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

Superbroad near to mid infrared luminescence from closo-deltahedral Bi53+ cluster in Bi5(GaCl4)3

Open Access Open Access

Abstract

Closo-deltahedral Bi53+ cluster in Bi5(GaCl4)3, which can be synthesized in benzene by oxidizing bismuth metal either with BiCl3 or GaCl3, respectively, can absorb ultraviolet, visible and infrared lights, and luminesce superbroadly in near to mid infrared (NMIR) spectral range from 1 to 3μm at room temperature. Slight geometry change of the cluster can lead to the redshift of emission peak. These observations may initialize the development of Bi-based NMIR light sources with superbroad emission spectrum, where Bi53+ or similar polycationic species act as activators. Disputable crystal structure of Bi5(GaCl4)3 was redefined by classic Rietveld refining analysis. Consistent with crystallographic data, excitation, emission, temporal decay and time-resolved infrared emission spectra all reveal only one type of luminescent centers, viz. Bi53+, in the compound. And a new absorption of Bi53+ was found at ~1100nm.

©2012 Optical Society of America

1. Introduction

The wonder metal bismuth exhibits wonder spectroscopy properties particularly when introduced into glasses and crystals [130]. For instance, when bismuth is controlled to “+2,” it can absorb ultraviolet and blue lights and emit intensive red lights. The Bi2+ doped crystals can then be used to improve color rendering index of commercial white light LEDs [811]. As bismuth goes lower than “+2,” it can emit lights broadly distributed in the spectral range from 1 to 1.6μm at the same time with considerable absorptions in visible and near infrared range [17,1226]. This endows the doped glasses a unique ability to provide high optical gain in the bandgap where traditional rare earth based devices cannot reach [27,27,28]. The intriguing spectral properties of such materials tempt scientists in materials and optics communities to perform intense researches on this topic. It is because of this tremendous input throughout the world that rapid progress has been enabled and witnessed within only two years from first lasering to efficient all fiber optical amplifiers and lasers [27,28]. Now, bismuth doped broadband amplifiers have been recognized as next generation devices after erbium doped fiber amplifier. However, despite of sequence success in fabrication of device and exploration of various potential laser materials, the nature of the optically active NIR emission centers remains unclear [1725].

This seemingly elemental problem is complicated partly by the diversity of valence states of monoatomic bismuth ion and the lapsable transformation between these states, and also by the strong proneness of bismuth to oligomerize or copolymerize into unusual homo- or hetero-nuclear clusters. The fascinating cluster species are often with novel unexpected chemical and optical behaviors, bonding and reactivity. For instance, neutral bismuth clusters can emit NIR light [14]. Strikingly, Bi53+ was found to even luminesce in superbroadly spectral range from about 1 to 4μm [29]. None of any rare earth and transition metal doped optical materials could be on a par with it (see below, Fig. 2(b) ). Nevertheless, yet to date, the studies carried out on the fluorescent properties of bismuth polycation groups have been very limited especially in the range of NMIR [2940].

 figure: Fig. 2

Fig. 2 (a) Raman spectrum of Bi5(GaCl4)3-A-36h. Vibration modes of Bi53+ are labeled in the figure. Inset is the polycation group of Bi53+; (b) Emission (curves 1-3) and excitation (curves 4-5) spectra of Bi5(GaCl4)3-A-36h at room temperature. Curve 1 is upon 470nm excitation, curve 2 is the emission spectrum calibrated to the detector spectral response, and both of them were recorded with liquid N2 cooled NIR photomultiplier (Hamamatsu R5509-72). Curve 3 was collected with a PbSe detector under the excitation of 808nm laser diode. Curves 4 and 5 were measured by monitoring the emissions at 1400 and 1500nm, respectively. Curve 4 is shifted vertically for clarity. For reference, curve 6 an emission spectrum of Er3+ doped fluorophosphate glass is listed upon 808nm excitation and the FWHM is much smaller than Bi5(GaCl4)3.

Download Full Size | PDF

Basic motivation for this work is to study the optical properties of bismuth polycations. It will in one hand help better understanding the origin of NIR fluorescence of bismuth doped materials; on the other hand, it will possibly prompt the finding of optical materials in new spectral regime for new laser sources.

When measuring Raman spectroscopy of Bi53+ in (1 – n – butyl – 3 – methylimidazonlium) Cl/AlCl3 ionic liquid ([BMIM]Cl/AlCl3) with 1064 nm laser excitation, Ruck et al observed strong luminescence disturbance over the spectral range of 1075-1852 nm [32]. This hints possible excited electronic states of Bi53+ lying around 1064nm. However, none of absorptions have been reported beyond 1000nm [2939]. Sun et al followed Ruck's approach and synthesized Bi5(AlCl4)3 where bismuth exists in an entity of Bi53+, and reported a broad fluorescence band spanning the range of 900-1600 nm with a FWHM >500 nm [14,15]. This agrees well with the finding by Ruck et al [32]. Very recently, we synthesized Bi5(AlCl4)3 via a molten salt route with NaAlCl4 as solvent and excess Lewis strong acid AlCl3 as stabilizer, and found it fluoresce in NIR-MIR at room temperature [29]. In the final products, however, Bi5(AlCl4)3 is miscellaneous with NaAlCl4 phase, and it seems physically impractical to separate them from each other within this framework. Subsequently Romanov et al found similar NMIR fluorescence of Bi53+ in chloride glass at 77K, and it is thermal quenched at room temperature [40]. In the glass Bi53+ coexists with Bi+ and Bi24+, which probably is due to the violent and indiscriminate reactivity of acidic molten salts as reaction medium [40].

Selection of Bi5(GaCl4)3 is prompted by these concerns:

  • (1) Inspired by the works of Ulvenlund et al and Lindsjö et al [35,37,41,42], the compound can be prepared in an organic liquid medium at room temperature; this offers an opportunity of conducting reduction and oxidation of bismuth under a rather mild condition in a more controlled way;
  • (2) Pure phase or even single crystal of the compound could be separated conveniently by remove of the solvent in regular procedure;
  • (3) Discrepancy of crystal structure of the compound was reported; for instance, Ulvenlund et al found the compound belonging to trigonal space group R-3c [35,37], while Lindsjö et al reported it crystallizing in the space group of R3c; and Lindsjö et al found two apical bismuth atoms of Bi53+ were not equivalent and this distorted Bi53+ from ideal D3h symmetry [41];
  • (4) Despite of the dispute, a common sense is that there is only one type of homonuclear deltahedral entity of Bi53+ in the compound as shown in Fig. 1(a) ; it belongs to closo-polyhedra according to Wade’s rules, where skeletal electron count equals to 4n + 2 (n is the number of atoms in the electron deficient polyhedron Mnx); this scenario is quite different from bismuth doped chloride glasses [40] and it can no doubt simplify the problem risen by the simultaneous existence of multiple centers.
     figure: Fig. 1

    Fig. 1 (a) Unit cell representation of Bi5(GaCl4)3 on the basis of Rietveld refining results; (b) XRD pattern (black circle) of Bi5(GaCl4)3-A-36h, corresponding Rietveld refining results (solid red line, Rp = 7.47%, Rwp = 14.6%, Rexp = 11.0%, GOF = 1.76, RB = 3.95%), Bragg positions (vertical green line) and difference profile (oliver line) between observed and calculated values. Inset is the sample before separation.

    Download Full Size | PDF

  • (5) No meaningful Raman spectrum of Bi5(GaCl4)3 has been recorded in the case of exciting radiation of 1064nm YAG: Nd laser [37].

To elucidate these queries, in this work, we will first synthesize the pure phase of Bi5(GaCl4)3 with benzene as solvent and mild oxidizer such as BiCl3 or GaCl3 for bismuth metal at room temperature, determine the crystal structure of Bi5(GaCl4)3 with the classic Rietveld refining process, analyze the spectroscopic properties of Bi53+ in Bi5(GaCl4)3 by collecting Raman spectrum, excitation spectra, and site-selectively excited emission spectra, time resolved NIR emission spectra as well as absorption spectrum beyond 1000nm, and study the thermal resistance to laser irradiation in the case of uncooling. The compound exhibits extraordinarily broad luminescence in near to mid infrared spectral region even at room temperature.

2. Experimental section

2.1 Materials and sample synthesis

Anhydrous GaCl3 (Alfa, 99.999%), BiCl3 (Aladin, 99.9%), bismuth metal powder (Aladin, 99.999%) and benzene (Sigma-Aldrich, AR) were used as received. All the operations were carried out in a glovebox under dry nitrogen. When molar ratio of GaCl3 to benzene X(GaCl3, benz)<0.3 and the solution is saturated with bismuth, it tends to form into two liquid phases. When X(GaCl3, benz) = 0.3 and molar ratio of bismuth to GaCl3 X(Bi, GaCl3) is 0.2, solution will produce a homogenous liquid phase [42]. In view of this, we kept X(GaCl3, benz) = 0.3 constant throughout the study. Though BiCl3 or GaCl3 has been suggested as suitable oxidizing species to prepare Bi5(GaCl4)3 using benzene as reaction medium, none of reaction rate has been evaluated before when the two oxidizers are used respectively [37,42]. Partially for this, two parallel experiments were designed according to the works by Ulvenlund et al [37,42]. Experiment (A) with BiCl3 as oxidizer selected the molar composition of 4.00Bi:1.00BiCl3: 8.00GaCl3:26.67 Benzene. 178.3mg GaCl3 was first dissolved in 0.3ml benzene, 105.8mg Bi was then added right after 39.9mg BiCl3. Reaction time was counted from zero to 96 hours. The mixture was rocked to improve the homogeneity. Experiment (B) with GaCl3 as oxidizer employed 1.00Bi: 5.00GaCl3: 16.67Benzene. 178.3mg GaCl3 and 42.3mg Bi were consequently dissolved in 0.3ml benzene, and further admixed by rocking. Reaction time changed from 0 to 120 hours. The reactions were all confined in evacuated and flame sealed Pyrex tubes. Both experiments were at room temperature. For convenience, samples were coded as Bi5(GaCl4)3-A(or B)-reaction time. For instance, Bi5(GaCl4)3-A-36h means that the sample was synthesized by Experiment (A) with reaction time of 36h. The solutions were deep reddish brown before separation as inset shown in Fig. 1(b).

2.2 Measurements and characterization

Because the solutions are too deep, any structures will be smoothed by the strong absorption. However, when the glass tubes are shaken, some reddish crystals will adhere on the inner side of the tube wall. By letting light through the wall thinner covered with the crystals, the absorption spectrum was measured with a Perkin Elmer Lambda-900 UV-VIS-NIR spectrometer. Before other measurements, the solution was centrifuged, and red powders were separated and washed with cyclohexane, and part of them was transferred and sealed in the same type of tube. The residual solution shows weak luminescence when compared to the powder. Static and dynamic excitation and emission spectra as well as emission lifetime were recorded with a high-resolution spectrofluorometer Edingburgh FLS 920 with a liquid nitrogen cooled NIR photomultiplier (Hamamatsu R5509-72). Due to limitation of the detector in the spectrometer, only time resolved NIR emission spectra could be achieved in the range from 900 to 1650nm. Photoluminescence between 1000 and 3500 nm was observed with a PbSe detector on a Horiba Jobin Yvon Triax 320 fluorometer, using an 808 nm laser diode as pump source. X-ray diffraction patterns of the samples were recorded with a Rigaku D/max-IIIA X-ray diffractometer (40kV, 1.2°min−1, 40 mA, Cu-Kα1, λ = 1.5405Å). Raman spectra were measured on a Horiba Jobin Yvon LabRAM Aramis spectrometer equipped with a 633 nm He-Ne laser. Signal was collected vertical to laser beam.

3. Results and discussion

3.1 Structural identification

When refining the crystal structure of Bi5(GaCl4)3 in R-3c, Lindsjö et al found a poorer result with significantly larger values of R factors [41]. The situation could be improved as the structure was evaluated with R3c space group. And goodness of fit (GOF) could be, thus, produced as 1.215. This led to the obvious change of cation geometry of Bi53+. The two apical bismuth atoms Bi(2) and Bi(2)’ of the entity, which are crystallographically equivalent in R-3c, became quite different. This yielded two different bond lengths 2.97 and 3.02Å between the equatorial atoms and the apical atoms. However, they could not find any experimental proof to confirm the symmetry. Ulvenlund et al [37] believed that it crystallized in space group R-3c and satisfactory residual values (Rp = 3.9%; Rwp = 5.2%; RB = 2.3% and GOF = 3.0) could be obtained when the strongest reflection (012) around 2θ = 10.47° was excluded during the refining.

To settle the divergence we prepared the samples with experiments A and B. The structures of the samples are rather analogous and, thus, only XRD pattern of sample Bi5(GaCl4)3-A-36h is shown exemplarily in Fig. 1(b). For the structural refining (with FullProf Suite), angular range is restricted to 9.8≤2θ≤42 and it includes the reflection (012). The start of the refining is based on the crystallographic data of Bi5(AlCl4)3. And stop condition of convergence is set as all parameter shifts are lower than 0.1 of the corresponding estimated standard deviation. The refinement converges with residual values of Rp = 7.47%, Rwp = 14.6%, Rexp = 11.0%, GOF = 1.76 and RB = 3.95%. The refining results are listed in Fig. 1(b) also with the difference profile between observed and calculated values. This comparison clearly suggests that Bi5(GaCl4)3 belongs to the same space group R-3c as Bi5(AlCl4)3, which consists with Ulvenlund et al. The analysis produces structural parameters shown in Table 1 , based on which Fig. 1(a) the unit cell of the compound is drawn, and bond distances and angles are calculated and listed in Table 2 along with those of Bi5(AlCl4)3.

Tables Icon

Table 1. Wyckoff positions (wp) and coordinates for Bi5(GaCl4)3

Tables Icon

Table 2. A comparison of unit cell parameters (Å), distances (Å) and angles (deg) between Bi5(GaCl4)3 and Bi5(AlCl4)3

Isostructural with Bi5(AlCl4)3, Bi5(GaCl4)3 comprises two types of bismuth atoms Bi(1) (Wyckoff position 18e) and Bi(2) (12c), one type of Ga atoms and two types of chlorine atoms Cl(1) and Cl(2) (see Table 1 and Fig. 1(a)). Occupation ratio is 3:2 for Bi(1) and Bi(2). These bismuth atoms form into a single type of Bi53+ polyhedron, as Ulvenlund et al had noted before composed of three equatorial Bi(1) and two equivalent apices Bi(2). Two Cl(1) and two Cl(2) ions coordinate one Ga ion and form into the tetrahedron group of GaCl4-. In each unit cell, there are on average six Bi53+ and eighteen GaCl4- polyhedra. The Bi53+ cluster preserves an ideal symmetry of D3h for which geometry the representation of Raman and Infrared active normal vibrations can be Γvib(D3h) = 2A1’ + A2” + 2E’ + E”. Among the modes, A1’, E’ and E” are Raman active [38]. A 633nm He-Ne laser rather than former 1064nm was selected as pump light for collecting Raman since the compound has less absorption at the wavelength as will be shown in the excitation spectrum. A good spectrum of Raman was therefore achieved and depicted as Fig. 2(a). Typical vibrations of trigonal bipyramid Bi53+ indeed appear at 54, 62, 99, 120, 124, 138, 151cm−1, which can be attributed to ν5(E’), ν4(E’), ν6(E”), ν2(A1’), ν2(A1’), ν1(A1’) and ν1(A1’) of Bi53+ [29,3235,38], respectively. This presents clear evidence for the symmetry of Bi53+ polycation. Rest Raman peaks at 87, 111 and 166cm−1 are due to lattice modes and ν2(E) of GaCl4- [32,38].

For Bi-Bi and Bi-Cl bonds, the distances are slightly longer in Bi5(GaCl4)3 as compared to Bi5(AlCl4)3. The distance between two bismuth apexes is elongated to 4.696(2)Å in the direction of c-axis. This elongation expands the angles Bi(2)-Bi(1)-Bi(2)’ and Bi(1)-Bi(1)’-Bi(2), and simultaneously suppresses Bi(1)-Bi(2)-Bi(1)’ (see Table 2). Cell parameters of Bi5(GaCl4)3 are refined as a = 11.886(0) Å and c = 30.136(0) Å, slightly larger than those of Bi5(AlCl4)3 which are a = 11.86 Å and c = 30.10 Å reported by Krebs et al [39], a = 11.8712 Å and c = 30.1203 Å by Ruck et al [32], and a = 11.8698 Å and c = 30.1133 Å by Cao et al [29]. These, we think, are due to the substitution of Ga3+ (Ionic radius is 0.47 Å for four coordination.) for Al3+ (Also for four coordinated ions, radius is 0.39 Å.) ions.

When Ulvenlund et al measured liquid X-ray scattering spectrum of a solution containing Bi5(GaCl4)3, they found [42] that the reduced radial distribution function of the sample reveals the structure of the building blocks Bi53+ and GaCl4-. Peaks at 2.24, 3.13, 3.58 and 4.70 Å can be found in Fig. 7 of [42] and they are due to the average distances of Ga-Cl, Bi-Bi (axial-equatorial and equatorial-equatorial) and Bi-Cl, and the distance of Bi(2)-Bi(2)’. They agree with our results (2.242, 3.142, 3.584, 4.696 Å) better than those (2.249, 3.170, 3.479, 4.649 Å) by Ulvenlund et al [37]. This probably is because our refining is more reasonable since it does not abandon the important reflection peaks such as the strongest (012).

3.2 Near to mid infrared (NMIR) luminescence from Bi5(GaCl4)3

When exposed to the light of 470 nm, broad near infrared luminescence was observed in the range of 900 to 1650nm from all examined samples. The emission spectrum of representative sample Bi5(GaCl4)3-A-36h is depicted as curve 1 of Fig. 2(b), and it is similar to Bi5(AlCl4)3 as reported in references [29,33,34]. The full width at half maximum (FWHM) is 555nm, comparable to 570nm and 510nm of Bi5(AlCl4)3 reported by Cao et al and Sun et al [29,33,34]. The sudden falloff around 1380nm and in 1500-1650nm is due to the spectral response drop of the NIR photomultiplier in the areas. After correction of curve 1 over the detector sensitivity curve 2 in Fig. 2(b) is resulted and the tail of the spectrum rises steeply in the wavelengths longer than 1380nm. This indicates a possible luminescence peak lying beyond 1650nm. To confirm it a broader band PbSe detector was summoned. Subsequent measurement testifies the presence of the band as clearly shown in curve 3 of Fig. 2(b). It peaks at ~1835nm with FWHM of 800nm and persists to ~3500nm. The emission peak of Bi5(GaCl4)3 is located at longer wavelength than ~1700nm of Bi5(AlCl4)3 [29]. This probably correlates with the slight change of geometry of Bi53+ in Bi5(GaCl4)3. As discussed above, the entity is elongated in the c-axis when compared to Bi5(AlCl4)3.

Curve 1 in Fig. 3(a) is the absorption spectrum of the representative sample Bi5(GaCl4)3-A-36h. Absorptions can be found at 430, 580, 854, 1137(sharp), respectively. An additional peak around ~1100nm can be traced though it is superimposed by the sharp peak at 1137nm. To assign these accurately, absorption spectrum of benzene (curve 2 of Fig. 3(a)) is listed as references. Peak at 1137 is from second overtone stretching of C-H in benzene (see curves 1 and 2 of Fig. 3(a)). Intense absorption blurs fine structures in the wavelengths shorter than 400nm, the excitation spectra, however, can differentiate these. Thus, the excitation spectra of the compound of Bi5(GaCl4)3 were recorded upon infrared emissions and illustrated as curves 4 and 5 in Fig. 2(b). The excitation spectrum shows no clear dependence on emission wavelength varying in the range of 1000 to 1600nm. This is possible due to the existence of only one type of Bi53+ emission centers in the compound as aforementioned (see Fig. 1(a)). Curves 4 and 5 are similar, and they are visually composed of two strong peaks at 473 and 766nm and two shoulders at 362 and 430nm. The peaks at 473 and 766nm are both broad and they contain the 854nm peak in curve 1 of Fig. 3(a) and other unresolved peaks which have been observed in the compound of Bi5(AlCl4)3. Due to instrumentation limit, excitation spectrum cannot be achieved in the range of wavelengths longer than 900nm. In all, by combining absorption and excitation spectra, it can be concluded that absorptions at 362, 430, 473, 580, 766, 854 and ~1100nm are from Bi53+. On the basis of the calculation by Corbett et al [36] on the entity of Bi53+, the former six can be assigned to e’→e’(3)(E’), e’(1)→a2’(E’), e’(1)→a1’(E’), a1’→e’(2)(E’), e’(1) →e’(2)(E’) and e”→e’(2)(E’), respectively. The last one was observed for the first time and, however, it cannot be assigned properly at the moment since the complete energy level diagram of Bi53+ is absent for beyond 900nm [36].

 figure: Fig. 3

Fig. 3 (a) Absorption spectra of (1) Bi5(GaCl4)3-A-36h and (2) benzene, I0 and I are the incident and transmitted radiation through the sample; (b) Decay curve of the emission at 1365nm from Bi5(GaCl4)3-A-36h excited by 470nm at room temperature. Red circles are fitted results by an equation I = 442.596exp(-t/8.67) with χ2 = 1.086.

Download Full Size | PDF

Excitation into either of these absorption bands does not lead to the shape change of the emission spectrum. This proves in the other way that they all belong to the same emission center. The situation is similar to Bi5(AlCl4)3 [29]. Figure 3(b) illustrates exemplarily the temporal decay of the emission at 1365nm upon 470nm excitation. The curve fits well the single exponential decay equation with χ2 = 1.086, and the fitting produces lifetime of 8.67μs, comparable to 5.96μs of Bi5(AlCl4)3 at room temperature [29,33,34]. Fig. 4(a) depicts the time resolved emission spectra of Bi5(GaCl4)3-A-36h excited by 470nm at room temperature. With increasing delay time, intensities of the emission monotonically decrease while the shape of the spectra does not change. This again evidences one type of infrared emission centers in the compound.

 figure: Fig. 4

Fig. 4 (a) Time resolved emission spectra of Bi5(GaCl4)3-A-36h excited by 470nm at room temperature. Delay times are marked beside the curves; (b) Dependence of the intensity of infrared emission@1835nm of Bi5(GaCl4)3-A-36h on pump power of 808nm laser diode.

Download Full Size | PDF

For potential practical applications such as laser gain medium, excited state absorption and appropriate pumping schemes are essential issues to be considered. Here, for instance, we examined the 808nm pump plan, since 808 nm diode laser is low cost and more importantly it is commercially accessible. The dependence of infrared luminescence on the pump power can to some extent reflect the information on the issues. When the pump power increases from 400mW to 1100mW, the emission intensity increases linearly (see Fig. 4(b)), so in this regime, no severe excited state absorption occurs. Nevertheless, when the laser power rises higher than 1100mW, emission intensity starts to fall slightly, what we think perhaps is due to thermal dissociation of the compound of Bi5(GaCl4)3, since its melting points is 321°C, five degree lower than the structural analogue Bi5(AlCl4)3 [37]. Similar phenomenon appeared in Bi5(AlCl4)3 [29]. In view of this, initializing laser operation with the crystal needs efficient cooling especially when the excitation power exceeds the W-regime.

3.3 Effect of oxidizing species on reaction rate of formation of Bi5(GaCl4)3

Ulvenlund et al found with GaCl3 as the oxidant that introduction of bismuth metal to a solution of GaCl3 and benzene could induce immediately a strongly colored liquid [35]. Rather, here, we observed different results. Since Bi5(GaCl4)3 can luminesce we can then use the luminescence signal as a formation indicator of the crystal. After reaction for 60 hours the solution becomes reddish brown and red precipitates appear at the bottom of the tube and typical infrared luminescence of Bi5(GaCl4)3 can be detected as shown in Fig. 5(a) . As the reaction continues, the luminescence becomes stronger. As the reaction time prolongs from 96 to 120 hours, the increase of the luminescence is fairly slight. Contrarily, with BiCl3 as the oxidizing species for bismuth metal, we found the crystal starts to come into being after reacting for 4hours as depicted in Fig. 5. Comparing the two experiments with different oxidizing reagents, we can find out that the rate of the reaction using BiCl3 oxidant is much faster than that of GaCl3 especially at the very beginning period.

 figure: Fig. 5

Fig. 5 (a) Dependence of fluorescence intensity of Bi5(GaCl4)3-A (-●-) and Bi5(GaCl4)3-B (-●-) on reaction time; (b) Evolution of emission spectrum of Bi5(GaCl4)3-A with reaction time. Curve 1: 2h; curve 2: 4h; curve 3: 8h; curve 4: 24h; curve 5: 36h; curve 6: 72h; curve 7: 96h.

Download Full Size | PDF

4. Conclusion

Clearly, the compound of Bi5(GaCl4)3 can be synthesized by redox reactions with BiCl3 or GaCl3 as the mild oxidizing species for bismuth metal in benzene. The rate of the former reaction is much speedier than the latter in the beginning phase. Addition of bismuth metal into GaCl3 + benzene solution did not initialize instant reaction, which is different from what Ulvenlund et al found. Rietveld refining process has identified that Bi5(GaCl4)3 belongs to trigonal space group R-3c with a larger lattice cell than structural analog Bi5(AlCl4)3. This is due to the substitution of Ga3+ for Al3+, and it can elongate the spacing between two bismuth apexes of Bi53+. The change of geometry of the entity doesn’t make the two apical bismuth atoms unequivalent as reported by Kloo et al; it also doesn’t destroy the ideal symmetry of Bi53+, that is, D3h as further verified by Raman spectrum of the compound. But it results in the redshift of emission peak as compared to Bi5(AlCl4)3. The closo-detahedra of Bi53+ exhibit extraordinary NMIR luminescence peaking at ~1835nm with the FWHM of 800nm and the lifetime of 8.67μs at room temperature. Consistent with crystallographic data, excitation, emission, dynamic decay and time resolved infrared emission spectra all prove only one type of emission centers in the crystal, that is Bi53+, which is intrinsically differing from those in bismuth doped glasses, as discussed in [29,40]. Except typical absorptions of Bi53+ reported previously, a new absorption was found at ~1100nm. Before thermally decomposing, Bi53+ doesn’t appear severe excited state absorption. This work opens a new direction to manipulate the fluorescence of polycationic species such as Bi53+ by controlling geometry of the clusters. Guided by it and choosing similar entities to Bi53+ as activators, we can in the future develop new glass systems, where they can be stabilized or prevail over other bismuth species. And this will become the solid foundation to develop practical devices working in the desired spectral range of NMIR, the frequencies where traditional rare earth based schemes cannot fulfill.

Acknowledgments

This work is financially supported by the National Natural Science Foundation of China (Grants no. 51132004 and 51072060), Fundamental Research Funds for the Central Universities (Grants no. 2011ZZ0001 and 2011ZP0002), Guangdong Natural Science Foundation (Grant no. S2011030001349), Fok Ying Tong Education Foundation (Grant No. 132004), and the Chinese Program for New Century Excellent Talents in University (Grant no. NCET-11-0158).

References and links

1. J. Xu, H. Zhao, L. Su, J. Yu, P. Zhou, H. Tang, L. Zheng, and H. Li, “Study on the effect of heat-annealing and irradiation on spectroscopic properties of Bi:alpha-BaB2O4 single crystal,” Opt. Express 18(4), 3385–3391 (2010), http://www.opticsinfobase.org/abstract.cfm?URI=oe-18-4-3385. [CrossRef]   [PubMed]  

2. I. A. Bufetov, M. A. Melkumov, S. V. Firstov, A. V. Shubin, S. L. Semenov, V. V. Vel’miskin, A. E. Levchenko, E. G. Firstova, and E. M. Dianov, “Optical gain and laser generation in bismuth-doped silica fibers free of other dopants,” Opt. Lett. 36(2), 166–168 (2011). [CrossRef]   [PubMed]  

3. A. V. Kir’yanov, V. V. Dvoyrin, V. M. Mashinsky, N. N. Il’ichev, N. S. Kozlova, and E. M. Dianov, “Influence of electron irradiation on optical properties of Bismuth doped silica fibers,” Opt. Express 19(7), 6599–6608 (2011), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-19-7-6599. [CrossRef]   [PubMed]  

4. V. Dvoyrin, V. Mashinsky, and E. Dianov, “Efficient bismuth-doped fiber lasers,” IEEE J. Quantum Electron. 44(9), 834–840 (2008). [CrossRef]  

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), http://www.opticsinfobase.org/oe/abstract.cfm?uri=oe-17-22-19345. [CrossRef]   [PubMed]  

6. S. Zhou, H. Dong, H. Zeng, G. Feng, H. Yang, B. Zhu, and J. Qiu, “Broadband optical amplification in Bi-doped germanium silicate glass,” Appl. Phys. Lett. 91(6), 061919 (2007). [CrossRef]  

7. I. Razdobreev and L. Bigot, “On the multiplicity of bismuth active centres in germano-aluminosilicate preform,” Opt. Mater. 33(6), 973–977 (2011). [CrossRef]  

8. 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), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-17-23-21169. [CrossRef]   [PubMed]  

9. 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), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-18-12-12852. [CrossRef]   [PubMed]  

10. M. Peng and L. Wondraczek, “Bi2+-doped strontium borates for white-light-emitting diodes,” Opt. Lett. 34(19), 2885–2887 (2009). [CrossRef]   [PubMed]  

11. 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]  

12. M. Peng, N. Zhang, L. Wondraczek, J. Qiu, Z. Yang, and Q. Zhang, “Ultrabroad NIR luminescence and energy transfer in Bi and Er/Bi co-doped germanate glasses,” Opt. Express 19(21), 20799–20807 (2011), http://www.opticsinfobase.org/abstract.cfm?URI=oe-19-21-20799. [CrossRef]   [PubMed]  

13. 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]  

14. M. Peng, J. Qiu, D. Chen, X. Meng, and C. Zhu, “Superbroadband 1310 nm emission from bismuth and tantalum codoped germanium oxide glasses,” Opt. Lett. 30(18), 2433–2435 (2005). [CrossRef]   [PubMed]  

15. M. Peng, J. Qiu, D. Chen, X. Meng, and C. Zhu, “Broadband infrared luminescence from Li2O-Al2O3-ZnO-SiO2 glasses doped with Bi2O3.,” Opt. Express 13(18), 6892–6898 (2005), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-13-18-6892. [CrossRef]   [PubMed]  

16. M. Peng and L. Wondraczek, “Bismuth-doped oxide glasses as potential solar spectral converters and concentrators,” J. Mater. Chem. 19(5), 627–630 (2009). [CrossRef]  

17. 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]  

18. 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]  

19. 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]  

20. Y. Arai, T. Suzuki, Y. Ohishi, S. Morimoto, and S. Khonthon, “Ultrabroadband near-infrared emission from a colorless bismuth-doped glass,” Appl. Phys. Lett. 90(26), 261110 (2007). [CrossRef]  

21. T. Suzuki and Y. Ohishi, “Ultrabroadband near-infrared emission from Bi-doped Li2O-Al2O3-SiO2 glass,” Appl. Phys. Lett. 88(19), 191912 (2006). [CrossRef]  

22. Z. Yang, Z. Liu, Z. Song, D. Zhou, Z. Yin, K. Zhu, and J. Qiu, “Influence of optical basicity on broadband near infrared emission in bismuth doped aluminosilicate glasses,” J. Alloy. Comp. 509(24), 6816–6818 (2011). [CrossRef]  

23. J. Ren, L. Yang, J. Qiu, D. Chen, X. Jiang, and C. Zhu, “Effect of various alkaline-earth metal oxides on the broadband infrared luminescence from bismuth-doped silicate glasses,” Solid State Commun. 140(1), 38–41 (2006). [CrossRef]  

24. A. Romanov, Z. Fattakhova, D. Zhigunov, V. Korchak, and V. 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]  

25. M. Peng, B. Wu, N. Da, C. Wang, D. Chen, C. Zhu, and J. Qiu, “Bismuth-activated luminescent materials for broadband optical amplifier in WDM system,” J. Non-Cryst. Solids 354(12-13), 1221–1225 (2008). [CrossRef]  

26. H. T. Sun, F. Shimaoka, Y. Miwa, J. Ruan, M. Fujii, J. Qiu, and S. Hayashi, “Sensitized superbroadband near-IR emission in bismuth glass/Si nanocrystal superlattices,” Opt. Lett. 35(13), 2215–2217 (2010). [CrossRef]   [PubMed]  

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

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

29. 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), http://www.opticsinfobase.org/oe/fulltext.cfm?uri=oe-20-3-2562&id=226693. [CrossRef]   [PubMed]  

30. W. Xu, M. Peng, Z. Ma, G. Dong, and J. Qiu, “A new study on bismuth doped oxide glasses,” Opt. Express 20(14), 15692–15702 (2012), http://www.opticsinfobase.org/oe/fulltext.cfm?uri=oe-20-14-15692&id=239280. [CrossRef]   [PubMed]  

31. 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]  

32. E. Ahmed, D. Kohler, and M. Ruck, “Room-temperature synthesis of bismuth clusters in ionic liquids and crystal growth of Bi5(AlCl4)3,” Z. Anorg. Allg. Chem. 635(2), 297–300 (2009). [CrossRef]  

33. 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]  

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

35. S. Ulvenlund, A. Wheatley, and L. Bengtsson, “Synthesis of main-group metal clusters in organic solvents,” J. Chem. Soc. Chem. Commun. 1(1), 59–60 (1995). [CrossRef]  

36. J. Corbett, “Homopolyatomic ions of the heavy post-transition elements. the preparation, properties and bonding of Bi5(AlCl4)3 and Bi4(AlCl4),” Inorg. Chem. 7(2), 198–208 (1968). [CrossRef]  

37. S. Ulvenlund, K. Ståhl, and L. Bengtsson-Kloo, “Structural and quantum chemical study of Bi53+ and isoelectronic main-group metal clusters. The crystal structure of pentabismuth(3+) tetrachlorogallate(III) refined from X-ray powder diffraction data and synthetic attempts on its antimony analogue,” Inorg. Chem. 35(1), 223–230 (1996). [CrossRef]   [PubMed]  

38. R. C. Burns, R. J. Gillespie, and W.-C. Luk, “The preparation, spectroscopic properties, and structure of the pentabismuth(3+) cation, Bi53+,” Inorg. Chem. 17(12), 3596–3604 (1978). [CrossRef]  

39. B. Krebs, M. Mummert, and C. 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]  

40. A. N. Romanov, Z. T. Fattakhova, A. A. Veber, O. V. Usovich, E. V. Haula, V. N. Korchak, V. B. Tsvetkov, L. A. Trusov, P. E. Kazin, and V. B. Sulimov, “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,” Opt. Express 20(7), 7212–7220 (2012), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-20-7-7212. [CrossRef]   [PubMed]  

41. M. Lindsjö, A. Fischer, and L. Kloo, “Improvements of and insights into the isolation of bismuth polycations from benzene solution – single-crystal structure determinations of Bi8[GaCl4]2 and Bi5[GaCl4]3,” Eur. J. Inorg. Chem. 2005(4), 670–675 (2005). [CrossRef]  

42. S. Ulvenlund, L. Bengtsson-Kloo, and K. Stahl, “Formation of subvalent bismuth cations in molten gallium trichloride and benzene solutions,” J. Chem. Soc., Faraday Trans. 91(23), 4223–4234 (1995). [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 (5)

Fig. 2
Fig. 2 (a) Raman spectrum of Bi5(GaCl4)3-A-36h. Vibration modes of Bi53+ are labeled in the figure. Inset is the polycation group of Bi53+; (b) Emission (curves 1-3) and excitation (curves 4-5) spectra of Bi5(GaCl4)3-A-36h at room temperature. Curve 1 is upon 470nm excitation, curve 2 is the emission spectrum calibrated to the detector spectral response, and both of them were recorded with liquid N2 cooled NIR photomultiplier (Hamamatsu R5509-72). Curve 3 was collected with a PbSe detector under the excitation of 808nm laser diode. Curves 4 and 5 were measured by monitoring the emissions at 1400 and 1500nm, respectively. Curve 4 is shifted vertically for clarity. For reference, curve 6 an emission spectrum of Er3+ doped fluorophosphate glass is listed upon 808nm excitation and the FWHM is much smaller than Bi5(GaCl4)3.
Fig. 1
Fig. 1 (a) Unit cell representation of Bi5(GaCl4)3 on the basis of Rietveld refining results; (b) XRD pattern (black circle) of Bi5(GaCl4)3-A-36h, corresponding Rietveld refining results (solid red line, Rp = 7.47%, Rwp = 14.6%, Rexp = 11.0%, GOF = 1.76, RB = 3.95%), Bragg positions (vertical green line) and difference profile (oliver line) between observed and calculated values. Inset is the sample before separation.
Fig. 3
Fig. 3 (a) Absorption spectra of (1) Bi5(GaCl4)3-A-36h and (2) benzene, I0 and I are the incident and transmitted radiation through the sample; (b) Decay curve of the emission at 1365nm from Bi5(GaCl4)3-A-36h excited by 470nm at room temperature. Red circles are fitted results by an equation I = 442.596exp(-t/8.67) with χ2 = 1.086.
Fig. 4
Fig. 4 (a) Time resolved emission spectra of Bi5(GaCl4)3-A-36h excited by 470nm at room temperature. Delay times are marked beside the curves; (b) Dependence of the intensity of infrared emission@1835nm of Bi5(GaCl4)3-A-36h on pump power of 808nm laser diode.
Fig. 5
Fig. 5 (a) Dependence of fluorescence intensity of Bi5(GaCl4)3-A (-●-) and Bi5(GaCl4)3-B (-●-) on reaction time; (b) Evolution of emission spectrum of Bi5(GaCl4)3-A with reaction time. Curve 1: 2h; curve 2: 4h; curve 3: 8h; curve 4: 24h; curve 5: 36h; curve 6: 72h; curve 7: 96h.

Tables (2)

Tables Icon

Table 1 Wyckoff positions (wp) and coordinates for Bi5(GaCl4)3

Tables Icon

Table 2 A comparison of unit cell parameters (Å), distances (Å) and angles (deg) between Bi5(GaCl4)3 and Bi5(AlCl4)3

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.