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

Effect of the replacement of Zn2+ with Mg2+ in Ca14Zn6Ga10O35:Mn4+

Open Access Open Access

Abstract

To obtain applicable red phosphor for plant lighting, a series of Ca14Zn6-xMgxGa9.86O35:0.14Mn4+ phosphors were synthesized by a conventional high-temperature solid-state method. The XRD, SEM, excitation and emission spectra, temperature-dependent emission spectra, decay time, internal quantum efficiency and luminous efficiency are investigated with the change of the doping concentration of Mg2+. The experiment results suggest that the luminescence properties and quantum efficiency of the phosphor can by improved by replacing Zn2+ with Mg2+ at appropriate doses, while the thermal resistance performance and decay time declined, and the internal quantum efficiency and luminous efficiency reached its maximum, when the Mg2+ concentration is 0.40. Ultimately, the appropriate doping concentration of Mg2+ was determined at x = 0.40, in which case, the phosphor has a promising application in future plant lighting.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

With the improvement of human living standard and quality, more and more people have access to fruits and vegetables of all seasons at any time. The approach to achieve this availability is plant lighting. It is reported that 410–500nm blue light affects chlorophyll synthesis, leaf shape, and flower bud formation; 610–700nm red light affects germination and plant growth; and 700–740nm far-red light is indispensable for plant photosynthesis and growth [1–3]. Unfortunately, commercial white light-emitting diode (LED) phosphors lack the far-red component, which makes them unsuitable for plant lighting. Therefore, developing appropriate far-red phosphors for plant lighting represents the general trend.

To date, Eu ion (Eu2+ and Eu3+)-activated red phosphors, such as M2Si5N8:Eu2+(M = Ca, Sr, Ba), (Sr, Ca)S:Eu3+, and Y2O2S:Eu3+, have been widely investigated because of their outstanding luminescence properties and high quantum yields [4–6]. However, their emission wavelengths are below 700nm instead of in the far-red region. Recently, transition element Mn4+-activated red phosphors have attracted great attention. It is well known that Mn4+ with a 3d3 electron configuration will exhibit 620–750nm emission bands under 200–600nm excitation when in an octahedral coordination environment [7]. Mn4+-doped germanate, titanate, aluminate, and fluoride red phosphors have been reported largely because the ionic radius of Ge4+, Ti4+, Al3+, or Si4+ in the crystal structure are quite similar to that of Mn4+, and these ions can therefore be displaced by Mn4+. However, most of their emission wavelengths are beyond the far-red region [8–13].

Recently, Ca14Zn6Al10O35(CZA):Mn4+ phosphor with 650–750nm far-red emission has been reported to be suitable for plant lighting [14–17]. There are several reports about improving its luminescence properties, such as by determining the appropriate Mn4+ doping concentration, synthesis temperature, and Mg2+ doping concentration, as well as by employing the oxygen-pressure method [17–19]. Despite a crystal structure similar to CZA:Mn4+, Ca14Zn6Ga10O35(CZG):Mn4+ possesses superior luminescence properties and better thermal stability [18,20,21]. Therefore, there is great potential to improve the luminescence properties of CZG:Mn4+ by an appropriate method, which has not yet been reported.

Herein, a series of Ca14Zn6-xMgxGa9.86O35:0.14Mn4+(x = 0.00, 0.20, 0.40, 0.60, 0.80, 1.00) phosphors have been prepared by a high-temperature solid-state method to explore the effect of replacing Zn2+ with Mg2+ and therefore determine the optimal Mg2+ doping concentration .

2. Experimental

2.1. Synthesis

A series of Ca14Zn6-xMgxGa9.86O35:0.14Mn4+(CZMG:Mn4+) (x = 0.00, 0.20, 0.40, 0.60, 0.80, 1.00) phosphors doping concentrations have been prepared through a simple solid-state reaction. First, the raw materials, including CaCO3 (99.99%), ZnO (99.99%), Ga2O3 (99.99%), MgO(99.99%) and MnCO3 (99.99%), were weighed based on the stoichiometric ratio and ground in an agate mortar for 30 min. Then, the mixture was inserted in a muffle furnace with 1473K calcination temperature and 6h calcination time. Finally, the samples were removed from the muffle furnace and ground again to pulverulent form.

2.2. Characterizations

The crystal structure of the samples was determined by an X-ray diffractometer (XRD, D8 Advance, Bruker, Germany) with Cu Kα radiation at 40 kV and 40 mA. The particle morphology and size were observed with a JMF-6700F scanning electron microscope (SEM). The excitation spectra, emission spectra, quantum efficiency, and temperature-dependent emission spectra, decay curves were recorded with a FluoroMax-4 spectrophotometer. Luminous efficiency are evaluated using an integrating sphere (PMS-50; Everfine Photo-E-Info Co. Ltd, Hangzhou, China) at an operating current of 20 mA.

3. Results and discussion

3.1 XRD and SEM analysis

Figure 1(a) displays the XRD patterns of CZMG:Mn4+ phosphor. It should be pointed out that all the diffraction peaks are in agreement with the standard ICSD card No.245649 when x is less than 0.60, which indicates that a small amount of Mg2+ substitution has no impact on the CZG host crystal structure, while the impurity peaks around 17°, 22° and 32° appear when x is greater than 0.60. Figure 1(b) is refined results for the XRD profile of the CZMG:Mn4+. The residual values of Rwp = 10.00%, Rp = 7.35% illustrates refined results are dependable and atom coordinates, thermal vibration parameters and fraction factors are basically in the same line with experimental values.

 figure: Fig. 1

Fig. 1 (a) XRD patterns of CZMG:Mn4+(x = 0.00, 0.20, 0.40, 0.60, 0.80, 1. 00); (b) Rietveld refinement of the XRD profile CZMG:Mn4+(x = 0.40).

Download Full Size | PDF

As we know, CZG belongs to cubic system with space group F23(196) and the standard cell parameters of it are a = 15.0794Å, Z = 4, V = 3428.88Å3. In order to corroborate the successful incorporation of Mg2+ in CZG host, the cell parameters are obtained by XRD Rietveld refinement as listed in Table 1. With the increase of Mg2+ concentration, the diminishing Unit-cell edge length and volume reveal that the substitution of Zn by Mg2+ is successful because the ionic radius of Mg2+ (0.72Å for CN = 6) is smaller than that of Zn2+ (0.74Å for CN = 6). In addition, the cell Volume difference of the phosphor with x less than 0.40 is far less than that of the phosphor with x more than 0.40, implying the substitution of Mg2+ ions is almost to a limit.

Tables Icon

Table 1. Cell parameters of Ca14Zn6-xMgxGa9.86O35:0.14Mn4+

Figure 2(a) and (b) is the SEM image of CZMG:Mn4+ (x = 0.00) and CZMG:Mn4+(x = 0.40) respectively. Two images show that the fine particles with few micrometers in size have an irregular morphology because of aggregation. Moreover, it is worth mentioning that the aggregation extent of CZMG (x = 0.40):Mn4+ is more serious than that of CZMG:Mn4+(x = 0.00).

 figure: Fig. 2

Fig. 2 SEM image of (a) CZMG:Mn4+ (x = 0.00) and (b) CZMG:Mn4+(x = 0.40).

Download Full Size | PDF

3.2 Luminescence properties of CZMG:Mn4+

The excitation spectra of the CZMG:Mn4+ phosphor monitored at 711nm were acquired in Fig. 3(a). Two intense excitation bands at 250–420nm and 420–550nm are observed. The former excitation band corresponds to the overlap of the Mn4+–O2– charge transfer and the 4A24T1 transition of Mn4+, and the latter region, in which the excitation peak is located at 460nm, corresponds to the 4A24T2 transition of Mn4+ [22–25]. It can also be known that the excitation intensity increases with the increase of Mg2+ concentration and then decreases sharply.

 figure: Fig. 3

Fig. 3 Normalized excitation (a) and emission (b) spectra of CZMG:Mn4+(x = 0.00, 0.20, 0.40, 0.60, 0.80, 1.00).

Download Full Size | PDF

The emission spectra of CZMG:Mn4+ excited at 460nm in Fig. 3(b). The emission band at 650nm-750nm produced by spin-forbidden 2E → 4A2 transition of Mn4+ is detected [22]. The emission intensity of the phosphor continuously increases with the increase in Mg2+ doping c oncentration until x = 0.40, then decreases. Moreover, there is a slight shift of the emission peaks toward longer wavelengths with the increase in Mg2+ .

Here, two factors account for the phenomenon that the Mg2+ substitution has a positive influence in luminous intensity. In reality, Mg2+ occupies the Zn2+ position at the center of the octahedron, the original Mn4+-Mn4+-O2- pairs tend to form Mn4+-Mg2+ pairs, which increases the transition probability from the ground state to the excited state and inhibits the energy transfer between Mn4+ and Mn4+, thus enhancing the luminescence [26,27], The other factor is that a spot of incorporated Mg2+ as a flux made the particle size and shape more uniform in Fig. 2(a) and Fig. 2(b), which leads to greater luminescence. With regard to reducing emission intensity subsequently, on the one hand, it is most likely that impurity phase(s) forms and increases when Mg2+ concentration more than 0.60 because impurity peaks increase in Fig. 1(a) and cell volume difference decreases in Table 1; on the other, substitutional defects lead to lattice distortion, and then damage the rigidity of crystal structure, increasing nonradiative transition rate. Additionally, the phenomenon of red shift of emission peak can be explained by the reason that the incorporation induced lattice shrinking and therefore crystal field weakening because Mg2+ radius less than that of Zn2+.

3.3 Thermal resistance performance of CZMG:Mn4+ and CZG:Mn4+

To evaluate the application of the phosphor in LED, the temperature-dependent emission spectra of CZMG:Mn4+(x = 0.00) and CZMG:Mn4+(x = 0.40) are presented in Fig. 4(a) and Fig. 4(b) respectively. It can be seen that their emission intensity gradually decreases with the increases in temperature from 318 K to 518 K due to thermal quenching behavior. The intensity decreases to only ~50% of the initial emission intensity when the temperature increases by 200°C, implying that both of them have low thermal quenching and outstanding thermal resistance performance. In order to study the effect of Mg2+ substitution on phosphor in detail, the activation energy ∆E is used as a specific evaluation indicator of temperature dependence, which can be obtained by the Arrhenius equation as follows:

IT=I0[1+cexp(ΔE/kT)]
where I0 represents the initial emission intensity, IT is the emission intensity at a given temperature, c and T are both constants, and k is the Boltzmann constant (8.617 × 10−5 eV/K). Based on the equation, the fitted linear relationship between ln[(I0/IT) – 1] and 1/kT is given in Fig. 4(c), and the activation energy ∆E, the abs of the slope was obtained as 0.2065 and 0.1868 for CZMG:Mn4+(x = 0.00) and CZMG:Mn4+(x = 0.40), which manifests that the thermal resistance performance of the phosphor is slightly reduced by the addition of Mg2+. The mechanism of this phenomenon is that the defects caused by Mg2+ substitution lead to poor rigidity of crystal structure, which makes the vibration relaxation or nonradiative transition stronger when the temperature goes up.

 figure: Fig. 4

Fig. 4 Temperature-dependent emission spectra of (a) CZMG:Mn4+(x = 0.00) and (b) CZMG:Mn4+(x = 0.40) excited at 460 nm, (c) Dependence of ln(I0/IT–1) on 1/kT.

Download Full Size | PDF

3.4 Luminescence decay properties of CZMG:Mn4+

Figure 5 shows the decay curves of the samples excited at 460 nm and monitored at 711 nm. All the sample curves are well fitted with the following single exponential formula:

I=I0exp(t/τ)+A
where I and I0 are the emission intensity of sample at a given time and the initial emission intensity, respectively, and τ is the lifetime. According to the equation, the decay time decreases from 2.72 ms to 2.25 ms with the increase in Mg2+ concentration. Note also that the decay time is 2.57ms when x = 0.40. As a matter of fact, lifetime of the phosphor depends on the following formula:
τ=1WR+WNR
In the formula, WR and WNR represent radiation transition rate constant and nonradiative transition rate constant respectively. In consequence, the phosphor lifetime decreasing most likely derives from increase in radiation transition rate and nonradiative transition rate caused by Mn4+-Mg2+ pairs formation and substitutional defects respectively.

 figure: Fig. 5

Fig. 5 Decay curves of CZMG:Mn4+(x = 0.00, 0.20, 0.40, 0.60, 0.80, 1.00) excited at 460 nm.

Download Full Size | PDF

3.5 Internal quantum efficiency and luminous efficiency of CZMG:Mn4+

The variation of internal quantum efficiency and luminous efficiency of the phosphor with different Mg2+ incorporation is depicted in Fig. 6(a). The internal quantum efficiency was calculated by following formula:

ηQE=LSERES
Where Ls is the emission spectrum of the sample, ER and ES are the excitation light with and without the sample in the integrating sphere. The quantum efficiency first increases from 61.7% to 64.3% and then decrease from 64.3% to 56.8% with the increasing concentration of Mg2+ from 0.00 to 1.00, which is positively correlated with the variation of emission intensity. To investigate effect of Mg2+ substitution on luminous efficiency, the red LED devices are fabricated by covering mixture on 460nm chip. The mixture is made up of 10% CZMG:Mn4+(x = 0.00, 0.20, 0.40, 0.60, 0.80, 1.00) and 90% glue respectively. The luminous efficiency in Fig. 6(a) from 14.67lm/w to 16.58lm/w corresponding the increase in Mg2+ concentration from 0.00 to 0.40, and then it decreases from maximum to 11.86lm/w as the concentration of Mg2+ continues to increase. The change in trends of internal quantum efficiency and luminous efficiency is synchronous on the whole. Figure 6(b) displays electroluminescence of device with 10% CZMG:Mn4+(x = 0.40), implying that the phosphor has potential application in plant lighting.

 figure: Fig. 6

Fig. 6 (a). The internal quantum efficiency and luminous efficiency of CZMG:Mn4+ as a function of Mg2+ concentration under 460nm excitation. (b) The electroluminescence spectra of CZMG:Mn4+(x = 0.40) based on 460nm chip under 20 mA current excitation The inset shows the photo of the phosphor coated LED.

Download Full Size | PDF

4. Conclusion

A series of red-emitting CZMG:Mn4+ phosphors were synthesized by a conventional high-temperature solid-state method, and their properties were investigated in detail. As the degree of Mg2+ replacing Zn2+ increases, the luminescence of the phosphor first increases and then decreases. The reason for the enhancement is that the addition of Mg2+ ions can make the phosphor particles more uniform and form Mn4+-Mg2+ pairs, while the decrease is due to the formation and increase of impurities with the increase of Mg2+ concentration. The redshift of the emission peak is caused by the weakening of the crystal field. Besides, the phosphor has good thermal resistance and Mg2+ substitution results in poor thermal resistance of the phosphor due to defects formation. The decrease in decay time from 2.72ms to 2.25ms is attributed to the increase in radiation transition rate and nonradiative transition rate. The internal quantum efficiency and luminous efficiency reach the maximum at 64.3% and 16.58lm/w respectively when the doping concentration of Mg2+ is 0.40. These results indicate that the optimum doping concentration of Mg2+ is determined to be 0.40 and CZMG:Mn4+ (x = 0.40) is an applicable phosphor for plant lighting.

Funding

National Natural Science Foundation of China (Nos. 51372172 and 51672192).

References and links

1. R. P. Cao, Z. H. Shi, G. J. Quan, T. Chen, S. L. Guo, Z. F. Hu, and P. Liu, “Preparation and luminescence properties of Li2MgZrO4: Mn4+ red phosphor for plant growth,” J. Lumin. 188, 577–581 (2017). [CrossRef]  

2. Z. Zhou, M. Xia, Y. Zhong, S. J. Gai, S. X. Huang, Y. Tian, X. Y. Lu, and N. Zhou, “Dy3+@ Mn4+ co-doped Ca14Ga10−mAlmZn6O35 far-red emitting phosphors with high brightness and improved luminescence and energy transfer properties for plant growth LED light,” J. Mater. Chem. C 5(32), 8201–8210 (2017). [CrossRef]  

3. Z. Zhou, J. Zheng, R. Shi, N. Zhang, J. Chen, R. Zhang, H. Suo, E. M. Goldys, and C. Guo, “Abinitiosite occupancy and far-red emission of Mn4+ in cubic-phase La (MgTi)1/2O3 for plant cultivation,” ACS Appl. Mater. Interfaces 9(7), 6177–6185 (2017). [CrossRef]   [PubMed]  

4. 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. Alloys Compd. 417(1–2), 273–279 (2006). [CrossRef]  

5. H. S. Kim, K. Machida, T. Horikawa, and H. Hanzawa, “Luminescence properties of CaAlSiN3: Eu2+ phosphor prepared by direct-nitriding method using fine metal hydride powders,” J. Alloys Compd. 633, 97–103 (2015). [CrossRef]  

6. Q. F. Wang, Y. Liu, Y. Wang, W. Wang, Y. Wan, G. G. Wang, and Z. G. Lu, “Considerable photoluminescence enhancement of LiEu(MoO4)(2) red phosphors via Bi and/or Si doping for white LEDs,” J. Alloys Compd. 625, 355–361 (2015). [CrossRef]  

7. R. Cao, Z. Shi, G. Quan, Z. Luo, P. Tang, H. Ao, and X. Yu, “Li3Mg2NbO6: Mn4+ red phosphor for light-emitting diode: Synthesis and luminescence properties,” Opt. Mater. 57, 212–216 (2016). [CrossRef]  

8. T. Jansen, J. Gorobez, M. Kirm, M. G. Brik, S. Vielhauer, M. Oja, N. M. Khaidukov, V. N. Makhov, and T. Justel, “Narrow Band Deep Red Photoluminescence of Y2Mg3Ge3O12:Mn4+, Li+ Inverse Garnet for High Power Phosphor Converted LED,” ECS Journal of Solid State Science and Technology 7(1), R3086–R3092 (2018). [CrossRef]  

9. G. Hu, X. Hu, W. Chen, Y. Cheng, Z. Liu, Y. Zhang, X. Liang, and W. Xiang, “Luminescence properties and thermal stability of red phosphor Mg2TiO4: Mn4+ additional Zn2+ sensitization for warm W-LEDs,” Mater. Res. Bull. 95, 277–284 (2017). [CrossRef]  

10. Z. Liu, M. Yuwen, J. Liu, C. Yua, T. Xuan, and H. Li, “Electrospinning, optical properties and white LED applications of one-dimensional CaAl12O19:Mn4+ nanofiber phosphors,” Ceram. Int. 43(7), 5674–5679 (2017). [CrossRef]  

11. J. Long, X. Yuan, C. Ma, M. Du, X. Ma, Z. Wen, R. Ma, Y. Wang, and Y. Cao, “Strongly enhanced luminescence of Sr4Al14O25:Mn4+ phosphor by co-doping B3+ and Na+ ions with red emission for plant growth LEDs,” RSC Advances 8(3), 1469–1476 (2018). [CrossRef]  

12. X. Y. Sun, Z. He, and X. Gu, “Synthesis, deep red emission and warm WLED applications of K2SiF6: Mn4+ phosphors,” J. Photochem. Photobiol. Chem. 350, 69–74 (2018). [CrossRef]  

13. Y. Liu, G. Gao, L. Huang, Y. Zhu, X. Zhang, J. Yu, B. S. Richards, T. Xuan, Z. Wang, and J. Wang, “Co-precipitation synthesis and photoluminescence properties of BaTiF6:Mn4+: an efficient red phosphor for warm white LEDs,” J. Mater. Chem. C Mater. Opt. Electron. Devices 6(1), 127–133 (2018). [CrossRef]  

14. X. Gao, W. Xia, T. Chen, X. Yang, X. Jina, and S. Xiao, “Conversion of broadband UV-visible light to near infrared emission by Ca14Zn6Al10O35: Mn4+, Nd3+/Yb3+,” RSC Advances 6(9), 7544–7552 (2016). [CrossRef]  

15. J. Chen, W. Zhao, N. Wang, Y. Meng, S. Yi, J. He, and X. Zhang, “Energy transfer properties and temperature-dependent luminescence of Ca14Zn6Al10O35:Dy3+, Mn4+ phosphors,” J. Mater. Sci. 51(9), 4201–4212 (2016). [CrossRef]  

16. L. Li, Y. Pan, Z. Chen, S. Huang, and M. Wu, “Tunable luminescence and energy transfer properties of Bi3+ and Mn4+ co-doped Ca14Zn6Al10O35 phosphors for agricultural applications,” RSC Advances 7(24), 14868–14875 (2017). [CrossRef]  

17. J. Y. Park, J. S. Joo, H. K. Yang, and M. Kwak, “Deep red-emitting Ca14Zn6Al10O35: Mn4+ phosphors for WLED applications,” J. Alloys Compd. 714, 390–396 (2017). [CrossRef]  

18. W. Lü, W. Lv, Q. Zhao, M. Jiao, B. Shao, and H. You,”A novel efficient Mn4+ activated Ca14Zn6Al10O35: phosphor: application in red-emitting and white LEDs,” Inorg. Chem. 53(22), 11985–11990 (2014). [CrossRef]   [PubMed]  

19. T. Hasegawa, S. W. Kim, T. Abe, S. Kumagai, R. Yamanashi, K. Seki, K. Uematsu, K. Toda, and M. Sato, “Improvement of emission intensity for near-infrared-emitting Ca14Zn6Al10O35:Mn4+ phosphor by oxygen-pressure method,” Chem. Lett. 45(9), 1096–1098 (2016). [CrossRef]  

20. K. Seki, K. Uematsu, K. Toda, and M. Sato, “Novel deep red emitting phosphors Ca14Zn6Al10O35:Mn4+ (M= Al3+ and Ga3+),” Chem. Lett. 43(8), 1213–1215 (2014). [CrossRef]  

21. C. Yang, Z. Zhang, G. Hu, R. Cao, X. Liang, and W. Xiang, “A novel deep red phosphor Ca14Zn6Al10O35:Mn4+ as color converter for warm W-LEDs: structure and luminescence properties,” J. Alloys Compd. 694, 1201–1208 (2017). [CrossRef]  

22. Y. Zhou, W. Zhao, J. Chen, and Z. Liao, “Highly efficient red emission and multiple energy transfer properties of Dy3+/Mn4+ co-doped Ca14Zn6Al10O35 phosphors,” RSC Advances 7(28), 17244–17253 (2017). [CrossRef]  

23. Y. Xu, Y. Zhang, L. Wang, M. Shi, L. Liu, and Y. Chen, “Red emission enhancement for CaAl12O19:Cr3+ and CaAl12O19: Mn4+ phosphors,” J. Mater. Sci. Mater. Electron. 28(16), 12032–12038 (2017). [CrossRef]  

24. R. Cao, Y. Ye, Q. Peng, G. Zheng, H. Ao, J. Fu, Y. Guo, and B. Guo, “Synthesis and luminescence characteristics of novel red-emitting Ba2TiGe2O8: Mn4+ phosphor,” Dyes Pigments 146, 14–19 (2017). [CrossRef]  

25. J. Deng, H. Zhang, X. Zhang, Y. Zheng, J. Yuan, H. Liu, Y. Liu, B. Lei, and J. Qiuc, “Ultrastable Red-emitting Phosphor-in-glass for Superior High-power Artificial Plant Growth LEDs,” J. Mater. Chem. C 6(7), 1738–1745 (2018). [CrossRef]  

26. Y. X. Pan and G. K. Liu, “Influence of Mg2+ on luminescence efficiency and charge compensating mechanism in phosphor CaAl12O19:Mn4+,” J. Lumin. 131(3), 465–468 (2011). [CrossRef]  

27. J. Zhong, D. Chen, X. Chen, K. Wang, X. Li, Y. Zhu, and Z. Ji, “Efficient rare-earth free red-emitting Ca2YSbO6:Mn4+,M(M = Li+, Na+, K+, Mg2+) phosphors for white light-emitting diodes,” Dalton Trans. 47(18), 6528–6537 (2018). [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 (a) XRD patterns of CZMG:Mn4+(x = 0.00, 0.20, 0.40, 0.60, 0.80, 1. 00); (b) Rietveld refinement of the XRD profile CZMG:Mn4+(x = 0.40).
Fig. 2
Fig. 2 SEM image of (a) CZMG:Mn4+ (x = 0.00) and (b) CZMG:Mn4+(x = 0.40).
Fig. 3
Fig. 3 Normalized excitation (a) and emission (b) spectra of CZMG:Mn4+(x = 0.00, 0.20, 0.40, 0.60, 0.80, 1.00).
Fig. 4
Fig. 4 Temperature-dependent emission spectra of (a) CZMG:Mn4+(x = 0.00) and (b) CZMG:Mn4+(x = 0.40) excited at 460 nm, (c) Dependence of ln(I0/IT–1) on 1/kT.
Fig. 5
Fig. 5 Decay curves of CZMG:Mn4+(x = 0.00, 0.20, 0.40, 0.60, 0.80, 1.00) excited at 460 nm.
Fig. 6
Fig. 6 (a). The internal quantum efficiency and luminous efficiency of CZMG:Mn4+ as a function of Mg2+ concentration under 460nm excitation. (b) The electroluminescence spectra of CZMG:Mn4+(x = 0.40) based on 460nm chip under 20 mA current excitation The inset shows the photo of the phosphor coated LED.

Tables (1)

Tables Icon

Table 1 Cell parameters of Ca14Zn6-xMgxGa9.86O35:0.14Mn4+

Equations (4)

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

IT=I0[1+cexp(ΔE/kT)]
I=I0exp(t/τ)+A
τ= 1 WR+WNR
ηQE= LS ER ES
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.