Facile Generation of A2B Corrole Radical Using Fe ... - ACS Publications

Mar 16, 2017 - in its monomeric form. On the other hand, Anand et al. have synthesized a stable and .... pentathiophene radical reported by Anand et a...
0 downloads 8 Views 2MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article http://pubs.acs.org/journal/acsodf

Facile Generation of A2B Corrole Radical Using Fe(III) Salts and Its Spectroscopic Properties Pinky Yadav, Pinki Rathi, and Muniappan Sankar* Department of Chemistry, Indian Institute of Technology Roorkee, Roorkee 247667, India S Supporting Information *

ABSTRACT: A carboxyphenyl-substituted corrole, 5,15dimesityl-10-(4′-carboxyphenyl)corrole (1), has been synthesized and characterized by UV−vis, fluorescence, 1H NMR spectroscopy, and electrospray ionization (ESI)-mass spectrometry (MS) techniques. An air-stable corrole radical (1•) was obtained with the addition of the Fe(III) salt to 1 in dimethyl sulfoxide (DMSO) and characterized by UV−vis, fluorescence, electron paramagnetic resonance (EPR), ESI-MS techniques, and density functional theory studies. The neutral corrole radical (1•) exhibited a sharp EPR signal at g = 2.006 in DMSO. The reduced bipyrrolic (N−C−C−N) dihedral angle (χ) of 1 from 19.11 to 7.07° leads to the release of angle strain, which is the driving force for the generation of 1•. Notably, transdimesityl groups prevent the dimerization or aggregation of the corrole radical. Further, 1• was converted to 1 by excess addition of Fe(II) salts in DMSO at 298 K.



of (Cor)H4+ and (Cor)H2− produces a neutral corrole, (Cor)H3, by an acid−base proton-transfer mechanism, as shown in the above scheme. Recently, the 3,17-dichloro-mesotrimesitylcorrole radical was reported by Bröring et al., which was further used to prepare the first-ever Zn(II) corrole radical.9b The mesityl groups prevent the aggregation or dimerization of radical species and stabilize the radical corrole in its monomeric form. On the other hand, Anand et al. have synthesized a stable and neutral 25π-pentathiophene radical species, which readily converts into aromatic and antiaromatic electronic circuits on adding appropriate chemical agents.9c Kobayashi et al. have reported an extremely air-stable 19πelectron azaporphyrin radical, which was prepared from the reduction of a cationic P(V) azaporphyrin.9d The central phosphorus(V) atom and the peripheral bulky tert-butyl groups stabilize the reduced state. Herein, we report the generation of stable corrole radical species (1•) of 5,15-dimesityl-10-(4′-carboxyphenyl)corrole (1) as well as 5,15-dimesityl-10-(4′-methoxycarbonylphenyl)corrole (MEC) by a chemical oxidation process in quantitative yield using FeIII salts in dimethyl sulfoxide (DMSO) at 298 K. The corrole radical was characterized by UV−vis, fluorescence, electrospray ionization (ESI)-mass spectrometry (MS), electron paramagnetic resonance (EPR) spectroscopic techniques and density functional theory (DFT) studies. Notably, 1 was regenerated by the excess addition of FeII salt to 1• in DMSO at 298 K.

INTRODUCTION In recent years, studies on tetrapyrrolic macrocycles have been given ardent interest to explore their structural, photophysical, and photochemical properties. Corrole is a structural counterpart of porphyrin, and it emanates from the corrin ring. Recently, corroles have emerged as an independent area of research1−4 due to their useful applications in catalysis, sensors, molecular electronics, solar cells, nonlinear optics, and photodynamic therapy.5,6 A number of free base corroles and their metal complexes have been synthesized and characterized over the last two decades after the invention of facile synthetic methods. Notably, β-functionalization such as halogenation, nitration, chlorosulfonation, and hydroformylation of the corrole periphery leads to the formation of novel and more sophisticated corroles.7 Different strategies are available to prepare symmetrical and unsymmetrical corroles in good yields.8

Redox chemistry of electron-rich corroles is of great interest due to the formation of radical and cationic/anionic corrole species. Recently, Kadish et al. have generalized the redox mechanism of meso-substituted free base corroles using cyclic voltammetry and spectroelectrochemical studies.9a For example, the oxidation of free base trimesitylcorrole (Cor)H3 yielded [(Cor)H3]•+, which is a strong acid and rapidly transforms into (•Cor)H2 and H+. Further, H+ reacts with another free base corrole and produces [(Cor)H4]+ ions. The electroreduction of (•Cor)H2 gives (Cor)H2−. The mixing © 2017 American Chemical Society

Received: November 27, 2016 Accepted: March 1, 2017 Published: March 16, 2017 959

DOI: 10.1021/acsomega.6b00430 ACS Omega 2017, 2, 959−965

ACS Omega

Article

Scheme 1. Synthetic Route for 5,15-Dimesityl-10-(4′-carboxyphenyl)corrole (1)



RESULTS AND DISCUSSION Synthesis and Characterization. The synthesis of 1 was carried out in two steps; the first step involves the synthesis of MEC by acid (HCl)-catalyzed condensation of 5-mesityldipyrromethane and methyl 4-formylbenzoate in an H2O/MeOH (1:2, v/v) mixture,8a followed by alkaline hydrolysis of MEC in 95% ethanol to afford 1 (45% yield) in the second step. The synthesized free base corrole, 1, was characterized by UV−vis, fluorescence, and NMR spectroscopic techniques and MS (Scheme 1).8b The free base corrole (1) exhibited a Soret band at 410 nm with a shoulder at 417 nm and three Q bands in the visible region located at 567, 602, and 638 nm in CH2Cl2 as shown in Figure S1.8b Figure S2 represents the 1H NMR spectra of MEC in CDCl3. Figures S3 and S4 illustrate the 1H NMR spectra of 1 in CDCl3 and DMSO-d6, respectively. The ESI mass spectrum of 1 in CH3CN is shown in Figure S5. To probe the influence of the carboxyphenyl group at the meso-position of the corrole ring, we carried out the cyclic voltammetric studies of 1 using a saturated calomel electrode as the reference electrode in CH2Cl2 at 298 K. 1 exhibited one-electron reversible oxidation and reduction as illustrated in Figure S6. The first oxidation potential of 1 was observed at 0.73 V, while the reduction potential was observed at −0.78 V at 298 K. Spectroscopic Studies. We used electronic absorption spectral studies to demonstrate the formation of 1•, which is produced from the A2B corrole (1) with the addition of 1 equiv of Fe(III) salt in DMSO at 298 K, as shown in Figure 1. Further, we have tested with MEC and 5,10,15-tritolylcorrole (TTC) by the addition of 1 equiv of Fe(III) salt in DMSO at

298 K. The spectral changes of MEC are comparable to those of 1 while adding Fe(III) salt with a prominent peak at 390 nm (Figure S8a), indicating the formation of a radical corrole, which is further confirmed by electron spin resonance studies (vide infra). However, TTC exhibited only marginal spectral changes, with a shoulder at 390 nm (Figure S8b), indicating the partial formation of a radical corrole. Herein, we have a brief discussion about the interaction of corrole 1 with FeIII ions in DMSO at 298 K. The observed UV−vis spectral pattern of 1• matches with that of the reported corrole radical by Kadish et al.9a and Bröring et al.,9b whereas no changes were observed with other metal ions (Figure S7). The appropriate redox reagents such as a one-electron oxidizing agent [Et3O+SbCl6−] and a one-electron reducing agent (KO2) were used to confirm the radical nature of 1. The addition of [Et3O+SbCl6−] to 1 in DMSO leads to the similar color and UV−vis spectral changes as FeIII (Figure S9a). The anionic species of corrole 1 was obtained with the successive addition of FeIII and KO2 (oneelectron reducing agent) in DMSO, as shown in Figure S9b. We have also tested the protonation version of 1 with TFA in DMSO, as shown in Figure S10. In general, protonated corroles, [CorH4]+, show a characteristic intense band between 632 and 694 nm. The lack of spectral signatures of the protonated corrole after addition of the Fe(III) salt indicates that the protonated corrole is not formed during the chemical oxidation. The spectral features of 1• are similar to those of the pentathiophene radical reported by Anand et al.9c Hence, the conversion of 1 to the neutral corrole radical (1•) may follow a radical mechanism (homolytic cleavage of N−H) as reported by Bröring et al.9b The driving force for the forward reaction in Scheme 2 is possibly due to the redox reaction between the electronically rich corrole (1) (having two mesityl groups) and the FeIII ion, which behaves as an oxidizing agent. The formed radical species is stabilized in the presence of electron-withdrawing carboxylic acid and carbomethoxy groups (-R effect) and a high polarity of DMSO. The ESI mass spectrum of 1• was found to be 1 mass unit less than the exact mass of 1, which confirms the formation of radical species (Figure S11). The simulated isotopic pattern of the corrole radical is matching with the experimentally observed one (Figure S11). The UV−vis spectrum of 1 exhibited a Soret band at 414 nm with a shoulder at 431 nm and Q bands at 574, 604, and 638 nm. The UV−vis spectral titration of 1 with Fe(III) perchlorate is depicted in Figure 1. An incremental addition of FeIII ions into the solution of 1 leads to a concomitant increase in the absorbance at 394 nm with a decrease in absorbance of Soret (414, 431sh) and Q bands (574, 604, and 638 nm) accompanied by a color change from purple to pale brown.

Figure 1. UV−vis spectral titration of 1 (18 μM) with FeIII (1.6−20 μM) in DMSO at 298 K. 960

DOI: 10.1021/acsomega.6b00430 ACS Omega 2017, 2, 959−965

ACS Omega

Article

Scheme 2. Schematic Representation of the Formation of 1• in the Presence of FeIII or [Et3O+SbCl6−] and Its Reversibility While Adding Excess of FeII in DMSO

Two isosbestic points were observed at 404 and 516 nm upon addition of FeIII ions into corrole 1, as shown in Figure 1. Fluorescence quenching was observed at 658 nm (excitation at 574 nm) while addition of FeIII ions to 1, as shown in Figure 2.

Figure 3. Electronic absorption spectral changes of 1 while adding FeCl3 and excess of FeCl2 in DMSO at 298 K.

EPR Spectral Studies of the Corrole Radical (1•). The formation of the corrole radical (1•) was characterized by EPR spectroscopy. We have carried out EPR experiments using the X-band EPR spectrometer in DMSO at 298 K as well as at lower temperatures. The EPR spectrum of 1 (1.5 mmol) in the presence of FeIII at different temperatures showed a sharp signal at g = 2.006 (∼ organic radical species), which corresponds to 1•, as depicted in Figures 4a and S14b. Similar results were obtained in the solidstate EPR spectrum of 1 + FeIII at 100 K, as shown in Figure 4b. The EPR spectrum of Fe(ClO4)3 in the solid state at 100 K (Figure S15) is quite different from that of 1•. These results clearly indicate the existence of a paramagnetic species, 1•. Similar EPR spectral changes were observed with the addition

Figure 2. Fluorescence spectral titration of 1 (18 μM) with FeIII (1.6− 20 μM) in DMSO at 298 K.

The fluorescence of 1 is completely quenched with 1 equiv of Fe(III) salt in DMSO at 298 K. The 1:1 stoichiometry of 1 with FeIII was calculated by a Job plot, as shown in Figure S12. The addition of excess FeCl2 or Fe(ClO4)2 to 1• in DMSO leads to a reverse absorption profile and also a color change from pale brown to purple due to one-electron reduction, which shows reversible redox behavior of 1, as depicted in Scheme 2 and Figures 3 and S13. 961

DOI: 10.1021/acsomega.6b00430 ACS Omega 2017, 2, 959−965

ACS Omega

Article

Figure 4. EPR spectra of 1• (a) in DMSO at 298 K and (b) in the solid state at 100 K.

of FeIII to MEC at variable temperatures under similar experimental conditions (Figure S16). Logic Gate Application. The convenient use of logic gates fascinated the attention of researchers in the construction of molecular switches and molecular keypad devices. Figure 5

dihedral angle (χ) leads to the release of strain present in the smaller corrole core due to the presence of three NH. The decrement in χ is the driving force for the generation of corrole radical 1•. Further, time-dependent (TD)-DFT calculations were performed to examine the absorption properties of 1•. The simulated electronic absorption spectral features using TDDFT studies matched with the experimentally observed absorption profile (Figure S17).



CONCLUSIONS

We have synthesized a carboxyphenyl-substituted corrole (1) and characterized by various spectroscopic techniques. 1 leads to a stable neutral radical species, 1•, in the presence of FeIII in DMSO as well as in the solid state. The observed UV−vis spectral features of 1• are in accordance with the reported literature. Further, complete fluorescence quenching of 1 was observed with 1 equiv of FeIII, and 1:1 stoichiometry was confirmed by the Job plot. The formation of radical species was supported by fluorescence quenching, decrement in lifetime, ESI-MS, and by using appropriate one-electron redox reagents. Further, 1• was reconverted to its precursor 1 by the addition of excess Fe(II) salt and to the corresponding protonated corrole, [(Cor)H4]+, by the addition of a trace of TFA in DMSO. 1• shows a sharp EPR signal at g = 2.006 in the presence of Fe(III) in DMSO as well as in the solid state. The angle strain in 1• reduced, and DFT studies suggest that the reduced bipyrrolic (N−C−C−N) dihedral angle (χ) of 1 from 19.11 to 7.07° in 1• is the driving force for the generation of the neutral corrole radical. The simulated absorption spectral features using TDDFT calculations match with the experimental one. Overall, the spectroscopic and theoretical studies suggest the formation of 1• while adding Fe(III) salt to 1 in DMSO.

Figure 5. Logic functions and truth table based on UV−vis spectral changes of 1 in the presence of FeIII and FeII ions in DMSO.

shows the truth table and the logic gate operation using UV− vis spectral changes of 1 at a Q-band of 574 nm (a peak at 574 nm: output = 1 and absence of a peak at 574 nm: output = 0) in the presence or absence of FeIII and FeII. We have used four input combinations: (0, 0) (0, 1) (1, 0), and (1, 1). An output signal was observed at the Q-band at 574 nm in the presence or absence of FeIII and FeII (input FeIII and FeII = 0 or 1). The absorbance at 574 nm was diminished (output = 0) in the presence of FeIII (input FeIII = 1 and FeII = 0), whereas its intensity was regained in the presence of excess Fe(II) salt in DMSO at 298 K. DFT Calculations. We optimized the geometry of 1 and 1• using B3LYP functional and the 6-31G basis set in the gas phase. The bond angle (red) and bond lengths (black) in the optimized structure of 1 and 1• are shown in Figure 6. One pyrrole ring of 1 is deviated from the 23-atom core, whereas the same attains planarity in the case of 1• with the removal of one proton from the corrole core. In addition, the bipyrrolic (N− C−C−N) dihedral angle (χ) in 1 was found to be 19.11°, which is reduced to 7.07° in 1•. The reduced N−C−C−N



EXPERIMENTAL SECTION Chemicals. Methyl 4-formylbenzoate and pyrrole were purchased from Alfa Aesar and used as received. CH2Cl2 was dried and distilled over P2O5. Iron perchlorate salts (M(ClO4)n· xH2O, M = FeII and FeIII) were used as received from Alfa Aesar. 5-Mesityldipyrromethane was synthesized using the literature method.8a Instrumentation. UV−vis and fluorescence spectra were recorded using a Cary 100 spectrometer and Hitachi F-4600 spectrofluorometer, respectively. All 1H NMR measurements were performed using a Bruker AVANCE 500 MHz 962

DOI: 10.1021/acsomega.6b00430 ACS Omega 2017, 2, 959−965

ACS Omega

Article

Figure 6. Optimized geometries of 1 and 1• using B3LYP functional and the 6-31G basis set in the gas phase. The bond angles and bond lengths are represented in red and black colors, respectively.

H, 6.03; N, 8.38; O, 4.78%. Found: C, 80.53; H, 6.34; N, 8.62; O, 4.51%. Synthesis of 10-(4-Carboxyphenyl)-5,15-dimesitylcorrole (1). In 30 mL of 95% ethanol, 30 mg of MEC was dissolved. To this, 0.075 g of NaOH in 1 mL of water was added and stirred at 70 °C for 2 h. Then, the reaction mixture was neutralized with 2N HCl, extracted with CHCl3, washed with water, and dried over Na2SO4. The crude product was recrystallized from MeOH/CHCl3 (2:98, v/v) and dried under vacuum. The yield was found to be 45% (0.012 g, 18.3 μmol). The observed spectral data of 1 are consistent with the literature values.8b UV−vis (CH2Cl2) λmax (nm): 410, 417 (sh), 567, 602, 638; 1 H NMR in CDCl3 (500 MHz): δ (ppm) 9.90 (s, 1H, COOH), 8.89 (d, 3JHH = 4.0 Hz, 2H, β-pyrrole-H), 8.46 (t, 4H, 3JHH = 8 Hz, carboxyphenyl-H and 3JHH = 4 Hz, β-pyrrole-H), 8.33 (d, 3 JHH = 4 Hz, 2H, β-pyrrole-H), 8.29 (d, 3JHH = 8 Hz, 2H, carboxyphenyl-H), 7.27 (s, 4H, m-mesityl-H), 2.60 (s, 6H, pmesityl-CH3-H), 1.93 (s, 12H, o-mesityl-CH3); ESI-MS found 655.30 [M + H]+ calcd 655.31. Anal. Calcd for C44H38N4O2: C, 80.71; H, 5.85; N, 8.56; O, 4.89%. Found: C, 80.31; H, 5.95; N, 8.25; O, 4.69%. Synthesis of 10-(4-Carboxyphenyl)-5,15-dimesitylcorrole Radical (1•). 1• was generated by mixing equimolar solutions of free base corrole (1) and Fe(III) salt in DMSO. An immediate color change was observed from purple to pale brown. EPR (DMSO, 9.5 GHz, giso = 2.006); UV−vis (DMSO) λmax (nm): 393 (sh), 417, 549, 574, 603; ESI-MS found 653.289 [M•] calcd 653.292.

spectrometer in CDCl3 and DMSO-d6.The ESI mass spectra were recorded using Bruker Daltanics-microTOF in the positive ion mode in CH3CN. EPR spectra were recorded using a Bruker X-band EPR instrument in DMSO. Electrochemical measurements were carried out with a CHI 620E electrochemical workstation. A three-electrode system was used that consisted of a GC working electrode, an Ag/AgCl reference electrode, and a Pt-wire counter electrode. The concentrations of all of the corroles used were approximately 1 mM. UV−vis and fluorescence titrations of 1 with FeIII ions were carried out in DMSO. Fluorescence lifetime measurements in the nanosecond time domain were recorded in DMSO using a Horiba Jobin Yvon “fluorocube fluorescence lifetime system” equipped with a NanoLED (635 nm) source. Theoretical calculations were carried out using B3LYP functional and the 6-31G basis set in the gas phase. Synthesis of MEC. 5-Mesityldipyrromethane (1 mmol, 264 mg) and methyl 4-formylbenzoate (0.5 mmol, 82 mg) were dissolved in MeOH (100 mL). To this, HClaq (36%, 5 mL) and H2O (50 mL) were added and stirred for 2 h at room temperature. Then, the mixture was extracted with CHCl3, and the organic layer was washed twice with water, dried over anhydrous Na2SO4, filtered, and diluted with 250 mL of CHCl3. p-Chloranil (369 mg, 1.5 mmol) was added and stirred overnight at room temperature. The reaction mixture was concentrated and purified on a silica column using CHCl3 as the eluent. Yield was found to be 8% (0.030 g, 0.04 mmol). UV−vis (CH2Cl2) λmax (nm): 408, 426 (sh), 568, 602, 639; 1 H NMR in CDCl3 (500 MHz): δ (ppm) 8.88 (d, 3JHH = 4.0 Hz, 2H, β-pyrrole-H), 8.50 (d, 3JHH = 4.5 Hz, 2H, β-pyrrole-H), 8.44 (d, 3JHH = 4.5 Hz, 2H, β-pyrrole-H), 8.39 (d, 3JHH = 8 Hz, 2H, esterphenyl-H), 8.33 (d, 3JHH = 4.5 Hz, 2H, β-pyrrole-H), 8.25 (d, 3JHH = 8.0 Hz, 2H, esterphenyl-H), 6.27 (s, 4H, mmesityl-H), 4.08 (s, 3H, COOCH3), 2.60 (s, 6H, p-mesitylCH3-H), 1.92 (s, 12H, o-mesityl-CH3); ESI-MS found 669.46 [M + H]+ calcd 669.32; Anal. Calcd for C45H40N4O2: C, 80.81;



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.6b00430. UV−vis, 1H NMR, EPR, and ESI mass spectra of MEC, 1, and 1• as well as CV and DPV of 1 (PDF) 963

DOI: 10.1021/acsomega.6b00430 ACS Omega 2017, 2, 959−965

ACS Omega



Article

10469−10478. (c) Ghosh, A. A Perspective of One-Pot Pyrrole− Aldehyde Condensations as Versatile Self-Assembly Processes. Angew. Chem., Int. Ed. 2004, 43, 1918−1931. (d) Steene, E.; Dey, A.; Ghosh, A. β-Octafluorocorroles. J. Am. Chem. Soc. 2003, 125, 16300−16309. (e) Bursa, B.; Barszcz, B.; Bednarski, W.; Lewtak, J. P.; Koszelewski, D.; Vakuliuk, O.; Gryko, D. T.; Wrobel, D. New meso-substituted corroles possessing pentafluorophenyl groupssynthesis and spectroscopic characterization. Phys. Chem. Chem. Phys. 2015, 17, 7411−7423. (f) Buckley, H. L.; Rubin, L. K.; Chrominski, M.; McNicholas, B. J.; Tsen, K. H. Y.; Gryko, D. T.; Arnold, J. Corroles That “Click”: Modular Synthesis of Azido- and PropargylFunctionalized Metallocorrole Complexes and Convergent Synthesis of a Bis-corrole Scaffold. Inorg. Chem. 2014, 53, 7941−7950. (g) Orłowski, R.; Vakuliuk, O.; Gullo, M. P.; Danylyuk, O.; Ventura, B.; Koszarna, B.; Tarnowska, A.; Jaworska, N.; Barbieri, A.; Gryko, D. T. Self-assembling corroles. Chem. Commun. 2015, 51, 8284−8287. (h) Einrem, R. F.; Gagnon, K. J.; Alemayehu, A. B.; Ghosh, A. Metal−Ligand Misfits: Facile Access to Rhenium−Oxo Corroles by Oxidative Metalation. Chem. − Eur. J. 2016, 22, 517−520. (I) Norheim, H.-K.; Capar, J.; Einrem, R. F.; Gagnon, K. J.; Beavers, C. M.; Vazquez-Lima, H.; Ghosh, A. Ligand noninnocence in FeNO corroles: insights from βoctabromocorrole complexes. Dalton Trans. 2016, 45, 681−689. (5) (a) Wang, Z.-L.; Xu, D.; Xu, J.-J.; Zhang, X.-B. Oxygen electrocatalysts in metal−air batteries: from aqueous to nonaqueous electrolytes. Chem. Soc. Rev. 2014, 43, 7746−7786. (b) Buckley, H. L.; Arnold, J. Recent developments in out-of-plane metallocorrole chemistry across the periodic table. Dalton Trans. 2015, 44, 30−36. (c) Fukuzumi, S. Electron transfer and catalysis with high-valent metaloxo complexes. Dalton Trans. 2015, 44, 6696−6705. (d) Barata, J. F. B.; Pinto, R. J. B.; Vaz Serra, V. I. R. C.; Silvestre, A. J. D.; Trindade, T.; Neves, M. G. P. M. S.; Cavaleiro, J. A. S.; Daina, S.; Sadocco, P.; Freire, C. S. R. Fluorescent Bioactive Corrole Grafted-Chitosan Films. Biomacromolecules 2016, 17, 1395−1403. (6) (a) Aviv-Harel, I.; Gross, Z. Aura of Coroles. Chem. − Eur. J. 2009, 15, 8382−8394. (b) Zhu, C.; Liang, J.-X. Theoretical insight into a novel zinc di-corrole dye with excellent photoelectronic properties for solar cells. New J. Chem. 2015, 39, 3624−3628. (c) Brennan, B. J.; Lam, Y. C.; Kim, P. M.; Zhang, X.; Brudvig, G. W. Photoelectrochemical Cells Utilizing Tunable Corroles. ACS Appl. Mater. Interfaces 2015, 7, 16124−16130. (d) Paolesse, R.; Natale, C. D.; Macagnano, A.; Sagone, F.; Scarselli, M. A.; Chiaradia, P.; Troitsky, V. I.; Berzina, T. S.; D’Amico, A. Langmuir-Blodgett Films of a Manganese Corrole Derivative. Langmuir 1999, 15, 1268−1274. (e) Santos, C. I. M.; Oliveira, E.; Barata, J. F. B.; Faustino, M. A. F.; Cavaleiro, J. A. S.; Neves, M. G. P. M. S.; Lodeiro, C. Corroles as anion chemosensors: exploiting their fluorescence behaviour from solution to solid-supported devices. J. Mater. Chem. 2012, 22, 13811−13819. (f) Santos, C. I. M.; Oliveira, E.; Barata, J. F. B.; Faustino, M. A. F.; Cavaleiro, J. A. S.; Neves, M. G. P. M. S.; Lodeiro, C. New Gallium(III) corrole complexes as colorimetric probes for toxic cyanide anion. Inorg. Chim. Acta 2014, 417, 148−154. (g) Garai, A.; Kumar, S.; Sinha, W.; Purohit, C. S.; Das, R.; Kar, S. A comparative study of optical nonlinearities of trans-A2B-corroles in solution and in aggregated state. RSC Adv. 2015, 5, 28643−28651. (h) Santos, C. I. M.; Barata, J. F. B.; Calvete, M. J. F.; Vale, L. S. H. P.; Dini, D.; Meneghetti, M.; Neves, M. G. P. M. S.; Faustino, M. A. F.; Tomé, A. C.; Cavaleiro, J. A. S. Synthesis and functionalization of corroles. An insight on their nonlinear optical absorption properties. Curr. Org. Synth. 2014, 11, 29−41. (i) Yadav, P.; Sankar, M. Synthesis, spectroscopic and electrochemical studies of phosphoryl and carbomethoxyphenyl substituted corroles, and their anion detection properties. Dalton Trans. 2014, 43, 14680−14688. (7) (a) Lemon, C. M.; Halbach, R. L.; Huynh, M.; Nocera, D. G. Photophysical Properties of β−Substituted Free-Base Corroles. Inorg. Chem. 2015, 54, 2713−2725. (b) Saltsman, I.; Mahammed, A.; Goldberg, I.; Tkachenko, E.; Botoshansky, M.; Gross, Z. Selective Substitution of Corroles: Nitration, Hydroformylation, and Chlorosulfonation. J. Am. Chem. Soc. 2002, 124, 7411−7420. (c) Vestfrid, J.; Goldberg, I.; Gross, Z. Tuning the Photophysical and Redox

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel: +91-1332-284753. ORCID

Muniappan Sankar: 0000-0001-6667-3759 Author Contributions

P.Y. and P.R. synthesized the compound and performed the analysis. P.Y. and M.S. prepared the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS M.S. sincerely thanks Board of Research in Nuclear Science (2012/37C/61/BRNS) and Science and Engineering Research Board (SB/FT/CS-015/2012) for funding. We are thankful to Dr. J. Sankar, Department of Chemistry, IISER Bhopal, India, and Dr. G. Rajaraman, Department of Chemistry, IIT Bombay, India, for EPR measurements. P.Y. thanks CSIR, Govt. of India for Senior Research Fellowship.



REFERENCES

(1) Johnson, A. W.; Kay, I. T. Corroles. Part 1 Synthesis. J. Chem. Soc. 1965, 306, 1620−1629. (2) (a) Paolesse, R.; Jaquinod, L.; Nurco, D. J.; Mini, S.; Sagone, F.; Boschi, T.; Smith, K. M. 5,10,15-Triphenylcorrole: A Product From a Modified Rothemund Reaction. Chem. Commun. 1999, 1307−1308. (b) Paolesse, R.; Nardis, S.; Sagone, F.; Khoury, R. G. Synthesis and Functionalization of meso-Aryl-Substituted Corroles. J. Org. Chem. 2001, 66, 550−556. (c) Gross, Z.; Galili, N.; Simkhovich, L.; Saltsman, I.; Botoshansky, M.; Blaser, D.; Boese, R.; Goldberg, I. Solvent-Free Condensation of Pyrrole and Pentafluorobenzaldehyde: A Novel Synthetic Pathway to Corrole and Oligopyrromethenes. Org. Lett. 1999, 1, 599−602. (d) Gross, Z.; Galili, N.; Saltsman, I. The First Direct Synthesis of Corroles From Pyrrole. Angew. Chem., Int. Ed. 1999, 38, 1427−1429. (3) (a) Orłowski, R.; Gryko, D.; Gryko, D. T. Synthesis of Corroles and Their Heteroanalogs. Chem. Rev. 2017, 117, 3102. (b) Paolesse, R.; Nardis, S.; Monti, D.; Stefanelli, M.; Di Natale, C. Porphyrinoids for Chemical Sensor Applications. Chem. Rev. 2017, 117, 2517. (c) Chatterjee, T.; Shetti, V. S.; Sharma, R.; Ravikanth, M. Heteroatom-Containing Porphyrin Analogues. Chem. Rev. 2017, 117, 3254. (d) Setsune, J.-I. 2,2′-Bipyrrole-Based Porphyrinoids. Chem. Rev. 2017, 117, 3044. (e) König, M.; Faschinger, F.; Reith, L. M.; Schöfberger, W. The evolution of corrole synthesis − from simple onepot strategies to sophisticated ABC-corroles. J. Porphyrins Phthalocyanines 2016, 20, 96−107. (f) Fang, Y.; Ou, Z.; Kadish, K. M. Electrochemistry of Corroles in Nonaqueous Media. Chem. Rev. 2017, 117, 3377. (g) Barata, J. F. B.; Neves, M. G. P. M. S.; Faustino, M. A. F.; Tome, A. C.; Cavaleiro, J. A. S. Strategies for Corrole Functionalization. Chem. Rev. 2017, 117, 3192. (h) Ding, Y.; Zhu, W.-H.; Xie, Y. Development of Ion Chemosensors Based on Porphyrin Analogues. Chem. Rev. 2017, 117, 2203. (i) Sarma, T.; Panda, P. K. Annulated Isomeric, Expanded, and Contracted Porphyrins. Chem. Rev. 2017, 117, 2785. (j) Zhang, W.; Lai, W.; Cao, R. Energy-Related Small Molecule Activation Reactions: Oxygen Reduction and Hydrogen and Oxygen Evolution Reactions Catalyzed by Porphyrinand Corrole-Based Systems. Chem. Rev. 2017, 117, 3717. (k) Teo, R. D.; Hwang, J. Y.; Termini, J.; Gross, Z.; Gray, H. B. Fighting Cancer with Corroles. Chem. Rev. 2017, 117, 2711. (4) (a) Capar, C.; Hansen, L. K.; Conradie, J.; Ghosh, A. βoctabromo-meso-tris(pentafluorophenyl)corrole: reductive demetalation-based synthesis of a heretofore inaccessible, perhalogenated freebase corrole. J. Porphyrins Phthalocyanines 2010, 14, 509−512. (b) Thomas, K. E.; Wasbotten, I. H.; Ghosh, A. Copper βOctakis(trifluoromethyl)corroles: New Paradigms for Ligand Substituent Effects in Transition Metal Complexes. Inorg. Chem. 2008, 47, 964

DOI: 10.1021/acsomega.6b00430 ACS Omega 2017, 2, 959−965

ACS Omega

Article

Properties of Metallocorroles by Iodination. Inorg. Chem. 2014, 53, 10536−10542. (d) Stefanelli, M.; Mandoj, F.; Nardis, S.; Raggio, M.; Fronczek, F. R.; McCandless, G. T.; Smith, K. M.; Paolesse, R. Corrole and nucleophilic aromatic substitution are not incompatible: a novel route to 2,3-difunctionalized copper corrolates. Org. Biomol. Chem. 2015, 13, 6611−6618. (e) Stefanelli, M.; Mancini, M.; Raggio, M.; Nardis, S.; Fronczek, F. R.; McCandless, G. T.; Smith, K. M.; Paolesse, R. 3-NO2-5,10,15-triarylcorrolato-Cu as a versatile platform for synthesis of novel 3-functionalized corrole derivatives. Org. Biomol. Chem. 2014, 12, 6200−6207. (8) (a) Koszarna, B.; Gryko, D. T. Efficient Synthesis of mesoSubstituted Corroles in a H2O-MeOH Mixture. J. Org. Chem. 2006, 71, 3707−3717. (b) Ngo, T. H.; Puntoriero, F.; Nastasi, F.; Robeyns, K.; Meervelt, L. V.; Campagna, S.; Dehaen, W.; Maes, W. Synthetic, Structural, and Photophysical Exploration of meso-PyrimidinylSubstituted AB2-Corroles. Chem. − Eur. J. 2010, 16, 5691−5705. (9) (a) Shen, J.; Shao, J.; Ou, Z.; Wenbo, E.; Koszarna, B.; Gryko, D. T.; Kadish, K. M. Electrochemistry and Spectroelectrochemistry of meso-Substituted Free-Base Corroles in Nonaqueous Media: Reactions of (Cor)H3, [(Cor)H4]+, and [(Cor)H2]−. Inorg. Chem. 2006, 45, 2251−2265. (b) Schweyen, P.; Brandhorst, K.; Wicht, R.; Wolfram, B.; Bröring, M. The corrole radical. Angew. Chem., Int. Ed. 2015, 54, 8213−8216. (c) Gopalakrishna, T. Y.; Reddy, J. S.; Anand, V. G. An Amphoteric Switch to Aromatic and Antiaromatic States of a Neutral Air-Stable 25p Radical. Angew. Chem., Int. Ed. 2014, 53, 10984−10987. (d) Yoshida, T.; Zhou, W.; Furuyama, T.; Leznoff, D. B.; Kobayashi, N. An Extremely Air-Stable 19π Porphyrinoid. J. Am. Chem. Soc. 2015, 137, 9258−9261.

965

DOI: 10.1021/acsomega.6b00430 ACS Omega 2017, 2, 959−965