Acessibilidade / Reportar erro

Synthesis and Characterization of Nb2O5:La3+,Eu3+ Phosphors Obtained by the Non-Hydrolytic Sol-Gel Process

Abstract

The phosphor-converted light-emitting diode technique is an important solid-state illumination strategy. Sulfide-based materials are the most often employed red phosphors, but they are chemically unstable and present lower efficiency in comparison to the blue and green phosphors. Therefore, it is important to find a new red phosphor source that can emit intense red light while absorbing light in the near ultraviolet (UV) spectral region. This paper describes the photoluminescence properties of Nb2O5:La3+,Eu3+ obtained by the non-hydrolytic sol-gel process. The X-ray results indicated that the thermal treatment allowed to obtain the different crystalline structures such as, the orthorhombic and monoclinic phases for the Nb2O5 and the orthorhombic phase for the La2Nb10O28. This polymorphism was also confirmed by the Raman spectroscopy. The luminescence spectra revealed the existence of the Eu3+ ions in both crystalline phases for the samples annealed at higher temperature, depending of the excitation wavelength. The emission spectrum showed that increasing the annealing temperature promotes the narrowing of all intraconfigurational f-f transitions for the Eu3+ ions, due to the structural changes. In addition, all samples present good CIE (International Illumination Committee) chromaticity coordinates when excited in the UV (275 and 394 nm), blue (465 nm) and green (525 nm) radiation.

Keywords:
non-hydrolytic sol-gel route; niobia; red phosphor; luminescence


Introduction

Currently, solid-state white light emitters (LEDs) are considered the next generation of solid-state light sources. Because LEDs offer advantages such as longer lifetime and low energy consumption, they constitute potential replacements for incandescent and fluorescent light bulbs.11 Huang, J.; Zhou, L.; Liang, Z.; Gong, F.; Han, J.; Wang, R.; J. Rare Earths2010 , 28, 356.,22 Neeraj, S.; Kijima, N.; Cheetham, A. K.; Chem. Phys. Lett. 2004 , 387, 2. Bearing in mind that illumination consumes about 33% of all the generated energy, the development of energy-saving systems has become fundamental from a technological standpoint.33 Thejo Kalyani, N.; Dhoble, S. J.; Renewable Sustainable Energy Rev. 2012 , 16, 2696.

Ultraviolet (UV) radiation provides high excitation energy, so a new strategy to obtain white light relies on the use of near UV radiation (ca. 400 nm) by LEDs coated with blue/green/red luminophores.44 Wu, Z. C.; Shi, J. X.; Wang, J.; Gong, M. L.; Su, Q.; J. Solid State Chem. 2006 , 179, 2356. An ultraviolet light emitting diode composed of indium gallium nitride (InGaN) or gallium nitride (GaN) present a high light emission efficiency in the light emission wavelengths from 370 nm to 410 nm, and particularly, has the highest light emission efficiency at wavelengths around 390 nm.55 Zhou, L.; Huang, J.; Gong, F.; Lan, Y.; Tong, Z.; Sun, J.; J. Alloys Compd. 2010 , 495, 268. Yttrium oxysulfide doped with Eu3+ (Y2O2S:Eu3+) is the red-light luminophore that is commercially available at present; however, it is less efficient than luminophores that emit blue and green light, has shorter lifetime under UV radiation, and is unstable because it releases hydrogen sulfide.66 Ferrari, J. L.; Cebim, M. A.; Pires, A. M.; dos Santos, M. A. C.; Davolos, M. R.; J. Solid State Chem. 2010 , 183, 2110.

7 Kodaira, C. A.; Stefani, R.; Maia, A. S.; Felinto, M. C. F. C.; Brito, H. F.; J. Lumin. 2007 , 127, 616.

8 Pires, A. M.; Serra, O. A.; Davolos, M. R.; J. Alloys Compd. 2004, 374, 181.
-99 Kutty, T. R. N.; Nag, A.; J. Mater. Chem. 2003 , 13, 2271. In this context, achieving novel and stable luminophores that can emit intense red light, exhibits excitation in the near UV region, and presents CIE (International Illumination Committee) chromaticity coordinates that meet the demands of the National Television Standard Committee (NTSC) is highly desirable.

To produce white light, the current phosphor materials based on near UV GaN-LEDs are BaMgAl10O17:Eu2+ (BAM) for blue, ZnS:(Cu+, Al3+) for green, and Y2O2S:Eu3+ for red. However, compared with the blue and green phosphors, the efficiency of Y2O2S:Eu3+ is about eight times lower and a mixture containing 10% blue, 10% green and 80% red is necessary to obtain appreciable color rendering.22 Neeraj, S.; Kijima, N.; Cheetham, A. K.; Chem. Phys. Lett. 2004 , 387, 2. Therefore, it is mandatory that scientists design new processes and search for novel materials to synthesize stable luminophores that can efficiently absorb near UV radiation while emitting red light. In this sense, niobium oxide doped with La3+ and Eu3+ has emerged as a promising material for multifunctional applications.

Nb2O5 is transparent over a wide range of wavelengths (0.35-9.0 µm), it has a wide band-gap (3.6 eV), it is stable under near UV radiation, it has a relatively low cut-off phonon energy (900 cm-1), high refractive index (2.4), high permittivity (29 to 200 depending of the crystalline phase) and undergoes polymorphic transformations induced by treatment temperature.1010 Aquino, F. T.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Opt. Mater. (Amsterdam, Neth.)2013 , 35, 387.

11 Aquino, F. T.; Pereira, R. R.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Mater. Chem. Phys. 2014 , 147, 751.

12 Palatnikov, M.; Shcherbina, O.; Sidorov, N.; Skab, I.; Bormanis, K.; Ukr. J. Phys. Opt.2012 , 13, 207.
-1313 Rooksby, H. P.; White, E. A. D.; Acta Crystallogr. 1963 , 16, 888. All these properties enhance the optical features of rare earth ions, since efficient radiative emission output requires a relatively high refractive index and a low nonradioactive decay for excited states. The amorphous niobium oxide (V) begins to crystallize in a low-temperature form called the (T) form or γ-phase (orthorhombic, space group: Pbam) at about 500 ºC. Crystallization occurs more rapidly at higher temperatures until about 800 ºC, where upon a transition to a medium-temperature (M) form or β-phase (monoclinic, space group: I4/mmm). This transition continues more rapidly at higher temperatures and heating for 4 h at 1000 ºC brings about a complete conversion. At even higher temperatures, a third transformation to a high-temperature (H) form or β;-phase (monoclinic, space group: P2/m) is observed. These polymorphic transitions take place slowly, at temperatures which are not well-defined, and are irreversible.1414 Emmenegger, F. P.; Robinson, M. L. A.; J. Phys. Chem. Solids1968 , 29, 1673.

15 Nowak, I.; Ziolek, M.; Chem. Rev. 1999 , 99, 3603.
-1616 Blanquart, T.; Niinistö, J.; Heikkilä, M.; Sajavaara, T.; Kukli, K.; Puukilainen, E.; Xu, C.; Hunks, W.; Ritala, M.; Leskelä, M.; Chem. Mater. 2012 , 24, 975.

Luminescent materials can absorb light from diverse sources and emit radiative electromagnetic energy in the visible, ultraviolet, and infrared regions.1717 Blasse, G.; Grabmaier, B. C; Luminescent Materials; Springer: Berlin, 1994. Among the various methodologies that are available to synthesize luminophores, the non-hydrolytic sol-gel process stands out as one of the most advantageous: it yields highly pure products (the metallic oxides originate in situ) with fewer pores; occurs at relatively low temperatures; allows for incorporation of organic species, to give more homogeneous hybrid materials; and is easier to reproduce.1818 Nassar, E. J.; Avila, L. R.; Pereira, P. F. S.; de Lima, O. J.; Rocha, L. A.; Mello, C.; Ciuffi, K. J.; Carlos, L. D.; Quim. Nova2005 , 28, 238.

In the traditional sol-gel process, the water molecules within the Eu3+ ion coordination sphere contribute to a non-radiative decay, which has increased the interest in alternative low-temperature processes that take place in the absence of water.1919 Ricci, G. P.; Rocha, Z. N.; Nakagaki, S.; Castro, K. A. D. F.; Crotti, A. E. M.; Calefi, P. S.; Nassar, E. J.; Ciuffi, K.; Appl. Catal., A2010 , 389, 147.,2020 Mutin, P. H.; Vioux, A.; Chem. Mater. 2009 , 21, 582. The non-hydrolytic sol-gel route is a versatile way to prepare inorganic oxides1818 Nassar, E. J.; Avila, L. R.; Pereira, P. F. S.; de Lima, O. J.; Rocha, L. A.; Mello, C.; Ciuffi, K. J.; Carlos, L. D.; Quim. Nova2005 , 28, 238. during which non-hydrolytic condensation reactions furnish oxides and hybrid organic-inorganic materials; the oxo bonds originate from oxygen atoms of donors other than water.2020 Mutin, P. H.; Vioux, A.; Chem. Mater. 2009 , 21, 582. This route provides strict control of stoichiometry, powder morphology, and phase purity; cations are distributed all over the polymeric structure. Thermal treatment at high temperatures (from 400 ºC) releases organic matter and produces well-ordered crystallites, which is important when one wishes to obtain high crystallinity and controlled distribution of the constituents within the crystalline network. The non-hydrolytic sol-gel process dismisses the need for solvents and may reduce or eliminate the formation of residual metal-OH groups.2121 Sommerdijk, N.; Wright, J. D.; J. Sol-Gel Sci. Technol.1998 , 13, 565.

It is known that particle morphology and size affect the luminescent properties of Nb2O5 matrices; that is, the features of the Nb2O5 material depend on the methodology used to prepare the solid. Authors have employed countless methodologies to obtain Nb2O5 matrixes, like sol-gel,1010 Aquino, F. T.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Opt. Mater. (Amsterdam, Neth.)2013 , 35, 387.,1111 Aquino, F. T.; Pereira, R. R.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Mater. Chem. Phys. 2014 , 147, 751. Pechini,2222 Falcomer, D.; Speghini, A.; Ibba, G.; Enzo, S.; Cannas, C.; Musinu, A.; Bettinelli, M.; J. Nanomater. 2007 , 2007, DOI http://dx.doi.org/10.1155/2007/94975.
http://dx.doi.org/10.1155/2007/94975...
,2323 Graça, M. P. F.; Meireles, A.; Nico, C.; Valente, M. A.; J. Alloys Compd. 2013 , 553, 177. chemical transport,1414 Emmenegger, F. P.; Robinson, M. L. A.; J. Phys. Chem. Solids1968 , 29, 1673. hydrothermal route2424 Luo, H.; Wei, M.; Wei, K.; Mater. Chem. Phys.2010 , 120, 6.,2525 Wei, M.; Wei, K.; Ichihara, M.; Zhou, H.; Electrochem. Commun.2008 , 10, 980. and polymerized complexes.2626 Bueno, P. R.; Faria, R. C.; Bulhões, L. O. S.; Solid State Ionics2005 , 176, 1175. Nevertheless, to date there are no reports on the use of the non-hydrolytic sol-gel technique to prepare these materials.

Thus, the present study employed the non-hydrolytic sol-gel route to prepare Nb2O5:La3+,Eu3+ matrices at Nb2O5:La3+(1.0%x%)Eu3+(x%) stoichiometric ratios, where x = 0.1 and 0.3 mol%, to obtain red phosphors with a wide excitation range and adequate CIE chromaticity coordinates. The samples underwent thermal treatment at 550, 750, 900, and 1200 ºC, for 4 h, followed by investigation of their luminescent properties.

Experimental

All the solvents and reagents were of commercial grades (Merck and Sigma-Aldrich) unless otherwise stated. NbCl5 was donated by Companhia Brasileira de Metalurgia e Mineração (CBMM).

Preparation of Nb2O5:La3+(1%-x%)Eu3+(x%)

The preparation of the gel was carried out in oven-dried glassware. The material was synthesized via modification of the method described by Acosta et al.2727 Acosta, S.; Corriu, R. J. P.; Leclercq, D.; Lefèvre, P.; Mutin, P. H.; Vioux, A.; J. Non-Cryst. Solids1994 , 170, 234. Niobium pentachloride (NbCl5, 5.55 mmol) was previously suspended in 5 mL of dry dichloromethane and 50 mL of ethanol; it was then reacted in the presence of lanthanum chloride (LaCl3, 5.55 × 10-2 mmol) under reflux at 110 ºC, under argon atmosphere. For the samples doped with Eu3+ ions, an amount of europium chloride (EuCl3, 5.55 × 10 -3 or 1.66 × 10 -2 mmol) was added in the precursor solution. The condenser was placed in a thermostatic bath at −5 ºC. The gel was formed after 4 h of reaction; after 0.5 h, a solid material started to precipitate. After reflux, the mixture was cooled and aged overnight in the mother liquor, at room temperature. The solvent was removed under vacuum, and the solid was dried at 100 ºC. The powders were annealed at 550, 750, 900, and 1200 ºC for 4 h.

Characterization

X-ray diffraction (XRD) was performed at room temperature using a Rigaku Geigerflex D/max-c diffractometer with monochromated CuKα radiation (λ = 1.54 Å). Diffractograms were recorded in the 2θ range from 10 to 80º at a resolution of 0.05º. Raman spectra of the materials were obtained in the Raman spectrometer Horiba Jobin Yvon model LabRAM HR 800, operating with a He-Ne laser at 632.81 nm equipped with a charged couple device (CCD) camera model DU420A-OE-325 from the same manufacturer. The spectrum was obtained in the region from 100 to 4000 cm-1 with a hole of 100 µm and length of 5 to 30 cycles. Photoluminescence data were obtained at room temperature, under continuous Xe lamp (450 W) excitation in a Horiba Jobin Yvon Fluorolog-3 spectrofluorimeter equipped with an excitation and emission double monochromator and a photomultiplier R 928 Hammatsu. The emission was collected at 90º from the excitation beam. The slits were placed at 2.0 and 0.5 nm for excitation and emission, respectively; the bandpass was 0.5 nm, and the integration time was 0.5 ms. G1227 emission filters were employed (transmittance 100% for λ > 450 nm). The decay curves were acquired using a phosphorimeter accessory equipped with a Xe Lamp (5 J per pulse). Particle morphology was investigated by scanning electron microscopy (SEM) on a Shimadzu microscope model VEGA 3 SB.

Results and Discussion

Figure 1 shows the XRD patterns of the samples sintered at different temperatures. Samples without thermal treatment do not display any peaks. Addition of increasing, but small amounts of the dopant Eu3+ does not shift the diffraction peaks, because the unit cell varies within the detection limit of the instrument. Samples sintered at 550 ºC present peaks indexed to the (T) phase of Nb2O5 with an orthorhombic structure (JCPDS 30-873) and the Nb3O7Cl orthorhombic structure (JCPDS 73-295). Samples treated at 750 ºC contain Nb2O5 polycrystalline phases indexed to the Nb2O5 orthorhombic phase (JCPDS 30-873) and tetragonal structure (JCPDS 18-911). The crystalline (M) phase of the Nb2O5 monoclinic structure (JCPDS 15-166) and the La2Nb10O28 orthorhombic phase (JCPDS 20-547) appear after sintering at 900 ºC. After the thermal treatment at 1200 ºC, a totally crystalline (H) phase indexed to the Nb2O5 monoclinic structure (JCPDS 37-1468) and the La2Nb10O28 orthorhombic phase (JCPDS 20-547) were identified.

Figure 1
X-ray diffractograms of samples treated at different temperatures: (a) Nb2O5:La(1.0%); (b) Nb2O5:La(0.9%)Eu(0.1%); and (c) Nb2O5:La(0.7%)Eu(0.3%).

Figure 2 illustrates the Raman spectra of the samples after calcination at different temperatures. Untreated samples display broad bands in the region of the 60-400 and 760-1100 cm-1 related to the precursor (Nb-Cl) and to the initial linkages between niobium and oxygen.1111 Aquino, F. T.; Pereira, R. R.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Mater. Chem. Phys. 2014 , 147, 751.,2828 Konings, R. J. M.; Booij, A. S.; Vib. Spectrosc.1994 , 6, 345. Confirming the data observed in X-ray diffraction, the results obtained by Raman spectroscopy confirmed the polymorphism of niobia, which has many Nb-O polyhedra, being NbO6, octahedrons, the most common presenting in the T, M and H phases in different degree of distortion. In the T phase, broad bands are always present due to the large number of hexa or hepta coordination of Nb per unit cell, which increased the distortion and the number of different sites occupied by niobium atoms. The M and H phases are very similar and less distorted.1010 Aquino, F. T.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Opt. Mater. (Amsterdam, Neth.)2013 , 35, 387.,1111 Aquino, F. T.; Pereira, R. R.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Mater. Chem. Phys. 2014 , 147, 751.

Figure 2
Raman spectra of samples treated at different temperatures: (a) Nb2O5:La(1.0%); (b) Nb2O5:La(0.9%)Eu(0.1%); and (c) Nb2O5:La(0.7%)Eu(0.3%).

The samples calcined at 550 and 750 ºC presented bands at around 130, 230 and 700 cm-1 related to the optical vibration of Nb2O5, to the bending modes of the Nb−O−Nb linkage (ν4) and to the vibrational mode related to the Nb−O symmetric stretching (ν1). These vibrations are characteristic of the (T)-phase of Nb2O5 as observed in the X-ray results and they are in agreement to the literature. A broad band between 560 and 790 cm-1 is related to the vibrations of niobium polyhedrons in the (T)-phase of the Nb2O5. In addition, there is a weak shoulder around 890 cm-1, which corresponds to the symmetric stretching mode of surface Nb=O. For the samples annealed at higher temperature, an intense peak at 992 cm-1 was assigned to the Nb−O symmetric stretching (ν1) of the NbO6 polyhedra, characteristic of the highly-ordered (M)- or (H)-phase.1111 Aquino, F. T.; Pereira, R. R.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Mater. Chem. Phys. 2014 , 147, 751.,2929 McConnell, A. A.; Anderson, J. S.; Rao, C. N. R.; Spectrochim. Acta, Part A1976 , 32, 1067. There are low intensity bands at 895 and 847 cm-1, which correspond to the symmetric stretching mode of surface Nb=O and to the vibrational mode (ν1) of the NbO4 tetrahedron, once that one of the 28 Nb atoms in each unit cell is present in a tetrahedral site, which occurs at some block junctions. Several bands were also observed between 620 and 680 cm-1, which are related to the transverse optic modes in different types of octahedral arrangements.1111 Aquino, F. T.; Pereira, R. R.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Mater. Chem. Phys. 2014 , 147, 751.,1515 Nowak, I.; Ziolek, M.; Chem. Rev. 1999 , 99, 3603.,2929 McConnell, A. A.; Anderson, J. S.; Rao, C. N. R.; Spectrochim. Acta, Part A1976 , 32, 1067.

Low intensity bands at 350, 470 and 548 cm-1 assigned to the mode (ν5) are related to the block structures and the band at around 347 cm-1 is ascribed to the corner-shared octahedra. The octahedral distortion provides the appearance of broad multiplet bands at 260-290 cm-1 assigned to the infrared-active modes ν3 and ν4; the bands at 204 and 265 cm-1 are characteristic of ν6 modes and exist as a broad multiplet, also because of octahedral distortion.1111 Aquino, F. T.; Pereira, R. R.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Mater. Chem. Phys. 2014 , 147, 751.,2929 McConnell, A. A.; Anderson, J. S.; Rao, C. N. R.; Spectrochim. Acta, Part A1976 , 32, 1067. As observed for the (T)-phase, the samples calcined at 900 and 1200 ºC (M- and H-phase) also presented a band at around 120 cm-1 but in this case, it probably arises from a more ordered structure. Finally, bands at 102 and 200 cm-1 were assigned to the La-O vibrational mode,3030 Boldish, S. I.; White, W. B.; Spectrochim. Acta, Part A1979 , 35, 1235.,3131 Weckhuysen, B. M.; Phys. Chem. Chem. Phys. 2003 , 5, 4351. due to the lanthanum niobate phase also detected by the X-ray diffraction.

SEM images (Figure 3) reveal some interesting morphologies. Non-calcined samples and samples sintered at 550 ºC (Figures 4a1,2, 4b1,2 and 4c1,2) present porous surface; surface particles are irregular. Samples treated at 750 ºC (Figures 3a3, 3b3 and 3c3) correspond to regular pellets with typical crystalline grain measuring around 0.2-0.7 µm. Higher sintering temperature increases the crystalline grain size; indeed, the samples calcined at 900 and 1200 ºC (Figures 3a4,5, 3b4,5 and 3c4,5) have crystals with particle size around 0.7-5.0 µm; the pellets are arranged as sheets with a thickness estimated of around 60 nm (Figure 4). This is in agreement with the values obtained by the Scherrer analysis,3232 Burton, A. W.; Ong, K.; Rea, T.; Chan, I. Y.; Microporous Mesoporous Mater. 2009 , 117, 75. which provided crystallite size of around 50 nm to the diffraction plane h k l (3 0 1) at 2θ = 17.25º for the samples annealed at 1200 ºC.

Figure 3
Scanning electron micrographs of samples (a) Nb2O5:La(1.0%); (b) Nb2O5:La(0.9%)Eu(0.1%); and (c) Nb2O5:La(0.7%)Eu(0.3%). The indexes 1-5 indicate the thermal treatment temperature of 100, 500, 750, 900, and 1200 ºC, respectively.
Figure 4
Scanning electron micrographs of samples: (a) Nb2O5:La(1.0%); (b) Nb2O5:La(0.9%)Eu(0.1%); and (c) Nb2O5:La(0.7%)Eu(0.3%) treated at 1200 ºC.

Figures 5 and 6 depict the luminescence spectra of the samples Nb2O5:La(0.9%)Eu(0.1%) and Nb2O5:La(0.7%)Eu(0.3%), respectively, calcined at different temperatures. All the samples display a broad band at 275 nm in the excitation spectra (λem: 614 nm, Figures 5a and 6a), irrespective of the thermal treatment temperature. This band was more intense for the Nb2O5:La(0.7%)Eu(0.3%) calcined at 750 and 900 ºC (Figure 5a). Many literature papers describing niobium oxide doped with lanthanides have reported that the charge transfer (CT) band emerges in the excitation spectrum. Zhang et al.3333 Zhang, X.; Meng, F.; Wei, H.; Seo, H. J.; Ceram. Int. 2013 , 39, 4063. detected the CT band at 280 nm for the Li5(La2.7Eu0.3)Nb2O12 oxide obtained by solid-state reaction; the most intense band of the ƒƒ transition appeared at 463 nm and corresponded to the 7F05D2 transition. The (La0.99Eui0.01)NbO4:Eu3+ matrix prepared by spray pyrolysis method display a large CT band between 240 and 310 nm, ascribed to CT band.3434 Freiria, G. S.; Nassar, E. J.; Verelst, M.; Rocha, L. A.; J. Lumin.2015 , DOI dx.doi.org/10.1016/j.jlumin.2015.06.022.
dx.doi.org/10.1016/j.jlumin.2015.06.022...
The YNbO4-EuNbO4 system produced via hydrothermal route also presented the CT band between 240 and 270 nm.3535 Hirano, M.; Dozono, H.; Mater. Res. Bull. 2014 , 50, 213. On the other hand, the nanocrystalline powder Ln3+ doped Nb2O5 obtained by the Pechini methodology did not have a CT band.3636 Hsiao, Y. J.; Fang, T. H.; Chang, Y. S.; Chang, Y. H.; Liu, C. H.; Ji, L. W.; Jywe, W. Y.; J. Lumin. 2007 , 126, 866. In our work, we hypothesize that this band is related to the La2Nb10O28 micro-crystalline phase; which was assign to the charge CT band resulting from the 2p orbital of oxygen to empty 4f orbitals of Eu3+ (Eu3+ → O2-), and to charge transfer from the ligand oxygen to the empty 4f orbital of the central niobium atom of the group NbO43- (Nb5+ → O2-).55 Zhou, L.; Huang, J.; Gong, F.; Lan, Y.; Tong, Z.; Sun, J.; J. Alloys Compd. 2010 , 495, 268.,3434 Freiria, G. S.; Nassar, E. J.; Verelst, M.; Rocha, L. A.; J. Lumin.2015 , DOI dx.doi.org/10.1016/j.jlumin.2015.06.022.
dx.doi.org/10.1016/j.jlumin.2015.06.022...
,3737 Mahesh, S. K.; Rao, P. P.; Francis, T. L.; Reshmi, V. R.; Koshy, P.; Mater. Lett.2014 , 120, 115. This fact is in agreement with the results obtained by the X-ray diffraction, that reveled the presence of the La2Nb10O28 as a secondary phase in the niobia host. It is important to note that in several crystals, such as lithium niobate, strontium barium niobate, and barium sodium niobate, the location of the Ln3+ in the lattice is not obvious, as the trivalent lanthanide ions cannot easily replace the constitutional cations (Li+, Na+, Sr2+, Ba2+, Nb5+) due to clear size and/or charge mismatches. This location has been addressed by several studies where optical and/or structural techniques have been employed.2222 Falcomer, D.; Speghini, A.; Ibba, G.; Enzo, S.; Cannas, C.; Musinu, A.; Bettinelli, M.; J. Nanomater. 2007 , 2007, DOI http://dx.doi.org/10.1155/2007/94975.
http://dx.doi.org/10.1155/2007/94975...
Nevertheless, it is still a matter of debate whether the Ln3+ can substitute for the smaller and higher charged Nb5+ cation in a crystalline lattice. Falcomer et al.2222 Falcomer, D.; Speghini, A.; Ibba, G.; Enzo, S.; Cannas, C.; Musinu, A.; Bettinelli, M.; J. Nanomater. 2007 , 2007, DOI http://dx.doi.org/10.1155/2007/94975.
http://dx.doi.org/10.1155/2007/94975...
reported that pentavalent niobium ions can be substituted by trivalent lanthanide ions in crystalline niobates, but this substitution is accompanied by a strong disorder around the Ln3+ ions.

Figure 5
Luminescence spectra of the sample Nb2O5:La(0.9%)Eu(0.1%) calcined at different temperatures: (a) excitation spectra (λem: 614 nm); (b) emission spectra (λexc: 275 nm); (c) emission spectra (λexc: 394 nm); (d) emission spectra (λexc: 465 nm) and (e) emission spectra (λexc: 525 nm).
Figure 6
Luminescence spectra of the sample Nb2O5:La(0.7%)Eu(0.3%) calcined at different temperatures: (a) excitation spectra (λem: 614 nm), (b) emission spectra (λexc: 275 nm), (c) emission spectra (λexc: 394 nm), (d) emission spectra (λexc: 465 nm) and (e) emission spectra (λexc: 525 nm).

In the excitation spectrum, a transition is observed only if this level is efficient in populating the emitting level and thus in generating luminescence. If an energy level is absent it means that this level is not efficient in absorbing the excitation light and/or is not able in populating the emitting level.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1. Transitions within the 7F ground term are only observed for Eu3+ ions doped in inorganic matrices with low phonon energies, since these transitions are otherwise masked by the much stronger overtones and combination bands of the vibrations of the host matrix or ligands.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1.

The sharp lines in the 380-600 nm range correspond to Eu3+ intraconfigurational 4f-4f transitions in the host lattices; the main excitation bands at 394, 465, 525 and 532 nm refer to the Eu3+transitions: 7F05L6, 5D2, 5D1 and 7F15D1, respectively. As the 7F05D1 is a magnetic dipole transition, its intensity does not change. However, the increase of the annealing temperature promotes an increase of crystallinity (from orthorhombic to monoclinic) changing the local site symmetries for the Eu3+ ions and that can provide some differences in the intensities of the others electric dipole transitions. Higher annealing temperatures promote a decrease of the relative intensity of the 7F05L6 transition and favors the 7F15D1 transitions.

The emission spectrum for both samples show different profiles depending of excitation wavelengths and annealing temperatures. The samples excited at CT band (275 nm) exhibit broad bands for all intraconfigurational f-f transitions of Eu3+ ions, as presented in Figures 5b and 6b. Similar profiles were observed for the samples annealed at 750 ºC and 900 ºC, which were different in comparison with the samples annealed at 550 ºC and 1200 ºC. These three different behaviors are explained by the different crystalline phase. All samples calcined at 550 ºC present inhomogeneous broad bands for the 5D07FJ (J = 0, 1, 2, 3 and 4) due to the presence of many sites of symmetry, which are different to each other. As mentioned above, it is characteristic of (T) phase of niobia, that presents in its unit cell, eight of the Nb ions are present in distorted octahedra, while another eight Nb ions occupy pentagonal bipyramids. The remaining 0.8 Nb ion per unit cell is located in interstitial 9-coordinated sites in the unit cell.1515 Nowak, I.; Ziolek, M.; Chem. Rev. 1999 , 99, 3603. The samples annealed at 750, 900 and 1200 ºC presented a narrowing in the emission bands in comparison with the samples calcined at 550 ºC. For the samples calcined at higher temperatures, the (M) and (H) phase of the Nb2O5, have a shear structure consisting of blocks of NbO6 octahedra (3 × 4 and 3 × 5) that share corners with octahedra in their own block and edges with octahedra in other blocks. One of the 28 Nb atoms in each unit cell is present in a tetrahedral site, which occurs at some block junctions.1515 Nowak, I.; Ziolek, M.; Chem. Rev. 1999 , 99, 3603. In addition, as the 5D07F4 transition is also sensitive to the local symmetry, the structure changes can be observed by the variation of its intensity. The emission spectra (Figures 5a and 6a) excited at 275 nm for these samples present similar profiles to those already published in the literature,11 Huang, J.; Zhou, L.; Liang, Z.; Gong, F.; Han, J.; Wang, R.; J. Rare Earths2010 , 28, 356.,3333 Zhang, X.; Meng, F.; Wei, H.; Seo, H. J.; Ceram. Int. 2013 , 39, 4063.,3434 Freiria, G. S.; Nassar, E. J.; Verelst, M.; Rocha, L. A.; J. Lumin.2015 , DOI dx.doi.org/10.1016/j.jlumin.2015.06.022.
dx.doi.org/10.1016/j.jlumin.2015.06.022...
,3636 Hsiao, Y. J.; Fang, T. H.; Chang, Y. S.; Chang, Y. H.; Liu, C. H.; Ji, L. W.; Jywe, W. Y.; J. Lumin. 2007 , 126, 866. which describe the preparation of niobates matrices doped with the Eu3+ ions, suggesting that under CT band excitation, the majority emitting ions are located in the lanthanum niobate phase.

The presence of the 5D07F0 transition is an indication that the Eu3+ ion occupies a site with Cnv, Cn or Cs symmetry.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1.,3939 Rocha, L. A.; Avila, L. R.; Caetano, B. L.; Molina, E. F.; Sacco, H. C.; Ciuffi, K. J.; Calefi, P. S.; Nassar, E. J.; Mater. Res. 2005 , 8, 361. In addition, this transition is also useful to confirming the presence of non-equivalent sites in a host structure, because due to the non-degeneracy of the 7F0 and 5D0 levels, it is expected only one peak for a single site or species. The observation of more than one peak indicates that more than one site or species is present. Slight differences between two sites can provide small energy differences between the two different peaks in the 5D07F0 region, and the presence of more than one site for the Eu3+ ions is verified by the asymmetric shape of this transition or as a shoulder.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1. Although this transition is strictly forbidden according to the standard Judd-Ofelt theory, it occurs due to J-mixing. J is the total angular quantum number and it indicates the relative orientation of the spin and the orbital momenta. The J quantum numbers are well defined in the free Eu3+ ion, but J-mixing is due to the crystal-field perturbation which causes a mixing of the wave functions of terms with different J values.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1.

The emission spectra excited directly in the Eu3+ ions transitions (Figures 5c-e and 6c-e) prompts the appearance of bands characteristic of the transition from the excited state 5D0 to the fundamental level 7FJ (J = 0, 1, 2, 3, and 4). The emission profiles for the samples excited at 394, 465 and 525 nm presented great differences in comparison to those excited at 275 nm, suggesting the existence of different local sites symmetry for the Eu3+ ions. As the lanthanum ions are present in low concentration in comparison with the niobium ions, the majority of Eu3+ ions are most probably located in the niobia phase. For these samples, it was observed a narrowing of the bands in function of the thermal treatment temperature. The samples calcined at 900 and 1200 ºC presented two bands in the 5D07F0 transition region at 577.5 and 582.6 nm, evidencing that at least two active sites exist. This fact corroborates with the X-ray diffraction results, which show distinct phases for the calcined samples. An increase of the 5D07F0 transition’s intensity was observed for the samples calcined at higher temperatures due to an increase system’s cristallinity. Similar results were published for Sr2TiO4:Eu3+, Ln2O2SO4 (Ln = La, Gd, Y), Gd2O3 and Y2O3 materials.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1.,4040 Blasse, G.; Chem. Phys. Lett.1975 , 33, 616.

41 Antic-Fidancev, E.; Hölsä, J.; Lastusaari, M.; J. Alloys Compd. 2002 , 341, 82.
-4242 Porcher, P.; Svoronos, D. R.; Leskelä, M.; Hölsä, J.; J. Solid State Chem. 1983 , 46, 101. In these cases, the high intensity of the 5D07F0 transition was ascribed to ordered crystal structure, which leads to large linear terms in the crystal-field potential.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1.

The 5D07F1 transition directly reflects the crystal-field splitting of the 7F1 level. In a high symmetry such as, cubic or icosahedral crystal-fields, the 7F1 level is not split. In hexagonal or tetragonal crystal-fields, it is split into a non-degenerate and a twofold degenerate crystal-field level and in orthorhombic or lower symmetries, the total removal of crystal field degeneracies results in three sublevels for 7F1.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1. The samples presented an increase of splitting for the 5D07F1, indicating an increase of the local symmetry for the Eu3+ ions.

The peak at 582.5 nm could in principle be assigned to the 5D07F1 transition or in our case, to the 5D07F0 transition. According to Antic-Fidancev, a correlation exists between the position of the 5D0 level of Eu3+ and the position of the barycenter of the 7F1 manifold.4343 Antic-Fidancev, E.; J. Alloy Compd. 2000 , 300-301, 2. If the 5D0 level is at 17300 cm-1 (578.0 nm), the barycenter of the 7F1 manifold is expected at about 400 cm-1, which is not in agreement with our experimental values (17316 cm-1 and 240 cm-1, respectively).

The most intense line appears at 614 nm and corresponds to the forced Eu3+ 5D07F2 electric transition, indicating the absence of inversion symmetry in the sites occupied by Eu3+ ions.55 Zhou, L.; Huang, J.; Gong, F.; Lan, Y.; Tong, Z.; Sun, J.; J. Alloys Compd. 2010 , 495, 268. All samples presented a narrowing and an unfolding of the 5D07F2 transition with an increase of the annealed temperature. That fact is related to the changing of structure crystal as indicated by the results discussed heretofore. Only Nb2O5:La(0.7%)Eu(0.3%) treated at 1200 ºC and excited at 275 nm presented unfolding of this transition around 619 nm, probably as a consequence of higher occupation of different active sites by the emitting ion, due to its larger concentration in the matrix.

The ratio between the integrated intensities of these two transitions, I0-2/I0-1, functions as a probe of the local cation surroundings. The ratio values decrease as a function of the thermal treatment temperature for both Nb2O5:La(0.9%)Eu(0.1%) and Nb2O5:La(0.7%)Eu(0.3%). For an example, the values found for the Nb2O5:La(0.9%)Eu(0.1%) sample excited at 465 nm were 6.18, 4.39, 2.64 and 2.02 for calcination temperatures of 550, 750, 900 and 1200 ºC, respectively; and for the Nb2O5:La(0.7%)Eu(0.3%) sample excited at 465 nm, the values ratio were 5.63, 4.72, 2.46 and 2.08 for calcination temperatures of 550, 750, 900, and 1200 ºC, respectively. The ratio between the intensities of the 5D07F2 and 5D07F1 transitions varies as a function of the thermal treatment temperature, showing that Eu3+ exists in a distorted (or asymmetric) cationic environment in the samples calcined at 550 and 750 ºC. At higher treatment temperatures, the Eu3+ ions are preferentially located close to more crystalline and symmetric regions.

The 5D07F3 transition presented a similar behavior to the 5D07F0 transition, with an increase of intensity in function of the thermal treatment for all samples. In general the 5D07F3 transition is very weak, because it is forbidden according to the Judd-Ofelt theory, and as the 5D07F0 transition, the 5D07F3 transition can only gain intensity via J-mixing.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1. The increase of its intensity for the higher tempetarure annealed samples indicates an increase of J-mixing and crystal-field perturbation, caused by the structural changes.

Other indicative of structural changes can be observed by the variation of the 5D07F4 transition. The intensity of this transition is determined not only by symmetry factors, but also by the chemical composition of the host matrix.3838 Binnemans, K.; Coord. Chem. Rev. 2015 , 295, 1. As can be observed in the emission spectrum, the 5D07F4 transition presented different profiles with an increase of the annealing temperature and the europium concentration. Comparing the samples calcined at different temperatures, it was observed an increase of intensity, a narrowing and an higher unfolding for this transition, caused by the structural changes in the matrix. These structural changes allowed to obtain a deep-red emitting phosphor (ca. 700 nm). This fact is important because if a larger color gamut is required, e.g., for high-quality TV, monitors for design applications, etc., the primaries, i.e., the light transmitted by the blue, green, and red sections of the color filter, must be more separated from each other.4444 Jüstel, T.; Möller, S. ; Winkler, H. ; Adam, W. ; Ullmann's Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag GmbH: Weinheim, 2012. DOI 10.1002/14356007.a15_519.pub2.
https://doi.org/10.1002/14356007.a15_519...
In addition, slight changes on the intensity and split of this transition were also observed with an increase of europoium concentration, as showed in the emission spectrum for the samples annealed at 900 and 1200 ºC in Figures 5b-e and 6b-e.

Figure 7 brings the decay curves for the emission from the 5D0 states for both samples, treated at different temperatures. Table 1 lists the excited state lifetimes. The decay does not follow a simple exponential law, confirming the distribution of symmetry sites for the Eu3+ ion along the sample and providing a mean lifetime (τexp). In general, the Eu3+ ion lifetime increases in function of the thermal treatment temperature, which causes an increase on the structural crystallinity of the host as also confirmed by the X-rays diffraction and by the narrowing of the intraconfigurational f-f transtitions of the Eu3+ ions in the emission spectrum.

Figure 7
Decay curves for the Eu3+ ion in the matrixes (a) Nb2O5:La(0.9%) Eu(0.1%) and (b) Nb2O5:La(0.7%)Eu(0.3%) treated at different temperatures.
Table 1
Judd-Ofelt intensity parameters (Ω2, Ω4), probability of spontaneous Einstein emission (ARAD), radiative and experimental lifetimes (τRAD, τEXP), and quantum efficiency (q) of the calcined Nb2O5:La(0.9%)Eu(0.1%) and Nb2O5:La(0.7%)Eu(0.3%) samples

The emission spectra allowed us to calculate the spontaneous emission coefficient (ARAD), the radiative lifetime (τRAD), the quantum efficiency (q), and the Judd-Ofelt (Ω2, Ω4) parameters (Table 1). The 5D07F1 transition is purely magnetically dipolar, and its radiative rate does not depend on the local field imposed by the environment. The relation A0-λ = A0-1(S0-λ/S0-1)(σλ1), where S0-λ is the area under the curve related to the 5D07Fλ transition obtained from the spectral data, and σλ is the energy barycenter of the 0-λ transition, furnishes the A0-λ values. Here, we assumed that A0-1= 50 s-1, and that 5D07F5,6 have negligible intensities. Although the calculations of radiative rates were based in lanthanide complexes in solutions, they have been used for lanthanides doped in solid state matrices as a form to evaluate its luminescent properties.4545 Rocha, L. A.; Ciuffi, K. J.; Sacco, H. C.; Nassar, E. J.; Mater. Chem. Phys.2004 , 85, 245.

46 Yan, B.; Li, Q. P.; Microporous Mesoporous Mater. 2014 , 196, 284.

47 Shao, Y.-F.; Yan, B.; Microporous Mesoporous Mater. 2014 , 193, 85.
-4848 Gibelli, E.; Kai, J.; Teotonio, E.; Felinto, M.; Brito, H.; J. Photochem. Photobiol., A2013 , 251, 154. The details of these calculation were given by Werts et al.4949 Werts, M. H. V.; Jukes, R. T. F.; Verhoeven, J. W.; Phys. Chem. Chem. Phys. 2002 , 4, 1542. The mean lifetime was calculated using the equation <τ> = ∑αii/∑αi; α and τ represent the pre-exponential factor or amplitude and lifetime of the decay components, respectively.5050 Basu, B. B. J.; Vasantharajan, N.; J. Lumin. 2008 , 128, 1701.

All the samples have presented similar q values between 30 and 40%. Rising calcination temperatures reduce ARAD and Ω2. Reisfeld and Jørgensen5151 Reisfeld, R.; Jørgensen, C. K. In Handbook on the Physics and Chemistry of Rare Earth; Gschneidner, K. A., Eyring, L., eds.; North Holland: Amsterdam, 1984. consider that Ω2 is a useful parameter, because it is sensitive to the local symmetry of the ligand field and bond covalency. The value of Ω2 increases as the local symmetry of the ligand field decreases and the bond covalency increases. The Ω4 and Ω6 are related to the viscosity and rigidity of the host medium in which the ions are situated, but there is no theoretical prediction for this sensibility to macroscopic properties.5151 Reisfeld, R.; Jørgensen, C. K. In Handbook on the Physics and Chemistry of Rare Earth; Gschneidner, K. A., Eyring, L., eds.; North Holland: Amsterdam, 1984.,5252 Barve, R. A.; Suriyamurthy, N.; Panigrahi, B. S.; Venkatraman, B.; Phys. B (Amsterdam, Neth.)2015 , 475, 156.

The increase of the annealed temperatures promoted a decrease of the Ω2 values for the samples, indicating an increase of the local symmetry for the Eu3+ ions.5353 Santos, J. G.; Dutra, J. D. L.; Alves, S.; de Sá, G. F.; da Costa, N. B.; Freire, R. O.; J. Braz. Chem. Soc. 2013 , 24, 236. This fact corroborates with the results obtained by the X-ray diffraction that shows a higher crystallinity phases for the samples calcined at higher temperatures. In addition, these structural changes were also observed in the emission spectra of the Eu3+ ions in different annealed samples. However, the low Ω2 and Ω4 values obtained suggest that a weakly polarizable and rigid chemical environment surrounds Eu3+.5353 Santos, J. G.; Dutra, J. D. L.; Alves, S.; de Sá, G. F.; da Costa, N. B.; Freire, R. O.; J. Braz. Chem. Soc. 2013 , 24, 236.

The stronger transition at 614 nm favors saturated CIE chromaticity. We generated chromaticity coordinates for the Nb2O5:La(0.9%)Eu(0.1%) and Nb2O5:La(0.7%)Eu(0.3%) samples using the software Spectra Lux 2.05454 Santa-Cruz, P. A.; Teles, F. S.; Spectra Lux Software; Ponto Quantico Nanodispositivos: Recife, 2003. and the respective emission spectra recorded at room temperature (Tables 2 and 3). The CIE chromaticity coordinates of all the samples excited at 465 nm lie above the NTSC standard values (x = 0.670 and y = 0.330).55 Zhou, L.; Huang, J.; Gong, F.; Lan, Y.; Tong, Z.; Sun, J.; J. Alloys Compd. 2010 , 495, 268. Moreover, the full width at half maximum (FWHM) values diminish with increasing treatment temperature: Nb2O5:La(0.9%)Eu(0.1%) calcined at 1200 ºC affords the lowest value among the investigated samples, 3.5 nm.

Table 2
Chromaticity coordinates and FWHM of the Eu3+ 5D07F2 transition in calcined Nb2O5:La(0.9%)Eu(0.1%) and Nb2O5:La(0.7%)Eu(0.3%) samples excited at 275 and 394 nm
Table 3
Chromaticity coordinates, quantum efficiency (q) and FWHM of the Eu3+ 5D07F2 transition in calcined Nb2O5:La(0.9%)Eu(0.1%) and Nb2O5:La(0.7%) Eu(0.3%) samples excited at 465 and 525 nm

Conclusions

In this work, we prepared Nb2O5: La3+, Eu3+ phosphors by the non-hydrolytic sol-gel process. The SEM images revealed that particles resembled pellets arranged into sheets with thickness of ca. 60 nm and the samples annealed above 900 ºC elicited a totally crystalline monoclinic phase. On the contrary of Y2O2S:Eu3+ which efficiently absorbs light at around 370 nm5555 Fu, Z.; Geng, Y.; Chen, H.; Zhou, S.; Yang, H. K.; Jeong, J. H.; Opt. Mater. (Amsterdam, Neth.)2008 , 31, 58.,5656 Thirumalai, J.; Jagannathan, R.; Trivedi, D. C.; J. Lumin. 2007 , 126, 353. the luminescence results shown that all samples presented a wide excitation reange, including the light emission region of a light emitting diode emitting UV (275, 394 nm), the blue (465 nm) and green (525 nm) light. The CIE chromaticity coordinates were above the NTSC standard values. Hence, these two phosphors are promising red components for LED applications.

  • FAPESP has sponsored the publication of this article.

Acknowledgments

The authors would like to deeply express our most sincere thanks to Jeannette Dexpert-Ghys (CEMES) for her profitable assistance and cooperation. The agencies CNPq, and CAPES (Brazilian research funding agencies), grant 2011/51858-0 Brazil-France cooperation program FAPESP-CNRS; 2011/09823-4 and 2012/11673-3 E.J.N.; 2011/15757-4 L.A.R.; São Paulo Research Foundation (FAPESP). The authors thank the Laboratory of Photonic Materials (Prof S. J. L. Ribeiro) for the use of the RAMAN spectrometer and CBMM (Companhia Brasileira de Metalurgia e Mineração) for donating NbCl5.

References

  • 1
    Huang, J.; Zhou, L.; Liang, Z.; Gong, F.; Han, J.; Wang, R.; J. Rare Earths2010 , 28, 356.
  • 2
    Neeraj, S.; Kijima, N.; Cheetham, A. K.; Chem. Phys. Lett 2004 , 387, 2.
  • 3
    Thejo Kalyani, N.; Dhoble, S. J.; Renewable Sustainable Energy Rev 2012 , 16, 2696.
  • 4
    Wu, Z. C.; Shi, J. X.; Wang, J.; Gong, M. L.; Su, Q.; J. Solid State Chem 2006 , 179, 2356.
  • 5
    Zhou, L.; Huang, J.; Gong, F.; Lan, Y.; Tong, Z.; Sun, J.; J. Alloys Compd 2010 , 495, 268.
  • 6
    Ferrari, J. L.; Cebim, M. A.; Pires, A. M.; dos Santos, M. A. C.; Davolos, M. R.; J. Solid State Chem 2010 , 183, 2110.
  • 7
    Kodaira, C. A.; Stefani, R.; Maia, A. S.; Felinto, M. C. F. C.; Brito, H. F.; J. Lumin 2007 , 127, 616.
  • 8
    Pires, A. M.; Serra, O. A.; Davolos, M. R.; J. Alloys Compd. 2004, 374, 181.
  • 9
    Kutty, T. R. N.; Nag, A.; J. Mater. Chem 2003 , 13, 2271.
  • 10
    Aquino, F. T.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Opt. Mater. (Amsterdam, Neth.)2013 , 35, 387.
  • 11
    Aquino, F. T.; Pereira, R. R.; Ferrari, J. L.; Ribeiro, S. J. L.; Ferrier, A.; Goldner, P.; Gonçalves, R. R.; Mater. Chem. Phys 2014 , 147, 751.
  • 12
    Palatnikov, M.; Shcherbina, O.; Sidorov, N.; Skab, I.; Bormanis, K.; Ukr. J. Phys. Opt.2012 , 13, 207.
  • 13
    Rooksby, H. P.; White, E. A. D.; Acta Crystallogr 1963 , 16, 888.
  • 14
    Emmenegger, F. P.; Robinson, M. L. A.; J. Phys. Chem. Solids1968 , 29, 1673.
  • 15
    Nowak, I.; Ziolek, M.; Chem. Rev 1999 , 99, 3603.
  • 16
    Blanquart, T.; Niinistö, J.; Heikkilä, M.; Sajavaara, T.; Kukli, K.; Puukilainen, E.; Xu, C.; Hunks, W.; Ritala, M.; Leskelä, M.; Chem. Mater 2012 , 24, 975.
  • 17
    Blasse, G.; Grabmaier, B. C; Luminescent Materials; Springer: Berlin, 1994.
  • 18
    Nassar, E. J.; Avila, L. R.; Pereira, P. F. S.; de Lima, O. J.; Rocha, L. A.; Mello, C.; Ciuffi, K. J.; Carlos, L. D.; Quim. Nova2005 , 28, 238.
  • 19
    Ricci, G. P.; Rocha, Z. N.; Nakagaki, S.; Castro, K. A. D. F.; Crotti, A. E. M.; Calefi, P. S.; Nassar, E. J.; Ciuffi, K.; Appl. Catal., A2010 , 389, 147.
  • 20
    Mutin, P. H.; Vioux, A.; Chem. Mater 2009 , 21, 582.
  • 21
    Sommerdijk, N.; Wright, J. D.; J. Sol-Gel Sci. Technol.1998 , 13, 565.
  • 22
    Falcomer, D.; Speghini, A.; Ibba, G.; Enzo, S.; Cannas, C.; Musinu, A.; Bettinelli, M.; J. Nanomater 2007 , 2007, DOI http://dx.doi.org/10.1155/2007/94975.
    » http://dx.doi.org/10.1155/2007/94975
  • 23
    Graça, M. P. F.; Meireles, A.; Nico, C.; Valente, M. A.; J. Alloys Compd 2013 , 553, 177.
  • 24
    Luo, H.; Wei, M.; Wei, K.; Mater. Chem. Phys.2010 , 120, 6.
  • 25
    Wei, M.; Wei, K.; Ichihara, M.; Zhou, H.; Electrochem. Commun.2008 , 10, 980.
  • 26
    Bueno, P. R.; Faria, R. C.; Bulhões, L. O. S.; Solid State Ionics2005 , 176, 1175.
  • 27
    Acosta, S.; Corriu, R. J. P.; Leclercq, D.; Lefèvre, P.; Mutin, P. H.; Vioux, A.; J. Non-Cryst. Solids1994 , 170, 234.
  • 28
    Konings, R. J. M.; Booij, A. S.; Vib. Spectrosc.1994 , 6, 345.
  • 29
    McConnell, A. A.; Anderson, J. S.; Rao, C. N. R.; Spectrochim. Acta, Part A1976 , 32, 1067.
  • 30
    Boldish, S. I.; White, W. B.; Spectrochim. Acta, Part A1979 , 35, 1235.
  • 31
    Weckhuysen, B. M.; Phys. Chem. Chem. Phys 2003 , 5, 4351.
  • 32
    Burton, A. W.; Ong, K.; Rea, T.; Chan, I. Y.; Microporous Mesoporous Mater 2009 , 117, 75.
  • 33
    Zhang, X.; Meng, F.; Wei, H.; Seo, H. J.; Ceram. Int 2013 , 39, 4063.
  • 34
    Freiria, G. S.; Nassar, E. J.; Verelst, M.; Rocha, L. A.; J. Lumin.2015 , DOI dx.doi.org/10.1016/j.jlumin.2015.06.022.
    » dx.doi.org/10.1016/j.jlumin.2015.06.022
  • 35
    Hirano, M.; Dozono, H.; Mater. Res. Bull 2014 , 50, 213.
  • 36
    Hsiao, Y. J.; Fang, T. H.; Chang, Y. S.; Chang, Y. H.; Liu, C. H.; Ji, L. W.; Jywe, W. Y.; J. Lumin 2007 , 126, 866.
  • 37
    Mahesh, S. K.; Rao, P. P.; Francis, T. L.; Reshmi, V. R.; Koshy, P.; Mater. Lett.2014 , 120, 115.
  • 38
    Binnemans, K.; Coord. Chem. Rev 2015 , 295, 1.
  • 39
    Rocha, L. A.; Avila, L. R.; Caetano, B. L.; Molina, E. F.; Sacco, H. C.; Ciuffi, K. J.; Calefi, P. S.; Nassar, E. J.; Mater. Res 2005 , 8, 361.
  • 40
    Blasse, G.; Chem. Phys. Lett.1975 , 33, 616.
  • 41
    Antic-Fidancev, E.; Hölsä, J.; Lastusaari, M.; J. Alloys Compd 2002 , 341, 82.
  • 42
    Porcher, P.; Svoronos, D. R.; Leskelä, M.; Hölsä, J.; J. Solid State Chem 1983 , 46, 101.
  • 43
    Antic-Fidancev, E.; J. Alloy Compd 2000 , 300-301, 2.
  • 44
    Jüstel, T.; Möller, S. ; Winkler, H. ; Adam, W. ; Ullmann's Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag GmbH: Weinheim, 2012. DOI 10.1002/14356007.a15_519.pub2.
    » https://doi.org/10.1002/14356007.a15_519.pub2
  • 45
    Rocha, L. A.; Ciuffi, K. J.; Sacco, H. C.; Nassar, E. J.; Mater. Chem. Phys.2004 , 85, 245.
  • 46
    Yan, B.; Li, Q. P.; Microporous Mesoporous Mater 2014 , 196, 284.
  • 47
    Shao, Y.-F.; Yan, B.; Microporous Mesoporous Mater 2014 , 193, 85.
  • 48
    Gibelli, E.; Kai, J.; Teotonio, E.; Felinto, M.; Brito, H.; J. Photochem. Photobiol., A2013 , 251, 154.
  • 49
    Werts, M. H. V.; Jukes, R. T. F.; Verhoeven, J. W.; Phys. Chem. Chem. Phys 2002 , 4, 1542.
  • 50
    Basu, B. B. J.; Vasantharajan, N.; J. Lumin 2008 , 128, 1701.
  • 51
    Reisfeld, R.; Jørgensen, C. K. In Handbook on the Physics and Chemistry of Rare Earth; Gschneidner, K. A., Eyring, L., eds.; North Holland: Amsterdam, 1984.
  • 52
    Barve, R. A.; Suriyamurthy, N.; Panigrahi, B. S.; Venkatraman, B.; Phys. B (Amsterdam, Neth.)2015 , 475, 156.
  • 53
    Santos, J. G.; Dutra, J. D. L.; Alves, S.; de Sá, G. F.; da Costa, N. B.; Freire, R. O.; J. Braz. Chem. Soc 2013 , 24, 236.
  • 54
    Santa-Cruz, P. A.; Teles, F. S.; Spectra Lux Software; Ponto Quantico Nanodispositivos: Recife, 2003.
  • 55
    Fu, Z.; Geng, Y.; Chen, H.; Zhou, S.; Yang, H. K.; Jeong, J. H.; Opt. Mater. (Amsterdam, Neth.)2008 , 31, 58.
  • 56
    Thirumalai, J.; Jagannathan, R.; Trivedi, D. C.; J. Lumin 2007 , 126, 353.

Publication Dates

  • Publication in this collection
    Nov 2015

History

  • Received
    09 June 2015
  • Accepted
    18 Sept 2015
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br