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

On the origin of near-IR luminescence in Bi-doped materials (II). Subvalent monocation Bi+ and cluster Bi53+ luminescence in AlCl3/ZnCl2/BiCl3 chloride glass

Open Access Open Access

Abstract

Broadband NIR photoluminescence (from 1000 to 2500 nm) was observed from partially reduced AlCl3/ZnCl2/BiCl3 glass, containing subvalent bismuth species. The luminescence consists of three bands, assigned to Bi+, Bi24+, and Bi53+ ions. The physical and optical characteristics of these centers and possible contribution to NIR luminescence from bismuth-doped oxide glasses are discussed.

©2012 Optical Society of America

1. Introduction

Bismuth-doped materials possess intriguing broadband Near-infrared (NIR) luminescence (1100-1700 nm), coinciding well with telecommunication window and show promises for new optical amplifiers and oscillators with outstanding characteristics [1,2]. A number of silicate [1], phosphate [3], borate [4], sulfide [5] and fluoride [6] glasses as well as RbPb2Cl5 [7], BaB2O4 [8], BaF2 [9], Ba2P2O7 [10], CsI [11], zeolite [12,13] and sodalite [14] crystals, became NIR luminescent being doped with bismuth. The luminescence of bismuth ions in common oxidation state +3 is well known [15], but it takes place in visible, not NIR, so another less trivial sources of NIR luminescence in bismuth-doped materials should be suggested. Recently, it was demonstrated by synproportionation reaction of metallic Bi with Bi3+ in glasses [16] and ionic liquids [17], that subvalent (i.e., lower than usual +3 oxidation state) bismuth species are responsible for NIR luminescence in these media.

Bismuth possesses an interesting ability to form numerous monoatomic and cluster polycation subvalent states: Bi2+ [18], Bi+ [1921], Bi24+ [2224], Bi3+ [25], Bi53+ [19,21,26], Bi5+, Bi62+ [27,28], Bi82+ [29,30], Bi95+ [20,24,31]. All of them can be formed in acidic media via synproportionation of Bi3+ with Bi0 or by Bi3+ reduction.

Usually several bismuth subvalent species exist in the media simultaneously, and their concentrations depend on equilibrium constants of interconversion [19,21], like

Bi53++Bi3+6Bi+

That is why several subvalent bismuth species can in principle contribute to NIR luminescence. Until now, only polycation Bi53+ was definitely shown to possess very broadband NIR luminescence in Bi5(AlCl4)3 crystals [32,33]. Bi+ monocation was also suggested as possible source of NIR luminescence [3,4], but in this case evidences are still indirect. So, real contribution of all bismuth subvalent states into NIR luminescence in different bismuth-doped materials is unclear.

Here we try to clarify this problem by studying the NIR luminescence from subvalent bismuth species in chloride glass with composition: AlCl3-ZnCl2-BiCl3. We have investigated this glass system for several reasons:

  • 1. AlCl3-ZnCl2 chloride melts possess enough Lewis acidity to facilitate the stabilization of subvalent bismuth species [19], if we try to produce them by controlled reduction of BiCl3, added to this melt. Previously, subvalent bismuth stabilization and NIR luminescence was also reported for acidic chloroaluminate ionic liquid [17] and another halide Lewis acidic systems like fluoride glasses [6].
  • 2. Absorption spectra of different bismuth subvalent species in molten chlorides are well studied [19,21] simplifying their characterization. Luminescence of Bi53+ polycation in chloroaluminate crystals was also characterized recently [32,33]. This information can help to assign luminescent centers in bismuth-containing chloride glass.
  • 3. It is known, that low temperature melting chloride glasses are easily formed in AlCl3-ZnCl2 system [34]. Low temperature simplifies the control on glass preparation in sealed fused silica cells. We have found that small addition of BiCl3 to this composition did not decrease its glass-forming ability and glassy AlCl3/ZnCl2/BiCl3 specimens could be produced with optical quality by cooling the melt to the room temperature. Subvalent bismuth species can be introduced into this glass, if proper amount of reductive agent is added to chloride melt.

2. Experimental

Dry AlCl3 (99.999%) ZnCl2 (99.999%) BiCl3 (99.998%) were used for glass preparation. All manipulations with the starting chlorides (weighing and transfer to fused silica cell) were performed in argon-filled glovebox (<2ppm H2O) due to their extreme hygroscopicity. The proper amounts of chlorides (60% molar ZnCl2, 38% molar AlCl3, 2% molar BiCl3, approximately 4 g batch) were placed into L-shaped fused silica cell (Fig. 1 ). Then, the cell was capped off at the standard-taper joint with special cap, equipped with a vacuum stopcock. It was taken from the glovebox, evacuated, filled with 200 Torr He and sealed at the preformed waist. The fused silica cell also had a little appendix, perpendicular to the plane of L figure. Before the main operations in the glovebox, this appendix was filled with molten zinc metal, which after solidification consolidated with appendix. This zinc metal could not form direct contact with melt, when the chloride composition was melted in the horizontally placed cell with raised appendix (Fig. 1). This isolated metal position allows slow reduction of Bi3+ in chloride melt during melting process with convenient control on the concentration of subvalent bismuth species, introduced into resulting glass.

 figure: Fig. 1

Fig. 1 Sealed L-shaped fused silica cell with thin layer of AlCl3/ZnCl2/BiCl3 glass. The glass orange color is due to the presence of subvalent bismuth species (mainly Bi53+ polycation).

Download Full Size | PDF

The filled cell was heated initially to 473K, then to 573K until moderately colored thermochromic melt was formed. The color and thermochromic properties of our specimens (violet color at 573K and orange at room temperature) was similar to previously reported for chloride melts, containing subvalent bismuth [19]. Upon cooling, the melt not only changes color, but also become viscous, and at this moment it can be spread in a thin layer (approximately 1 mm thick) over the wall of short compartment of L-shaped cell. The walls of this short compartment were flattened and overall dimensions of compartment allow its easy insertion into standard socket of spectrophotometer.

After the final solidification of chloride melt at room temperature it forms a thin transparent glass layer, attached to the wall of the cell. At 77K the glass became very fragile; it was detached from the wall and usually became broken into several pieces. At the room temperature these pieces slowly stick together and to the walls of the cell, indicating, that glass transition of the material occurs near the room temperature.

The optical absorption spectra of AlCl3/ZnCl2/BiCl3 glass were obtained with Shimadzu UV-3600 UV-VIS-NIR spectrophotometer. Photoluminescence spectra were measured using an ARC SpectraPro SP-305 monochromator combined with an ARC ID-441C InGaAs or ID-443 InSb detector for the ranges of 900-1600 nm and 1600-2500 nm respectively. Emission decay curves were obtained with Hamamatsu G8372-01 InGaAs photodiode coupled with a wideband (100 Hz–5 MHz) low-noise preamplifier mounted on the exit slit of the monochromator. Osram HBO 150W Xe-lamp combined with LOMO MDR-12 monochromator was used as variable wavelength light source to obtain luminescence excitation spectra. The measurements were performed at 300 and 77K temperatures. All the obtained spectra were corrected for the spectral response of the system.

3. Results and discussion

The optical absorption spectrum of AlCl3/ZnCl2/BiCl3 glass, prepared, as described above is shown in Fig. 2(a) .

 figure: Fig. 2

Fig. 2 (a) Absorption spectrum of AlCl3/ZnCl2/BiCl3 glass, containing subvalent bismuth species; (b) Photoluminescence spectra of AlCl3/ZnCl2/BiCl3 glass (300K), excited at different wavelengths normalized with respect to the 1320 nm emission.

Download Full Size | PDF

The spectrum contains overlapped bands at 390, 430, 500, 750, 850 nm, corresponding well to the spectra of Bi53+ polycation in chloride solvents, reported previously [19,21]. In acidic chloride melts with excess Bi3+ the polycation Bi53+ usually exists in equilibrium with only one another subvalent bismuth form: the monocation Bi+ (Eq. (1)) [19]. The high temperature and increasing acidity shifts equilibrium to the preferable Bi+ formation. In our case the dominant subvalent form is Bi53+, since the temperature of our specimens during the spectrum recording is low during spectrum recording. At elevated temperatures (300°C) the Bi+ concentration in the melt became much higher, changing the color of the melt to violet, but upon cooling to the room temperature the Bi53+ turn to be dominant. The equilibrium reaction (Eq. (1)) is not frozen in the glass even at ambient conditions due to the low glass transition temperature [34]. Only a very weak absorption line due to Bi+ at 660 nm is evident in absorption spectrum at room temperature, whereas the second (and stronger) Bi+ absorption line at 580 nm is obscured by overlapping with overwhelming Bi53+ absorption. From reported molar absorptivities of Bi53+ and Bi+ [21] we can estimate their concentration ratio as [Bi53+]/[Bi+]≈10 at the room temperature. These subvalent Bi53+ and Bi+ cations are formed in AlCl3/ZnCl2/BiCl3 melt by reduction of BiCl3 with metal Zn:

BiCl3+ZnBi++ZnCl2+Cl
5BiCl3+6ZnBi53++6ZnCl2+3Cl

These reactions proceed slowly in the cell even without direct contact of Zn metal with BiCl3-containing chloride melt, may be because of non-negligible BiCl3 vapor pressure at our glass melting temperature. The progress of reaction is seen as melt coloration became deeper.

The NIR luminescence spectra of subvalent bismuth containing AlCl3/ZnCl2/BiCl3 glass at the room temperature are shown in Fig. 2(b). It is evident, that luminescence consists of two bands, peaked at 1080 and 1300 nm, excited differently by 450 and 532 nm radiation. At room temperature no luminescence was observed in the range 1700-2500nm with InSb detector and at the highest sensitivity of the measurement system. After glass cooling to 77K those two bands in luminescence spectrum became stronger and sharper and still another very broad luminescence band, covering interval from 1300 to 2500 nm (maximum at 2000 nm) had appeared (Fig. 3(a) ).

 figure: Fig. 3

Fig. 3 (a) Photoluminescence spectra of AlCl3/ZnCl2/BiCl3 glass, measured at 77 and 300K (excitation wavelength λex = 532 nm). The spectrum at 77K composed from two separate measurements: at short waves (with InGaAs detector), and at long waves (with InSb detector); (b) Photoluminescence excitation spectra, measured for three luminescence bands. Spectra of the first and the second bands (λem = 1060 nm and λem = 1300nm, respectively) were measured at 300K. Excitation spectrum of the third band (λem = 1900 nm) was measured at 77K.

Download Full Size | PDF

The luminescence excitation spectra of all the three bands are shown in Fig. 3(b). All excitation spectra rise at λ<350 nm (nonspecific feature) due to excitation of charge transfer transition of chlorine-coordinated Bi3+ with subsequent energy transfer to all NIR-luminescent species. Besides this common UV feature, the excitation spectrum for the first band (measured at 1060 nm emission) consists of two well resolved peaks (at 580 and 660 nm), which are characteristic to Bi+ absorption spectrum in chloroaluminate and chlorogallate systems [19,21].

The luminescence excitation spectrum of 1300 nm band (second band) is represented by the single peak at 580 nm and it doesn’t match well both Bi+ and Bi53+ absorption spectra. Yet, this excitation spectrum coincides well with absorption (with single maximum at 580 nm), reported for Bi+ in reduced BiCl3 melts [35]. The presence of Bi+ in reduced BiCl3 melt was demonstrated both by thermodynamic and spectroscopic methods [3537], although its absorption spectrum was very different from more studied and properly assigned spectrum of Bi+ in chloroaluminate and chlorogallate melts [19,21,38]. It was noticed, however, that both thermodynamic and spectroscopic methods could not distinguish between genuine monocation Bi+ and its probable complex with arbitrary number of Bi3+ from solvent BiCl3 excess [25]. So, the absorption at 580 nm in reduced BiCl3 melt, as well as 1300 nm luminescence with excitation peak at 580 nm in our glass can both originate from Bi+ complex: Bi+·nBi3+. There is one known example of such clusters: Bi24+ (n = 1), reported by X-ray crystallography investigations of different systems [2224], although no study of possible NIR luminescence from these systems has been undertaken until today. Clusters with n>1 unlikely exists due to strong electrostatic repulsion from the large overall charge. That’s why, it is reasonable to consider only dimer Bi24+ as a possible emitter of 1300 nm luminescence band. This hypothesis also agrees well with the observation that in some bismuth-doped materials luminescence intensity is proportional to the square of introduced bismuth concentration [39]. We also should remark, that all bismuth polycations possesses some degree of covalent bonding with surrounding counterions, so, notations like Bi24+ always represent considerable simplification, especially for the chalcogen [22] and trifluoroacetate [23] systems. Generally speaking, terms like Bi24+ should be perceived here in a broad sense - as a designation for dimer of bismuth in +2 valence state.

Finally, the luminescence excitation spectrum for long-wavelength emission (measured at 1900 nm, third band) coincides well with features of Bi53+ absorption spectrum (peak near 800 nm and steep rising at λ<500 nm). Two peaks, characteristic for Bi+ excitation are also present in 1900 nm luminescence excitation spectrum. This fact can be explained by initial Bi+ excitation with possible energy transfer from high excited state of Bi+ to Bi53+ polycation. This identification of the long-wavelength emission center as Bi53+ is in complete agreement with recent luminescence measurements from Bi53+ in Bi5(AlCl4)3 crystal [33]. In this system very broad luminescence continuum, reaching 3500 nm was observed. Here we report the first observation of broadband Bi53+ luminescence in glass materials.

The temporal decay of the luminescence in all three bands had been studied at 77K (Fig. 4 ) with an excitation at 532nm. The room temperature data for the temporal decay were too noisy to be correctly approximated.

 figure: Fig. 4

Fig. 4 Low temperature (77K) luminescence decay curves measured at 1350nm(a), 1900nm(b) and 1050nm(c).

Download Full Size | PDF

It was found, that single exponential curve poorly represents all obtained decay data. Decay curve for the emission in the first band (Bi+ luminescence, Fig. 4(c)) can be best represented as a sum of three exponents. The fast component has the lifetime (τ1) <0.2μs, which is beyond the response of the detector and cannot be resolved in our system. The slow component in the first band at 77K has a lifetime (τ3)~266μs which is similar to Bi+ luminescence in chloride crystal RbPb2Cl5 [7]. The positions of emission maxima are also coincided well.

Decay data for the second and the third luminescence bands are well approximated by so called “stretched” exponent I ~exp[-(t/τ)0.5]. One of the possible explanations for such temporal dependence is that dipole-dipole energy transfer quenching is observed in this case [40]. However at this moment we have not enough data to confirm or refute this hypothesis.

Cao et al. [33] demonstrated, that Bi53+ luminescence decay in Bi5(AlCl4)3 was a single exponent with lifetimes changing from 23 μs (10K) to 6 μs(300K). Not only the lifetime, but also the luminescence intensity diminished at room temperature, indicating increased quenching. In our glass the luminescence of Bi53+ (third band) is completely quenched at the room temperature, while at 77K its lifetime can be estimated as τ~18μs which is comparable with Cao et al. data.

Steep decreasing of luminescence intensity and lifetimes at warming to the room temperature allows suggesting the multiphonon relaxation as a primary mechanism of Bi53+ excited state depopulation or possible relaxation via the conical intersection. The fact, that at the room temperature the luminescence from Bi53+ is completely quenched in the chloride glass, whereas it is still observed in Bi5(AlCl4)3 crystal, argues for multiphonon relaxation, since phonon spectrum in glasses are broadened (compared to crystals), speeding up quenching process. However depletion mechanism of observed luminescent excited levels cannot be clearly determined from our measurements and further investigations are needed.

The most interesting feature of Bi53+ luminescence is the unprecedented spectral width. Recently, we had calculated (by spin-orbit CI calculation with relativistic pseudopotential) the energies and oscillator strengths for Bi53+ excited states up to 4 eV [41]. It was shown, that at the ground state equilibrium geometry (trigonal bipyramid, D3h point symmetry group) the lowest excited state is E” (1.64 eV) with forbidden electrodipole transition to the ground state A1'. The polycation Bi53+ in double degenerate state E” can undergo a geometric distortion from D3h bipyramidal geometry due to Jahn-Teller effect. The possibility of Jahn-Teller effect in Bi53+ lowest excited state can explain the broad luminescence spectrum and large Stokes shift, since equilibrium geometry of Bi53+ is very different in ground and excited luminescent states. The additional experimental observations (such as temperature dependence of Bi53+ luminescence spectrum), analysis of possible distortion of Bi53+ geometry in the ground and excited states by the crystal field along with the theoretical investigations of Bi53+ potential energy surfaces should be considered to clarify the possibility of Jahn-Teller effect in Bi53+ lowest excited state.

4. From chloride to oxide glasses

Let us try to compare the luminescence bands assignment, obtained for the chloride glass, with the situation in oxide glasses, attracting much practical interest. The lack of luminescence from Bi53+ at room temperature, relatively short luminescence lifetime at 77K and different spectral range (red-shifted, relative to luminescence in oxide glasses) indicate that Bi53+ hardly contributes to room temperature NIR luminescence in oxide glasses. If we assume, that multiphonon relaxation is the main quenching process for excited Bi53+ at room temperature, this quenching will be even more severe in the oxide matrices (relative to chlorides) since phonon spectrum spans to higher energies here. On the other hand, demonstration of broadband luminescence from Bi53+ in chloride glass indicates, that bismuth and other post-transition metal polycations could be attractive active centers for broadband amplifiers and oscillators. Previously, another heavy metal polycation Hg32+ was shown to be luminescent, although in the visible range [42].

The luminescence in the first band (from Bi+) and in the second band (presumably from Bi24+) in our glass resembles the situation in oxide glass. Indeed, it was demonstrated [43], that germanate glasses have two luminescent bands: shortwave (1100 nm) with two-maxima (550, 720 nm) in excitation spectrum, resembling excitation spectrum of Bi+ in our glass, and longwave (1300 nm), with excitation profile with one maximum near 550 nm similar to what we assume for Bi24+. Increasing of bismuth concentration in germanate glass diminished luminescence at 1100 nm, but increased luminescence at 1300 nm. It is exactly what we should expect, supposing 1100 nm luminescence originating from Bi+ and 1300 nm one from Bi24+, since increased Bi3+ concentration shift equilibrium toward Bi24+:

Bi++Bi3+Bi24+

In SiO2-Al2O3 bismuth doped glasses relatively shortwave 1100 nm NIR luminescence is also characterized by two peaks in excitation spectrum (at 705 and 510 nm) [44]. By analogy with chloride glass we can suppose that this emission also originates from Bi+ monocation.

4. Conclusions

Lewis acidic chloride glass AlCl3/ZnCl2/BiCl3, doped with subvalent bismuth was prepared. At the room temperature the dominant subvalent bismuth form was Bi53+ polycation, while Bi+ monocation was a minor constituent. Despite this fact, the NIR luminescence at room temperature originates only from Bi+ (peak at 1080 nm) and, probably, from its association product with Bi3+: dimer Bi24+ (peak at 1300 nm). At 77K broadband luminescence from Bi53+ emerges, spanning the range from 1300 to 2500 nm. Comparison of the luminescence characteristics allows suggesting, that Bi+ and Bi24+ also contribute to luminescence of oxide-based glasses, while Bi53+ does not. This situation exemplifies one possible problem with bismuth-doped glasses: after the preparation they can incorporate several subvalent bismuth species simultaneously and some of them, being non-luminescent, are either useless or introduce optical losses to active media. To circumvent this problem, the proper preparation conditions should be chosen. For example, to maximize the NIR luminescent Bi+ concentration and minimize unnecessary Bi53+ the glass transition temperature and glass acidity should be as high as possible.

Acknowledgments

Financial support from both the Russian Ministry of Education and Science and OOO “Dimonta” contract 07.514.11.4059 of October 12, 2011 is gratefully acknowledged.

References and links

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

2. E. M. Dianov, “Bi-doped glass optical fibers: is it a new breakthrough in laser materials?” J. Non-Cryst. Solids 355(37-42), 1861–1864 (2009). [CrossRef]  

3. X. G. Meng, J. R. Qiu, M. Y. Peng, D. P. Chen, Q. Z. Zhao, X. W. Jiang, and C. S. Zhu, “Near infrared broadband emission of bismuth-doped aluminophosphate glass,” Opt. Express 13(5), 1628–1634 (2005). [CrossRef]   [PubMed]  

4. X. G. Meng, J. R. Qiu, M. Y. Peng, D. P. Chen, Q. Z. Zhao, X. W. Jiang, and C. S. Zhu, “Infrared broadband emission of bismuth-doped barium-aluminum-borate glasses,” Opt. Express 13(5), 1635–1642 (2005). [CrossRef]   [PubMed]  

5. 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). [CrossRef]   [PubMed]  

6. A. N. Romanov, E. V. Haula, Z. T. Fattakhova, A. A. Veber, V. B. Tsvetkov, D. M. Zhigunov, V. N. Korchak, and V. B. Sulimov, “Near-IR luminescence from subvalent bismuth species in fluoride glass,” Opt. Mater. 34(1), 155–158 (2011). [CrossRef]  

7. A. G. Okhrimchuk, L. N. Butvina, E. M. Dianov, N. V. Lichkova, V. N. Zagorodnev, and K. N. Boldyrev, “Near-infrared luminescence of RbPb2Cl5:Bi crystals,” Opt. Lett. 33(19), 2182–2184 (2008). [CrossRef]   [PubMed]  

8. 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]  

9. J. Ruan, L. Su, J. Qiu, D. Chen, and J. Xu, “Bi-doped BaF2 crystal for broadband near-infrared light source,” Opt. Express 17(7), 5163–5169 (2009). [CrossRef]   [PubMed]  

10. M. Peng, B. Sprenger, M. A. Schmidt, H. G. L. 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). [CrossRef]   [PubMed]  

11. 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]  

12. H.-T. Sun, A. Hosokawa, Y. Miwa, F. Shimaoka, M. Fujii, M. Mizuhata, S. Hayashi, and S. Deki, “Strong ultra-broadband near-infrared photoluminescence from bismuth-embedded zeolites and their derivatives,” Adv. Mater. (Deerfield Beach Fla.) 21(36), 3694–3698 (2009). [CrossRef]  

13. H.-T. Sun, Y. Sakka, Y. Miwa, N. Shirahata, M. Fujii, and H. Gao, “Spectroscopic characterization of bismuth embedded Y zeolites,” Appl. Phys. Lett. 97(13), 131908 (2010). [CrossRef]  

14. H.-T. Sun, M. Fujii, Y. Sakka, Z. Bai, N. Shirahata, L. Zhang, Y. Miwa, and H. Gao, “Near-infrared photoluminescence and Raman characterization of bismuth-embedded sodalite nanocrystals,” Opt. Lett. 35(11), 1743–1745 (2010). [CrossRef]   [PubMed]  

15. W. A. Runciman, “Absorption and emission spectra of bismuth-activated phosphors,” Proc. Phys. Soc. A 68(7), 647–649 (1955). [CrossRef]  

16. A. N. Romanov, Z. T. Fattakhova, D. M. Zhigunov, V. N. Korchak, and V. B. Sulimov, “On the origin of near-IR luminescence in Bi-doped materials (I). Generation of low-valence bismuth species by Bi3+ and Bi0 synproportionation,” Opt. Mater. 33(4), 631–634 (2011). [CrossRef]  

17. H.-T. Sun, Y. Sakka, M. Fujii, N. Shirahata, and H. Gao, “Ultrabroad near-infrared photoluminescence from ionic liquids containing subvalent bismuth,” Opt. Lett. 36(2), 100–102 (2011). [CrossRef]   [PubMed]  

18. M. A. Hamstra, H. F. Folkerts, and G. Blasse, “Materials chemistry communications. Red bismuth emission in alkaline-earth-metal sulfates,” J. Mater. Chem. 4(8), 1349–1350 (1994). [CrossRef]  

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

20. R. M. Friedman and J. D. Corbett, “Synthesis and structural characterization of bismuth(1+)nonabismuth(5+)hexachlorohafnate(IV), BiBi9(HfCl6)3,” Inorg. Chem. 12(5), 1134–1139 (1973). [CrossRef]  

21. S. Ulvenlund, L. Bengtsson-Kloo, and K. Ståhl, “Formation of subvalent bismuth cations in molten gallium trichloride and benzene solution,” J. Chem. Soc., Faraday Trans. 91, 4223–4234 (1995). [CrossRef]  

22. H. Kalpen, W. Hönle, M. Somer, U. Schwarz, K. Peters, H. G. von Schnering, and R. Blachnik, “Bismut(II)-chalkogenometallate(III) Bi2M4X8, Verbindungen mit Bi24+-Hanteln (M=Al, Ga; X=S,Se),” Z. Anorg. Allg. Chem. 624(7), 1137–1147 (1998). [CrossRef]  

23. E. V. Dikarev and B. Li, “Rational syntheses, structure, and properties of the first bismuth(II) carboxylate,” Inorg. Chem. 43(11), 3461–3466 (2004). [CrossRef]   [PubMed]  

24. B. Wahl and M. Ruck, “Ag3Bi14Br21: ein Subbromid mit Bi24+-Hanteln und Bi95+-Polyedern – Synthese, Kristallstruktur und chemische Bindung,” Z. Anorg. Allg. Chem. 634(15), 2873–2879 (2008). [CrossRef]  

25. J. D. Corbett, F. C. Albers, and R. A. Sallach, “An electromotive force studies of solutions of bismuth in bismuth (III) chloride at 240°C,” Inorg. Chim. Acta 2, 22–26 (1968). [CrossRef]  

26. B. Krebs, M. Mummert, and C. J. Brendel, “Characterization of the Bi53+ cluster cation: preparation of single crystals, crystal and molecular structure of Bi5(AlCl4)3,” J. Less Common Met. 116(1), 159–168 (1986). [CrossRef]  

27. M. Ruck, “Bi34Ir3Br37: Ein pseudosymmetrisches Subbromid aus Bi5+ und Bi62+ Polykationen sowie [IrBi6Br12] und [IrBi6Br13]2– - Clusteranionen,” Z. Anorg. Allg. Chem. 624(3), 521–528 (1998). [CrossRef]  

28. M. Ruck and S. Hampel, “Stabilization of homonuclear Bi5+ and Bi62+ polycations by cluster anions in the crystal structures of Bi12−xIrCl13−x, Bi12−xRhCl13−x and Bi12−xRhBr13−x,” Polyhedron 21(5-6), 651–656 (2002). [CrossRef]  

29. N. J. Bjerrum and G. P. Smith, “Lower oxidation states of bismuth. Bi82+ formed in aluminum chloride-sodium chloride melts,” Inorg. Chem. 6(11), 1968–1972 (1967). [CrossRef]  

30. J. Beck, C. J. Brendel, L. A. Bengtsson-Kloo, B. Krebs, M. Mummert, A. Stankowski, and S. Ulvenlund, “The crystal structure of Bi8(AlCl4)2 and the crystal structure, conductivity and theoretical band structure of Bi6Cl7 and related subvalent bismuth halides,” Chem. Ber. 129(10), 1219–1226 (1996). [CrossRef]  

31. A. Hershaft and J. D. Corbett, “The crystal structure of bismuth subchloride. Identification of the ion Bi95+,” Inorg. Chem. 2(5), 979–985 (1963). [CrossRef]  

32. H.-T. Sun, Y. Sakka, H. Gao, Y. Miwa, M. Fujii, N. Shirahata, Z. Bai, and J.-G. Li, “Ultrabroad near-infrared photoluminescence from Bi5(AlCl4)3 crystal,” J. Mater. Chem. 21(12), 4060–4063 (2011). [CrossRef]  

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

34. S. Pedersen, “Viscosity, structure and glass formation in the AlCl3-ZnCl2 system,” Ph.D thesis (Institutt for Kjemi, Norges Tekniskurn-Naturvitenskaplige Universitet, 2001).

35. C. R. Boston and G. P. Smith, “Spectra of dilute solutions of bismuth metal in molten bismuth trihalides. I. Evidence for two solute species in the system bismuth-bismuth trichloride,” J. Phys. Chem. 66(6), 1178–1181 (1962). [CrossRef]  

36. C. R. Boston, G. P. Smith, and L. C. Howick, “Spectra of dilute solutions of bismuth metal in molten bismuth trihalides. II. Formulation of solute equilibrium in bismuth trichloride,” J. Phys. Chem. 67(9), 1849–1852 (1963). [CrossRef]  

37. L. E. Topol, S. J. Yosim, and R. A. Osteryoung, “E.M.F. measurements in molten bismuth-bismuth trichloride solutions,” J. Phys. Chem. 65(9), 1511–1516 (1961). [CrossRef]  

38. H. L. Davis, N. J. Bjerrum, and G. P. Smith, “Ligand field theory of p2,4 configurations and its application to the spectrum of Bi+ in molten salt media,” Inorg. Chem. 6(6), 1172–1178 (1967). [CrossRef]  

39. B. I. Denker, B. I. Galagan, V. V. Osiko, I. L. Shulman, S. E. Sverchkov, and E. M. Dianov, “Factors affecting the formation of near infrared-emitting optical centers in Bi-doped glasses,” Appl. Phys. B 98(2-3), 455–458 (2010). [CrossRef]  

40. N. A. Alexeev, V. P. Gapontsev, M. E. Zhabotinskii, V. B. Kravchenko, and Yu. P. Rudnitskii, Laser Phosphate Glasses (Nauka, Moscow, 1980), Chap. 3.

41. A. N. Romanov, O. A. Kondakova, D. N. Vtyurina, A. V. Sulimov, and V. B. Sulimov, “Calculation of excited states properties for Bi53+ polycation by the spin-orbit configuration interaction method,” Num. Meth. Prog. 12, 443–449 (2011).

42. H. Kunkely and A. Vogler, “On the origin of the photoluminescence of mercurous chloride,” Chem. Phys. Lett. 240(1-3), 31–34 (1995). [CrossRef]  

43. X. Guo, H. Li, L. Su, P. Yu, H. Zhao, Q. Wang, J. Liu, and J. Xu, “Study on multiple near-infrared luminescent centers and effects of aluminum ions in Bi2O3–GeO2 glass system,” Opt. Mater. 34(4), 675–678 (2012). [CrossRef]  

44. S. V. Firstov, V. F. Khopin, I. A. Bufetov, E. G. Firstova, A. N. Guryanov, and E. M. Dianov, “Combined excitation-emission spectroscopy of bismuth active centers in optical fibers,” Opt. Express 19(20), 19551–19561 (2011). [CrossRef]   [PubMed]  

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 (4)

Fig. 1
Fig. 1 Sealed L-shaped fused silica cell with thin layer of AlCl3/ZnCl2/BiCl3 glass. The glass orange color is due to the presence of subvalent bismuth species (mainly Bi53+ polycation).
Fig. 2
Fig. 2 (a) Absorption spectrum of AlCl3/ZnCl2/BiCl3 glass, containing subvalent bismuth species; (b) Photoluminescence spectra of AlCl3/ZnCl2/BiCl3 glass (300K), excited at different wavelengths normalized with respect to the 1320 nm emission.
Fig. 3
Fig. 3 (a) Photoluminescence spectra of AlCl3/ZnCl2/BiCl3 glass, measured at 77 and 300K (excitation wavelength λex = 532 nm). The spectrum at 77K composed from two separate measurements: at short waves (with InGaAs detector), and at long waves (with InSb detector); (b) Photoluminescence excitation spectra, measured for three luminescence bands. Spectra of the first and the second bands (λem = 1060 nm and λem = 1300nm, respectively) were measured at 300K. Excitation spectrum of the third band (λem = 1900 nm) was measured at 77K.
Fig. 4
Fig. 4 Low temperature (77K) luminescence decay curves measured at 1350nm(a), 1900nm(b) and 1050nm(c).

Equations (4)

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

Bi 5 3 + + Bi 3 + 6Bi +
BiCl 3 + Zn Bi + + ZnCl 2 + Cl
5BiCl 3 + 6Zn Bi 5 3 + + 6ZnCl 2 + 3Cl
Bi + + Bi 3 + Bi 2 4 +
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.