10 to 300K. This work deepens our understanding on the intriguing nature of bismuth luminescence. It should be helpful for design of efficient bismuth laser materials in future.

Acknowledgments

This work was financially supported by National Natural Science Foundation of China (Grant No. 51322208), Guangdong Natural Science Foundation for Distinguished Young Scholars (Grant No. S20120011380), The Department of Education of Guangdong Province (Grant No. 2013gjhz0001), Fundamental Research Funds for the Central Universities, and Fok Ying Tong Education Foundation (Grant No. 132004).

References and links

1. A. Mori, Y. Ohishi, and S. Sudo, “Erbium-doped tellurite glass fibre laser and amplifier,” Electron. Lett. 33(10), 863–864 (1997). [CrossRef]  

2. Y. Fujimoto and M. Nakatsuka, “Optical amplification in bismuth-doped silica glass,” Appl. Phys. Lett. 82(19), 3325–3326 (2003). [CrossRef]  

3. M. Peng, J. Qiu, D. Chen, X. Meng, I. Yang, X. Jiang, and C. Zhu, “Bismuth- and aluminum-codoped germanium oxide glasses for super-broadband optical amplification,” Opt. Lett. 29(17), 1998–2000 (2004). [CrossRef]   [PubMed]  

4. E. Dianov, V. Dvoyrin, V. Mashinsky, A. Umnikov, M. Yashkov, and A. Gur’yanovb, “CW bismuth fibre laser,” Quantum Electron. 35(12), 1083–1084 (2005). [CrossRef]  

5. H. Guo, Z. Li, H. Qian, Y. Hu, and I. N. Muhammad, “Seed-mediated synthesis of NaY F4:Y b, Er/NaGdF4 nanocrystals with improved upconversion fluorescence and MR relaxivity,” Nanotechnology 21(12), 125602 (2010). [CrossRef]   [PubMed]  

6. M. A. Hughes, T. Akada, T. Suzuki, Y. Ohishi, and D. W. Hewak, “Ultrabroad emission from a bismuth doped chalcogenide glass,” Opt. Express 17(22), 19345–19355 (2009), https://www.osapublishing.org/oe/abstract.cfm?uri=oe-17-22-19345. [CrossRef]   [PubMed]  

7. S. Khonthon, S. Morimoto, Y. Arai, and Y. Ohishi, “Redox equilibrium and NIR luminescence of Bi2O3-containing glasses,” Opt. Mater. 31(8), 1262–1268 (2009). [CrossRef]  

8. M. Peng, Q. Zhao, J. Qiu, and L. Wondraczek, “Generation of emission centers for broadband NIR luminescence in bismuthate glass by femtosecond laser irradiation,” J. Am. Ceram. Soc. 92(2), 542–544 (2009). [CrossRef]  

9. Q. Sheng, Q. Zhou, and D. Chen, “Efficient methods of obtaining good optical properties in Yb-Bi co-doped phosphate glasses,” J. Mater. Chem. C Mater. Opt. Electron. Devices 1(18), 3067–3071 (2013). [CrossRef]  

10. X. Wang, Q. Sheng, L. Hu, and J. Zhang, “Observation of broadband infrared luminescence in a novel Bi-doped P2O5–B2O3–Al2O3 glass,” Mater. Lett. 66(1), 156–158 (2012). [CrossRef]  

11. Z. Song, C. Li, Y. Li, Z. Yang, D. Zhou, Z. Yin, X. Wang, Q. Wang, T. Hau, Z. Zhao, Y. Yang, X. Yu, and J. Qiu, “The influence of alkali ions size on the superbroadband NIR emission from bismuth-doped alkali aluminoborophosphsilicate glasses,” Opt. Mater. 35(1), 61–64 (2012). [CrossRef]  

12. R. Cao, M. Peng, L. Wondraczek, and J. Qiu, “Superbroad near-to-mid-infrared luminescence from Bi53+ in Bi5(AlCl4)3.,” Opt. Express 20(3), 2562–2571 (2012). [CrossRef]   [PubMed]  

13. I. Razdobreev, L. Bigot, V. Pureur, A. Favre, G. Bouwmans, and M. Douay, “Efficient all-fiber bismuth-doped laser,” Appl. Phys. Lett. 90(3), 031103 (2007). [CrossRef]  

14. S. Kivisto, J. Puustinen, M. Guina, O. Okhotnikov, and E. Dianov, “Tunable modelocked bismuth-doped soliton fibre laser,” Electron. Lett. 44(25), 1456–1458 (2008). [CrossRef]  

15. I. Bufetov and E. Dianov, “Bi-doped fiber lasers,” Laser Phys. Lett. 6(7), 487–504 (2009). [CrossRef]  

16. M. Peng, N. Da, S. Krolikowski, A. Stiegelschmitt, and L. Wondraczek, “Luminescence from Bi2+-activated alkali earth borophosphates for white LEDs,” Opt. Express 17(23), 21169–21178 (2009), https://www.osapublishing.org/oe/abstract.cfm?uri=oe-17-23-21169. [CrossRef]   [PubMed]  

17. M. Peng and L. Wondraczek, “Photoluminescence of Sr2P2O7:Bi2+ as a red phosphor for additive light generation,” Opt. Lett. 35(15), 2544–2546 (2010). [CrossRef]   [PubMed]  

18. L. Su, J. Yu, P. Zhou, H. Li, L. Zheng, Y. Yang, F. Wu, H. Xia, and J. Xu, “Broadband near-infrared luminescence in γ-irradiated Bi-doped α-BaB2O4 single crystals,” Opt. Lett. 34(16), 2504–2506 (2009). [CrossRef]   [PubMed]  

19. L. Su, H. Zhao, H. Li, L. Zheng, G. Ren, J. Xu, W. Ryba-Romanowski, R. Lisiecki, and P. Solarz, “Near-infrared ultrabroadband luminescence spectra properties of subvalent bismuth in CsI halide crystals,” Opt. Lett. 36(23), 4551–4553 (2011). [CrossRef]   [PubMed]  

20. M. Peng, B. Sprenger, M. A. Schmidt, H. G. Schwefel, and L. Wondraczek, “Broadband NIR photoluminescence from Bi-doped Ba2P2O7 crystals: insights into the nature of NIR-emitting Bismuth centers,” Opt. Express 18(12), 12852–12863 (2010), https://www.osapublishing.org/oe/abstract.cfm?uri=oe-18-12-12852. [CrossRef]   [PubMed]  

21. J. Zheng, M. Peng, F. Kang, R. Cao, Z. Ma, G. Dong, J. Qiu, and S. Xu, “Broadband NIR luminescence from a new bismuth doped Ba2B5O9Cl crystal: evidence for the Bi0 model,” Opt. Express 20(20), 22569–22578 (2012), https://www.osapublishing.org/oe/abstract.cfm?uri=oe-20-20-22569. [CrossRef]   [PubMed]  

22. M. Peng, G. Dong, L. Wondraczek, L. Zhang, N. Zhang, and J. Qiu, “Discussion on the origin of NIR emission from Bi-doped materials,” J. Non-Cryst. Solids 357(11–13), 2241–2245 (2011). [CrossRef]  

23. M. Peng, C. Zollfrank, and L. Wondraczek, “Origin of broad NIR photoluminescence in bismuthate glass and Bi-doped glasses at room temperature,” J. Phys. Condens. Matter 21(28), 285106 (2009). [CrossRef]   [PubMed]  

24. V. Sokolov, V. Plotnichenko, and E. Dianov, “The origin of near-IR luminescence in bismuth-doped silica and germania glasses free of other dopants: First-principle study,” Opt. Mater. Express 3(8), 1059–1074 (2013). [CrossRef]  

25. V. Sokolov, V. Plotnichenko, V. Koltashev, and E. Dianov, “Centres of broadband near-IR luminescence in bismuth-doped glasses,” J. Phys. D Appl. Phys. 42(9), 095410 (2009). [CrossRef]  

26. A. Romanov, A. Veber, Z. Fattakhova, O. Usovich, E. Haula, L. Trusov, P. E. Kazin, V. N. Korchak, V. B. Tsvetkov, and V. B. Sulimov, “Subvalent bismuth monocation Bi+ photoluminescence in ternary halide crystals KAlCl4 and KMgCl3,” J. Lumin. 134, 180–183 (2013). [CrossRef]  

27. A. Romanov, A. Veber, Z. Fattakhova, D. Vtyurina, M. Kouznetsov, K. Zaramenskikh, I. S. Lisitsky, V. N. Korchak, V. B. Tsvetkov, and V. B. Sulimov, “Spectral properties and NIR photoluminescence of Bi+ impurity in CsCdCl3 ternary chloride,” J. Lumin. 149, 292–296 (2014). [CrossRef]  

28. N. Bjerrum, C. Boston, and G. Smith, “Lower oxidation states of bismuth. Bi+ and Bi53+ in molten salt solutions,” Inorg. Chem. 6(6), 1162–1172 (1967). [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 (2)

Fig. 1
Fig. 1 (a) Emission spectra (λex = 273nm) of Ba2(1-x%)B5O9Cl: 2x%Bi (x = 0.1, 0.3, 0.5, 0.7, 1.0, 1.5, 2.0, 3.0, 5.0) treated in N2/H2 for 0.5h, and inset demonstrates in situ reduction; Orange balls are for barium atoms, yellow for boron, deep blue for oxygen, green for chlorine, and blue for H2 atmosphere; (b) excitation spectra (λem = 655nm) of Ba1.99B5O9Cl: 1%Bi treated in N2/H2 for 0.5h. Visible excitations were enlarged for clarity; (c) decay curves (λem = 655nm, λex = 273nm) of Ba1.99B5O9Cl: 1%Bi (0.5h treatment in N2/H2) at different temperatures as indicated and (d) the dependence of the emission lifetimes on temperature (green ball) and bismuth concentration (black ball).
Fig. 2
Fig. 2 (a) Emission spectra (λex = 330nm) of Ba1.99B5O9Cl: 1%Bi treated in N2/H2 for different time; (b) emission spectra (λex = 330nm) of Ba2(1-x%)B5O9Cl: 2x%Bi (x = 0.1, 0.5, 1.0, 1.5, 5.0) treated in N2/H2 for 1h detected by visible and NIR photomultipliers; The spectra were rescaled and combined; (c) excitation spectra (λem = 790nm, λem = 970nm) of Ba1.99B5O9Cl: 1%Bi treated in N2/H2 for 1h and 0h; (d) decay curves of Ba1.99B5O9Cl: 1%Bi treated in H2/N2 for 1h.
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.