Eu2+ inserted in β-Si3xAlxOxN4x is a material that shows exceptional promise as a green-emitting phosphor. Synchrotron X-ray and neutron scattering, in conjunction with first-principles calculations and Eu L3 X-ray absorption measurements, yield a consistent picture of the composition, and the favorable position for Eu2+ substitution in the crystal structure. The Debye temperature ΘD, which is a proxy for structural rigidity relating to effectiveness as a phosphor, is very high for the starting β-Si3N4 framework and is determined to decrease only slightly for the small amounts of Al3+ and O2− co-substitution that are required for charge balance associated with Eu2+ insertion.

The continued replacement of traditional illumination sources with solid-state lighting (SSL) is driven by their exceptional efficiency compared to incandescent and fluorescent lamps and the absence of toxic-heavy metals, such as mercury.1–3 One of many strategies for assembling a SSL device is to integrate ultraviolet (UV) or near-UV light emitting diodes (LED) with phosphors emitting in the blue, green, and red regions of the visible spectrum.2,3 Eu2+ inserted in β-Si3xAlxOxN4x is a material that shows exceptional promise as a green-emitting phosphor for such applications.4,5 The compound crystallizes in the β-Si3N4 crystal structure (space group 163, P63), shown in Fig. 1(a), with three-dimensional corner-sharing of (Si,Al)(O,N)4 tetrahedra. This high symmetry structure contains one crystallographically independent cation position (Wyckoff 6c) and two independent anion positions (Wyckoff 6c and 2b).

FIG. 1.

(a) The β-SiAlON framework, depicting a corner-sharing, three-dimensional tetrahedral network that provides a highly rigid host structure, projected nearly down the c-axis. (b) A contour plot of the excitation (y-axis) and emission (x-axis) attests to the broad excitation spectrum and near constant emission wavelength.

FIG. 1.

(a) The β-SiAlON framework, depicting a corner-sharing, three-dimensional tetrahedral network that provides a highly rigid host structure, projected nearly down the c-axis. (b) A contour plot of the excitation (y-axis) and emission (x-axis) attests to the broad excitation spectrum and near constant emission wavelength.

Close modal

Incorporating Eu2+ in the β-SiAlON host produces an ideal excitation and emission spectrum for use in SSLs, as highlighted by the contour plot in Fig. 1(b). The excitation band is strong and broad, spanning from the UV (<300 nm) to the blue-green region of the visible spectrum (∼475 nm), allowing excitation by a variety of commercial LEDs. Intense emission, with quantum yields that are 80% or greater, occurs in the green region of the visible spectrum (λem = 540 nm) and corresponds to the allowed Eu2+ 5d to 4f (8Hj8S7/2) transition.

The substitution of the luminescent center in most rare-earth inserted phosphors occurs on an obvious crystallographic site based on size and valence. In β-SiAlON, a substitution site is not obvious due to the size mismatch between Eu2+ and the cations in the structure. As a result, the inclusion of the rare-earth must occur at an interstitial, and the Si:Al or N:O ratios must change to maintain electroneutrality. Rare-earth solubility is limited (<2 mol. %),6 making direct identification of the interstitial site extremely challenging by traditional characterization techniques like X-ray diffraction. Scanning transmission electron microscopy (STEM) has tentatively suggested that Eu2+ occupies the large hexagonal channels along the [001] direction coordinated by SiN4 tetrahedra (Wyckoff position 2a).7 The structural studies carried out here are intended to enhance understanding of such Eu2+ insertion in this unique phosphor.

Recent research has suggested that highly efficient phosphors, such as β-SiAlON:Eu2+, contain densely packed polyhedral units and high symmetry, which result in structures8,9 that limit photoluminescence quenching and other non-radiative relaxation pathways. Because structural rigidity in a solid is rather subjective, a proxy such the Debye temperature (ΘD) proves very useful.10 In the case of stoichiometric β-Si3N4, ΘD is experimentally measured to be very high, close to 900 K, which indicates great promise for luminescence applications.11 However, the inclusion of the luminescent center (Eu2+) to produce the phosphor requires co-substitution of Al3+ and O2−, potentially impacting ΘD. A second goal reported here is to demonstrate that the changes in ΘD can be monitored by first-principles density functional theory (DFT).

β-SiAlON:Eu2+ can be readily prepared using many synthetic methods.12–14 However, the concentration of Al3+ and O2− can be strictly controlled using a high-pressure and high-temperature route. For the studies carried out here, a mixture of α-Si3N4, Al2O3, and Eu2O3 powders was ground with an alumina mortar and pestle following the stoichiometric ratio β-Si3xAlxOxN4x:Eu2+. Powders were reacted in a BN crucible at 1950 °C for 12 h under 0.92 MPa of N2 (purity >99.9995%). The products were then annealed for 8 h in an Ar atmosphere, ground to a fine powder, and washed with a mixture of HNO3 and HF.

This route produces a pure-phase product for samples containing Eu2+ as well as Eu2+-free compounds, as identified by a Rietveld refinement of high-resolution synchrotron X-ray powder diffraction data [Fig. 2(a)] measured at beamline 11-BM, Advanced Photon Source.15–17 Due to the similar X-ray scattering power of the substituted cations and anions (e.g., Al3+ for Si4+ and O2− for N3−), the total composition was not refined based on the X-ray diffraction data. Fortunately, the neutron scattering path lengths of N (9.36 fm) and O (5.80 fm) are sufficiently different to unequivocally distinguish composition and site occupancy of the anions.18 Because Eu has a large neutron absorption cross-section (4530 b for 0.0253 eV), a sample was prepared following the same synthetic protocol, maintaining the total composition while omitting Eu2+ to prevent absorption. Neutron powder diffraction was performed at NPDF19 (Los Alamos Neutron Science Center, Los Alamos National Laboratory); time-of-flight data were collected at 295 K using four sets of detector banks, located at ±40°,±90°,±119°, and ±140°.

FIG. 2.

Combined Rietveld refinement with (a) powder synchrotron X-ray diffraction and (b) neutron powder diffraction. Neutron diffraction is required to accurately refine the N:O ratio on the two crystallographic anion sites. The determined composition is Si2.875Al0.125O0.16(1)N3.84(1).

FIG. 2.

Combined Rietveld refinement with (a) powder synchrotron X-ray diffraction and (b) neutron powder diffraction. Neutron diffraction is required to accurately refine the N:O ratio on the two crystallographic anion sites. The determined composition is Si2.875Al0.125O0.16(1)N3.84(1).

Close modal

Rietveld refinement of neutron scattering data confirms the incorporation of a small oxygen concentration (∼5%) on both crystallographic anion sites [Fig. 2(b)]. At such small O2− content, there is not an obvious site preference. Although the anion concentration was refined, the scattering contrast between Al3+ (3.45 fm) and Si4+ (4.15 fm) is not sufficient to allow refinement of the cations, so the ratio of cations was fixed as the nominally loaded composition. The final refined composition was Si2.875Al0.125O0.16(1)N3.84(1). Comparing the refined anionic O2−:N3− ratio with the loaded cation Si4+:Al3+ ratio indicates that the as prepared compound is nearly charge neutral within standard deviation.

To determine the substitution site of Eu2+ in the bulk, a combined approach of first-principles computational modeling and X-ray absorption spectroscopy was employed. DFT calculations based on the Vienna ab initio Simulation Package (VASP)20,21 were used to generate the potential energy surface for β-Si3N4 (space group P63), as illustrated in Fig. 3. Visualization of the (001) plane reveals two favorable interstitial sites for Eu substitution: a 12-coordinate site (Wyckoff 2a) and a second 12-coordinate site located in the center of the unit cell (Wyckoff 6c). The polyhedral volume of the 6c site (∼20 Å3) is half the size of the 2a site (∼40 Å3), making it more likely for Eu2+ to incorporate in the large void.

FIG. 3.

The potential energy surface of the (001) plane for β-Si3N4 calculated using density functional theory. The low energy regions shown in violet indicate the most favorable interstitial position for the incorporation of Eu2+.

FIG. 3.

The potential energy surface of the (001) plane for β-Si3N4 calculated using density functional theory. The low energy regions shown in violet indicate the most favorable interstitial position for the incorporation of Eu2+.

Close modal

The location of Eu2+ in the crystal structure was experimentally confirmed by measuring the extended X-ray absorption fine structure (EXAFS) at the Eu L3 edge. Fluorescence spectra at 295 K and 20 K were collected at the Advanced Photon Source beamline 20-BM, and all data were processed using the Demeter software suite,22 which employs IFEFFIT and FEFF6. Additional experimental details are presented in the supplementary material.23 The absorption energy is consistent with Eu2+, and the X-ray absorption near-edge structure (XANES) shows that a majority (>90%) of Eu is in the divalent oxidation state.

The starting model for the EXAFS refinement was a DFT-optimized 2 × 2 × 4 supercell of β-Si3N4 with an interstitial Sr atom. Sr (rSr2+=1.31Å) was used instead of Eu (rEu2+=1.30Å) due to the difficulty of treating f electrons within DFT. The supercell was optimized until the residual forces were less than 10–2 eV/Å. Using this model, Δr of the outer-shell N was initially >0.1 Å, so FEFF calculations were performed on an expanded unit cell (with the unit cell parameters increased by 2%) to ensure proper representation of atomic potentials by muffin-tin radii implemented in FEFF.

Fitting the Eu L3 EXAFS demonstrates that the spectrum is consistent with Eu2+ in the distorted 2a interstitial site of β-Si3N4 (Fig. 4). The coordination environment and structural parameters around Eu2+ are presented in Table I. Notably, the Debye-Waller factors (σN2,σSi2) are quite large and do not change markedly between 295 K and 20 K. These large mean-squared displacements are dominated by static disorder in the structure owing to the distortion of the β-Si3N4 structure required to incorporate Eu2+. This is apparent in the intensity of the oscillations of the EXAFS, as the intensity is not enhanced at low temperature. Additionally, the first coordination shell (Na + Nb) determined by EXAFS is larger than the DFT-optimized structure. This finding is supported by analysis of the Eu2+ bond valence sum (BVS),24 which is 2.0 at 20 K (BVSDFT = 2.3), indicating optimal bonding.

FIG. 4.

Eu L3 EXAFS of β-SiAlON:Eu2+ collected at 295 K. The EXAFS is consistent with a structural model where Eu2+ lies in the large interstitial 2a site in the β-Si3N4 structure (χ2 = 9.5, χred2=1.6, and R-factor = 0.0020). (a) Reciprocal space k2-weighted EXAFS. (b) Real-space k2-weighted magnitude and real component of the EXAFS. The dashed green line represents the Fourier transform and fitting window.

FIG. 4.

Eu L3 EXAFS of β-SiAlON:Eu2+ collected at 295 K. The EXAFS is consistent with a structural model where Eu2+ lies in the large interstitial 2a site in the β-Si3N4 structure (χ2 = 9.5, χred2=1.6, and R-factor = 0.0020). (a) Reciprocal space k2-weighted EXAFS. (b) Real-space k2-weighted magnitude and real component of the EXAFS. The dashed green line represents the Fourier transform and fitting window.

Close modal
TABLE I.

Structural parameters around Eu2+ in β-SiAlON:Eu determined by modeling Eu L3 EXAFS collected at 295 K and 20 K. The coordination shell is shown with the corresponding number of equivalent atoms, N. x, y, and z are constrained parameters.

ShellNrDFT (Å)r (Å)Diff. (%)
295 Ka 
Na 2.492 2.577(8)x 3.4(3) 
Nb 2.807 2.898(8)x 3.2(3) 
Sia 2.822 2.861(3) 1.4(1) 
Sib 3.116 3.129(4) 0.4(1) 
3.813 3.97(2) 4.2(7) 
Si 3.986 4.002(8) 0.4(2) 
20 Kb 
Na 2.492 2.490(9) −0.1(4) 
Nb 2.807 2.89(1) 3.1(4) 
Sia 2.822 2.876(8) 1.9(3) 
Sib 3.116 3.108(5) −0.3(1) 
3.813 3.96(4) 4(1) 
Si 3.986 3.98(1) −0.2(3) 
ShellNrDFT (Å)r (Å)Diff. (%)
295 Ka 
Na 2.492 2.577(8)x 3.4(3) 
Nb 2.807 2.898(8)x 3.2(3) 
Sia 2.822 2.861(3) 1.4(1) 
Sib 3.116 3.129(4) 0.4(1) 
3.813 3.97(2) 4.2(7) 
Si 3.986 4.002(8) 0.4(2) 
20 Kb 
Na 2.492 2.490(9) −0.1(4) 
Nb 2.807 2.89(1) 3.1(4) 
Sia 2.822 2.876(8) 1.9(3) 
Sib 3.116 3.108(5) −0.3(1) 
3.813 3.96(4) 4(1) 
Si 3.986 3.98(1) −0.2(3) 
a

σN2=0.021(7)Å2;σSi2=0.011(3)Å2; BVS(Eu) = 1.84(5); S02 = 1.00(3)y; ΔE0 = 5.7(4) eVz.

b

σN2=0.013(6)Å2;σSi2=0.008(3)Å2; BVS(Eu) = 2.05(5); S02=1.00(3)y,ΔE0=2.8(4)eVz.

DFT calculations of elastic constants25 allowed ΘD to be approximated.10 The substitution of Al3+ and O2− in β-SiAlON was modeled following a previously determined site preference4,26–28 for β-Si3xAlxOxN4x (x = 0, 0.5, 1, 1.5, 2, and 2.5). Incorporating Al3+ and O2− decreases the structural rigidity, decreasing ΘD from 956 K for x = 0 to 710 K for x = 2.5 (Fig. 5). This is a result of phonon mode softening due to chemical disorder and the lower valences of Al3+ and O2− in the crystal structure. At the low level of substitution experimentally studied here (x = 0.125), the reduction in ΘD is not significant, with interpolation suggesting a value near ΘD = 920 K. Rietveld refinement of neutron diffraction data provides reliable atomic anisotropic displacement parameters (ADPs) necessary for an experimental estimate of ΘD at the high-temperature limit to compare with calculation.8,23,29 For this sample, it is estimated that ΘD ≈ 725 K. Although this is lower than calculated by DFT, significant static disorder and site mixing between Si4+ and Al3+ are missed by the Rietveld analysis and likely account for such a sharp drop in the experimental ΘD. Nevertheless, staying at low x values is clearly advantageous to maximize the quantum yield, and it has also been shown to narrow emission and improve color in β-SiAlON:Eu2+.30 

FIG. 5.

Evolution of the calculated ΘD as a function of Al and O co-substitution x in β-Si3xAlxOxN4x (x = 0, 0.5, 1, 1.5, 2, and 2.5). The horizontal line corresponds to x = 0.125, the composition experimentally studied here.

FIG. 5.

Evolution of the calculated ΘD as a function of Al and O co-substitution x in β-Si3xAlxOxN4x (x = 0, 0.5, 1, 1.5, 2, and 2.5). The horizontal line corresponds to x = 0.125, the composition experimentally studied here.

Close modal

In summary, the position of Eu2+ in the crystal structure was confirmed by fitting the Eu L3 EXAFS using a DFT-optimized starting structure; the Eu2+ luminescent center occupies a distorted 12-coordinate interstitial site in the parent β-Si3N4 crystal structure (Wyckoff 2a). There are minor differences in bond lengths determined by EXAFS and DFT, but the presence of local distortions around Eu2+ in the interstitial is captured well by both approaches. DFT calculations of ΘD have been employed as a proxy for structural rigidity and suggest that at the low level of (Al/O) substitution required to incorporate small amounts of Eu2+, the structure remains rigid and should retain high performance.

M.W.G. thanks the Natural Sciences and Engineering Council of Canada for support through a NSERC Postgraduate Scholarship, and the U.S. Department of State for an International Fulbright Science & Technology Award. The research made use of the shared facilities at the Materials Research Laboratory (MRL) (DMR-1121053). Calculations were conducted at the UCSB Center for Scientific Computing, supported by the California Nanosystems Institute (NSF CNS-0960316), Hewlett-Packard, and the MRL. This work benefited from the use of the Advanced Photon Source at Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences (DE-AC02-06CH11357), and neutron scattering at the Lujan Center, funded by DOE Office of Basic Energy Sciences at the time of measurement; LANL is operated by Los Alamos National Security LLC (DE-AC52-06NA25396).

1.
H. A.
Höppe
,
Angew. Chem., Int. Ed.
48
,
3572
(
2009
).
2.
S.
Pimputkar
,
J. S.
Speck
,
S. P.
DenBaars
, and
S.
Nakamura
,
Nat. Photonics
3
,
180
(
2009
).
3.
N. C.
George
,
K. A.
Denault
, and
R.
Seshadri
,
Annu. Rev. Mater. Res.
43
,
481
(
2013
).
4.
N.
Hirosaki
,
C.
Kocer
,
S.
Ogata
, and
K.
Tatsumi
,
Phys. Rev. B
71
,
104105
(
2005
).
5.
R.-J.
Xie
,
N.
Hirosaki
,
H.-L.
Li
,
Y. Q.
Li
, and
M.
Mitomo
,
J. Electrochem. Soc.
154
,
J314
(
2007
).
6.
T.
Nakayasu
,
T.
Yamada
,
I.
Tanaka
,
H.
Adachi
, and
S.
Goto
,
J. Am. Ceram. Soc.
80
,
2525
(
1997
).
7.
K.
Kimoto
,
R.-J.
Xie
,
Y.
Matsui
,
K.
Ishizuka
, and
N.
Hirosaki
,
Appl. Phys. Lett.
94
,
041908
(
2009
).
8.
N. C.
George
,
A.
Birkel
,
J.
Brgoch
,
B.-C.
Hong
,
A. A.
Mikhailovsky
,
K.
Page
,
A.
Llobet
, and
R.
Seshadri
,
Inorg. Chem.
52
,
13730
(
2013
).
9.
N. C.
George
,
A. J.
Pell
,
G.
Dantelle
,
K.
Page
,
A.
Llobet
,
M.
Balasubramanian
,
G.
Pintacuda
,
B. F.
Chmelka
, and
R.
Seshadri
,
Chem. Mater.
25
,
3979
(
2013
).
10.
J.
Brgoch
,
S. P.
DenBaars
, and
R.
Seshadri
,
J. Phys. Chem. C
117
,
17955
(
2013
).
11.
S.
Dodd
,
M.
Cankurtaran
,
G.
Saunders
, and
B.
James
,
J. Mater. Sci.
36
,
2557
(
2001
).
12.
M.
Alcalá
,
J.
Criado
,
F.
Gotor
, and
C.
Real
,
J. Mater. Sci.
41
,
1933
(
2006
).
13.
J.
Niu
,
K.
Harada
,
I.
Nakatsugawa
, and
T.
Akiyama
,
Ceram. Int.
40
,
1815
(
2014
).
14.
X.
Yi
and
T.
Akiyama
,
J. Alloys Compd.
495
,
144
(
2010
).
15.
J.
Wang
,
B. H.
Toby
,
P. L.
Lee
,
L.
Ribaud
,
S. M.
Antao
,
C.
Kurtz
,
M.
Ramanathan
,
R. B.
Von Dreele
, and
M. A.
Beno
,
Rev. Sci. Instrum.
79
,
085105
(
2008
).
16.
A. C.
Larson
and
R. B. V.
Dreele
,
Los Alamos National Laboratory Report LAUR
,
2000
, p.
86
.
17.
B. H.
Toby
,
J. Appl. Crystallogr.
34
,
210
(
2001
).
18.
V. F.
Sears
,
Neutron News
3
,
26
(
1992
).
19.
T.
Proffen
,
T.
Egami
,
S. J. L.
Billinge
,
A. K.
Cheetham
,
D.
Louca
, and
J. B.
Parise
,
Appl. Phys. A.: Mater. Sci. Process.
74
,
S163
(
2002
).
20.
G.
Kresse
and
J.
Furthmuller
,
Phys. Rev. B
54
,
11169
(
1996
).
21.
G.
Kresse
and
D.
Joubert
,
Phys. Rev. B
59
,
1758
(
1999
).
22.
B.
Ravel
and
M.
Newville
,
J. Synchrotron Radiat.
12
,
537
(
2005
).
23.
See supplementary material at http://dx.doi.org/10.1063/1.4901104 for details of the EXAFS fitting and Debye temperature determination.
24.
I. D.
Brown
,
Chem. Rev.
109
,
6858
(
2009
).
25.
Y. Le
Page
and
P.
Saxe
,
Phys. Rev. B
65
,
104104
(
2002
).
26.
C. M.
Fang
and
R.
Metselaar
,
J. Mater. Chem.
13
,
335
(
2003
).
27.
L.
Benco
,
J.
Hafner
,
Z.
Lenčš
, and
P.
Šajgalk
,
J. Am. Ceram. Soc.
86
,
1162
(
2003
).
28.
W.-Y.
Ching
,
M.-Z.
Huang
, and
S.-D.
Mo
,
J. Am. Ceram. Soc.
83
,
780
(
2000
).
29.
B. T. M.
Willis
and
A. W.
Pryor
,
Thermal Vibrations in Crystallography
(
Cambridge University Press
,
1975
).
30.
K.
Takahashi
,
R.-J.
Xie
, and
N.
Hirosaki
,
Electrochem. Solid-State Lett.
14
,
E38
(
2011
).

Supplementary Material