DOI:
10.1039/C5RA15641F
(Paper)
RSC Adv., 2015,
5, 74790-74801
Blue-silica by Eu2+-activator occupied in interstitial sites
Received
5th August 2015
, Accepted 12th August 2015
First published on 12th August 2015
Abstract
A blue-emitting SiO2:Eu2+ compound has been successfully synthesized and characterized. The PL intensity of SiO2:Eu0.0022+ compound is about 24 times higher than that of the O-defective SiO2 compound (without activators), which emits blue light. The valence state of the Eu ions responsible for the highly enhanced blue emission was determined to be Eu2+ using reference materials (EuCl2 and EuCl3) and XPS measurements. The Eu2+-activator ions occupied in the interstitial sites of the SiO2 matrix were confirmed by FT-IR, XPS, and 29Si MAS-NMR spectroscopy. Even though the void spaces formed structurally in both α-quartz and α-cristobalite can accommodate Eu2+ ions (ionic radius = 1.25 Å at CN = 8), SiO2:Eu2+ compound fired at 1300 °C under a hydrogen atmosphere is destined to be deficient in O or Si atoms, indicating the formation of the wider void spaces in the SiO2 crystal lattice. A sputtered depth profile of SiO2-related compounds obtained by time-of-flight secondary ion mass spectrometry (TOF-SIMS) corroborates the O-defective SiO2 induced by hydrogen. In particular, the interatomic potentials, depending on the interstitial positions of Eu atoms in α-cristobalite and α-quartz, are calculated based on Lennard-Jones and coulomb potentials. For α-cristobalite, the minimum potential value is −51.47 eV and for α-quartz, the value is 221.8 eV, which reveals that the Eu2+-activator ions more preferably enter the interstitial sites of α-cristobalite than those of α-quartz. Thanks to the stable Eu2+-activator ions enclosed by Si–O linkages, the SiO2:Eu0.0022+ compound emits blue light and its PL emission intensity is about 24 times higher than that of the O-defective SiO2 compound. This phosphor material could be a platform for modeling a new phosphor and for application in the solid-state lighting field.
1 Introduction
Silicone (27.2 wt%) is the most abundant element in the earth's crust after oxygen (45.5%), and silicone never occurs free but invariably occurs combined with 4-coordinated oxygen in nature. The [SiO4] unit may occur as an individual group or be linked into chains, ribbons, sheets, or three-dimensional frameworks.1 Silica, or silicon dioxide (SiO2) naturally occurs in both crystalline and amorphous forms. The various forms of the crystalline silica are α-quartz, β-quartz, α-tridymite, β-tridymite, α-cristobalite, β-cristobalite, keatite, coesite, stishovite, and moganite.2 The most abundant form of silica is α-quartz, which is the most thermodynamically stable form of the crystalline silica in ambient conditions. Quartz has been used for a long time (several thousand years) in jewelry as a gem stone, and it is used extensively in electronics as well as the optical component industries.3 It is very interesting to note that Zolensky et al. reported blue quartz phenocrysts from Llano rhyolite (llanite), Llano County, Texas, and derived its coloration from Rayleigh scattering by abundant submicrometer-sized ilmenite inclusions, without experimental evidence.4 According to their results, the concentrations of rare-earth element (REE) in blue quartz were determined as 0.910 ppm (Ce), 0.108 ppm (Sm), 0.016 ppm (Eu), 0.029 ppm (Tb), 0.332 ppm (Yb), and 0.064 ppm (Lu). It is well-known that in phosphor materials, various activators, such as Nd3+, Pr3+, Sm3+, Eu3+, Eu2+, Ce3+, Tb3+ and Yb3+, have been widely used for flat panel displays and optoelectronic devices.5–8 Among them, the divalent europium ions (Eu2+) have been considered as a very important and useful activator, which exhibits broad emission bands between the ultraviolet (UV) and red spectral ranges, associated with the 4f65d1–4f7 transition.9–11 Considering various activator ions, it is presumed that the origin of blue coloration in quartz from Llano rhyolite (llanite), Llano County, Texas, may be due to the Eu2+ or Ce3+ ions occupying the crystal lattice of quartz. Zolensky et al., however, argued that the blue quartz originates from Rayleigh scattering by the ilmenite inclusions from the large difference in the ionic radii of Si4+ (0.40 Å) and Eu2+ (1.31 Å), precludes direct substitution.4 It is very interesting that the blue emission induced by the O-related defects (without activator ions) in SiO2-related compounds was reported by several groups.12–18 Uchino et al. proposed that the blue-emitting center in oxidized porous silicone and nano-SiO2 materials is a metastable defect pair consisting of
Si(O2) and
Si. On the basis of the density functional theory calculation, which was in good agreement with the peak positions of the PLE from the blue-emitting materials.14 McCrate et al. also presented that the intrinsic oxygen vacancy defect (OVD) on planar fused silica was experimentally detected by titration with fluorescent probe molecules such as perylene-3-methanamine and 3-vinyl perylene.18 Blue emission by Ce3+ ions occupied in the interstitial site of the AlN structure was reported by Liu et al.19 According to their results, an octahedral interstitial site of the AlN compound with a würtzite structure was proposed to be the site for Ce3+ ions because the ionic size of Ce3+ ions is much larger than that of Al3+. It should be noted that there are interstitial sites formed structurally in the crystal lattice of α-quartz (or α-cristobalite) and possibilities for Ce3+ or Eu2+ activator ions to enter the interstitial sites of α-quartz (or α-cristobalite), which probably results in the blue coloration in α-quartz (or α-cristobalite). In this study, we report the synthesis and characterization of a blue silica induced by Eu2+-activator ions occupied in interstitial sites. In particular, the Eu2+-doping mechanism in the SiO2 matrix is discussed using a photoluminescence (PL) apparatus, Rietveld refinements, Nano-Secondary Ion Mass Spectrometry (Nano-SIMS), X-ray photoelectron spectroscopy (XPS), 29Si magic-angle-spinning nuclear magnetic resonance (MAS-NMR) spectrometer, and theoretical calculations of the interatomic potential based on Lennard-Jones potential and coulomb potential.
2 Experimental
The polycrystalline SiO2:Eu compounds have been prepared from a mixture of SiO2 (quartz) and Eu2O3 using NH4Cl as a flux and lubricant under 4% H2–Ar mixture gas between 1000 °C and 1500 °C for 6 h. Alumina boats covered with lids were used to maximize collision time among the starting materials (SiO2, NH4Cl, and Eu2O3), and to prevent the loss of unreacted NH4Cl from a rapid out-gassing. In the firing step, each mixture was initially heated at 450 °C for 1 h, and then subsequently heated at a final temperature for 3 h (heating rate: 3 °C min−1). Powder X-ray diffraction measurements of SiO2:Eu were carried out using an X-ray diffractometer equipped with a graphite monochromator (DMAX-2200PC, Rigaku). A step scan mode was employed in a 2θ range of 10–110° with a step size of 0.02° and counting time of five second for each step. Structure refinements were carried out by the Rietveld method using the Fullprof program20 with pseudo-Voïgt peak shapes and refined backgrounds. The photoluminescence (PL) spectra were obtained at room temperature using a fluorescent spectrophotometer with a 150 W xenon lamp under an operating voltage of 350 V (Fluorometer FS-2, Scinco). The reflectance spectra were obtained using a UV-visible spectrophotometer (UV-2600, Shimadzu) with BaSO4 as a reference. Fourier-transform infrared spectroscopy (FT-IR) was performed using a FT-IR spectrophotometer (IRTracer-100, Shimadzu) with a resolution range of ±0.5 cm−1 and a KBr medium prepared with 0.5 wt% of sample. The Nano-Secondary Ion Mass Spectrometry (NanoSIMS) measurements were done using a rastering Cs+ primary ion beam on the Cameca Ametek NanoSIMS 50L at the Arizona State University. The beam current at the sample was lowered to ∼0.48 pA by choosing a small diaphragm to obtain a fine beam. Although a low current was used for analysis, the new capillary Cs source produced a high primary beam density and high secondary ion count rates during imaging. Negative secondary ions of 24Si− and 16O− were measured simultaneously using electron multipliers in the multi-collection mode. The measurement of NanoSIMS for all samples was performed after pre-sputtering for 5 min. All samples for the NanoSIMS measurement were prepared as follows: 500 mg of sample was dispersed in 5 mL of absolute ethanol by ultrasound sonication for 10 min. The obtained milky suspension after sonication was dripped by a pipet onto a stainless-steel substrate (1 cm diameter) and dried under ambient conditions. This was repeated until a white thick coating was obtained on the substrate. The oxidation states of the elements were analyzed by X-ray photoelectron spectroscopy (XPS, ESCALAB 250) with a monochromatic Al Kα X-ray source (hν = 1486.6 eV) at the Busans Center of Korea Basic Science Institute (KBSI). The obtained binding energies (BEs) were calibrated with that of an adventitious carbon (C 1s) core level peak at 284.6 eV as a reference. All 29Si single-pulse magic angle spinning (SP MAS) spectra were acquired using a Unity INOVA 600 MHz NMR instrument (Agilent Technologies, USA) with a cross polarization (CP) MAS probe for 5 mm zirconia rotors at room temperature, a 30° pulse corresponding to 2 μs, a pulse repetition delay of 20 s, a spectral width of 120 kHz, and a spinning rate of 10 kHz. The chemical shifts were referenced to external tetramethylsilane at 0 ppm.
3 Results and discussion
3.1 Structural characterization
Structural refinements of the samples were carried out using a Rietveld program. The background was fitted with a polynomial function and the peak shape was modelled using a pseudo-Voïgt function. The observed, calculated, and difference patterns from the Rietveld refinements of the X-ray diffraction data are shown in Fig. 1. The profile of undoped SiO2 fired at 1300 °C under a 4% H2–Ar atmosphere contains α-cristobalite as the major phase (90%) and quartz as the second phase (10%). The atomic coordinates of the both structures were initially selected from a data base (α-cristobalite: ICSD #75300, α-quartz: ICSD # 647436). The relative proportion of phases, final structural parameters (lattice parameters, refined atomic positions, and isotropic temperature factors for all atoms) and reliability factors are listed in Table 1. In comparison with the undoped SiO2, the Eu-doped SiO2 compound showed no substantial difference in the crystal structure except for the temperature factors. For instance, the isotropic temperature factors, Biso of Si and O atoms are 1.18(3) Å2 and 1.36(6) Å2 for the α-cristobalite phase in undoped SiO2, while those are 1.39(3) Å2 and 1.78(6) Å2 in doped SiO2:Eu0.002. The large temperature factors for SiO2:Eu0.002 imply that the Eu doping into SiO2 leads to the increase in the structural disorder. Table 2 shows the weight percentages (%) of α-quartz and α-cristobalite as a function of firing temperature and Eu content. It is evident that the weight% of α-cristobalite is increased with increasing firing temperature and Eu content.
 |
| Fig. 1 Rietveld refinement of the powder XRD profiles for SiO2:Eu0.002 (a) and undoped SiO2 (b). The diffraction pattern is composed of the peaks from α-cristobalite as a main phase and α-quartz as a minor phase. Measured data, fitted results, expected reflection positions, and the difference between measured and fitted results are expressed as red open circles, black solid lines, black vertical lines, and blue solid lines, respectively. The peak intensities over 40 degrees in 2θ have been magnified 10 times for clarity. | |
Table 1 Relative proportion of phases, structural parameters and reliability factors for SiO2:Eu0.002 and SiO2
Compound |
SiO2:Eu0.002 |
SiO2:No-Eu |
Phase |
α-Cristobalite |
α-Quartz |
α-Cristobalite |
α-Quartz |
Weight present (%) |
|
89.2(4) |
10.8(13) |
90.1(5) |
9.9(17) |
Rwp (%) |
|
11.1 |
13.0 |
RBragg |
|
2.63 |
7.60 |
2.89 |
6.69 |
Rf (%) |
|
2.98 |
4.53 |
3.17 |
4.61 |
χ2 |
|
1.51 |
2.12 |
Space group |
|
P41212 (no. 92) |
P3221 (no. 154) |
P41212 (no. 92) |
P3221 (no. 154) |
Lattice |
a |
4.9768(2) |
4.9030(5) |
4.9732(3) |
4.9041(6) |
Parameter (Å) |
c |
6.9490(4) |
5.3831(8) |
6.9362(4) |
5.3851(9) |
Cell volume (Å3) |
|
172.12(14) |
112.07(3) |
171.55(16) |
112.16(3) |
Compound |
SiO2:Eu0.002 (α-cristobalite) |
SiO2:No-Eu (α-cristobalite) |
|
SiO2:Eu0.002 (α-quartz) |
SiO2:No-Eu (α-quartz) |
Si, 4a |
|
|
Si, 3a |
|
|
x |
0.2967(2) |
0.2966(2) |
x |
0.470(2) |
0.469(2) |
y |
0.2967(2) |
0.2966(2) |
y |
0 |
0 |
z |
0 |
0 |
z |
2/3 |
2/3 |
Site occupancies |
1 |
1 |
Site occupancies |
1 |
1 |
Biso (Å2) |
1.39(3) |
1.18(3) |
Biso (Å2) |
0.5(1) |
0.4(1) |
O, 8b |
|
|
O, 6c |
|
|
x |
0.2412(4) |
0.2428(5) |
x |
0.439(2) |
0.433(2) |
y |
0.0985(4) |
0.1012(5) |
y |
0.279(2) |
0.276(2) |
z |
0.1756(3) |
0.1757(3) |
z |
0.786(2) |
0.783(2) |
Site occupancies |
1 |
1 |
Site occupancies |
1 |
1 |
Biso (Å2) |
1.78(6) |
1.36(6) |
Biso (Å2) |
2.1(3) |
1.7(4) |
Table 2 Unit cell parameters and phase content between α-quartz and α-cristobalite in silicas fired under a 4% H2–Ar atmosphere
Compound |
α-Cristobalite (weight%) |
a (Å) |
c (Å) |
α-Quartz (weight%) |
a (Å) |
c (Å) |
SiO2 (Aldrich chemical) |
— |
— |
— |
100% |
4.91202(13) |
5.40345(17) |
SiO2 (1500 °C) |
100% |
4.97114(19) |
6.9254(3) |
— |
— |
— |
SiO2 (1300 °C) |
90.03% |
4.9731(3) |
6.9362(4) |
9.97% |
4.9043(6) |
5.3844(10) |
SiO2:Eu0.002 (1300 °C) |
89.20% |
4.9767(2) |
6.9491(4) |
10.80% |
4.9035(6) |
5.38214 |
SiO2:Eu0.05 (1300 °C) |
99.07% |
4.9733(2) |
6.9341(3) |
0.93% |
4.90(5) |
5.38(9) |
3.2 Photoluminescence spectra
Fig. 2 presents the PL spectra of SiO2:Eu0.01 synthesized under a 4% H2–Ar atmosphere as a function of firing temperature. The highest excitation and emission intensity are observed in the SiO2:Eu0.01 compound synthesized at 1300 °C that is an optimal firing temperature. The excitation spectra of SiO2:Eu0.01 compounds monitored at 438 nm, consist of broad bands between 220 nm and 420 nm, which may be ascribed to the allowed 4f7–4f65d transitions of Eu2+.21–23 The emission spectra monitored under the 323 nm excitation show symmetric bands centered around 438 nm, which is associated with blue-emission. The excitation and emission spectra of SiO2:Eux compounds synthesized at 1300 °C as a function of Eu content are shown in Fig. 3. The PL maximum intensity is observed at x = 0.002, then decreased with increasing Eu concentration. The optimum concentration of Eu in the SiO2:Eu compound is 0.002, which implies that SiO2 crystal lattice is difficult to accommodate the large amount of Eu2+-activator ions. It should be noted that NH4Cl is widely used as a flux and lubricant chemical in solid-state reactions because it has a relatively low melting point (340 °C) and boiling point (520 °C). In this study, NH4Cl plays an important role in SiO2:Eu compounds as a flux and lubricant, which forces Eu2+ to enter the interstitial sites of the SiO2 matrix. As shown in Fig. 4, the PL intensity of SiO2:Eu0.0022+ is increased with the higher mole ratio between NH4Cl and SiO2 (mole (NH4Cl)/mole (SiO2)). The maximum PL intensity is observed at the mole ratio 4. It can be noted that the PL pattern of SiO2 (No-Eu, 4 mole NH4Cl) is similar to that of SiO2:Eu0.0022+ except the difference of the PL intensities, indicating the blue emissions induced by the O-related defects (without activator ions) in SiO2-related compounds12–18 as mentioned in the introduction. It is very remarkable that the PL intensity of SiO2:Eu0.0022+ is about 24 times as high as that of SiO2 (No-Eu, 4 mole NH4Cl) fired at 1300 °C under a 4% H2–Ar atmosphere, which is probably due to the Eu ions occupying the SiO2 matrix. Fig. 5 presents the PL spectra of SiO2:Eux2+ (x = 0.002, 0.005, and 0.010) with and without NH4Cl addition. For SiO2:Eu0.0022+ compound, it prominently shows that the emission intensity (at 438 nm) of the resulting materials after NH4Cl addition is 1.4 times higher than that of the resulting materials without NH4Cl addition. This result implies that NH4Cl shows a considerable effect on the PL intensity and plays an important role as a flux and lubricant in SiO2:Eux2+ compounds. Fig. 6 presents the diffuse reflectance spectra of SiO2 and SiO2:Eux2+ (x = 0.002 and 0.05). When Eu2+ ions are occupied into the SiO2 matrix, the broad absorption bands are shown between 220 nm and 450 nm, whereas absorption bands are not shown in α-quartz and α-cristobalite in the region. This may indicate the broad absorption bands between 220 nm and 450 nm are associated with the 4f → 5d transition of Eu2+. Moreover, the absorption bands of SiO2:Eu0.0022+ are still stronger than that of SiO2:Eu0.052+, which is consistent with the PL previous results, as shown in Fig. 3. The luminescent behavior of the SiO2:Eu0.0022+ compound synthesized at 1300 °C under a 4% H2–Ar atmosphere with the mole ratio = 4 (mole (NH4Cl)/mole (SiO2)) was compared with a commercial BaMgAl10O17:Eu2+ (BAM:Eu2+) phosphor (obtained from Nichia corp., Japan). The excitation and emission spectra of SiO2:Eu0.0022+ and BAM:Eu2+ phosphor are shown in Fig. 7. The excitation and emission spectra are very similar except for the difference in the PL intensity. The relative emission intensity of SiO2:Eu0.0022+ monitored at 323 nm is about 40% compared to a commercial BAM:Eu2+. The Commission International de I'Eclairage (CIE) coordinates of SiO2:Eu0.0022+ and BAM:Eu2+ monitored under UV light at 300 nm are x = 0.145, y = 0.068 and x = 0.143, y = 0.065, respectively, as shown in the inset of Fig. 7. According to CIE coordinates, the SiO2:Eu0.0022+ compound emits deep blue.
 |
| Fig. 2 Excitation and emission spectra of SiO2:Eu0.01 as a function of firing temperature. | |
 |
| Fig. 3 Excitation and emission spectra of SiO2:Eux2+ as a function of Eu content. | |
 |
| Fig. 4 Excitation and emission spectra of SiO2:Eu0.0022+ as a function of NH4Cl mole. | |
 |
| Fig. 5 Excitation and emission spectra of SiO2:Eux2+ with and without NH4Cl (4 moles). | |
 |
| Fig. 6 Diffuse reflectance spectra of SiO2-related compounds. | |
 |
| Fig. 7 PL spectra of SiO2:Eu0.0022+ and a commercial BaMgAl10O17:Eu2+ (BAM) phosphor (obtained from Nichia corp., Japan). The inset shows the CIE chromaticity for two phosphors depending on the excitation UV light. | |
3.3 Infrared spectroscopy
Fig. 8 presents the IR spectra of SiO2-related compounds. The absorption band at 1092 cm−1 in the spectra of α-quartz is associated with Si–O asymmetrical stretching vibrations, those at 799 and 780 cm−1 with Si–O symmetrical stretching vibrations, that at 696 cm−1 with Si–O symmetrical bending vibrations, and those at 512 and 460 cm−1 with Si–O asymmetrical bending vibrations.24 After firing α-quartz at 1500 °C, the IR spectrum of α-quartz is completely transformed into that of α-cristobalite with a notable new band at 621 cm−1 corresponding to Si–O asymmetrical bending vibrations.25 In SiO2:Eu2+ compounds fired at 1300 °C under a 4% H2–Ar atmosphere, there is no considerable wavenumber shift in IR modes except that the lower absorbance and greater broadness of IR modes compared with those of α-cristobalite. The IR modes in the SiO2 compound (without Eu) fired at 1300 °C under a 4% H2–Ar atmosphere reveals that the Eu ions occupied in the SiO2 matrix might have an effect on the SiO2 internal modes. As a consequence of the interaction, (Eu2+⋯[O–Si–O]4−⋯Eu2+), the local symmetry of the SiO2 matrix may partially collapse, which results in the modification of the [SiO4] internal modes, i.e., lower absorbance and greater broadness of IR modes because there is not enough Eu content to induce the chemical shift of the vibrational modes of the SiO2:Eu compounds.26 It should be noted that the variation of the IR modes by Eu ion doped in a SiO2 matrix means that there is a pseudo-covalent bond character between the Eu ion and O ion sharing with SiO4 tetrahedral units. Thus, this may indicate that Eu ions are occupied in the SiO2 interstitial sites formed structurally, particularly, in wider ones formed under a 4% H2–Ar atmosphere.
 |
| Fig. 8 Infrared spectra of SiO2-related compounds. | |
3.4 X-ray photoelectron and solid NMR spectroscopy
To examine the valence state of Eu ions occupied into the SiO2 matrix, XPS analysis was performed. All the XPS spectra were fitted after a Shirley background correction. Fig. 9 presents the wide-scan XPS spectra of EuCl2 (Eu2+) and EuCl3 (Eu3+) reference compounds. It should be pointed out that it is very difficult to obtain fully reduced Eu2+ in metal oxide-phosphor compounds based on the standard reduction potential (Eu3+/Eu2+ = −0.36 V vs. standard hydrogen electrode (SHE)),27 which indicates that the reduction of Eu3+ to Eu2+ requires an annealing process at high temperature (≥1000 °C) under a reducing atmosphere such as H2 or H2–Ar mixture gas. In addition, the thermodynamically unstable Eu2+ ions are prone to be easily oxidized by foreign molecules, such as H2O and O2, in particular, under X-ray irradiation in XPS measurements. As shown in the inset of Fig. 9, the fact that the new band around 128 eV is observed in EuCl2 compound supports the partial oxidation of Eu2+ to Eu3+. Fig. 10 presents Eu 3d and Eu 4d XPS spectra of EuCl2 and EuCl3. The Eu 4d binding energies of EuCl3 (top at right) are assigned to the Eu3+ 4d5/2 (137.3 eV) and Eu3+ 4d3/2 (142.9 eV).28–30 The spin-orbital splitting value for Eu3+ ion is 5.6 eV, which is in good agreement with that previously reported.28 The Eu 4d binding energies of EuCl2 (bottom at right) are assigned to Eu2+ 4d5/2 (128.4 eV), Eu2+ 4d3/2 (134.2 eV), Eu3+ 4d5/2 (136.7 eV), and Eu3+ 4d3/2 (142.3 eV), which are consistent with XPS results as previously reported.31–36 Notably, Jiang et al. synthesized EuCl2 nanoprisms and nanorods and presented Eu 4d XPS spectra, as follows:36 Eu2+ 4d5/2 (128.2 eV), Eu2+ 4d3/2 (∼134 eV), Eu3+ 4d5/2 (∼136 eV), and Eu3+ 4d3/2 (142.2 eV), which is in good agreement with our results. Moreover, they mentioned that Eu 4d XPS spectrum showed the clear presence of Eu2+ (128.2 eV and ∼134 eV), although the Eu2+ ions on the surface were oxidized to Eu3+ (∼136 eV and 142.2 eV). As shown in Fig. 10 (bottom at right), the spin-orbital splitting values of Eu3+ and Eu2+ ion (for EuCl2) are 5.6 eV and 5.8 eV, respectively. Interestingly, the Eu2+ 4d5/2 peak at 128.4 eV consists of two components with the spin–orbit splitting value of 5.8 eV, which may be attributed to the different sites of the Eu2+ 4d5/2 induced by the partial oxidation of Eu2+. From the XPS measurement for EuCl2 and EuCl3 reference compounds, it clearly shows that EuCl2 (Eu2+) compound is more sensitive than EuCl3 (Eu3+) in oxidation induced by X-ray irradiation. By the aid of the Eu 4d binding energy for EuCl2 and EuCl3 reference compounds, the XPS 3d binding energies of reference compounds could be precisely assigned, as presented in Fig. 10. The Eu 3d binding energies of EuCl3 reference compound (top at left) are assigned to the Eu3+ 3d5/2 (1136.5 eV) and Eu3+ 3d3/2 (1166.1 eV) with the small amount of Eu2+ state at lower binding energy. On the other hand, the peaks of the Eu2+ 3d binding energies for EuCl2 reference compound (bottom at left) are strongly intensified compared with those of the EuCl3 compound, i.e., Eu2+ 3d5/2 (1126.3 eV) and Eu2+ 3d3/2 (1156.2 eV). In particular, it is reasonable that the Eu2+ valence state in EuCl2 compound is prone to be easily oxidized under X-ray irradiation, as presented in Eu 3d as well as Eu 4d XPS spectra. It should be noted that the difference in 3d binding energies between Eu2+ and Eu3+ (9.7 eV) is somewhat larger than that of the 4d binding energies (8.3 eV). It can be considered that the 3d electrons are closer to the nucleus than the 4d elections and thus their binding energies are more strongly affected (ΔB.E ∝ 1/r). To examine the valence state of Eu ions occupied into the interstitial sites of SiO2, XPS measurements were performed and its results are presented in Fig. 11. As the PL spectrum (top at right) of SiO2:Eu0.052+ compound synthesized under 5% H2–Ar gas mixture shows the characteristic of Eu2+, 4f65d → 4f7 transition with a broad band between 380 nm and 520 nm corresponds to the blue emission. The Eu 3d XPS binding energies of SiO2:Eu0.052+ consist of four components, Eu2+ 3d5/2 (1126.1 eV), Eu3+ 3d5/2 (1135.8 eV), Eu2+ 3d3/2 (1155.9 eV), and Eu3+ 3d3/2 (1165.4 eV). Considering the PL emission spectrum, it is evident that the Eu2+ ions in the SiO2:Eu0.052+ compound is easily oxidized under X-ray irradiation. After thermal treatment of SiO2:Eu0.052+ compound under O2 atmosphere at 1300 °C for 3 h, the PL emission spectrum (bottom at right) was obtained. As its PL behavior shows the characteristic of Eu3+ with a line spectrum at 616 nm corresponding to the red emission, it is evident that the Eu2+ ions that occupy the SiO2 matrix are all nearly oxidized to Eu3+. Thus, the Eu3+ 3d XPS spectrum (bottom at left) is predominantly obtained after thermal treatment of SiO2:Eu0.052+ compound under an O2 atmosphere at 1300 °C for 3 h. From the XPS measurement, a valuable information was obtained that the valence state of Eu ions occupied into the SiO2 matrix is predominantly Eu2+, associated with the blue emission. However, XPS as a surface analysis technique, cannot determined the total presence of Eu3+ in the SiO2:Eu0.052+ compound even though there exists a small amount of Eu3+. Thus, XPS results evidently imply that the Eu2+ ions stabilized on the SiO2 matrix are easily oxidized under X-ray irradiation in the XPS measurements. 29Si MAS-NMR spectra of the SiO2 compounds are given in Fig. 12. The chemical shifts for α-quartz (a) and α-cristobalite ((b), after firing α-quartz fired at 1500 °C under 4% H2/Ar) are −107.7 ppm and −109.3 ppm, respectively, which is in good agreement with those as previously reported.37–39 α-Quartz (a) and α-cristobalite (b) show essentially no “side-bands” and no “chemical shifts”, indicating that there are neither different Si sites nor metal cations, as shown in Fig. 12 α-Quartz is transformed into the mixed phase (c) with 10 wt% of α-quartz and 90 wt% of α-cristobalite after firing at 1300 °C under a 4% H2–Ar atmosphere, and the chemical shift of the mixed phase is −109.5 ppm, which is similar to that of α-cristobalite. It can be noted that in SiO2:Eu2+ compounds, the spinning side-bands appearance depends on the Eu content; two weak side bands in SiO2:Eu0.0022+ fired at 1500 °C under 4% H2–Ar (d) and SiO2:Eu0.0022+ fired at 1300 °C under 4% H2–Ar (e), and six intense ones in SiO2:Eu0.052+ fired at 1300 °C under 4% H2–Ar (f). It is well-known that the spinning side-bands originated from the dipolar interactions between the nuclear spin and the unpaired electron spins of the paramagnetic ions,40–44 in this system between 29Si spin and the unpaired electron spins of the Eu2+ ions. Thus, the fact that there are spinning side-bands in MAS-NMR spectra of SiO2:Eu2+ compounds supports that the Eu2+ ions are occupied with a pseudo-covalent bond character in the interstitial sites of the SiO2 crystal lattice.
 |
| Fig. 9 Wide scan XPS spectra of reference compounds, EuCl2 and EuCl3. | |
 |
| Fig. 10 High-resolution Eu 3d and 4d XPS spectra of reference compounds EuCl2 and EuCl3. | |
 |
| Fig. 11 High-resolution Eu 3d XPS spectra and PL emission spectra of SiO2:Eu0.052+ and SiO2:Eu0.053+. | |
 |
| Fig. 12 29Si MAS-NMR spectra of SiO2-related compounds; (a) α-quartz, (b) α-cristobalite after firing α-quartz at 1500 °C under H2, (c) SiO2 fired at 1300 °C under H2, (d) SiO2:Eu0.0022+ fired at 1500 °C under H2, (e) SiO2:Eu0.0022+ fired at 1300 °C under H2, and (f) SiO2:Eu0.052+ fired at 1300 °C under H2. The spinning side-bands are marked by arrows. | |
3.5 Hydrogen assisted SiO2-defects: nano-SIMS measurements
It is well-known that a reducing atmosphere plays an important role in controlling the oxygen vacancies as well as charge valences of metal ions in metal oxides. Therefore, the additional charges should be compensated for in order to incorporate metal ions of lower valence into the interstitial sites as well as to replace Si4+ with metal ions of lower valence in the SiO2 matrix. The types of the point defects in silicon dioxide can be classified into intrinsic and extrinsic point defects. Intrinsic point defects involve vacancies caused by the missing host atoms and self-interstitials by additional host atoms at an interstitial position. Extrinsic point defects relate to foreign atoms different from the host crystal. Defects in a perfect silicon dioxide could include O (or Si) vacancies, their interstitials, Si–Si (or O–O) homobonds, and under-coordinated silicons or oxygens.45 Many authors consider hydrogen to be an intrinsic defect because it has been commonly found in silicon dioxide.46–50 Mysovsky et al.47 argued that the oxygen-deficient vacancy combined a hydrogen atom, E′4 center consists of a hydrogen substituting for an oxygen atom in α-quartz, as the following reaction: |
Si–O–Si + H → Si–H + Si + O (interstitial)
| (1) |
The neutral oxygen vacancy (ODC(I)) could be converted to
Si–O–H groups by thermal reaction with hydrogen molecules according to the following reaction:51,52
|
Si–Si + H2 → 2{ Si–O–H}
| (2) |
The formation mechanism of hydrogenic trapped-hole species in α-quartz was proposed by Nuttall et al.,48 which indicates that the hydrogen ions (4H+) incorporated into α-quartz structure substitute for Si4+ as the following reaction:
|
4{ Si–O}–Si (defect) + 4H+ → 4{ Si–O–H}
| (3) |
The variation of the O/Si signal ratios of SiO2 compounds is clearly seen depending on the firing condition as presented in Fig. 13. A sputtered depth profile of SiO2 compounds obtained by time-of-flight secondary ion mass spectrometry (TOF-SIMS) corroborates the O-defective SiO2 induced by hydrogen as discussed above. The O/Si signal ratio of α-quartz as received (Aldrich chemical) is the highest value, whereas those of SiO2:Eu0.0022+ and SiO2:Eu0.052+ fired at 1300 °C under a H2 atmosphere are lower than that of commercially available α-quartz, indicating that the oxygen atoms are predominantly missing at a high temperature under a hydrogen atmosphere. Moreover, α-cristobalite obtained after firing α-quartz at 1500 °C under a H2 atmosphere shows the constant and lowest O/Si ratio. Although our Nano-SIMS analysis does not allow for the quantification of H content in the samples, the observed O/Si signal ratios are consistent with this formation mechanism.
 |
| Fig. 13 O/Si signal ratios determined by TOF-SIMS depth profile in SiO2-related compounds. | |
3.6 Interstitial and vacancy mechanism of Eu2+ doping in SiO2 matrix
The incorporation of Eu2+ ions into a SiO2 matrix may be explained in three different ways. The first explanation is that Eu atoms can replace silicon atoms forming a substitutional solid solution in a SiO2 crystal. However, the direct substitution of Si4+ with Eu2+ is difficult because of the difference in the ionic radii of Si4+ (0.26 Å) and Eu2+ (1.30 Å).53 Moreover, to the best of our knowledge, there has been no report on 4-coordinated Eu2+ (or Eu3+) in metal oxides as well as organometallic complexes. As a second explanation, the Eu atoms can distribute in the boundary of SiO2 grains to form microscopic regions of europium oxide. The europium oxide can react with SiO2 grains to generate Eu-containing silica compounds such as EuSiO3 and Eu2SiO4. Several authors have reported the formation of EuSiO3 and Eu2SiO4 on Eu-doped silica films in the studies.54–56 Li et al. reported that Eu-doped SiO2 films annealed in N2 at a temperature higher than 800 °C exhibited a broad emission band between 400 and 800 nm, which was attributed to the formation of EuSiO3 giving rise to an intense yellow luminescence.54 In Eu-doped SiO2 films, the broad emission peak centered at 610 nm between 500 and 750 nm was observed due to the Eu2+ ions stabilized in the Eu2SiO4 and EuSiO3 crystalline structures.56 In our Eu2+-doped SiO2 compounds, the broad emission bands centered at 440 nm between 380 and 540 nm are observed, indicating blue-emission. Moreover, Rietveld analyses within the Eu-doping range from 0.002 to 0.05 show that there is no trace of impurities, which agree with the phase diagram of the EuO–SiO2 system.57 As a third explanation, we speculate that the Eu atoms can incorporate into the structural void spaces (interstitial sites). Both α-cristobalite and α-quartz have open structures, as shown in Fig. 14. As the void spaces formed in the α-cristobalite structure are somewhat larger than those in the α-quartz structure, it is presumed that Eu2+ ions preferably occupy the interstitial sites of α-cristobalite. Compared with the ionic radius between Eu2+ (1.25 Å at CN = 8) and Eu3+ (1.066 Å at CN = 8),53 Eu3+ ions are more favorable to the void spaces of SiO2 matrix on the condition that no other phases is formed in the EuO–SiO2 solid-solution. However, in the Eu-doped SiO2 compound fired at 1300 °C in air, the impurity phase of Eu2SiO5 (at the Eu-doping range of 0.01) is clearly observed in the XRD pattern and this compound exhibited a faint reddish-emission attributed to the formation of Eu3+ ions, as shown in Fig. 15, which indicates that in an air atmosphere, Eu3+ ions cannot occupy the interstitial sites of SiO2 matrix and react with SiO2 giving rise to the formation of Eu2SiO5 in an air atmosphere within the Eu-doping ranges from 0.002 to 0.05. In contrast, Eu2+ ions can enter the interstitial sites at 1300 °C under a 4% H2–Ar atmosphere. Thus, it is presumed that Eu2+ ions can be stabilized in the interstitial sites of SiO2 matrix through the oxygen vacancies and partial fragmentation of tetrahedron linkages caused by hydrogen assistance. It should be noted that Eu2O3 (without adding SiO2) was fired at 1300 °C under a 4% H2–Ar atmosphere and obtained a red-emitting Eu2O3 with a monoclinic structure. Based on the experimental results, the Eu2+ doping mechanism in SiO2 matrix could be postulated as follows: (i) the reduction of Eu2O3 (Eu3+) to EuO (Eu2+) at a high temperature (≥1000 °C) under a hydrogen atmosphere, (ii) the entrance of Eu2+ ions in the interstitial sites of the SiO2 matrix through the oxygen vacancies and partial fragmentation of tetrahedron linkages caused by hydrogen assistance, (iii) the block of subsequent oxidation of the Eu2+ ions enclosed by Si–O linkages compared with the isolated Eu2+ species.
 |
| Fig. 14 The schematic of the crystal structures for α-cristobalite (top) and α-quartz (bottom). Si and O atoms are represented by blue and red spheres. Structural voids are represented by grey spheres with radius of 1.25 Å (Eu2+ at CN = 8). | |
 |
| Fig. 15 XRD pattern of SiO2:Eu0.013+ fired at 1300 °C in an O2 atmosphere. The inset shows the PL emission spectra. | |
3.7 Theoretical comparison of interstitial positions of Eu atom in α-cristobalite and α-quartz: interatomic potential
We theoretically calculate the interatomic potentials depending on the interstitial positions of Eu atom in α-cristobalite and α-quartz and compare their energy values. As shown in Fig. 16, α-cristobalite and α-quartz have large interstitial void spaces. When an Eu atom is located in the void spaces, the interatomic chemical forces between atoms exert. Atoms have strong repulsion forces when their distance is shorter than sum of their atomic radii. Ions have electric repulsion or attraction, depending on their charges.
 |
| Fig. 16 Eu2+ positions in the interstitial site of α-cristobalite (a) and α-quartz (b). Blue, red, and purple dots are Si, O, and Eu atoms respectively. | |
The former can be modeled by the Lennard-Jones potential58 and the latter can be modeled by a coulomb potential.58 The potential can be written as follows:
|
 | (4) |
where first sum is the Lennard-Jones potential of nearby ions and the second sum is the coulomb potential. Index
i is related to the surrounding atoms in the crystal (
εi: bond dissociation energy,
i: the location of
ith atom,
ri: the minimum potential distance between Eu
2+ and
ith atom,
ZEu: charge number of Eu
2+,
Zi: charge number of
ith atom). There are two types of atoms around Eu, O and Si. For O, we choose
εi as the Eu–O bond dissociation energy 473 kJ mol
−1 = 4.903 eV.
59 r is chosen as sum of ionic radius of Eu
2+ (1.30 Å) and crystal radius O
2− (1.34 Å),
Zi for O charge number is −1 by considering the half ionic character. For Si, it does not form a bond with Eu
2+ ions. We choose
ri as sum of ionic radius of Eu
2+ (1.30 Å) and crystal radius of Si (0.26 Å) and determine the value
εi as 4.892 eV, which gives same repulsion force as Eu and O Lennard-Jones force when atomic distance is compressed to 80% of their sum of radii.
Zi for Si charge number is 4, O charge number −2 and
ZEu is 2. The local minimum point of this potential in three dimension can be found numerically, by successively following the negative gradient of potential (force), which gives the steepest descent. Starting from the initial position, search next potential minimum in the finite interval of the force direction, and repeat the procedure until it converges. In the α-cristobalite case, the minimum position is shown in
Fig. 16. Crystal unit cell parameters are given by (
a,
b,
c) = (4.978, 4.978, 6.948) (Å), (
α,
β,
γ) = (90, 90, 120) (°) and the Eu minimum position is (0.26758, 0.25590, 0.50357) in the unit cell coordinate. Nearby oxygen coordinates are (0.1032, 0.2397, 0.8216), (0.7397, 0.3968, 0.5716), (0.2397, 0.1032, 0.1784) and (0.3968, 0.7397, 0.4284). Their distances from Eu atoms are 2.504, 4.325, 2.769 and 4.293 Å. Thus, the calculated minimum potential value is −51.47 eV. In the α-quartz case, the minimum position is shown in
Fig. 16. Crystal unit cell parameters are given by (
a,
b,
c) = (4.912, 4.912, 5.404) (Å), (
α,
β,
γ) = (90, 90, 120) (°) and the Eu minimum position is (−0.0301, 0.0014, 0.6674) in unit cell coordinates. Nearby oxygen coordinates are (0.4136, 0.2676, 0.7857), (0.1460, −0.2676, 0.5476), (−0.4136, −0.1460, 0.8801), and (−0.2676, 0.1460, 0.4524). Their distances from Eu are 2.005, 2.014, 2.008 and 2.011 Å. Minimum potential value is calculated to be 221.8 eV. The results suggest that it is more difficult to put Eu atoms into α-quartz than α-cristobalite. In the α-cristobalite case, the minimum potential is negative as well as the distances between atoms at the minimum point are large. Only two oxygens around 2.5 and 2.7 Å and next nearest oxygen distances are larger than 4 Å. In the α-quartz case, the potential value is positive, which implies instability. Moreover, the nearby oxygen distances are almost identically 2 Å at a minimum potential point, showing a quite regular and closely packed structure. It is reasonable to infer that the Interstitial positioning of Eu is more easily done in the α-cristobalite case than the α-quartz case.
4 Conclusions
Blue-emitting silica by Eu2+-activator ion occupied in the interstitial sites of SiO2 matrix has been successfully synthesized and characterized. The PL excitation spectra of SiO2:Eu0.01 compounds monitored at 438 nm consist of broad bands between 220 nm and 420 nm, which may be ascribed to the allowed 4f7–4f65d transitions of Eu2+, which is in good agreement with XPS results. The emission spectra, monitored under the 323 nm excitation, show symmetric bands centered around 438 nm, which is associated with blue-emission. From FT-IR spectra of SiO2 compounds with and without Eu2+-activator ions, it is clearly observed that the lower absorbance and greater broadness in the IR modes of SiO2:Eu2+ compounds compared with those of α-quartz and α-cristobalite without Eu2+-activator ions indicate that the Eu2+-activator ions are well located in the interstitial sites of the SiO2 matrix with a pseudo-covalent bond character between Eu ion and O ion sharing with SiO4 tetrahedral units. 29Si MAS-NMR spectra corroborate the IR results interpretation. The fact that there are spinning side-bands in MAS-NMR spectra of SiO2:Eu2+ compounds supports that the Eu2+ ions are occupied with a pseudo-covalent bond character in the interstitial sites of the SiO2 crystal lattice. The interatomic potentials depending on the interstitial positions of the Eu atom in α-cristobalite and α-quartz are calculated using Lennard-Jones potential and coulomb potential. For α-cristobalite, the minimum potential value is −51.47 eV, and for α-quartz, the value is 221.8 eV, which reveals that the Eu2+-activator ions more preferably enter the interstitial sites of α-cristobalite than those of α-quartz. Due to the Eu2+-activator ions occupied in the SiO2 matrix, the PL intensity of the SiO2:Eu0.0022+ compound is about 24 times as high as that of the O-defective SiO2 compound and the relative emission intensity of SiO2:Eu0.0022+ monitored at 300 nm is about 40% compared to a commercial BAM:Eu2+ (obtained from Nichia Co. Ltd in Japan). This phosphor material could be a breakthrough for modeling a new phosphor and for applications in the solid state lighting field.
Acknowledgements
This research was financially supported by the Ministry of Education (MOE) and National Research Foundation of Korea (NRF) through the Human Resource Training Project for Regional Innovation (2014H1C1A1066859). One of the authors, S. J. Kim acknowledges the support from the National Research Foundation (NRF) of Korea (Grant No. 2009-0094046). Dr Seen Ae Chae at the KBSI is acknowledged for carrying out NMR experiments. We would like to thank NSFs Major Research Instrumentation (NSFARRA award #0960334) and Arizona State University for acquisition and installation of the Nano-SIMS 50L.
References
- N. N. Greenwood and A. Earnshaw, Chemistry of the Elements, Pergamon Press, Oxford, 1984 Search PubMed.
- A. F. Wells, Structural Inorganic Chemistry, Clarendon Press, Oxford, 1984 Search PubMed.
- IARC, IARC Monogr Eval Carcinog Risks Hum., 1997, 68, 1, PMID:9303953.
- M. E. Zolensky, P. J. Sylvester and J. B. Paces, Am. Mineral., 1988, 73, 313–323 CAS.
- G. Dantelle, M. Mortier, D. Vivien and G. Patriarche, Chem. Mater., 2005, 17, 2216–2222 CrossRef CAS.
- S. Sivakumar, F. C. J. M. van Veggel and M. Raudsepp, J. Am. Chem. Soc., 2005, 127, 12464–12465 CrossRef CAS PubMed.
- D. Starick, S. I. Golovkova and A. M. Gurvic, J. Lumin., 1988, 40&41, 199–200 Search PubMed.
- F. E. Swindells, J. Electrochem. Soc., 1954, 101, 415–418 CrossRef CAS.
- R. Stefani, A. D. Maia, E. E. S. Teotonio, M. A. F. Monteiro, M. C. F. C. Felinto and H. F. Brito, J. Solid State Chem., 2006, 179, 1086–1092 CrossRef CAS PubMed.
- X. Piao, T. Horikawa, H. Hanzawa and K. Machida, Appl. Phys. Lett., 2006, 88, 161908–161913 CrossRef PubMed.
- W. B. Im, Y. K. Kim and D. Y. Jeon, Chem. Mater., 2006, 18, 1190–1195 CrossRef CAS.
- D. P. Yu, Q. L. Hang, Y. Ding, H. Z. Zhang, Z. G. Bai, J. J. Wang, Y. H. Zou, W. Qian, G. C. Xiong and S. Q. Feng, Appl. Phys. Lett., 1998, 73, 3076–3078 CrossRef CAS PubMed.
- L. D. Carlos, V. de Zea Bermudez, R. A. Sá Ferreira, L. Marques and M. M. Assunção, Chem. Mater., 1999, 11, 581–588 CrossRef CAS.
- T. Uchino, N. Kurumoto and N. Sagawa, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 233203–233206 CrossRef.
- A. Aboshi, N. Kurumoto, T. Yamada and T. Uchino, J. Phys. Chem. C, 2007, 111, 8483–8488 CAS.
- S. Banerjee and A. Datta, Langmuir, 2010, 26, 1172–1176 CrossRef CAS PubMed.
- S. Yang, W. Li, B. Cao, H. Zeng and W. Cai, J. Phys. Chem. C, 2011, 115, 21056–21062 CAS.
- J. M. McCrate and J. G. Ekerdt, Chem. Mater., 2014, 26, 2166–2171 CrossRef CAS.
- T. C. Liu, H. Kominami, H. F. Greer, W. Zhou, Y. Nakanishi and R. S. Liu, Chem. Mater., 2012, 24, 3486–3492 CrossRef CAS.
- J. Rodriguez-Carvajal, Fullprof 2000: A Rietveld Refinement and Pattern Matching Analysis Program, Laboratoire Léon Brillouin (CEA-CNRS), April 2008 Search PubMed.
- S. Zhang, Y. Nakai, T. Tsuboi, Y. Huang and H. J. Seo, Chem. Mater., 2011, 23, 1216–1224 CrossRef CAS.
- W. B. Im, Y. I. Kim, H. S. Yoo and D. Y. Jeon, Inorg. Chem., 2009, 48, 557–564 CrossRef CAS PubMed.
- K. Inoue, N. Hirosaki, R. J. Xie and T. Takeda, J. Phys. Chem. C, 2009, 113, 9392–9397 CAS.
- J. Hlavay, K. Jonas, S. Elek and J. Inczedy, Clays Clay Miner., 1978, 26, 139–143 CAS.
- A. M. Pires and M. R. Davolos, Chem. Mater., 2001, 13, 21–27 CrossRef CAS.
- D. Kim, D. Park, N. Oh, J. Kim, E. D. Jeong, S. J. Kim, S. Kim and J. C. Park, Inorg. Chem., 2015, 54, 1325–1336 CrossRef CAS.
- R. C. Weast, Handbook of Chemistry and Physics, CRC Press, 1990 Search PubMed.
- L. Li, G. Li, Y. Che and W. Su, Chem. Mater., 2000, 12, 2567–2574 CrossRef CAS.
- S. Fujihara and K. Tokumo, Chem. Mater., 2005, 17, 5587–5593 CrossRef CAS.
- Y. P. Du, Y. W. Zhang, L. D. Sun and C. H. Yan, J. Phys. Chem. C, 2008, 112, 12234–12241 CAS.
- E. J. Cho and S. J. Oh, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59(R15), 613–616 Search PubMed.
- J. Qi, T. Matsumoto, M. Tanaka and Y. Masumoto, J. Phys. D: Appl. Phys., 2000, 33, 2074–2078 CrossRef CAS.
- L. Cario, P. Palvadeau, A. Lafond, C. Deudon, Y. Moëlo, B. Corraze and A. Meerschaut, Chem. Mater., 2003, 15, 943–950 CrossRef CAS.
- D. Y. Lu, M. Sugano, X. Y. Sun and W. H. Su, Appl. Surf. Sci., 2005, 242, 318–325 CrossRef CAS PubMed.
- C. Caspers, M. Müller, A. X. Gray, A. M. Kaiser, A. Gloskovskii, C. S. Fadley, W. Drube and C. M. Schneider, Phys. Status Solidi RRL, 2011, 5, 441–443 CrossRef CAS PubMed.
- W. Jiang, Z. Bian, C. Hong and C. Huang, Inorg. Chem., 2011, 50, 6862–6864 CrossRef CAS PubMed.
- E. Lippmaa, M. Mӓgi, G. Engelhardt and A. R. Grimmer, J. Am. Chem. Soc., 1980, 102, 4889–4893 CrossRef CAS.
- J. V. Smith and C. S. Blackwell, Nature, 1983, 303, 223–225 CrossRef CAS PubMed.
- K. A. Smith, R. J. Kirkpatrick, E. Oldfield and D. M. Henderson, Am. Mineral., 1983, 68, 1206–1215 CAS.
- W. Huang, M. Schopfer, C. Zhang, R. C. Howell, L. Todaro, B. A. Gee, L. C. Francesconi and T. Polenova, J. Am. Chem. Soc., 2008, 130, 481–490 CrossRef CAS PubMed.
- S. L. Poulsen, V. Kocaba, G. L. Saoût, H. J. Jakobsen, K. L. Scrivener and J. Skibsted, Solid State Nucl. Magn. Reson., 2009, 36, 32–44 CrossRef CAS PubMed.
- S. Ivanova, E. Zhecheva, R. Stoyanova, D. Nihtianova, S. Wegner, P. Tzvetkova and S. Simova, J. Phys. Chem. C, 2011, 115, 25170–25182 CAS.
- A. Ferreira, D. Ananias, L. D. Carlos, C. M. Morais and J. Rocha, J. Am. Chem. Soc., 2003, 125, 14573–14579 CrossRef CAS PubMed.
- M. B. Ley, D. B. Ravnsbæk, Y. Filinchuk, Y. S. Lee, R. Janot, Y. W. Cho, J. Skibsted and T. R. Jensen, Chem. Mater., 2012, 24, 1654–1663 CrossRef CAS.
- S. Basu, Crystalline Silicon-Properties and Uses, InTech, 2011 Search PubMed.
- R. A. Weeks, Phys. Rev., 1963, 130, 570–576 CrossRef CAS.
- A. S. Mysovsky, P. V. Sushko, S. Mukhopadhyay, A. H. Edwards and A. L. Shluger, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 69, 85202 CrossRef.
- R. H. D. Nuttall and J. A. Weil, Solid State Commun., 1980, 33, 99–102 CrossRef CAS.
- A. C. McLaren, R. F. Cook, S. T. Hyde and R. C. Tobin, Phys. Chem. Miner., 1983, 9, 79–94 CrossRef CAS.
- P. E. Blöchl and J. H. Stathis, Phys. Rev. Lett., 1999, 83, 372–375 CrossRef.
- H. Imai, K. Arai, H. Imagawa, H. Hosono and Y. Abe, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 38, 12772–12775 CrossRef CAS.
- H. Hosono, Y. Abe, H. Imagawa, H. Imai and K. Arai, Phys. Rev. B: Condens. Matter Mater. Phys., 1991, 44, 12043–12045 CrossRef CAS.
- R. D. Shannon, Acta Crystallogr., 1976, 32, 751–767 CrossRef.
- D. Li, X. Zhang, L. Jin and D. Yang, Opt. Express, 2010, 18, 27191–27196 CrossRef CAS PubMed.
- J. Qi, T. Matsumoto, M. Tanaka and Y. Masumoto, Electrochem. Solid-State Lett., 2000, 3, 239–241 CrossRef CAS PubMed.
- L. Li, J. Zheng, Y. Zuo, B. Cheng and Q. Wang, Nanoscale Res. Lett., 2013, 8, 194 CrossRef PubMed.
- M. W. Shafer, J. Appl. Phys., 1965, 36, 1145–1151 CrossRef CAS PubMed.
- P. Atkins and J. Paula, Physical Cehmistry, Oxford University Press, 8th edn, 2006 Search PubMed.
- Y. Luo, Comprehensive handbook of chemical bonding energies, CRC Press, 2007 Search PubMed.
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.