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

Theoretical analysis and experiment on Eu reduction in alumina optical materials

Open Access Open Access

Abstract

Density functional theory simulation with a U correction (DFT + U) was adopted to predict the proper experimental conditions required to generate divalent europium (Eu2+) in α-Al2O3 host material consolidated via vacuum consolidation. Accordingly, samples of 0.1 at% Eu2+:α-Al2O3 were obtained by gelcasting, followed by high vacuum sintering process (<10−3 Pa). The room temperature photoluminescence excitation and emission spectra of this composition were examined and characteristic broad band positioned in the blue part of spectrum was detected under UV excitation. This luminescence decay trace had two-exponential character and the average decay time of 0.1 at% Eu2+:α-Al2O3 was 0.080ms.

© 2016 Optical Society of America

1. Introduction

Optical materials based on rare earth activation center, such as neodymium and europium ions, are critical for a wide range of applications including phosphors [1–3], lasers [4], and scintillator materials [5, 6] for ionizing radiation detection [7]. Their optical properties are very sensitive to the host lattice, i.e., crystal field strength, covalency, cation size, etc., and therefore considerable attention has been paid to the search for host materials with high performance and reliability. Many of the recently developed hosts are halides [6, 8–14]. They provide high light output and reasonable scintillation decay time, but suffer from sensitivity to moisture and low mechanical strength, which has substantially limited their applications, especially as scintillators. Oxide ceramics, on the other hand, possess high chemical stability and high transparency, and thus can be ideal candidates for the next generation of scintillator materials. For example, it was recently reported that Eu-doped garnet (Y3Al5O12) scintillators show very promising scintillation properties [15]. The widely available α-Al2O3 has also been considered for cryogenic phonon detector designated for the Dark Matter experiment [16, 17]. Indeed, these materials can also provide good mechanical properties, excellent refractory character, high chemical stability and optical transmission over a wide range of wavelengths.

However, efficient identification of high-performance host oxide with proper fabrication conditions is challenging. The chemical state of many rare earth activation centers (i.e., valence) is sensitive to the synthesis condition, and this is coupled with the stringent sintering conditions required to achieve transparency in polycrystalline oxide materials. Predictive method based on computations are also lacking, due to the difficulty in accurately capture the electronic behavior, particularly for the f-electrons, in different atomic environments, at realistically low doping concentration (tenths of a percent), and as a function of experimental conditions. It is therefore no surprise that determining the correct experimental conditions so far has been largely relying on trial and error. In this paper, we present a protocol to use ab-initio simulations based on density functional theory (DFT) to guide the experimental synthesis of rare earth doped transparent polycrystalline ceramics. Using Eu2+ doped α-Al2O3 as an example, we present the evidence that sophisticated DFT simulations can provide critical guidance for assessing experimental feasibility and selecting effective synthesis conditions. This provides the basis for efficient predictive tools based on computations that would enable high-throughput screening of the similar materials. For validation, Eu:Al2O3 ceramics were synthesized by gelcasting techniques and vacuum sintering method with photoluminescence properties and decay time characterized here.

2. Method

2.1 Density functional theory simulations with U correction (DFT + U)

The ab initio simulations were performed with VASP, a plane wave DFT package [18]. PBE exchange and correlation functional was used with the plane wave cutoff set to 500 eV. To properly describe the f-electrons of Eu ions, we applied the DFT + U method with an effective U value of 6 eV, according to several previous simulation studies of Eu-containing compounds [19–21]. We paid particular attention to the issues of unphysical sub-orbital occupations raised in recent DFT + U studies, which can lead to incorrect energetics and ionic structures in some materials [22–25]. With a simple scheme proposed recently, which involves ramping the U value iteratively during ionic relaxation [24, 25], we did not find sub-orbital occupations significantly more stable than the ones obtained by the standard method for the systems studied here, although more thorough investigation might be needed.

The target doping concentration of synthesized Eu:Al2O3 is 0.1 at. % is well below the dilute solution limit. For example, the probability of a Eu substitution have no other Eu substitution in a nearest neighbor configuration, i.e., all three nearest cation sites around the Eu are occupied by Al ions, is (99.9%)3 = 99.7% in a random distribution. Similarly, in the case of Eu2+ compensated with oxygen vacancy, the probability of a Eu dopant in the α-Al2O3 lattice (assuming six coordinated, see detailed discussions below) does not have neighboring oxygen vacancy is 99.8%. Therefore, it is necessary to consider the defects as isolated. We note, however, our simulations indeed found considerable association between certain defects, especially in the case of Eu2+ and oxygen vacancy (binding energy of 2.5 eV in a 3 × 3 × 1 supercell). Such association would inevitably begin to manifest when the doping concentration increases, and should be taken into consideration in heavily doped cases.

Within the plane wave DFT frame, the periodic system containing isolated defects must be dealt with care. Especially, charged defects, which are neutralized with electron jellium in the present study, can considerably slow the convergence against system size [26]. To resolve this issue, we adopted a correction scheme proposed by Hine et al [27]. In this scheme, a series of supercells of different sizes and shapes containing the same defect are simulated, and the corrected energy is obtained by extrapolating a linear relationship between the energy and the Madelung potential of each supercell (calculated as a function of the defect charge and cell geometry) to infinite system size. For each charged defect in the present study, we fully relaxed four hexagonal corundum supercells (2 × 2 × 1, 2 × 2 × 2, 3 × 3 × 1, 2 × 2 × 3) to extrapolate the corrected defect energy. For Eu3+ substitution, no addition charge is involved and the results from 3 × 3 × 1 supercell (270 atoms) are used. For all the structures, relaxation was conducted until forces on all the atoms are below 0.01 eV/Å.

2.2 Synthesis and characterization of Eu: Al2O3

0.1 at. % Eu: Al2O3 were synthesized by gelcasting technique. Vacuum sintering method were adopted here to provide a cost-effective route to manipulate the valence state of Eu in Eu:Al2O3. Eu was introduced in the form of Eu2O3 (99.5%, Sigma-Aldrich) and high-purity α-Al2O3 powder (>99%) (CR-10, Baikowski, Annecy, France, D50 = 0.45 μm) was used as the matrix material. After drying and de-binding of gelled Eu:Al2O3, the material is sintered in a vacuum furnace (10−3-10−4 Pa) at the temperature 1400-1850 °C for 8 hours with the final sample diameter around 3 mm and thickness in the range of 3-5 mm.

After sintering, the room temperature photoluminescence emission and excitation spectra of Eu: Al2O3 were recorded in an apparatus consisting of a Xenon lamp, two grating monochromators and a synchronous light detection system (Fluorolog-Tau-3, Jobin Yvon Inc. Horiba, Japan). The measurements were performed with internal corrections and compensation for the light source. Decay curves were also recorded using a pulsed Nd:YAG laser at 266 nm or 355 nm (Spectron Laser System SL802G, Rugby, UK) with a pulse energy of approximately 5 mJ, 10 Hz repetition rate and 5 ns duration.

3. Results and discussion

3. 1 Density functional theory simulations with U correction (DFT + U) of Eu2+/Eu3+:α-Al2O3

We first investigated the defect mechanism for Eu doping in α-Al2O3. Mössbauer measurement has indicated that the Eu3+ dopant in α-Al2O3 has a coordination number of six [28]. There have also been previous work hypothesizing that the Eu3+ dopant would substitute Al3+, rather than occupy interstitial sites, based on the bonding covalency and similar X-ray diffraction pattern between pure and doped alumina [28, 29]. In this study, we examined both substitution and interstitial mechanisms. The preferred one was found to be the substitution of Al3+ by Eu3+ as described by the following defect reaction in Eq. (1):

Eu2O3Al2O32EuAl×+3Oo

This should come as no surprise even though both Al3+ substitution and octahedral interstitial (tetrahedral interstitial site is too small for large ion such as Eu3+) provide the same coordination environment for Eu3+. The interstitial mechanism places Eu3+ adjacent to Al3+ ions, causing strong lattice distortions. It also needs to be charge compensated by O2- interstitials or Al3+ vacancies, incurring large energetic penalties.

Interestingly, Eu2+ dopant also prefers a similar substitution mechanism, even though both substitution and interstitial require other defects (considering the most likely mechanisms: one O2- vacancy per two Eu2+ substitutions versus one O2- interstitial per one Eu2+ interstitial). The defect reaction equation, describing the reduction of Eu3+ to Eu2+ in Al2O3 host lattice, can therefore be written as Eq. (2):

Eu2O3Al2O32Eu(2+)Al'+VO..+32O2(g)

It is worth noting that both isolated Eu3+ and Eu2+ substations have six-coordinated environments. For both cases, the first coordination shell is well-defined and similar to that of Al3+ in Al2O3, with two sets of Eu-O bond lengths. It can be seen that, in spite of the substantial size difference between europium and aluminum ions, isolated Eu dopant can be constricted by the host lattice and assume the environment of host cations, at least at a low dopant concentration. This can also apply to other host materials for selecting desired local environment for the luminescence dopant.

As both Eu3+ and Eu2+ would assume the same lattice site, the reduction reaction of Eu3+ to Eu2+ can be written as the following Eq. (3), by combining Eq. (1) and (2):

Eu2O3Al2O32Eu(2+)Al'+VO..+32O2(g)

The reduction enthalpy for this reaction can then be calculated in Eq. (4):

Er=E2Eu2++EVO..+μO(T,P)E2Eu3+

where (E2Eu2+E2Eu3+) is the difference in lattice energy between Eu2+ and Eu3+ doped structures, and μO(T,P) is calculated as:

μO(T,P)=12EO2+ΔμOΟ(T)+kBTln(pO2)

where EO2 is the energy of O2 molecule at 0K, which was calculated from DFT by placing an isolated O2 molecule in the simulation box. The standard chemical potential of oxygen at higher temperatures can be calculated by correcting EO2 with ΔμOΟ(T), which is obtained from thermodynamic tables [30]. The DFT + U simulations calculated the value of (E2Eu2++EVO..+12EO2E2Eu3+) as 3.50 eV. Using this value, the reduction level can be estimated as a function of oxygen partial pressure according to reaction kinetics theory. Ignoring the vibrational entropy contribution, the temperature-dependence equilibrium constant follows the Boltzmann distribution:

[EuEu']2[VO..][EuEu×]2[OO×]=exp(ErkT)

where [EuEu'], [EuEu×], [VO..], [OO×] are equilibrium concentrations of Eu2+, Eu3+, oxygen vacancies, and oxygen ions, respectively, and they satisfy:

[EuEu']=2[VO..]

and, considering that [VO..]+[OO×] gives the total concentration of oxygen sites in the host Al2O3:

[EuEu']+[EuEu×][VO..]+[OO×]=[EuEu']+[EuEu×]32[AlAl×]=23cEu 

where cEu is the doping concentration of Eu. Solving Eqs. (6), (7) and (8), the Eu2+ concentration can be obtained. The results are plotted in Fig. 1. The Eu2+ concentration shows a dependence on oxygen partial pressure with a characteristic exponent of −1/6. At high oxygen partial pressure, reduction is limited. Typical vacuum furnaces for high temperature sintering can maintain oxygen partial pressure in the range of 0.21 × (10−3-10−4) Pa, which is marked with the dotted lines. It can be seen that, for p(O2) in this range, a high temperature above 1700℃ can reduce majority (>95%) of the Eu3+ to Eu2+. Such a high temperature can also encourage oxygen diffusion so that the reduction equilibrium can be reached efficiently.

 figure: Fig. 1

Fig. 1 Calculated Eu2+ concentration (in log10(c), where c = [Eu2+]/[Eu]) versus oxygen partial pressure (in log10(PO), where PO is oxygen partial pressure in Pa) in the temperature range between 1400~1900 °C. Vertical dashed lines mark the range of oxygen partial pressure used in vacuum synthesize. The horizontal dashed line indicates c = 95%.

Download Full Size | PDF

The electronic band structures of Eu3+ and Eu2+ doped α-Al2O3 are also calculated from the DFT + U simulations, as shown in Fig. 2. We observe no substantial differences between the electronic bands contributed by the host material, α-Al2O3 between the two cases. The defect states contributed by the dopants in the band gap are well defined, and show no obvious characteristics of the α-Al2O3 symmetry. Therefore, we expect the luminescence in both cases to retain its characteristic spectrum. However, it is worth noting that the effect of oxygen vacancy, especially the association between the dopant and vacancy, is not considered here. This could potentially cause alterations of the spectrum of the Eu2+ doped material, which warrants further investigations.

 figure: Fig. 2

Fig. 2 Band structure of Eu3+ (a) and Eu2+ (b) doped alumina. The Fermi energy is set to 0 eV. For comparison, the two plots are arranged so that the energy levels contributed mainly by alumina are aligned. For clarity, unoccupied and occupied bands are shown in blue. Spin up and spin down levels are plotted with solid and dotted lines, respectively.

Download Full Size | PDF

3.2 Photoluminescence characterization of Eu2+/Eu3+:Al2O3

3.2.1 Room temperature photoluminescence emission and excitation of Eu2+/Eu3+:Al2O3

Room temperature photoluminescence excitation and emission spectra of Eu: Al2O3 were recorded in Fig. 3 and Fig. 4. The photoluminescence excitation spectrum of Eu2+: Al2O3 (Fig. 3) shows two over-lapped excitation peaks with the centers at 277 nm and 330 nm, coming from the emission behavior of Eu2+. The blue emission of Eu:Al2O3 comes from the transition of Eu2+ between 4f65d1 and 4f7 (8S7/2) (Eu2+(4f7)+hνEu2+*(4f65d1)). Such f-d transition and broadband emission are usually strongly dependent on the local atomic environment of the activation center. Gaussian fitting analysis of the excitation spectrum of vacuum-sintered Eu2+: Al2O3 (monitored at 430 nm, Fig. 3) showed two broad excitation peaks at 277 nm and 330 nm, respectively, indicating that the Eu2+ ions are located in low-symmetry sites with a broad crystal field distribution [31] and a large field splitting of 5d band [29]. Such excitation spectrum is different from those measured for Eu2+:Al2O3 powder [29] and Eu2+:Al2O3 film [32]. It is possible that the fabrication process has led to changes in the local symmetry of the Eu2+ activator. As shown by the DFT + U calculations, the Eu2+ dopants substitute Al3+, accompanied by oxygen vacancies. The oxygen vacancies may form F+-centers [33], and contribute to broad excitation and emission spectra in the UV region. Additionally, strong association may exist between dopants and vacancies. Under favorable synthesis conditions, e.g. slow cooling rate, defect clusters such as Eu(2+)Al'VO.. may be obtained, which leads to a change in the local symmetry of the Eu2+ activator and, consequently, the crystal field felt by Eu2+ ions. Indeed, the bulk samples may experience slower cooling in the center than powders and films even when the same nominal cooling rate is used [34]. Other factors, such as phase separation, defect segregation at grain boundary, etc., could also contribute to the changes in the local atomic environment of the activation centers, which warrants further investigations.

 figure: Fig. 3

Fig. 3 Room temperature photoluminescence excitation spectra of vacuum-sintered 0.1at% Eu2+: Al2O3 monitored at 430 nm with corresponding Gaussian fitting curves.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Room temperature photoluminescence emission spectra of vacuum-sintered 0.1at% Eu2+: Al2O3 and air-sintered 0.1at% Eu3+: Al2O3 excited with 277 nm and 330 nm (dark line: Eu2+:Al2O3; red line: Eu3+:Al2O3).

Download Full Size | PDF

Air-sintered Eu:Al2O3 was also examined under excitation by UV light (277 nm), which showed the characteristic photoluminescence peaks of Eu3+ at 588 nm (5D0-7F1) and 611 nm (5D0-7F2) [35–37]. In comparison, under excitation with UV light (277 nm and 330 nm), vacuum sintered Eu:Al2O3 only showed one broad peak in the wavelength range of 330-600 nm with a peak center near 430 nm but none of the characteristic peaks of Eu3+. This indicates that the reduction of Eu3+ —› Eu2+ in the Al2O3 host was achieved by vacuum sintering at temperatures of 1750-1850 °C, as predicted by the DFT + U calculations.

3.2.2 Room temperature photoluminescence emission spectra of Eu: Al2O3 sintered at varying vacuum-sintering temperature

To further validate the DFT + U calculations, samples were vacuum sintered at different temperatures using the same procedure. The measured emission spectra were shown in Fig. 5. As expected, the intensity of the characteristic Eu3+ photoluminescence peak at 611 nm (5D0-7F2) decreases with increasing vacuum sintering temperature. In comparison, the emission intensity of Eu2+ in the blue light region increases with higher vacuum sintering temperature. The presence of Eu3+ is evident at and below 1700 °C, suggesting the calculated Eu2+ concentration may be overestimated, which requires further quantification. Nonetheless, the DFT + U calculations provided adequate guidance for selecting the correct sintering condition. It should also be noted that, low porosity and good sintering quality in the samples, a prerequisite for optical applications that require high transmittance, have been obtained at temperatures above 1700 °C.

 figure: Fig. 5

Fig. 5 Photoluminescence spectra of vacuum-sintered 0.1at% Eu2+: Al2O3 at room temperature excited with 277 nm as a function of vacuum sintering temperature (normalized to the same height at the spectra maxima).

Download Full Size | PDF

3.2.3 Luminescence decay properties of Eu2+:Al2O3

The luminescence decay curve of the vacuum sintered (1750°C-1850°C) 0.1 at. % Eu2+:Al2O3 measured at room temperature is shown in Fig. 6, monitored at 490 nm with an excitation wavelength of 266 nm. The luminescence decay curves were fitted by double exponentials (I = A1* exp (-t12) + A2* exp (-t22)) with the equation parameter shown in Fig. 6. The decay time of the Eu2+ d-f transition is on the order of microseconds. The average decay time of the Eu2+ in Al2O3 after vacuum-sintering was calculated from the fitting process to be approximately 0.080 ms.

 figure: Fig. 6

Fig. 6 Decay curve of 0.1at% Eu2+: Al2O3 (sintering temperature: 1750°C-1850°C) monitored at 490 nm with the excitation 266 nm at room temperature.

Download Full Size | PDF

4. Summary

Overall, it can be seen that ab-initio calculations employing an improved DFT + U method for rare earth elements successfully guided the fabrication of Eu in α-Al2O3 host. The simulation results showed that Eu2+:Al2O3 can be fabricated by gelcasting and a subsequent vacuum sintering process under a vacuum atmosphere of 10−3-10−4 Pa in the temperature region of 1750°C-1850 °C. The resultant materials show strong photoluminescence behavior with an average lifetime of 0.080 ms. Additionally, the DFT + U calculations also provided details regarding the lattice distortions and defect associations. The presented method can be easily applied to other rare earth dopants and other host materials.

Acknowledgments

We gratefully acknowledge the US Air Force Office of Scientific Research (AFSOR) (contract FA9550-14-1-0155) for funding and supporting this research.

References and links

1. G. Blasse and A. Bril, “Fluorescence of Eu2+ activated alkaline-earth aluminates,” Philips Res. Rep. 23, 201–206 (1968).

2. 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]  

3. 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]  

4. E. Bayer and G. Schaack, “Two-photon absorption of CaF2:Eu2+,” Phys. Status Solidi41(2), 827–835 (1970) (b). [CrossRef]  

5. G. Patton, F. Moretti, A. Belsky, K. Al Saghir, S. Chenu, G. Matzen, M. Allix, and C. Dujardin, “Light yield sensitization by X-ray irradiation of the BaAl4O7:Eu(2+)ceramic scintillator obtained by full crystallization of glass,” Phys. Chem. Chem. Phys.16(45), 24824–24829 (2014) (pp.). [CrossRef]   [PubMed]  

6. Y. Wu, L. A. Boatner, A. C. Lindsey, M. Zhuravleva, S. Jones, J. D. Auxier II, H. L. Hall, and C. L. Melcher, “Defect Engineering in SrI2:Eu2+ Single Crystal Scintillators,” Cryst. Growth Des. 15(8), 3929–3938 (2015). [CrossRef]  

7. M. S. Alekhin, J. T. de Haas, K. W. Kramer, I. V. Khodyuk, L. de Vries, and P. Dorenbos, “Scintillation properties and self-absorption in SrI2: Eu2,+” in Nuclear Science Symposium Conference Record (NSS/MIC) (IEEE, 2010), pp. 1589–1599.

8. N. Cherepy, B. Sturm, O. B. Drury, T. Hurst, S. Sheets, L. Ahle, C. Saw, M. Pearson, S. A. Payne, A. Burger, L. A. Boatner, J. O. Ramey, E. V. van Loef, J. Glodo, R. Hawrami, W. M. Higgins, K. S. Shah, and W. W. Moses, “SrI2 scintillator for gamma ray spectroscopy,” Proc. SPIE 7449, 74490F (2009). [CrossRef]  

9. G. A. Appleby, A. Edgar, and G. V. M. Williams, “Structure and photostimulated luminescent properties of Eu-doped M2BaX4 (M=Cs, Rb; X=Br, Cl),” J. Appl. Phys. 96(11), 6281–6285 (2004). [CrossRef]  

10. G. Gundiah, E. D. Bourret-Courchesne, G. Bizarri, S. M. Hanrahan, A. Chaudhry, A. Canning, W. W. Moses, and S. E. Derenzo, “Scintillation properties of Eu2+-activated barium fluoroiodide,” in Nuclear Science Symposium Conference Record (NSS/MIC) (IEEE, 2009), 1575–1578.

11. E. D. Bourret-Courchesne, G. Bizarri, R. Borade, Z. Yan, S. M. Hanrahan, G. Gundiah, A. Chaudhry, A. Canning, and S. E. Derenzo, “Eu2+-doped Ba2CsI5, a new high-performance scintillator,” Nucl. Instrum. Methods Phys. Res. Sect. A 612(1), 138–142 (2009). [CrossRef]  

12. M. Kirm, Y. Chen, S. Neicheva, K. Shimamura, N. Shiran, M. True, and S. Vielhauer, “VUV spectroscopy of Eu doped LiCaAlF6 and LiSrAlF6 crystals,” Phys. Status Solidi, C Conf. Crit. Rev. 2(1), 418–421 (2005). [CrossRef]  

13. A. Fukabori, “Comparative analysis of scintillation characteristics derived from different emission mechanisms in BaCl2,” J. Appl. Phys. 117(15), 153106 (2015). [CrossRef]  

14. J. Glodo, H. Lingertat, C. Brecher, K. S. Shah, and A. Lempicki, “A new ceramic scintillator for neutron detection: CaF2: Eu2+/6LiF,” in Nuclear Science Symposium Conference Record (NSS/MIC) (IEEE, 2005) 1, 112–115.

15. M. Sugiyama, T. Yanagida, Y. Fujimoto, Y. Yokota, A. Ito, M. Nikl, T. Goto, and A. Yoshikawa, “Basic study of Eu2+-doped garnet ceramic scintillator produced by spark plasma sintering,” Opt. Mater. 35(2), 222–226 (2012). [CrossRef]  

16. V. B. Mikhailik, H. Kraus, M. Balcerzyk, W. Czarnacki, M. Moszyński, M. S. Mykhaylyk, and D. Wahl, “Low-temperature spectroscopic and scintillation characterization of Ti-doped Al2O3,” Nucl. Instrum. Methods Phys. Res, Sect. A 546, 523–534 (2005).

17. G. Angloher, M. Bruckmayer, C. Bucci, M. Bühler, S. Cooper, C. Cozzini, P. DiStefano, F. Von Feilitzsch, T. Frank, D. Hauff, T. Jagemann, J. Jochum, V. Jörgens, R. Keeling, H. Kraus, M. Loidl, J. Marchese, O. Meier, U. Nagel, F. Pröbst, Y. Ramachers, A. Rulofs, J. Schnagl, W. Seidel, I. Sergeyev, M. Sisti, M. Stark, S. Uchaikin, L. Stodolsky, H. Wulandari, and L. Zerle, “Limits on WIMP dark matter using sapphire cryogenic detectors,” Astropart. Phys. 18(1), 43–55 (2002). [CrossRef]  

18. C. A. Arguello, D. L. Rousseau, and S. P. S. Porto, “First-order Raman effect in wurtzite-type crystals,” Phys. Rev. 181(3), 1351–1363 (1969). [CrossRef]  

19. J. L. Bettis, M. H. Whangbo, J. Köhler, A. Bussmann-Holder, and A. R. Bishop, “Lattice dynamical analogies and differences between SrTiO3 and EuTiO3 revealed by phonon-dispersion relations and double-well potentials,” Phys. Rev. B 84(18), 184114 (2011). [CrossRef]  

20. J. Gegner, T. C. Koethe, H. Wu, Z. Hu, H. Hartmann, T. Lorenz, T. Fickenscher, R. Pöttgen, and L. H. Tjeng, “Electronic structure of RAuMg and RAgMg (R=Eu,Gd,Yb),” Phys. Rev. B 74(7), 073102 (2006). [CrossRef]  

21. K. Koga, H. Anno, K. Akai, M. Matsuura, and K. Matsubara, “First-principles study of electronic structure and thermoelectric properties for guest substituted clathrate compounds Ba6R2Au6Ge40 (R=Eu or Yb),” Mater. Trans. 48(8), 2108–2113 (2007). [CrossRef]  

22. J. P. Allen and G. W. Watson, “Occupation matrix control of d- and f-electron localisations using DFT + U,” Phys. Chem. Chem. Phys. 16(39), 21016–21031 (2014). [CrossRef]   [PubMed]  

23. B. Dorado, B. Amadon, M. Freyss, and M. Bertolus, “DFT+U calculations of the ground state and metastable states of uranium dioxide,” Phys. Rev. B 79(23), 235125 (2009). [CrossRef]  

24. B. Meredig, A. Thompson, H. A. Hansen, C. Wolverton, and A. van de Walle, “Method for locating low-energy solutions within DFT+U,” Phys. Rev. B 82(19), 195128 (2010). [CrossRef]  

25. B. Wang, X. N. Xi, and A. N. Cormack, “Chemical Strain and Point Defect Configurations in Reduced Ceria,” Chem. Mater. 26(12), 3687–3692 (2014). [CrossRef]  

26. H. P. Komsa, T. T. Rantala, and A. Pasquarello, “Finite-size supercell correction schemes for charged defect calculations,” Phys. Rev. B 86(4), 045112 (2012). [CrossRef]  

27. N. D. M. Hine, K. Frensch, W. M. C. Foulkes, and M. W. Finnis, “Supercell size scaling of density functional theory formation energies of charged defects,” Phys. Rev. B 79(2), 024112 (2009). [CrossRef]  

28. J. Li, Y. Shi, J. Gong, and G. Chen, “Mössbauer study of amorphous Al2O3:Eu3+,” J. Mater. Sci. Lett. 16(9), 743–744 (1997). [CrossRef]  

29. N. Rakov and G. S. Maciel, “Photoluminescence analysis of α-Al2O3 powders doped with Eu3+ and Eu2+ ions,” J. Lumin. 127(2), 703–706 (2007). [CrossRef]  

30. M. W. Chase, JANAF thermochemical tables, American Chemical Society and the American Institute of Physics for the National Bureau of Standards, New York, 1985.

31. G. Blasse, W. Wanmaker, J. Ter Vrugt, and A Bril, “Fluorescence of Eu2+ activated alkaline-earth aluminates,” Philips Res. Rep 23, 189–200 (1968).

32. A. Pillonnet, A. Pereira, O. Marty, and C. Champeaux, “Valence state of europium doping ions during pulsed-laser deposition,” J. Phys. D Appl. Phys. 44(37), 375402 (2011). [CrossRef]  

33. S. Mochizuki, T. Nakanishi, Y. Suzuki, and K. Ishi, “Reversible photoinduced spectral change in Eu2O3 at room temperature,” Appl. Phys. Lett. 79(23), 3785–3787 (2001). [CrossRef]  

34. J. A. Zasadzinski, “A new heat transfer model to predict cooling rates for rapid freezing fixation,” J. Microsc. 150(2), 137–149 (1988). [CrossRef]  

35. G. Hirata, N. Perea, M. Tejeda, J. A. Gonzalez-Ortega, and J. McKittrick, “Luminescence study in Eu-doped aluminum oxide phosphors,” Opt. Mater. 27(7), 1311–1315 (2005). [CrossRef]  

36. N. Rakov, G. S. Maciel, W. B. Lozano, and C. B. de Araújoa, “Investigation of Eu3+ luminescence intensification in Al2O3 powderscodoped with Tb3+ and prepared by low-temperature direct combustion synthesis,” Appl. Phys. Lett. 88(8), 081908 (2006). [CrossRef]  

37. N. Rakov, F. E. Ramos, G. Hirata, and M. Xiao, “Strong photoluminescence and cathodoluminescence due to f-f transitions in Eu3+ doped Al2O3 powders prepared by direct combustion synthesis and thin films deposited by laser ablation,” Appl. Phys. Lett. 83(2), 272–274 (2003). [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 Calculated Eu2+ concentration (in log10(c), where c = [Eu2+]/[Eu]) versus oxygen partial pressure (in log10(PO), where PO is oxygen partial pressure in Pa) in the temperature range between 1400~1900 °C. Vertical dashed lines mark the range of oxygen partial pressure used in vacuum synthesize. The horizontal dashed line indicates c = 95%.
Fig. 2
Fig. 2 Band structure of Eu3+ (a) and Eu2+ (b) doped alumina. The Fermi energy is set to 0 eV. For comparison, the two plots are arranged so that the energy levels contributed mainly by alumina are aligned. For clarity, unoccupied and occupied bands are shown in blue. Spin up and spin down levels are plotted with solid and dotted lines, respectively.
Fig. 3
Fig. 3 Room temperature photoluminescence excitation spectra of vacuum-sintered 0.1at% Eu2+: Al2O3 monitored at 430 nm with corresponding Gaussian fitting curves.
Fig. 4
Fig. 4 Room temperature photoluminescence emission spectra of vacuum-sintered 0.1at% Eu2+: Al2O3 and air-sintered 0.1at% Eu3+: Al2O3 excited with 277 nm and 330 nm (dark line: Eu2+:Al2O3; red line: Eu3+:Al2O3).
Fig. 5
Fig. 5 Photoluminescence spectra of vacuum-sintered 0.1at% Eu2+: Al2O3 at room temperature excited with 277 nm as a function of vacuum sintering temperature (normalized to the same height at the spectra maxima).
Fig. 6
Fig. 6 Decay curve of 0.1at% Eu2+: Al2O3 (sintering temperature: 1750°C-1850°C) monitored at 490 nm with the excitation 266 nm at room temperature.

Equations (8)

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

E u 2 O 3 A l 2 O 3 2E u Al × +3 O o
E u 2 O 3 A l 2 O 3 2Eu (2+) Al ' + V O .. + 3 2 O 2 (g)
E u 2 O 3 A l 2 O 3 2Eu (2+) Al ' + V O .. + 3 2 O 2 (g)
E r = E 2E u 2+ + E V O .. + μ O (T,P) E 2E u 3+
μ O ( T,P )= 1 2 E O 2 +Δ μ O Ο (T)+ k B Tln( p O 2 )
[ E u Eu ' ] 2 [ V O .. ] [ E u Eu × ] 2 [ O O × ] =exp( E r kT )
[ E u Eu ' ]=2[ V O .. ]
[ E u Eu ' ]+[ E u Eu × ] [ V O .. ]+[ O O × ] = [ E u Eu ' ]+[ E u Eu × ] 3 2 [ A l Al × ] = 2 3 c Eu  
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.