Please use this identifier to cite or link to this item: https://rima.ufrrj.br/jspui/handle/20.500.14407/14597
Full metadata record
DC FieldValueLanguage
dc.contributor.authorAreas, Esther Saraiva
dc.date.accessioned2023-12-22T03:03:22Z-
dc.date.available2023-12-22T03:03:22Z-
dc.date.issued2018-02-01
dc.identifier.citationAREAS, Esther Saraiva. Síntese de novos complexos de cobalto contendo híbrido de cumarina--cetoéster: investigação da citotoxidade e reatividade em solução. 2018. 115 f. Dissertação (Programa de Pós-Graduação em Química) - Instituto de Ciências Exatas, Universidade Federal Rural do Rio de Janeiro, [Seropédica - RJ] .por
dc.identifier.urihttps://rima.ufrrj.br/jspui/handle/20.500.14407/14597-
dc.description.abstractNeste trabalho realizou-se a síntese de novos complexos de cobalto do tipo [Co(L)Am](ClO4)n, onde L (etil 3-(7-(dietilamina)-2-oxo-2H-cromen-3-il)-3-oxopropanoato) é um ligante fluorescente híbrido de cumarina--cetoéster na forma aniônica e Am é uma amina auxiliar TPA (Tris(2-piridilmetil)amina) ou Py2en (bis(2-piridilmetil)-1,2-diamino). A metodologia de síntese de 1b-CoII, [CoII(L)TPA]ClO4, levou à formação da espécie de CoII, sendo confirmado por análise elementar de CHN, condutividade molar e difração de raios X (DRX). O íon complexo 1b-CoII apresentou-se como uma espécie catiônica e hexacoordenada no estado sólido, contendo a cumarina--cetoéster na forma L1- como O,O doador, e o TPA neutro como N,N’,N’,N’ doador. Para o complexo análogo 2b-CoIII, [CoIII(L)TPA](ClO4)2, a metodologia utilizada a partir de cis-[CoIII(TPA)Cl2]ClO4 levou à formação da espécie de CoIII, entretanto, a análise elementar de CHN e RMN de 1H também indicaram a presença dos precursores da reação no sólido isolado. O complexo 2c-CoIII, [CoIII(L)Py2en](ClO4)2, foi obtido partindo-se da mesma metodologia empregada para 1b-CoII, no entanto, a presença de uma amina mais básica (Py2en) levou à formação da espécie de CoIII. Dados de análise elementar de CHN, condutividade molar, espectroscopia de infravermelho e RMN de 1H apontaram para uma estrutura em que o ligante L1- também interage com o metal através das carbonilas da porção -cetoéster. Cálculos DFT mostraram que a banda em 448 nm no espectro de UV-Vis de 1b-CoII, obtido em tampão (pH = 7,4), é referente à transição  centrada em L1- mas para 2c-CoIII a banda em 455 nm apresenta-se como uma mistura de  e TCLM. Estudos de voltametria cíclica mostraram que os complexos exibem um processo quasi-reversível associado ao par CoIII/CoII, com Epc na ordem 2c-CoIII < 1b-CoII < 2b-CoIII, que foi associado às propriedades eletrônicas das aminas auxiliares. O maior caráter aceptor  do TPA facilita o ganho de elétrons pelo cobalto nos complexos 1b-CoII e 2b-CoIII, enquanto que o Py2en, que possui maior caráter doador , levou a um potencial menor para 2c-CoIII. HL exibiu um processo catódico associado à redução de carbonila, o qual não sofreu deslocamento após a complexação. Ensaios de reatividade de 2c-CoIII em solução tampão (pH 7,4) monitorados por espectroscopia de fluorescência mostraram que o ligante é liberado em presença de um agente redutor (ascorbato de sódio) durante 8,5 h, onde a liberação foi mais pronunciada em atmosfera de argônio, sugerindo que a dissociação de L1- ocorre preferencialmente após a redução do metal. 2c-CoIII apresentou-se estável em solução tampão sem agente redutor. A investigação da citotoxidade de HL, 1b-CoII e 2c-CoIII contra linhagens B16-F10 (melanoma murino metastático), 4T1 (melanoma mamário murino) e BHK-21 (célula não tumoral de rim de hamister) mostraram que HL não foi ativo, mas a coordenação favoreceu sua citotoxicidade, na maioria dos casos. 2c-CoIII exibiu melhor atividade global quando comparado com 1b-CoII, apresentando citotoxicidade seletiva às células de câncer (45,9 ± 5,2 M e 39,1 ± 6,6 M para B16-F10 e 4T1, respectivamente) e sendo inativo contra células normais (71,2 ± 6,2 M).por
dc.description.sponsorshipCoordenação de Aperfeiçoamento de Pessoal de Nível Superior, CAPES, Brasil.por
dc.formatapplication/pdf*
dc.languageporpor
dc.publisherUniversidade Federal Rural do Rio de Janeiropor
dc.rightsAcesso Abertopor
dc.subjectcomplexos de Cobaltopor
dc.subjectcumarinaspor
dc.subjectpró-fármaco biorredutívelpor
dc.subjectCobalt complexeseng
dc.subjectcoumarinseng
dc.subjectbioreductive prodrugseng
dc.titleSíntese de novos complexos de cobalto contendo híbrido de cumarina--cetoéster: investigação da citotoxidade e reatividade em soluçãopor
dc.title.alternativeSynthesis of novel cobalt complexes containing coumarin--ketoester hybrid: investigation of the cytotoxicity and reactivity in solution.eng
dc.typeDissertaçãopor
dc.description.abstractOtherIn this work it was synthesized novel cobalt complexes of the type [Co(L)(Am)](ClO4)n, where L (ethyl 3-(7-(diethylamino)-2-oxo-2H-chromen-3-yl)-3-oxopropanoate) is a fluorescent hybrid ligand of coumarin--ketoester in its anionic form and Am is the auxiliary amine TPA (Tris(2-pyridylmethyl)amine) or Py2en (bis(2-pyridylmethyl)-1,2-diamine). The synthetic procedure to obtain 1b-CoII, [CoII(L)TPA]ClO4, led to the formation of the CoII specie, which was confirmed by CHN elemental analysis, molar conductivity and X-ray diffraction (XRD). The 1b-CoII ion complex is a cationic hexacoordinated specie in the solid state, containing the coumarin--ketoester in the L1- form, acting as O,O-donor, and a neutral TPA as N,N',N',N'-donor. For analogous complex 2b-CoIII, [CoIII(L)TPA](ClO4)2, the methodology used from cis-[CoIII(TPA)Cl2]ClO4 led to the formation of the CoIII species, however, CHN elemental analysis and 1H NMR also indicated the presence of the precursors in the isolated solid. The complex 2c-CoIII, [CoIII(L)Py2en](ClO4)2, was obtained from the same methodology used for 1b-CoII, however, the presence of a more basic amine (Py2en) favored the CoIII specie. Elemental analysis, molar conductivity, infrared and 1H NMR spectroscopy data indicated a structure of the type [CoIII(L)(Py2en)](ClO4)2, where L1- also interacts with the metal through the carbonyl groups of the -ketoester moiety. DFT calculations showed that the band at 448 nm in the UV-Vis spectra of 1b-CoII, obtained in buffer solution (pH 7,4), refers to the transition centered on L1-, and for 2c-CoIII the band at 455 nm is related to a mixture of and LMCT transitions. Cyclic voltammetry showed that the complexes exhibit a quasi-reversible process associated with the CoIII/CoII pair, with Epc in the order 2c-CoIII < 1b-CoII < 2b-CoIII, which was associated to the electronic properties of the auxiliary amine. The higher -acceptor character of the TPA facilitates the electron gain by the cobalt in 1b-CoII and 2b-CoIII, while Py2en, which has a larger -donor character, leads to a lower potential for 2c-CoIII. HL exhibits a cathodic process associated to the reduction of the carbonyl groups, which did not shift after coordination. Reactivity assays of 2c-CoIII in buffer solution (pH 7,4) monitored by fluorescence spectroscopy indicated that the ligand was released in the presence of a reducing agent (sodium ascorbate) for 8.5 h, which was most pronounced in argon atmosphere, suggesting that dissociation of L1- occurs preferentially after reduction of the metal. 2c-CoIII was stable in buffer solution without reducing agent. The cytotoxicity investigation of HL, 1b-CoII and 2c-CoIII against B16-F10 (metastatic murine melanoma), 4T1 (murine mammary melanoma) and BHK-21 (non-tumor cell of hamster kidney) showed that the free ligand was not active, but the coordination favored its cytotoxicity in most cases. 2c-CoIII exhibited the best overall activity when compared to 1b-CoII, presenting selective cytotoxicity to cancer cells (45.9 ± 5.2 M e 39.1 ± 6.6 M para B16-F10 e 4T1, respectively) and being inactive against normal cells (71.2 ± 6.2 M).eng
dc.contributor.advisor1Neves, Amanda Porto
dc.contributor.advisor1ID055.195.997-50por
dc.contributor.advisor1Latteshttp://lattes.cnpq.br/7460226353493536por
dc.contributor.referee1Neves, Amanda Porto
dc.contributor.referee2Herbst, Marcelo Hawrylak
dc.contributor.referee3Lanznaster, Maurício
dc.creator.ID130.493.877-85por
dc.creator.Latteshttp://lattes.cnpq.br/4784063066540234por
dc.publisher.countryBrasilpor
dc.publisher.departmentInstituto de Ciências Exataspor
dc.publisher.initialsUFRRJpor
dc.publisher.programPrograma de Pós-Graduação em Químicapor
dc.relation.references[1] RENFREW, A. K., O'NEILL, E. S., HAMBLEY, T.W. Harnessing the properties of cobalt coordination complexes for biological application. Coord. Chem. Rev., Sydney, v. 00, p. 0-0, 2017. ISSN 10.1016/j.ccr.2017.11.027. [2] SAKAR, T., BANERJEE, S. E., HUSSAIN, A. Remarkable visible light-triggered cytotoxicity of mitochondria targeting mixed-ligand cobalt(III) complexes of curcumin and phenanthroline bases binding to human serum albumin. RSC Adv., v. 5, p. 16641-16654, 2015. [3] MUNTEANU, C. R; SUNTHARALINGAM, K., Advances in cobalt complexes as anticancer agents. Dalton Trans.,v. 44, p. 13796–13808, 2015. [4] GRAF, N.; LIPPARD, S. J., Redox activation of metal-based prodrugs as a strategy for drug delivery. Adv. Drug Deliver Rev., v. 64, p. 993–1004, 2012. [5] WANG, Z.; WU, Q.; LI, J.; QIU, S.; CAO, D.; XU, Y.; LIU, Z.; YU, X.; SUN, Y., Two benzoyl coumarin amide fluorescence chemosensors for cyanide anions, Spectrochim. Acta A, v. 183, p. 1–6, 2017. [6] SOUZA, G. A., Caracterização fotofísica e morfológica de estruturas peptídicas contendo composto fluorescente para aplicação ambiental, 2014, Universidade Federal de Goiás, Instituto de Química, Goiânia, 2014. Dissertação de Mestrado. [7] MONTAGNER, C.; SAVI, L. A.; SILVA, I. T.; KRATZ, J. M.; SMÂNIA, E. F. A.; MONACHE, F. D.; SMÂNIA, A. Jr.; SIMÕESI, C. M. O. Vitro Cytotoxic Screening of Coumarins. Lat. Am. J. Pharm., v. 30, p. 1669-1674, 2011. [8] THAKUR, A.; SINGLA, R.; JAITAK, V., Coumarins as anticancer agents: A review on synthetic strategies, mechanism of action and SAR studies, Eur. J. Med. Chem., v. 101, p. 476-495, 2015. [9] YAMAMOTO, N.; RENFREW, A.K.; KIM, B.J.; BRYCE, N.S.; HAMBLEY, T.W., Dual Targeting of Hypoxic and Acidic Tumor Environments With a Cobalt (III) Chaperone Complex. J. Med. Chem., v. 55, p. 11013-11021, 2012. [10] YAMAMOTO, N.; DANOS, S.; BONNITCHA, P. D.; FAILES, T. W.; NEW, E. J.; HAMBLEY, T. W., Cellular uptake and distribution of cobalt complexes of fluorescent ligands. J. Biol. Inorg. Chem., v. 13, p. 861–871, 2008. [11] LEE, J. H.; JEONG, R.; IK-SOO SHIN; KIM, H.; HONG, J., Fluorescence Turn-On Sensor for Cyanide Based on a Cobalt(II)−Coumarinylsalen Complex. Org. Lett., v. 12, p. 764–767, 2010. 97 [12] VITÓRIO, F.; MOREIRA, T. P.; CASTRO, R. N.; GUEDES, G. P.; GRAEBINAB, C. S.; KUMMERLE, A. E., Synthesis and mechanism of novel fluorescent coumarin–dihydropyrimidinone dyads obtained by the Biginelli multicomponent reaction, New J. Chem., v. 39, p. 2323-2332, 2015. [13] BAILE, M. B.; KOLHE, N. S.; DEOTARSE P. P.; JAIN, A. S.; KULKARNI, A. A., Metal Ion Complex -Potential Anticancer Drug- A Review. Int. J. Pharm., v. 4, p. 59-66, 2015. [14] PUCCI, D.; CRISPINI, A.; MENDIGUCHÍA, B. S.; PIRILLO, S.; GHEDINI, M.; MORELLI, S.; DE BARTOLO, L., Improving the bioactivity of Zn(II)-curcumin based complexes, Dalton Trans., v. 42, p. 9679–9687, 2013. [15] Food and Drugs Administration. Disponível em: https://www.accessdata.fda.gov/scripts/cder/daf/. Acesso em: 05/01/2018. [16] Agência Nacional de Vigilância Sanitária. Disponível em: https://consultas.anvisa.gov.br/#/medicamentos/. Acesso em: 05/01/2018. [17] BAR-SELA, G.; EPELBAUM.; R.; SCHAFFER, M., Curcumin as an anti-cancer agent: review of the gap between basic and clinical applications, Curr. Med. Chem., v. 17, p. 190–197, 2010. [18] BANSAL, S. S.; GOEL, M.; AQIL, F.; VADHANAM, M. V.; GUPTA, R. C., Advanced drug delivery systems of curcumin for cancer chemoprevention, Cancer Prev. Res., v. 4, p. 1158–1171, 2011. [19] CASTONGUAY, A.; DOUCET, C.; JUHAS, M.; MAYSINGER, D., New Ruthenium(II)–Letrozole Complexes as Anticancer Therapeutics, J. Med. Chem., v. 55, p. 8799–8806, 2012. [20] HUANG, Q.-M.; WANG, S.-W.; LI, Q.; PAN, W.; DENG, P.-X.; ZHOU, H.; PAN, Z.-Q., Synthesis and characterization of curcumin bridged porphyrins as photosensitizers, Chem. J. Chin. Univers., v. 33, p. 732-737, 2012. [21] ANNARAJ, J.; SRINIVASAN, S.; PONVEL, K.M.; ATHAPPAN, P. R.; Mixed ligand copper(II) complexes of phenanthroline/bipyridyl and curcumin diketimines as DNA intercalators and their electrochemical behavior under Nafion® and clay modified electrodes, J. Inorg. Biochem., v. 99, p. 669-676, 2005. [22] ALIAGA-ALCALDE, N.; MARQUÉS-GALLEGO, P.; KRAAIJKAMP, M.; HERRANZ-LANCHO, C.; DULK, H.; GÖRNER, H.; ROUBEAU, O.; TEAT, S. J.; WEYHERMÜLLER, T.; REEDIJK, J., Copper Curcuminoids Containing Anthracene 98 Groups: Fluorescent Molecules with Cytotoxic Activity. Inorg. Chem., v. 49, p. 9655–9663, 2010. [23] PRÖHLA, M.; SCHUBERTA, ULRICH S.; WEIGAND, W.; GOTTSCHALDT, M., Metal complexes of curcumin and curcumin derivatives for molecular imaging and anticancer therapy, Coord. Chem. Rev., v. 307, p. 32-41, 2016. [24] GRAF, N.; LIPPARD, S. J., Redox activation of metal-based prodrugs as a strategy for drug delivery, Adv. Drug Deliv. Rev., v. 64, p. 993–1004, 2012. [25] REISNER, E.; ARION, V. B.; KEPPLER, B. K.; POMBEIRO, ARMANDO J. L., Electron-transfer activated metal-based anticancer drugs. Inorg. Chim. Acta, v. 361, p. 1569–1583, 2008. [26] WERMUTH, C. G.; GANELLIN, C. R.; LINDBERG, P.; MITSCHER, L. A., Glossary of terms used in medicinal chemistry (IUPAC Recommendations 1998), Pure Appl. Chem., v. 70, p. 1129-1135, 1998. [27] PIRES, B. M., GIACOMIN, L. C., CASTRO, F. A.V., CAVALCANTI, A. DOS S., PEREIRA, M. D., BORTOLUZZI, A. J., FARIA, R. B., SCARPELLINI, M., Azido- and Chlorido-cobalt complex as carrier-prototypes for antitumoral prodrugs. J. Inorg. Biochem., v. 157, p. 104–113, 2016. [28] FILES, T. W., HAMBLEY, T. W. Models of Hypoxia Activated Prodrugs: Co(III) Complexes of Hydroxamics Acids. Dalton Trans., p. 1895-1901, 2005. [29] BROWN, J. M. E WILSON, W. R., Exploiting tumour hypoxia in cancer treatment. Nat Rev. cancer, v. 4, p. 437-447, 2004. [30] WILSON, W. R. E HAY, M. P., Targeting hypoxia in cancer therapy. Nat. Rev. Cancer, v. 11, p. 393-410, 2011. [31] BUSTAMANTE, F. L. S.; SOUZA, E. T.; LANZNASTER, M.; SCARPELLINI, M., Complexos ativados por hipóxia: uma estratégia para o combate ao câncer. Rev. Virtual Quim., v. 1, p. 138-148, 2009. [32] ZHANG, X.; LI, X.; YOU, Q.; ZHANG, X.; Prodrug strategy for cancer cell-specific targeting: A recent overview. Eur. J. Med. Chem., v. 139, p. 542-563, 2017. [33] MISTRY, I. N.; THOMAS, M.; CALDER, E. D. D.; CONWAY, S. J.; HAMMOND, E. M., Clinical Advances of Hypoxia-Activated Prodrugs in Combination with Radiation Therapy. Int. J. Radiation Oncol. Biol. Phys., v. 98, p. 1183-1196, 2017. 99 [34] BUGLYÓA, P.; KACSIR, I.; KOZSUP, M.; NAGY, I.; NAGY, S.; BÉNYEI, A. C.; KOVÁTS, E.; FARKAS, E., Tuning the redox potentials of ternary cobalt(III) complexes containing various hydroxamates. Inorg. Chim. Acta, 2017, http://dx.doi.org/10.1016/j.ica.2017.07.026. [35] GARCIA, C. V., PARRILHA, G. L., RODRIGUES, B. L., BARBEIRA, P. J. S., CLARKE, R. M., T., BERALDO, H., Cobalt(III) Complexes with 2-acetylpyridine-derived Schiff bases: studies investigating ligand release upon reduction, Polyhedron, v. 124, p. 86–95, 2017. [36] KIM, B. J.; HAMBLEY, T. W.; BRYCE, N. S., Visualising the hypoxia selectivity of cobalt(III) prodrugs, Chem. Sci., v. 2, p. 2135-2149, 2011. [37] ABREU, F. C.; FERRAZ, P. A. DE L.; GOULART, M. O. F., Some Applications of Electrochemistry in Biomedical Chemistry. Emphasis on the Correlation of Electrochemical and Bioactive Properties, J. Braz. Chem. Soc., v. 13, p. 19-35, 2002. [38] HEFFERN, M. C.; YAMAMOTO, N.; HOLBROOK, R. J.; ECKERMANN, A. L.; MEADE, THOMAS J., Cobalt Derivatives as Promising Therapeutic Agents. Curr. Opin. Chem. Biol., v. 17, p. 189–196, 2013. [39] FARRER, N. J.; SALASSA, L.; SADLER, P. J., Photoactivated chemotherapy (PACT): the potential of excitedstate d-block metals in medicine. Dalton Trans., v. 28, p. 10690–10701, 2009. [40] ROY, S.; SAHA, S.; MAJUMDAR, R.; DIGHE, R. R.; JEMMIS, E. D.; CHAKRAVARTY, A. R., Cobalt(II) complexes of terpyridine bases as photochemotherapeutic agents showing cellular uptake and photocytotoxicity in visible light, Dalton Trans., v. 40, p. 1233-1242, 2011. [41] SAHA, S.; MAJUMDAR, R.; DIGHE, R. R.; CHAKRAVARTY, A. R., Enhanced photodynamic effect of cobalt(III) dipyridophenazine complex on thyrotropin receptorexpressing HEK293 cells. Metallomics, v. 2, p. 754-765, 2010. [42] GREEN, B. P.; RENFREW, A. K.; GLENISTER, A.; TURNER, P.; HAMBLEY, TREVOR W., The influence of the ancillary ligand on the potential of cobalt(III) complexes to act as chaperones for hydroxamic acid-based drugs, Dalton Trans., v. 46, p. 15897-15918, 2017. [43] GREENWOOD, N. N., EARNSHAW, A. Chemistry of the Elements. 2ª edição, Editora Butterwoeth-Heinemann, 1997. 100 [44] BONNITCHA, P. D., KIM, B. J., HOCKING, R. K., CLEGG, J. K., TURNER, P., NEVILLEA, S. M. E HAMBLEY, T. W., Cobalt complexes with tripodal ligands: implications for the design of drug chaperones. Dalton Trans., v. 41, p. 11293-11304, 2012. [45] AHN, G-O.; BOTTING, K. J.; PATTERSON, A. V.; WARE, D. C.; TERCEL, M.; WILSON, W. R., Radiolytic and cellular reduction of a novel hypoxia-activated cobalt(III) prodrug of a chloromethylbenzindoline DNA minor groove alkylator, Biochem. Pharmacol., v. 71, p. 1683–1694, 2006. [46] WARE, D. C., BROTHERS, P. J., CLARK, G. R., DENNY, W. A., PALMER, B. D. E WILSON, W. R., Synthesis, structures and hypoxia-selective cytotoxicity of cobalt(III) complexes containing tridentate amine and nitrogen mustard ligands. J. Chem. Soc., Dalton Trans., p. 925–932, 2000. [47] RENFREW, A. K., BRYCE, N. S. E HAMBLEY, T. W., Delivery and release of curcumin by a hypoxia-activated cobalt chaperone: a XANES and FLIM study. Chem. Sci., v. 4, p. 3731- 3739, 2013. [48] CHANG, J. Y., STEVENSON, R. J., LU, G. L., BROTHERS, P. J., CLARK, G. R., DENNY, W. A. E WARE, D. C., Dalton Trans., v. 39, p. 11535–11550, 2010. [49] BUSTAMANTE, F. L. S., MIRANDA, F. S., CASTRO, F. A.V., RESENDE, J. A. L. C., PEREIRA, M. D., LANZNASTER, M., A study on the properties and reactivity of naphthoquinone–cobalt(III) prototypes for bioreductive prodrugs. J. Inorg. Biochem., v. 132, p. 37–44, 2014. [50] FYLAKTAKIDOU, K. C.; HADJIPAVLOU-LITINA, D. J.; LITINAS, K. E.; NICOLAIDES, D. N., Natural and synthetic coumarin derivatives with anti-inflammatory/ antioxidant activities. Curr. Pharm., v. 10, p. 3813-3833, 2004. [51] BANSAL, Y.; SETHI, P.; BANSAL, G. Coumarin: a potential nucleus for anti-inflammatory molecules. Med. Chem. Res., v. 22, p. 3049–3060, 2013. [52] KASHMAN, Y.; GUSTAFSON, K. R.; FULLER, R. W.; CARDELLINA, J. H.; MCMAHON, J. B.; CURRENS, M. J.; BUCKHEIT, R. W.; HUGHES, S. H.; CRAGG, G. M.; BOYD, M. R. J., HIV inhibitory natural products. Part 7. The calanolides, a novel HIV-inhibitory class of coumarin derivatives from the tropical rainforest tree, Calophyllum lanigerum. Med. Chem., v. 35, p. 2735–2743, 1992. [53] SHIKISHIMA, Y.; TAKAISHI, Y.; HONDA, G.; ITO, M.; TAKFDA, Y.; KODZHIMATOV, O. K.; ASHURMETOV, O.; LEE, K. H., Chemical constituents of Prangos tschiniganica; structure elucidation and absolute configuration of coumarin and 101 furanocoumarin derivatives with anti-HIV activity, Chem. Pharm. Bull., v. 49, p. 877-880, 2001. [54] OSTROV, D. A.; HERNANDEZ PRADA, J. A.; CORSINO, P. E.; FINTON, K. A.; LE, N.; ROWE, T.C., Discovery of novel DNA gyrase inhibitors by high-throughput virtual screening. Antimicrob. Agents Chemother., v. 51, p. 3688-3698, 2007. [55] GORMLEY, N. A.; ORPHANIDES, G.; MEYER, A.; CULLIS, P. M.; MAXWELL, A., The Interaction of Coumarin Antibiotics with Fragments of the DNA Gyrase B Protein. Biochemistry, v. 35, p. 5083-5092, 1996. [56] CURINI, M.; EPIFANO, F.; MALTESE, F.; MARCOTULLIO, M. C.; GONZALES, S. P.; RODRIGUEZ, J. C., Synthesis of Collinin, an Antiviral Coumarin, Aust. J. Chem., v. 56, p. 58-61, 2003. [57] ANAND, P.; SINGH, B.; SINGH, N., A review on coumarins as acetylcholinesterase inhibitors for Alzheimer’s disease. Bioorg. Med. Chem., v. 20, p. 1175-1180, 2012. [58] PIAZZI, L.; CAVALLI, A.; COLIZZI, F.; BELLUTI, F.; BARTOLINI, M.; MANCINNI, F.; RECANATINI, M.; ANDRISANA, V.; RAMPA, A., Multi-target-directed coumarin derivatives: hAChE and BACE1 inhibitors as potential anti-Alzheimer compounds, Bioorg. Med. Chem. Lett., v. 18, p. 423-426, 2008. [59] RIVEIRO, M. E.; MOGLIONI, A.; VAZQUEZ, R.; GOMEZ, N.; FACORRO, G.; PIEHL, L.; DE CELIS, E. R.; SHAYO, C.; DAVIO, C., Structural insights into hydroxycoumarin-induced apoptosis in U-937 cells. Bioorg. Med. Chem., v. 16, p. 2665-2675, 2008. [60] THAKUR, A.; SINGLA, R.; JAITAK, V., Coumarins as anticancer agents: A review on synthetic strategies, mechanism of action and SAR studies. Eur. J. Med. Chem., v. 101, p. 476-495, 2015. [61] SARKER, S. D.; NAHAR, L., Progress in the Chemistry of Naturally Occurring Coumarins. Prog. Chem. Org. Nat. Prod., v. 106, p. 241-304, 2017. [62] MUSA, M. A.; COOPERWOOD, JOHN S.; KHAN, M. O. F., A Review of Coumarin Derivatives in Pharmacotherapy of Breast Cancer, Curr. Med. Chem., v. 15, p. 2664–2679, 2008. 102 [63] Food and Drugs Administration. Disponível em: https://www.accessdata.fda.gov/scripts/cder/daf/index.cfm?event=BasicSearch.process. Acessado em: 04/01/2018. [64] Agência Nacional de Vigilância Sanitária. Disponível em: https://consultas.anvisa.gov.br/#/medicamentos/q/?substancia=22353. Acessado em: 04/01/2018. [65] Bulário eletrônico da Agência Nacional de Vigilância Sanitária. Disponível em: http://www.anvisa.gov.br/datavisa/fila_bula/index.asp. Acessado em: 04/01/2018. [66] SANDHU, S.; BANSAL, Y.; SILAKARI, O.; BANSAL, G., Coumarin Hybrids as Novel Therapeutic Agents, Bioog. Med. Chem., v. 22, p. 3806-3814, 2014. [67] MARCU, M. G.; SCHULTE, T. W.; NECKERS, L., Novobiocin and related coumarins and depletion of heat shock protein 90-dependent signaling proteins, J. Natl Cancer Inst., v. 92, p. 242-248, 2000. [68] LEWIS, R. J.; SINGH, O. M.; SMITH, C. V.; SKARZYNSKI, T.; MAXWELL, A.; WONACOTT, A. J.; WIGLEY, D. B., The nature of inhibition of DNA gyrase by the coumarins and the cyclothialidines revealed by X-ray crystallography, EMBO J., v. 15, p. 1412-1420, 1996. [69] MARCU, M. G.; CHADLI, A.; BOUHOUCHE, I.; CATELLI, M.; NECKERS, L. M. J., The heat shock protein 90 antagonist novobiocin interacts with a previously unrecognized ATP-binding domain in the carboxyl terminus of the chaperone, Biol. Chem., v. 275, p. 37181-37186, 2000. [70] KUSUMA, B. R.; KHANDELWAL, A.; GUA, W.; BROWN, D.; LIU, W.; VIELHAUER, G.; HOLZBEIERLEIN, J.; BLAGG, B. S. J., Synthesis and biological evaluation of coumarin replacements of novobiocin as Hsp90 inhibitors, Bioorg. Med. Chem., v. 22, p. 1441–1449, 2014. [71] DEVJI, T.; REDDY, C.; WOO, C.; AWALE, S.; KADOTA, S.; CARRICO-MONIZ, D., Pancreatic anticancer activity of a novel geranylgeranylated coumarin derivative, Bioorg. Med. Chem. Lett., v. 21, p. 5770-5773, 2011. [72] EMAMI, S.; DADASHPOU, S. CURRENT developments of coumarin-based anti-cancer agents in medicinal chemistry. Eur. J. Med. Chem., v. 102, p. 611-630, 2015. 103 [73] REHMAN, S. U.; CHOHAN, Z. H.; GULNAZ, F.; SUPURAN, C. T., In-vitro antibacterial, antifungal and cytotoxic activities of some coumarins and their metal complexes, J. Enzym. Inhib. Med. Chem., v. 20, p. 333–340, 2005. [74] REFAT, M. S.; EL-DEEN, I. M.; ANWER, Z. M.; EL-GHOL, S., Bivalent Transition metal complexes of coumarin-3-yl thiosemicarbazone devivatives: Spectroscopic, antibacterial activity and thermogravimetric studies. J. Mol. Struc., v. 920, p. 149–162, 2009. [75] RAJ, K. M., MRUTHYUNJAYASWAMY, B. H. M., Synthesis, spectroscopic characterization, electrochemistry and biological activity evaluation of some metal (II) complexes with ONO donor ligands containing indole and coumarin moieties. J. Saudi Chem Soc., v. 21, p. S202–S218, 2017. [76] MAHMOOD, K., MUDDASSAR, M., TAHIR, M. N. E SHAFIQ, Z., Coumarin-based thiosemicarbazones as potent urease inhibitors: synthesis, solid state selfassembly and molecular docking. RSC Adv., v. 6, p. 63886-63894, 2016. [77] GUEOGJIAN, K. Funcionalização de Cumarinas Via Reação de Acoplamento de Suziki-Miyaura de Sais de Organotrifluoroboratos de Sódio, 2011, Faculdade de Ciências Farmacêuticas da Universidade de São Paulo, Dissertação de mestrado, USP, São Paulo, 2011. [78] SCHULMAN, S. G., Fluorescence and phosphorescence spectroscopy: physicochemical principles and practice, 1ª edição, University of Florida, Editora Pergamon Press, 1977. [79] JAMESON, D. M., Introduction to Fluorescence, Nova Iorque, Editora CRC Press, 2014. [80] FABBRIZZI, L., LICCHELLI, M., PALLAVICINI, P., SACCHI, D., TAGLIETTI, A., Sensing of transition metals through fluorescence quenching or enhancement. A review. Analyst., v. 12, p. 1763-1768, 1996. [81] HARIHARAN, C.; VIJAYSREE, V.; MISHRA, A.K., Quenching of 2,5-diphenyloxazole (PPO) fluorescence by metal ions, J. Lumin., v. 75, p. 205-211, 1997. [82] RAJ, K. M., MRUTHYUNJAYASWAMY, B. H. M., Synthesis, spectroscopic characterization, electrochemistry and biological activity evaluation of some metal (II) complexes with ONO donor ligands containing indole and coumarin moieties. J. Saudi Chem. Soc., v. 21, p. S202–S218, 2017. 104 [83] FAILES, T. W.; CULLINANE, C.; DIAKOS, C. I.; YAMAMOTO, N.; LYONS, J. G.; HAMBLEY, T. W., Studies of a Cobalt (III) Complex of the MMP Inhibitor Marimastat: A Potential Hypoxia-Activated Prodrug, Chem. Eur. J., v. 13, p. 2974–2982, 2007. [84] WHITTAKER, M.; FLOYD, C. D.; BROWN, P.; GEARING, A. J. H., Design and therapeutic application of matrix metalloproteinase inhibitors., Chem. Rev., v. 99, p. 2735-2776, 1999. [85] MICHAELIDES, M. R.; CURTIN, M. L., Recent advances in matrix metalloproteinase inhibitors research, Curr. Pharm. Des., v. 5, p. 787– 819, 1999. [86] SHINHABABU, A. K.; THAKKER, D. R., Prodrugs of anticancer agents, Adv. Drug Deliver. Rev., v. 19, p. 241–273, 1996. [87] MANAZANARO, S.; VINCENT, M. J.; MARTIN, M. J.; SALVADOR-TORMO, N.; PÉREZ, J. M.; BLANCO, M. DEL M.; AVENDAÑO, C.; MENENDEZ, J. C.; FUENTE, J. A., Synthesis and biological evaluation of new 1,5-diazaanthraquinones with cytotoxic activity. Bioorg. Med. Chem., v. 12, p. 6505–6515, 2004. [88] WHEATE, N. J.; BRODIE, C. R.; COLLINS, J. G.; KEMP, S.; WRIGHT, J. R. A., DNA Intercalators in Cancer Therapy: Organic and Inorganic Drugs and Their Spectroscopic Tools of Analysis, Mini-Rev. Med. Chem., v. 7, p. 627–648, 2007. [89] MUELLER, S. O.; LUTZ, W. K.; STOPPER, H., Factors affecting the genotoxic potency ranking of natural anthraquinones in mammalian cell culture systems, Mutat. Res, v. 414, p. 125– 129, 1998. [90] DOWNWARD, A. M.; MOORE, E. G.; HARTSHORN, R. M., Photoinduced ligand release in a ruthenium(II)-cobalt(III) heterodinuclear system, Chem. Commun., v. 47, p. 7692–7694, 2011. [91] BRADLEY, D.; WILLIAMS, G.; LAWTON, M., Drying of Organic Solvents: Quantitative Evaluation of the Efficiency of Several Desiccants. J. Org. Chem., v. 12, p. A–D, 2010. [92] HUREAU, C.; BLONDIN, G.; CHARLOT, M-F.; PHILOUZE, C.; NIERLICH, M.; CÉSARIO, M.; ANXOLABÉHÈRE-MALLART, E., Synthesis, Structure, and Characterization of New Mononuclear Mn(II) Complexes. Electrochemical Conversion into New Oxo-Bridged Mn2(III,IV) Complexes. Role of Chloride Ions. Inorg. Chem., v. 44, p. 3669−3683, 2005. 105 [93] BRUKER (2007). APEX2 v2014.5-0. Bruker AXS Inc., Madison, Wisconsin, USA. [94] BRUKER (2013). SAINT v8.34A. Bruker AXS Inc., Madison, Wisconsin, USA. [95] SHELDRICK, G. M., SADABS, Program for Empirical Absorption Correction of Area Detector Data, University of Göttingen, Germany, 1996. [96] SHELDRICK, G. M., Crystal structure refinement with SHELXL, Acta Crystallogr., Sect. C: Struct. Chem., v. 71, p. 3–8, 2015. [97] MACRAE, C. F., BRUNO, I. J., CHISHOLM, J. A., EDGINGTON, P. R, P. PIDCOCK M., RODRIGUEZ-MONGE, E., L., STREEK, R. T., J. WOOD, V. P. A., WOOD, Mercury CSD 2.0 – new features for the visualization and investigation of crystal structures, J. Appl. Cryst., v. 41, p. 466–470, 2008. [98] TURBOMOLE V6.6 2014, A development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989-2007, TURBOMOLE GmbH, since 2007; available from http://www.turbomole.com [99] FURCHE, F.; AHLRICHS, R., Adiabatic time-dependent density functional methods for excited state properties, J. Chem. Phys., v. 117, p. 7433–7447, 2002. [100] ADAMO, C.; BARONE, V., Toward reliable density functional methods without adjustable parameters: The PBE0 model, J. Chem. Phys., v. 110, p. 6158–6170, 1999. [101] WEIGEND, F.; AHLRICHS, R., Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy, Phys. Chem. Chem. Phys., v. 7, p. 3297–3305, 2005. [102] SOUZA, I. C. A.; FARO, L. V.; PIEPHIRO, C. B.; GONZAGA, D. T. G.; SILVA, F. C.; FERREIRA, V. F.; MIRANDA, F. S.; SCARPELLINIC, M.; LANZNASTER, M., Investigation of cobalt(III)-triazole systems as prototypes for hypoxia-activated drug delivery, Dalton Trans. Comm., 2016. DOI: 10.1039/c6dt02456d. [103] KLAMT, A.; SCHÜÜRMANN, G., COSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradient, J. Chem. Soc., Perkin Trans., v. 2, p. 799–805, 1993. [104] ChemCraft program. http://www.chemcraftprog.com. Accessado em: 07.11.2017. [105] GAGNE, R. R.; KOVAL, C. A.; LISENSKY, G. C., Ferrocene as an internal standard for electrochemical measurements, Inorg. Chem., v. 19, p. 2854–2855, 1980. [106] CHAVES, J. D. S.; TUNES, L. G.; FRANCO, C. H. J.; FRANCISCO, T. M.; CORRÊA, C. C.; MURTA, S. M. F.; MONTE-NETO, R. L.; SILVA, H.; FONTES, A. P. S.; 106 ALMEIDA, M. V., Novel gold(I) complexes with 5-phenyl-1,3,4-oxadiazole-2-thione and phosphine as potential anticancer and antileishmanial agentes, Eur. J. Med. Chem., v. 127, p. 727-739, 2017. [107] MOSMANN, T., Rapid Colorimetric Assay for Cellular Growth and Survival: Application to Proliferation and Cytotoxicity Assays, J. lmmunol. Methods, v. 65, p. 55-63, 1983. [108] LOHANNE, J., NEVES, A. P., iniciação científica, 2014. [109] AREAS, E. S., BRONSATO, B. J. S., PEREIRA, T. M., GUEDES, G. P., MIRANDA, F. DA S., KUMMERLE, A. E., CRUZ, A. G. B. E NEVES, A. P., Novel CoIII complexes containing fluorescent coumarin-N-acylhydrazone hybrid ligands: synthesis, crystal structures, solution studies and DFT calculations, Spectrochim. Acta A, v. 187, p. 130-142, 2017. [110] GEARY, W. J., The use of conductivity measurements in organic solvents for the characterization of coordination compounds, Coord. Chem. Rev., v. 7, p. 81–122, 1971. [111] OTTER, C. A.; HARTSHORN, R. M., Preparation and photochemistry of cobalt(III) amino and amino acidato complexes containing tripodal polypyridine ligands, Dalton Trans., p. 150–156, 2004. [112] CHAKRABORTY, B., PAINE, T. K., Synthesis and characterization of cobalt(II)– salicylate complexes derived from N4-donor ligands: Stabilization of a hexameric water cluster in the lattice host of a cobalt(III)–salicylate complex. Inorg. Chim. Acta, v. 378, p. 231–238, 2011. [113] MASSOUD, S. S., PERKINS, R. S., LOUKA, F. R., WU XU, ROUX, A. L., DUTERCQ, Q., FISCHER, R. C., MAUTNER, F. A., HANDA, M., HIRAOKA, Y., KREFT, G. L., BORTOLOTTOE, T. E TERENZI, H., Efficient hydrolytic cleavage of plasmid DNA by chloro-cobalt(II) complexes based on sterically hindered pyridyl tripod tetraamine ligands: synthesis, crystal structure and DNA cleavage. Dalton Trans., v. 43, p. 10086–10103, 2014. [114] AHUMADA G.; FUENTEALBA, M.; ROISNEL, T.; KAHLAL, S.; CÓRDOVA, R.; CARRILLO, D.; SAILLARD, J-Y.; HAMON, J-R.; MANZUR, C., Spectroscopic, structural, electrochemical and computational studies of some new 2-Thienyl-containing β-diketonate complexes of cobalt(II), nickel(II) and copper(II), J. Mol. Struc., v. 1150, p. 531-539, 2017. 107 [115] KULAI I.; MALLET-LADEIRA, S., Synthesis and molecular structure of 2-bromo-N-(4-7-(diethylamino)-coumarin-3-yl)phenyl)propenamide: New coumarin-based fluorescent ATRP initiator, J. Mol. Struc., v. 1104, p. 14-18, 2016. [116] YU, T.; MENG, J.; ZHANG, P.; ZHAO, Y.; ZHANG, H.; FAN, D.; CHEN, L.; QIU, Y., Synthesis, crystal structures and photoluminescence of 7-(N,N′-diethylamino)-3-phenylcoumarin derivatives, Spectrochim. Acta A, v. 75, p. 1036–1042, 2010. [117] JANIAK, C., A critical account on – stacking in metal complexes with aromatic nitrogen-containing ligands, J. Chem. Soc., Dalton Trans., p. 3885–3896, 2000. [118] ARCOS-RAMOS, R.; MALDONADO-DOMÍNGUEZ, M.; ORDÓÑEZ-HERNÁNDEZ, J.; ROMERO-ÁVILA, M.; FARFÁN, N.; CARREÓN-CASTRO, M. del P.; 3-Substituted-7-(diethylamino)coumarins as molecular scaffolds for the bottom-up self-assembly of solids with extensive -stacking, J. Mol. Struc., v. 1130, p. 914-921, 2017. [119] SHRINER, R. L.; HERMANN, C. K. F.; MORRILL, T. C.; CURTIN, D. Y.; FUSON, R. C., The Systematic Identification of Organic Compounds, 8ª edição, editora Wiley John Wiley & Sons. INC, 2004. [120] BASANAGOUDA, M., JAMBAGI, V. B., BARIGIDAD, N. N., LAXMESHWAR, S. S., NARAYANACHAR, V. D., Synthesis, structureeactivity relationship of iodinated-4aryloxymethyl-coumarins as potential anti-cancer and antimycobacterial agentes. Eur. J. Med. Chem., v. 74, p. 225-233, 2014. [121] DRAGO, R. S., Physical Methods for Chemists, 2ª edição, Universidade de Flórida, editora Surfside, 1997. [122] ZHANG, H.; YU, T.; ZHAO, Y.; FAN, D.; CHEN, L.; QIU, Y.; QIAN, L.; ZHANG, K.; YANG, C., Crystal structure and photoluminescence of 7-(N,N’-diethylamino)-coumarin-3-carboxylic acid, Spectrochim. Acta A, v. 69, p. 1136–1139, 2008. [123] YU, T.; ZHANG, P.; ZHAO, Y.; ZHANG, H.; MENG, J.; FAN, D., Synthesis, characterization and high-efficiency blue electroluminescence based on coumarin derivatives of 7-diethylamino-coumarin-3-carboxamide, Organic Electron., v. 10, p. 653–660, 2009. [124] VALLE-BOURROUET, G.; UGALDE-SALDÍVAR, V. M.; GÓMEZ, M. ORTIZ-FRADE, L. A.; GONZÁLEZ, I.; FRONTANA, C., Magnetic interactions as a stabilizing factor of semiquinone species of lawsone by metal complexation, Electrochim. Acta, v. 55, p. 9042–9050, 2010. 108 [125] MARTINOVIC, J.; WYK, J. V.; MAPOLIE, S.; JAHEDA, N.; BAKER, P.; IWUOHA, E., Electrochemical and spectroscopic properties of dendritic cobalto-salicylaldiimine DNA biosensor, Electrochim. Acta, v. 55, p. 4296–4302, 2010. [126] GOTTLIEB, H. E.; KOTLYAR, V.; NUDELMAN, A., NMR Chemical Shifts of Common Laboratory Solvents as Trace Impurities, J. Org. Chem., v. 62, p. 7512-7515, 1997. [127] SOUZA, E. T.; CASTRO, L. C.; CASTRO, F. A. V.; VISENTIN, L. C.; PIEPHIRO, C. B.; PEREIRA, M. D.; MACHADO, S. P.; SCARPELLINI, M., Synthesis, characterization and biological activities of mononuclear Co(III) complexes as potential bioreductively activated prodrugs, J. Inorg. Biochem., v. 103, p. 1355–1365, 2009. [128] FIELD, L. D.; STEREPHLL, S.; KALMAN, J. R., Organic Structures from Spectra, 4ª edção, Editora Wiley, 2008. [129] PEREIRA, T. M., VITÓRIO, F., AMARAL, R. C., ZANONI, K. P. S., IHA, N. Y. M. E KUMMERLE, A. E., Microwave-assisted synthesis and photophysical studies of novel fluorescent N-acylhydrazone and semicarbazone-7-OH-coumarin dyes. J. Chem., v. 40, p. 8846- 8854, 2016. [130] MIRANDA, F. S.; RONCONI, C. M.; SOUSA, M. O. B.; SILVEIRA, G. Q.; VARGAS, M. D., 6-Aminocoumarin-naphthoquinone conjugates: design, synthesis, photophysical and electrochemical properties and DFT calculations. J. Braz. Chem. Soc., v. 25, 2014. [131] ZHANG, L.; XIA, Y.; LI, M.; LI, D.; HOU, R., Synthesis, photophysical and electrochemical properties of symmetric silicon-linked coumarin-oxadiazole derivatives, Tetrahedron, v. 72, p. 7438-7442, 2016. [132] MIYANO, D. M.; LIMA, T.; SIMÕES, F. R.; LA-SCALEA, M. A.; OLIVEIRA, H. P. M.; CODOGNOTO, L., Electrochemical Study of Simple Coumarin and its Determination in Aqueous Infusion of Mikania glomerate. J. Braz. Chem. Soc., v 25, p. 602-609, 2014. [133] NUÑEZ-VERGARA, L. J., PARDO-JIMÉNEZ, V., BARRIENTOS, C., OLEA-AZAR, C. A., ENCINA, P. A. N., SQUELLA, J. A., Dihydropyridine-fused and pyridine-fused coumarins: Reduction on a glassy carbon electrode in dimethylformamide, Electrochim. Acta, v. 85, p. 336–344, 2012. [134] KUHN, A.; ESCHWEGE, K. G. V.; CONRADIE, J., Electrochemical and density functional theory modeled reduction of enolized 1,3-diketones, Electrochim. Acta, v. 56, p. 6211–6218, 2011. 109 [135] STANIC, Z.; VOULGAROPOULOS, A.; GIROUSIA, S., Electroanalytical Study of the Antioxidant and Antitumor Agent Curcumin. Electroanalysis, v. 20, p. 1263–1266, 2008. [136] AHMADI, F.; ALIZADEHA, A.A.; SHAHABADI, N.; RAHIMI-NASRABADI, M., Study binding of Al–curcumin complex to ds-DNA, monitoring by multispectroscopic and voltammetric techniques. Spectrochim. Acta A, v. 79, p. 1466–1474, 2011. [137] HUSSAIN, A.; LAHIRI, D.; BEGUM, M. S. A.; SAHA, S.; MAJUMDAR, R.; DIGHE, R. R.; CHAKRAVARTY, A. R., Photocytotoxic Lanthanum(III) and Gadolinium(III) Complexes of Phenanthroline Bases Showing Light-Induced DNA Cleavage Activity. Inorg. Chem., v. 49, p. 4036–4045, 2010. [138] CHANA, S. L-F.; LAMA, T. L.; YANG, C.; LAIA, J.; CAO, B.; ZHOU, Z.; ZHU, Q., Cobalt(II) tris(2-pyridylmethyl)amine complexes [Co(TPA)X]+ bearing coordinating anion (X = Cl, Br,I and NCS): synthesis and application for carbon dioxide reduction, Polyhedron, v. 125, p. 156–163, 2017. [139] NARULKARA, D. D.; SRIVASTAVAB, A. K.; BUTCHERC, R. J.; ANSYD, K. M.; DHURI, S. N., Synthesis and characterization of N3Py2 ligand-based cobalt(II), nickel (II) and copper(II) catalysts for efficient conversion of hydrocarbons to alcohols, Inorg. Chim. Acta, v. 467, p. 405–414, 2017. [140] OURARI, A.; MESSALI, S.; BOUZERAFA, B.; OUENNOUGHI, Y.; AGGOUN, D.; MUBARAK, M. S.; STRAWSINE, L. M.; PETERS, D. G., Synthesis, characterization, and electrochemical behavior of a cobalt(II) salen-like complex, Polyhedron, v. 97, p. 197–201, 2015. [141] SHAVALEEV, N. M.; KESSLER, F.; GRÄTZEL, M.; NAZEERUDDIN, M. K., Redox properties of cobalt(II) complexes with azole-pyridines, Inorg. Chim. Acta, v. 407, p. 261–268, 2013. [142] RIGA, B. A.; NEVES, M. D.; MACHADO, A. E. H.; ARAÚJO, D. M. S.; SOUZA, J. R.; NASCIMENTO, O. R.; SANTANA, V. T.; CAVALHEIRO, C. C. S.; CARVALHO-JR, V. P.; GOI, B. E., Synthesis of cobalt(II)-a-diimines complexes and their activity as mediators in organometallic mediated radical polymerization of vinyl acetate, Inorg. Chim. Acta, v. 471, p. 620–629, 2018. [143] R. A. KRAUSE AND K. KRAUSE, Chemistry of bipyridyllike ligands. 2. Mixed complexes of ruthenium(II) with 2-(phenylazo)pyridine: a new .pi.-bonding probe, Inorg. Chem., v. 21, p. 1714–1720, 1982. 110 [144] JOHNSON, C. R.; SHEPHERD, R. E., Synth. React. Inorg. Met.-Org. Chem., v. 14, p. 339–353, 1984. [145] MEITES, L., Handbook of Analytical Chemistry; editora McGraw Hill, New York, 1963.por
dc.subject.cnpqQuímicapor
dc.thumbnail.urlhttps://tede.ufrrj.br/retrieve/65362/2018%20-%20Esther%20Saraiva%20Areas.pdf.jpg*
dc.originais.urihttps://tede.ufrrj.br/jspui/handle/jspui/4713
dc.originais.provenanceSubmitted by Sandra Pereira (srpereira@ufrrj.br) on 2021-05-30T03:16:49Z No. of bitstreams: 1 2018 - Esther Saraiva Areas.pdf: 3476100 bytes, checksum: 355155d6cdd5cfd22596f5162099ebcd (MD5)eng
dc.originais.provenanceMade available in DSpace on 2021-05-30T03:16:49Z (GMT). No. of bitstreams: 1 2018 - Esther Saraiva Areas.pdf: 3476100 bytes, checksum: 355155d6cdd5cfd22596f5162099ebcd (MD5) Previous issue date: 2018-02-01eng
Appears in Collections:Mestrado em Química

Se for cadastrado no RIMA, poderá receber informações por email.
Se ainda não tem uma conta, cadastre-se aqui!

Files in This Item:
File Description SizeFormat 
2018 - Esther Saraiva Areas.pdf2018 - Esther Saraiva Areas3.39 MBAdobe PDFThumbnail
View/Open


Items in DSpace are protected by copyright, with all rights reserved, unless otherwise indicated.