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

Crystal structure and variation of luminescence properties of (Ba,Ca)Si7N10:Eu2+ as a function of the Eu and Ca concentration

Open Access Open Access

Abstract

Cyan-emitting (Ba,Ca)Si7N10:Eu2+ was synthesized using a high-temperature graphite furnace. Crystallographic data for BaSi7N10:0.1Eu2+ were obtained by Rietveld refinement and discussed compared with un-doped BaSi7N10. The excitation spectra peaked at 368 nm and decay times were apparently different from those of earlier papers, and these differences were explained by the structural analyses. The substitution of Ca2+ ions for Ba2+ ions increased the emission intensity by approximately 130%, as this resulted in local lattice modification. The findings indicate that (Ba,Ca)Si7N10:Eu2+ phosphors are highly stable and well suited for use in white LEDs pumped by near ultraviolet sources.

© 2014 Optical Society of America

1. Introduction

Nitride/oxynitride phosphors have been investigated extensively for use in white light emitting diodes (LEDs), because these materials exhibit high quantum efficiencies, high thermal and chemical stabilities, and low thermal quenching properties. For example, Eu2+-doped M2Si5N8 (M: Ba, Sr) [16] and CaAlSiN3 [79] are known to be red-emitting phosphors, while Eu2+-doped β-sialon [10] and (Sr,Ca,Ba)Si2O2N2 [11] are used as green phosphors. It was recently reported that Sr2Si(O1-xNx)4 powders doped with Eu2+ ions exhibit red emissions [12, 13].

In addition to the above-mentioned nitride phosphors, BaSi7N10:Eu2+ has also been suggested as a cyan-emitting phosphor [1417]. BaSi7N10 is a highly condensed ternary silicon nitride. It has a network structure composed of corner- and edge-sharing SiN4 tetrahedra; the Ba2+ ions occupy the cavities within the network structure of [Si7N10]2-. A Ba atom is coordinated by 13 nitrogen atoms, with the Ba–N distances being 2.91 – 3.53 Å, while the coordination number (CN) of Ba can be 12, if only the nitrogen atoms corresponding to Ba–N distances of approximately 3.50 Å are considered [14,18,19].

Earlier studies have reported that the photoluminescence (PL) spectrum of BaSi7N10:Eu2+ contains excitation and emission bands that peak at approximately 300 – 330 nm and 475–500 nm, respectively [1417]. Further, both band types are located at wavelengths shorter (i.e., at higher energies) than those in the case of red-emitting Ba2Si5N8:Eu2+ powders [1,2]. This difference in wavelengths is attributable to the fact that the crystal field splitting of the Eu2+ ions in BaSi7N10 is weaker. The CN of Ba (13 N atoms) in BaSi7N10 is approximately twice as large as those for the Ba(I) and Ba(II) sites in Ba2Si5N8, which were 6 and 7, respectively, resulting in the volume of the polyhedron in BaSi7N10 being much larger and the average bond length being longer. BaSi7N10:Eu2+ powders have been prepared by various methods: the nitridation of each metal source and the solid-state reaction process [14], the sol-gel process [15], the nitridation of metal alloys [16], and the conventional solid-state reaction process using barium nitrates, Eu2O3, and Si3N4 [17].

In spite of the high potential of BaSi7N10:Eu2+ for use in white LEDs, there have been few reports on its luminescent properties, in contrast to the case for other nitride phosphors. In this study, we prepared BaSi7N10:Eu2+ powders through a simple and low-cost process that involved the use of a high-temperature graphite furnace. We also investigated the effects of the Ba content in the starting mixtures and the substitution of Ca2+ ions at the Ba2+ sites on the luminescent spectrum and intensity exhibited by the synthesized powders.

2. Experimental details

The (Ba,Ca)Si7N10:Eu2+ powders were synthesized using a high-temperature graphite furnace. BaCO3 (High Purity Chemicals, 99.95%), CaCO3 (High Purity Chemicals, 99.95%), Si3N4 (Ube Industry, Ltd., E10, 99.9%), and Eu2O3 (Grand Chemicals & Materials, 99.99%) were used as the starting materials. The mixtures of (x-y-z)BaCO3yCaCO3–7/3Si3N4–0.5zEu2O3 were ball milled for 24 h, calcined, and fired at 1600 °C for 5 h in a nitrogen (99.999%) atmosphere using graphite crucibles. The crystalline phases of the synthesized powders were determined with an X-ray diffractometer (XRD, Rigaku MiniFlex II) using CuKα radiation (λ = 1.5406 Å); Si powders were used as the reference. Rietveld refinement was performed using X'pert HighScore Plus (PANalytical B. V.). The particle morphologies and composition of the powders were analyzed using a field emission scanning electron microscopy (FE-SEM, JEOL JSM-6500F) with an energy dispersive spectrometry (EDS) attachment. The average atomic ratios were estimated by calibrating the measured ones using a reference powder consisting of BaCO3, CaCO3, Si3N4, Eu2O3. The nitrogen and oxygen contents (wt%) of the powders were measured using a nitrogen/oxygen analyzer (LECO, TC600). The PL spectra, quantum efficiencies (QEs), and thermal quenching were determined using by a PL system (PSI, Darsa 5000) with a 500 W xenon lamp as an excitation source. The external QE is the ratio of the number of photons of the emission to that of the excitation light irradiated on a phosphor, while the internal QE is the ratio of the number of photons of the emission to that of the excitation light absorbed by a phosphor. QEs were measured using an integrating sphere within the PL system, and BaSO4 was used as a reference. The decay curves were measured with a time-resolved PL (TRPL) streak-scope system (Hamamatsu), using a 374 nm pulsed laser diode with a pulse duration of 65 ps. All the spectral measurements were performed at room temperature.

3. Results and discussion

3.1. Crystal structure

A number of BaSi7N10:0.1Eu2+ powders were synthesized using different amounts of BaCO3 in the starting mixtures of (x-0.1)BaCO3–7/3Si3N4–0.05Eu2O3. The XRD patterns of the synthesized powders are shown in Fig. 1(a). For x = 0.6, the XRD pattern was predominantly composed of peaks corresponding to BaSi7N10 (JCPDS #89-6751). A number of weak unknown peaks (◆), which were attributed to BaCO3 being present in the starting mixtures in an insufficient amount, were also observed. For x = 0.8 – 1.2, a pure BaSi7N10 phase was observed; its XRD patterns did not changed much over the range of x values. In addition, the unknown peaks noticed in the case of x = 0.6 disappeared completely.

 figure: Fig. 1

Fig. 1 (a) XRD patterns of the samples prepared with (x-0.1)BaCO3, (b) experimental (red dots) and calculated (blue solid line) XRD profiles of Rietveld refinement for BaSi7N10:0.1Eu2+ powders corresponding to x = 0.8, and (c) a difference between the experimental and calculated data. Blue bars represent Bragg reflections.

Download Full Size | PDF

The variation in the Ba/Si atomic ratio (RBa/Si) of the prepared powders is shown in Fig. 2; the estimated RBa/Si values were higher than the theoretical ones, and the powders prepared with x = 0.8 had the same RBa/Si value (~0.13 ± 3%) as the theoretical one (~0.13) with the stoichiometric composition. The difference between the estimated and theoretical RBa/Si values increased gradually with an increase in x. This phenomenon can be ascribed to the formation of a volatile SiO phase in the reductive atmosphere resulting from the use of the graphite furnace and crucibles. The measured nitrogen content of the BaSi7N10:0.1Eu2+ powders corresponding to x = 0.8 was approximately 30.5 wt%; this value was almost the same as the one calculated (29.5 wt%), if one takes into consideration errors in measurement. On the other hand, the measured oxygen content was 0.26 wt%. These results demonstrate that the simple and low-cost process employed in the present study perfectly suited for the synthesis of BaSi7N10 powders. Possible reactions for the nitridation are speculated to be as follows: BaCO3 decomposed into BaO and CO2 (g) during calcination, and then, BaO reacted with C to afford Ca and CO (g); in this case, the C and CO partial pressures in the reacting chamber were assumed to be appropriately controlled. However, oxygen production from Eu2O3 was probably inevitable to a certain extent, resulting in the local incorporation of oxygen ions in the BaSi7N10 lattice, but it was negligible owing to a small amount of Eu2O3 in the starting mixtures. G. Anoop et al. prepared BaSi7N10:Eu2+ powders by the conventional solid-state reaction process using BaCO3 at 1600 °C under the reduction atmosphere, but the detailed preparation conditions, the atomic composition, and the nitrogen content were not reported [17].

 figure: Fig. 2

Fig. 2 Variation in the Ba/Si atomic ratios (RBa/Si) of the samples prepared with (x-0.1)BaCO3.

Download Full Size | PDF

Rietveld refinement was performed for BaSi7N10:0.1Eu2+ powders corresponding to x = 0.8; XRD profiles are shown in Figs. 1(b) and 1(c), and crystallographic data are summarized in Table 1 compared with those of a reference (BaSi7N10, JCPDS #89-6751). A refinement of the site occupancy factor (SOF) confirmed that the refined composition of this sample was Ba0.9Eu0.1Si7N10, and that each crystallographic site was fully occupied. These findings indicated that 10 mol% Eu2+ ions were successfully substituted for the Ba2+ sites, and the BaSi7N10:0.1Eu2+ pure nitride with the stoichiometric composition could be obtained by the synthesis process of this study. The refined composition was just in accordance with that estimated by the results of EDS and N/O analysis. Ba0.9Eu0.1Si7N10 has a monoclinic system with a space group of Pc, while its lattice parameters a, b, and c slightly decreased when compared with those of BaSi7N10, because the ionic radii for Eu2+ and Ba2+ are evidently different: 1.35 Å (CN = 10) and 1.52 Å (CN = 10) for Eu2+ and Ba2+, respectively. However, the difference in a unit cell volume (V) and an axis angle β is insignificant. Selected interatomic distances (Eu–N) of Ba0.9Eu0.1Si7N10 are listed in Table 2.The shortest and longest distances of Eu–N are 2.910 Å and 3.490 Å, respectively, which are very nearly equal to or slightly shorter than those of Ba-N for BaSi7N10 (2.913 Å and 3.529 Å, respectively). In summary, the findings indicated that the changes in the lattice parameters and interatomic distances were insignificant, even though the smaller Eu2+ ions were substituted for the Ba2+ ions. This was due to the rigidity of the networks of BaSi7N10. Pilet al. suggested that, when compared with those of BaSi7N10, the high rigidity of the highly condensed [Si7N10]2- networks of SrSi7N10 resulted in the slight change in the crystal parameters in spite of the smaller ionic size of Sr2+ (1.61 Å for Ba2+ and 1.44 Å for Sr2+, CN = 12) [20].

Tables Icon

Table 1. Crystallographic data for BaSi7N10 and BaSi7N10:0.1Eu2+ (x = 0.8).

Tables Icon

Table 2. Interatomic distances (Å) of Eu-N for Ba0.9Eu0.1Si7N10.

3.2. Photoluminescence of (Ba,Ca)Si7N10:Eu2+

The PL excitation (PLE) and PL spectra of the various powders are shown in Fig. 3(a) and 3(b). The PLE spectra ranged from 250 nm to 430 nm, and their intensities changed with x. Li et al. have suggested that the asymmetric PLE band noticed in the spectrum of BaSi7N10:Eu2+ is attributable to the overlapping Gaussian sub-bands at approximately 296, 332, and 363 nm [14]. The PLE spectra observed in the present study could also be decomposed into three sub-bands (A, B, and C), which are represented by dotted lines in Fig. 3(a); either the A-band or the B-band was the strongest one, depending on the value of x. Compared to those in the case of x = 0.6 – 0.8, the B-bands for x = 1.0 – 1.2 were significantly blue-shifted, while the A-bands were shifted slightly toward shorter wavelengths. These blue-shifts were related to the compositions of the powders.

 figure: Fig. 3

Fig. 3 (a) PLE and (b) emission spectra of the samples prepared with (x-0.1)BaCO3.

Download Full Size | PDF

The ratios of the intensities of the A- and B-bands (IR = IA-band/IB-band) are shown in Fig. 4. For x = 0.6–0.8, the IR values were greater than one, but decreased significantly for x = 1.0 or greater, becoming less than one. When compared with the variation in the Ba/Si ratio (Fig. 2), it was evident that the A-band was stronger than the B-band in the Ba-deficient and the stoichiometric regions, and vice versa in the Si-deficient region. This result demonstrated that the energy transition of Eu2+ ions from the ground state to the excited energy levels E1, E2, and E3, which corresponded to the A-, B-, and C-bands, respectively, was greatly affected by the crystal structure of BaSi7N10. In summary, the BaSi7N10:0.1Eu2+ powder with the stoichiometric composition (i.e., the one corresponding to x = 0.8) exhibited the strongest excitation intensity at approximately 368 nm, indicating that it has potential for use in white LEDs pumped by near-ultraviolet (nUV) sources.

 figure: Fig. 4

Fig. 4 Variation in the IR values of the samples prepared with (x-0.1)BaCO3.

Download Full Size | PDF

The PLE spectra obtained in the present study were completely different from those listed in earlier studies [1417], which reported that the strongest excitation bands occurred at approximately 280–330 nm; these corresponded to the C- or B-band in Fig. 3(a). The Eu2+ ions substituted for the Ba2+ ions occupy the cavities within the network structure of [Si7N10]2- and are coordinated by 13 nitrogen atoms. Therefore, the crystal-field splitting of the 5d level of the Eu2+ ions depends strongly on the nitrogen and oxygen contents. Fang et al. calculated the electronic structure of BaSi7N10 using first-principles calculations [18]. They confirmed that the top of the valence band of the compound was dominated by N [2] atoms (coordinated by two Si atoms), while the conduction band was determined mainly by the Ba 6s states. The experimentally determined optical band gap of BaSi7N10 is roughly 4 – 4.8 eV [14,18]. Thus, the highest excited energy level of the Eu2+ ions (C-band at 296 nm, ~4.2 eV) lie fully or partly within the conduction band. This suggests that the energy-level scheme of the Eu2+ ion is closely related to the crystal structure of BaSi7N10. Thus, it is likely that the differences in the PLE spectra reported in previous studies and those determined in the present study are attributable to the differences in the compositions, crystal defects, and nitrogen (oxygen) contents of the investigated compounds. The effects of the crystal defects are explained in detail in the discussion on the decay time (Fig. 8).

The dominant-peak wavelengths (DPWs) of the PL emission spectra (Fig. 3(b)) were 478 nm, irrespective of the value of x and the excitation wavelength used (A-, B-, and C-bands). This indicates that there is only one activation enter (Eu2+ sites) surrounded by the 13 nitrogen atoms. The intensity of the emission was dependent on the composition (Fig. 2) and the phase purity of the compounds (Fig. 1) and was the strongest for x = 0.8.

The QE and thermal quenching of BaSi7N10:0.1Eu2+ (x = 0.8) were investigated; commercialized (Sr,Ba)-orthosilicate green phosphor was used as the reference. The internal and external QEs were approximately 107% and 79%, respectively, of those of the reference. This demonstrates that BaSi7N10:Eu2+ has high internal QE, while its external QE can be improved further; this can be accomplished by optimizing the synthesis process and the post-synthesis treatments. The effects of the temperature on the emission intensity (i.e., the degree of thermal quenching) and the corresponding spectra of BaSi7N10:0.1Eu2+ (x = 0.8) are shown in Figs. 5(a) and 5(b), respectively. The emission intensity decreased gradually over temperatures of 20–180 °C and remained at approximately 80% of the initial intensity, whereas that of the reference phosphor at 180 °C dropped abruptly to 65% of its initial intensity. Thus, the compound exhibited low thermal quenching, being comparable to those of other nitride phosphors. As can be seen from the Fig. 5(b), the DPWs remained almost constant, and the full width at half maximum (FWHM) of the emission band broadened slightly, by approximately 3.56 nm, when the temperature was increased from 20 °C to 180 °C. These results indicate that BaSi7N10:Eu2+ exhibits highly stable chromaticity and brightness even at the high temperatures. On the other hand, earlier papers reported that the increase in the temperature resulted in a slight blue-shift of the DPWs [1517].

 figure: Fig. 5

Fig. 5 (a) Thermal quenching and (b) spectral variation of BaSi7N10:0.1Eu2+ (x = 0.8) powders.

Download Full Size | PDF

The effects of the substitution of Ca2+ ions for Ba2+ ions on the luminescence were also investigated. The variations in the emission intensity and DPWs of (Ba,Ca)Si7N10:0.1Eu2+ powders, which were prepared from a starting mixture consisting of (0.7–y)BaCO3yCaCO3–7/3Si3N4–0.05Eu2O3, are shown in Figs. 6(a) and 6(b), respectively. The XRD patterns of the powders confirmed that no secondary phases were produced for y ≤ 0.2. With an increase in the Ca content, that is, for an increase in y from 0 to 0.1, the emission intensity increased by approximately 130%, and then sharply decreased at y = 0.2. The DPW shifted continuously to a higher value (i.e., it underwent a red-shift) for y ≤ 0.1, but a sudden blue shift was observed at y = 0.2. The Commission International de L'Eclairage (CIE) chromaticity coordinates of the powders are shown in the inset of Fig. 6(b). They were all in the greenish-blue region (cyan) and varied with the amount of Ca2+, which led to changes in the DPW.

 figure: Fig. 6

Fig. 6 Variation in (a) the emission intensity and (b) DPWs of (Ba,Ca)Si7N10:0.1Eu2+ powders with Ca (y). (Inset: CIE chromaticity coordinates).

Download Full Size | PDF

These phenomena can be explained by the changes induced in the crystal parameters of the compounds. The radius of the Ca2+ ion (r = 1.34 Å for CN = 12) is much smaller than that of the Ba2+ ion (r = 1.61 Å for CN = 12). In addition, both ions exhibit different chemical properties (i.e., they have different electronegativities and ionic bond energies). These differences result in structural modification when Ca2+ ions replace Ba2+ ions. The crystal parameters of (Ba,Ca)Si7N10:0.1Eu2+ are shown in Fig. 7. When compared to the refined parameters of BaSi7N10:0.1Eu2+, the changes of the lattice constants of (Ba,Ca)Si7N10:0.1Eu2+ for y = 0.05 were insignificant, β was lower, and the unit cell volume was higher. Further, for y = 0.05 – 0.20, a, b, and the unit cell volume decreased continuously; however, c and β increased abruptly at y = 0.20. These finding demonstrate that the substitution of Ca2+ ions for Ba2+ ions causes local lattice distortion. As a result, the crystal field surrounding the Eu2+ ions is the optimal one, with the strongest emission intensity being observed at y = 0.1, as shown in Fig. 6(a). The relationship between the crystal field strength, Dq, and ionic size is Dq ∝ R−5, where R is the length of the bond between the central ion (Eu2+) and a ligand ion (N3-). Accordingly, the substitution of the smaller Ca2+ ions for Ba2+ ions increases the strength of the crystal field, resulting in the red-shift observed for y = 0 – 0.1, as shown in Fig. 6(b). However, the unexpected blue-shift at y = 0.2 might be attributable to the marked increase in c and β (see Fig. 7).

 figure: Fig. 7

Fig. 7 Crystal parameters of the (Ba,Ca)Si7N10:0.1Eu2+ powders prepared with different amounts of Ca (y). (a) Lattice constants a and b, (b) lattice constant c, (c) axial angle β, and (d) unit cell volume.

Download Full Size | PDF

The decay time of Io/e (Io: initial emission intensity at 481 nm) of (Ba,Ca)Si7N10:0.1Eu2+ (x = 0.8, y = 0.1) was approximately 0.7 μs, as can be seen from Fig. 8. This value is in the range that is typical for the 5d → 4f energy transition of the Eu2+ ion in nitride phosphors, while the single exponential curve suggests the presence of a single activation center (see Fig. 3). Earlier studies report significantly longer decay times [15,16], but do not describe the underlying mechanism. For example, Qin et al. have suggested that barium vacancies (VBa”) and positive defects are created during the synthesis process owing to local oxidation and that these act as hole- and electron-trapping centers, respectively, leading to longer-lasting phosphorescence [16]. Consequently, it can be speculated that the differences in the decay time and the PLE spectra obtained in the present study and those reported in previous works [1416] are probably ascribable to the differences in the crystal compositions, defects, and nitrogen contents of the synthesized compounds.

 figure: Fig. 8

Fig. 8 Decay curve of (Ba,Ca)Si7N10:0.1Eu2+ (x = 0.8, y = 0.1).

Download Full Size | PDF

The PL emission spectra of the (Ba,Ca)Si7N10:zEu2+ (x = 0.8, y = 0.1) samples with different Eu2+ concentrations (z) are shown in Fig. 9. The emission intensity increases continuously for z ≤ 0.15 and then drops sharply at z = 0.25, owing to a concentration-quenching effect. The change in the DPW with an increase in z was significant, because the large Stokes shift (~6,340 cm−1) does not allow the shift of the emission band, which is attributable to the re-absorption mechanism caused by overlapping absorption and emission bands [15]. On the other hand, at z = 0.25, a weak red band was observed at approximately 650 nm, in addition to the main emission band. The results of the XRD analyses indicated that no impurity phase was present in the compound corresponding to z = 0.25. Nevertheless, the most probable reason for the red emission is the presence of a Ba2Si5N8:Eu2+ phase, which might exist in amounts smaller than the detection limit of the XRD analysis system. A similar phenomenon has been reported by Li et al., who found that Ba2Si5N8:Eu2+ as well as BaSi7N10:Eu2+ appeared as an impurity phase at high Eu2+ concentration (20 at%) and that this phase resulted in orange emission [15].

 figure: Fig. 9

Fig. 9 PL emission spectra of the (Ba,Ca)Si7N10:zEu2+ powders (x = 0.8, y = 0.1) prepared with different Eu2+ concentrations (z).

Download Full Size | PDF

4. Summary

(Ba,Ca)Si7N10:Eu2+ powders were synthesized from mixtures consisting of (x-y-z)BaCO3yCaCO3–7/3Si3N4–0.5zEu2O3 using a high-temperature graphite furnace at normal nitrogen atmosphere. Rietveld refinement confirmed that stoichiometric and pure BaSi7N10:Eu2+ powders were obtained at x = 0.8. The excitation spectra of the powders were composed of three sub-bands, whose intensities varied with the BaCO3 content; the strongest 368 nm excitation band was observed at x = 0.8, resulting in the strongest 478 nm emission. There were distinct differences in the PLE spectra observed in the present study and those reported previously. These differences are probably attributable to the differences in the compositions, crystal defects, and nitrogen (oxygen) contents of the synthesized compounds. The substitution of Ca2+ ion for Ba2+ ions increased the intensity of the emission by approximately 130%, owing to local lattice modification. Finally, the powders exhibited low thermal quenching, high QEs, and stable chromaticities, indicating that (Ba,Ca)Si7N10:Eu2+ phosphors show promise for use in white LEDs pumped by nUV sources.

Acknowledgment

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MEST) (2011-0013554). Thanks to Dr. S. –H. Jeong at KBSI Gwangju-center for the TRPL measurement.

References and links

1. H. A. Höppe, H. Lutz, P. Morys, W. Schnick, and A. Seilmeier, “Luminescence in Eu2+-doped Ba2Si5N8: fuorescence, thermoluminescence, and upconversion,” J. Phys. Chem. Solids 61(12), 2001–2006 (2000). [CrossRef]  

2. Y. Q. Li, J. E. J. van Steen, J. W. H. van Krevel, G. Botty, A. C. A. Delsing, F. J. DiSalvo, G. de With, and H. T. Hintzen, “Luminescence properties of red-emitting M2Si5N8:Eu2+ (M = Ca, Sr, Ba) LED conversion phosphors,” J. Alloy. Comp. 417(1-2), 273–279 (2006). [CrossRef]  

3. B. Lei, K. Machida, T. Horikawa, H. Hanzawa, N. Kijima, Y. Shimomura, and H. Yamamoto, “Reddish-orange long-lasting phosphorescence of Ca2Si5N8:Eu2+,Tm3+ phosphor,” J. Electrochem. Soc. 157(6), J196–J201 (2010). [CrossRef]  

4. J. Li, B. Lei, J. Qin, Y. Liu, and X. Liu, “Temperature-dependent emission spectra of Ca2Si5N8:Eu2+, Tm3+ phosphor and its afterglow properties,” J. Am. Ceram. Soc. 96(3), 873–878 (2013). [CrossRef]  

5. V. D. Luong, W. Zhang, and H.-R. Lee, “Preparation of Sr2Si5N8:Eu2+ for white light-emitting diodes by multi-step heat treatment,” J. Alloy. Comp. 509(27), 7525–7528 (2011). [CrossRef]  

6. G. Kirakosyan and D. Y. Jeon, “Low-temperature synthesis Sr2Si5N8:Eu2+ red-emitting phosphor by modified solid-state metathesis approach and its photoluminescent characteristics,” J. Electrochem. Soc. 159(2), J29–J33 (2012). [CrossRef]  

7. J. Li, T. Watanabe, H. Wada, T. Setoyama, and M. Yoshimura, “Synthesis of Eu-doped CaAlSiN3 from ammonometallates: effects of sodium content and pressure,” J. Am. Ceram. Soc. 92(2), 344–349 (2009). [CrossRef]  

8. X. Piao, K. Machida, T. Horikawa, H. Hanzawa, Y. Shimomura, and N. Kijima, “Preparation of CaAlSiN3:Eu2+ phosphors by the self-propagating high-temperature synthesis and their luminescent properties,” Chem. Mater. 19(18), 4592–4599 (2007). [CrossRef]  

9. Y. W. Jung, B. Lee, S. P. Singh, and K. S. Sohn, “Particle-swarm-optimization-assisted rate equation modeling of the two-peak emission behavior of non-stoichiometric CaAl(x)Si(7-3x)/4N3:Eu2+ phosphors,” Opt. Express 18(17), 17805–17818 (2010). [CrossRef]   [PubMed]  

10. N. Hirosaki, R.-J. Xie, K. Kimoto, T. Sekiguchi, Y. Yamamoto, T. Suehiro, and M. Mitomo, “Characterization and properties of green-emitting β-SiAlON:Eu2+ powder phosphors for white light-emitting diodes,” Appl. Phys. Lett. 86(21), 211905 (2005). [CrossRef]  

11. V. Bachmann, C. Ronda, O. Oeckler, W. Schnick, and A. Meijerink, “Color point tuning for (Sr,Ca,Ba)Si2O2N2:Eu2+ for white light LEDs,” Chem. Mater. 21(2), 316–325 (2009). [CrossRef]  

12. J. Park, S. J. Lee, and Y. J. Kim, “Evolution of luminescence of Sr2-y-zCazSi(O1-xNx)4:yEu2+ with N3-, Eu2+, and Ca2+ substitutions,” J. Cryst. Growth Des. 13(12), 5204–5210 (2013). [CrossRef]  

13. S. J. Lee, S.-H. Hong, and Y. J. Kim, “Synthesis and luminescent properties of (Sr,M)2Si(O1-xNx)4:Eu2+(M: Mg2+, Ca2+, Ba2+),” J. Electrochem. Soc. 159(5), J163–J167 (2012). [CrossRef]  

14. Y. Q. Li, A. C. A. Delsing, R. Metslaar, G. de With, and H. T. Hintzen, “Photoluminescence properties of rare-earth activated BaSi7N10,” J. Alloy. Comp. 487(1-2), 28–33 (2009). [CrossRef]  

15. H. L. Li, R. J. Xie, G. H. Zhou, N. Hirosaki, and Z. Sun, “A cyan-emitting BaSi7N10 : Eu2 + phosphor prepared by gas reduction and nitridation for UV-Pumping White LEDs,” J. Electrochem. Soc. 157(7), J251–J255 (2010). [CrossRef]  

16. J. Qin, H. Zhang, B. Lei, H. Dong, Y. Liu, H. Meng, M. Zheng, and Y. Xiao, “Preparation and afterglow properties of highly condensed nitridosilicate BaSi7N10:Eu2+ phosphor,” J. Lumin. 152, 230–233 (2014). [CrossRef]  

17. G. Anoop, D. W. Lee, D. W. Suh, S. L. Wu, K. M. Ok, and J. S. Yoo, “Solid-state synthesis, structure, second-harmonic generation, and luminescent properties of noncentrosymmetric BaSi7N10:Eu2+ phosphors,” J. Mater. Chem. C 1(31), 4705–4712 (2013). [CrossRef]  

18. C. M. Fang, H. T. Hintzen, and G. de With, “First-principles electronic structure calculations of BaSi7N10 with both corner- and edge-sharing SiN4 tetrahedra,” J. Alloy. Comp. 336(1-2), 1–4 (2002). [CrossRef]  

19. H. Huppertz and W. Schnick, “Edge-sharing SiN4 tetrahedra in the highly condensed nitridosilicate BaSi7N10,” Chemistry 3(2), 249–252 (1997). [CrossRef]   [PubMed]  

20. G. Pilet, H. A. Höppe, W. Schnick, and S. Esmaeilzadeh, “Crystal structure and mechanical properties of SrSi7N10,” Solid State Sci. 7(4), 391–396 (2005). [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 (9)

Fig. 1
Fig. 1 (a) XRD patterns of the samples prepared with (x-0.1)BaCO3, (b) experimental (red dots) and calculated (blue solid line) XRD profiles of Rietveld refinement for BaSi7N10:0.1Eu2+ powders corresponding to x = 0.8, and (c) a difference between the experimental and calculated data. Blue bars represent Bragg reflections.
Fig. 2
Fig. 2 Variation in the Ba/Si atomic ratios (RBa/Si) of the samples prepared with (x-0.1)BaCO3.
Fig. 3
Fig. 3 (a) PLE and (b) emission spectra of the samples prepared with (x-0.1)BaCO3.
Fig. 4
Fig. 4 Variation in the IR values of the samples prepared with (x-0.1)BaCO3.
Fig. 5
Fig. 5 (a) Thermal quenching and (b) spectral variation of BaSi7N10:0.1Eu2+ (x = 0.8) powders.
Fig. 6
Fig. 6 Variation in (a) the emission intensity and (b) DPWs of (Ba,Ca)Si7N10:0.1Eu2+ powders with Ca (y). (Inset: CIE chromaticity coordinates).
Fig. 7
Fig. 7 Crystal parameters of the (Ba,Ca)Si7N10:0.1Eu2+ powders prepared with different amounts of Ca (y). (a) Lattice constants a and b, (b) lattice constant c, (c) axial angle β, and (d) unit cell volume.
Fig. 8
Fig. 8 Decay curve of (Ba,Ca)Si7N10:0.1Eu2+ (x = 0.8, y = 0.1).
Fig. 9
Fig. 9 PL emission spectra of the (Ba,Ca)Si7N10:zEu2+ powders (x = 0.8, y = 0.1) prepared with different Eu2+ concentrations (z).

Tables (2)

Tables Icon

Table 1 Crystallographic data for BaSi7N10 and BaSi7N10:0.1Eu2+ (x = 0.8).

Tables Icon

Table 2 Interatomic distances (Å) of Eu-N for Ba0.9Eu0.1Si7N10.

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.