Aditya Sharma*a,
Mayora Varshneya,
Hyun-Joon Shinb,
Keun Hwa Chaea and
Sung Ok Won*a
aAdvanced Analysis Centre, Korea Institute of Science and Technology, Seoul-02792, South Korea. E-mail: adityaiuac@gmail.com; sowon@kist.re.kr
bPohang Accelerator Laboratory (POSTECH), Pohang-37673, South Korea
First published on 14th November 2017
In this study, spectroscopic investigations are employed to quantify the Ga distribution over the tetrahedral/octahedral sites and to assimilate the luminescence properties in the barely reported γ-Ga2.67O4:Sm nanoparticles. O K-edge XANES are convincing the decreased electron density in the outermost hybrid orbits of O 2p and Ga 4p, and the movement of p electrons to the inner hybrid orbits with s character under the distorted oxygen environment of γ-Ga2.67O4:Sm nanoparticles. The Ga K/L-edge XANES and Sm L3/M5,4-edge XANES results have confirmed Ga3+ and Sm3+ ions, respectively, in the γ-Ga2.67O4:Sm nanoparticles. Quantitative determination of the cation distribution is performed by applying Ga K-edge XANES data analysis, which is further substantiated by an EXAFS data simulation, and conveys a Ga-O4 tetrahedra/Ga-O6 octahedra ratio (Ga(t)/Ga(o)) of ∼0.9, ∼1.4, ∼1.5 and ∼1.6 for the β-Ga2O3, γ-Ga2.67O4, γ-Ga2.67O4:5Sm and γ-Ga2.67O4:10Sm samples, respectively, which signify the Sm doping induced deformation of Ga-O6 octahedra via the formation of oxygen defects. Multiple luminescence centres, facilitated by the Ga-O6 octahedra distortion, O vacancies and Sm3+ ions, have helped in a significant enhancement of the emission characteristics. Our spectroscopy investigations, with the help of XANES and EXAFS, may open new opportunities to the mechanistic understanding of metal-O polyhedra alteration induced fruition of physical/chemical properties in the less explored meta stable phases of other complex oxides.
Most of the reports on γ-Ga2O3 have considered the cubic unit cell of this compound;8,10–12 however, a defective spinel structure of γ-Ga2O3 (space group; Fdm) was recently identified by a neutron diffraction technique with four different site occupancies of Ga: tetrahedral (8a), octahedral (16d), tetrahedral (48f) and octahedral (16c).3,4 Unlike the normal spinel-structure stoichiometry in γ-Ga2O3 (ideally, it should be Ga3O4), among every 18 cationic sites, two sites are vacant in order to retain charge neutrality.3,4 Pair distribution function analysis was also employed to refine the γ-Ga2O3; however, the discrepancy between the average and local structure could not be resolved.4 If such complications were not enough during the simulation of the γ-Ga2O3 structural details, investigations on the local structure determination would further be complicated in the cation deficient spinel phases of gallium oxides, such as γ-Ga3−δO4. The cation deficient γ-Ga3−δO4 phase is very similar to that of a hypothetical spinel Ga3O4 compound; however, less investigation has been made because of its poor thermal stability during its synthesis.
Theoretical studies have shown that the band structure of Ga2O3, near the Fermi level, is formed by the O 2p and Ga 4s states. The top of the valence band consists mainly of the O 2p states, with a width of 7 eV, and the bottom of the conduction band is formed by Ga 4s states.13–15 Ga d orbitals do not contribute to the band structure near the Fermi level, because of the 3d10 electronic configuration of Ga d orbitals in Ga2O3, but facilitate a deep valence band along with the O 2s orbitals.13,14 The octahedrally coordinated Ga atoms possess higher coordination with the surrounding O atoms, compared with the tetrahedrally coordinated Ga atoms, and, thus, the degree of covalency is higher for the tetrahedrally coordinated Ga atoms in the Ga2O3 lattice.13–15 The band structure and, hence, the electronic/optical properties of Ga2O3 are not only sensitive to the doping of foreign element at Ga lattice sites16,17 but may also be detonated by the distortion of the Ga–O environment or alteration in the Ga site occupancy. A small change in the Ga tetrahedral/octahedral site occupancy is expected to affect the orbital hybridization of frontier Ga 4s/4p and O 2p orbitals which, obviously, can regulate the band structure of the compound and consequently electronic/optical properties.
In the previous reports, nano/micro sized β-Ga2O3 particles have been extensively studied for their synthesis,9,18–23 band structure,15 optical18,19 and catalyst properties.21–23 The other polymorphous phases of gallium oxide can demonstrate exceptional metal–oxygen orbital hybridization and a perceptibly diverse band structure near the Fermi level because of the miscellaneous Ga site occupancy in the Ga2O3 lattice; however, their synthesis, crystal structure analysis, optical properties and local atomic/electronic structure investigations are ineffectually reported. Thus, an atomistic level investigation on the unidentified or less studied polymorphous phases, such as the cation deficient γ-Ga3−δO4, is desirable to gain control over and fabricate their nanostructures for light-emitting applications. Though the synthesis of small sized nanostructures is important for catalytic21–23 and optical properties,10–12,18–20 the tiny nanostructures introduce complications in the structural and electronic structure investigations because of the significant surface artefacts.10–12,18–20 The sensitivity of the powder XRD techniques becomes poor to the structural details when the particle size reaches up to a few nanometres.24,25 Moreover, for the case of nanocrystalline Ga2O3 polymorphs, the diverse occupancies of Ga atoms and cation/anion deficiencies facilitate a complex local atomic structure to the compound which may not be completely elucidated by the conventional XRD technique.2,3 Therefore, it is necessary to apply modern techniques which not only probe the crystal structure ‘locally’ but also provide valuable information on the lattice disorders upon defect formation, foreign element doping and alteration in the metal element occupancy. In this regard, X-ray absorption spectroscopy (XAS), with its two branches of X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS), becomes the important elemental characterization technique and has been known to be highly sensitive to the local electronic/atomic structure properties and effectively overcomes the issue of sensitivity to the structure and electronic structure characterization of small sized, poorly crystalline or embedded nanostructures.26–28 In the last few years, probing the local symmetry of the diluted doped elements,29,30 quantification of the valence state changes of host metal ions upon foreign element doping,31–33 changes in the orbital occupancy,34,35 and variation in the crystal field effects36,37 have been studied for various compounds using the XANES. On the other side, EXAFS, due to the sensitivity to the local atomic structure, has become a powerful probing tool for determining the variation in the coordination number and bond lengths of the constituent elements of the probed material.32–40 Besides providing detailed experimental and simulated results on the Ga2O3 polymorphous compounds, the XANES–EXAFS analysis presented in this article addresses various aspects. Firstly, O K-edge XANES spectra were examined to specifically look into the diversity in the oxygen environment of γ-Ga3−δO4 and β polymorphous phases and to assimilate the effect of Sm insertion on the metal 4s/4p and O 2p orbital hybridization. Secondly, quantitative determination of the cation distribution is performed by using Ga K-edge XANES and EXAFS to estimate the Ga-O4 tetrahedra and Ga-O6 octahedra proportions in β-Ga2O3, γ-Ga3−δO4 and Sm doped γ-Ga3−δO4. To our knowledge, the Ga-O6/Ga-O4 octahedra/tetrahedra deformation under Sm incorporation and enhancement in the emission characteristics of γ-Ga3−δO4:Sm nanoparticles are barely covered in the literature. Our spectroscopy investigations may provide mechanistic insights into the modification in the orbital hybridization of Ga 4s/4p and O 2p, Ga local site occupancy, defect formation and emission properties of the metastable γ-Ga3−δO4 phase.
Fig. 1 (a) XRD patterns of pure and Sm doped γ-Ga2.67O4. (b) XRD patterns of the samples annealed at 600 °C and 900 °C. Spectra are vertically shifted for clarity of the data. |
In previous reports, synthesis with DI water and liquid ammonia solution could only grow GaOOH kinds of phases which further transformed into α-Ga2O3, γ-Ga2O3 and β-Ga2O3 phases upon high-temperature annealing.18,19 Thus, the formation of the γ-Ga2.67O4 phase in the present study, at moderate temperate, is due to the modified non-aqueous synthesis procedure with the amendment of a reactive hydrazine monohydrate precursor. No diffraction peak matched with the other polymorphous of gallium oxide (i.e., α-Ga2O3, β-Ga2O3, δ-Ga2O3 and ε-Ga2O3) or trivial phases of metallic Sm, SmO and Sm2O3 and, thus, the XRD results strengthen the manufacturing of single phase γ-Ga2.67O4:Sm in the present study. Significant XRD peak broadening and diminishing of peak intensity is seen in the Sm doped samples and could be due to a decrease of particle size and/or lattice deformation. We have calculated the crystallite size from all of the samples using the Scherrer relation; D = 0.9λ/βCosθ (D is the particle size, λ is the wavelength of the X-rays used, β is the full width at half maximum of the diffraction peak, and the instrument broadening effect was not included in such calculations). Thus, the calculated sizes are 4.3 nm, 2.9 nm, 2.1 nm, 13.2 nm and 32.5 nm for the γ-Ga2.67O4, γ-Ga2.67O4:5Sm, γ-Ga2.67O4:10Sm, 600 °C annealed and 900 °C annealed samples, respectively. The XRD patterns also give an impression that there is a lower angle shifting in the diffraction peaks of the Sm doped samples, compared with pure γ-Ga2.67O4 (this can be easily seen in the γ-Ga2.67O4:10Sm sample). Thus, d(311) was examined and found to be 2.488(6) Å, 2.495(2) Å and 2.519(2)Å for the γ-Ga2.67O4, γ-Ga2.67O4:5Sm and γ-Ga2.67O4:10Sm samples, respectively. The marginal increase in the d(311) values for Sm doped samples may arise due to the occupancy of larger Sm3+ ions (ionic radii ∼ 109.8 pm) at the lattice position of smaller Ga3+ ions (ionic radii, ∼78.0 pm). The decrease in crystallite size and increase in the d(311) values indicate the inhibition of crystalline growth and structural perturbation in the crystalline lattice of gallium oxide upon Sm3+ ion doping. Here we expect that the insertion of Sm3+ ions in the crystalline lattice of gallium oxide produces a strain (due to the larger ionic radii of Sm3+ over Ga3+) which, conceivably, alters the metal–oxygen polyhedra and leads to a decline in the crystallinity of the material. Moreover, in previous reports, doping of a foreign element evidenced the non-stoichiometry of the metal–oxygen chains or segregation of the doped ions at surface/grain boundaries, which prevents the crystalline growth of the particle.12,42 The non-stochiometric metal–oxygen grain boundaries, which can be formed by lattice-strain induced diverse metal–oxygen stacking or segregation of dopants, may increase the amorphous content in the γ-Ga2.67O4:Sm nanoparticles, leading to broadening of the XRD peaks and a decrease in the peak intensity. In our previous reports,32,33 a similar trend in the XRD peak broadening and increase in the cell parameters/d(hkl) was observed for the Eu and Hf doped CeO2 NPs and was ascribed to the Eu/Hf induced lattice disorders in CeO2. Raman measurements are also convincing of the poor crystallinity in the Sm doped samples and have also ruled out a suspected phase formation (Fig. S1 and related discussion in the ESI†). The decrease in the particle size and the poor crystallinity are also seen in the TEM and FFT images of the γ-Ga2.67O4:5Sm and γ-Ga2.67O4:10Sm samples and tally with the findings from XRD results (Fig. S2 and related discussion in the ESI†). XRD patterns in Fig. 1(b) are certainly dissimilar with respect to the XRD patterns of γ-Ga2.67O4 and suggest a different phase formation in the 600 °C and 900 °C annealed samples. XRD patterns of the 600 °C and 900 °C annealed samples are fairly matched with the monoclinic unit cell of the β-Ga2O3 phase (space group C2/m; JCPDS#02-2603). Hereafter, the 600 °C and 900 °C annealed samples are named as Ga2O3_600 and Ga2O3_900, respectively. Raman studies have also confirmed the β-Ga2O3 phase formation in the annealed sample (see Fig. S1 and related discussion in the ESI†). In previous reports, β-Ga2O3 phase formation from the other polymorphous phases was achieved at or above 550 °C.18,19 Therefore, our findings are in accordance with the reported results and are convincing of the annealing induced β-Ga2O3 phase formation in the Ga2O3_600 and Ga2O3_900 samples. Although the β-Ga2O3 phase is present in both of the annealed samples, narrower XRD peaks with a higher intensity are seen in the Ga2O3_900 sample. XRD peak intensification in the Ga2O3_900 sample is due to the growth of larger sized particles and stabilization of the β-Ga2O3 phase under the high-temperature treatment. Larger sized particles and higher crystallinity is clearly seen in the TEM and FFT images of Ga2O3_600 and Ga2O3_900 samples (see Fig. S2 and related discussion in the ESI†).
By considering the Δ0 of the samples, we expect that the fractional occupancy of Ga atoms at tetrahedral sites is larger in the case of γ-Ga2.67O4 polymorphs and the octahedral sites are the favourable sites of Ga atoms in the β phasic samples. In a previous study, the preferred octahedral sites occupancy for the dopants were reported in spinel phase MnFe2O4 nanoparticles.50
Here, we may hypothesize that the Sm doping in the γ-Ga2.67O4 lattice offers more tetrahedral sites to the metal atoms at the expense of octahedral sites. Under the condition of substitutional doping of Sm at the Ga sites, if we assume that isovalent Sm3+ ions take the octahedral sites of Ga3+ ions, O vacancies are expected to form because of the disorder/strain produced via imbalance of ionic radii and electronegativity of Sm3+ and Ga3+ ions. Formation of O vacancies has also been reported in the isovalent (i.e., Sn4+, Hf4+ and Zr4+) element doped CeO2 nanoparticles (Ce4+ ions in CeO2), and was ascribed to the imbalance of ionic radii and lattice distortion.32,33,51 In the case of Ga-O4 tetrahedra, there are only four O ligand approaches to the Ga atoms (see the schematic in Fig. 3(iii)). Similarly, six O ligands form Ga-O6 octahedra with the Ga atoms (see the schematic in Fig. 3(iv)). Reduction of metal-O6 octahedra into metal-O4 tetrahedra is possible via O vacancy formation and is shown in the Fig. 3(v). Further analysis at Ga K-edge XANES and EXAFS is helpful to elaborate such a hypothesis of an increase in the metal-O4 tetrahedral sites and decrease in the metal-O6 octahedral sites in the Sm doped γ-Ga2.67O4 and is discussed in the following sections.
The Ga K-edge XANES originates from the 1s → 4p transitions and can be used for understanding the inconsequential local electronic structure differences among the polymorphous Ga2O3 compounds.48,49,52–54 Fig. 4(a) shows the Ga K-edge XANES spectra of γ-Ga2.67O4, γ-Ga2.67O4:5Sm and γ-Ga2.67O4:10Sm samples while Fig. 4(c) shows the second derivates of the Ga K-edge XANES spectra. Here, we have analysed the Ga K-edge XANES spectra by concerning the two aspects; (i) valence state change of Ga3+ ions and (ii) perturbation in the Ga lattice site occupancy (i.e., change in the Ga-O6 octahedra and Ga-O4 tetrahedra proportions), if any, upon Sm doping or annealing conditions. The edge-energy position and other spectral features of the pure and Sm doped γ-Ga2.67O4 samples fairly match each other (can be seen in the second derivate spectra) and nullify any valence state change upon Sm doping. Similar to the Ga K-edge XANES, the Sm L-edge XANES spectra and second derivate spectra of the γ-Ga2.67O4:5Sm and γ-Ga2.67O4:10Sm samples do not show any noteworthy changes and, thus, the possibility of valence state change of Sm ions can be neglected. The Ga K-edge and Sm L-edge XANES results were further cross-checked with the Ga L-edge and Sm M5,4-edge XANES and convey the +3 valence state of Ga and Sm ions (see Fig. S3 and S4 in the ESI† and discussion therein). Sm L-edge and Sm M-edge XANES are convincing of Sm3+ ions in the samples but the existence of Sm at the surface/grain boundaries of NPs may not be explained within the periphery of the XANES results. Interesting spectral changes were observed in the Ga K-edge XANES spectra, while we have analysed the white-line region of γ-Ga2.67O4 and Ga2O3_900 samples, by normalizing the white-line peak intensity (see the inset of Fig. 4(a)). It is evidenced from the figure that the Ga2O3_900 sample (i.e., β-Ga2O3 phase) exhibits a broad white line peak compared with the sharp peak in γ-Ga2.67O4. It seems that the white-line peak position of the β-Ga2O3 sample is at a slightly higher energy (∼2.2 eV) compared with the γ-Ga2.67O4 sample. It is known for the transition metal compounds that the edge-energy and white-line peak position progressively shift towards a higher energy with increasing valence state (cation charge) of metal ions.55 In the present study, the edge-energy positions are nearly identical; however, the white line peak positions give the impression of positive energy shifting for the β-Ga2O3 compound with respect to γ-Ga2.67O4. In a previous report, the Al K-edge XANES also showed white-line peak shifting towards a higher energy with an increase of the oxygen coordination number of Al from 4 to 6 (i.e., sodalite to kyanite transformation).56 Similarly, a Ga K-edge XANES spectrum of α-Ga2O3 has been reported with a +2 eV shifting in the white-line peak position with respect to GaAs despite the same valence state of the Ga ions (i.e., Ga3+) in both of the compounds.52 The existence of a higher energy white-line peak in α-Ga2O3 and a lower energy white-line peak in GaAs was ascribed to the diverse site occupancy of Ga atoms.48,52 In the GaAs compound, the Ga atoms occupy regular tetrahedral sites, whereas in α-Ga2O3 the Ga atoms are in distorted octahedral sites.52 By considering the previous reports and findings of our Ga K-edge XANES, the low energy white-line peak in the present study of γ-Ga2.67O4 sample can be anticipated as a favourable tetrahedral occupancy of Ga atoms, whereas the broadened or positive energy shifted white-line peak in the Ga2O3_900 sample may be an indication of mixed occupancy of the Ga atoms at tetrahedral and octahedral sites.
To understand the quantitative Ga occupancy in Ga2O3_900, γ-Ga2.67O4, and Sm doped γ-Ga2.67O4, we have deconvoluted the white line peak of Ga K-edge XANES and present it in Fig. 5(a–d). The deconvolution was done by assuming the two components in the main peak region (i.e., 10355 eV to 10385 eV) and the fittings were done by applying the Gaussian curve fitting method.48,49 The details of the fitting procedure and the other parameters like peak energy, full width at half maxima of the peaks and the fractional percentages of tetrahedral and octahedral Ga atoms (i.e., Ga(t)% and Ga(o)%) in various samples are provided in Table S1 of the ESI.† According to previous reports,48,49,52 the Gaussian 1 (magenta colour) and Gaussian 2 (olive colour) in Fig. 5 can be regarded as the contribution from the tetrahedral (Ga(t)) and octahedral (Ga(o)) occupancy of Ga atoms, respectively. The integrated areas of Gaussian 1 and Gaussian 2 have been applied to determine the fractional percentage of Ga(t) and Ga(o) from various samples. Detailed calculations are provided in the ESI.† Our calculations convey that the Ga2O3_900 sample exhibits 48.4% and 51.6% of Ga(t) and Ga(o) atoms, respectively (i.e., Ga(t)/Ga(o) ratio of ∼0.92). In previous studies, a 1:1 ratio of Ga(t):Ga(o) has been reported for the commercial β-Ga2O3 compound.48,49 A slightly smaller Ga(t):Ga(o) ratio from the Ga2O3_900 sample may be due to the higher occupancy of Ga atoms on octahedral sites. On the other hand, the γ-Ga2.67O4 sample has shown 58.8% and 41.2% fractional percentages of Ga(t) and Ga(o), respectively (i.e., Ga(t)/Ga(o) ratio of ∼1.4). Ga(t)/Ga(o) ratios were further increased for the γ-Ga2.67O4:5Sm (∼1.51) and γ-Ga2.67O4:10Sm (∼1.62) samples. The highest Ga(t)/Ga(o) ratio was reported to be ∼1.8 for the poorly crystalline and small sized Ga2O3 samples48 and is in agreement with our results. Ga K-edge XANES results imply that the γ phasic samples have preferred tetrahedral sites of Ga species and the substitution of Sm leads to much octahedral site deformation. Population of Ga atoms at tetrahedral (Ga-O4) sites at the expense of octahedral (Ga-O6) sites may facilitate O ion vacancy in the γ-Ga2.67O4 lattice leading to more unoccupied O 2p hybridized orbitals in the Sm doped samples, as illustrated in the O K-edge XANES. The smaller Δ0 in the O K-edge spectra is analogous to the higher Ga(t):Ga(o) ratio for the pure and Sm doped samples and is convincing of the higher metal-O4 tetrahedra in the γ-Ga2.67O4, γ-Ga2.67O4:5Sm and γ-Ga2.67O4:10Sm samples, compared with β-Ga2O3.
Name | Ga sites | Shell | N | R (Å) | σ2 (Å2) | E0 (eV) | Ga(t)/Ga(o) |
---|---|---|---|---|---|---|---|
Ga2O3_900 | Gao | Ga–O | 5.4 | 1.92(3) | 0.012(2) | 3.0 | 0.85 |
Ga–Ga | 5.4 | 3.09(3) | 0.009(5) | ||||
Gat | Ga–O | 3.0 | 1.82(7) | 0.047(0) | |||
Ga–Ga | 6.1 | 3.34(7) | 0.008(0) | ||||
γ-Ga2O3 | Gao | Ga–O | 4.1 | 2.01(7) | 0.026(4) | 2.5 | 1.17 |
Ga–Ga | 4.7 | 2.97(1) | 0.007(4) | ||||
Gat | Ga–O | 3.2 | 1.85(0) | 0.002(0) | |||
Ga–Ga | 5.7 | 3.34(2) | 0.012(4) | ||||
γ-Ga2O3:5Sm | Gao | Ga–O | 3.7 | 2.04(9) | 0.011(8) | 2.2 | 1.37 |
Ga–Ga | 3.9 | 3.01(4) | 0.008(9) | ||||
Gat | Ga–O | 3.4 | 1.90(8) | 0.007(1) | |||
Ga–Ga | 5.1 | 3.45(0) | 0.047(0) | ||||
γ-Ga2O3:10Sm | Gao | Ga–O | 3.5 | 2.06(9) | 0.038(5) | 2.7 | 1.59 |
Ga–Ga | 3.5 | 3.11(3) | 0.007(0) | ||||
Gat | Ga–O | 3.3 | 1.92(7) | 0.004(8) | |||
Ga–Ga | 4.5 | 3.46(9) | 0.018(8) |
It is also noticeable from Fig. 6(b) that the intensity of the Ga–Ga shell (between the 2–3.4 Å) is quite smaller than that of the Ga–O shell intensity, while it was quite comparable with that of the Ga2O3_900 sample. This is due to the fact that the Ga2O3_900 sample has a high crystallinity and larger particle size compared with the γ-Ga2.67O4 sample. Diminishing the Ga–Ga shell intensity in Fig. 6(c and d), for the γ-Ga2.67O4:Sm samples, could be related to the net decrease in the nanoparticle size and increase in the amorphous content in the samples. In previous reports, a continuous decrease in the EXAFS oscillations with a decrease in the nanoparticle size is also reported.59 It is noticeable from Table 1 that the Ga–O and Ga–Ga bond distances are longer for the Sm doped sample compared with pure Ga2.67O4. The net increase in the Ga–O or Ga–Ga bond distances is, certainly, due to the structural perturbation in the crystalline lattice of gallium oxide upon Sm3+ ion doping. In the XRD studies, we have noticed an enlargement of d(311) values for the Sm3+ ion doped samples. Thus, the enlarged Ga–O and Ga–Ga bond distances are the consequences of the insertion of larger Sm3+ ions at the lattice position of Ga3+ ions. In previous reports,48,49 Ga–O shell coordination numbers were applied to calculate the fractional Ga(t) and Ga(o). We have also evaluated the Ga(t)/Ga(o) = (N/4)/(N/6) ratio from the coordination number of Ga–O shell.48 It is noticeable from Table 1 that the Ga(t)/Ga(o) ratios, for all of the samples, are nearly the same as obtained from the Ga K-edge XANES. In the previous EXAFS analysis,48 the maximum Ga(t)/Ga(o) ratio was evaluated to be ∼1.7 for the poorly crystalline and small sized (2–6 nm) Ga2O3 NP samples and is in accordance with our findings. Our XANES and EXAFS results are, thus, convincing of favourable tetrahedral sites of the Ga atoms in γ-Ga2.67O4 which are further increased in the Sm doped samples leading to higher Ga(t)/Ga(o) ratios.
Fig. 7 Room temperature PL spectra of (a) pure γ-Ga2.67O4, (b) γ-Ga2.67O4:5Sm, (c) γ-Ga2.67O4:10Sm, (d) Ga2O3_600 and (e) Ga2O3_900 samples. |
While comparing the PL spectra of γ-Ga2.67O4 nanoparticles and the annealed samples (Ga2O3_600 and Ga2O3_900), it is noticeable that the annealed samples show a dominating green emission at ∼548 nm and a negligible blue emission band. The red shift in the PL spectra, comparing with pure γ-Ga2.67O4, indicates a smaller average energy separation between the recombining pairs in the annealed samples. The annealing induced ripening and fusion of particles, as evidenced in the TEM study, may have hidden the surface defect induced shallow energy levels in the energy gap of the material, leading to the quenching of the blue emission from the Ga2O3_600 and Ga2O3_900 samples. Only the interior defects , which have a smaller energy separation of donor–acceptor pairs or more Coulomb interactions, are expected to be responsible for the green emission from the Ga2O3_600 and Ga2O3_900 samples. An interesting variation can be seen in the PL spectra of Sm doped samples: (i) a low-intensity band evolves in the deep violet region (∼396 nm), (ii) the intensity of the ∼434 nm band is increased and (iii) intense bands appear between 550 nm and 675 nm. Taken together, XRD, XAS and PL data indicate that Sm doping in γ-Ga2.67O4 nanoparticles causes a change in DAP recombination because of several factors: (i) a decrease in the average particle size, which provides more surface related defect states,63 (ii) alteration in the Ga and O local lattice site occupancy, which gives diverse defect levels in the compound18–20 and (iii) the existence of the Sm3+ ion in the γ-Ga2.67O4 lattice, which provides distinct defect states and intra-4f transitions.64,65 The broad range of luminescence from the γ-Ga2.67O4:Sm nanoparticles can be understood by knowing the energy and charge transfer process in distinct electronic states.
In Fig. 8, we present a schematic of the DAP and Sm3+ ion transition induced luminescence in the γ-Ga2.67O4:Sm nanoparticles. Excitation by the 325 nm wavelength from a He–Cd laser leads to electronic transitions and trapping of photo excited electrons and holes in the donor and acceptor states, respectively. Recombination of electrons and holes leads to typical emission spectra from γ-Ga2.67O4 nanoparticles. The energy (or colour) of emitted radiation is governed by the coulombic interaction between the charge donor and acceptors and can be expressed as;60–62 E = Eg − (Edonor + Eacceptor) + Ecoulombic ± nEphonon. Here, Eg is the band gap energy of the material, Edonor and Eacceptor are the donor and acceptor binding energies, respectively. The Ephonon term represents the energies of phonons involved in the radiative transitions and the coulombic interaction term (Ecoulombic = e2/4πεr) is the deciding term for the emitted energy (or colour) from a DAP transition. Since the average donor–acceptor separation (r) becomes narrower for the smaller sized particles,60–62 a higher emission energy can be expected from the above relation for a smaller sized particle. The observed violet emission (∼395.5 nm) and increased intensity of the ∼434 nm band in the Sm doped samples may be the consequence of an Sm doping induced decrease of particle size and increase in the defect concentration. Our O K-edge and Ga K-edge XANES/EXAFS are convincing of O and Ga defect formation and Ga lattice site occupancy alteration upon the insertion of Sm3+ ions. These kinds of defect states are certainly responsible for the intensification of the ∼434 nm band.
Donor and acceptor states not only provide DAP induced radiative transitions but also play a role in the sensitization of the Sm emission by nonradiative transfer of the DAP exciton energy to the doped rare earth ions (shown by a dashed black line in Fig. 8).12 The 4f electrons of rare-earth ions in the host materials are quite similar to the free ions and the emission characteristics are determined by the crystalline fields of the host compound around the rare-earth ions.12,64 In several reports, the 3+ valence state of rare-earth ions was reduced to 2+, even under in situ XAS data collection, leading to intriguing electronic structure and luminescence properties.12,65 In the present study, our Sm L-edge and Sm M5,4-edge XANES, along with the multiplet calculations, have confirmed Sm3+ ions in the samples, and therefore Sm3+ reduction into Sm2+ can be neglected and the observed luminescence properties are the consequences of defect formation and the Sm3+ ion related 4G5/2 → 6Hj transition induced emission properties. Despite the enhancement in the defect induced luminescence bands (violet-blue-green bands) the intensity of the Sm3+ ions induced transitions,64,65 4G5/2 → 6H5/6 (yellow emission), 4G5/2 → 6H7/6 (orange emission) and 4G5/2 → 6H9/6 (red emission), are also enhanced, conveying that the luminescence properties of γ-Ga2.67O4 can be engineered by the formation of controlled defect states and incorporation of variable Sm3+ ion concentration. The photoluminescence spectra of the samples were converted to the Commission International de I'Eclairage (CIE-1913) chromaticity diagram and are presented in the Fig. 9. It is visible from the CIE colour chromaticity diagram that the pure γ-Ga2.67O4 nanoparticle sample exhibits DAP induced cyan-green emission characteristics (CIE coordinates, 0.26, 0.43) while the Ga2O3_600 (CIE coordinates, 0.36, 0.45) or Ga2O3_900 (CIE coordinates, 0.37, 0.46) samples exhibits green yellow emission which is analogous to the previous reports where the increase of particle size had demonstrated the red-shifted PL bands.60,61 By considering the standard CIE coordinates for the white light (i.e., 0.333, 0.333),66 we can state that the Sm doped γ-Ga2.67O4 nanoparticles have shown a nearly white line emission, i.e., ranging from violet-blue to orange-red emission. The CIE coordinates for the γ-Ga2.67O4:5Sm and γ-Ga2.67O4:10Sm samples are (0.28, 0.34) and (0.30, 0.35), respectively.
Fig. 9 CIE colour chromaticity diagram of sample 1 (pure γ-Ga2.67O4), sample 2 (γ-Ga2.67O4:5Sm), sample 3 (γ-Ga2.67O4:10Sm), sample 4 (Ga2O3_600) and sample 5 (Ga2O3_900). |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c7ra10341g |
This journal is © The Royal Society of Chemistry 2017 |