Acessibilidade / Reportar erro

New Europium(III)-TTA Complex Containing 2-Pyrrolidone as Coligand to Application as Luminescent Sensor: PbII and Ethanol in Gasoline

Abstract

This work describes the synthesis and application of an europium(III) complex, [Eu(TTA)3(2-pyr) (H2O)], containing 2-thenoyltrifuoroacetonate (TTA) and 2-pyrrolidone (2-pyr) as ligands, obtained as white solid and soluble in ethanol. In solution, the complex showed red emission, characteristic of the EuIII emission in the solid state (λexc = 375 nm and λem = 612 nm). The complex was tested with PbII, CdII, and HgII ions in water. The effect of metal ions was monitored and evaluated by hypersensitive transition 5D0 7F2 (612 nm). The results show that PbII suppresses the complex luminescence and present high sensitivity, according to the values for the Stern-Volmer constant (K SV = 2300 L mol−1), showing good linearity, i.e., determination coefficient (R 2) of 0.9913, and low limit of detection (LOD = 6.03 µM). Luminescence quenching increased with the increasing concentration of PbII. These results indicate that the synthesized complex is a potential luminescent sensor for detecting PbII in a simple and fast way, being applicable for routine environmental analysis. When applied as an ethanol sensor in gasoline, the complex hypersensitive transition intensity decreased with the increasing volume of ethanol in gasoline, reaching the values of R 2 = 0.9815 and LOD = 4.94% v/v.

Keywords:
europium(III); 2-pyrrolidone; lanthanides; luminescent sensor


Introduction

Due to their luminescent properties, lanthanide ions compounds arouse the interest of the industry in developing new technologies.11 Rostra, J. G.; Ferrer, F. J.; Espinós, J. P.; Elipe, A. R. G.; Yubero, F.; ACS Appl. Mater. Interfaces 2017, 9, 16313., 22 Du, P.; Bharat, L. K.; Yu, J. S.; J. Alloys Compd. 2015, 633, 37., 33 Ling, Q.; Song, Y.; Ding, S. J.; Zhu, C.; Chan, D. S. H.; Kwong, D. L.; Kang, E. T.; Neoh, K. G.; Adv. Mater. 2005, 17, 455., 44 Reichardt, C.; Schneider, K. R. A.; Sainuddin, T.; Wächtler, M.; McFarland, S. A.; Dietzek, B.; J. Phys. Chem. A 2017, 121, 5635., 55 Muller, G.; Dalton Trans. 2009, 44, 9692., 66 Jia, J.; Zhang, Y.; Zheng, M.; Shan, C.; Yan, H.; Wu, W.; Gao, X.; Cheng, B.; Liu, W.; Tang, Y.; Inorg. Chem. 2018, 57, 300., 77 Kumar, P.; Kanika; Singh, S.; Lahon, R.; Gundimeda, A.; Gupta, G.; Gupta, B. K.; J. Lumin. 2018, 196, 207., 88 Kuncewicz, J.; Dąbrowski, J. M.; Kyzioł, A.; Brindell, M.; Łabuz, P.; Mazuryk, O.; Macyk, W.; Stochel, G.; Coord. Chem. Re v. 2019, 398, 113012. Lanthanides are particularly interesting for the spectroscopic properties of their trivalent ions, resulting from the 4f-4f electron transition.99 Kariaka, N. S.; Smola, S. S.; Rusakova, N. V.; Dyakonenko, V. V.; Shishkina, S. V.; Sliva, T. Y.; Trush, V. A.; Amirkhanov, V. M.; J. Lumin. 2020, 223, 117187. Although prohibited by the Laporte rule, the Judd-Ofelt theory postulates that these transitions occur due to the influence of coordination geometrical symmetry, which in most cases has asymmetric structures.1010 Rosa, P. P. F.; Kitagawa, Y.; Hasegawa, Y.; Coord. Chem. Rev. 2020, 406, 213153. The coordination number of lanthanide compounds is usually greater than six,1111 Cotton, S. A.; Raithby, P. R.; Coord. Chem. Rev. 2017, 340, 220., 1212 You, L. X.; Hao, J. H.; Qi, D.; Xie, S. Y.; Wang, S. J.; Xiong, G.; Dragutan, I.; Dragutan, V.; Ding, F.; Sun, Y. G.; Inorg. Chim. Acta 2019, 497, 119075., 1313 Karraker, D. G.; J. Chem. Educ. 1970, 47, 424. presenting point groups with coordination number equal to seven, such as mono-capped octahedron (C3v) or a monocapped trigonal prism (C2v),1414 Yanagisawa, K.; Nakanishi, T.; Kitagawa, Y.; Seki, T.; Akama, T.; Kobayashi, M.; Taketsugu, T.; Ito, H.; Fushimi, K.; Hasegawa, Y.; Eur. J. Inorg. Chem. 2015, 28, 4769. equal to eight, such as square antiprism (D4d) and bi-capped trigonal prism (C2v),1515 Burdett, J. K.; Hoffmann, R.; Fay, R. C.; Inorg. Chem. 1978, 17, 2553. or even equal to nine, as occurs with tri-capped trigonal prism (C3h) or capped square antiprism (C4v).1616 Kuramochi, Y.; Nakagawa, T.; Yokoo, T.; Yuasa, J.; Kawai, T.; Hasegawa, Y.; Dalton Trans. 2012, 41, 6634., 1717 Bünzli, J. C. G.; Chauvin, A. S.; Vandevyver, C. D. B.; Bo, S.; Comby, S.; Ann. N. Y. Acad. Sci. 2008, 1130, 97.

As f orbitals are shielded by the 5s and 5p orbitals, 4f-4f transitions are little influenced by the ligand feld, generating narrow-band emissions similar to those of the free metal, which are always located in the same region.1818 Heffern, M. C.; Matosziuk, L. M.; Meade, T. J.; Chem. Rev. 2014, 114, 4496.

Lanthanides have a low molar absorption coefficient. As the direct excitation of lanthanide ion is not very efficient for emitting light, the use of ligands is necessary to perform the ligand to metal energy transfer process known as the antenna effect.1919 Braun, F.; Comba, P.; Grimm, L.; Herten, D. P.; Pokrandt, B.; Wadepohl, H.; Inorg. Chim. Acta 2018, 484, 464., 2020 Zhang, J. W.; Wang, C. R.; Liu, W. H.; Xu, S.; Liu, B. Q.; Inorg. Chim. Acta 2020, 508, 119648., 2121 Bai, C.; Fu, X. Y.; Hu, H. M.; He, S.; Wang, X.; Xue, G. L.; Inorg. Chim. Acta 2020, 506, 119550.

From this process, the energy absorbed by the ligand passes from low-energy singlet to the high-energy singlet (S0 → S1), possibly involving an intersystem crossing of the high energy singlet to the triple excited state of the ligand (S0 → T1). Thus, this process may imply energy transfer from S1 or T1 to each emitting level of the lanthanide, especially regarding the triplet state for having a relative longer lifetime.2222 Chen, W. T.; J. Solid State Chem. 2019, 284, 121160., 2323 Ning, Y.; Zhu, M.; Zhang, J. L.; Coord. Chem. Rev. 2019, 399, 213028., 2424 Abdelhamid, H. N.; Wilk-Kozubek, M.; El-Zohry, A. M.; Gómez, A. B.; Valiente, A.; Martín-Matute, B.; Mudring, A. V.; Zou, X.; Microporous Mesoporous Mater. 2019, 279, 400., 2525 Jia, J. H.; Li, Q. W.; Chen, Y. C.; Liu, J. L.; Tong, M. L.; Coord. Chem. Rev. 2019, 378, 365. Among the lanthanides, EuIII ions are capable of emitting red light, TbIII green light, SmIII orange light, and TmIII blue light in the visible region, whereas DyIII and the SmIII emit lights in the near-infrared region.2626 Maouche, R.; Belaid, S.; Benmerad, B.; Bouacida, S.; Freslon, S.; Daiguebonne, C.; Suffren, Y.; Calvez, G.; Bernot, K.; Roiland, C.; Pollès, L.; Guillou, O.; Inorg. Chim. Acta 2020, 501, 119309., 2727 Sun, Y.; Li, Q.; Wei, S.; Zhao, R.; Han, J.; Ping, G.; J. Lumin. 2020, 225, 117241., 2828 Stan, C. S.; Rosca, I.; Sutiman, D.; Secula, M. S.; J. Rare Earths 2012, 30, 401., 2929 Zhao, Y.; Wang, S.; Han, Y.; Zhang, J.; Liu, C.; Hu, X.; Zhang, Z.; Wang, L.; J. Lumin. 2020, 223, 117253., 3030 Yang, D.; Liao, L.; Zhang, Y.; Guo, Q.; Mei, L.; Liu, H.; J. Lumin. 2020, 224, 117285., 3131 Baklanova, Y. V.; Maksimova, L. G.; Lipina, O. A.; Tyutyunnik, A. P.; Zubkov, V. G.; J. Lumin. 2020, 224, 117315., 3232 Herrmann, A.; Friedrich, D.; Zscheckel, T.; Rüssel, C.; J. Lumin. 2019, 214, 116550.

In its excited state ( 5D0 ca. 17267 cm 1), the EuIII ion have lower energies than most organic ligand triplet states, thus being widely used.3333 Kaczmarek, M.; J. Lumin. 2020, 222, 117174. Likewise, β-diketones ligands are among the most used to obtain compounds with high light emission efficiency since their structures present chelating capacities and are favored by the π → π* transitions.3434 Binnemans, K.; Chem. Rev. 2009, 109, 4283., 3535 Sizov, V. S.; Komissar, D. A.; Metlina, D. A.; Aminev, D. F.; Ambrozevich, S. A.; Nefedov, S. E.; Varaksina, E. A.; Metlin, M. T.; Mislavskií, V. V.; Taydakova, I. V.; Spectrochim. Acta, Part A 2020, 225, 117503., 3636 Duan, Y. Y.; Wu, D. F.; Chen, H. H.; Wang, Y. J.; Li, L.; Gao, H. L.; Cu, J. Z.; Inorg. Chim. Acta 2020, 499, 119203., 3737 Berezhnytska, O. S.; Savchenko, I. O.; Ivakha, N. B.; Rogovtsov, O. O.; Trunova, O. K.; Rusakova, N. V.; J. Mol. Struct. 2020, 1201, 127160., 3838 Bhat, S. A.; Iftikhar, K.; Dyes Pigm. 2020, 179, 108383., 3939 Borges, A. S.; Fulgêncio, F.; Silva, J. G.; Santos, T. A. R.; Diniz, R.; Windmöller, D.; Magalhães, W. F.; Araujo, M. H.; J. Lumin. 2019, 205, 72., 4040 Borges, A. S.; Caliman, E. V.; Dutra, J. D. L.; Silva, J. G.; Araujo, M. H.; J. Lumin. 2016, 170, 654.

The literature reported many luminescent lanthanide compounds for various applications. Lucena et al.4141 Lucena, M. A. M.; Sá, G. F.; Rodrigues, M. O.; Alves, S.; Talhavini, M.; Weber, I. T.; Anal. Methods 2013, 5, 705. reported a ZnAl1.95Ln0.05O4 (where Ln = EuIII or TbIII) enabled the marking of ammunition and clearly identified luminescent gunshot residue (GSR). The identification can be performed visually by direct UV irradiation of any surface containing GSR particles. Destefani et al.4242 Destefani, C. A.; Motta, L. C.; Costa, R. A.; Macrino, C. J.; Bassane, J. F. P.; Filho, J. F. A.; Silva, E. M.; Greco, S. J.; Carneiro, M. T. W. D.; Endringer, D. C.; Romão, W.; Microchem. J. 2016, 124, 195. reported the evaluation of acute toxicity of the complex [Eu(PIC)3(NMK)3], which has as ligands picric acid (PIC) and N-methyl-caprolactam (NMK), applied as luminescent marker for the visual identification of GSR. Devi et al.4343 Devi, R.; Rajendran, M.; Singh, K.; Pala, R.; Vaidyanathan, S.; J. Mater. Chem. C 2021, 9, 6618. synthesized a series of smart luminescent EuIII complexes with five different ligands [Eu(DBM)3ligand] using functionalized phenanthro-imidazole derivatives as the neutral ligands and dibenzoylmethanate (DBM) as the anionic ligand and red light-emitting diodes (LEDs) were fabricated by integrating a near-ultraviolet (NUV). Wang et al.4444 Wang, Q.; Long, M. Y.; Lv, C. Y.; Xin, S. P.; Han, X. G.; Jiang, W.; Food Control 2020, 109, 106894. reported europium(III) chelated nanoparticles (EuNPs) conjugated to monoclonal antibodies specific to the O-specific polysaccharide fractions of Escherichia coli O157:H7. The EuNPs were used as fuorescent nanoparticle immunochromatographic strips, which enable rapid and quantitative detection in food samples. Zmojda et al.4545 Zmojda, J.; Kochanowicz, M.; Miluski, P.; Golonko, P.; Baranowska, A.; Ragi, T.; Dorosz, J.; Kuwik, M.; Pisarski, W.; Pisarska, J.; Szal, R.; Mach, G.; Starzyk, B.; Lesniak, M.; Sitarz, M.; Dorosz, D.; Materials 2020, 13, 2817. reported luminescent studies on germanate glasses doped with EuIII ions for photonic applications. Zhang et al.4646 Zhang, D.; Zhang, Y.; Wang, Z.; Zheng, Y.; Zheng, X.; Gao, L.; Wang, C.; Yang, C.; Tang, H.; Li, Y.; J. Lumin. 2021, 229, 117706. reported a biodegradable film enabling visible light excitation of hexanuclear EuIII complex [Eu6(TTA)18CTP-TPY] (TTA = 2-thenoyltrifluoroacetone; CTP-TPY = 2,2,4,4,6,6-hexakis(4-([2,2’:6’,2”-terpyridin]-4’-yl)phenoxy)-1,3,5,2-λ 5,4λ 5,6λ 5-triazatriphosphinine) for various applications. Li et al.4747 Li, R. F.; Zhang, T.; Liu, X. F.; Feng, X.; Inorg. Chem. Commun. 2016, 73, 170. synthesized a luminescent europium metal-organic framework [Eu(Hbptc)(H2O)3]n (Hbptc = biphenyl-2,3,3’,5’-tetracarboxylate) for selective sensing of pollutant small organic molecules. Zhou et al.4848 Zhou, J.; Li, B.; Qi, A.; Shi, Y.; Qi, J.; Xu, H.; Chen, L.; Sens. Actuators, B 2019, 305, 127462. reported a ZnSe quantum dot based ion imprinting technology for fuorescence detecting CdII and PbII ions on a 3D rotary paper based microfuidic chip. Sun et al.4949 Sun, Z.; Li, H.; Sun, G.; Guo, J.; Ma, Y.; Li, L.; Inorg. Chim. Acta 2018, 469, 51. reported a lanthanide metal-organic frameworks (Ln-MOFs) {[Ln(L)(ox)0.5(H2O)2]·H2O}n (Ln = Pr, Nd, Sm and Eu) constructed by 5-hydroxyisophthalic acid (H2L) and oxalate (ox) through solvothermal methods as luminescent sensor to acetone and CuII. Chen et al.5050 Chen, Z.; Yu, X.; Li, X.; Ye, Q.; Zhou, K.; Cai, Y.; Huang, L.; Li, Y.; Zeng, C.; Inorg. Chem. Commun. 2020, 112, 107744. synthesized a three lanthanide metal-organic frameworks (LnMOFs) [Ln2L3(H2O)2]n (LnIII = EuIII, TbIII and GdIII) adopting the ligand of cyclobutane-1,1-dicarboxylic acid (H2L). The study indicates that EuIII complex is a luminescence sensor for methanol. Zhang et al.5151 Zhang, F.; Ma, J.; Huang, S.; Li, Y.; Spectrochim. Acta, Part A 2020, 228, 117816. reported a lanthanide MOF (LnMOF) of [Tb(HIP)(H2O)5]·(H2O)·(HIP)1/2 (Tb-HIP, where HIP is 5-hydroxyisophthalate) for detecting picric acid and macrodantin. Zhai et al.5252 Zhai, B.; Li, Z. Y.; Wub, Z. L.; Cui, J. Z.; Inorg. Chem. Commun. 2016, 71, 23. reported an europium metal-organic framework (Eu-MOF), [Eu(L)(OAc)(DMA)]n (DMA = N,N-dimethylacetamide) as luminescent probe for detecting AlIII. Li et al.5353 Li, R.; Zhang, Y.; Liu, X.; Chang, X.; Feng, X.; Inorg. Chim. Acta 2020, 502, 119370. reported lanthanide polymers {[Ln2(bpda)3(H2O)3]·H2O}n (Ln = Tb or Dy) that have been synthesized with 2,2’-bipyridine-4,4’-dicarboxylic acid (H2bpda) as fluorescent sensor to HgII. Zheng et al.5454 Zheng, M.; Tan, H.; Xie, Z.; Zhang, L.; Jing, X.; Sun, Z.; ACS Appl. Mater. Interfaces 2013, 5, 1078. synthesized an europium based metal-organic framework (Eu-MOF), EuL3 (L = 4’-(4-carboxyphenyl)-2,2’:6′,2”-terpyridine), under hydrothermal conditions, and used it as a solid luminescence sensor for FeIII. Yan et al.5555 Yan, Q.; Chen, Z. H.; Xue, S. F.; Han, X. Y.; Lin, Z. Y.; Zhang, S.; Shi, G.; Zhang, M.; Sens. Actuators, B 2020, 268, 108. synthesized an oleic acid (OA)-capped β-NaYF4: 20%Yb, 0.5%Tm upconversion nanoparticles (UCNPs) [OA-UCNPs] for CuII sensing. Xu et al.5656 Xu, X. Y.; Yan, B.; Sens. Actuators, B 2016, 222, 347. reported a EuIII functionalized Zr-based metal-organic framework as fluorescent probe for CdII detection in aqueous environment. Lou et al.5757 Lou, T.; Chen, Z.; Wang, Y.; Chen, L.; ACS Appl. Mater. Interfaces 2011, 3, 1568. reported a “blue-to-red” colorimetric method for determination of HgII and AgI based on stabilization of gold nanoparticles (AuNPs) by redox formed metal coating in the presence of ascorbic acid (AA). Lian et al.5858 Lian, C.; Chen, Y.; Li, S.; Hao, M.; Gao, F.; Yang, L.; J. Alloys Compd. 2017, 702, 303. reported a {[Yb(TTTPC)·(H2O)2]·3Cl·NO3·0.5DMA·6H2O}n synthesized in conventional aqueous solutions with H3TTTPC ligands (H3TTTPC = 1,1’,1’’-(2,4,6-trimethylbenzene-1,3,5-triyl(methylene))-tris(pyridine-4-carboxylic acid), DMA = N,N-dimethylacetamide) for highly selective and sensitive luminescent sensor for PbII over mixed metal ions. Yang et al.5959 Yang, L.; Song, S.; Zhang, H.; Zhang, W.; Bu, Z.; Ren, T.; Synth. Met. 2011, 161, 2230. synthesized nanoporous coordination polymers {[La2(PDA)3(H2O)4]·H2O} (PDA = pyridine-2,6-dicarboxylate) and {[Pr2(PDA)3(H2O)3]·H2O} as selective luminescent probes of PbII, CaII and CdII ions. Fonseca et al.6060 Fonseca, R. R. F.; Gaspar, R. D. L.; Raimundo Jr., I. M.; Luz, P. P.; J. Rare Earths 2019, 37, 225. reported a terbium(III)-based metal-organic framework as a luminescent sensor to detect the adulteration of ethanol fuel with methanol.

The literature reported the great versatility of the use of lanthanides with 2-thenoyltrifuoroacetone (TTA) and others coligands. Xue et al.6161 Xue, B.; Zhang, Z.; Sund, Y.; Wang, J.; Jiang, H.; Du, M.; Chi, C.; Li, X.; Carbohydr. Polym. 2018, 186, 176. reported lanthanide complexes [Yb(fac)3(H2O)2, Yb(TTA)3(H2O)2, Nd(TTA)3(H2O)2] (fac = 1,1,1,5,5,5-hexafluoro-2,4-pentanedione) functionalized nanofibrillated cellulose (Ln-NFC) nanopapers with near-infrared (NIR) luminescence to bring a brilliant future for UV flters. Teotonio et al.6262 Teotonio, E. E. S.; Fett, G. M.; Brito, H. F.; Sá, W. M. F. G. F.; Felinto, M. C. F. C.; Santos, R. H. A.; J. Lumin. 2008, 128, 190. reported a [Eu(TTA)2(NO3)(TPPO)2] (TPPO = triphenylphosphine oxide) (bis-TTA complex) and [Eu(TTA)3(TPPO)2] (tris-TTA complex) where was evaluated the photoluminescent and triboluminescent behavior. Li et al.6363 Li, Y. J.; Yan, B.; Li, Y.; Microporous Mesoporous Mater. 2010, 131, 82. reported a novel organic-inorganic mesoporous luminescent hybrid containing LnIII (EuIII, TbIII) complexes covalently attached to the functionalized ordered mesoporous SBA-15 (Santa Barbara amorphous-15), which were designated as Ln(TTA-SBA-15)3bpy and Ln(DBM-SBA-15)3bpy (bpy = 2,2’-bipyridine), respectively, obtained by sol-gel process. Li et al.6464 Li, Y.; Jiao, J.; Yan, P.; Liu, L.; Wang, J.; Wang, Y.; Huang, L.; Liu, J.; Belfore, L. A.; Scr. Mater. 2018, 152, 1. synthesized nanocomplexes NaGdF4:Yb,Er@SiO2@Eu(TTA)3Phen (2-thenoyltrifluoroacetone, TTA), (1,10-phenanthroline monohydrate, Phen) showing the ligand-sensitized EuIII complexes attached to the outer surface of SiO2. The green (542 nm) and red (610 nm) light were excited by NIR and UV illumination, respectively. In this context, this work uses a EuIII compound with 2-thenoyltrifuoroacetonate (TTA) and 2-pyrrolidone (2-pyr) ligands as a luminescent sensor to identify PbII in water and ethanol in gasoline.

Experimental

Materials

Europium(III) oxide (Eu2O3) and the ligands 2-theoyltrifluoroacetonate (TTA = C8H4F3O2S) and 2-pyrrolidone (2-pyr = C4H7NO) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Eu2O3 was converted to its respective chloride using concentrated hydrochloric acid.3939 Borges, A. S.; Fulgêncio, F.; Silva, J. G.; Santos, T. A. R.; Diniz, R.; Windmöller, D.; Magalhães, W. F.; Araujo, M. H.; J. Lumin. 2019, 205, 72.

Characterization of complex

The europium(III) compound with TTA, [Eu(TTA)3(H2O)2], was synthesized according to an adapted literature procedure.6565 Charles, R. G.; Ohlmann, R. C.; J. Inorg. Nucl. Chem. 1965, 27, 255. The percentage of europium in the complex was determined by complexometric titration with 0.01 mol L−1 standard EDTA (ethylenediamine tetraacetic acid) solution using ortho-xylenol orange as indicator. Carbon, hydrogen, and nitrogen (CHN) elemental analysis was performed using the PerkinElmer 2400 series II Elemental Analysis Instrument (Waltham, Massachusetts, USA). Infrared spectra were obtained from the spectral region of 4000 to 400 cm−1 using a KBr tablet in transmittance mode and a PerkinElmer FTIR GX spectrometer (Waltham, Massachusetts, USA) at room temperature. Thermogravimetry analysis (TGA) was performed using a TG 60-H Shimadzu thermobalance (Kyoto, Japan) at a flow rate of 50 mL min-1 and a heating rate of 10 ºC min−1, under the temperature range of 25 to 800 ºC and ambient atmosphere. Solid-state Ultraviolet-Visible (UV-Vis) spectroscopy analysis was performed in the 200-800 nm region using a PerkinElmer spectrometer (Waltham, Massachusetts, USA). Photoluminescence spectra (excitation and emission) were obtained using a FLUOROLOG3 ISA/JobinYvon Horiba (Edison, New Jersey, USA) spectrofuorometer equipped with a Hamamatsu R928P photomultiplier, the SPEX 1934 D phosphorimeter, a 450 W Xe lamp, and a pulsed 150 W Xe-Hg lamp.

Synthesis of the [Eu(TTA)3(2-pyr)(H2O)] complex

[Eu(TTA)3(2-pyr)(H2O)] was synthesized according to an adapted literature procedure.4040 Borges, A. S.; Caliman, E. V.; Dutra, J. D. L.; Silva, J. G.; Araujo, M. H.; J. Lumin. 2016, 170, 654. First, 2-pyr (0.20 g, 2.35 mmol) was added dropwise to a methanol solution (30 mL) of [Eu(TTA)3(H2O)2] (1.20 g, 1.40 mmol) and left stirring overnight. The resulting white solid was washed with water to remove the excess 2-pyr ligand, dried, and stored at room temperature. The scheme represents the synthesis of the complex (Figure 1).

Figure 1
Synthetic route of [Eu(TTA)3(2-pyr)(H2O)].

Selectivity study of metals

Ions from heavy metals are major polluters, posing significant risks both to human health and the environment. Ions such as NiII, HgII, ZnII, CdII, PbII, CuII, and CrIV are not easily degraded when compared to conventional organic pollutants. Most of them are highly toxic and cancerous, raising concern due to their cumulative capacity in living organisms.6666 Zou, Y.; Wang, X.; Khan, A.; Wang, P.; Liu, Y.; Alsaedi, A.; Hayat, T.; Wang, X.; Environ. Sci. Technol. 2016, 50, 7290., 6767 Wu, D.; Chen, L.; Lee, W.; Ko, G.; Yin, J.; Yoon, J.; Coord. Chem. Rev. 2018, 354, 74. Considering that, we investigated the effect of metal ions on the luminescence of the [Eu(TTA)3(2-pyr)(H2O)] complex using three heavy metals, namely PbII, CdII, and HgII. For each metal, aqueous solutions were prepared from their respective nitrates in a 2.41 × 10−3 mol L−1 concentration. An ethanolic solution of the complex was also prepared in a 6.50 × 10−5 mol L−1 concentration. Then, 0.500 mL of each metal aqueous solution was individually mixed with 2.500 mL of the solution containing [Eu(TTA)3(2-pyr)(H2O)]. The mixtures were stirred and analyzed in the spectrofluorometer.

Detection of ethanol content in gasoline

Type ‘A’ gasoline (without additives) and ethanolic solution of the complex in the concentration of 6.50 × 10-5 mol L−1 were used in the experiment. Gasoline was mixed with ethanol in proportions ranging from 0-100% (v/v), stirred, and analyzed on the spectrofuorometer. This experiment sought to evaluate the behavior of the ethanol/complex when mixed with gasoline to verify whether it would be applicable as a luminescent sensor for identifying gasoline adulteration by photoluminescence.

Results and Discussion

Characterization of [Eu(TTA)3(2-pyr)(H2O)]

The results obtained from carbon, hydrogen, and nitrogen (CHN) elemental analysis and complexometric EDTA titration6868 Lyle, S. J.; Rahaman, M. M.; Talanta 1963, 10, 1177. suggest the formation of the stoichiometric compound [Eu(TTA)3(2-pyr)(H2O)].

C28H21EuF9NO8S3 calcd. (%): EuIII 16.49, C 35.98, H 2.24, N 1.45, measured: EuIII 16.54, C 36.61, H 2.30, N 1.52; FTIR ν / cm−1 (Figure S1, Supplementary Information (SI) section) νassC-O (1536(s)), νsC-O (1623(s)), νC-N (1496(s)); TTA free: νsC-O (1654(s)) and 2-pyr free: νsC-O (1662(s)), νC–N (1467(s)). Molar conductance in acetonitrile: 4.41 cm 2 Ω−1 mol−1. Yield 85%. Based on the molar conductance in acetonitrile, we may rightfully infer that the compound is nonelectrolyte, thus indicating that the three TTA are linked to the first coordination sphere of the complex, similarly to the [Eu(TTA)3(NMC)(H2O)] (NMC = N-methyl-6-caprolactam) studied by Borges et al.3939 Borges, A. S.; Fulgêncio, F.; Silva, J. G.; Santos, T. A. R.; Diniz, R.; Windmöller, D.; Magalhães, W. F.; Araujo, M. H.; J. Lumin. 2019, 205, 72., 4040 Borges, A. S.; Caliman, E. V.; Dutra, J. D. L.; Silva, J. G.; Araujo, M. H.; J. Lumin. 2016, 170, 654.

The Fourier transformed infrared spectroscopy (FTIR) (Figure S1) showed that the band associated with the symmetrical stretch, νsC–O in 1623 cm−1, suffered a shift to lower wavenumbers concerning the TTA (1654 cm−1) and 2-pyr (1643 cm−1), suggesting a coordination with EuIII ion. The [Eu(TTA)3(2-pyr)(H2O)] complex also presented the band associated with the asymmetric stretch, νassC–O (1536 cm−1), absent in the free ligand spectra, corroborating with ligands coordination. Both TTA and 2-pyr present νC=O, but they cannot be distinguished because they occur in the same region. The presence of the stretch vibration νC–N and its wavenumbers shift to larger values concerning free 2-pyrrolidone ligand suggests that this ligand is coordinated. Thus, the νsC–O displacement in TTA and 2-pyr suggest that these ligands are coordinated by carbonyl oxygen. Also, the spectrum of the complex presented the characteristic water νO–H vibrational mode in the 3000-3500 cm−1 region, suggesting the presence of water.6969 Ugale, A.; Kalyani, N. T.; Dhoble, S. J.; Mater. Sci. Energy Technol. 2020, 3, 51.

The thermogravimetry analysis (Figure S2, SI section) presented three main events of mass loss. From 25 to 76 ºC, the compound is very stable. The first mass loss occurred from 76 to 120 ºC, as an endothermal event. This first stage was related to the water output of coordination (experimental: 1.98%, calculated: 1.96%), corroborating the suggested stoichiometry. The following two steps are exothermic and occurred with peak temperatures at 289 and 456 °C, related to the thermal degradation of organic ligands, 2-tenoiltrifuoracetone and 2-pyrrolidone. The second weight loss (58.99%) occurred from 120 to 395 °C, which was attributed to the decomposition of two TTA and 2-pyr molecules. The third stage with about 20.91% mass loss started at 395 °C, finished at 540 °C, the mass loss percentage was near the loss of one TTA molecule from the complex. Complete degradation of the EuIII complex occurs at 540 ºC, with a total mass loss of 81.88% (residue mass: 18.12%). The total mass loss calculated was 80.99%, considering that the residue is Eu2O3 (calculated: 19.01%).

The absorption spectrum in the ultraviolet-visible (UV-Vis) region (Figure S3, SI section) showed bands with maximum absorption at 268 nm, attributable to π → π* intraligand electronic transitions, and an intense band in the 330 nm region, attributable to n → π* transitions. The excitation spectrum (Figure 2a) showed an intense broad band in the 250-450 nm assigned to S0 S1 and bands assigned to the 4f-4f transitions overlapped by the ligand band. The 4f-4f transitions are originated from the 7F0 ground state to the excited levels 5L J: 5L9 (360 nm), 5H4 (379 nm), 5L7 (385 nm), 5L6 (393 nm) and 5D3 (415 nm) excited states and remain approximately invariable in each EuIII complex. So, the 4f-4f transitions overlapped by the ligand band can be also an efficient channel for the photoluminescence for the compounds.7070 Oliveira, T. C.; Santos, H. P.; Lahoud, M. G.; Franco, D. F.; Freire, R. O.; Dutra, J. D. L.; Cuin, A.; Lima, J. F.; Marques, F. L.; J. Lumin. 2017, 181, 196. And it is possible to identify the bands 7F0 5D3 (451 nm), 7F0 5D2 (466 nm), 7F0 5D1 (526 nm), 7F0 5D0 (580 nm).4040 Borges, A. S.; Caliman, E. V.; Dutra, J. D. L.; Silva, J. G.; Araujo, M. H.; J. Lumin. 2016, 170, 654.

Figure 2
(a) Excitation with λem = 612 nm and (b) emission with λexe = 375 nm spectra of the complex in solid-state and (c) emission with λexe = 375 nm spectra of the complex (6 50 × 10−5 mol L−1) in ethanolic solution.

At room temperature, the emission spectrum (Figure 2b) of the compound in solid-state showed the regions of the intraconfigurational transitions, 5D0 7FJ (J = 0, 1, 2, 3, and 4), with maximum emission at 612 nm, 5D0 7F2 followed by 591 nm, 5D0 7F1, 652 nm, 5D0 7F3, and 700 nm, 5D0 7F4. The presence of the 5D0 7F2 transition, a hypersensitive electric dipole transition whose intensity is higher than that of 5D0 7F1 (allowed by a magnetic dipole transition), suggests that the compound does not have an inversion center and that EuIII is in chemical environment of low symmetry.7171 Chitnis, D.; Kalyani, N. T.; Dhoble, S. J.; Optik 2017, 130, 237. Experimental intensity parameters, Ωλ, for the [Eu(TTA)3(2-pyr)(H2O)] complex were determined from the emission spectra using the following equation 1.7272 Marques, L. F.; Santos, H. P.; Oliveira, K. A.; Botezine, N. P.; Freitas, M. C. R.; Freire, R. O.; Dutra, J. D. L.; Martins, J. S.; Legnani, C.; Quirino, W. G.; Machado, F. C.; Inorg. Chim. Acta 2016, 458, 28.

(1) Ω λ = 4 e 2 ω 3 A 0 λ 3 h c 3 χ 7 F J U ( λ ) 5 D 0 2

where e is the electron elementary charge, ω is the angular frequency, h is the Planck’s constant divided by 2π, c is the speed of light and χ is the Lorentz local field correction term, given by χ = n(n 2 + 2) 2/9 and 7FJ ║ U(λ) 5D0 2 is a squared reduced matrix element with value of 0.0032 for the 5D0 7F2 transition and 0.0023 for the 5D0 7F4. The refractive index (n) has been assumed equal to 1.5. In this work, the 5D0 7F6 transition was not observed experimentally, consequently, the experimental Ω6 parameter could not be estimated. The spontaneous emission coefficient, A01 = 0.31 × 10−11(n) 301) 3, leading to an estimated value around 50 s 1 for the refractive index (n) defined above.7272 Marques, L. F.; Santos, H. P.; Oliveira, K. A.; Botezine, N. P.; Freitas, M. C. R.; Freire, R. O.; Dutra, J. D. L.; Martins, J. S.; Legnani, C.; Quirino, W. G.; Machado, F. C.; Inorg. Chim. Acta 2016, 458, 28. In equation 2, the A0J term, where J = 2 and 4, represents the spontaneous emission coefficients of the 5D0 7F2 and 5D0 7F4 transitions, which can be calculated from 5D0 7F1 reference transition (magnetic dipole mechanism), therefore this transition is practically insensitive to chemical environment changing, equation 2.7373 Donegá, C. D.; Alves Jr., S.; Sá, G. F.; J. Alloys Compd. 1997, 250, 422.

(2) A 0 J = σ 0 λ S 0 λ S 01 σ 0 λ A 01

where S01 and S are the areas under the curves of the 5D0 7F1 and 5D0 7F J transitions, with σ01 and σ being their energy barycenters, respectively.7373 Donegá, C. D.; Alves Jr., S.; Sá, G. F.; J. Alloys Compd. 1997, 250, 422.

The lifetimes (τ) of the [Eu(TTA)3(2-pyr)(H2O)] complex were obtained from the photoluminescence decay curves (Figure S4, SI section) using the equation I(t) = I(0) exp(−t/τ) and a curve fitting program, where I(t) is the intensity at time t, t is time and I(0) is the intensity in the initial time equal to zero (t = 0). The quantum efficiency of the 5D0 emitting level (η) for the EuIII complex are presented in Table 1 and were calculated from the ratio Arad/Atotal, where Arad and Atotal are radiative and total rates assigned to the decay processes of the 5D0 emitting level, respectively. In this case, the Arad values were obtained by summing the radiative spontaneous coefficients due to the 5D0 7FJ transitions, while Atotal were determined from the lifetime values of the 5D0 emitting level (τ) by considering the reciprocal relation between these properties (Atotal = 1/τ).7474 Marques, L. F.; Cuin, A.; Carvalho, G. S. G.; Santos, M. V.; Ribeiro, S. J. L.; Machado, F. C.; Inorg. Chim. Acta 2016, 441, 67.

Table 1
Experimental values of intensity parameters (Ωλ), radiative (Arad), nonradiative (Anrad) rates, lifetimes (τ), emission quantum efficiencies (η) of the 5D0 emitting level determined for complexes [Eu(TTA)3(2-pyr)(H2O)] and compared to literature

The Table 1 shows the experimental values, obtained from the emission spectra, for the quantum efficiency (η), radiative (Arad) and nonradiative (Anrad) rates of spontaneous emission, lifetimes (τ) and the Judd-Ofelt intensity parameters (Ω2 and Ω4), for the EuIII ion in the complex.

The experimental data reveal that the substitution of the water molecule by a lactam increases the quantum efficiency of the europium complexes as noted by Borges et al.4040 Borges, A. S.; Caliman, E. V.; Dutra, J. D. L.; Silva, J. G.; Araujo, M. H.; J. Lumin. 2016, 170, 654. The value of η determined for this complex (η = 44%) is not larger owing to the presence of one water molecule in the first coordinated sphere of the EuIII ion. This luminescence quenching is an effect of the vibronic coupling of the higher-energy OH oscillators and 5D0 emitting level causing a non-radiative dissipation of the energy. When compared with another complex reported in the literature [Eu(TTA)3(NMC)(H2O)], changes occur in the radiative coefficient (Arad), the values of η are mainly associated to changes in the non-radiative rates (Anrad). The experimental intensity parameters Ω2 and Ω4 of [Eu(TTA)3(2-pyr)(H2O)] present different values from those of [Eu(TTA)3(NMC)(H2O)] synthesized by Borges et al.,4040 Borges, A. S.; Caliman, E. V.; Dutra, J. D. L.; Silva, J. G.; Araujo, M. H.; J. Lumin. 2016, 170, 654. indicating that EuIII ions are in different chemical environments (Table 1). According to the literature, the Ω2 value is the most infuenced by small angular changes in the local geometry. This effect, together with changes in the ligating atom polarizability (α), has been used to rationalize the hypersensitivity of certain 4f-4f transitions to changes in the chemical environment.4040 Borges, A. S.; Caliman, E. V.; Dutra, J. D. L.; Silva, J. G.; Araujo, M. H.; J. Lumin. 2016, 170, 654.

Selectivity study of metals

The emission spectrum of the mixture containing 2.50 mL of the complex in ethanolic solution and 0.500 mL of pure water (Figure 2c) showed the intraconfigurational transitions 5D0 7F1, 5D0 7F2, 5D0 7F3, and 5D0 7F4. We used the hypersensitive transition 5D0 7F2 (612 nm) as a reference to monitor possible metal-induced disturbances in the luminescence of the complex. As for the emission spectrum of the complex (Figure 3a), we verified that PbII decreased intensity by 12% at the reference transition (Figure 3b). Thus, the metal-induced luminescence quenching followed the order: PbII > CdII > HgII. The effect of PbII in the presence of other metals was evaluated by mixtures of 0.500 mL of each metal at the concentration of 2.41 × 10−3 mol L−1, it was kept the same final volumes and equimolar concentrations of the ions to avoid the effect of dilution on luminescence.

Figure 3
(a) Emission spectra (λex = 375 nm) and (b) fluorescence intensity of the transition 5D0 7F2 (612 nm) of [Eu(TTA)3(2-pyr)(H2O)] in ethanol solution with PbII, CdII and HgII in aqueous solutions.

Figure 3a shows that the luminescence intensity of the [Eu(TTA)3(2-pyr)(H2O)] solution decreases with all elements studied (PbII, HgII, and CdII). In the presence of PbII, the intensity of transition of HgII and CdII significantly decreases when compared to the elements alone.

We evaluated the change in luminescence intensity at 612 nm in the presence of PbII at different concentrations (1.45 × 10−4, 2.42 × 10−4, 4.83 × 10−4, 7.25 × 10−4, 9.66 × 10−4, 1.93 × 10−3 and 2.41 × 10−3 mol L−1) (Figure 4a) and found luminescence quenching to significantly increase with the increase in PbII concentration (Figure 4). The greater quenching caused by PbII may be explained by the energy absorption competition at the same wavelength of the ligand or by the complex-PbII interaction. The UV-Vis absorption spectrum of PbII in aqueous solution (Figure S5, SI section) showed an overlap with the complex excitation spectrum (Figure 2a), evincing the absorption competition between PbII and the europium(III) complex. Moreover, PbII ion electronic state has an empty p orbital, making it prone to become an electron-acceptor and thus weakening fluorescence.7575 Yan, Y. T.; Zhang, W. Y.; Zhang, F.; Cao, F.; Yang, R. F.; Wang, Y. Y.; Hou, L.; Dalton Trans. 2018, 47, 1682., 7676 Song, X. Z.; Wang, Y. X.; Yan, J. W.; Chen, X.; Meng, Y. L.; Tan, Z.; ChemistrySelect 2018, 3, 9564.

Figure 4
(a) Emission spectra (λex = 375 nm) monitoring the intensity of the 5D07F2 transition (at 612 nm) of [Eu(TTA)3(2-pyr)(H2O)] in aqueous solution with PbII at different concentrations (b) versus I0/I, and (c) luminescence of the mixtures with PbII at different concentrations, illuminated with a UV lamp at 365 nm at room temperature.

The ratio I0/I versus PbII concentration (Figure 4b) provides a coefficient of determination, R 2 = 0.9913, and the equation may be adjusted as (I0/I) − 1 = 2.3 × 10 3[PbII] − 0.7074. With that, the ratio approaches the Stern-Volmer equation: (I0/I) − 1 = K SV [M], where I0 and I are the 5D0 7F2 band emission intensities before and after metal addition, respectively; [M] is the molar concentration of the metal ion; and K SV is a Stern-Volmer constant that indicates quenching intensity.7777 Zeng, X.; Zhang, Y.; Zhang, J.; Hu, H.; Wu, X.; Long, Z.; Hou, X.; Microchem. J. 2017, 134, 140. Based on the approximation, the K SV value for [Eu(TTA)3(2-pyr)(H2O)] in the presence of PbII was equal to 2300 L mol−1. Limit of detection (LOD) was calculated using the equation LOD = 3σ/k, where σ is the blank measurement standard deviation and k is the slope (I0/I) vs. [PbII],7878 Zhang, M. H.; Zhang, L. L.; Xiao, Z. Y.; Zhang, Q. H.; Wang, R. M.; Dai, F. N.; Sun, D. F.; Sci. Rep. 2016, 6, 20672. reaching the value of 6.03 µM. In a study conducted by Lin et al.,7979 Lin, J.; Cheng, Q.; Zhou, J.; Lin, X.; Reddy, R. C. K.; Yang, T.; Zhang, G.; J. Solid State Chem. 2019, 270, 339. the authors evaluated an europium(III) coordination polymer as a sensor for PbII, finding K SV to be equal to 33014 L mol−1 and LOD to 90.78 µM in water. Lin et al.7979 Lin, J.; Cheng, Q.; Zhou, J.; Lin, X.; Reddy, R. C. K.; Yang, T.; Zhang, G.; J. Solid State Chem. 2019, 270, 339. synthesized europium(III) coordination polymer Eu-CPs, {[Eu2(PBA)3(H2O)3]∙DMF∙3H2O}n, this compound exhibited 3D framework, what increases the weak bonding interaction between nitrogen atoms and PbII. This interaction perturbs the electronic structure of the ligand and reduce the energy transfer efficiency from the ligand to EuIII centers. In the present work [Eu(TTA)3(2-pyr) (H2O)] compound did not exhibit 3D framework.

Sensor to detect ethanol in gasoline

To evaluate the effect of the gasoline (ethanol mixture in the 0-100% (v/v) range) on the luminescence of the complex, we monitored the 5D0 7F2 transition (612 nm) in the emission spectra (Figure 5a) of [Eu(TTA)3(2-pyr)(H2O)] in the ethanol-gasoline mixture (% v/v). The increase in the proportion of gasoline by ethanol volume significantly affected the 5D0 7F2 transition, increasing luminescence (Figure 5a). This may be explained by the fact that the chemical environment where the complex is located consists mostly of an apolar solvent, minimizing energy losses by a non-radioactive process. Quenching probability through OH oscillators resonance increases with a greater amount of ethanol,55 Muller, G.; Dalton Trans. 2009, 44, 9692. as shown in Figure 5b.

Figure 5
(a) Emission spectra (λex = 375 nm) monitoring the intensity of the 5D0 7F2 transition (612 nm) of [Eu(TTA)3(2-pyr)(H2O)] in the ethanol-gasoline mixture (% v/v). (b) Luminescence of the complex mixed with different volumes of ethanol in gasoline, under ultraviolet irradiation at 365 nm at room temperature and (c) I0/I ratio versus ethanol concentration in gasoline (% v/v).

The ethanol-gasoline mixtures (% v/v) (Figure 5b) presented the characteristic luminescence of the complex in the red region when illuminated with a UV lamp (at 365 nm). Based on the regulation that determines the permitted volume of ethanol in gasoline,8080 Romanel, S. A.; Cunha, D. A.; Castro, E. V. R.; Barbosa, L. L.; Microchem. J. 2018, 140, 31. we blended ethanol in the following concentrations: 20, 30, 40 50, and 60% v/v. Concentrations were related by the I0/I ratio, where I0 and I are the emission intensities of the 5D0 7F2 band before and after adding ethanol to gasoline, respectively. With that, we obtained a determination coefficient (R 2) equal to 0.9815, and calculated LOD (4.94% v/v) using the following equation: LOD = 3σ/k, where σ is the blank measurement standard deviation and k is the slope (I0/I) vs. ethanol (% v/v).7878 Zhang, M. H.; Zhang, L. L.; Xiao, Z. Y.; Zhang, Q. H.; Wang, R. M.; Dai, F. N.; Sun, D. F.; Sci. Rep. 2016, 6, 20672. Our results indicate that using the complex as a luminescent marker to identify ethanol in gasoline is a promising alternative for fuel quality control.

Conclusions

Our study verifed that the use of [Eu(TTA)3(2-pyr)(H2O)] complex as luminescent sensor is a promising alternative for detecting PbII. We evaluated PbII concentration based on the luminescence quenching of the complex, finding a great sensitivity based on the values found for K SV = 2300 L mol−1, R 2 = 0.9913, and LOD = 6.03 µM. Our results reveal that the complex could be used for identifying and quantifying PbII. We also tested the use of the complex as a luminescent marker to identify gasoline adulteration with the addition of ethanol, reaching the values of R 2 = 0.9815 and LOD = 4.94% v/v. These values indicate that the complex is sensitive and efficient, being promising for fuel quality control.

  • Supplementary Information
    Supplementary data (spectral data, FTIR and UV-Vis, TG/DTG, and DTA) are available free of charge at http://jbcs.sbq.org.br as a PDF file.

Acknowledgments

The authors thank CAPES (23038.007083/2014-40), FAPES (edital CNPq/FAPES No. 23/2018, Programa de Apoio a Núcleos Emergentes, PRONEM), and CNPq (422555/2018-5, and 305359/2017-7) for financial support. The authors would also like to thank the Núcleo de Competências em Química do Petróleo and LabPetro for the use of their installations. This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior, Brasil (CAPES) - Finance Code 1.

References

  • 1
    Rostra, J. G.; Ferrer, F. J.; Espinós, J. P.; Elipe, A. R. G.; Yubero, F.; ACS Appl. Mater. Interfaces 2017, 9, 16313.
  • 2
    Du, P.; Bharat, L. K.; Yu, J. S.; J. Alloys Compd. 2015, 633, 37.
  • 3
    Ling, Q.; Song, Y.; Ding, S. J.; Zhu, C.; Chan, D. S. H.; Kwong, D. L.; Kang, E. T.; Neoh, K. G.; Adv. Mater. 2005, 17, 455.
  • 4
    Reichardt, C.; Schneider, K. R. A.; Sainuddin, T.; Wächtler, M.; McFarland, S. A.; Dietzek, B.; J. Phys. Chem. A 2017, 121, 5635.
  • 5
    Muller, G.; Dalton Trans. 2009, 44, 9692.
  • 6
    Jia, J.; Zhang, Y.; Zheng, M.; Shan, C.; Yan, H.; Wu, W.; Gao, X.; Cheng, B.; Liu, W.; Tang, Y.; Inorg. Chem. 2018, 57, 300.
  • 7
    Kumar, P.; Kanika; Singh, S.; Lahon, R.; Gundimeda, A.; Gupta, G.; Gupta, B. K.; J. Lumin. 2018, 196, 207.
  • 8
    Kuncewicz, J.; Dąbrowski, J. M.; Kyzioł, A.; Brindell, M.; Łabuz, P.; Mazuryk, O.; Macyk, W.; Stochel, G.; Coord. Chem. Re v. 2019, 398, 113012.
  • 9
    Kariaka, N. S.; Smola, S. S.; Rusakova, N. V.; Dyakonenko, V. V.; Shishkina, S. V.; Sliva, T. Y.; Trush, V. A.; Amirkhanov, V. M.; J. Lumin. 2020, 223, 117187.
  • 10
    Rosa, P. P. F.; Kitagawa, Y.; Hasegawa, Y.; Coord. Chem. Rev. 2020, 406, 213153.
  • 11
    Cotton, S. A.; Raithby, P. R.; Coord. Chem. Rev. 2017, 340, 220.
  • 12
    You, L. X.; Hao, J. H.; Qi, D.; Xie, S. Y.; Wang, S. J.; Xiong, G.; Dragutan, I.; Dragutan, V.; Ding, F.; Sun, Y. G.; Inorg. Chim. Acta 2019, 497, 119075.
  • 13
    Karraker, D. G.; J. Chem. Educ. 1970, 47, 424.
  • 14
    Yanagisawa, K.; Nakanishi, T.; Kitagawa, Y.; Seki, T.; Akama, T.; Kobayashi, M.; Taketsugu, T.; Ito, H.; Fushimi, K.; Hasegawa, Y.; Eur. J. Inorg. Chem. 2015, 28, 4769.
  • 15
    Burdett, J. K.; Hoffmann, R.; Fay, R. C.; Inorg. Chem. 1978, 17, 2553.
  • 16
    Kuramochi, Y.; Nakagawa, T.; Yokoo, T.; Yuasa, J.; Kawai, T.; Hasegawa, Y.; Dalton Trans. 2012, 41, 6634.
  • 17
    Bünzli, J. C. G.; Chauvin, A. S.; Vandevyver, C. D. B.; Bo, S.; Comby, S.; Ann. N. Y. Acad. Sci. 2008, 1130, 97.
  • 18
    Heffern, M. C.; Matosziuk, L. M.; Meade, T. J.; Chem. Rev. 2014, 114, 4496.
  • 19
    Braun, F.; Comba, P.; Grimm, L.; Herten, D. P.; Pokrandt, B.; Wadepohl, H.; Inorg. Chim. Acta 2018, 484, 464.
  • 20
    Zhang, J. W.; Wang, C. R.; Liu, W. H.; Xu, S.; Liu, B. Q.; Inorg. Chim. Acta 2020, 508, 119648.
  • 21
    Bai, C.; Fu, X. Y.; Hu, H. M.; He, S.; Wang, X.; Xue, G. L.; Inorg. Chim. Acta 2020, 506, 119550.
  • 22
    Chen, W. T.; J. Solid State Chem. 2019, 284, 121160.
  • 23
    Ning, Y.; Zhu, M.; Zhang, J. L.; Coord. Chem. Rev. 2019, 399, 213028.
  • 24
    Abdelhamid, H. N.; Wilk-Kozubek, M.; El-Zohry, A. M.; Gómez, A. B.; Valiente, A.; Martín-Matute, B.; Mudring, A. V.; Zou, X.; Microporous Mesoporous Mater. 2019, 279, 400.
  • 25
    Jia, J. H.; Li, Q. W.; Chen, Y. C.; Liu, J. L.; Tong, M. L.; Coord. Chem. Rev. 2019, 378, 365.
  • 26
    Maouche, R.; Belaid, S.; Benmerad, B.; Bouacida, S.; Freslon, S.; Daiguebonne, C.; Suffren, Y.; Calvez, G.; Bernot, K.; Roiland, C.; Pollès, L.; Guillou, O.; Inorg. Chim. Acta 2020, 501, 119309.
  • 27
    Sun, Y.; Li, Q.; Wei, S.; Zhao, R.; Han, J.; Ping, G.; J. Lumin. 2020, 225, 117241.
  • 28
    Stan, C. S.; Rosca, I.; Sutiman, D.; Secula, M. S.; J. Rare Earths 2012, 30, 401.
  • 29
    Zhao, Y.; Wang, S.; Han, Y.; Zhang, J.; Liu, C.; Hu, X.; Zhang, Z.; Wang, L.; J. Lumin. 2020, 223, 117253.
  • 30
    Yang, D.; Liao, L.; Zhang, Y.; Guo, Q.; Mei, L.; Liu, H.; J. Lumin. 2020, 224, 117285.
  • 31
    Baklanova, Y. V.; Maksimova, L. G.; Lipina, O. A.; Tyutyunnik, A. P.; Zubkov, V. G.; J. Lumin. 2020, 224, 117315.
  • 32
    Herrmann, A.; Friedrich, D.; Zscheckel, T.; Rüssel, C.; J. Lumin. 2019, 214, 116550.
  • 33
    Kaczmarek, M.; J. Lumin. 2020, 222, 117174.
  • 34
    Binnemans, K.; Chem. Rev. 2009, 109, 4283.
  • 35
    Sizov, V. S.; Komissar, D. A.; Metlina, D. A.; Aminev, D. F.; Ambrozevich, S. A.; Nefedov, S. E.; Varaksina, E. A.; Metlin, M. T.; Mislavskií, V. V.; Taydakova, I. V.; Spectrochim. Acta, Part A 2020, 225, 117503.
  • 36
    Duan, Y. Y.; Wu, D. F.; Chen, H. H.; Wang, Y. J.; Li, L.; Gao, H. L.; Cu, J. Z.; Inorg. Chim. Acta 2020, 499, 119203.
  • 37
    Berezhnytska, O. S.; Savchenko, I. O.; Ivakha, N. B.; Rogovtsov, O. O.; Trunova, O. K.; Rusakova, N. V.; J. Mol. Struct. 2020, 1201, 127160.
  • 38
    Bhat, S. A.; Iftikhar, K.; Dyes Pigm. 2020, 179, 108383.
  • 39
    Borges, A. S.; Fulgêncio, F.; Silva, J. G.; Santos, T. A. R.; Diniz, R.; Windmöller, D.; Magalhães, W. F.; Araujo, M. H.; J. Lumin. 2019, 205, 72.
  • 40
    Borges, A. S.; Caliman, E. V.; Dutra, J. D. L.; Silva, J. G.; Araujo, M. H.; J. Lumin. 2016, 170, 654.
  • 41
    Lucena, M. A. M.; Sá, G. F.; Rodrigues, M. O.; Alves, S.; Talhavini, M.; Weber, I. T.; Anal. Methods 2013, 5, 705.
  • 42
    Destefani, C. A.; Motta, L. C.; Costa, R. A.; Macrino, C. J.; Bassane, J. F. P.; Filho, J. F. A.; Silva, E. M.; Greco, S. J.; Carneiro, M. T. W. D.; Endringer, D. C.; Romão, W.; Microchem. J. 2016, 124, 195.
  • 43
    Devi, R.; Rajendran, M.; Singh, K.; Pala, R.; Vaidyanathan, S.; J. Mater. Chem. C 2021, 9, 6618.
  • 44
    Wang, Q.; Long, M. Y.; Lv, C. Y.; Xin, S. P.; Han, X. G.; Jiang, W.; Food Control 2020, 109, 106894.
  • 45
    Zmojda, J.; Kochanowicz, M.; Miluski, P.; Golonko, P.; Baranowska, A.; Ragi, T.; Dorosz, J.; Kuwik, M.; Pisarski, W.; Pisarska, J.; Szal, R.; Mach, G.; Starzyk, B.; Lesniak, M.; Sitarz, M.; Dorosz, D.; Materials 2020, 13, 2817.
  • 46
    Zhang, D.; Zhang, Y.; Wang, Z.; Zheng, Y.; Zheng, X.; Gao, L.; Wang, C.; Yang, C.; Tang, H.; Li, Y.; J. Lumin. 2021, 229, 117706.
  • 47
    Li, R. F.; Zhang, T.; Liu, X. F.; Feng, X.; Inorg. Chem. Commun. 2016, 73, 170.
  • 48
    Zhou, J.; Li, B.; Qi, A.; Shi, Y.; Qi, J.; Xu, H.; Chen, L.; Sens. Actuators, B 2019, 305, 127462.
  • 49
    Sun, Z.; Li, H.; Sun, G.; Guo, J.; Ma, Y.; Li, L.; Inorg. Chim. Acta 2018, 469, 51.
  • 50
    Chen, Z.; Yu, X.; Li, X.; Ye, Q.; Zhou, K.; Cai, Y.; Huang, L.; Li, Y.; Zeng, C.; Inorg. Chem. Commun. 2020, 112, 107744.
  • 51
    Zhang, F.; Ma, J.; Huang, S.; Li, Y.; Spectrochim. Acta, Part A 2020, 228, 117816.
  • 52
    Zhai, B.; Li, Z. Y.; Wub, Z. L.; Cui, J. Z.; Inorg. Chem. Commun. 2016, 71, 23.
  • 53
    Li, R.; Zhang, Y.; Liu, X.; Chang, X.; Feng, X.; Inorg. Chim. Acta 2020, 502, 119370.
  • 54
    Zheng, M.; Tan, H.; Xie, Z.; Zhang, L.; Jing, X.; Sun, Z.; ACS Appl. Mater. Interfaces 2013, 5, 1078.
  • 55
    Yan, Q.; Chen, Z. H.; Xue, S. F.; Han, X. Y.; Lin, Z. Y.; Zhang, S.; Shi, G.; Zhang, M.; Sens. Actuators, B 2020, 268, 108.
  • 56
    Xu, X. Y.; Yan, B.; Sens. Actuators, B 2016, 222, 347.
  • 57
    Lou, T.; Chen, Z.; Wang, Y.; Chen, L.; ACS Appl. Mater. Interfaces 2011, 3, 1568.
  • 58
    Lian, C.; Chen, Y.; Li, S.; Hao, M.; Gao, F.; Yang, L.; J. Alloys Compd. 2017, 702, 303.
  • 59
    Yang, L.; Song, S.; Zhang, H.; Zhang, W.; Bu, Z.; Ren, T.; Synth. Met. 2011, 161, 2230.
  • 60
    Fonseca, R. R. F.; Gaspar, R. D. L.; Raimundo Jr., I. M.; Luz, P. P.; J. Rare Earths 2019, 37, 225.
  • 61
    Xue, B.; Zhang, Z.; Sund, Y.; Wang, J.; Jiang, H.; Du, M.; Chi, C.; Li, X.; Carbohydr. Polym. 2018, 186, 176.
  • 62
    Teotonio, E. E. S.; Fett, G. M.; Brito, H. F.; Sá, W. M. F. G. F.; Felinto, M. C. F. C.; Santos, R. H. A.; J. Lumin. 2008, 128, 190.
  • 63
    Li, Y. J.; Yan, B.; Li, Y.; Microporous Mesoporous Mater. 2010, 131, 82.
  • 64
    Li, Y.; Jiao, J.; Yan, P.; Liu, L.; Wang, J.; Wang, Y.; Huang, L.; Liu, J.; Belfore, L. A.; Scr. Mater. 2018, 152, 1.
  • 65
    Charles, R. G.; Ohlmann, R. C.; J. Inorg. Nucl. Chem. 1965, 27, 255.
  • 66
    Zou, Y.; Wang, X.; Khan, A.; Wang, P.; Liu, Y.; Alsaedi, A.; Hayat, T.; Wang, X.; Environ. Sci. Technol. 2016, 50, 7290.
  • 67
    Wu, D.; Chen, L.; Lee, W.; Ko, G.; Yin, J.; Yoon, J.; Coord. Chem. Rev. 2018, 354, 74.
  • 68
    Lyle, S. J.; Rahaman, M. M.; Talanta 1963, 10, 1177.
  • 69
    Ugale, A.; Kalyani, N. T.; Dhoble, S. J.; Mater. Sci. Energy Technol. 2020, 3, 51.
  • 70
    Oliveira, T. C.; Santos, H. P.; Lahoud, M. G.; Franco, D. F.; Freire, R. O.; Dutra, J. D. L.; Cuin, A.; Lima, J. F.; Marques, F. L.; J. Lumin. 2017, 181, 196.
  • 71
    Chitnis, D.; Kalyani, N. T.; Dhoble, S. J.; Optik 2017, 130, 237.
  • 72
    Marques, L. F.; Santos, H. P.; Oliveira, K. A.; Botezine, N. P.; Freitas, M. C. R.; Freire, R. O.; Dutra, J. D. L.; Martins, J. S.; Legnani, C.; Quirino, W. G.; Machado, F. C.; Inorg. Chim. Acta 2016, 458, 28.
  • 73
    Donegá, C. D.; Alves Jr., S.; Sá, G. F.; J. Alloys Compd. 1997, 250, 422.
  • 74
    Marques, L. F.; Cuin, A.; Carvalho, G. S. G.; Santos, M. V.; Ribeiro, S. J. L.; Machado, F. C.; Inorg. Chim. Acta 2016, 441, 67.
  • 75
    Yan, Y. T.; Zhang, W. Y.; Zhang, F.; Cao, F.; Yang, R. F.; Wang, Y. Y.; Hou, L.; Dalton Trans. 2018, 47, 1682.
  • 76
    Song, X. Z.; Wang, Y. X.; Yan, J. W.; Chen, X.; Meng, Y. L.; Tan, Z.; ChemistrySelect 2018, 3, 9564.
  • 77
    Zeng, X.; Zhang, Y.; Zhang, J.; Hu, H.; Wu, X.; Long, Z.; Hou, X.; Microchem. J. 2017, 134, 140.
  • 78
    Zhang, M. H.; Zhang, L. L.; Xiao, Z. Y.; Zhang, Q. H.; Wang, R. M.; Dai, F. N.; Sun, D. F.; Sci. Rep. 2016, 6, 20672.
  • 79
    Lin, J.; Cheng, Q.; Zhou, J.; Lin, X.; Reddy, R. C. K.; Yang, T.; Zhang, G.; J. Solid State Chem. 2019, 270, 339.
  • 80
    Romanel, S. A.; Cunha, D. A.; Castro, E. V. R.; Barbosa, L. L.; Microchem. J. 2018, 140, 31.

Edited by

Editors handled this article: Jaísa Fernandes Soares and Humberto O. Stumpf (Associate)

Publication Dates

  • Publication in this collection
    24 Jan 2022
  • Date of issue
    2022

History

  • Received
    09 June 2021
  • Accepted
    27 Sept 2021
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