Abstract

Ground-state and time-dependent density functional theory (TDDFT) calculations with the long-range-corrected, Coulomb-attenuating CAMY-B3LYP exchange-correlation functional and large, all-electron STO-TZ2P basis sets have been used to examine the potential “inverse hypercorrole” character of meso-p-nitrophenyl-appended dicyanidocobalt(III) corrole dianions. The effect is most dramatic for 5,15-bis(p-nitrophenyl) derivatives, where it manifests itself in intense NIR absorptions. The 10-aryl groups in these complexes play a modulatory role, as evinced by experimental UV–visible spectroscopic and electrochemical data for a series of 5,15-bis(p-nitrophenyl) dicyanidocobalt(III) corroles. TDDFT (CAMY-B3LYP) calculations ascribe these features clearly to a transition from the corrole’s a2u-like HOMO (retaining the D4h irrep used for metalloporphyrins) to a nitrophenyl-based LUMO. The outward nature of this transition contrasts with the usual phenyl-to-macrocycle direction of charge transfer transitions in many hyperporphyrins and hypercorroles; thus, the complexes studied are aptly described as inverse hypercorroles.
Short abstract
TDDFT calculations lend support to the idea of “inverse hypercorroles”, in which outward corrole-to-meso-aryl transitions result in intense NIR absorption bands.
Introduction
Gouterman’s four-orbital model conceptualizes the classic optical spectra of porphyrins as transitions from two near-degenerate highest occupied molecular orbitals (HOMOs), which transform as a1u and a2u in a D4h metalloporphyrin, to two degenerate lowest unoccupied molecular orbitals (LUMOs), which transform as eg.1−5 In a “normal” porphyrin, these four molecular orbitals (MOs) are well-separated from all other occupied and unoccupied orbitals. Hyperporphyrins are a diverse class of porphyrin derivatives with red-shifted optical spectra in which the frontier orbitals are modified in one of a myriad ways that lower the HOMO–LUMO gap.6 Two common mechanisms underlying hyper spectra involve admixture of transition metal d orbitals or of substituent-based orbitals into porphyrin-based MOs. In certain cases, the very identity of the HOMOs and LUMOs may be altered, and these may correspond to orbitals unrelated to the four-orbital model. A classic example of the latter scenario is found in diprotonated meso-tetrakis(p-aminophenyl)porphyrin, [H4TAPP]2+ in which the HOMO is almost exclusively localized on the electron-rich p-aminophenyl groups.7−10 The lowest-energy transition in this system is then an aminophenyl-to-porphyrin charge transfer transition. As it happens, charge transfer transitions underlie many, if not most, cases of hyperporphyrin spectra.6
In the early days of corrole chemistry,11−14 it was shown that simple corrole derivatives also conform to the four-orbital model.15 Soon, several cases of hypercorroles emerged, consisting of noninnocent transition metal meso-triarylcorroles in which the major optical transitions are thought to involve a significant degree of aryl-to-corrole charge transfer character.16,17 Protonated meso-tris(p-aminophenyl)corrole, an analogue of [H4TAPP]2+, was also shown to exhibit a hyper spectrum.18 In a recent Perspective article,6 Wamser and Ghosh considered the possibility of what might be termed inverse hyper spectra in which outward electron flow from a porphyrin or corrole core to the meso-aryl substituents results in strongly red-shifted optical spectra. Recently, Osterloh et al.19 have suggested, based on UV–vis-NIR absorption and electrochemical evidence, that meso-nitrophenyl-appended dicyanidocobalt(III) corrole dianions should qualify as inverse hypercorroles. In the absence of modern quantum chemical studies, however, the theoretical basis of the inverse hypercorrole description has remained uncertain and speculative.
We accordingly undertook a state-of-the-art ground-state and time-dependent density functional theory (DFT and TDDFT20,21) study of four meso-nitrophenyl-appended dicyanidocobalt(III) corroles (Scheme 1), namely {Co[TPC](CN)2}2– (C0), {Co[(5,15-P)(10-pNO2P)C](CN)2}2– (C1), {Co[(5,15-pNO2P)(10-P)C](CN)2}2– (C2), and {Co[TpNO2PC](CN)2}2– (C3), where P = phenyl, C = corrole, and TPC = triphenylcorrole. As described below, the results confirm the formulation of these systems as authentic inverse hypercorroles. Additional experimental data on the effects of substituents on the hypercorrole spectra are also included.
Scheme 1. Structures of Species Computationally Modeled in This Study.
In the Cn notation used, C refers ‘computational modeling’ and the numeral n to the number of nitrophenyl groups in the species.
Experimental Section
Starting Materials
All chemicals and solvents were of the highest grade available and were used without further purification. Benzonitrile (PhCN) was purchased from Sigma-Aldrich and distilled from P4O10 under a vacuum prior to use. Tetra-n-butyl-ammonium perchlorate (TBAP, for electrochemical analysis, ≥ 99.0%) and 95.0% tetra-n-butyl-ammonium cyanide (TBACN, 95%) were purchased from Sigma-Aldrich and stored in a desiccator until used.
UV–Visible Spectroscopy
UV–visible spectra of the synthesized compounds were recorded on a Varian Cary 50 or a Hewlett-Packard model 8453 diode array spectrophotometer, and quartz cells with an optical path length of 10 mm were used.
NMR Spectroscopy
1H NMR spectra were recorded in CDCl3 on a Bruker AVANCE NEO spectrometer (400 and 500 MHz). The measurements were made at the PACSMUB-WPCM technological platform, which relies on the “Institut de Chimie Moléculaire de l’Université de Bourgogne” and Welience “TM”, a Burgundy University private subsidiary. For DMSO-ligated cobalt corroles, gaseous NH3 was added to enhance the resolution of the spectra.
Mass Spectrometry
Mass spectra were recorded on a Bruker Microflex LRF MALDI Tandem TOF Mass Spectrometer using dithranol as the matrix or on an LTQ Orbitrap XL (Thermo) instrument in the ESI mode (for the HRMS spectra). Corroles S1 and S4 were prepared as described in the literature.19
General Procedure for the Synthesis of Free-Base Cobalt Corroles
The 5-(4-nitrophenyl)dipyrromethane (5.62 mmol, 1 equiv) and the appropriate benzaldehyde (2.81 mmol, 0.5 equiv) were dissolved in 560 mL of methanol. Then, a solution of HCl (36%, 28.0 mL) in H2O (560 mL) was added, and the reaction mixture was stirred at room temperature for 2 h. The mixture was extracted with chloroform, and the organic phase was washed three times with water, dried, and completed to 1.5L. p-Chloranil (1.5 equiv) was added, and the reaction mixture was stirred overnight at room temperature protected from light. Then 7.0 mL of hydrazine was added, and the mixture was further stirred for 30 min. After that, the solvent was evaporated and filtered on a dicalite plug. The compound thus obtained was purified by silica column with CHCl3 as the eluent. The crude compound was further recrystallized with DCM and heptane to afford crystal powder. The solid was filtered and dried under vacuum.
General Procedure for the Synthesis of Mono-DMSO Cobalt Corroles
Each free-base corrole (1.0 equiv) was added to a solution of cobalt acetate tetrahydrate (1.2 equiv) in DMSO (20 mL) in a round-bottom flask, after which the reaction mixture was stirred at 80 °C for 40 min and then cooled to room temperature. The crude mixture was poured into a cold NaCl aqueous solution (0.8 M), the resulting suspension was filtered, and the desired mono-DMSO cobalt corrole (Scheme 2) was washed five times with water (centrifugation) and dried overnight under vacuum.
Scheme 2. Mono-DMSO Cobalt Corroles Employed in This Study.
Metallocorrole S2
The free-base corrole was synthesized according to a published procedure.22 The mono-DMSO metallocorrole S2 was synthesized according to the general procedure starting from 100.1 mg of free-base corrole in 96% yield (115.1 mg). UV–visible (DCM): λmax (ε x 10–3 M–1 cm–1) 386.9 (70.24), 561.0 (13.47) nm. 1H NMR (500 MHz, CDCl3 + NH3 (g)) δ (ppm) 9.26 (d, 3JH–H = 4.5 Hz, 2H), 8.99 (d, 3JH–H = 4.5 Hz, 2H), 8.81 (m, 4H), 8.63 (d, 3JH–H = 8.5 Hz, 4H), 8.46 (d, 3JH–H = 8.5 Hz, 4H), 8.32 (d, 3JH–H = 8.0 Hz, 2H), 7.99 (d, 3JH–H = 8.0 Hz, 2H), 2.57 (s, 6H), −6.59 (s, 6H). 19F NMR (470 MHz, CDCl3 + NH3 (g)) δ −61.89 (s, 3F). MS (MALDI-TOF) m/z = 740.05 [M-DMSO]+., 740.08 calcd for C38H20CoF3N6O4. HR-MS (ESI): m/z = 740.0822 [M-DMSO]+., 740.0825 calcd for C38H20CoF3N6O4.
Metallocorrole S3
The free-base corrole was synthesized according to a published procedure.23 The mono-DMSO metallocorrole S3 was synthesized according to the general procedure starting from 50 mg of free-base corrole in 92% yield (55.3 mg). UV–visible (DCM): λmax (ε x 10–3 M–1 cm–1) 390.0 (55.24), 560.0 (10.16) nm. 1H NMR (500 MHz, CDCl3 + NH3 (g)) δ (ppm) 9.30 (d, 3JH–H = 4.5 Hz, 2H), 9.02 (d, 3JH–H = 4.5 Hz, 2H), 8.89 (d, 3JH–H = 4.5 Hz, 2H), 8.85 (d, 3JH–H = 4.5 Hz, 2H), 8.66 (d, 3JH–H = 8.5 Hz, 4H), 8.49 (d, 3JH–H = 8.5 Hz, 4H), 8.44 (d, 3JH–H = 8.0 Hz, 2H), 8.32 (d, 3JH–H = 8.0 Hz, 2H), 4.09 (s, 3H), 2.60 (s, 6H), −6.72 (s, 6H). MS (MALDI-TOF) m/z = 730.21 [M-DMSO]+., 730.10 calcd for C39H23CoN6O4. HR-MS (ESI): m/z = 730.1035 [M-DMSO]+., 730.1006 calcd for C39H23CoN6O4.
Metallocorrole S5
The free-base corrole was synthesized according to the general procedure using 341.3 μL of 4-methoxybenzaldehyde. Yield: 42% (76.3 mg). UV–visible (DCM): λmax (ε × 10–3 M–1cm–1) 442.0 (52.54), 594.0 (16.72), 656.0 (17.13) nm. 1H NMR (500 MHz, CDCl3 + NH3 (g)) δ (ppm) 9.07 (d, 3JH–H = 4.5 Hz, 2H), 8.86 (d, 3JH–H = 4.5 Hz, 2H), 8.69 (d, 3JH–H = 8.5 Hz, 4H), 8.66 (d, 3JH–H = 4.5 Hz, 2H), 8.62 (d, 3JH–H = 4.5 Hz, 2H), 8.55 (d, 3JH–H = 8.5 Hz, 4H), 8.09 (d, 3JH–H = 8.0 Hz, 2H), 7.32 (d, 3JH–H = 8.0 Hz, 2H), 4.10 (s, 3H). MS (MALDI-TOF) m/z = 646.20 [M]+., 646.20 calcd for C38H26N6O5. HR-MS (ESI): m/z = 647.2032 [M + H]+, 647.2037 calcd for C38H27N6O5.
The mono-DMSO metallocorrole S5 was synthesized according to the general procedure from 50.3 mg of free-base corrole in 91% yield (55.4 mg). UV–visible (DCM): λmax (ε × 10–3 M–1cm–1) 391.0 (47.38), 563.0 (4.43) nm. 1H NMR (400 MHz, CDCl3 + NH3 (g)) δ (ppm) 9.31 (d, 3JH–H = 4.5 Hz, 2H), 9.02 (d, 3JH–H = 4.5 Hz, 2H), 8.95 (d, 3JH–H = 4.5 Hz, 2H), 8.85 (d, 3JH–H = 4.5 Hz, 2H), 8.66 (d, 3JH–H = 8.5 Hz, 4H), 8.51 (d, 3JH–H = 8.5 Hz, 4H), 8.14 (d, 3JH–H = 8.0 Hz, 2H), 7.31 (d, 3JH–H = 8.0 Hz, 2H), 4.10 (s, 3H), 2.61 (s, 6H), −6.82 (s, 6H). MS (MALDI-TOF) m/z = 702.13 [M-DMSO]+., 702.11 calcd for C38H23CoN6O5. HR-MS (ESI): m/z = 702.1053 [M-DMSO]+., 702.1056 calcd for C38H23CoN6O5.
Metallocorrole S6
The free-base corrole was synthesized according to the general procedure using 418.7 mg of 4-(dimethylamino)benzaldehyde. Yield: 3.3% (61.9 mg). UV–visible (DCM): λmax (ε × 10–3 M–1 cm–1) 446.9 (45.06), 598.0 (14.02), 666.9 (15.9) nm. 1H NMR (500 MHz, CDCl3 + NH3 (g)) δ (ppm) 9.01 (d, 3JH–H = 4.5 Hz, 2H), 8.81 (d, 3JH–H = 4.5 Hz, 2H), 8.69 (d, 3JH–H = 4.5 Hz, 2H), 8.66 (d, 3JH–H = 8.5 Hz, 4H), 8.58 (d, 3JH–H = 4.5 Hz, 2H), 8.54 (d, 3JH–H = 8.5 Hz, 4H), 8.04 (d, 3JH–H = 8.0 Hz, 2H), 7.11 (d, 3JH–H = 8.0 Hz, 2H), 3.22 (s, 6H). MS (MALDI-TOF) m/z = 659.13 [M]+., 659.23 calcd for C39H29N7O4. HR-MS (ESI): m/z = 660.2352 [M + H]+, 660.2354 calcd for C39H30N7O4.
The mono-DMSO metallocorrole S6 was synthesized according to the general procedure starting from 19.5 mg of free-base corrole in 42% yield (9.8 mg). UV–visible (DCM): λmax (ε × 10–3 M–1cm–1) 390.0 (25.02), 568.1 (5.22), nm. 1H NMR (500 MHz, CDCl3 + NH3 (g)) δ (ppm) 9.27 (d, 3JH–H = 4.5 Hz, 2H), 8.99 (m, 4H), 8.83 (d, 3JH–H = 4.5 Hz, 2H), 8.64 (d, 3JH–H = 8.5 Hz, 4H), 8.49 (d, 3JH–H = 8.5 Hz, 4H), 8.08 (d, 3JH–H = 8.0 Hz, 2H), 7.13 (d, 3JH–H = 8.0 Hz, 2H), 3.22 (s, 6H), 2.59 (s, 6H), −6.70 (s, 6H). MS (MALDI-TOF) m/z = 715.18 [M-DMSO]+., 715.14 calcd for C39H26CoN7O4. HR-MS (ESI): m/z = 715.1376 [M-DMSO]+., 715.1373 calcd for C39H26CoN7O4.
General Procedure for the Generation of Dicyanidocobalt Corroles
Dicyanidocobalt corrole dianions were generated in situ by dissolving the mono-DMSO metallocorroles S1–S6 in benzonitrile containing 0.1 M tetra(n-butyl)ammonium perchlorate (TBAP) followed by the addition of 100 equiv of tetra(n-butyl)ammonium cyanide (TBACN).
Electrochemistry
Cyclic voltammetry was carried out at 298 K in benzonitrile (purified as described as earlier24) using an EG&G Princeton Applied Research (PAR) 173 potentiostat/galvanostat. A homemade three-electrode cell was used for all electrochemical measurements. A three-electrode system was used in each case and consisted of a glassy carbon working electrode. A platinum wire served as the auxiliary electrode and a saturated calomel electrode as the reference electrode, which was separated from the bulk of the solution by means of a salt bridge of low porosity which contained the solvent-supporting electrolyte (TBAP) mixture.
DFT and TDDFT Calculations
Geometry optimizations were carried out with scalar-relativistic DFT using the zeroth order regular approximation (ZORA25) to the Dirac equation, the OLYP26,27 functional augmented with the Grimme’s D328,29 dispersion correction, all-electron Slater-type ZORA TZ2P basis sets, fine integration grids, and tight criteria for the SCF cycles and geometry optimizations, as implemented in the ADF program system.30 A C2 symmetry constraint was used for all four species studied (Scheme 2). Solvation was modeled with COSMO (conductor-like screening model31−34) with acetonitrile as the solvent. The optimized geometries so obtained were used for range-separated TDDFT calculations with the CAMY-B3LYP35−37 exchange-correlation functionals and the same basis sets.
Results
UV–Vis–NIR and Electrochemical Studies
One of the most notable cases of hypercorrole spectra has been recently documented for dicyanidocobalt(III) 5,15-bis(p-nitrophenyl)corroles and 5,10,15-tris(p-nitrophenyl)corrole.19 The complexes exhibit an intense absorption in the 780–850 nm range, which is absent in analogous bis-cyano ligated triarylcorrole complexes that do not carry p-nitrophenyl groups at the lateral 5,15-meso-substituents. Thus, a 10-(p-nitrophenyl) substituent, by itself, does not lead to a similarly striking hypercorrole spectrum.38−40 In the same vein, 5,15-(m-nitrophenyl) substituents do not quite lead to a dramatic hypercorrole spectrum.19 Herein, we have further clarified the role of 10-substituents via the examination of a series of 5,15-bis(p-nitrophenyl)-10-(p-X-phenyl)corrole complexes, {CoIII[(pNO2P)2(pXP)C](CN)}2–, where the para substituent (X) at the 10-position ranges across those shown in Scheme 1.
As shown in Figure 1 and summarized in Table 1, the lowest-energy absorption for the dicyanidocobalt(III) complexes in benzonitrile was found to shift from 798 nm for X = NO2 to 847 nm for X = NMe2 (Table 1 and Figure 1). Cyclic voltammetry measurements suggest that the shift reflects a modest elevation in the orbital energy of the HOMO by about 130 mV, going from X = NO2 to X = NMe2, while the LUMO, presumably localized in the 5,15-p-nitrophenyl groups, remains essentially constant in terms of orbital energy (Figure 2). As discussed below, DFT calculations nicely confirm this conclusion. Furthermore, both the energy of the lowest-energy NIR absorption band and the oxidation potential increase linearly with the Hammett substituent constants of the para substituents, with excellent correlation coefficients (Figure 3a). The linear shifts of the oxidation potentials with σpara strongly suggest that HOMO has the same qualitative character across all of the species studied. Understandably, the NIR absorption energies and oxidation potentials also exhibit a linear relationship (Figure 3b).
Figure 1.

UV–vis spectra of cobalt 5,15-di(4-nitrophenyl)corroles (at ∼10–5 M) all in PhCN containing 0.1 M TBAP with 100 equiv of added TBACN.
Table 1. UV–visible Spectral Data for Dicyanidocobalt(III) 5,15-Bis(p-nitrophenyl)-10-(p-X-phenyl)corroles Generated In Situ from Precursors S1–S6 (See Scheme 2 for Structures) and TBACN (100 equiv) in PhCNa.
| λ, nm (ε × 10–4 M–1 cm–1) | |||||
|---|---|---|---|---|---|
| precursor | Xb | σp | visible region | NIR region | |
| S1 | NO2 | 0.78 | 438 (3.1) | 591 (1.8) | 798 (2.2) |
| S2 | CF3 | 0.54 | 447 (3.3) | 576 (1.2) | 807 (1.9) |
| S3 | CO2CH3 | 0.45 | 448 (3.2) | 581 (1.2) | 809 (2.0) |
| S4 | H | 0.00 | 445 (3.3) | 581 (1.0) | 826 (1.8) |
| S5 | OCH3 | –0.27 | 445 (3.4) | 581 (0.9) | 834 (1.8) |
| S6 | N(CH3)2 | –0.83 | 445 (3.4) | 583 (0.9) | 847 (1.6) |
The diagnostic inverse hypercorrole maxima are indicated in bold.
X = the para substituent at the 10-meso phenyl position.
Figure 2.

Cyclic voltammograms of cobalt 5,15-di(4-nitrophenyl)corroles in PhCN/0.1 M TBAP with 100 equiv of added TBACN. The reduction process at −1.10 V in blue corresponds to overlapping electron additions at the two or three meso-nitrophenyl groups. Scan rate: 0.1 V/s.
Figure 3.

(a) Hammett plots for the lowest-energy absorption band (above) and E1/2 for the first oxidation process (below), for measurements in 0.1 M TBAP in PhCN with 100 equiv of added TBACN and (b) plot of wavenumber for the lowest energy absorption band vs E1/2 for the first oxidation process in PhCN/0.1 M TBAP with 100 equiv of added TBACN.
DFT and TDDFT Calculations
The experimental data were modeled with state-of-the-art DFT and range-separated TDDFT calculations on the four species depicted in Scheme 1. For complexes C1–C3, the nitrophenyl-based LUMOs are considerably below the corrole-based LUMO, resulting in dramatically lower HOMO–LUMO gaps relative to those of the unadorned TPC complex, C0, as shown in Figure 4. Paralleling experimental measurements, the lowest DFT HOMO–LUMO gap was found for the 5,15-bis(p-nitrophenyl) complex with an unsubstituted 10-phenyl group, C2, closely followed by the tris(p-nitrophenyl) complex, C3. A slightly higher HOMO–LUMO gap is predicted for 10-nitrophenyl complex C1 with unsubstituted 5,15-phenyl groups, while a much higher HOMO–LUMO gap, understandably, is found for unadorned TPC complex C0. In other words, the HOMO–LUMO gaps will follow the order: C2 < C3 < C1 < C0. CAMY-B3LYP TDDFT calculations assign the lowest-energy absorption of each system to an overwhelming HOMO-to-LUMO transition. Understandably, the TDDFT transition energies (Table 2 and Figures 5 and 6) mirror the order of Kohn–Sham HOMO–LUMO gaps.
Figure 4.

CAMY-B3LYP-D3/STO-TZ2P-COSMO frontier MO energy levels along with C2 irreps (a and b).
Table 2. CAMY-B3LYP-D3/STO-TZ2P TDDFT Results, Including Wavelengths (λ), Oscillator Strengths (f), MO Compositions, and Symmetries.
| E (eV) | λ (nm)a | f | MO compositionb | state symmetry | ||
|---|---|---|---|---|---|---|
| weight (%) | from | to | ||||
| Complex C0 | ||||||
| 1.88 | 659.2 (a) | 0.56 | 94.6 | HOMO | LUMO | B |
| 3.5 | HOMO–1 | LUMO + 1 | B | |||
| 2.26 | 548.0 (b) | 0.03 | 58.6 | HOMO–1 | LUMO | A |
| 39.2 | HOMO | LUMO + 1 | A | |||
| 2.58 | 481.2 (c) | 1.58 | 58.5 | HOMO | LUMO + 1 | A |
| 39.2 | HOMO–1 | LUMO | A | |||
| 2.99 | 414.6 (d) | 1.04 | 91.7 | HOMO–1 | LUMO + 1 | B |
| 3.5 | HOMO | LUMO | B | |||
| Complex C1 | ||||||
| 1.64 | 758.0 (e) | 0.89 | 94.8 | HOMO | LUMO | A |
| 2.5 | HOMO | LUMO + 2 | A | |||
| 1.96 | 632.9 (f) | 0.34 | 84.0 | HOMO | LUMO + 1 | B |
| 8.9 | HOMO–1 | LUMO | B | |||
| 5.1 | HOMO–1 | LUMO + 2 | B | |||
| 2.35 | 527.5 (g) | 0.34 | 84.3 | HOMO–1 | LUMO | B |
| 11.5 | HOMO | LUMO + 1 | B | |||
| 2.37 | 523.5 (h) | 0.36 | 74.9 | HOMO–1 | LUMO + 1 | A |
| 22.0 | HOMO | LUMO + 2 | A | |||
| 2.71 | 457.8 (i) | 0.76 | 72.8 | HOMO | LUMO + 2 | A |
| 22.9 | HOMO–1 | LUMO + 1 | A | |||
| 3.11 | 398.4 (j) | 0.81 | 86.7 | HOMO–1 | LUMO + 2 | B |
| 4.8 | HOMO–1 | LUMO | B | |||
| Complex C2 | ||||||
| 1.48 | 839.7 (k) | 1.24 | 89.5 | HOMO | LUMO | B |
| 7.9 | HOMO | LUMO + 2 | B | |||
| 1.82 | 679.4 | 0.01 | 75.1 | HOMO | LUMO + 1 | A |
| 13.3 | HOMO–1 | LUMO | A | |||
| 5.3 | HOMO | LUMO + 3 | A | |||
| 2.01 | 616.5 (l) | 0.36 | 69.3 | HOMO–1 | LUMO | A |
| 19.6 | HOMO | LUMO + 1 | A | |||
| 8.8 | HOMO–1 | LUMO + 2 | A | |||
| 2.30 | 539.9 (m) | 0.31 | 43.8 | HOMO–1 | LUMO + 1 | B |
| 40.4 | HOMO | LUMO + 2 | B | |||
| 7.7 | HOMO–1 | LUMO + 3 | B | |||
| 5.3 | HOMO | LUMO | B | |||
| 2.41 | 514.4 (n) | 0.12 | 50.5 | HOMO–1 | LUMO + 1 | B |
| 43.8 | HOMO | LUMO + 2 | B | |||
| 2.63 | 471.3 (o) | 0.25 | 77.5 | HOMO | LUMO + 3 | A |
| 9.0 | HOMO–1 | LUMO | A | |||
| 7.8 | HOMO–1 | LUMO + 2 | A | |||
| 2.84 | 436.2 (p) | 0.90 | 76.0 | HOMO–1 | LUMO + 2 | A |
| 13.8 | HOMO | LUMO + 3 | A | |||
| 6.5 | HOMO–1 | LUMO | A | |||
| 3.11 | 398.9 (q) | 0.56 | 86.0 | HOMO–1 | LUMO + 3 | B |
| 5.1 | HOMO | LUMO + 2 | B | |||
| Complex C3 | ||||||
| 1.59 | 778.4 (r) | 1.04 | 86.4 | HOMO | LUMO | B |
| 10.1 | HOMO | LUMO + 3 | B | |||
| 1.74 | 714.1 (s) | 0.42 | 88.6 | HOMO | LUMO + 1 | A |
| 4.6 | HOMO | LUMO + 4 | A | |||
| 2.05 | 604.8 (t) | 0.59 | 63.7 | HOMO–1 | LUMO | A |
| 17.0 | HOMO | LUMO + 2 | A | |||
| 12.9 | HOMO–1 | LUMO + 3 | A | |||
| 2.12 | 583.8 | 0.00 | 79.3 | HOMO | LUMO + 2 | A |
| 14.7 | HOMO–1 | LUMO | A | |||
| 2.26 | 549.2 (u) | 0.46 | 69.5 | HOMO–1 | LUMO + 1 | B |
| 14.6 | HOMO | LUMO + 3 | B | |||
| 7.9 | HOMO–1 | LUMO + 4 | B | |||
| 2.46 | 504.9 (v) | 0.06 | 64.1 | HOMO | LUMO + 3 | B |
| 20.5 | HOMO–1 | LUMO + 1 | B | |||
| 6.4 | HOMO–1 | LUMO + 2 | B | |||
| 5.7 | HOMO | LUMO | B | |||
| 2.61 | 475.2 | 0.01 | 90.7 | HOMO–1 | LUMO + 2 | B |
| 4.0 | HOMO | LUMO + 3 | B | |||
| 2.72 | 455.6 | 0.03 | 65.6 | HOMO | LUMO + 4 | A |
| 16.5 | HOMO–1 | LUMO + 3 | A | |||
| 11.5 | HOMO–1 | LUMO | A | |||
| 2.88 | 430.8 (w) | 0.76 | 65.3 | HOMO–1 | LUMO + 3 | A |
| 25.0 | HOMO | LUMO + 4 | A | |||
| 3.17 | 391.4 (x) | 0.50 | 84.0 | HOMO–1 | LUMO + 4 | B |
| 4.7 | HOMO–1 | LUMO + 1 | B | |||
| 4.3 | HOMO | LUMO + 3 | B | |||
Figure 5.
Simulated TD-CAMY-B3LYP-D3/STO-TZ2P-COSMO optical spectra (oscillator strengths vs wavelength in nm) in dichloromethane. The vertical lines represent calculated transitions which have then been broadened with Gaussians to generate the simulated spectra. The peak labels are cross-referenced in Table 2, which lists the MO compositions of the peaks in question. The MOs themselves are visually depicted in Figure 6.
Figure 6.
Selected CAMY-B3LYP frontier MOs of dianions C0–C3, with C2 irreps (a, b) and orbital energies in eV.
For C0–C3, the largest redshift of the lowest-energy absorption is observed for dicyanido-cobalt 5,15-bis(p-nitrophenyl)-10-phenylcorrole, i.e., complex C2 (expt 826 nm in Table 1 and Figure 1; calc peak k at 839.7 nm in Figure 5 and Table 2) The second largest redshift for the NIR transition, both experimentally and theoretically, is exhibited by complex C3 (expt 798 nm in Table 1 and Figure 1; calc peak r at 778.4 nm in Figure 5 and Table 2) The third spot, experimentally, is occupied by the 5,15-dimesityl-10-(p-nitrophenyl)corrole complex (expt 732 nm19), which has been modeled here as complex C1 (calc peak e at 758 nm in Figure 5 and Table 2). The least red-shifted complex is that of meso-tris(p-t-butylphenyl)corrole (expt 696 nm19), which has been modeled here with TPC, i.e., complex C0 (calc peak a at 659 nm in Figure 5 and Table 2).
The calculations also permit plausible assignments of the remainder of the optical spectra. The experimentally studied species all exhibit an absorption in the 575–595 nm range, which appears as a shoulder in the majority of cases but as a distinct peak for {Co[TpNO2PC](CN)2}2– (C3) (Figure 1). This feature appears to correspond to essentially a (HOMO–1)-to-nitrophenyl transition (peak l at 616.5 nm for C2 and peak t at 604.8 nm for C3) in Figure 5 and Table 2, where HOMO–1 can be identified with the corrole analogue of the Gouterman a1u orbital of porphyrins. The relatively higher intensity of the feature for C3 appears to reflect additional charge transfer character mixing into the overall composition of the transition, as a result of the presence of the 10-nitrophenyl group. Finally, a relatively normal Gouterman-type four-orbital composition is indicated for the intense Soret-like features under 500 nm, i.e., peaks i and j for C1, peaks o–q for C2, and peaks w and x for C3 (see Figure 5 and Table 2).41
Overall, the calculated transition energies are in impressive, semiquantitative agreement with experimental absorption maxima, within the resolution of solution-phase spectroscopic measurements and allow for multiple conformations and details of solvation that we have not accounted for in our calculations. The agreement is all the more remarkable in that, for C1–C3, the lower-energy transitions largely involve corrole-to-nitrophenyl charge transfer character. TDDFT calculations routinely struggle with predicting the energetics of charge transfer transitions. Key to our success in the present study has been our earlier groundwork on hyperporphyrin systems,10,42 which established the importance of using an appropriate solvation model and a range-separated functional such as CAMY-B3LYP that provides improved description for charge transfer transitions.35−37 It will indeed be interesting to see how well the present methods perform vis-à-vis other anionic hyperporphyrin systems such as O-deprotonated meso-tetrakis(p-hydroxyphenyl)corrole.43,44
Discussion
The UV–vis-NIR spectra of dicyanidocobalt 5,15-bis(p-nitrophenyl) corroles, where the 10-position can vary, may be viewed as paradigms of inverse hyper spectra, with clean macrocycle-to-substituent charge-transfer transitions in the near-infrared. Such transitions reflect a clean LUMO switch in these systems (relative to other meso-triarylcorroles and tetraarylporphyrins), from macrocycle- to meso-aryl-based, as a result of the relatively weak electronic coupling between the macrocycle and the significantly twisted (i.e., out-of-plane) aryl substituents. These systems may be contrasted with β-formyl-,45 dicyanovinyl- and dicyanobutadienyl- metallocorroles46,47 in which strong NIR absorptions primarily reflect an extension of the corrole’s conjugation, with varying contributions of corrole-to-substituent charge transfer character.
It is interesting to reflect on the role of the metal center in engendering inverse hypercorrole spectra for meso-nitrophenyl-appended corroles. Some of us have suggested that an innocent corrole macrocycle is critical.19 For example, whereas neutral Cu[TpNO2PC] (the Cu analogue of C3), in which the corrole is thought to be noninnocent,48−53 does not exhibit much of a hypercorrole spectrum (in the form of strong NIR absorption), the innocent54,55 anionic species {Cu[TpNO2PC]}− exhibits a pronounced inverse hypercorrole spectrum similar to C3. Yet, an innocent meso-nitrophenyl-appended corrole, though possibly necessary for a pronounced inverse hypercorrole spectrum, does not guarantee one. Thus, innocent monocyanido analogues of the dicyanido cobalt corroles studied here do not exhibit an equally pronounced inverse hypercorrole effect. Nor, for that matter, do p-nitrophenyl-appended cobalt(III)-triphenylphosphine corroles.56 In the same vein, meso-p-nitrophenyl groups by themselves do not appear to elicit much of a hyperporphyrin effect in charge-neutral metalloporphyrins57 (although extending the conjugation with meso-p-nitrophenylethynyl groups does engender large spectral redshifts58). The dianionic character of the dicyanido complexes studied here and the sizable, negative formal charge on the cobalt (in spite of the + III oxidation state59) play a critical role in engendering the observed hypercorrole spectra.
The importance of the overall negative charge on the metal–corrole fragment and of anionic axial ligands immediately suggests applications of nitrophenyl-appended porphyrins and corroles as anion sensors. A handful of applications to the selective sensing of neutral ligands and heavy metal cations (such as Hg2+60 and Ru3+61) have already been reported in the literature; these systems, however, exhibit only modest, if any, NIR absorption and are at best viewed as incipient inverse hypercorroles. With the concept of an “inverse hypercorrole” authenticated both experimentally and theoretically as a result of this work, there is clearly considerable room for creativity in the design of new anion chemosensors. In the same vein, biocompatible inverse hyperporphyrins and hypercorroles, on account of their NIR emission, may lend themselves to applications in photomedicine, as new dyes for photodynamic and photothermal therapies and as physiological oxygen sensors.13,14
Conclusions
In a 2022 Perspective on The Hyperporphyrin Concept,6 we noted that hyper spectra arising via macrocycle-to-meso-aryl charge transfer were unknown. Herein, state-of-the-art TDDFT calculations have supported the formulation of meso-p-nitrophenyl-appended dicyanidocobalt(III) corroles as paradigmatic “inverse hypercorroles”. The intense NIR absorptions of these corroles are ascribed to a transition from the corrole HOMO (with a porphyrin a2u-like shape) to a nitrophenyl-based LUMO. The hypercorrole effect, as measured by the redshift and intensity of the NIR absorption, is particularly dramatic for 5,15-bis(p-nitrophenyl)-substituted complexes, with para substituents on the 10-phenyl group exercising a modulating influence. The simplicity of inverse hypercorrole design (involving meso-p-nitrophenyl substituents and anionic axial ligands) provides an attractive alternative to the traditional approach to NIR-absorbing porphyrinoids in which the macrocycle’s π-system is extended by conjugating substituents or arene annulation. Accordingly, we harbor the hope that applications of inverse hypercorroles to areas such as anion sensing and photomedicine will emerge in relatively short order, an exciting prospect from the perspective of the present study.
Acknowledgments
This research was supported by the Research Council of Norway (grant no. 262229 to A.G.), the South African National Research Foundation (grant nos. 129270 and 132504 to J.C.), the Robert A. Welch Foundation (K.M.K., Grant E-680), the CNRS (UMR UB-CNRS 6302), the “Agence Nationale de la Recherche” (ANR project MIPEnz-Decontam, grant no. ANR-20-CE39-0016), the “Université de Bourgogne”, the “Conseil Régional de Bourgogne”, and the European Union. The authors thank the “Plateforme d’Analyse Chimique et de Synthèse Moléculaire de l’Université de Bourgogne” (PACSMUB, http://www.wpcm.fr) for access to spectroscopy instrumentation. The authors also thank Dr. Quentin Bonnin and Mrs. Marie-José Penouilh (“Université de Bourgogne”, PACSMUB) for HRMS analysis. Mrs Sandrine Pacquelet is warmly acknowledged for synthetic contributions (synthesis of corroles and their precursors). This article is dedicated to Professor Carl Wamser of Portland State University, Portland, Oregon, on the occasion of his 80th birthday.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.4c00344.
MALDI/TOF LRMS, ESI HRMS spectra, and 1H NMR spectra of cobalt complexes S1–S6. Optimized coordinates of the DFT calculations (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- For an earlier version of this study, see: Osterloh W. R.; Desbois N.; Conradie J.; Gros C.; Kadish K.; Ghosh A.. Inverse Hypercorroles. ChemRxiv, submitted 29 January 2024. DOI: 10.26434/chemrxiv-2024-6672l. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gouterman M. Spectra of porphyrins. J. Mol. Spectrosc. 1961, 6, 138–163. 10.1016/0022-2852(61)90236-3. [DOI] [Google Scholar]
- Gouterman M.; Wagniére G. H.; Snyder L. C. Spectra of Porphyrins. Part II. Four-Orbital Model. J. Mol. Spectrosc. 1963, 11, 108–115. 10.1016/0022-2852(63)90011-0. [DOI] [Google Scholar]
- Ghosh A. An Exemplary Gay Scientist and Mentor: Martin Gouterman (1931–2020). Angew. Chem., Int. Ed. 2021, 60, 9760–9770. 10.1002/anie.202012840. [DOI] [PubMed] [Google Scholar]
- Gouterman M.Optical Spectra and Electronic Structure of Porphyrins and Related Rings. In The Porphyrins; Dolphin D., Ed.; Academic Press: New York, 1978; Vol. III, Part A, pp. 1–165. [Google Scholar]
- Wamser C. C.; Ghosh A. The Hyperporphyrin Concept: A Contemporary Perspective. JACS Au 2022, 2, 1543–1560. 10.1021/jacsau.2c00255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ojadi E. C. A.; Linschitz H.; Gouterman M.; Walter R. I.; Lindsey J. S.; Wagner R. W.; Droupadi P. R.; Wang W. Sequential Protonation of meso-(p-(Dimethylamino)phenyl)porphyrins: Charge-Transfer Excited States Producing Hyperporphyrins. J. Phys. Chem. 1993, 97, 13192–13197. 10.1021/j100152a025. [DOI] [Google Scholar]
- Wasbotten I. H.; Conradie J.; Ghosh A. Electronic absorption and resonance Raman signatures of hyperporphyrins and nonplanar porphyrins. J. Phys. Chem. B 2003, 107, 3613–3623. 10.1021/jp014417i. [DOI] [Google Scholar]
- Rudine A. B.; DelFatti B. D.; Wamser C. C. Spectroscopy of Protonated Tetraphenylporphyrins with Amino/Carbomethoxy Substituents: Hyperporphyrin Effects and Evidence for a Monoprotonated Porphyrin. J. Org. Chem. 2013, 78, 6040–6049. 10.1021/jo400742f. [DOI] [PubMed] [Google Scholar]
- Conradie J.; Wamser C. C.; Ghosh A. Understanding Hyperporphyrin Spectra: TDDFT Calculations on Diprotonated Tetrakis(p-aminophenyl)porphyrin. J. Phys. Chem. A 2021, 125, 9953–9961. 10.1021/acs.jpca.1c06621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ghosh A. Electronic Structure of Corrole Derivatives: Insights from Molecular Structures, Spectroscopy, Electrochemistry, and Quantum Chemical Calculations. Chem. Rev. 2017, 117, 3798–3881. 10.1021/acs.chemrev.6b00590. [DOI] [PubMed] [Google Scholar]
- Kumar A.; Kim D.; Kumar S.; Mahammed A.; Churchill D. G.; Gross Z. Milestones in Corrole Chemistry: Historical Ligand Syntheses and Post-Functionalization. Chem. Soc. Rev. 2023, 52, 573–600. 10.1039/D1CS01137E. [DOI] [PubMed] [Google Scholar]
- Mahammed A.; Gross Z. Milestones and Most Recent Advances in Corrole’s Science and Technology. J. Am. Chem. Soc. 2023, 2023 (145), 12429–12445. 10.1021/jacs.3c00282. [DOI] [PubMed] [Google Scholar]
- Di Natale C.; Gros C. P.; Paolesse R. Corroles at Work: A Small Macrocycle for Great Applications. Chem. Soc. Rev. 2022, 51, 1277–1335. 10.1039/D1CS00662B. [DOI] [PubMed] [Google Scholar]
- Ghosh A.; Wondimagegn T.; Parusel A. B. J. Electronic Structure of Gallium, Copper, and Nickel Complexes of Corrole. High-Valent Transition Metal Centers Versus Noninnocent Ligands. J. Am. Chem. Soc. 2000, 122, 5100–5104. 10.1021/ja9943243. [DOI] [Google Scholar]
- Alemayehu A. B.; Conradie J.; Ghosh A. A First TDDFT Study of Metallocorrole Electronic Spectra: Copper meso-Triarylcorroles Exhibit Hyper Spectra. Eur. J. Inorg. Chem. 2011, 12, 1857–1864. 10.1002/ejic.201001026. [DOI] [Google Scholar]
- Ganguly S.; Ghosh A. Seven Clues to Ligand Noninnocence: The Metallocorrole Paradigm. Acc. Chem. Res. 2019, 52, 2003–2014. 10.1021/acs.accounts.9b00115. [DOI] [PubMed] [Google Scholar]
- Thomassen I. K.; Ghosh A. Protonation-Induced Hyperporphyrin Spectra of meso-Aminophenylcorroles. ACS Omega 2020, 5, 9023–9030. 10.1021/acsomega.0c01068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Osterloh W. R.; Desbois N.; Gros C. P.; Kadish K. M. Hypercorroles Formed via the Tail that Wagged the Dog: Charge Transfer Interactions from Innocent Corroles to Meso-Nitrophenyl Substituents. Inorg. Chem. 2022, 61, 20576–20586. 10.1021/acs.inorgchem.2c03425. [DOI] [PubMed] [Google Scholar]
- Casida M. E.; Huix-Rotllant M. Progress in Time-Dependent Density-Functional Theory. Annu. Rev. Phys. Chem. 2012, 63, 287–323. 10.1146/annurev-physchem-032511-143803. [DOI] [PubMed] [Google Scholar]
- Maitra N. T. Charge Transfer in Time-Dependent Density Functional Theory. J. Phys.: Condens. Matter 2017, 29, 423001 10.1088/1361-648X/aa836e. [DOI] [PubMed] [Google Scholar]
- Yang J.; André L.; Desbois N.; Gros C. P.; Brandès S. 2D/3D Covalent Organic Frameworks Based on Cobalt Corroles for CO Binding. Mater. Today Chem. 2023, 28, 101357 10.1016/j.mtchem.2022.101357. [DOI] [Google Scholar]
- Yadav O.; Varshney A.; Kumar A.; Ratnesh R. K.; Mehata M. S. A2B-corroles: Fluorescence Signaling Systems for Sensing Fluoride Ions. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018, 202, 207–213. 10.1016/j.saa.2018.05.051. [DOI] [PubMed] [Google Scholar]
- Kadish K. M.; Anderson J. E. Purification of Solvents for Electroanalysis: Benzonitrile; Dichloromethane; 1,1-Dichloroethane and 1,2-Dichloroethane. Pure Appl. Chem. 1987, 59, 703–714. 10.1351/pac198759050703. [DOI] [Google Scholar]
- Van Lenthe E. V.; Snijders J. G.; Baerends E. J. The Zero-Order Regular Approximation for Relativistic Efects: The Effect of Spin–Orbit Coupling in Closed Shell Molecules. J. Chem. Phys. 1996, 105, 6505–6516. 10.1063/1.472460. [DOI] [Google Scholar]
- Handy N. C.; Cohen A. Left-Right Correlation Energy. J. Mol. Phys. 2001, 99, 403–412. 10.1080/00268970010018431. [DOI] [Google Scholar]
- Lee T C.T.; Yang W. T.; Parr R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron-Density. Phys. Rev. B 1988, 37, 785–789. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
- Grimme S. Density Functional Theory with London Dispersion Corrections. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1, 211–228. 10.1002/wcms.30. [DOI] [Google Scholar]
- Grimme S.; Antony J.; Ehrlich S.; Krieg H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
- Velde G. T.; Bickelhaupt F. M.; Baerends E. J.; Guerra C. F.; van Gisbergen S. J. A.; Snijders J. G.; Ziegler T. Chemistry with ADF. J. Comput. Chem. 2001, 22, 931–967. 10.1002/jcc.1056. [DOI] [Google Scholar]
- 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. 1993, 2, 799–805. 10.1039/P29930000799. [DOI] [Google Scholar]
- Klamt A. Conductor-like Screening Model for Real Solvents: A New Approach to the Quantitative Calculation of Solvation Phenomena. J. Phys. Chem. 1995, 99, 2224–2235. 10.1021/j100007a062. [DOI] [Google Scholar]
- Klamt A.; Jonas V. Treatment of the Outlying Charge in Continuum Solvation Models. J. Chem. Phys. 1996, 105, 9972–9981. 10.1063/1.472829. [DOI] [Google Scholar]
- Pye C. C.; Ziegler T. An Implementation of the Conductor-Like Screening Model of Solvation within the Amsterdam Density Functional Package. Theor. Chem. Acc. 1999, 101, 396–408. 10.1007/s002140050457. [DOI] [Google Scholar]
- Seth M.; Ziegler T. Range-Separated Exchange Functionals with Slater-Type Functions. J. Chem. Theory Comput. 2012, 8, 901–907. 10.1021/ct300006h. [DOI] [PubMed] [Google Scholar]
- Note that CAMY-B3LYP is the Yukawa form of CAM-B3LYP37 with a = 0.19 and b = 0.46 (and g = 0.34) but with the Yukawa potential as the switching function, as opposed to the Coulomb potential attenuated by the complementary error function.
- Yanai T.; Tew D. P.; Handy N. C. A New Hybrid Exchange-Correlation Functional Using the Coulomb-Attenuating Method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. 10.1016/j.cplett.2004.06.011. [DOI] [Google Scholar]
- The relatively muted role of 10-substituents has been noted before.39,40
- Alemayehu A.; Conradie M. M.; Ghosh A. Electronic Absorption Spectra of Copper Corroles: Unexpected Substituent Effects in Trans-meso-A2B-triarylcorrole Complexes. J. Porphyr. Phthalocyanines 2012, 16, 695–704. 10.1142/S108842461250085X. [DOI] [Google Scholar]
- Pomarico G.; Galloni P.; Mandoj F.; Nardis S.; Stefanelli M.; Vecchi A.; Lentini S.; Cicero D. O.; Cui Y.; Zeng L.; Kadish K. M.; Paolesse R. 5,10,15-Triferrocenylcorrole Complexes. Inorg. Chem. 2015, 54, 10256–10268. 10.1021/acs.inorgchem.5b01575. [DOI] [PubMed] [Google Scholar]
- Rhoda H. M.; Crandall L. A.; Geier G. R. III; Ziegler C. J.; Nemykin V. N. Combined MCD/DFT/TDDFT study of the electronic structure of axially pyridine coordinated metallocorroles. Inorg. Chem. 2015, 54, 4652–4662. 10.1021/ic502946t. [DOI] [PubMed] [Google Scholar]
- Ghosh A.; Conradie J. The Dog That Didn’t Bark: A New Interpretation of Hypsoporphyrin Spectra and the Question of Hypsocorroles. J. Phys. Chem. A 2021, 125, 9962–9968. 10.1021/acs.jpca.1c08425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Manna B. K.; Bera S. C.; Rohatgi-Mukherjee K. K. Effect of solvent and pH on the spectral characteristics of meso-tetrakis(p-hydroxyphenyl)porphyrin in dimethylformamide and dimethylformamide + water mixed solvents. Spectrochim. Acta A Mol. Biomol. Spectrosc. 1995, 51, 1051–1060. 10.1016/0584-8539(94)00163-6. [DOI] [Google Scholar]
- Lu G.; Zhang X.; Cai X.; Fang Y.; Zhu M.; Zhu W.; Ou Z.; Kadish K. M. Synthesis, structural characterization and protonation/deprotonation of hydroxyl-substituted free-base tetraphenyl-porphyrins in nonaqueous media. J. Porphyr. Phthalocyanines 2013, 17, 941–953. 10.1142/S1088424613500557. [DOI] [Google Scholar]
- Einrem R. F.; Jonsson E. T.; Teat S. J.; Settineri N. S.; Alemayehu A. B.; Ghosh A. Regioselective Formylation of Rhenium-Oxo and Gold Corroles: Substituent Effects on Optical Spectra and Redox Potentials. RSC Adv. 2021, 11, 34086–34094. 10.1039/D1RA05525A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yadav I.; Osterloh W. R.; Kadish K. M.; Sankar M. Synthesis, Spectral, Redox, and Sensing Studies of β-Dicyanovinyl-Appended Corroles and Their Metal Complexes. Inorg. Chem. 2023, 62, 7738–7752. 10.1021/acs.inorgchem.3c00341. [DOI] [PubMed] [Google Scholar]
- Yadav I.; Sankar M. Panchromatic and Perturbed Absorption Spectral Features and Multiredox Properties of Dicyanovinyl- and Dicyanobutadienyl-Appended Cobalt Corroles. Inorg. Chem. 2023, 62, 19956–19970. 10.1021/acs.inorgchem.3c02666. [DOI] [PubMed] [Google Scholar]
- Alemayehu A. B.; Gonzalez E.; Hansen L. K.; Ghosh A. Copper corroles are inherently saddled. Inorg. Chem. 2009, 48, 7794–7799. 10.1021/ic900744v. [DOI] [PubMed] [Google Scholar]
- Alemayehu A. B.; Hansen L. K.; Ghosh A. Nonplanar, noninnocent, and chiral: a strongly saddled metallocorrole. Inorg. Chem. 2010, 49, 7608–7610. 10.1021/ic1008736. [DOI] [PubMed] [Google Scholar]
- Thomas K. E.; Wasbotten I. H.; Ghosh A. Copper β-octakis(trifluoromethyl)corroles: new paradigms for ligand substituent effects in transition metal complexes. Inorg. Chem. 2008, 47, 10469–10478. 10.1021/ic801101k. [DOI] [PubMed] [Google Scholar]
- Thomas K. E.; Beavers C. M.; Ghosh A. Molecular structure of a gold β-octakis (trifluoromethyl)-meso-triarylcorrole: an 85 difference in saddling dihedral relative to copper. Mol. Phys. 2012, 110, 2439–2444. 10.1080/00268976.2012.695031. [DOI] [Google Scholar]
- Lim H.; Thomas K. E.; Hedman B.; Hodgson K. O.; Ghosh A.; Solomon E. I. X-ray absorption spectroscopy as a probe of ligand noninnocence in metallocorroles: the case of copper corroles. Inorg. Chem. 2019, 58, 6722–6730. 10.1021/acs.inorgchem.9b00128. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thomas K. E.; Settineri N. S.; Teat S. J.; Steene E.; Ghosh A. Molecular Structure of Copper and μ-Oxodiiron Octafluorocorrole Derivatives: Insights into Ligand Noninnocence. ACS Omega 2020, 5, 10176–10182. 10.1021/acsomega.0c01035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ou Z.; Shao J.; Zhao H.; Ohkubo K.; Wasbotten I. H.; Fukuzumi S.; Ghosh A.; Kadish K. M. Spectroelectrochemical and ESR studies of highly substituted copper corroles. J. Porphyr. Phthalocyanines 2004, 8, 1236–1247. 10.1142/S1088424604000593. [DOI] [Google Scholar]
- Thomas K. E.; Vazquez-Lima H.; Fang Y.; Song Y.; Gagnon K. J.; Beavers C. M.; Kadish K. M.; Ghosh A. Ligand noninnocence in coinage metal corroles: a silver knife-edge. Chem.—Eur. J. 2015, 21, 16839–16847. 10.1002/chem.201502150. [DOI] [PubMed] [Google Scholar]
- Li B.; Ou Z.; Meng D.; Tang J.; Fang Y.; Liu R.; Kadish K. M. Cobalt Triarylcorroles Containing One, Two or Three Nitro Groups. Effect of NO2 Substitution on Electrochemical Properties and Catalytic Activity for Reduction of Molecular Oxygen in Acid Media. J. Inorg. Biochem. 2014, 136, 130–139. 10.1016/j.jinorgbio.2013.12.014. [DOI] [PubMed] [Google Scholar]
- Charge-neutral meso-nitrophenyl-appended porphyrins do not exhibit pronounced hyperporphyrin spectra:; Fang Y.; Jiang X.; Ou Z.; Michelin C.; Desbois N.; Gros C. P.; Kadish K. M. Redox properties of nitrophenylporphyrins and electrosynthesis of nitrophenyl-linked Zn porphyrin dimers or arrays. J. Porphyr. Phthalocyanines 2014, 18, 832–841. 10.1142/S1088424614500540. [DOI] [Google Scholar]
- LeCours S. M.; Phillips C. M.; de Paula J. C.; Therien M. J. Synthesis, Transient Absorption, and Transient Resonance Raman Spectroscopy of Novel Electron Donor-Acceptor Complexes: [5,15-Bis[(4′-nitrophenyl)ethynyl]-10, 20-diphenylporphinato]copper(II) and [5-[[4′-(dimethylamino)phenyl]ethynyl]-15-[(4′-nitrophenyl)ethynyl]-10, 20-diphenylporphinato]copper(II). J. Am. Chem. Soc. 1997, 119, 12578–12589. 10.1021/ja964436j. [DOI] [PubMed] [Google Scholar]
- Parkin G. Valence, Oxidation Number, and Formal Charge: Three Related but Fundamentally Different Concepts. J. Chem. Educ. 2006, 83, 791–799. 10.1021/ed083p791. [DOI] [Google Scholar]
- Pariyar A.; Bose S.; Chhetri S. S.; Biswas A. N.; Bandyopadhyay P. Fluorescence signaling systems for sensing Hg(II) ion derived from A2B-corroles. Dalton Trans. 2012, 41, 3826–3831. 10.1039/c2dt12158a. [DOI] [PubMed] [Google Scholar]
- Kangwanwong T.; Pluempanupat W.; Parasuk W.; Keenan H. E.; Songsasen A. Using 5,10,15,20-tetra(4-nitrophenyl)porphyrin as a fluorescent chemosensor to determine Ru3+. ScienceAsia 2012, 38, 278–282. 10.2306/scienceasia1513-1874.2022.48.278. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




