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

High efficient blue emission of Ce3+ activated Ca4P2O9 phosphor for white LEDs

Open Access Open Access

Abstract

Intense blue emission from Ca4P2O9:Ce3+ phosphor was observed with CIE chromaticity coordinates (0.157, 0.094) and a high external quantum yield up to 64.8%. Concentration quenching effect of Ca4P2O9:xCe3+ indicated that the optimal Ce3+ concentration was at x = 0.005, and the energy transfer among the nearest neighbor ions accounted for this quenching. Good thermal stability of the optimal phosphor was also shown, since 90% of the initial intensity (at 300 K) remained as the temperature increased from 300 to 400 K. Ca4P2O9:0.005Ce3+ can act as a promising blue-emitting phosphor for white LEDs.

© 2013 Optical Society of America

1. Introduction

White LED (Light Emitting Diode) is a promising strategy for energy conservation in solid-state lighting due to the advantages such as high efficiency, long lifetimes, low power consumptions, and mercury-free design. In general, white LEDs are generated by not only combining the light of several colored LEDs, but also comprising a blue or ultraviolet LED with phosphors (called phosphor-converted white LEDs).

The first demonstration of phosphor-converted white LEDs was achieved by combination of a blue LED chip and a yellow phosphor YAG:Ce (Y3Al5O12:Ce3+) [1]. However, the application of YAG:Ce-converted white LEDs faces some disadvantages such as low color rendering index (CRI), saturation of luminous efficacy at high drive levels [2], and color drift by different drive conditions [3]. These problems may be solved by comprising one ultraviolet LED and two phosphors, because of their high luminous flux and color stability under high operating current [2, 3]. For example, Narukawa et al. have reported a white LED, which is combined near-ultraviolet LED, blue phosphors, and yellow phosphors, exhibits better performances of color temperature, color rendering, and luminous efficacy, in comparison with that comprised by a blue LED and a yellow phosphor [2]. Phosphor-converted white LEDs has also been achieved by using an ultraviolet LED with tricolor (red, green, and blue) phosphors [4]. Based on this idea, high CRI (>95) has been obtained by mixing multichromatic phosphors with near ultraviolet LEDs [5]. Under this condition, high efficient blue emitting phosphors are essential for white LEDs with high CRI.

BaMgAl10O17:Eu2+ is a traditional efficient blue phosphor for fluorescent lamps and also can act as a blue component in white LEDs. The internal quantum efficiency is over 90%, and the external quantum efficiency is close to 70% [6]. However, BaMgAl10O17:Eu2+ degrades more rapidly due to the oxidation of Eu2+ from heat treatment in air during lamp processing [7], and the concentration of Eu2+ is much high (10 mol %), which raises the cost.

Recently, there are some reports of new efficient blue phosphors for white LEDs such as Ca2PO4Cl:Eu2+ [8], NaSrBO3:Ce3+ [6], and Sr5(PO4)2SiO4:Ce3+ [9]. Lately, Ca4P2O9:Eu2+ has been reported as a red-emitting phosphor for solid-state lighting [10]. Generally, Ce3+ has a relatively smaller crystal field splitting in d orbital level than Eu2+ in the same host matrix, and Ce3+ activated phosphors usually exhibit high efficiency such as the commercial yellow phosphor YAG:Ce. Thus, efficient green or blue emitting may be got from Ce3+ doped Ca4P2O9.

In this study, luminescence properties of Ca4P2O9:Ce3+ and (Ca,Sr)4P2O9:Ce3+ were investigated. Ca4P2O9:Ce3+ was found to exhibit good thermal stability and high efficiency, suggesting that it can be as a candidate blue phosphor for white LEDs.

2. Experimental

All specimens were prepared by a solid-state reaction method using CaCO3 (AR), SrCO3 (AR), NH4H2PO4 (AR), and CeO2 (4N) as starting materials. Stoichiometric amounts of the raw materials were weighed and mixed in an agate mortar. The mixed powders were first sintered at 900 °C for 3 h in air to decompose the carbonate. After a second homogenization in the mortar the samples were heated up to 1400 °C and kept at this temperature for 6 h in reductive atmosphere (5 vol% H2 and 95 vol% N2).

The crystalline phase of products were identified using X-ray diffraction (XRD; Ultima III, Rigaku) with Cu Kα radiation (λ = 0.154 nm). The photoluminescence spectra were recorded with VARIAN Cary Eclipse. Photoluminescence quantum yield was measured using the integrating sphere approach on a Horiba Jobin Yvon FM-4P-TCSPC instrument. During the measurement of quantum yield, a Xe lamp was used and the excitation wavelength was centered at 340 nm with a half band width of 5 nm. The emission spectrum was from 350 to 600 nm. The integral excitation intensity and the integral emission intensity were measured by the integrating sphere. The quantum yield of the phosphor was determined by the ratio of the number of photons in the emission spectrum to that in the excitation spectrum.

3. Results and discussion

Ca4P2O9 is crystallized in a monoclinic unit cell (space group P21) with lattice parameters of a = 7.018 Å, b = 11.98 Å, c = 9.469 Å, and β = 90.88°. All the Ce-doped Ca4P2O9 samples have similar XRD pattern to pure phase Ca4P2O9 (JCPDS No. 25-1137), and no impurity phase is observed, as shown in Fig. 1(a). Ca2+ ions in Ca4P2O9 can be divided into two types, one eight-coordinated Ca2+ and seven seven-coordinated Ca2+ [10, 11]. The phase of (Ca,Sr)4P2O9:Ce3+ binary system were investigated, and their XRD patterns of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-1) are displayed in Fig. 1(b). The samples remain the monoclinic Ca4P2O9 phase for y = 0-0.1 samples with XRD peaks shifted to lower angle, since the ionic radius of Sr2+ ion is larger than that of Ca2+ ion. The XRD peaks around 28° and 32° for the samples (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0.2-0.3) shift change significantly, accompanying with the disappearance of some diffraction peaks ascribed to Ca4P2O9 phase. When y equals 0.4, diffraction peaks of Sr5(PO4)3OH emerge obviously. In the case of the samples with y = 0.5, 0.75 and 1, the diffraction peaks are mainly ascribed to Sr5(PO4)3OH, and two peaks (marked with asterisk) of impurity SrO exist in the sample with y = 1. For the samples with y = 0.4 and 0.5, some peaks attributed to solid solution phase remain, especially for the sample with y = 0.4. It can be concluded that pure Sr4P2O9 is hardly obtained under the existence of water vapor which is from the environment and the reaction products [12].

 figure: Fig. 1

Fig. 1 (a) XRD patterns of Ca4P2O9:xCe3+ (x = 0-0.01) compared with the indexed JCPDS No. 25-1137 for Ca4P2O9; (b) XRD patterns of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-1) compared with the indexed JCPDS No. 25-1137 (Ca4P2O9), No. 20-1208 (Sr4P2O9) and No. 33-1348 (Sr5(PO4)3OH).

Download Full Size | PDF

Figure 2(a) displays the photoluminescence excitation and emission spectra of the Ca4P2O9: Ce3+ phosphor at room temperature. A strong excitation band and a relatively weak one from Ce3+ excitation appeared at about 340 and 270 nm, respectively. The 5d orbital is strongly affected by the crystal field of the host matrix and splits into several sub-bands. Thus, these two excitation bands at about 340 and 270 nm are caused by the transitions from the ground state to different 5d splitting state of Ce3+. The strong excitation band of ca. 340 nm indicates that this phosphor can be pumped effectively by UV-LED. The broad emission spectrum from 360 to 520 nm shows non-Gaussian symmetry. As mentioned above, there exist two types of Ca2+ ions, eight-coordinated Ca2+ and seven-coordinated Ca2+, and Ce3+ ions are expected to randomly occupy Ca2+ ion sites [10]. Different substitution sites of Ce3+ may lead to different emission band (shape or center position), furthermore, the emission peaks are due to the electronic transitions from the lowest 5d level to two different 4f ground state levels. Thus, coupling the possible effects of different substitution sites and different transition levels, a non-Gaussian symmetric emission band is observed.

 figure: Fig. 2

Fig. 2 (a) Photoluminescence excitation (monitored at 420 nm) and emission (excited at 340 nm) spectra of Ca4P2O9:0.005Ce3+; inset shows the emission intensity (I) as a function of the Ce3+ concentration (x mol %); (b) the dependence of lg (I/x) on lg (x) according to Eq. (2).

Download Full Size | PDF

The photoluminescence emission intensity of Ca4P2O9:xCe3+ depends on the Ce3+ concentration, as shown in the inset of Fig. 2(a). The emission intensity (excited at 340 nm) increases with the Ce3+ doping concentration until a maximum reaches at x = 0.005. As the Ce3+ concentration further increases, the emission intensity decreases due to concentration quenching process. In general, the concentration quenching originates from the energy migration among the activator ions. The excitation energy will be lost at a killer center in the energy migration process, resulting in luminescence quenching. The critical distance (Rc), pointed out by Blasse [13], is defined as the distance for which the probability of energy transfer equals the probability of radiative emission of the activator ions. According to Eq. (1),

Rc=2[3V4πxcZ]13
where V is the volume of the unit cell; xc is the critical concentration of Ce3+; and Z is the number of formula units in the unit cell. Rc is calculated to be 42.1 Å, using V = 779.1 Å3, Z = 4, xc = 0.005. Many interactions account for the energy transfer such as exchange interaction, radiation reabsorption, or electric multipolar interaction. In this case, the exchange interaction plays no role for the energy migration between Ce3+ ions in Ca4P2O9:xCe3+ phosphor, since exchange interaction is generally dominant only for short distance (typical critical distance of 5 Å) in a forbidden transition [10, 14]. And the mechanism of radiation reabsorption is only effective when the fluorescence spectra are broadly overlapping [15]. Thus, radiation reabsorption does not occur either. Consequently, process of energy transfer is due to the electric multipolar interaction including dipole–dipole, dipole–quadrupole and quadrupole–quadrupole, respectively [16]. The strength of the multipole-multipole interactions can be determined from the change in the emission intensity if the energy transfer occurs between the same sorts of activators [10]. The emission intensity per activator ion can be expressed by the following Eq. (2) [17]:
Ix=k[1+β(x)θ3]1,
where I is the emission intensity; x is the activator concentration; k and β are constants for the same interaction for a given host crystal; θ = 3 represents for the energy transfer among nearest-neighbor ions [18], while θ = 6, 8, and 10 are for dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole interactions, respectively [17, 18]. The relationship of lg (I/x) versus lg (x), shown in Fig. 2(b), is linear and the slope is −0.93. The value of θ is found to be 2.79 using the above Eq. (2), which is approximately equal to 3. This result indicates that the energy transfer among the nearest neighbor ions is the main mechanism of concentration quenching of Ca4P2O9:xCe3+.

Excitation and normalized emission spectra of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-0.4) are shown in Fig. 3. The excitation spectra in Fig. 3(a) are composed of two excitation bands which are similar to the excitation spectrum in Fig. 2(a), accompanying with the decrease of the excitation intensity with Sr content (y). The only difference is that the weak excitation band position at short wavelength has changed when y exceeds 0.2, which is consistent with the structure change in Fig. 1(b). The emission spectra shape and intensity (inset in Fig. 3(b)) are almost the same for the samples with y = 0-0.1. As y further increases to 0.2, the emission peak is broadened, accompanying with dramatic decrease of the intensity (inset in Fig. 3(b)), but almost the same shape has been observed for the samples with y = 0.2-0.4. It is noted that the change of the emission shape and intensity of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-0.4) depends on the change of (Ca1-ySry)4P2O9:0.005Ce3+ structure shown in Fig. 1(b). The compound Sr4P2O9 was reported by Bauer and Balz, who found it to be isostructural with the analogous calcium compound Ca4P2O9 [12]. The samples with low concentrations of Sr remain the monoclinic structure analogous to Ca4P2O9, which leads to almost the same emission spectra for y = 0-0.1 (Fig. 3). As the Sr concentration increases to some value (y>0.2 in this case), obvious change in the structure of (Ca1-ySry)4P2O9 has been observed, especially for y>0.4 (Sr5(PO4)3OH dominates). Thus, the emission spectra of (Ca1-ySry)4P2O9:0.005Ce3+ are broadened. However, when y exceeds 0.4, Sr5(PO4)3OH has formed, and the emission can hardly be observed.

 figure: Fig. 3

Fig. 3 (a) Excitation spectra of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-0.4); (b) The normalized emission spectra of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-0.4) under the excitation of 340 nm, the inset shows the emission intensity versus y.

Download Full Size | PDF

It is to our interest that why the emission intensity of (Ca1-ySry)4P2O9:0.005Ce3+ for y = 0.2-0.4 decreases dramatically compared with those for y = 0-0.1. Two main possible mechanisms may be ascribed to this decrease. One is the existence of OH group in the lattice, quenching the excitation energy; the other is the difference of the activation energy for the thermal quenching of Ce3+ which is due to the structure difference. Hydroxyl-quenching influences obviously the luminescence of rare earth ions [19]. Rare earth ions coupled to OH groups can be as quenchers [20]. The more quenchers are, the more possible excitation energy quenches through energy transfer from the activators to quenchers. On the other side, the change of the matrix structure has great effect on the 5d orbitals of Ce3+. And difference of the activation energy for the thermal quenching of Ce3+ may be caused by structure change. Thus, the decrease of luminescence with the increase in the Sr concentration may due to large thermal quenching at room temperature. To investigate this, temperature quenching experiment of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0 and 0.25) was performed from 77 to 400 K (Fig. 4). The temperature dependence of the luminescence can be described by the Arrhenius equation as follows:

IT=I01+cexp(Ea/kT),
where I0 is the initial intensity at 77 K; IT is the intensity at a given temperature T; c is a constant; Ea is the activation energy for thermal quenching; and k is Boltzmann’s constant.

 figure: Fig. 4

Fig. 4 Temperature dependent emission intensity of (Ca1-ySry)4P2O9:0.005Ce3+ for y = 0 and 0.25 respectively.

Download Full Size | PDF

The obtained Ea by fitting the experiment data for the sample with y = 0 and 0.25 is 0.033 and 0.027 eV, respectively. The activation energy of Ca4P2O9:0.005Ce3+ is very close to that of (Ca0.75Sr0.25)4P2O9:0.005Ce3+. Hence, the decrease of the luminescence of (Ca1-ySry)4P2O9:0.005Ce3+ with higher Sr concentration (y>0.2) is not ascribed to the difference of the activation energy for the thermal quenching. Therefore, the existence of OH groups account for the decrease in the luminescence (inset in Fig. 3(b)).

Small thermal quenching is a critical factor for phosphors in the character of chromaticity and brightness in lighting at high temperature. Thermal quenching of the luminescent property of Ca4P2O9:0.005Ce3+ from room temperature to 450 K has been investigated as shown in Fig. 5. No obvious shift is observed in the emission band, suggesting stable chromaticity coordinates by varying the temperature. The relative integrated emission intensity of Ca4P2O9:0.005Ce3+ decreases marginally with the increase of the temperature, and remains 90% of the original intensity (at 300 K) when at 400 K. As the temperature increases further, the relative integrated emission intensity begins to drop obviously. Stable chromaticity property and small decrease of the intensity with temperature indicate good thermal stability of Ca4P2O9:0.005Ce3+ phosphor.

 figure: Fig. 5

Fig. 5 Temperature dependence of the integrated emission intensity normalized with respect to the value at 300 K.

Download Full Size | PDF

Bright blue emission of Ca4P2O9:0.005Ce3+ is shown to our naked eyes with CIE chromaticity coordinates (0.157, 0.094). And the chromaticity can be tuned by forming the solid solution of Ca4P2O9 and Sr4P2O9. Cyanine color is observed from the emission of (Ca0.75Sr0.25)4P2O9:0.005Ce3+ with CIE chromaticity coordinates (0.164, 0.168). The colors of these two phosphors are displayed in Fig. 6. The optimal phosphor has a high external quantum yield up to 64.8% (excited at 340 nm), which is very close to that of commercial product BaMgAl10O17:Eu2+ (69.6%) (excited at 365 nm) [6]. However, the activation center concentration in Ca4P2O9:Ce3+ (0.5 mol %) in this study is much lower than that of BaMgAl10O17:Eu2+ (11.6 mol %) [21]. Ca4P2O9:Ce3+ can act as an efficient promising blue phosphor for white LEDs.

 figure: Fig. 6

Fig. 6 CIE chromaticity coordinates of Ca4P2O9:0.005Ce3+ and (Ca0.75Sr0.25)4P2O9:0.005Ce3+ phosphors.

Download Full Size | PDF

4. Conclusion

Ca4P2O9:Ce3+, as a blue-emitting phosphor, was synthesized by a solid-state reaction. A non-Gaussian symmetric broad emission band centered at 420 nm with CIE chromaticity coordinates (0.157, 0.094) was observed. Concentration quenching effect of Ca4P2O9:xCe3+ showed that the optimal concentration is at x = 0.005. The quenching mechanism is due to the energy transfer among the nearest neighbor ions. The chromaticity can be tuned by substituting part of Sr for Ca, changing the structure of this system. And Cyanine color was obtained with CIE chromaticity coordinates (0.164, 0.168) from the phosphor (Ca0.75Sr0.25)4P2O9:0.005Ce3+. Ca4P2O9:Ce3+ phosphor displayed a good thermal stability, and 90% of the initial intensity at room temperature (300 K) remained as the temperature increased to 400 K. A high external quantum yield, 64.8%, was also performed in this phosphor, which indicated that Ca4P2O9:Ce3+ can be as a promising candidate blue-emitting phosphor for white LEDs.

Acknowledgments

This work is supported by National Basic Research Program of China (973 Program, 2013CB632404, 2011CB933303), a Project Funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions, NCET-12-0268, BK2011056, and NSFC (Nos. 11174129 and 51272101). The author (Z. Z) also thanks Jiangsu Province Innovation Platform for Superiority Subject of Environmental Science and Engineering.

References and links

1. S. Nakamura and G. Fasol, The Blue Laser Diode: GaN Based Light Emitters and Lasers (Springer, 1997).

2. Y. Narukawa, I. Niki, K. Izuno, M. Yamada, Y. Murazaki, and T. Mukai, “Phosphor-conversion white light emitting diode using InGaN near-ultraviolet chip,” Jpn. J. Appl. Phys. 41(Part 2, No. 4A), L371–L373 (2002). [CrossRef]  

3. Y. Xu, L. Chen, Y. Li, G. Song, Y. Wang, W. Zhuang, and Z. Long, “Phosphor-conversion white light using InGaN ultraviolet laser diode,” Appl. Phys. Lett. 92(2), 021129 (2008). [CrossRef]  

4. C. H. Kuo, J. K. Sheu, S. J. Chang, Y. K. Su, L. W. Wu, J. M. Tsai, C. Liu, and R. Wu, “n-UV + blue/green/red white light emitting diode lamps,” Jpn. J. Appl. Phys. 42(Part 1, No. 4B), 2284–2287 (2003). [CrossRef]  

5. K. Takahashi, N. Hirosaki, R. J. Xie, M. Harada, K. I. Yoshimura, and Y. Tomomura, “Luminescence properties of blue La1-x CexAl(Si6-zAlz)(N10-zOz)(z~1) oxynitride phosphors and their application in white light-emitting diode,” Appl. Phys. Lett. 91, 091923 (2007). [CrossRef]  

6. W. R. Liu, C. H. Huang, C. P. Wu, Y. C. Chiu, Y. T. Yeh, and T. M. Chen, “High efficiency and high color purity blue-emitting NaSrBO3:Ce3+ phosphor for near-UV light-emitting diodes,” J. Mater. Chem. 21(19), 6869–6874 (2011). [CrossRef]  

7. S. Oshio, T. Matsuoka, S. Tanaka, and H. Kobayashi, “Mechanism of luminance decrease in BaMgAl10O17:Eu2+ phosphor by oxidation,” J. Electrochem. Soc. 145(11), 3903–3907 (1998). [CrossRef]  

8. Y. C. Chiu, W. R. Liu, C. K. Chang, C. C. Liao, Y. T. Yeh, S. M. Jang, and T. M. Chen, “Ca2PO4Cl:Eu2+: an intense near-ultraviolet converting blue phosphor for white light-emitting diodes,” J. Mater. Chem. 20(9), 1755–1758 (2010). [CrossRef]  

9. Z. Tao, Y. Huang, and H. J. Seo, “Blue luminescence and structural properties of Ce3+-activated phosphosilicate apatite Sr5(PO4)2(SiO4),” Dalton Trans. 42(6), 2121–2129 (2013). [CrossRef]   [PubMed]  

10. D. Deng, H. Yu, Y. Li, Y. Hua, G. Jia, S. Zhao, H. Wang, L. Huang, Y. Li, C. Li, and S. Xu, “Ca4(PO4)2O:Eu2+ red-emitting phosphor for solid-state lighting: structure, luminescent properties and white light emitting diode application,” J. Mater. Chem. C 1(19), 3194–3199 (2013). [CrossRef]  

11. B. Dickens, W. Brown, G. Kruger, and J. Stewart, “Ca4(PO4)2O, tetracalcium diphosphate monoxide. Crystal structure and relationships to Ca5 (PO4) 3OH and K3Na(SO4)2,” Acta Crystallogr. B 29(10), 2046–2056 (1973). [CrossRef]  

12. E. R. Kreidler and F. A. Hummel, “Phase relations in the system SrO-P2O5 and the influence of water vapor on the formation of Sr4P2O9,” Inorg. Chem. 6(5), 884–891 (1967). [CrossRef]  

13. G. Blasse, “Energy transfer between inequivalent Eu2+ ions,” J. Solid State Chem. 62(2), 207–211 (1986). [CrossRef]  

14. Z. Xia, R.-S. Liu, K.-W. Huang, and V. Drozd, “Ca2Al3O6F:Eu2+: a green-emitting oxyfluoride phosphor for white light-emitting diodes,” J. Mater. Chem. 22(30), 15183–15189 (2012). [CrossRef]  

15. Y. C. Chiu, W. R. Liu, C. H. Huang, Y. T. Yeh, and S. M. Jang, “Luminescence properties, crystal structure and LED package of potential blue-emitting phosphors--Ca2BN2F:Eu2+.,” Opt. Express 20(25), 27361–27366 (2012). [CrossRef]   [PubMed]  

16. D. L. Dexter, “A theory of sensitized luminescence in solids,” J. Chem. Phys. 21(5), 836–850 (1953). [CrossRef]  

17. L. Ozawa and P. M. Jaffe, “The mechanism of the emission color shift with activator concentration in Eu+3 activated phosphors,” J. Electrochem. Soc. 118(10), 1678–1679 (1971). [CrossRef]  

18. L. Van Uitert, “Characterization of energy transfer interactions between rare earth ions,” J. Electrochem. Soc. 114(10), 1048–1053 (1967). [CrossRef]  

19. T. Moon, S. T. Hwang, D. R. Jung, D. Son, C. Kim, J. Kim, M. Kang, and B. Park, “Hydroxyl-quenching effects on the photoluminescence properties of SnO2:Eu3+ nanoparticles,” J. Phys. Chem. C 111(11), 4164–4167 (2007). [CrossRef]  

20. A. J. Faber, D. R. Simons, Y. Yan, and H. de Waal, “Photoluminescence quenching by OH in Er- and Pr-doped glasses for 1.5 and 1.3 μm optical amplifiers,” Proc. SPIE 2290, 80–88 (1994). [CrossRef]  

21. K. B. Kim, Y. I. Kim, H. G. Chun, T. Y. Cho, J. S. Jung, and J. G. Kang, “Structural and optical properties of BaMgAl10O17:Eu2+ phosphor,” Chem. Mater. 14(12), 5045–5052 (2002). [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 (6)

Fig. 1
Fig. 1 (a) XRD patterns of Ca4P2O9:xCe3+ (x = 0-0.01) compared with the indexed JCPDS No. 25-1137 for Ca4P2O9; (b) XRD patterns of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-1) compared with the indexed JCPDS No. 25-1137 (Ca4P2O9), No. 20-1208 (Sr4P2O9) and No. 33-1348 (Sr5(PO4)3OH).
Fig. 2
Fig. 2 (a) Photoluminescence excitation (monitored at 420 nm) and emission (excited at 340 nm) spectra of Ca4P2O9:0.005Ce3+; inset shows the emission intensity (I) as a function of the Ce3+ concentration (x mol %); (b) the dependence of lg (I/x) on lg (x) according to Eq. (2).
Fig. 3
Fig. 3 (a) Excitation spectra of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-0.4); (b) The normalized emission spectra of (Ca1-ySry)4P2O9:0.005Ce3+ (y = 0-0.4) under the excitation of 340 nm, the inset shows the emission intensity versus y.
Fig. 4
Fig. 4 Temperature dependent emission intensity of (Ca1-ySry)4P2O9:0.005Ce3+ for y = 0 and 0.25 respectively.
Fig. 5
Fig. 5 Temperature dependence of the integrated emission intensity normalized with respect to the value at 300 K.
Fig. 6
Fig. 6 CIE chromaticity coordinates of Ca4P2O9:0.005Ce3+ and (Ca0.75Sr0.25)4P2O9:0.005Ce3+ phosphors.

Equations (3)

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

R c =2 [ 3V 4π x c Z ] 1 3
I x =k [ 1+β ( x ) θ 3 ] 1 ,
I T = I 0 1+cexp( E a / kT ) ,
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.