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

Temperature dependent red luminescence from a distorted Mn4+ site in CaAl4O7:Mn4+

Open Access Open Access

Abstract

Thermal luminescence quenching behavior of a phosphor is essential for application in phosphor converted white light emitting diodes (pc-WLEDs) because the phosphor layer can be heated up to 473K in a working high power WLEDs. Here, we have confirmed indeed a red luminescence of Mn4+ substituting for calcium sites rather than tetrahedral aluminum sites in CaAl4O7:Mn which can be synthesized in pure phase even with boron acid as flux, and examined the low and high temperature luminescent properties in the range of 10 to 500K. We have revealed as well as thermal quenching mechanism that distorted octahedral Mn4+ sites suffer severe thermal quenching. This work, thus, hints a strategy to find a new Mn4+ phosphor with better resistance to thermal impact in the future.

©2013 Optical Society of America

1. Introduction

Recently, WLEDs have been paid increasing attention due to the advantages over traditional incandescent or fluorescent lightings such as high luminous efficiency, long lifetime, energy saving and therefore environmental benefits etc, and for these, they have been promoted in the widespread applications for instance in domestic and commercial lighting, automobiles, communications, imaging, agriculture, and medicine [13]. Currently for fabrication of pc-WLEDs, the mainstream technology is to combine a single LED chip with one or more conversion phosphors [26]. For the purpose, desirable phosphors are highly needed to efficiently convert the near ultraviolet (typically 380 nm to 420 nm) or blue emission (typically 450 nm to 480 nm) from (In,Ga)N LED chip to visible lights. Among three primary color phosphors, searching a new red phosphor is one of the big challenges in the field of luminescent materials particularly when a white lighting is preferred in some occasions with a warm perception, a high color rendering ability and full color gamut [26]. It, thus, has triggered a series of findings on new red phosphors (e.g. activated by rare earth, transition metal, divalent bismuth, or oxygen related defects etc as well as nanostructured composite [716]. This has led to the discovery of the outstanding Eu2+ doped oxynitrides and nitrides phosphors [2,3,9,17].

The general requirements of potential red phosphor for instance for a blue chip scheme are: (1) significant absorption in blue but not in green; (2) high quantum efficiency of the red emission; (3) good resistance of the emission to thermal impact; (4) good tolerance to ambient; and (5) compatibility with present and future environmental regulations, for instance, free of toxic species, green compositions. Some of them cannot be met simultaneously by previous approaches. For an instance, Eu2+-doped red nitride phosphors M2Si5N8:Eu2+ (M = Ca, Sr, Ba) can absorb blue and green lights [2,3,6,17]. Thus, it is unavoidable to reabsorb the co-existent green emission of WLEDs.

Recently Mn4+ doped phosphors were recognized as one of the promising alternatives, because Mn4+ usually exhibits broad and strong absorption between 300 and 480 nm and emits light between 600 nm to 760 nm once it is introduced into an octahedral site [1822]. The luminescent properties remain quite similar from host to host because the lowest excited state energy, 2E(t23), barely changes upon variation of the crystal field. Lately Park et al reported CaAl4O7:Mn4+ as a new good candidate of red phosphor for WLEDs using UV chips [21]. This argument, however, needs reconsideration not only because a confirmation has to be done on whether CaAl4O7:Mn4+ is a red phosphor, but also because the properties of the phosphor are not well known as we check it over with all the requirements listed above. A single phase of CaAl4O7:Mn was not actually synthesized with boron acid H3BO3 as flux though the phase dominated over the other phases (α-Al2O3 and CaAl12O19) in some cases [21]. The authors commented that the contribution of minor phase could be negligible to the red luminescence of the sample (see Fig. 1 of [21]). The comment, however, is questionable in respect of the facts: (1) the red luminescence of Mn4+ in Al2O3 and CaAl12O19 has been confirmed undoubtedly, and particularly the luminescence of CaAl12O19:Mn4+ is very efficient [18,19]; (2) no red luminescence was observed from Mn doped sample where CaAl4O7 and Al2O3 coexisted [21]. This is one of the impetuses to the present study. Another motivation is to understand the capacity of the title compound to resist temperature quenching if it was a red phosphor. It is inspired by the fact that the temperature of the layer capped on top of a blue diode can reach 150-200°C due to the heat rouse from the p-n junction and the phosphor layer as the device works [2,3,16,22].

 figure: Fig. 1

Fig. 1 (a) XRD patterns of CaAl4O7: 0.1%Mn prepared at different temperature in air with 2%H3BO3 as flux and reference pattern of CaAl4O7 (ICSD #34487) at the bottom. (b) XRD pattern (-o-) of the sample of CaAl4O7: 0.1%Mn prepared at 1400°C, Rietveld refining results (), Bragg reflections (|) and the profile difference between experimental and calculated values (). Inset shows the lattice cell of the compound viewed along c; blue ball: Ca, green ball: Al, red ball: O; purple polyhedra: CaO7, dark cyan polyhedra: AlO4. (c) Exemplary excitation and emission spectra of CaAl4O7: 0.1%Mn4+ prepared at 1400 (curves 1 and 2) and 1600 °C (curves 3 and 4). Calcium sites which could stabilize and accommodate Mn4+: (d) ideal CaO6 octahedron in the compound of CaO, where Ca-O bond length is 2.407(6)Å; (e) CaO7 polyhedron in CaAl4O7, where one Ca-O(1) bond is 2.488(9)Å, two Ca-O(2) bond are 2.367(6)Å, two Ca-O(2) bond are 2.925(6)Å and two Ca-O(3) bond are 2.310(2)Å.

Download Full Size | PDF

We have shown, here, it is possible to synthesize the pure phase of CaAl4O7:Mn even when H3BO3 is used as flux. We have confirmed the red luminescence from it due to the Mn4+ substitution for calcium sites. This is rather different from what Park et al [21] reported on as well as the spectroscopic properties. Meanwhile, we have inspected the temperature dependent luminescence in a wide range from 10K to 500K, and discussed the mechanism for the severe thermal quenching.

2. Experimental procedure

The samples were synthesized by typical high temperature solid state reaction with analytical reagents CaCO3, Al(OH)3, MnCO3 and H3BO3 as raw materials. Individual 10-gram batches were weighed according to CaAl4O7: 0.1%Mn, where 2mol% extra H3BO3 was added as flux. The samples were mixed homogeneously in an agate mortar, and moved to alumina crucibles for subsequent processes. They were heated up to 700°C with a rate of 1.2°C/min and held for 5h to thermally decompose the starting reagents, and then cooled down to room temperature naturally. They, after intermediate grinding to improve the mixing homogeneity, were heated at a rate of 2°C/min to 1300, 1400, 1500 and 1600°C, respectively, and sintered at each temperature for 5h, and finally cooled down to room temperature. All the operations were performed in air.

X-ray diffraction (XRD) patterns of the samples were recorded by a Rigaku D/max-IIIA X-ray diffractometer (40 kV, 1.2° min−1, 40 mA, Cu-Kα1, λ = 1.5405 Å). Static excitation and emission spectra, dynamic emission decay spectra, the emission spectra at 10 to 500K were measured by an Edinburgh FLS 920 instrument equipped with a red-sensitive photomultiplier (Hamamatsu R928P) and a closed cycle helium cryostat. Excitation curves were corrected over the lamp intensity with a silicon photodiode, and the emission curves were corrected by the PMT spectral response. Measurements were performed at room temperature unless otherwise specified.

3. Results and discussion

When prepared at 1200°C, Park et al found the phases of α-Al2O3 and CaAl4O7 coexisted in Mn doped sample [21]. This changed as the preparation temperature rose up to 1300, 1400 or 1500 °C. The secondary minor phase became into CaAl12O19 instead of α-Al2O3. The scenario is very different in our case. Figure 1(a) depicts XRD pattern of CaAl4O7: 0.1%Mn prepared at different temperature in air. Comparison to reference pattern of CaAl4O7 (ICSD #34487) reveals the nature of pure phase in all the samples. For achieving crystallographic data for the following discussion, we analyzed for example the XRD pattern of the sample CaAl4O7: 0.1%Mn prepared at 1400°C with the classic Rietveld refining technique. The refining starts with the crystallographic data of CaAl4O7 (ICSD #34487) and it converges with residual values of Rp = 10.50%, Rwp = 14.10%, and Rexp = 7.62%. The structural refining results are depicted in Fig. 1(b) along with the different profile between experimental and calculated values. The comparison confirms that the sample can be indexed in a monoclinic C12/c1 space group. The refining calculation produced lattice parameters of a = 12.893(9)Å, b = 8.888(8)Å, c = 5.446(9)Å, β = 106.91(3) and cell volume V = 597.2707 Å3, well consistent with reference data of CaAl4O7 (ICSD #34487).

As illuminated with either UV or Blue lights, all the samples exhibit typical Mn4+ red luminescence. And this, however, differs from what Park et al reported [21]. They observed that the excitation spectrum consisted of three peaks at 335, 380 and 467nm, the emission spectrum was peaked at 656nm and the two UV excitation peaks exhibited comparable intensity particularly for their sample prepared at 1300 °C. This was very similar to CaAl12O19:Mn4+ as noticed by Park et al [21], and it was possibly due to the large amount of CaAl12O19 phase in the samples. The blue excitation peak showed fine structure similar to xenon lamp, possibly due to improper calibration [21]. Figure 1(c) exemplarily illustrates excitation and emission spectra of our samples prepared at different temperatures. The excitation spectrum comprises also three peaks of Mn4+ but at ~330, ~390, and ~464nm. The intensity of ~390nm shoulder peak is roughly half of the strongest peak at ~330nm, and the blue peak after well calibrated does not appear structured. The emission spectrum peaks at ~655nm upon 330nm excitation. It is noticed that the preparation temperature has influences on emission intensity and lifetime but not the spectral shape (see Fig. 1(c)). For instance, as the temperature rises up from 1300 to 1600°C. the lifetime prolongs from 642 to 963μs.

Park et al reported red luminescent Mn4+ was substituted for tetrahedral Al3+ sites in the sample CaAl4O7: Mn4+ [21]. This contradicts what have been reported so far, and only Mn4+ in octahedral crystal field can emit red light [1820, 22]. In the compound as revealed by the refining, there are one type of seven-coordinate calcium sites Ca(1), two types of tetrahedral aluminum sites Al(1) and Al(2), and four types of oxygen sites O(1), O(2), O(3) and O(4) (shown as inset of Fig. 1(b)). So clearly, even if Mn4+ could substitute for either Al(1) or Al(2) in view of size match, the tetrahedral Mn4+ ions cannot red luminesce. The rest host cation sites that Mn4+ ions possibly reside will be Ca(1). Actually reports have confirmed that octahedral Mn4+ ions can disolvedly substitute for the Mg or Ca sites in MgO and CaO which are fully built up by MgO6 and CaO6 polyhedra, respectively [18]. For ideal CaO6 polyhedra as depicted in Fig. 1(d), length of each Ca-O bond is 2.407(6)Å, and one Ca site and four O sites share equatorial plane. The polyhedron of CaO7 in CaAl4O7 could be regarded as a distorted version of CaO6 polyhedron (see Fig. 1(e)) once the two O(2) atoms which connect to Ca in the longest bond were treated as an imaginary atom entity. When Mn4+ replaces Ca(1), local field perhaps is distorted due to increased Coulomb interaction, and it turns into typical octahedral field eventually as evidenced by Fig. 1(c) and Fig. 2.

 figure: Fig. 2

Fig. 2 (a) Excitation spectra (λem = 652nm), (b) emission spectra (λex = 325nm) and (c) decay curves (λem = 652nm, λex = 325nm) of CaAl4O7: 0.1%Mn4+ at different temperatures as indicated. (d) The temperature dependence of the integrated emission intensities (light green ellipse pattern: “red star” for λex = 325nm; “blue ball” for λex = 470nm) and lifetimes (light blue ellipse pattern: “red star” for λex = 325nm and λem = 652nm; “blue ball” for λex = 470nm and λem = 652nm). The green line through the data points are fits to τ(T) = τ0/(1 + τ0Ce-Ea/kT) with τ0 = 1435μs, C = 1.47, and Ea = 0.196eV. The goodness of fitting is 96.0%.

Download Full Size | PDF

The luminescent properties of the phosphor of CaAl4O7: Mn4+ show clear temperature dependence (Fig. 2). At 10K, the excitation spectrum does not appear fine structured, and it comprises three peaks at 325nm, 381nm and 452nm, due to the transitions of 4A24T1 (partly Mn4+-O2- charge transfer transition), 4A22T2, 4A24T2, respectively [1820,22]. These peaks redshift obviously as the temperature increases (Fig. 2(a)). For instance, the strongest peak lies at ~325nm at 10K, and it goes to ~330nm, ~336nm, ~342nm at 300K, 400K, and 500K, respectively.

The emission spectra show also redshift as Fig. 2(b) depicts. At low temperature, e.g. 10K, anti-Stokes feature disappears, and two sharp peaks were observed at 15649cm−1 and 15337cm−1 along with Stokes sidebands. The two peaks show different red shifts as the temperature rises up. The former peak shift is 48cm−1 and 97cm−1, and the shift of the latter is 70cm−1 and 116cm−1 when the temperature increases from 10K to 300K and 450K, respectively. This implies that the two peaks are possible associated with different Mn4+ emission centers. According to the criterion developed by Brik and Srivastava [18], the low energy peak of 15337cm−1 should be due to a Mn4+ site with higher covalency.

Regardless of consequent charge compensation and distortion after substitution of Mn4+ for Ca(1) site in CaAl4O7: Mn4+, the symmetry of the Mn4+ site is C2. At the site without inversion symmetry, electric dipole character can be introduced into the forbidden transition of 2E→4A2 by mixing 3d3 with odd parity configurations via crystal field. Therefore, zero phonon line of the transition is expected strong at the site of C2. Indeed, it is the case in CaAl4O7: Mn4+ where the peak at 15337cm−1 (652nm) dominates the spectrum at 10K. At higher temperature, vibronic anti-Stokes and particularly Stokes sidebands are evidently enhanced, which reflects the phonon-involved nature of these transitions (see Fig. 2(b)).

Emission decay behavior depends strongly on temperature (Fig. 2(c)). At lower temperature, decay curve fits well single exponential decay equation, and the fitting deteriorates as the temperature increases. The lifetime reduces significantly from 1578μs@10K to 13μs@500K for instance upon excitation of 325nm. Figure 2(d) summarizes the dependences of emission lifetime and intensity on temperature. And they show similar trend for excitation schemes of either UV or blue. The temperature at which the intensity decreases to half of the value at 10K is about 200K. At 500K, only very weak luminescence of Mn4+ was recorded. Significant lifetime shortening gives evidence for both cases of excitation that there is the intrinsic severe quenching in CaAl4O7: Mn4+. The decrease in lifetime is accompanied by the decrease of emission intensity.

One of the mechanisms for Mn4+ luminescence quenching might be quenching by thermally activated crossover from the first excited state of 2E to the ground state 4A2. For evaluating activation energy (Ea) for thermal quenching, we fit the temperature dependence of emission lifetime rather than intensity to a modified Arrhenius equation [6]: τ(T) = τ0/(1 + τ0Ce-Ea/kT), where τ(T) is the emission decay time at temperature T (K), τ0 is the emission decay time at 0K, C is a rate constant for the process. This is because emission intensity when compared to lifetime is very susceptible to various factors. Best fitting leads to Ea = 0.196eV. It is much lower than 0.6eV of SrSi2O2N2:Eu2+, which can well explain why luminescence quenching occurs in CaAl4O7: Mn4+ but not SrSi2O2N2:Eu2+ at 500K [6].

When inspecting the excitation spectra at 450 and 500K, we noticed there appears a sharp peak at ~406nm and it is intensified as temperature increases. The peak comes from the transition of 6A1(S)→4E(G), 4A1(G) of a typical octahedral Mn2+ site [7,23]. The trace evidence reveals the thermal induced reduction happens in situ from Mn4+ to Mn2+. This means that valence change might be one of the causes for luminescence quenching of Mn4+. And it could lead to permanent thermal degradation of the phosphor. In addition, negative defects, which are possible created for charge compensation due to aliovalent substitution of Mn4+ for host cations, might also play as luminescence quencher.

4. Conclusions

Indeed, CaAl4O7: Mn4+ is a red phosphor and it can be synthesized in pure phase in presence of boric acid as flux. The red luminescence comes from a distorted octahedral Mn4+ sites due to substitution for Ca(1) sites in the compound. And it, however, suffers from strong thermal quenching, revealed by the low quenching temperature and Ea. The dominating quenching mechanism may be the thermally activated crossover from the first excited state to the ground state and the in situ reduction of Mn4+ to Mn2+ at higher temperature. This work reveals to us that CaAl4O7: Mn4+ would be less practically valuable in the application of pc-WLEDs if it was not functionally engineered in surface, size, morphology etc. Meanwhile, it implies a general strategy to explore a new Mn4+ phosphor with better performances in future, that is, the host compound should better comprise regular octahedral sites that Mn4+ could occupy.

Acknowledgments

M. Peng would like to thank Prof. J. Wang and Dr. X. Zhang for kind helps on high temperature measurements. This work was financially supported by National Natural Science Foundation of China (Grant No. 51072060), Guangdong Natural Science Foundation for Distinguished Young Scholars (Grant No. S20120011380), SRF for ROCS SEM and Chinese Program for New Century Excellent Talents in University (Grant No. NCET-11-0158).

References and links

1. S. Pimputkar, J. Speck, S. DenBaars, and S. Nakamura, “Prospects for LED lighting,” Nat. Photonics 3(4), 180–181 (2009). [CrossRef]  

2. R. J. Xie, N. Hirosaki, K. Sakuma, and N. Kimura, “White light-emitting diodes (LEDs) using (oxy) nitride phosphors,” J. Phys. D Appl. Phys. 41(14), 144013 (2008). [CrossRef]  

3. P. Smet, A. Parmentier, and D. Poelman, “Selecting Conversion Phosphors for White Light-Emitting Diodes,” J. Electrochem. Soc. 158(6), R37–R54 (2011). [CrossRef]  

4. G. Zhang, J. Wang, Y. Chen, and Q. Su, “Two-color emitting of Ce3+ and Tb3+ co-doped CaLaGa3S6O for UV LEDs,” Opt. Lett. 35(14), 2382–2384 (2010). [CrossRef]   [PubMed]  

5. Y. Chen, J. Wang, C. Liu, X. Kuang, and Q. Su, “A host sensitized reddish-orange Gd2MoO6:Sm3+ phosphor for light emitting diodes,” Appl. Phys. Lett. 98(8), 081917 (2011). [CrossRef]  

6. V. Bachmann, A. Meijerink, and C. Ronda, “Luminescence properties of SrSi2AlO2N3 doped with divalent rare-earth ions,” J. Lumin. 129(11), 1341–1346 (2009). [CrossRef]  

7. Y. Liu, X. Zhang, Z. Hao, Y. Luo, X. Wang, and J. Zhang, “Generating yellow and red emissions by co-doping Mn2+ to substitute for Ca2+ and Sc3+ sites in Ca3Sc2Si3O12:Ce3+ green emitting phosphor for white LED applications,” J. Mater. Chem. 21(41), 16379–16384 (2011). [CrossRef]  

8. M. Smedskjaer, J. Wang, and Y. Yue, “Tunable photoluminescence induced by thermal reduction in rare earth doped glasses,” J. Mater. Chem. 21(18), 6614–6620 (2011). [CrossRef]  

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

10. M. Peng, B. Sprenger, M. A. Schmidt, H. G. Schwefel, and L. Wondraczek, “Broadband NIR photoluminescence from Bi-doped Ba2P2O7 crystals: insights into the nature of NIR-emitting Bismuth centers,” Opt. Express 18(12), 12852–12863 (2010). [CrossRef]   [PubMed]  

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

12. 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]  

13. M. Peng and L. Wondraczek, “Photoluminescence of Sr2P2O7:Bi2+ as a red phosphor for additive light generation,” Opt. Lett. 35(15), 2544–2546 (2010). [CrossRef]   [PubMed]  

14. C. Lin, Y. Luo, H. You, Z. Quan, J. Zhang, J. Fang, and J. Lin, “Sol-gel-derived BPO4/Ba2+ as a new efficient and environmentally-friendly bluish-white luminescent material,” Chem. Mater. 18(2), 458–464 (2006). [CrossRef]  

15. Z. Xia, J. Zhuang, A. Meijerink, and X. Jing, “Host composition dependent tunable multicolor emission in the single-phase Ba2Ln1-zTbz(BO3)2Cl:Eu phosphors,” Dalton Trans. 42(18), 6327–6336 (2013). [CrossRef]   [PubMed]  

16. Y. Zhao, C. Riemersma, F. Pietra, R. Koole, C. M. Donegá, and A. Meijerink, “High-temperature luminescence quenching of colloidal quantum dots,” ACS Nano 6(10), 9058–9067 (2012). [CrossRef]   [PubMed]  

17. R. Xie, N. Hirosaki, T. Suehiro, F. Xu, and M. Mitomo, “A simple, efficient synthetic route to Sr2Si5N8: Eu2+-based red phosphors for white light-emitting diodes,” Chem. Mater. 18(23), 5578–5583 (2006). [CrossRef]  

18. M. Brik and A. Srivastava, “On the optical properties of the Mn4+ ion in solids,” J. Lumin. 133, 69–72 (2013). [CrossRef]  

19. Y. X. Pan and G. K. Liu, “Enhancement of phosphor efficiency via composition modification,” Opt. Lett. 33(16), 1816–1818 (2008). [CrossRef]   [PubMed]  

20. S. Okamoto and H. Yamamoto, “Luminescent-efficiency Improvement by alkaline-earth fluorides partially replacing MgO in 3.5MgO-0.5MgF2-GeO2:Mn4+ deep-red phosphors for light emitting diodes,” J. Electrochem. Soc. 157(3), J59–J63 (2010). [CrossRef]  

21. J. Park, G. Kim, and Y. Kim, “Luminescent properties of CaAl4O7 powders doped with Mn4+ ions,” Ceram. Int. 39, S623–S626 (2013). [CrossRef]  

22. M. Peng, X. Yin, P. Tanner, C. Liang, P. Li, Q. Zhang, and J. Qiu, “Orderly-layered tetravalent manganese-doped strontium aluminate Sr4Al14O25:Mn4+: an efficient red phosphor for warm white light emitting diodes,” J. Am. Ceram. Soc. n/a (2013), doi:. [CrossRef]  

23. N. Da, M. Peng, S. Krolikowski, and L. Wondraczek, “Intense red photoluminescence from Mn2+-doped (Na+; Zn2+) sulfophosphate glasses and glass ceramics as LED converters,” Opt. Express 18(3), 2549–2557 (2010). [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 (2)

Fig. 1
Fig. 1 (a) XRD patterns of CaAl4O7: 0.1%Mn prepared at different temperature in air with 2%H3BO3 as flux and reference pattern of CaAl4O7 (ICSD #34487) at the bottom. (b) XRD pattern (-o-) of the sample of CaAl4O7: 0.1%Mn prepared at 1400°C, Rietveld refining results (), Bragg reflections (|) and the profile difference between experimental and calculated values (). Inset shows the lattice cell of the compound viewed along c; blue ball: Ca, green ball: Al, red ball: O; purple polyhedra: CaO7, dark cyan polyhedra: AlO4. (c) Exemplary excitation and emission spectra of CaAl4O7: 0.1%Mn4+ prepared at 1400 (curves 1 and 2) and 1600 °C (curves 3 and 4). Calcium sites which could stabilize and accommodate Mn4+: (d) ideal CaO6 octahedron in the compound of CaO, where Ca-O bond length is 2.407(6)Å; (e) CaO7 polyhedron in CaAl4O7, where one Ca-O(1) bond is 2.488(9)Å, two Ca-O(2) bond are 2.367(6)Å, two Ca-O(2) bond are 2.925(6)Å and two Ca-O(3) bond are 2.310(2)Å.
Fig. 2
Fig. 2 (a) Excitation spectra (λem = 652nm), (b) emission spectra (λex = 325nm) and (c) decay curves (λem = 652nm, λex = 325nm) of CaAl4O7: 0.1%Mn4+ at different temperatures as indicated. (d) The temperature dependence of the integrated emission intensities (light green ellipse pattern: “red star” for λex = 325nm; “blue ball” for λex = 470nm) and lifetimes (light blue ellipse pattern: “red star” for λex = 325nm and λem = 652nm; “blue ball” for λex = 470nm and λem = 652nm). The green line through the data points are fits to τ(T) = τ0/(1 + τ0Ce-Ea/kT) with τ0 = 1435μs, C = 1.47, and Ea = 0.196eV. The goodness of fitting is 96.0%.
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.