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

Intense red photoluminescence from Mn2+-doped (Na+; Zn2+) sulfophosphate glasses and glass ceramics as LED converters

Open Access Open Access

Abstract

We report on intense red fluorescence from Mn2+-doped sulfophosphate glasses and glass ceramics of the type ZnO-Na2O-SO3-P2O5. As a hypothesis, controlled internal crystallization of as-melted glasses is achieved on the basis of thermally-induced bimodal separation of an SO3-rich phase. Crystal formation is then confined to the relict structure of phase separation. The whole synthesis procedure is performed in air at ≤ 800 °C. Electron spin resonance and Raman spectroscopy indicate that Mn2+ species are incorporated on Zn2+ sites with increasingly ionic character for increasing concentration. Correspondingly, in the glasses, increasing MnO content results in decreasing network polymerization. Stable glasses and continuously increasing emission intensity are observed for relatively high dopant concentration of up to 3 mol.%. Recrystallization of the glass results in strongly increasing emission intensity. Dynamic emission spectroscopy reveals only on type of emission centers in the glassy material, whereas three different centers are observed in the glass ceramic. These are attributed to octahedrally coordinated Mn2+ in the residual glass phase and in crystalline phosphate and sulfate lattices, respectively. Relatively low crystal field strength results in almost ideal red emission, peaking around 625 nm. Excitation bands lie in the blue-to-green spectral range and exhibit strong overlap. The optimum excitation range matches the emission properties of GaN- and InGaN-based light emitting devices.

©2010 Optical Society of America

1. Introduction

Divalent manganese ions play an important role as active centers in inorganic phosphors, mainly for lighting applications. They exhibit a 3d5 electronic configuration and photoemission usually occurs due to the transition 4T1(G) → 6A1(S). Therefore, the position of the emission band is strongly dependent on the field strength of the surrounding lattice. Typically, the emission peak lies between about 500 and 700 nm with a bandwidth of several tens of nanometers [13]. As a consequence, fluorescence from Mn2+ dopants can be used, for instance, for probing coordination numbers or other structural data in both glasses and crystalline materials [1,4]. On the other hand, the strong dependence of emission behaviour on host lattice makes the Mn2+ center a highly versatile and, thus, attractive component in the development of novel phosphors that emit in the green-to-red spectral range (e.g [5].). However, particularly for applications in solid-state lighting, knowledge on active centers other than rare-earth ions remains comparably limited. Their optical properties, on the other hand, often suggest significant potential as primary phosphor in combination with soft UV and blue excitation sources or as secondary phosphors to improve color rendering index (CRI) and perception of white light emitting devices (WLEDs) [6,7].

Contrary to the often employed solid state reaction for the production of inorganic phosphor materials, in the present study we are focussing on the glass ceramic route. That is, production of a glass by conventional melting and quenching of a mixture of raw materials, and controlled recrystallization of that glass in a secondary heating procedure. In a general way, this procedure offers various advantages over conventional solid state reactions, particularly related to compositional variance, dopant concentration and homogeneity, microstructural design, forming and production of microspheres, and recycling. In this context, as a means to stabilize Mn2+ over trivalent manganese, we have chosen a phosphate glass matrix. Phosphate glasses are long known as excellent hosts for high concentrations of transition metal ions as well as rare earth ions, what has made them the most important system for glass lasers, but also high-performance filters. They also exhibit relatively low optical basicity, what enables stabilization of polyvalent ions at lower redox states than are normally found in other glasses. Their major drawback, however, is their chemical stability and resistance to (atmospheric) water, what often renders them useless for any applications where they cannot be protected adequately. In the present case, another problem is presented by the desired controlled recrystallization of the as-melted glass. For that, internal nucleation (versus nucleation and crystal growth only at the surface) must be achieved, what cannot easily be done in phosphate glasses with relatively high kinetic fragility. Furthermore, even if effective heterogeneous nucleation agents were available, as additional components in the potential material, their optical and redox properties might strongly alter those of the final phosphor. Noteworthy, this prevents the use of many of the commonly [8] used nucleant species such as, e. g., TiO2, metallic nanoparticles, SnO2 or CeO2. Therefore, a different strategy was applied in the present case that resulted in a unique material. Two competing network forming units, PO4 3- and SO4 2-, were combined in a glass with relatively low total amount of network former (and high kinetic fragility). It is demonstrated that in this case, phase separation can be induced, resulting in controlled, homogeneous and finely distributed precipitation of an SO3-rich phase. This then gives rise to laterally confined crystallization of (Zn,Na2)SO3:Mn2+. The resulting glass ceramic exhibits intense red luminescence, in dependence on Mn2+ concentration and composition of the precursor glass, peaking between about 610 and 650 nm with a bandwidth FWHM of about 110 nm.

2. Experimental

Glass samples were prepared by conventional melting and quenching, using analytical grade ZnO, Na2CO3, ZnSO4·7H2O, NH4H2PO4 and MnCO3 as raw materials and adding ZnSO4·7H2O to obtain SO3 in excess of 3 mol.% to compensate for volatilization losses. Batches were homogenized by grinding mixtures of appropriate composition in an agate mortar. MnO-containing batches were first pre-calcined for 3 h at 300 °C, and subsequently melted at 750 °C for 1 h in alumina crucibles, using an alumina rod for stirring. For comparison, undoped glasses were melted in Au crucibles. Melting in alumina crucibles resulted in contamination of the glasses with < 0.2 mol.% Al2O3 (determined by electron dispersive spectroscopy, EDS). Noteworthy, this contamination had no observable effect on rheology, crystallization or optical properties of the derived glasses. Glass slabs of ~20 g were obtained by quenching on a graphite plate at room temperature. Nominal sample compositions are give in Table 1 . The glass transition temperature T g of the samples was estimated from differential scanning calorimetry (DSC, Netzsch DSC 404 F1) according to Ref [9]. All glass samples appeared highly transparent, what can be taken as a reliable indicator for 2 + as the dominant valence state of the manganese ions [Fig. 1(a) ]. Analyses of optical emission and excitation behavior were performed with a spectrofluorometer equipped with double monochromators (Czerny-Turner) in excitation and emission (Fluorolog3, Horiba Jobin Yvon, spectral resolution of ~0.1 nm), using a 400 W Xe-lamp for static and a 75 W Xe-flashlamp for dynamic analyses, respectively, as excitation sources. Raman spectra were collected on a Nicolet Almega XR dispersive Raman spectrometer. The crystallization process was studied in situ, using a high-temperature X-ray diffractometer (XRD) at a heating rate of 5 K/min. Finally, electron spin resonance spectra (ESR) of Mn-doped samples were recorded on an X-band microwave spectrometer at a frequency of 9.7 GHz (Bruker ESR 300E). With the exception of XRD, all measurements were carried out at room temperature.

Tables Icon

Table 1. Nominal compositions and basic physical data of examined glasses.

 figure: Fig. 1

Fig. 1 Optical excitation (monitoring wavelength of emission peak) and emission spectra (excitation wavelength of 409 nm), (A), and corresponding electronic band structure (B) of Mn2+-doped sulfophosphate glasses of type SP11. Labels in (A) indicate concentration of MnO in mol.%. Inset: Photograph of samples SP11_30 (top), SP11_07 (middle), SP11_01 (bottom).

Download Full Size | PDF

3. Results and discussion

3.1 Photoluminescence in Mn2+-doped precursor glasses

The electronic structure of manganese (3d5) is relatively well known. It comprises the excited terms 4G, 4P, 4D and 4F [10]. For Mn2+, the ground state is 6A1(S) and at least five transitions can readily be identified in the optical excitation spectra of Mn-doped sulfophosphate glasses [Fig. 1(a)], peaking at 348.9 nm, 360.7 nm, 409.4 nm, 421.9 and 504.4 nm, respectively. These correspond to the transitions 6A1(S) → 4E(D), 6A1(S) → 4T2(D), 6A1(S)→ 4A1(G), 4E, 6A1(S) → 4T2(G) and 6A1(S) → 4T1(G), respectively [Fig. 1(b)]. A further excitation band can be detected at ~317.5 nm, corresponding to 6A1(S) → 4T1(P), if the concentration of MnO exceeds ~0.5 mol.%. Noteworthy, for increasing manganese content, samples exhibit increasing violet coloration [inset of Fig. 1(a)]. The origin of this lies in an absorption band at ~490 nm which can readily be assigned to traces of Mn3+ (3d4; 5E →5T2 [11] (note that Mn3+ does not exhibit optical emission in the 600-nm spectral range). Due to strong overlap particularly of the lower three bands (excitation to 4A1(G), 4T2(G) and 4T1(G), respectively), excitation can in principle be performed with a relatively broad blue source, i.e. from about 400 to 530 nm. On the other hand, the band around 409 nm (4E(G) →4A1(S)) appears most effective. The ratio between blue and UV excitation, I 409nm to I 360nm increases with increasing Mn2+ concentration. In the same time, emission intensity increases continuously, surprisingly without notable concentration quenching for up to 3 mol.% of dopant concentration (noteworthy, concentration quenching was observed for SO3-free zinc phosphate glasses at MnO > 2.5 mol.% upon excitation at 410 nm [4]).

Emission always occurs due to 4T1(G) → 6A1(S), and a significant red-shift as well as increasing emission intensity are observed with increasing concentration of Mn2+ [Fig. 1(a)]. Linear regression of the integrated emission intensity I versus manganese concentration c MnO leads to I(a.u.) = 9.86397 × c MnO(mol%) with an R-value of 0.98757. The line width at half maximum (FWHM) is ~110 nm for all samples. In a first consideration, the position of the emission band suggests octahedral coordination of the Mn2+ ions (while tetrahedrally coordinated Mn2+ usually results in green emission [12]). Noteworthy, changing the SO3 to P2O5 ratio at constant MnO concentration (samples SP6_07; SP11_07; SP16_07 and SP19_07) results in a slight blue-shift of emission for increasing amount of SO3. Considering the d5 Tanabe-Sugano diagram, this indicates decreasing crystal field strength with increasing SO3.

ESR spectra of doped glasses are shown in Fig. 2(a) , revealing the typical [13,14] fingerprint of Mn2+ (no resonance was observed in undoped glasses). It comprises a sextet hyperfine line structure. As a result of concentration quenching, this sub-structure disappears at MnO ≥ 2 mol.% [15,16]. Obtained spectra indicate increasing hyperfine interaction with increasing Mn2+ concentration and, hence, increasingly ionic character of the bonds between Mn2+ and its surrounding ligands [17,18]. The increasing interaction between Mn2+-species may further be taken as an explanation for the observed red-shift in optical emission with increasing concentration [19].

 figure: Fig. 2

Fig. 2 ESR (A) and Raman spectra of Mn2+-doped sulfophosphate glasses of type SP11. Labels indicate concentration of MnO (mol.%). Inset: Intensity ratio of the Raman bands at 1050 cm−1 and 992 cm−1.

Download Full Size | PDF

Raman spectra further confirm this conclusion [Fig. 2(b)]. They exhibit four major bands: at 1410 cm−1 (symmetric P = O stretch [20], ), 1050 cm−1 (symmetric stretching vibration in PO3 [20]), 992 cm−1 (υ 1-vibration in SO4 2- [21,22]) and 750 cm−1 (stretching mode in bridging oxygen [21]). Relative to the Raman active SO4 2- band, the νs-PO3 band decreases with increasing Mn2+-concentration (inset of Fig. 2). Considering charge balance, ionic radii and coordination, it is reasonable to assume that Mn2+ ions are incorporated in the lattice on Zn2+ sites (note that Zn2+ is in octahedral coordination [23], and above luminescence data indicates octahedral coordination for Mn2+, too). An increase in (Zn2+ + Mn2+) then results in further decreasing polymerization of the P2O5-network [21]. This is reflected by decreasing intensity of the 1050 cm−2 and 750 cm−2 bands.

3.2 Sulfophosphate glass ceramics

In terms of emission efficiency, the preferred environment of active centers should usually be crystalline rather than glassy. To fully benefit from the several advantages of glass forming, but to also obtain high quantum yield, however, controlled recrystallization – the formation of a glass ceramic – may be employed. Thereby, the question whether or not nucleation and crystal growth can be initiated in the bulk material or only at free surfaces is decisive for any potential application. As noted before, bulk (internal) nucleation is usually achieved via the use of chemical agents that promote heterogeneous nucleation, often by epitaxial growth of the desired phase on the nucleant phase. In the present case, as with most phosphate systems, however, such nucleation agents are on the one side not known. On the other side, common choices would have significant impact on the optical properties as well as on the redox distribution of the manganese ions. Therefore, a different approach was pursued that has originally been described by Beall for the controlled precipitation of mullite from an aluminosilicate system [24]: A glass is formed from two glass formers so that its composition lies within the mixability gap of the system, but phase separation is kinetically prevented. Via re-heating this glass, the kinetic barrier is overcome and phase separation occurs, what results in a significant shift of the chemical (and rheological) properties of the demixed phases. Then, crystallization may occur in the demixed phase, either homogenously or at the newly-created interface. Crystal precipitation remains confined to the so-called relict structure [24] of the previous separation process and can thus be controlled to occur finely distributed over the bulk material. This approach was applied in the present case. Upon reheating the glass, the presence of competing network forming structural units PO4 3- and SO4 2- in comparable quantities triggers precipitation of an SO4 2--rich droplet phase. From this phase, crystallization is initiated. This process can be observed in situ by high-temperature XRD (Fig. 3 ). In dependence on SO3-concentration, crystallization is roughly described as a two-step process. Starting at ~370°C (SP16) and ~390°C (SP19), respectively, crystalline sulfate and sulfophosphate phases are initially precipitated. These comprise dominantly Na2SO3 and (Na2, Zn)SO3. At this point, crystallization occurs internally, what is clearly evidenced by gradually evolving opalescence of the samples if heat-treated in this temperature regime. Noteworthy, this step cannot be observed in sample SP6. Assumedly, in that case, SO3-concentration lies within the solubility limit of the phosphate matrix. The second crystallization process is characterized by the precipitation of mixed (Na, Zn) phosphate phases, dominantly via surface crystallization. This second process resembles typical devitrification behaviour of phosphate glasses, occurring in either sample at temperatures between ~470 °C (SP6) and 400 °C (SP19). It is undesired in the sense that it cannot be controlled, does not occur homogeneously over the bulk sample and prevents simple transformation of a glass body into a glass ceramic while maintaining its external geometry (e.g. fiber). Hence, the observed process window for the preparation of glass ceramics lies in the temperature range of ~370°C to ~390°C. If the relatively high heating rates and short observation times, respectively, that were applied in the HT-XRD analyses are taken into account, these temperature limits should be reduced by about 10-20 K.

 figure: Fig. 3

Fig. 3 In situ X-ray diffraction patterns as taken during heating samples SP6, SP16 and SP19 (varying SO4 2-/PO4 3- ratio). Labels indicate respective temperature. Spectra are initially dominated by sulfate phases Na2SO4, JCPDS 086-0800 (red arrows); ZnSO4, JCPDS 070-1255 (green dotted line in SP16), later by various phosphate phases (blue arrows).

Download Full Size | PDF

As a confirmation of the motivating assumption, the resulting glass ceramics exhibit several times higher emission intensity, as compared to the as-melted glass at equivalent composition and MnO-concentration (Fig. 4 ). Besides this significant intensity increase, on a second view, the emission spectrum of the crystallized sample appears slightly blue-shifted and strongly asymmetric. A more detailed view on this is obtained if emission spectra for various excitation wavelengths are considered (Fig. 5 ). Qualitatively, on the crystallized sample, a shift of the peak emission wavelength between three dominant positions can be observed, i.e. ~620 nm, 628 nm and 616 nm for an excitation wavelength of ~350 nm, 409 nm and 500 nm, respectively [Fig. 5(a)]. On the other hand, in the as-made glass, no dependence of emission peak position on excitation wavelength could be detected [Fig. 5(b)]. This indicates that in the glass ceramic, emission centers are present in three different ligand fields. Compared to as-made glasses (0.7 mol.% MnO, emission peak at ~625 nm), two of the three positions result in a slight blue shift of emission, thus indicating lower crystal field strength [25]. The third position practically equals that of a glassy environment with low SO3 concentration. This is consistent with principle expectations that arise with the synthesis of the glass ceramics. Precipitation of sulfate phases results in a decrease of SO3 concentration in the residual glass phase. Noteworthy, emission as a function of SO3-content exhibits a slight red-shift for decreasing SO3 (Fig. 4). Then, the glass ceramic emission peak at 628 nm is attributed to Mn2+-centers in the residual glass phase. Upon crystallization, Mn2+ can be incorporated on Zn2+-sites in zinc phosphate as well as zinc sulfate lattices, what results, in consistence with XRD data, in two principle emission centers. In both cases, Zn2+ and, hence, Mn2+ remains in octahedral coordination. Therefore, the extent of the observed blue-shift is relatively small. Considering the ionic radii of S6+~29 pm and P5+~38 pm, respectively, a stronger field would be expected to act on Mn2+ in a phosphate lattice as compared to a sulfate lattice. Then, emission at 620 nm should be attributed to the phosphate species (higher crystal field strength), while emission at 616 nm originates from Mn2+ in sulfate environment. A corresponding picture is revealed by dynamic emission data (Fig. 6 ). Delayed spectroscopy reveals changes in shape and position of the dominant emission peak with increasing delay time. Namely, in accordance with above arguments, the emission spectrum is dominated, in sequence, by peaks at 620 nm, 628 nm and 616 nm, respectively [Fig. 6(a)]. This is taken as another evidence for the presence of at least three emission centers, and is confirmed by the decay kinetic which clearly does not follow a first order exponential equation [Fig. 6(b)], particularly in the beginning period. The effective lifetime (1/e times initial intensity) is 14.0 and 16.8 ms for glass and glass ceramic, respectively.

 figure: Fig. 4

Fig. 4 Optical emission spectra of as-made glass SP16_10 and corresponding glass ceramic (heat treated for 4h at 380 °C). Inset: Photographs of both samples (A) under ambient light and (B) under 40 W UV-A broadband illumination.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Optical emission spectra of glass ceramic (A) and glass (B) SP11_07 for varying excitation wavelength (labels).

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Time-resolved delay curves of emission from glass ceramic (A) and luminescence decay curves (B) of glass and glass ceramic.

Download Full Size | PDF

4. Conclusions

In summary, we report on intense red fluorescence from Mn2+-doped sulfophosphate glasses and glass ceramics of the type ZnO-Na2O-SO3-P2O5. Due to the low optical basicity of the phosphate matrix, stabilization of Mn2+ over Mn3+ can readily be achieved in this system, even for high dopant concentrations, when preparing the glasses in air. Through thermally induced bimodal separation of a sulphate-rich phase, finely distributed precipitation of (Na, Zn) sulfate phases by internal crystallization is achieved at temperatures around 350-390 °C, and crystal growth is confined to the relict structure of phase separation. At higher temperatures, a secondary crystallization process may be initiated. In that case, dominantly zinc and (Zn,Na) phosphates are precipitated by surface crystallization. In both glasses and glass ceramics, photoemission occurs at around 620-650 nm with a bandwidth of ~110 nm. Emission is red-shifted for increasing Mn2+ concentration, and blue-shifted for increasing ratio of SO3:P2O5. Emission intensity is linearly increasing with increasing dopant concentration up to 3.0 mol.%, and no quenching effects could be observed for this concentration range. In the glassy state, only one type of emission centers could be detected, i.e. Mn2+ on octahedral Zn2+ sites. The position of the emission peak and shifts associated with changes in matrix composition and dopant concentration can clearly be attributed to changes in ligand field strength. As indicated by ESR and Raman spectroscopy, increasing Mn2+ content leads to further depolymerization of the glass lattice, and the character of bonds between Mn2+ and the lattice becomes increasingly ionic. Recrystallization of the glass and, hence, formation of a glass ceramic results in a significant increase of emission intensity. Then, dynamic emission spectroscopy reveals three different emission centers. On the basis of crystal field splitting, these are attributed to octahedrally coordinated Mn2+ in the residual glass phase and in crystalline phosphate and sulfate lattices, respectively. Excitation bands lie in the blue-to-green spectral range and exhibit strong overlap, and the optimum excitation range matches the emission properties of GaN- and InGaN-based LEDs. Noteworthy, in the same time as emission intensity increases with dopant concentration, the ratio between blue and UV excitation increases as well, what presents another significant advantage for application in WLED devices.

Compared to conventional solid-state synthesis of alternative materials, the glass ceramic route offers unique advantages with respect to composition, dopant concentration, formation of microspheres and recycling.

Acknowledgements

The authors gratefully acknowledge funding from the German Excellence Initiative within the cluster “Engineering of Advanced Materials – Hierarchical Structure Formation for Functional Devices.”

References and links

1. R. Reisfeld, A. Kisilev, and C. K. Jorgensen, “Luminescence of manganese(II) in 24 phosphate glasses,” Chem. Phys. Lett. 111(1-2), 19–24 (1984). [CrossRef]  

2. B. Sudhakar Reddy, N. O. Gopal, K. V. Narasimhulu, C. Linga Raju, J. L. Rao, and B. C. V. Reddy, “EPR and optical absorption spectral studies on Mn2+ ions doped in potassium thiourea bromide single crystals,” J. Mol. Struct. 751(1-3), 161–167 (2005). [CrossRef]  

3. I. E. C. MacHado, L. Prado, L. Gomes, J. M. Prison, and J. R. Martinelli, “Optical properties of manganese in barium phosphate glasses,” J. Non-Cryst. Solids 348, 113–117 (2004). [CrossRef]  

4. M. C. Flores J, U. Caldino G, J. Hernández A, E. Camarillo G, E. Cabrera B, H. del Castillo, A. Speghini, M. Bettinelli, and H. Murrieta S, “Study of Mn2+ luminescence in Zn(PO3)2 glasses,” Phys. Status Solidi C 4(3), 922–925 (2007). [CrossRef]  

5. S. Yuan, Y. Yang, X. Zhang, F. Tessier, F. Cheviré, J. L. Adam, B. Moine, and G. Chen, “Eu2+ and Mn2+ codoped Ba2Mg(BO3)2--new red phosphor for white LEDs,” Opt. Lett. 33(23), 2865–2867 (2008). [CrossRef]   [PubMed]  

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

7. 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). [CrossRef]   [PubMed]  

8. L. Wondraczek and P. Pradeau, “Transparent hafnia-containing β-quartz glass ceramics: nucleation and crystallization behavior,” J. Am. Ceram. Soc. 91(6), 1945–1951 (2008). [CrossRef]  

9. L. Wondraczek, H. Behrens, Y. Yue, J. Deubener, and G. W. Scherer, “Relaxation and Glass Transition in an Isostatically Compressed Diopside Glass,” J. Am. Ceram. Soc. 90(5), 1556–1561 (2007). [CrossRef]  

10. J. Sugar and C. Corliss, “Atomic energy levels of the iron period elements: potassium through nickel,” J. Phys. Chem. Ref. Data 14, 1–664 (1985).

11. A. J. Faber, A. van Die, G. Blasse, and F. van der Weg, “Luminescence of manganese of different valencies in oxide glasses,” Phys. Chem. Glasses 28, 150–155 (1987).

12. V. B. Mikhailik, “VUV sensitization of Mn2+ emission by Tb3+ in strontium aluminate phosphor,” Mater. Lett. 63(9-10), 803–805 (2009). [CrossRef]  

13. C. Sumalatha, B. Sreedhar, M. Yamazaki, and K. Kojima, “Electron paramagnetic resonance and optical absorption spectra of Mn(II) ions in silica sol-gel,” Phys. Chem. Glasses 38, 206–210 (1997).

14. J. Qiu, C. Zhu, T. Nakaya, J. Si, K. Kojima, F. Ogura, and K. Hirao, “Space-selective valence state manipulation of transition metal ions inside glasses by a femtosecond laser,” Appl. Phys. Lett. 79(22), 3567–3569 (2001). [CrossRef]  

15. H. W. de Wijn and R. F. van Balderen, “Electron Spin Resonance of Manganese in Borate Glasses,” J. Chem. Phys. 46(4), 1381–1387 (1967). [CrossRef]  

16. D. L. Griscom and R. E. Griscom, “Paramagnetic Resonance of Mn2+ in Glasses and Compounds of the Lithium Borate System,” J. Chem. Phys. 47(8), 2711–2722 (1967). [CrossRef]  

17. J. S. van Wieringen, “Paramagnetic resonance of divalent manganese incorporated in various lattices,” Discuss. Faraday Soc. 19, 118–126 (1955). [CrossRef]  

18. B. Vaidhyanathan, C. Prem Kumar, J. L. Rao, and K. J. Rao, “Spectroscopic investigations of manganese ions in microwave-prepared NaPO3—PbO glasses,” J. Phys. Chem. Solids 59(1), 121–128 (1998). [CrossRef]  

19. C. R. Ronda and T. Amrein, “Evidence for exchange-induced luminescence in Zn2SiO4:Mn,” J. Lumin. 69(5-6), 245–248 (1996). [CrossRef]  

20. R. K. Brow, “Review: the structure of simple phosphate glasses,” J. Non-Cryst. Solids 263–264(1-2), 1–28 (2000). [CrossRef]  

21. E. I. Kamitsos, M. A. Karakassides, and G. D. Chryssikos, “A vibrational study of lithium sulfate based fast ionic conducting borate glasses,” J. Phys. Chem. 90(19), 4528–4533 (1986). [CrossRef]  

22. M. Ganguli and K. J. Rao, “Studies of ternary Li2SO4–Li2O–P2O5 glasses,” J. Non-Cryst. Solids 243(2-3), 251–267 (1999). [CrossRef]  

23. K. J. Rao and H. G. K. Sundar, “Electrical conductivity studies in K2SO4-Na2SO4-ZnSO4 glasses and the mixed alkali effect,” Phys. Chem. Glasses 21, 216–220 (1980).

24. W. Höland, and G. H. Beall, “Glass ceramic technology,” American Ceramic Society, Westerville, OH, USA, 2002.

25. F. N. Su and Z. Deng, “Influence of chemical environment on the optical properties in transition metal ions doped materials,” J. Fluoresc. 16(1), 43–46 (2006). [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 (6)

Fig. 1
Fig. 1 Optical excitation (monitoring wavelength of emission peak) and emission spectra (excitation wavelength of 409 nm), (A), and corresponding electronic band structure (B) of Mn2+-doped sulfophosphate glasses of type SP11. Labels in (A) indicate concentration of MnO in mol.%. Inset: Photograph of samples SP11_30 (top), SP11_07 (middle), SP11_01 (bottom).
Fig. 2
Fig. 2 ESR (A) and Raman spectra of Mn2+-doped sulfophosphate glasses of type SP11. Labels indicate concentration of MnO (mol.%). Inset: Intensity ratio of the Raman bands at 1050 cm−1 and 992 cm−1.
Fig. 3
Fig. 3 In situ X-ray diffraction patterns as taken during heating samples SP6, SP16 and SP19 (varying SO4 2-/PO4 3- ratio). Labels indicate respective temperature. Spectra are initially dominated by sulfate phases Na2SO4, JCPDS 086-0800 (red arrows); ZnSO4, JCPDS 070-1255 (green dotted line in SP16), later by various phosphate phases (blue arrows).
Fig. 4
Fig. 4 Optical emission spectra of as-made glass SP16_10 and corresponding glass ceramic (heat treated for 4h at 380 °C). Inset: Photographs of both samples (A) under ambient light and (B) under 40 W UV-A broadband illumination.
Fig. 5
Fig. 5 Optical emission spectra of glass ceramic (A) and glass (B) SP11_07 for varying excitation wavelength (labels).
Fig. 6
Fig. 6 Time-resolved delay curves of emission from glass ceramic (A) and luminescence decay curves (B) of glass and glass ceramic.

Tables (1)

Tables Icon

Table 1 Nominal compositions and basic physical data of examined glasses.

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.