Acessibilidade / Reportar erro

Low Temperature Synthesis of Luminescent RE2O3:Eu3+ Nanomaterials Using Trimellitic Acid Precursors

Abstract

[RE(TLA)·(H2O)n:Eu3+] (RE3+: Y, Gd and Lu; TLA: trimellitic acid) precursor complexes were synthesized by an one step aqueous co-precipitation method. After annealing for 1 h, RE2O3:Eu3+ nanophosphors were formed through the benzenetricarboxylate low temperature thermolysis method (500-1000 ºC). The compounds were characterized by using different techniques [elemental analysis (CHN), Fourier transform infrared spectroscopy (FTIR), thermogravimetry (TG/DTG), X-ray powder diffraction (XPD) and scanning electron microscope (SEM)]. The XPD data indicated that the Y2O3:Eu3+ materials have crystallite size range from 11 to 62 nm. The SEM and transmission electron microscopy (TEM) images show that the annealed materials keep morphological similarities with the precursor complexes. The photoluminescence properties were studied based on the excitation and emission spectra, and luminescence decay lifetimes of the 5D0 emitting level of the Eu3+ ion. The experimental intensity parameters (Ωλ), lifetimes (τ), as well as radiative (Arad) and non-radiative (Anrad) decay rates were calculated and discussed. The RE2O3:Eu3+ phosphors (RE: Y3+ and Lu3+) annealed at 500 to 1000 ºC have emission quantum efficiency (intrinsic quantum yield) values from 60 to 82%, indicating that this material can be potentially used for optical markers applications.

Keywords:
low temperature method; benzenetricarboxylate precursors; rare earth sesquioxides; photoluminescence materials


Introduction

Polycarboxylate ligands have a wide variety of structure providing large range of chemical properties when combined with metal ions. It has been drawing the attention in the areas such as metal framework systems (MOF),11 Mustafa, D.; Silva, I. G. N.; Bajpe, S. R.; Martens, J. A.; Kirschhock, C. E. A.; Breynaert, E.; Brito, H. F.; Dalton Trans.2014 , 43 , 13480.,22 Jiblaoui, A.; Leroy-Lhez, S.; Ouk, T.-S.; Grenier, K.; Sol, V.; Bioorg. Med. Chem. Lett. 2015, 25, 355. selective markers for medical applications,33 Choi, J. R.; Tachikawa, T.; Fujitsuka, M.; Majima, T.; Langmuir2010 , 26 , 10437. magnetic materials,44 Pan, Z.-R.; Xu, J.; Yao, X.-Q.; Li, Y.-Z.; Guo, Z.-J.; Zheng, H.-G.; CrystEngComm2011 , 13 , 1617. gas storage,55 Atwood, D. A.; The Rare Earth Elements: Fundamentals and Applications ; EIC Books; Wiley: Lexington, 2013. drug delivery,66 Avichezer, D.; Schechter, B.; Arnon, R.; React. Funct. Polym.1998 , 36 , 59. precursors for materials,77 Galvão, S. B.; Lima, A. C.; de Medeiros, S. N.; Soares, J. M.; Paskocimas, C. A.; Mater. Lett.2014 , 115 , 38. etc.

Rare earth (RE) containing materials show a versatility for application in the areas of science and technology specially in catalysis, permanent magnets in hybrid cars batteries,88 Sugimoto, S.; J. Phys. D: Appl. Phys.2011 , 44 , 064001.,99 Ma, H.; Okuda, J.; Macromolecules2005 , 38 , 2665. electroluminescent materials, persistent phosphors, structural probes, luminescent markers, display panels, etc.1010 Hong, Z. R.; Liang, C. J.; Li, R. G.; Li, W. L.; Zhao, D.; Fan, D.; Wang, D. Y.; Chu, B.; Zang, F. X.; Hong, L. S.; Lee, S. T.; Adv. Mater.2001 , 13 , 1241.

11 Trojan-Piegza, J.; Niittykoski, J.; Hölsä, J.; Zych, E.; Chem. Mater.2008 , 20 , 2252.

12 Gawryszewska, P.; Sokolnicki, J.; Legendziewicz, J.; Coord. Chem. Rev.2005 , 249 , 2489.

13 Cotton, S.; Spectrochim. Acta, Part A1990 , 46 , 1797.

14 Kumar, P.; Dwivedi, J.; Gupta, B. K.; J. Mater. Chem. C2014 , 2 , 10468.
-1515 Moine, B.; Bizarri, G.; Opt. Mater.2006 , 28 , 58. Most of those applications are consequence of their intrinsic characteristic: sharp intraconfigurational 4fN transitions, archiving high monochromatic emission colors and a wide range of emissions, from infrared to ultraviolet,1616 Brito, H. F.; Malta, O. L.; Felinto, M. C. F. C.; Teotonio, E. E. S. In The Chemistry of Metal Enolates, Part 1 ; John Wiley & Sons: West Sussex, England, 2009. e.g., Nd3+, Eu3+, Gd3+, Tb3+ and Tm3+ ions which emit in the infrared, red, ultraviolet, green and blue regions, respectively.

One very important feature of the RE3+ is their 4f-4f transitions, forbidden by the Laporte's rule. Associated to that, the shielding from the chemical environment by the filled 5s and 5p sub-shells1717 De Sá, G. F.; Malta, O. L.; Donegá, C. M.; Simas, A. M.; Longo, R. L.; Santa-Cruz, P. A.; da Silva, E. F.; Coord. Chem. Rev.2000 , 196 , 165. over the 4f electrons lead to a characteristic sharp lines spectra with small absorptivity and emission intensities. Taking into account the RE3+ intraconfigurational transitions, these ions can be divided in four groups depending on their spectroscopic features:

(i ) Sc3+(3d0), Y3+(4d0), La3+ (4f0) and Lu3+(4f14) where the 4f electrons are non-optically active due to their completely empty or fully occupied subshells;1616 Brito, H. F.; Malta, O. L.; Felinto, M. C. F. C.; Teotonio, E. E. S. In The Chemistry of Metal Enolates, Part 1 ; John Wiley & Sons: West Sussex, England, 2009.

(ii ) Gd3+(4f7) is a singular case due to its half-filled 4f layer, and therefore very stable. The energy difference between the lower emitting level (6P7/2) and the fundamental level (8S7/2) is approximately 32000 cm−1 opening the opportunity for its application as inorganic matrices. Due to the chemical similarity with other RE3+ ions, it is extensively used to study the emission of the ligands in coordination complexes;

(iii ) Sm3+(4f5), Eu3+(4f6), Tb3+ (4f8) and Dy3+(4f9): in these ions, the energy gap between the emitting and the lower levels are large enough to reduce the non-radiative decay process and accept energy from the ligands, interconfigurational transitions or charge transfer bands excited levels (Figure 1);

Figure 1
Partial energy diagram of trimellitic acid (TLA) ligand from [RE(TLA)·(H2O)n] (RE3+: Y, Gd and Lu) precursor (singlet and triplet states), Eu3+ ion and RE2O3:Eu3+ (LMCT) state.

(iv ) Ce3+(4f1), Pr3+(4f2), Nd3+ (4f3), Ho3+(4f10), Er3+(4f11), Tm3+(4f12) and Yb3+ (4f13): in these ions the energy gap between the emitting and lower levels are small, increasing the non-radiative decay process usually mediated by high energy vibrational modes in ligands (typically water molecules) or matrices (oxycarbonates, hydroxides, etc.). In these cases, the process accounts for the decreasing in the final emission efficiency.

To overcome the small absorptivity coefficients, luminescence sensitizers can be used to absorb and transfer the energy efficiently to the RE ions, keeping their desirable atomic characteristics. This phenomenon is a key feature in design of luminescent materials.1616 Brito, H. F.; Malta, O. L.; Felinto, M. C. F. C.; Teotonio, E. E. S. In The Chemistry of Metal Enolates, Part 1 ; John Wiley & Sons: West Sussex, England, 2009.,1818 Biggemann, D.; Mustafa, D.; Tessler, L. R.; Opt. Mater.2006 , 28 , 842.,1919 Souza, E. R.; Silva, I. G. N.; Teotonio, E. E. S.; Felinto, M. C. F. C.; Brito, H. F.; J. Lumin.2010 , 130 , 283.

In inorganic matrices such as vanadates, molybdates, tungstates and sesquioxides containing RE3+ ion, generally is observed an efficient energy transfer from the ligand metal charge transfer (LMCT) band to the metal ions. In the special case, the Eu3+ ion shows a high absorption intensity arising from the allowed LMCT transition, yielding a high intensity luminescence.2020 Kodaira, C. A.; Brito, H. F.; Felinto, M. C. F. C.; J. Solid State Chem.2003 , 171 , 401.

In solid state reactions, typically, is necessary high temperatures and long reaction time periods to prepare luminescent materials. This way to synthesize materials is known as ceramic method, which promotes heterogeneous distribution of the activator ion within the matrix and generate materials with high crystallite and particle sizes. Alternative methods to obtain materials in milder reaction conditions as: sol-gel, combustion or Pechini methods,2121 Huang, H.; Xu, G. Q.; Chin, W. S.; Gan, L. M.; Chew, C. H.; Nanotechnology2002 , 13 , 318.,2222 Aitasalo, T.; Dereń, P.; Hölsä, J.; Jungner, H.; Lastusaari, M.; Niittykoski, J.; Stręk, W.; Radiat. Meas.2004 , 38 , 515. are key to overcome the experimental limitation and improve their properties.

This report demonstrate the synthesis, characterization and optical properties of [RE(TLA):Eu3+(x mol%)] complexes (RE3+: Y, Gd and Lu; x: 0.1, 0.5, 1.0, and 5.0 mol%) and their low temperature annealing into the high luminescent RE2O3:Eu3+ phosphors. All the precursor complexes and resulting nanophosphors were characterized by elemental analysis (CHN), Fourier transform infrared (FTIR), thermogravimetry (TG), derivative thermogravimetry (DTG), X-ray powder diffraction (XPD) and scanning electron microscopy (SEM). The photoluminescence properties of the doped materials were studied based on the excitation and emission spectra and luminescence decay curves of the Eu3+ ion 5D0 excited level.

Experimental

High purity RE2O3 (RE3+: Y, Eu, Gd and Lu; CSTARM, 99.99%, China) were used to prepare the respective RECl3. (H2O)6 salts by reaction with concentrated HCl solution until total decomposition (ca. 60-80 ºC) of the solid and final pH close to 6. The trimellitic acid (TLA) (in the form of 1,2,4-benzenetricarboxylic acid 1,2-anhydride or 1,3-dihydro-1,3-dioxo-5-isobenzofurancarboxylic acid; Aldrich, 97%, Germany) was solubilized in water by drop-wise addition of 1 mol L−1 sodium hydroxide up to pH close to 6.

For the preparation of the [RE(TLA):Eu3+] complexes, 50 mL of RECl3(aq) (0.05 mol L−1) was slowly added to a 200 mL solution of Na3(TLA)(aq) (0.0125 mol L−1) at 1:1 molar ratio at ca. 100 ºC. The reaction mixture was refluxed for 1 h, the precipitate was filtered and washed four times with distilled water, dried and stored at reduced pressure.

The [RE(TLA):Eu3+] complexes obtained are non-hygroscopic, white crystalline powders, stable in air. The RE2O3:Eu3+ nanophosphors were obtained by annealing the [RE(TLA):Eu3+] complexes at 500, 600, 700, 800, 900 and 1000 ºC in a static air atmosphere, resulting in RE2O3:Eu3+ nanophosphors.

Elemental analyses were performed with a Perkin-Elmer CHN 2400 analyzer. The FTIR were acquired from 400 to 4000 cm−1 in KBr pallets form by using a Bomem MB100 FTIR. Thermogravimetry was performed from 30 to 900 ºC (heating ramp of 5 ºC min−1, synthetic air dynamic atmosphere) in a TA HI-RES TGA 2850 equipment. The XPD patterns were obtained in a Miniflex Rigaku II equipment (CuKα1) from 5 to 70º (2θ). The SEM micrographs were recorded in a JEOL JSM 7401F field emission scanning electron microscope. The transmission electron microscope (TEM) micrographs were recorded in a JEOL USA JEM-2100 LaB6 transmission electron microscope.

The luminescence study was based on the excitation and emission spectra recorded at room (300 K) and liquid nitrogen (77 K) temperatures. The measurements were performed in a SPEX-Fluorolog 3 instrument with double monochromators in front face mode (22.5º) using a 450 W Xenon lamp as excitation source. Luminescence decay curves were obtained by using a 150 W pulsed lamp and recorded in a SPEX 1934D phosphorimeter.

Results and Discussion

Characterization

A combination of elemental and thermogravimetric analysis (Table S1 and Figure 2) suggests an 1:1 molar ratio between the RE3+ ion and TLA ligand ([RE(TLA)·(H2O)n:Eu3+]; n: 4, 4 and 3 for Y3+, Gd3+ and Lu3+, respectively).2323 Silva, I. G. N.; Rodrigues, L. C. V.; Souza, E. R.; Kai, J.; Felinto, M. C. F. C.; Hölsä, J.; Brito, H. F.; Malta, O. L.; Opt. Mater.2015 , 40 , 41. The TG curves of coordination compounds show a water molecules mass-loss in the temperature interval between 50 and 230 ºC. Although the organic moiety decomposition of the complexes presents only one single-step between 450 and 570 ºC. In this case, it was used annealing temperature of 500 ºC during 1 h, in order to eliminate all the organic part leading to formation of the RE2O3:Eu3+ luminescent material.

Figure 2
TG and DTG curves (a) [RE(TLA)] and (b) [RE(TLA):Eu3+ (5.0 mol%)] (RE3+: Y, Gd and Lu).

The infrared absorption spectra (Figure S1) present similar spectral profile for the RE3+ complexes and Eu3+- doped matrices. The absorption bands between 1300 and 1600 cm–1 in the FTIR spectra of [RE(TLA)·(H2O)n:Eu3+] are assigned to the carboxylate symmetric νs(C=O) and asymmetric νas(C=O) stretching modes, respectively.1919 Souza, E. R.; Silva, I. G. N.; Teotonio, E. E. S.; Felinto, M. C. F. C.; Brito, H. F.; J. Lumin.2010 , 130 , 283.,2424 Nakamoto, K.; Infrared and Raman Spectra of Inorganic and Coordination Compounds ; John Wiley & Sons: New York, 1997.,2525 Silva, I. G. N.; Kai, J.; Felinto, M. C. F. C.; Brito, H. F.; Opt. Mater.2013 , 35 , 978. The narrow absorption peak around 3070 cm–1 is assigned to the C–H bond stretching of the [RE(TLA):Eu3+] complexes and the broad band between 3100-3700 cm–1 correspond to the O–H stretching from the water molecules.2626 Łyszczek, R.; J. Therm. Anal. Calorim.2007 , 90 , 533.

The sharp absorption bands around 510 and 580 cm−1 correspond to the characteristic RE3+−O stretching vibration. It is worth mentioning that the broad bands from 1250 to 1600 cm−1 are assigned to stretching mode of oxycarbonate remainder from the decomposition of the organic moistly of TLA and decreases with increasing annealing temperature (Figure S1), due to oxycarbonate decomposition.2323 Silva, I. G. N.; Rodrigues, L. C. V.; Souza, E. R.; Kai, J.; Felinto, M. C. F. C.; Hölsä, J.; Brito, H. F.; Malta, O. L.; Opt. Mater.2015 , 40 , 41. The broad absorption band located from 2800 to 3700 cm−1 is assigned to the superficial hydroxyl groups in the nanomaterials. Therefore, the RE2O3:Eu3+ materials originated from the [RE(TLA)] precursor complexes present similar chemical behavior compared to the sesquioxides prepared from the [RE(TMA)] complexes as reported by Silva et al.2323 Silva, I. G. N.; Rodrigues, L. C. V.; Souza, E. R.; Kai, J.; Felinto, M. C. F. C.; Hölsä, J.; Brito, H. F.; Malta, O. L.; Opt. Mater.2015 , 40 , 41.

The X-ray diffraction patterns of the [RE(TLA):Eu3+] complexes are similar to the powder diffraction patterns (PDF) for [Gd(TLA)]:Eu3+ and [Y(TLA)]:Eu3+] (00-056-1733) and [Lu(TLA)]:Eu3+] (00-058-1915), Y3+ and Gd3+ complexes are isomorphs. Consequently, there is no change in position or formation of new diffraction peaks at different concentrations of the dopants. This result is consistent with the Vegard's rule2727 Shannon, R. D.; Acta Crystallogr.1976 , 32 , 751.,2828 Vegard, L.; Z. Phys.1921 , 5 , 17. which suggests a formation of a solid solution between the Eu3+ dopant and the RE3+ in the host matrices due to the high similarity in the radii of these RE3+ ions.2727 Shannon, R. D.; Acta Crystallogr.1976 , 32 , 751.

The XPD patterns of the annealed materials at 500, 600, 700, 800, 900 and 1000 ºC (Figure 3) reveal a formation of RE2O3:Eu3+ in a cubic phase crystallization with the Ia3 space group.2929 Hölsä, J.; Turkki, T.; Thermochim. Acta1991 , 190 , 335. The absence of 2θ shift and reflections of impurities in the patterns of the RE2O3:Eu3+ indicates the formation of pure RE3+ sesquioxides. The XPD data of the Y2O3, Gd2O3 and Lu2O3 matrices (Figure 3) are very similar. Slight differences in the (222) reflection around 28º, moving to higher 2θ values with decreasing of the ionic radius of the RE3+ in the matrix, as predicted by Bragg's law.3030 Davolos, M. R.; Feliciano, S.; Pires, A. M.; Marques, R. F. C.; Jafelicci, M.; J. Solid State Chem.2003 , 171 , 268.

Figure 3
XPD patterns of (a) Y2O3:Eu3+; (b) Gd2O3:Eu3+ and (c) Lu2O3:Eu3+ (1.0 mol%) materials annealed for 1 h at different temperatures; reference pattern: PDF: 86-2477 and 86-2475, respectively.

The average crystal size of the doped materials was estimated from the powder diffraction data by using the Scherrer's formula (Figure 4).2323 Silva, I. G. N.; Rodrigues, L. C. V.; Souza, E. R.; Kai, J.; Felinto, M. C. F. C.; Hölsä, J.; Brito, H. F.; Malta, O. L.; Opt. Mater.2015 , 40 , 41.,3131 Klug, H. P.; Alexander, L. E.; X-ray Diffraction Procedures for Polycrystalline and Amorphous Materials ; Wiley: New York, 1975. The crystallite size of the RE2O3 materials increases as function of the RE3+ radius and annealing temperature. This behavior can be assigned to the higher reactivity of the Gd2O3, with lower melting point (2339 ºC) compared to Y2O3 (2410 ºC) and Lu2O3 (2427 ºC).3232 Adachi, G.; Imanaka, N.; Chem. Rev.1998 , 1479. Therefore, the sintering process is favored for the gadolinium matrix due to the dependence of the partial melting of the nanocrystals.2323 Silva, I. G. N.; Rodrigues, L. C. V.; Souza, E. R.; Kai, J.; Felinto, M. C. F. C.; Hölsä, J.; Brito, H. F.; Malta, O. L.; Opt. Mater.2015 , 40 , 41.

Figure 4
Correlation between the sesquioxides crystallite size and annealing temperature for RE2O3:Eu3+ (1.0 mol%) materials.

The narrowing of the diffraction peaks of RE2O3:Eu3+ (1.0 mol%) (RE3+: Y, Gd and Lu) phosphors presented in the XPD patterns (Figure 3) as function of the annealing temperature, indicates that the crystallite size increases from 11, 17, 18, 37, 46 and 62 nm as the annealing temperature increases from 500, 600, 700, 800, 900 and 1000 ºC (Y2O3), respectively (Figure 4). This behavior is related to the sintering of the nanocrystallites favored at high temperatures. Although the Gd2O3:Eu3+ annealed at 1000 ºC was also included in this work, the Scherrer's formula is recommended only for crystallite sizes up to 200 nm (Figure 4).

The SEM images of the [RE(TLA):Eu3+ (1.0 mol%)] precursors shows rods and a flower like morphologies (stacking of micrometric sheets of the material) for Y3+/Gd3+ (Figures 5a and 5b) and Lu3+ complexes, respectively (Figure 5c). After annealing up to 1000 ºC, the RE2O3:Eu3+ materials retained the original morphology of the correspond precursor complex (Figures 5d-5f). The nanosesquioxides exhibit higher porosity due to the decomposition of the organic moiety. This property is important for the design of nanomaterials with controlled morphology. Since it is possible to modify the complex morphologies, the desired nanoparticle shapes can be obtained by choosing the suitable synthetic method and reaction conditions.3333 Ren, H.; Liu, G.; Song, X.; Hong, G.; Cui, Z.; Proc. SPIE 6029, ICO20: Materials and Nanostructures2006 , 60291S.,3434 Wang, F.; Deng, K.; Wu, G.; Liao, H.; Liao, H.; Zhang, L.; Lan, S.; Zhang, J.; Song, X.; Wen, L.; J. Inorg. Organomet. Polym. Mater.2012 , 22 , 680.

Figure 5
SEM images of [RE(TLA):Eu3+ (1.0 mol%)] precursor complexes (a; b; c); RE2O3:Eu3+ (1.0 mol%) phosphor annealed during 1 h (d; e; f) and TEM images of RE2O3:Eu3+ (1.0 mol%) annelaed during 1 h at 1000 ºC (g; h).

The TEM micrographs (Figures 5g and 5h) show the cubic shape of the crystallites with high crystallinity. The particles retained the shape of the precursor agglomerates, shown in the SEM microscopy. At higher magnification no defects were observed in the crystals (except for the edges and crystallite contact points), suggesting the formation of a solid solution between the Eu3+ ions and the host matrices, compatible with the similar RE3+ ionic radii and chemical behavior of the Eu3+ and RE3+ matrices.

Photophysical properties of materials

[RE(TLA):Eu3+] precursor complexes

The excitation spectra of [RE(TLA):Eu3+ (x mol%)] (RE3+: Y, Gd and Lu) compounds were obtained by monitoring the hypersensitive transition 5D07F2 (619 nm) at 77 K (Figure 6). For all the complexes, the absorption bands are dominated by a high intensity broad TLA ligand band centered at 295 nm assigned to the S0→ S1 transition, indicating an efficient energy transfer TLA → Eu3+. The sharp peaks are assigned to the absorption of the Eu3+ ion originated from the ground state 7F0 to the 5L6 and 5D2 excited levels. The excitation spectra of the [RE(TLA):Eu3+ (x mol%)] (RE3+: Y and Gd) compounds show similar profiles suggesting that this system presents equivalent chemical environments around RE3+ ions and optical behaviors. On the other hand, [Lu(TLA):Eu3+ (x mol%)] shows slightly different spectral profile. For all [RE(TLA):Eu3+] systems, the 7F05L6 transition (25445 cm–1 for [Y(TLA):Eu3+ (5.0 mol%)]) exhibits the highest intensity among the intraconfigurational transitions in the excitation spectra.

Figure 6
The (a) excitation spectra of [RE(TLA):Eu3+ (5.0 mol%)] (RE3+: Y, Gd and Lu), with emission monitored at 616 nm; (b) emission spectra, with excitation at 295 nm, recorded at 77 K and (c) correlation between [RE(TLA):Eu3+ (x%)] lifetimes and Eu3+-doping concentrations (x: 0.1, 0.5, 1.0 and 5.0 mol%).

The emission spectra of the [RE(TLA):Eu3+ (x mol%)] complexes (RE3+: Y, Gd and Lu), were recorded under excitation in the TLA ligand band (ca. 295 nm) at 77 K, to reduce the vibronic coupling compared to the room temperature case. The emission energy levels of 5D07FJ transitions (J = 0-4) of the Eu3+ ion, can be attributed as the following (in cm–1): 7F0 (17270); 7F1 (16920); 7F2, (16210); 7F3 (15337) and 7F4 (14350), based at the [Y(TLA):Eu3+ 5.0 mol%)]. The efficient energy transference TLA → Eu3+ ion is evidenced by the absence of ligand broad emission band in the emission spectra in the spectral range from the 400 to 700 nm.

Using the optical data obtained from the emission spectra, it is possible to calculate the radiative rates (A0→J) from the 5D07FJ transitions using equation 1:1616 Brito, H. F.; Malta, O. L.; Felinto, M. C. F. C.; Teotonio, E. E. S. In The Chemistry of Metal Enolates, Part 1 ; John Wiley & Sons: West Sussex, England, 2009.,1717 De Sá, G. F.; Malta, O. L.; Donegá, C. M.; Simas, A. M.; Longo, R. L.; Santa-Cruz, P. A.; da Silva, E. F.; Coord. Chem. Rev.2000 , 196 , 165.

where σ0→1 and σ0→2,4 correspond to the energy barycenter of the 5D07F1 and 5D07F2,4 transitions, respectively. The S0→1 and S0→J are the areas calculated under the emission of the spectral curve corresponding to the 5D07F1 and 5D07FJ transitions, respectively.3535 Teotonio, E. E. S.; Fett, G. M.; Brito, H. F.; Faustino, W. M.; de Sá, G. F.; Felinto, M. C. F. C.; Santos, R. H. A.; J. Lumin.2008 , 128 , 190. Since the magnetic dipole 5D07F1 transition is almost insensitive to changes with the chemical environment around the Eu3+ ion, the A0→1 rate can be used as an internal standard to determine the A0→J coefficients for Eu3+ containing compounds.1616 Brito, H. F.; Malta, O. L.; Felinto, M. C. F. C.; Teotonio, E. E. S. In The Chemistry of Metal Enolates, Part 1 ; John Wiley & Sons: West Sussex, England, 2009.

The lifetime (τ) of the luminescent compounds were obtained from the luminescence decay curve using a first order exponential decay, with excitation at the 7F05L6 band. The emission quantum efficiency (η, or intrinsic quantum yield, 3636 Bünzli, J. C. G.; Coord. Chem. Rev.2015 , 293 , 19. of the 5D0 emitting level is determined according to equation 2:, as it has been defined by Bünzli)

where the total decay rate, Atot = 1/τ = Arad + Anrad and the Arad = ∑J A0→J. The Arad and Anrad quantities are the radiative and non-radiative rates, respectively. Table 1 shows the experimental values of the radiative (Arad), non-radiative (Anrad) rates and 5D0 emitting level emission quantum efficiency (η).

Table 1
Experimental values of intensity parameters (Ωλ), radiative (Arad) and non-radiative (Anrad) rates, emission lifetimes and emission quantum efficiencies of the 5D0 emitting level determined for the [RE(TLA):Eu3+ (x mol%)] (RE3+: Y, Gd and Lu) phosphors based on the emission spectra recorded at 77 K

The [RE(TLA):Eu3+ (x mol%)] lifetime values (Table 1 and Figure 6c) show higher values for Gd3+ and Y3+ containing complexes when compared to the Lu3+ ion case. On the other hand, there are no changes in the lifetime behavior doping with an increasing concentration from 0.1, 0.5, 1.0 and 5.0 mol%, within the same system.

The 5D07F2 and 5D07F4 transitions can be used to estimate the experimental intensity parameters (Ωλ, λ = 2 and 4). The Ω6 intensity parameter is not included in this study since the 5D07F6 transition was not observed for these systems. The coefficient of spontaneous emission, A, is given by equation 3:

where, χ = n (n + 2)2/9 is the Lorentz local field correction and n is the refractive index of the medium (refractive index used: 1.5 for all [RE(TLA):Eu3+] complexes and between 1.5 and 1.6 for RE2O3:Eu3+ materials). The squared reduced matrix elements 〈7FJ||U(λ)||5DJ2 are 0.0032 and 0.0023 calculated for J = 2 and 4, respectively.3535 Teotonio, E. E. S.; Fett, G. M.; Brito, H. F.; Faustino, W. M.; de Sá, G. F.; Felinto, M. C. F. C.; Santos, R. H. A.; J. Lumin.2008 , 128 , 190.,3737 Carlos, L. D.; Messaddeq, Y.; Brito, H. F.; Ferreira, R. A. S.; Bermudez, V. Z.; Ribeiro, S. J. L.; Adv. Mater.2000 , 12 , 594.

The Ωλ parameters depend mainly on the local geometry, bonding atoms and polarizabilities in the first coordination sphere of the RE3+ metal ion, and are governed by both forced electric dipole (FED) and dynamic coupling (DC) mechanisms. Moura et al.3838 Moura, R. T.; Carneiro Neto, A. N.; Longo, R. L.; Malta, O. L.; J. Lumin.2015 , in press , DOI: 10.1016/j.jlumin.2015.08.016.
https://doi.org/10.1016/j.jlumin.2015.08...
reported that the Ω2 parameter values are very sensitive to small angular changes in the local coordination geometry (much more than the Ω4,6 parameters). This spectroscopic behavior is associated with the hypersensitivity of certain 4f-4f transitions, to changes in the chemical environment, that are usually ruled by the Ω2 intensity parameter. On the other hand, the Ω4 and Ω6 values are most sensitive the chemical bond distances to the ligating atoms around the lanthanide ion. Indeed, as concluded by Moura et al. ,3838 Moura, R. T.; Carneiro Neto, A. N.; Longo, R. L.; Malta, O. L.; J. Lumin.2015 , in press , DOI: 10.1016/j.jlumin.2015.08.016.
https://doi.org/10.1016/j.jlumin.2015.08...
covalency in the ion-ligand bonding becomes more important with the increasing rank of the Ωλ, supporting the idea that the Ω4 and Ω6 parameters are better probes then Ω2 to quantify covalency in these compounds.

The Ωλ (λ = 2 and 4) parameter values for the [RE(TLA):Eu3+(x mol%)] compounds (x = 0.1, 0.5, 1.0 and 5.0 mol%) are presented in Table 1. The Ω2 values (ca. 6 × 10–20 cm2) found for these doped complexes are systematically larger than the [RE(TMA):Eu3+] (RE3+: Y and Lu) anhydrous complexes (ca. 2 × 10–20 cm2) values [RETMA] reported by Silva et al .3939 Silva, I. G. N.; Mustafa, D.; Andreoli, B.; Felinto, M. C. F. C.; Malta, O. L.; Brito, H. F.; J. Lumin.2015 , in press , DOI: 10.1016/j.jlumin.2015.04.047.
https://doi.org/10.1016/j.jlumin.2015.04...
reflecting the higher hypersensitive character of the 5D07F2 transition.2323 Silva, I. G. N.; Rodrigues, L. C. V.; Souza, E. R.; Kai, J.; Felinto, M. C. F. C.; Hölsä, J.; Brito, H. F.; Malta, O. L.; Opt. Mater.2015 , 40 , 41.,4040 Ferreira, R. A. S.; Nobre, S. S.; Granadeiro, C. M.; Nogueira, H. I. S.; Carlos, L. D.; Malta, O. L.; J. Lumin.2006 , 121 , 561.,4141 Silva, I. G. N.; Brito, H. F.; Souza, E. R.; Mustafa, D.; Felinto, M. C. F. C.; Carlos, L. D.; Malta, O. L.; Z. Z. Naturforsch., B: J. Chem. Sci.2013 , 69b , 231.

The emission quantum efficiency values of the [RE(TLA):Eu3+ (x mol%)] are lower for the complexes containing Y3+ and Gd3+ (η ca. 10%) and Lu3+ (η ca. 6%) ions, which indicate a strong non-radiative decay pathway mediated by water molecules (Table 1). It is also observed that increasing the Eu3+ concentration from 0.1 to 5.0 mol% produces no change in the emission quantum efficiency values, suggesting that the luminescence quenching concentration effect is not operative for these systems.

RE2O3:Eu3+ materials

The excitation spectra of RE2O3:Eu3+ annealed phosphors (RE3+: Y, Gd and Lu) were recorded at 77 K in the spectral range from 200 to 590 nm, with the emission monitored at 613 nm (Figure 7 and Figure S3). They show the presence of a broad absorption band centered around (ca. 39000 cm–1) assigned to the O2−(2p) → Eu3+(4f6) LMCT transition. Besides, the narrow absorption bands arisen from 4f-4f transitions from the RE3+ ion (ca. 17000 to 34000 cm–1) are observed.

The excitation spectra recorded at 300 K (Figure S4) show the presence of the overlapped 7F05D1 and 7F15D1 transitions (ca. 19000 cm–1) allowed by magnetic-dipole mechanism (∆J = 0, ±1, but 0 ↔ 0 is forbidden) for both the C2 and S6 symmetries. This optical results are due to the thermal population of the 7F1 level that are in agreement with the results previously reported for RE2O3:Eu3+.4242 Zych, E.; Karbowiak, M.; Domagala, K.; Hubert, S.; J. Alloys Compd.2002 , 341 , 381.,4343 Karbowiak, M.; Zych, E.; Holsa, J.; J. Phys.: Condens. Matter2003 , 15 , 2169. The absorption bands assigned to the 7F05D2 transition allowed by induced electric dipole and dynamic coupling mechanisms were observed from 21500 to 21900 cm–1. In addition, a weak absorption band around 24100 cm–1 is assigned to the forbidden 7F05D3 transition (by ∆J selection rules) as a result of the relaxation of the selection rule due to the J-mixing effects in the 7FJ manifolds. Moreover, the other absorption bands (Figure 7) originated from 4f-4f transitions of the Eu3+ ion were observed such as (in nm): the 7F05L6 (394), 5G2−6 (387), 5L7,8 (376), 5D4 (363), 5HJ', 5FJ', 5IJ' and 3P0 (between 286 and 335).

Figure 7
The (a) excitation spectra of Y2O3:Eu3+ (1.0 mol%), with emission monitored at 613 nm; (b) emission spectra, with excitation at 260 nm, recorded at 77 K.

It is worth mentioning that the excitation spectra of the Gd2O3:Eu3 present the characteristic strong absorption (nm): 8S7/26P7/2 (313), 8S7/26P5/2 (307) and 8S7/26P3/2 (302) transitions, indicating efficient energy transfer from the Gd3+ to the Eu3+ ion upper levels.4444 Buijs, M.; Meyerink, A.; Blasse, G.; J. Lumin.1987 , 37 , 9. The 8S7/26IJ(J = 7/2,9/2,17/2) (276) transitions overlap with the LMCT band. This high intensity absorption band indicates an efficient Gd3+ to Eu3+ energy transfer.4545 Macedo, A. G.; Ferreira, R. A. S.; Ananias, D.; Reis, M. S.; Amaral, V. S.; Carlos, L. D.; Rocha, J.; Adv. Funct. Mater.2010 , 20 , 624.

The luminescent materials prepared by the benzenetricarboxylate method present comparable excitation features, indicating the reproducibility of the method even when using different benzenetricarboxylate (BTC) ligands.2323 Silva, I. G. N.; Rodrigues, L. C. V.; Souza, E. R.; Kai, J.; Felinto, M. C. F. C.; Hölsä, J.; Brito, H. F.; Malta, O. L.; Opt. Mater.2015 , 40 , 41.

The emission spectra of the RE2O3:Eu3+ (RE3+: Y, Gd and Lu) annealed at temperatures from 500 to 1000 ºC were recorded at 77 K from 400 to 750 nm, under excitation in the LMCT band at 260 nm (Figure 7). All the spectra exhibit only the sharp lines arising from the 5D0,1,2,37F0-6, transitions of the Eu3+ ion. All materials show only one emission line assigned to 5D07F0 transition (ca. 17270 cm–1) of the C2 site of the cubic C-type. The 5D07F1 transition is present in both sites in the region of 16666, 16846 and 17015 cm–1 as well at 16770 and 17165 cm–1 originating from the C2 and S6 sites.4242 Zych, E.; Karbowiak, M.; Domagala, K.; Hubert, S.; J. Alloys Compd.2002 , 341 , 381.,4343 Karbowiak, M.; Zych, E.; Holsa, J.; J. Phys.: Condens. Matter2003 , 15 , 2169.

As reported by Boyer et al.46 and Meltzer et al. ,4747 Meltzer, R. S.; Feofilov, S. P.; Tissue, B.; Yuan, H. B.; Phys. Rev. B1999 , 60 , R14012. the refractive index (n) of the bulk RE2O3:Eu3+ is around 1.9 and the 5D0 lifetime (τ) of europium ion is 1.0 ms. On the other hand, these values can be different in the case of the RE2O3 nanostructured materials, with average sizes around 20-30 nm (crystallite size inferior to the wavelength of exciting radiation). Moreover, the morphology and surface/volume ratio of the nanoparticles may play a role in the profile of the decay curves.

The radiative rate (A01) of the 5D07F1 transition of Eu3+ ion (allowed by the magnetic dipole mechanism) is formally insensitive to the ligand field environment. Therefore it can be used as a reference transition whose value is 50 s–1 assuming a refractive index equal to 1.6.1717 De Sá, G. F.; Malta, O. L.; Donegá, C. M.; Simas, A. M.; Longo, R. L.; Santa-Cruz, P. A.; da Silva, E. F.; Coord. Chem. Rev.2000 , 196 , 165.,4848 Whiffen, R. M. K.; Antić, Ž.; Speghini, A.; Brik, M. G.; Bártová, B.; Bettinelli, M.; Dramićanin, M. D.; Opt. Mater.2014 , 36 , 1083.,4949 Santa-Cruz, P. A.; Teles, F. S.; Spectra Lux Software v.2.0 Beta , Ponto Quântico Nanodispositivos, RENAMI, 2003 . Based on this value, the refractive indices were determined and compared to the lifetime and crystallite size values reported previously.4646 Boyer, J. C.; Vetrone, F.; Capobianco, J. A.; Speghini, A.; Bettinelli, M.; J. Phys. Chem. B2004 , 108 , 20137.

47 Meltzer, R. S.; Feofilov, S. P.; Tissue, B.; Yuan, H. B.; Phys. Rev. B1999 , 60 , R14012.

48 Whiffen, R. M. K.; Antić, Ž.; Speghini, A.; Brik, M. G.; Bártová, B.; Bettinelli, M.; Dramićanin, M. D.; Opt. Mater.2014 , 36 , 1083.

49 Santa-Cruz, P. A.; Teles, F. S.; Spectra Lux Software v.2.0 Beta , Ponto Quântico Nanodispositivos, RENAMI, 2003 .

50 Binnemans, K.; Coord. Chem. Rev.2015 , 295 , 1.
-4846 Boyer, J. C.; Vetrone, F.; Capobianco, J. A.; Speghini, A.; Bettinelli, M.; J. Phys. Chem. B2004 , 108 , 20137. The experimental intensity parameters (Ω2,4) and lifetimes (0.8-1.9 ms) values were obtained using the effective refractive index values between 1.5 and 1.6. The values of the experimental intensity parameters (Ω2,4) the radiative (Arad) and non-radiative (Anrad) rates and emission quantum efficiencies (η) of the 5D0 emitting level of the RE2O3:Eu3+ are presented in Table 2.

Table 2
Experimental values of intensity parameters (Ωλ), radiative (Arad) and non-radiative (Anrad) rates, emission lifetimes and emission quantum efficiencies of the 5D0 emitting level determined for the RE2O3:Eu3+ (1.0 mol%) (RE3+: Y, Gd and Lu) phosphors, annealed for 1 hour, based on the emission spectra recorded at 77 K

The values for Ω2 (ca. 12) and Ω4 (ca. 2-3) are very similar in the same matrix (Table 2) for different annealing temperatures as shown in the spectral profiles (Figure 7b).5050 Binnemans, K.; Coord. Chem. Rev.2015 , 295 , 1. These results are a reflection of the observed emission intensity variations of the 5D07F2 transition of the Eu3+ ion. This optical behavior demonstrates that the Eu3+ ion acts as efficient luminescence probe even for the samples annealed at different temperatures. In addition, Ω2 and Ω4 values are also comparable changing the RE3+ matrix, due to the similarity in the radii in the lanthanide series.

The experimental intensity parameter values for the phosphors using the TLA ligand as precursor are smaller for all the systems, as compared to those originated from the TMA ligand, especially for the of Gd3+ matrix.1919 Souza, E. R.; Silva, I. G. N.; Teotonio, E. E. S.; Felinto, M. C. F. C.; Brito, H. F.; J. Lumin.2010 , 130 , 283.

According to Table 2, the RE2O3:Eu3+ phosphors present an emission quantum efficiency values varying from 37 to 82% with the annealing temperature of 500-1000 ºC. Among the materials, the Lu2O3:Eu3+(1.0 mol%) with annealing at 900 ºC present the highest emission quantum efficiency (η = 82%). This phenomenon is probably associated to the removal of oxycarbonate from the matrices with increasing the annealing temperature. It is important to mention that the RE2O3:Eu3+ phosphors prepared by the benzenetricarboxylate method using the TLA ligand is cheaper than compared with the TMA ligand.

The Commission Internationale de l'Eclairage (CIE) chromaticity coordinates generated from the emission spectra of Eu3+ doped RE2O3 (Figure 8) are x: 0.650 and y: 0.335.4949 Santa-Cruz, P. A.; Teles, F. S.; Spectra Lux Software v.2.0 Beta , Ponto Quântico Nanodispositivos, RENAMI, 2003 . The color coordinates show virtually no change for different sesquioxide matrices, concentration or annealing temperature. The phosphors containing Gd3+, Y3+, Lu3+ ions exhibit the same characteristic nearly monochromatic emission. The images of the Y2O3:Eu3+ (1.0 mol%) nanomaterials under UV irradiation show identical strong red emission for all the phosphors annealed at temperatures from 500 to 1000 ºC.

Figure 8
CIE diagram (center) and images and of [RE(TLA):Eu3+] (left) and RE2O3:Eu3+ (right), excitated at 254 nm.

Conclusions

[RE(TLA):Eu3+] complexes present low total decomposition temperature of the organic moistly producing the high luminescent RE2O3:Eu3+ materials at 500 ºC. The benzenetricarboxylate method is reliable, efficient and reproducible for the synthesis of phosphors at low temperature. The red emission of the RE2O3:Eu3+ materials (RE3+: Y3+, Gd3+ and Lu3+) arise mainly from the C2 symmetry site. The large values of the Ω2 experimental parameters corroborates with the high intensity of the 5D07F2 transition. Besides, these materials can act as efficient red light conversion devices in the studied Eu3+-concentration range. Finally, the RE2O3:Eu3+ phosphors prepared by the benzenetricarboxylate method using the [RE(TLA):Eu3+] present lower emission quantum efficiency (η close to 80%) than from the [RE(TMA):Eu3+] precursor complexes (η close to 90%). However they are cheaper, becoming an efficient and more economically viable system potentially usable as optical markers.

  • FAPESP has sponsored the publication of this article.
  • Supplementary Information
    Supplementary information is available free of charge at http://jbcs.sbq.org.br as PDF file.

Acknowledgments

The authors acknowledge financial support from Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) and Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP).

References

  • 1
    Mustafa, D.; Silva, I. G. N.; Bajpe, S. R.; Martens, J. A.; Kirschhock, C. E. A.; Breynaert, E.; Brito, H. F.; Dalton Trans.2014 , 43 , 13480.
  • 2
    Jiblaoui, A.; Leroy-Lhez, S.; Ouk, T.-S.; Grenier, K.; Sol, V.; Bioorg. Med. Chem. Lett. 2015, 25, 355.
  • 3
    Choi, J. R.; Tachikawa, T.; Fujitsuka, M.; Majima, T.; Langmuir2010 , 26 , 10437.
  • 4
    Pan, Z.-R.; Xu, J.; Yao, X.-Q.; Li, Y.-Z.; Guo, Z.-J.; Zheng, H.-G.; CrystEngComm2011 , 13 , 1617.
  • 5
    Atwood, D. A.; The Rare Earth Elements: Fundamentals and Applications ; EIC Books; Wiley: Lexington, 2013.
  • 6
    Avichezer, D.; Schechter, B.; Arnon, R.; React. Funct. Polym.1998 , 36 , 59.
  • 7
    Galvão, S. B.; Lima, A. C.; de Medeiros, S. N.; Soares, J. M.; Paskocimas, C. A.; Mater. Lett.2014 , 115 , 38.
  • 8
    Sugimoto, S.; J. Phys. D: Appl. Phys.2011 , 44 , 064001.
  • 9
    Ma, H.; Okuda, J.; Macromolecules2005 , 38 , 2665.
  • 10
    Hong, Z. R.; Liang, C. J.; Li, R. G.; Li, W. L.; Zhao, D.; Fan, D.; Wang, D. Y.; Chu, B.; Zang, F. X.; Hong, L. S.; Lee, S. T.; Adv. Mater.2001 , 13 , 1241.
  • 11
    Trojan-Piegza, J.; Niittykoski, J.; Hölsä, J.; Zych, E.; Chem. Mater.2008 , 20 , 2252.
  • 12
    Gawryszewska, P.; Sokolnicki, J.; Legendziewicz, J.; Coord. Chem. Rev.2005 , 249 , 2489.
  • 13
    Cotton, S.; Spectrochim. Acta, Part A1990 , 46 , 1797.
  • 14
    Kumar, P.; Dwivedi, J.; Gupta, B. K.; J. Mater. Chem. C2014 , 2 , 10468.
  • 15
    Moine, B.; Bizarri, G.; Opt. Mater.2006 , 28 , 58.
  • 16
    Brito, H. F.; Malta, O. L.; Felinto, M. C. F. C.; Teotonio, E. E. S. In The Chemistry of Metal Enolates, Part 1 ; John Wiley & Sons: West Sussex, England, 2009.
  • 17
    De Sá, G. F.; Malta, O. L.; Donegá, C. M.; Simas, A. M.; Longo, R. L.; Santa-Cruz, P. A.; da Silva, E. F.; Coord. Chem. Rev.2000 , 196 , 165.
  • 18
    Biggemann, D.; Mustafa, D.; Tessler, L. R.; Opt. Mater.2006 , 28 , 842.
  • 19
    Souza, E. R.; Silva, I. G. N.; Teotonio, E. E. S.; Felinto, M. C. F. C.; Brito, H. F.; J. Lumin.2010 , 130 , 283.
  • 20
    Kodaira, C. A.; Brito, H. F.; Felinto, M. C. F. C.; J. Solid State Chem.2003 , 171 , 401.
  • 21
    Huang, H.; Xu, G. Q.; Chin, W. S.; Gan, L. M.; Chew, C. H.; Nanotechnology2002 , 13 , 318.
  • 22
    Aitasalo, T.; Dereń, P.; Hölsä, J.; Jungner, H.; Lastusaari, M.; Niittykoski, J.; Stręk, W.; Radiat. Meas.2004 , 38 , 515.
  • 23
    Silva, I. G. N.; Rodrigues, L. C. V.; Souza, E. R.; Kai, J.; Felinto, M. C. F. C.; Hölsä, J.; Brito, H. F.; Malta, O. L.; Opt. Mater.2015 , 40 , 41.
  • 24
    Nakamoto, K.; Infrared and Raman Spectra of Inorganic and Coordination Compounds ; John Wiley & Sons: New York, 1997.
  • 25
    Silva, I. G. N.; Kai, J.; Felinto, M. C. F. C.; Brito, H. F.; Opt. Mater.2013 , 35 , 978.
  • 26
    Łyszczek, R.; J. Therm. Anal. Calorim.2007 , 90 , 533.
  • 27
    Shannon, R. D.; Acta Crystallogr.1976 , 32 , 751.
  • 28
    Vegard, L.; Z. Phys.1921 , 5 , 17.
  • 29
    Hölsä, J.; Turkki, T.; Thermochim. Acta1991 , 190 , 335.
  • 30
    Davolos, M. R.; Feliciano, S.; Pires, A. M.; Marques, R. F. C.; Jafelicci, M.; J. Solid State Chem.2003 , 171 , 268.
  • 31
    Klug, H. P.; Alexander, L. E.; X-ray Diffraction Procedures for Polycrystalline and Amorphous Materials ; Wiley: New York, 1975.
  • 32
    Adachi, G.; Imanaka, N.; Chem. Rev.1998 , 1479.
  • 33
    Ren, H.; Liu, G.; Song, X.; Hong, G.; Cui, Z.; Proc. SPIE 6029, ICO20: Materials and Nanostructures2006 , 60291S.
  • 34
    Wang, F.; Deng, K.; Wu, G.; Liao, H.; Liao, H.; Zhang, L.; Lan, S.; Zhang, J.; Song, X.; Wen, L.; J. Inorg. Organomet. Polym. Mater.2012 , 22 , 680.
  • 35
    Teotonio, E. E. S.; Fett, G. M.; Brito, H. F.; Faustino, W. M.; de Sá, G. F.; Felinto, M. C. F. C.; Santos, R. H. A.; J. Lumin.2008 , 128 , 190.
  • 36
    Bünzli, J. C. G.; Coord. Chem. Rev.2015 , 293 , 19.
  • 37
    Carlos, L. D.; Messaddeq, Y.; Brito, H. F.; Ferreira, R. A. S.; Bermudez, V. Z.; Ribeiro, S. J. L.; Adv. Mater.2000 , 12 , 594.
  • 38
    Moura, R. T.; Carneiro Neto, A. N.; Longo, R. L.; Malta, O. L.; J. Lumin.2015 , in press , DOI: 10.1016/j.jlumin.2015.08.016.
    » https://doi.org/10.1016/j.jlumin.2015.08.016
  • 39
    Silva, I. G. N.; Mustafa, D.; Andreoli, B.; Felinto, M. C. F. C.; Malta, O. L.; Brito, H. F.; J. Lumin.2015 , in press , DOI: 10.1016/j.jlumin.2015.04.047.
    » https://doi.org/10.1016/j.jlumin.2015.04.047
  • 40
    Ferreira, R. A. S.; Nobre, S. S.; Granadeiro, C. M.; Nogueira, H. I. S.; Carlos, L. D.; Malta, O. L.; J. Lumin.2006 , 121 , 561.
  • 41
    Silva, I. G. N.; Brito, H. F.; Souza, E. R.; Mustafa, D.; Felinto, M. C. F. C.; Carlos, L. D.; Malta, O. L.; Z. Z. Naturforsch., B: J. Chem. Sci.2013 , 69b , 231.
  • 42
    Zych, E.; Karbowiak, M.; Domagala, K.; Hubert, S.; J. Alloys Compd.2002 , 341 , 381.
  • 43
    Karbowiak, M.; Zych, E.; Holsa, J.; J. Phys.: Condens. Matter2003 , 15 , 2169.
  • 44
    Buijs, M.; Meyerink, A.; Blasse, G.; J. Lumin.1987 , 37 , 9.
  • 45
    Macedo, A. G.; Ferreira, R. A. S.; Ananias, D.; Reis, M. S.; Amaral, V. S.; Carlos, L. D.; Rocha, J.; Adv. Funct. Mater.2010 , 20 , 624.
  • 46
    Boyer, J. C.; Vetrone, F.; Capobianco, J. A.; Speghini, A.; Bettinelli, M.; J. Phys. Chem. B2004 , 108 , 20137.
  • 47
    Meltzer, R. S.; Feofilov, S. P.; Tissue, B.; Yuan, H. B.; Phys. Rev. B1999 , 60 , R14012.
  • 48
    Whiffen, R. M. K.; Antić, Ž.; Speghini, A.; Brik, M. G.; Bártová, B.; Bettinelli, M.; Dramićanin, M. D.; Opt. Mater.2014 , 36 , 1083.
  • 49
    Santa-Cruz, P. A.; Teles, F. S.; Spectra Lux Software v.2.0 Beta , Ponto Quântico Nanodispositivos, RENAMI, 2003 .
  • 50
    Binnemans, K.; Coord. Chem. Rev.2015 , 295 , 1.

Data availability

Publication Dates

  • Publication in this collection
    Dec 2015

History

  • Received
    26 Aug 2015
  • Accepted
    16 Nov 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