Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2025 Jun 9;64(24):12242–12253. doi: 10.1021/acs.inorgchem.5c01593

Rhenium-Selenido Corroles: Reflections on 5d Metalloporphyrins and Metallocorroles as Triplet Emitters and Photosensitizers

Abraham B Alemayehu , Jeanet Conradie †,, Simon Larsen , Bjørn Cicerôn Lukas Pérez , Nicholas S Settineri §, Abhik Ghosh †,*
PMCID: PMC12188564  PMID: 40489243

Abstract

Following the successful synthesis of rhenium-oxido, rhenium-imido, and rhenium-sulfido corroles, a series of para-X-substituted rhenium-selenido triarylcorroles (X = OCH3, CH3, H, F, CF3) have been prepared in 70–76% yields. A one-pot, two-step procedure was used, analogous to that used for ReS corroles, involving rhenium insertion via high-temperature reaction of Re2(CO)10 and free-base corroles in refluxing decalin, followed by exposure to PCl3 (which is thought to deoxygenate ReO corroles formed under the reaction) and elemental selenium. An analogous procedure, however, failed to yield rhenium-tellurido corroles, presumably reflecting, according to DFT calculations, the limited stability of these species. Unlike ReO corroles, ReS and ReSe corroles were found not to exhibit phosphorescence in the NIR-I regime (600–1000 nm); nor did they sensitize singlet oxygen formation. In the hope of obtaining a broader perspective of this negative result, we also examined ReN porphyrins and found them not to phosphoresce or to sensitize singlet oxygen formation. These results were explained by DFT and TDDFT calculations in terms of low-energy triplet states, which are not energetic enough to excite molecular oxygen to its lowest singlet state. Whether some of the new Re corroles exhibit phosphorescence in the NIR-II regime remains an interesting question for future investigation .


graphic file with name ic5c01593_0012.jpg


graphic file with name ic5c01593_0010.jpg

Introduction

The interactions of porphyrin-type ligands and rhenium have been studied for over a half-century. , In a striking, early discovery, Tsutsui and coworkers reported that porphyrins can act as binucleating ligands toward rhenium, yielding binuclear [Por]­[Re­(CO)3]2 complexes, in which the two rhenium atoms straddle opposite faces of the porphyrin, with each metal coordinated to three porphyrin nitrogens and two of the porphyrin nitrogens coordinated to both metals. , Key examples of other coordination motifs involving porphyrins include six-coordinate rhenium-oxido and five-coordinate rhenium-nitrido porphyrins. A striking rhenium-porphyrin interaction was reported in 1998, in which attempted rhenium insertion into meso-tetrakis­(trifluoromethyl)­porphyrin resulted in an unexpected ring contraction affording ReVO meso-tris­(trifluoromethyl)­corrole. The latter compound was the first example of a corrole derivative of a 5d transition metal; today, many such complexes are known and recognized for their luminescence properties and potential application in photomedicine. Subsequently, with the availability of one-pot corrole syntheses, ReVO corroles could be synthesized simply and in high yields via the interaction with free-base meso-triarylcorroles and Re2(CO)10 in a high-boiling solvent. Tinkering with the reaction conditions (especially the temperature) led to additional coordination motifs, including metal–metal quadruple-bonded rhenium corrole dimers , and rhenium biscorrole sandwich compounds. Still other coordination motifs were obtained by the introduction of additional reagents into the reaction medium; key examples include rhenium-imido and rhenium-sulfido corroles. In this study, we investigated the effect of selenium and tellurium-based additives and the successful synthesis and characterization of rhenium-selenido corroles. The latter constitute rare examples of stable, structurally characterized terminal rhenium-selenido complexes.

The availability of a wide range of rhenium­(V) porphyrin and corrole derivatives encouraged us to undertake a comparative photophysical study of ReO, Re-imido, ReS, and ReSe corroles and ReN porphyrins. Unfortunately, unlike ReO corroles (which exhibit near-infrared phosphorescence at room temperature, singlet oxygen sensitization, and photocytotoxicity toward multiple cell lines, , ReS and ReSe corroles and ReN porphyrins proved nonemissive in the 600–1000 nm range and also incapable of sensitizing singlet oxygen formation. A ground state and time-dependent density functional theory (DFT and TDDFT) analysis of a selection of metalloporphyrins and metallocorroles, however, yielded deep insights into their widely varying photophysical behavior. These insights are likely to be a valuable aid in the future design of new triplet emitters and photosensitizers (Chart ).

1. Complexes Synthesized As Part of This Work.

1

Results and Discussion

Discovery and Optimization of Synthetic Methods

In a recent paper reporting the synthesis of stable rhenium-sulfido corroles, we presented DFT-based thermochemical arguments in favor of similarly stable rhenium-selenido corroles. That prediction has now been realized via a simple, one-pot, two-step synthesis of ReSe corroles. Rhenium insertion was first accomplished (typically an overnight process) via the high-temperature interaction of a free-base meso-tris­( p -X-phenyl)corrole, H3[TpXPC] (X = OCH3, CH3, H, F, CF3; Chart 1), Re2(CO)10, and potassium carbonate in refluxing decalin (∼180 °C). In the next step, phosphorus trichloride and elemental selenium powder was introduced into the reaction mixture, and the reaction was continued for another ∼ 4 h, at the end of which Re­[TpXPC]­(Se) complexes were isolated as air-stable solids in yields of 70.0–75.4%.

High-resolution mass spectra, 1H NMR spectroscopy (Figure ) and three single-crystal X-ray diffraction analyses (Table and Figure ) provided unambiguous proof of composition and structure of the new compounds. To freeze out interconversion of the meso-aryl o,ó and m,ḿ peaks, fully assigned 1H NMR spectra were recorded at 243 K, by now a standard practice for square-pyramidal metallocorroles. Given the rarity of terminal rhenium-selenido complexes (in contrast to molybdenum- and tungsten-selenido complexes, which are well-represented in the literature and in the Cambridge Structural Database), the Re–Se distances were of considerable interest for us. The observed distances, 2.1881(5)-2.2157(5) Å, are slightly shorter than the value, 2.2668(9) Å, obtained for a tetrahedral Re­[β-diketiminato]­(NPh)­(Se) complex. Interestingly, the Re–Se distances observed for our complexes are about halfway between the sums of Pyykkö’s covalent radii for double and triple bonds. (The double and triple bond radii for Re are 1.19 and 1.10 Å, respectively; for Se, the values are, each, 1.07 Å. , The observation suggests that the Re–Se bond has substantial, but not quite full, triple bond character. Natural bond orbital (NBO) analyses of ReCh corroles (Ch = O, S, Se, i.e., a chalcogen) lend support to such a conclusion. As shown in Figure , while the occupancies of Re-Ch π -NBOs are roughly independent of the chalcogen, the corresponding π* occupancies are significantly higher for the heavier chalcogens (relative to oxygen), consistent with a lower overall bond order. Other aspects of the structures, such as the Re–N distances (involving the corrole nitrogens) and the displacement of the Re from mean plane of the corrole nitrogens, are relatively unremarkable and do not appear to merit comment.

1.

1

1H NMR spectra of (a) Re­[TpCH3PC]­(Se) and (b) Re­[TpOCH3PC]­(Se) in CD2Cl2 at 243 K.

1. Crystal Data and Structure Refinement Parameters.

Compound Re[TpOCH3PC](Se) Re[TpCH3PC](Se) Re[TpCF3PC](Se)
CCDC deposition number 2435739 2435738 2435737
Method SC-XRD SC-XRD SC-XRD
Chemical formula C40 H29 N4 O3Se Re C40 H29 N4 Se Re C40 H20 F9 N4 Se Re
Formula weight 878.83 830.83 992.76
Crystal system Orthorhombic Triclinic Monoclinic
Crystal dimensions 0.060 × 0.060 × 0.050 0.080 × 0.040 × 0.010 0.060 × 0.020 × 0.010
Space group Pccn P-1 P21/c
λ (Å) 0.7288 0.7288 0.7288
a (Å) 26.194(2) 13.1157(7) 16.1470(12)
b (Å) 15.2424(13) 15.9836(8) 15.2888(11)
c (Å) 16.4190(13) 16.7850(9) 14.0172(10)
α (°) 90 66.324(3) 90
β(°) 90 87.379(3) 94.796(3)
γ (°) 90 80.778(3) 90
Z 8 4 4
V3) 6555.6(10) 3180.2(3) 3448.3(4)
Temperature (K) 100(2) 100(2) 100(2)
Density (g/cm3) 1.781 1.735 1.912
Measured reflections 186678 120210 99985
Unique reflections 8218 24423 8632
Parameters 445 835 537
Restraints 6 0 87
R int 0.0590 0.0524 0.0683
θ range (°) 1.585 – 29.205 1.359 – 34.301 1.298 – 29.190
R1 [I ≥ 2σ(I)], wR 2 [all data] 0.0627, 0.1582 0.0551, 0.0964 0.0449, 0.0954
S (GooF) all data 1.073 1.056 1.057
Max/min residue (e/Å3) 3.655 /-2.032 4.301 /-5.454 1.339/-1.751

2.

2

X-ray structures (top and side views) of Re­[TpXPC]Se complexes. Selected distances (Å): (a) Re­[T p CF 3 PC]­(Se). Re1–N1 1.985(4), Re1–N2 2.011(3), Re1–N3 2.023(3), Re1–N4 1.984(4), Re1–Se1 2.1881(5). (b) Re­[T p CH 3 PC]­(Se). Re1A-N1A 1.979(3), Re1A-N2A 2.007(3), Re1A-N3A 1.998(3), Re1A-N4A 2.000(4), Re1A-Se1A 2.2157(5); 1.998(3), Re1A-N4A 2.000(4), Re1A-Se1A 2.2157(5); Re1B–N1B 1.983(3), Re1B–N2B 2.009(3), Re1B–N3B 2.002(3), Re1B–N4B 1.983(3), Re1B–Se1B 2.2094(5). (c) Re­[T p OCH 3 PC]­(Se). Re1–N1 2.007(5), Re1–N2 1.986(5), Re1–N3 2.003(5), Re1–N4 2.006(6), Re1–Se1 2.2101(7). The thermal ellipsoids correspond to 50% probability. Positional disorder and solvent molecules have been omitted for clarity.

3.

3

Re-Ch (Ch = O, Se) π and π * NBO occupancies based on OLYP-D3/ZORA-STO-TZ2P calculations. (Note that the x direction is parallel to the direct pyrrole–pyrrole linkage.).

It may be of some interest to view the present syntheses through a geochemical lens. Although inorganic chemists often think of rhenium as an oxophilic element, rhenium occurs as a trace constituent of both oxide minerals such as columbite, (Fe,Mn)­(Ta,Nb)2O6, and the sulfide mineral molybdenite, MoS2. Indeed a recently developed, quantitative scale of thiophilicity (based on element-chalcogen bond energies) ranks rhenium as equally oxophilic and thiophilic. The successful synthesis of stable ReO, ReS, and ReSe corroles appears eminently consonant with such a view.

Electrochemical, Optical, and Photophysical Studies

Cyclic voltammetry, UV–vis absorption spectroscopy and photophysical studies yielded a fascinating system of electronic-structural insights (Table ). On the whole, the results closely parallel those obtained for ReS corroles.

2. Electronic Absorption Maxima (λ max, Nm) and Redox Potentials (V vs SCE) of Re­[TpXPC]­(Y), for Y = Se, S, O, and NPh.

Compound λmax E 1/2(ox2) E 1/2(ox1) E 1/2(red1) E 1/2(red2) ΔE ref
Re[TpCF3PC](Se) 280, 344, 393, 471, 577 1.58 1.14 –1.16 –1.69 2.30 This work
Re[TpFPC](Se) 280, 344, 391, 471, 574 1.45 0.98 –1.17 –1.70 2.15
Re[TPC](Se) 282, 345, 393, 430, 577 1.44 0.95 –1.20 –1.74 2.15
Re[TpCH3PC](Se) 280, 343, 393, 438, 577 1.43 0.94 –1.21 –1.77 2.15
Re[TpOCH3PC](Se) 283, 346, 393, 439, 578 1.32 0.91 –1.23 –1.79 2.14
Re[TpCF3PC](S) 276, 340, 393, 459, 571 1.59 1.07 –1.15 –1.68 2.22
Re[TpFPC](S) 276, 341, 391, 459, 571 1.54 1.00 –1.22 –1.79 2.22
Re[TPC](S) 277, 339, 391, 459, 572 1.52 0.96 –1.23 –1.80 2.19
Re[TpCH3PC](S) 276, 339, 392, 459, 575 1.46 0.92 –1.25 –1.84 2.17
Re[TpOCH3PC](S) 273, 344, 394, 459, 574 1.35 0.89 –1.27 –1.85 2.16
Re[TpCF3PC](O) 438, 585 - 1.10 –1.16 - 2.26 24
Re[TpFPC](O) 438, 585 - 1.01 –1.23 - 2.24
Re[TPC](O) 439, 585 - 0.98 –1.26 - 2.24
Re[TpCH3PC](O) 440, 587 - 0.94 –1.29 - 2.23
Re[TpOCH3PC](O) 441, 592 - 0.93 –1.29 - 2.22
Re[TpCF3PC](NPh) 434, 577 1.24 0.97 –1.29 - 2.26 31
Re[TpFPC](NPh) 434, 575 1.15 0.88 –1.36 - 2.24
Re[TPC](NPh) 434, 576 1.18 0.86 –1.38 - 2.24
Re[TpCH3PC](NPh) 434, 578 1.12 0.82 –1.40 - 2.22
Re[TpOCH3PC](NPh) 435, 578 1.03 0.78 –1.41 - 2.19

Cyclic voltammetry revealed two reversible oxidations and at least one reversible reduction (as well as second quasi-reversible reduction) for each ReSe corrole (Figure ). The complexes were found to exhibit relatively high first oxidation potentials, 0.89 to 1.07 V, and relatively low first reduction potentials, −1.27 to −1.15 V, vs the SCE, as might be expected for an electronically innocent corrole macrocycle with a high-valent central metal ion. The electrochemical HOMO–LUMO gap (i.e., the algebraic difference between the first oxidation and reduction potentials) of 2.2 V is also essentially the same as that observed for ReO, ReS, OsN, Au , and other electronically innocent metallocorroles. This HOMO–LUMO gap is also consistent with that inferred from the lowest-energy optical absorption maxima of the compounds (Table and Figure ). These observations might be naively interpreted as indicative of a purely ligand-based HOMO and a purely ligand-based LUMO, as expected from Gouterman’s four-orbital model. [According to this model, the two π HOMOs of a porphyrin are near-degenerate, as are the two LUMOs, and these four MOs are energetically well-separated from all other occupied and unoccupied MOs; although originally formulated for porphyrins, the model has been found to hold for many simple corroles such as Ga and Au corroles. , We shall see that such a conclusion would be entirely fallacious: the valence electronic structures of ReS and ReSe corroles are utterly at odds with the four-orbital model. Indeed, the broad, smeared-out features in the 370–650 nm region of the optical spectra of ReS and ReSe corroles already hint at a distinctly non-Gouterman-type MO architecture.

4.

4

Cyclic voltammograms of Re­[TpXPC]­(Se) in 0.1 M solutions of tetrabutylammonium perchlorate in anhydrous dichloromethane; scan rate = 100 mV.s–1.

5.

5

UV–vis spectra in dichloromethane: (a) the Re­[TpXPC]­(Se) series; (b) Re­[TPC]­(Ch) (Ch = O, S, Se). Sample concentrations were in the range 4.0 ± 0.5 mM. See Table for a listing of peak maxima.

Additional electronic-structural clues came from photophysical studies. Since ReO corroles were previously found to exhibit moderate NIR phosphorescence and to efficiently generate singlet oxygen, , the potential emissive properties of the ReS and ReSe corroles were investigated. Somewhat disappointingly, both higher- and lower-energy excitation of their toluene solutions failed to elicit any emission in the 600–1000 nm range. Since emission properties (particularly phosphorescence) are often significantly enhanced at low temperatures, the measurements were repeated at 77K in toluene:tetrahydrofuran (2:3 v/v) frozen glass, but to no avail. Finally, we tested both ReS and ReSe corroles for singlet oxygen sensitization in ethanol using 9,10-dimethylanthracene as a trap (with excitation at 570 nm). None was observed.

Given that many iridium porphyrins exhibit intense phosphorescence, relative to iridium corroles, which are only weakly phosphorescent, we chose to examine rhenium-nitrido porphyrins. Several were prepared according to literature procedures, but only two, based on electron rich meso-tetrakis­(p-X-phenyl)­porphyrins (X = Me, OMe) proved stable enough for photophysical studies. A number of Re­(O)­(Cl) porphyrins were also prepared, but these too proved rather unstable, precluding reliable photophysical measurements. , Disappointingly, unlike their Ir counterparts, the ReN porphyrins did not exhibit phosphorescence (in the 600–1000 nm range) or singlet oxygen sensitization.

DFT and TDDFT Calculations

To make sense of the wide range of photophysical properties of rhenium porphyrins and corroles, and of 5d metalloporphyrins − ,− and metallocorroles − ,,− ,− in general, we undertook a DFT and TDDFT survey of over a dozen complexes. The B3LYP* Kohn–Sham MO energy level diagrams (Table , Figures , and S11–S13) alone provided certain insights into why phosphorescence and singlet-oxygen sensitization are observed for certain species, but not for others. Thus, an unusually low HOMO–LUMO gap was found for unsubstituted ReN porphyrin, suggesting that the necessarily low-energy triplet state would not phosphoresce in our spectral window (600–1000 nm) or excite molecular oxygen to its lowest singlet state (which has an energy of 1.2 eV above the triplet ground state). Second, the MO architecture of ReS and ReSe corrole exhibits little resemblance to Gouterman’s four-orbital model; the LUMOs of these complexes are substantially metal-based and profoundly different from those of ReO corrole (Figure ), which qualitatively accounts for the dramatic differences in UV–vis spectral profile among different Re-chalcogenido corroles (Figure b).

3. Calculated Triplet Energies (eV) from Regular Scalar-Relativistic (SR) and Time-Dependent (TD) Scalar- (SR) and Spin-Orbit (SO) Relativistic B3LYP* Calculations.

  SR-B3LYP*
TD-SR-B3LYP*
TD-SO-B3LYP*
  E T1 E T1 E S1 Lowest excitation energy f
Porphyrins          
Re[Por](N) 0.83 0.55 0.58 0.56 3.18 × 10–8
Re[Por](O)(F) 1.25 1.19 1.45 1.19 2.93 × 10–7
Ir[Por](Me) 1.64 1.55 1.77 1.46 1.02 × 10–5
Pd[Por] 1.84 1.79 2.46 1.79 1.43 × 10–8
Pt[Por] 1.93 1.89 2.35 1.89 2.35 × 10–7
Corroles          
Re[Cor](O) 1.58 1.54 2.06 1.54 6.90 × 10–8
Re[Cor](S) 0.98 0.98 1.38 0.99 5.69 × 10–7
Re[Cor](Se) 0.88 0.87 1.28 0.87 5.42 × 10–7
Ru[Cor](N) 1.50 1.46 1.82 1.46 3.63 × 10–7
Os[Cor](N) 1.56 1.52 2.28 1.52 1.60 × 10–9
Ir[Cor](py)2 1.51 1.47 2.10 1.47 1.64 × 10–6
Pt[Cor](Ph)(py) 1.51 1.47 2.25 1.47 1.90 × 10–8
Au[Cor] 1.57 1.52 2.42 1.52 1.46 × 10–8
a

These triplet energies are adiabatic singlet–triplet gaps and were obtained from single point scalar-relativistic B3LYP* calculations on symmetry-unconstrained OLYP-D3 optimized geometries for M S = 0 and 1.

b

Both scalar-relativistic and spin–orbit time-dependent B3LYP* calculations were carried out on symmetry-unconstrained OLYP-D3 optimized geometries for M S = 1. The lowest excitation energies thus correspond to phosphorescence from a geometry-relaxed triplet state.

c

The spin–orbit relativistic oscillator strengths f are expected to be proportional to phosphorescence quantum yields.

6.

6

B3LYP*/STO-ZORA-TZ2P MO energy level diagrams of Re porphyrins and corroles based on OLYP-D3 ground-state optimized geometries. Cor and Por refer to unsubstituted corrolato­(3-) and porphyrinato­(2-) ligands, respectively. Doubly occupied and unoccupied MO energy levels are indicated in blue and red, respectively.

7.

7

B3LYP*/STO-ZORA-TZ2P frontier MOs (with a surface isovalue of 0.05 e/Å3) of Re porphyrins and corroles based on OLYP-D3 ground-state optimized geometries.

To explain the observed photophysical behavior of a wider range of complexes (Table ), we chose two computational approaches. In the first, ground-state DFT methods were used to calculate single-point “singlet” and “triplet” energies at the “triplet” geometry (obtained with scalar-relativistic OLYP-D3 calculations). Given the longer time scale of phosphorescence, the singlet–triplet gap thus obtained at a relaxed “triplet” geometry difference corresponds to the phosphorescence energy. (Note that we are using the expressions “singlet” and “triplet” here in a loose sense; more accurately, we carried out geometry optimizations for M S = 0 and 1.) In the second approach, TDDFT calculations were carried out on singlet reference states with optimized triplet geometries. Again, the triplet energies thus obtained correspond to phosphorescence energies. Both scalar-relativistic and spin–orbit ZORA Hamiltonians were used in the TDDFT calculations. Reassuringly, the triplet state energetics obtained with the different methods (Table ) proved largely mutually consistent. To our satisfaction, the results also went a long way toward rationalizing the observation or otherwise of phosphorescence and singlet oxygen sensitization by key species.

Table indicates substantial HOMO–LUMO gaps for ReO, , OsN, Ir, PtPh and Au corrole as well as for Pd, Pt, and IrMe porphyrin, consistent with the observation of NIR phosphorescence and singlet oxygen sensitization for substituted analogues of these complexes. In contrast, dramatically lower triplet energies are predicted for ReS and ReSe corrole, as well as for ReN porphyrin, explaining their lack of phosphorescence within our spectral window, as well as their lack of singlet oxygen activity. Given that ReS and ReSe corroles exhibit relatively normal optical and electrochemical HOMO–LUMO gaps (see preceding section and Table ), the low triplet energies can only be attributed to strong orbital relaxation effects in the triplet states. Strong evidence for such a scenario comes from scalar-relativistic B3LYP* spin density profiles of the ReS and ReSe corrole (Figure ), which are found to be largely concentrated on the Re-chalcogenido units, even though the LUMOs of the singlet states are substantially corrole-based (Figure ).

8.

8

B3LYP*/STO-ZORA-TZ2P triplet spin density profiles (with a surface isovalue of 0.004 e/Å3) based on symmetry-unconstrained OLYP-D3 optimized geometries for M S = 1.

Note that spin–orbit TDDFT calculations also yield oscillator strengths that may be expected to correlate with observed phosphorescence quantum yields (Table ). (Since phosphorescence is an inherently spin–orbit coupling-based phenomenon, such oscillator strengths cannot be obtained scalar-relativistic TDDFT calculations.) In fact, only a qualitative correlation is observed and that too among isostructural molecules. Thus, the calculations correctly predict a higher oscillator strength for Pt­[Por] relative to Pd­[Por], consistent with observed phosphorescence quantum yields for Pt and Pd porphyrins. In the same vein, the calculations predict a higher oscillator strength for IrMe porphyrin relative to Ir­(py)2 corrole, again consistent with experimental results. For each of these pairs of complexes, the higher oscillator strength appears to be associated with a higher degree of metal character in the LUMO or in the triplet spin density, which should lead to a greater impact of spin–orbit coupling on the rate of phosphorescent emission. Certain TDDFT results, however, do not appear to correlate with experimental findings. Thus, unusually high phosphorescence oscillator strengths are predicted for the two Ir complexes studied, relative to most of the other metalloporphyrins and metallocorroles, which appears to be a somewhat unphysical result. It will be interesting to see whether methodological improvements yield more accurate phosphorescence oscillator strengths in the foreseeable future.

Attempted Synthesis of Re-tellurido Corroles

Intrigued by reports of several tungsten-tellurido complexes, we also attempted to synthesize rhenium-tellurido corroles. Toward that end, we substituted elemental tellurium powder for selenium in the second step of the synthesis. UV–vis spectroscopy did not yield any indication of novel species; high-resolution mass spectrometry also did not indicate the formation of a ReTe corrole. Substituting TeCl4 for tellurium powder, and using an equimolar mixture of elemental tellurium and TeCl4 (which are known to react to yield TeCl2), proved similarly fruitless. These negative results may suggest that ReTe corroles are thermodynamically less stable than their lighter-chalcogen congeners (in line with tellurium’s reputation as “a maverick among the chalcogens”, a proposition that we chose to test with DFT calculations.

We accordingly revisited our earlier OLYP-D3/ZORA-STO-TZ2P calculations on the energetics of the following reactions, where {Re­[Cor]}2 refers to metal–metal quadruple-bonded, unsubstituted rhenium corrole dimer.

Re[Cor](O)+Me3P=1/2{Re[Cor]}2+Me3PO;ΔE=0.59eV
Re[Cor](S)+Me3P=1/2{Re[Cor]}2+Me3PS;ΔE=0.54eV
Re[Cor](Se)+Me3P=1/2{Re[Cor]}2+Me3PSe;ΔE=0.48eV
Re[Cor](Te)+Me3P=1/2{Re[Cor]}2+Me3PTe;ΔE=0.58eV

As shown above, the reaction energies (not corrected for zero-point energies) appear to be roughly independent of the chalcogen. The result, however, does not imply that the Re-chalcogenido corroles are all equally stable, but rather that their stabilities track those of the corresponding phosphine chalcogenides. The latter is a most valuable clue: while phosphine oxides, sulfides, and selenides are stable compounds, , phosphine tellurides are known to be thermodynamically unstable (even though a handful have been reported). At this point, our failure to generate ReTe corroles may reflect their limited stability. That said, we have far from exhausted all reasonable avenues that could lead to the successful isolation of ReTe corroles.

Concluding Remarks

The main conclusions of this study are as follows.

1. The high-temperature insertion of rhenium into free-base meso-triarylcorroles, followed by exposure to elemental selenium, has yielded a series of stable rhenium-selenido corroles in 70–75% yields. The synthesis is similar to that reported recently for rhenium-sulfido corroles. The stability of ReO, ReS, and ReSe corroles is a testament to rhenium’s chalcophilic character, specifically its unique status as a metal that is almost equally oxophilic and thiophilic. Rhenium-tellurido corroles, however, resisted synthesis, a reflection, potentialy, of their limited stability.

2. Unlike ReO corroles, which exhibit room-temperature phosphorescence and efficiently sensitize singlet oxygen formation, ReS and ReSe corroles proved nonemissive in the NIR-I regime and also did not exhibit singlet oxygen formation; a similar negative result was also obtained for ReN porphyrins. Whether some of these complexes might emit in the NIR-II regime remains an interesting question for the future.

3. Based on a relatively wide-ranging DFT and TDDFT survey of 4d and 5d metalloporphyrins and metallocorroles, the lack of near-infrared (<1000 nm) phosphorescence and singlet oxygen activity of ReS and ReSe corroles, and of ReN porphyrins, may be attributed to low triplet-state energies that are insufficient for singlet oxygen generation. Such calculations are likely to aid in the design of new triplet emitters and photosensitizers.

Experimental Section

Materials and Methods

The majority of chemicals were purchased from Merck. Free-base triarylcorroles were prepared according to literature procedures. , Rhenium-nitride porphyrins studied were prepared according to literature methods in two steps. The first step involved refluxing trichlorobenzene solutions of free-base porphyrins and ReCl5 to give rhenium-oxido porphyrins. The second step involved heating chloroform-ethanol solutions of rhenium-oxido porphyrins and hydrazine hydrate to yield rhenium-nitrido porphyrins.

UV–visible-NIR spectra were recorded on a Cary 8454 spectrophotometer. 1H NMR spectra were recorded on a 400 MHz Bruker Avance III HD spectrometer equipped with a 5 mm BB/1H SmartProbe at 298 K in CDCl3 and 243 K in CD2Cl2 and referenced to residual CHCl3 at 7.26 ppm and CH2Cl2 at 5.31 ppm. High-resolution electrospray ionization mass spectra (HR-ESI-MS) were recorded on an LTQ Orbitrap XL spectrometer in the positive mode.

Cyclic voltammetry was carried out at ambient temperature with a Gamry Reference 620 potentiostat equipped with a three-electrode system: a 3 mm disk glassy carbon working electrode, a platinum wire counter electrode, and a saturated calomel reference electrode (SCE).Tetra­(n-butyl)­ammonium perchlorate was used as the supporting electrolyte. Anhydrous CH2Cl 2 (Aldrich) was used as the solvent. The electrolyte solution was purged with argon for at least 2 min prior to all measurements, which were carried out under an argon blanket. The glassy carbon working electrode was polished using a polishing pad and 0.05-μm polishing alumina from ALS, Japan. All potentials were referenced to the SCE.

Re­[TpXPC]­(Se)

To a 50 mL two-necked round-bottom flask fitted with a reflux condenser and containing decalin (15 mL) and a magnetic stirring bar were added a free-base corrole, H3[TpXPC] (0.17 mmol), Re2(CO)10 (221.8 mg, 0.34 mmol), and potassium carbonate (140 mg). The contents were deoxygenated with a flow of argon and then refluxed overnight with constant stirring under argon. Phosphorus trichloride (148.7 μL, 10 equiv) and elemental selenium, powder (13.4 mg, 5 equiv) were then added and the reaction was continued under reflux (i.e., at ∼ 180 °C) for ∼ 4 h. The color of the reaction slowly turned to brown and completion of the reaction was monitored by UV–vis spectroscopy and mass spectrometry. Upon cooling, the reaction mixture was loaded directly on to a silica gel column with n-heptane as the mobile phase. The decalin was first removed by eluting with pure n-heptane. Different solvent mixtures were then used to elute the various Re­[TpXPC]­(Se) derivatives: 3:1 n-heptane/dichloromethane for X = CF3, H, CH3 and F and 1:1 n-heptane/dichloromethane for X = OCH3. All fractions with λmax ∼ 393 nm were collected and evaporated to dryness. The products were further purified with a second round of column chromatography and finally with preparative thin-layer chromatography, all with the same solvent system as in the first round. Yields and analytical details for the different complexes are given below. Note that the NMR assignments (o1, o2) and (m1, m2) refer to symmetry-distinct (diastereotopic) ortho and meta protons for a given aryl group.

Re­[TpCF3PC]­(Se)

Yield 123.2 mg (0.124 mmol, 72.9%). UV–vis (CH2Cl2) λmax [nm, ε x 10–4 (M–1cm–1)]: 280 (5.25), 344 (3.20), 393 (4.82), 471 (2.77), 577 (1.68). 1H NMR (400 MHz, CD2Cl2, – 30 °C): δ 9.64 (d, 2H, 3 J HH = 4.4 Hz, β-H); 9.31 (d, 2H, 3 J HH = 4.6 Hz, β-H); 9.27 (d, 2H, 3 J HH = 4.9 Hz, β-H); 9.11 (d, 2H, 3 J HH = 4.6 Hz, β-H); 8.74 (d, 2H, 3 J HH = 7.8 Hz, 5,15-o1-Ph); 8.64 (d, 1H, 3 J HH = 7.8 Hz. 10-o1-Ph); 8.19 (d, 2H, 3 J HH = 8.1 Hz 5,15-m1-Ph); 8.13 (t, 3H, 3 J HH = 8.7 Hz, 5,15-o2-Ph 10-m1-Ph); 8.03 (t, 2H, 3 J HH = 8.1 Hz 5,15-Ph); 7.97 (s, 2H, 10-m2 and o2-Ph). MS (ESI): [M+H]+ = 995.0338 (expt), 995.0342 (calcd for M = C40H20N4F9SeRe).

Re­[TpFPC]­(Se)

Yield 108.3 mg (0.128 mmol, 75.4%). UV–vis (CH2Cl2): λmax (nm), [ε x 10–4 (M–1cm–1)]: 280 (5.04), 344 (3.18), 391 (4.76), 471 (2.60); 574(1.59). 1H NMR (400 MHz, CD2Cl2, – 30 °C): δ 9.63 (d, 2H, 3 J HH = 4.5 Hz, β-H); 9.31 (d, 2H, 3 J HH = 4.5 Hz, β-H); 9.28 (d, 2H, 3 J HH = 5.0 Hz, β-H); 9.11 (d, 2H, 3 J HH = 5.0 Hz, β-H); 8.58 (t, 2H, 3 J HH = 7.1 Hz, 5,15-o1-Ph); 8.49 (d, 1H, 3 J HH = 7.1 Hz, 10-o1-Ph); 7.97 (t, 2H, 3 J HH = 7.4 Hz, 5,15-o2-Ph); 7.80 (t, 2H, 3 J HH = 7.2 Hz, 10-o2-Ph); 7.67 – 7.55 (m, 3H, 5,10,15-m1-Ph); 7.48 (td, 2H, 3 J HH = 8.8, 2.6 Hz, 10-m2-Ph); 7.41 (dt, 1H, 3 J HH= 8.7, 4.9 Hz, 10-m2-Ph). MS (ESI): [M+H]+ = 845.043 (expt), 845.0437 (calcd for M = C37H20F3N4SeRe).

Re­(TPC)­(Se)

Yield 96.4 mg (0.122 mmol, 71.7%). UV–vis (CH2Cl2): λmax (nm), [ε x 10–4 (M–1cm–1)]: 282 (4.74), 345 (3.17), 393 (4.40), 430 (3.03); 577 (1.59). 1H NMR (400 MHz, CD2Cl2, – 30 °C): δ 9.67 (d, 2H, 3 J HH = 4.5 Hz, β-H); 9.36 (d, 2H, 3 J HH = 4.6 Hz, β-H); 9.31 (d, 2H, 3 J HH = 4.9 Hz, β-H); 9.13 (d, 2H, 3 J HH = 4.9 Hz, β-H); 8.63 (d, 2H, J = 7.6 Hz, 5,15-o1-Ph); 8.53 (d, 1H, J = 7.5 Hz, 10-o1-Ph); 8.04 (d, 2H, J = 6.9 Hz, 5,15-o2-Ph); 7.94 (t, 2H, J = 7.3 Hz, 5,15-m2-Ph); 7.90 – 7.77 (m, 8H, 10-o2-Ph, 5,10,15-p-Ph and 5,15-m2-Ph); 7.73 (d, 1H, J = 7.5 Hz, 10-m2-Ph). MS (ESI): [M+H]+ = 791.0718 (expt), 791.0720 (calcd for M = C37H23N4SeRe).

Re­[TpCH3PC]­(Se)

Yield 104.0 mg (0.125 mmol, 73.5%). UV–vis (CH2Cl2): λmax (nm), [ε x 10–4 (M–1cm–1)]: 280 (4.74), 343 (2.97), 339 (2.86), 393 (4.41), 438 (3.29); 577 (1.51). 1H NMR (400 MHz, CD2Cl2, – 30 °C): δ 9.64 (d, 2H, 3 J HH = 4.5 Hz, β-H); 9.34 (d, 2H, 3 J HH = 4.4 Hz, β-H); 9.30 (d, 2H, 3 J HH = 4.9 Hz, β-H); 9.13 (d, 2H, 3 J HH = 4.9 Hz, β-H); 8.50 (dd, 2H, 3 J HH = 7.7, 2.0 Hz, 5,15-o1-Ph); 8.39 (dd, 1H, 3 J HH = 7.6, 2.0 Hz, 10-o1-Ph); 7.92 (dd, 2H, 3 J HH = 7.6, 1.9 Hz, 5,15-o2-Ph); 7.74 (d, 3H, 3 J HH = 7.4 Hz, 5,15-m1-Ph); 7.71 (d, 1H, 3 J HH = 6.36 Hz, 10-o2-Ph); 7.68 (d, 2H, 3 J HH = 7.8 Hz, 10-m2-Ph); 7.58 (d, 2H, 3 J HH = 7.8 Hz, 5,15-m2-Ph); 7.51 (d, 1H, 3 J HH = 7.8 Hz, 10-m2-Ph); 2.68 (s, 6H, 5,15-p-CH3); 2.66 (s, 3H, 10-p-CH3). MS (ESI): [M+H]+ = 833.1185 (expt), 833.1190 (calcd for M = C40H29N4SeRe).

Re­[TpOCH3PC]­(Se)

Yield 104.7 mg (0.119 mmol, 70.0%). UV–vis (CH2Cl2): λmax (nm), [ε x 10–4 (M–1cm–1)]: 283 (5.13), 346 (3.21), 393 (4.66), 439 (3.41); 578 (1.55). 1H NMR (400 MHz, – 30 °C): δ 9.63 (d, 2H, 3 J HH = 4.5 Hz, β-H); 9.34 (d, 2H, 3 J HH = 4.4 Hz, β-H); 9.32 (d, 2H, 3 J HH = 5.0 Hz, β-H); 9.15 (d, 2H, 3 J HH = 4.9 Hz, β-H); 8.54 (dd, 2H, 3 J HH = 8.4, 2.3 Hz, 5,15-o1-Ph); 8.42 (dd, 1H, 3 J HH = 8.3, 2.3 Hz, 10-o1-Ph); 7.96 (dd, 2H, 3 J HH = 8.4, 2.3 Hz, 5,15-o2-Ph); 7.77 (dd, 1H, 3 J HH = 8.3, 2.3 Hz, 10-o2-Ph); 7.45 (dd, 2H, 3 J HH = 8.4, 2.8 Hz, 5,15-m1-Ph); 7.39 (dd, 1H, 3 J HH = 8.4, 2.7 Hz, 10-m1-Ph); 7.30 (dd, 2H, 3 J HH = 8.4, 2.7 Hz, 5,15-m2-Ph); 7.23 (d, 1H, 3 J HH = 8.4, 2.7 Hz, 10-m2-Ph); 4.06 (s, 6H, 5,15-p-OCH3); 4.04 (s, 3H, 10-p-OCH3). MS (ESI): [M+H]+ = 881.1035 (expt), 881.1038 (calcd for M = C40H29O3N4SeRe).

Sample Preparation for Crystallography

Crystals were grown by slow diffusion of methanol into concentrated solutions of the samples in dichloromethane.

Single-Crystal X-ray Diffraction Analyses

X-ray data were collected on beamline 12.2.1 at the Advanced Light Source, Lawrence Berkeley National Laboratory. Samples were mounted on MiTeGen Kapton loops and placed in a 100(2) K nitrogen cold stream provided by an Oxford Cryostream 800 Plus low-temperature apparatus on the goniometer head of a Bruker D8 diffractometer equipped with a PHOTON II CPAD detector operating in shutterless mode. Diffraction data were collected using synchrotron radiation monochromated using silicon(111) to a wavelength of 0.7288(1)­Å. An approximate full-sphere of data was collected using a combination of φ and ω scans with scan speeds of one second per degree. The structures were solved by intrinsic phasing (SHELXT and refined by full-matrix least-squares on F2 (SHELXL-2014. All non-hydrogen atoms were refined anisotropically. Hydrogen atoms were geometrically calculated and refined as riding atoms.

Photophysical Studies

The complexes in question were screened for potential emission using a Fluorolog 3 spectrometer, with a setup identical to that reported previously. Singlet oxygen assays were also performed as described in the same report.

Computational Methods

All geometry optimizations were carried out with the scalar-relativistic zeroth-order regular approximation (ZORA) Hamiltonian, OLYP , functional, Grimme’s D3 dispersion correction, ZORA TZ2P all-electron relativistic basis sets, carefully tested fine integration grids and tight criteria for the SCF cycles and geometry optimizations, as implemented in the ADF 2019 program system. Time-dependent B3LYP* calculations were carried out on OLYP optimized geometries using both scalar-relativistic (SR) and spin–orbit (SO) ZORA Hamiltonians.

Supplementary Material

ic5c01593_si_001.pdf (7.6MB, pdf)

Acknowledgments

This work was supported in part by grant no. 324139 of the Research Council of Norway (AG). Single-crystal X-ray diffraction analyses were carried out at the Advanced Light Source, which is a DOE Office of Science User Facility under contract no. DE-AC02-05CH11231. We are deeply grateful to Prof. Sergey M. Borisov (Graz University of Technology), who performed the photophysical studies reported herein.

All data generated or analyzed in this study are included in this published article and its Supporting Information.

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.5c01593.

  • Mass spectra, selected photophysical data, and optimized DFT coordinates (PDF)

The authors declare no competing financial interest.

References

  1. Alemayehu, A. ; Conradie, J. ; Larsen, S. ; Pérez, B. C. L. ; Settineri, N. S. ; Ghosh, A. . Rhenium-Selenido Corroles: Reflections on 5d Metalloporphyrins and Metallocorroles as Triplet Emitters and Photosensitizers. ChemRxiv. 2025, 10.26434/chemrxiv-2025-k12gb. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Chatterjee T., Ravikanth M.. Rhenium complexes of porphyrinoids. Coord. Chem. Rev. 2020;422:213480. doi: 10.1016/j.ccr.2020.213480. [DOI] [Google Scholar]
  3. Majumder S., Borah B. P., Bhuyan J.. Rhenium in the core of porphyrin and rhenium bound to the periphery of porphyrin: synthesis and applications. Dalton Trans. 2020;49:8419–8432. doi: 10.1039/D0DT00813C. [DOI] [PubMed] [Google Scholar]
  4. Cullen D., Meyer E., Srivastava T. S., Tsutsui M.. Unusual metalloporphyrins. XIV. Structure of [meso-tetraphenylporphinato]­bis­[tricarbonylrhenium­(I)] J. Am. Chem. Soc. 1972;94(21):7603–7605. doi: 10.1021/ja00776a069. [DOI] [Google Scholar]
  5. Ostfield D., Tsutsui M.. Novel Metalloporphyrins – Syntheses and Implications. Acc. Chem. Res. 1974;7:52–58. doi: 10.1021/ar50074a004. [DOI] [Google Scholar]
  6. Buchler, J. W. ; Dreher, C. ; Künzel, F. M. . Synthesis and coordination chemistry of noble metal porphyrins. In Metal Complexes with Tetrapyrrole Ligands III. Structure and Bonding; Springer: Berlin, Heidelberg, 1995, Vol. 84. 10.1007/BFb0111330. [DOI] [Google Scholar]
  7. Tse M. K., Zhang Z. Y., Chan K. S.. Synthesis of an Oxorhenium­(V) Corrolate from Porphyrin with Detrifluoromethylation and Ring Contraction. Chem. Commun. 1998:1199–1200. doi: 10.1039/A802033G. [DOI] [Google Scholar]
  8. Buckley H. L., Arnold J.. Recent Developments in Out-of-Plane Metallocorrole Chemistry across the Periodic Table. Dalton Trans. 2015;44:30–36. doi: 10.1039/C4DT02277G. [DOI] [PubMed] [Google Scholar]
  9. Alemayehu A. B., Vazquez-Lima H., Gagnon K. J., Ghosh A.. Tungsten Biscorroles: New Chiral Sandwich Compounds. Chem.-Eur. J. 2016;22:6914–6920. doi: 10.1002/chem.201504848. [DOI] [PubMed] [Google Scholar]
  10. Alemayehu A. B., Gagnon K. J., Terner J., Ghosh A.. Oxidative Metalation as a Route to Size-Mismatched Macrocyclic Complexes: Osmium Corroles. Angew. Chem., Int. Ed. 2014;53:14411–14414. doi: 10.1002/anie.201405890. [DOI] [PubMed] [Google Scholar]
  11. Alemayehu A. B., McCormick L. J., Vazquez-Lima H., Ghosh A.. Relativistic Effects on a Metal–Metal Bond: Osmium Corrole Dimers. Inorg. Chem. 2019;58(4):2798–2806. doi: 10.1021/acs.inorgchem.8b03391. [DOI] [PubMed] [Google Scholar]
  12. Palmer J. H., Day M. W., Wilson A. D., Henling L. M., Gross Z., Gray H. B.. Iridium corroles. J. Am. Chem. Soc. 2008;130:7786–7787. doi: 10.1021/ja801049t. [DOI] [PubMed] [Google Scholar]
  13. Alemayehu A. B., Vazquez-Lima H., Beavers C. M., Gagnon K. J., Bendix J., Ghosh A.. Platinum Corroles. Chem. Commun. 2014;50:11093–11096. doi: 10.1039/C4CC02548B. [DOI] [PubMed] [Google Scholar]
  14. Alemayehu A. B., Ghosh A.. Gold Corroles. J. Porphyrins Phthalocyanines. 2011;15:106–110. doi: 10.1142/S1088424611003045. [DOI] [Google Scholar]
  15. Rabinovitch E., Goldberg I., Gross Z.. Gold (I) and Gold­(III) Corroles. Chem.-Eur. J. 2011;17:12294–12301. doi: 10.1002/chem.201102348. [DOI] [PubMed] [Google Scholar]
  16. Thomas K. E., Alemayehu A. B., Conradie J., Beavers C., Ghosh A.. Synthesis and Molecular Structure of Gold Triarylcorroles. Inorg. Chem. 2011;50:12844–12851. doi: 10.1021/ic202023r. [DOI] [PubMed] [Google Scholar]
  17. Mahammed A., Gray H. B., Gross Z.. Silver Anniversary of the Renaissance in Metallocorrole Chemistry. Chem. Rev. 2025;125(5):2809–2845. doi: 10.1021/acs.chemrev.4c00764. [DOI] [PubMed] [Google Scholar]
  18. Teo R. D., Hwang J. Y., Termini J., Gross Z., Gray H. B.. Fighting Cancer with Corroles. Chem. Rev. 2017;117:2711–2729. doi: 10.1021/acs.chemrev.6b00400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Mahammed A., Gross Z.. Corroles as Triplet Photosensitizers. Coord. Chem. Rev. 2019;379:121–132. doi: 10.1016/j.ccr.2017.08.028. [DOI] [Google Scholar]
  20. Lemon C. M.. Corrole Photochemistry. Pure Appl. Chem. 2020;92:1901–1919. doi: 10.1515/pac-2020-0703. [DOI] [Google Scholar]
  21. Di Natale C., Gros C. P., Paolesse R.. Corroles at Work: A Small Macrocycle for Great Applications. Chem. Soc. Rev. 2022;51:1277–1335. doi: 10.1039/D1CS00662B. [DOI] [PubMed] [Google Scholar]
  22. Alemayehu A. B., Thomas K. E., Einrem R. F., Ghosh A.. The Story of 5d Metallocorroles: From Metal–Ligand Misfits to New Building Blocks for Cancer Phototherapeutics. Acc. Chem. Res. 2021;54:3095–3107. doi: 10.1021/acs.accounts.1c00290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Nardis S., Mandoj F., Stefanelli M., Paolesse R.. Metal Complexes of Corrole.Coord . Chem. Rev. 2019;388:360–405. doi: 10.1016/j.ccr.2019.02.034. [DOI] [Google Scholar]
  24. Orłowski R., Gryko D., Gryko D. T.. Synthesis of Corroles and Their Heteroanalogs. Chem. Rev. 2017;117:3102–3137. doi: 10.1021/acs.chemrev.6b00434. [DOI] [PubMed] [Google Scholar]
  25. 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. doi: 10.1002/chem.201504307. [DOI] [PubMed] [Google Scholar]
  26. Alemayehu A. B., Einrem R. F., McCormick-McPherson L. J., Settineri N. S., Ghosh A.. Synthesis and Molecular Structure of Perhalogenated Rhenium-Oxo Corroles. Sci. Rep. 2020;10(1):19727. doi: 10.1038/s41598-020-76308-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. 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. doi: 10.1039/D1RA05525A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Johannessen K. E., Johansen M. A. L., Einrem R. F., McCormick McPherson L. J., Alemayehu A. B., Borisov S. M., Ghosh A.. Influence of Fluorinated Substituents on the Near-Infrared Phosphorescence of 5d Metallocorroles. ACS Org. Inorg. Au. 2023;3:241–245. doi: 10.1021/acsorginorgau.3c00016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Alemayehu A. B., McCormick-McPherson L. J., Conradie J., Ghosh A.. Rhenium Corrole Dimers: Electrochemical Insights into the Nature of the Metal–Metal Quadruple Bond. Inorg. Chem. 2021;60(11):8315–8321. doi: 10.1021/acs.inorgchem.1c00986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Osterloh W. R., Conradie J., Alemayehu A. B., Ghosh A., Kadish K. M.. The Question of the Redox Site in Metal–Metal Multiple-Bonded Metallocorrole Dimers. ACS Org. Inorg. Au. 2023;3:35–40. doi: 10.1021/acsorginorgau.2c00030. [DOI] [Google Scholar]
  31. Alemayehu A. B., Abernathy M. J., Conradie J., Sarangi R., Ghosh A.. Rhenium Biscorrole Sandwich Compounds: XAS Evidence for a New Coordination Motif. Inorg. Chem. 2023;62:8467–8471. doi: 10.1021/acs.inorgchem.3c00632. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Alemayehu A. B., Teat S. J., Borisov S. M., Ghosh A.. Rhenium-Imido Corroles. Inorg. Chem. 2020;59:6382–6389. doi: 10.1021/acs.inorgchem.0c00477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Alemayehu A. B., Settineri N. S., Lanza A. E., Ghosh A.. Rhenium-Sulfido and -Dithiolato Corroles: Reflections on Chalcophilicity. Inorg. Chem. 2024;63:24787–24796. doi: 10.1021/acs.inorgchem.4c04091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Buchler J. W., De Cian A., Fischer J., Kruppa S. B., Weiss R.. Metal complexes with tetrapyrrole ligands LVII. Synthesis, spectra, and structure of nitridorhenium­(V) porphyrins. Chem. Ber. 1990;123(12):2247–2253. doi: 10.1002/cber.19901231204. [DOI] [Google Scholar]
  35. Borisov S. M., Einrem R. F., Alemayehu A. B., Ghosh A.. Ambient-temperature near-IR phosphorescence and potential applications of rhenium-oxo corroles. Photochem. Photobiol. Sci. 2019;18:1166–1170. doi: 10.1039/c8pp00473k. [DOI] [PubMed] [Google Scholar]
  36. Einrem R. F., Alemayehu A. B., Borisov S. M., Ghosh A., Gederaas O. A.. Amphiphilic Rhenium-Oxo Corroles as a New Class of Sensitizers for Photodynamic Therapy. ACS Omega. 2020;5:10596–10601. doi: 10.1021/acsomega.0c01090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Lohrey T. D., Cortes E. A., Fostvedt J. I., Oanta A. K., Jain A., Bergman R. G., Arnold J.. Diverse Reactivity of a Rhenium (V) Oxo Imido Complex: [2+ 2] Cycloadditions, Chalcogen Metathesis, Oxygen Atom Transfer, and Protic and Hydridic 1, 2-Additions. Inorg. Chem. 2020;59(59):11096–11107. doi: 10.1021/acs.inorgchem.0c01589. [DOI] [PubMed] [Google Scholar]
  38. Pyykkö P., Riedel S., Patzschke M.. Triple-Bond Covalent Radii. Chem.Eur. J. 2005;11:3511–3520. doi: 10.1002/chem.200401299. [DOI] [PubMed] [Google Scholar]
  39. Pyykkö P.. Additive Covalent Radii for Single-, Double-, and Triple-Bonded Molecules and Tetrahedrally Bonded Crystals: A Summary. J. Phys. Chem. A. 2015;119:2326–2337. doi: 10.1021/jp5065819. [DOI] [PubMed] [Google Scholar]
  40. Christy A. G.. Quantifying Lithophilicity, Chalcophilicity and Siderophilicity. Eur. J. Mineral. 2018;30:193–204. doi: 10.1127/ejm/2017/0029-2674. [DOI] [Google Scholar]
  41. Romão C. C., Kühn F. E., Herrmann W. A.. Rhenium (VII) oxo and imido complexes: synthesis, structures, and applications. Chem. Rev. 1997;97:3197–3246. doi: 10.1021/cr9703212. [DOI] [PubMed] [Google Scholar]
  42. Dilworth J. R.. Rhenium chemistry – then and now. Coord. Chem. Rev. 2021;436:213822. doi: 10.1016/j.ccr.2021.213822. [DOI] [Google Scholar]
  43. Braband H., Benz M., Spingler B., Conradie J., Alberto R., Ghosh A.. Relativity as a Synthesis Design Principle: A Comparative Study of [3+ 2] Cycloaddition of Technetium (VII) and Rhenium (VII) Trioxo Complexes with Olefins. Inorg. Chem. 2021;60:11090–11097. doi: 10.1021/acs.inorgchem.1c00995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Kepp K. P.. A Quantitative Scale of Oxophilicity and Thiophilicity. Inorg. Chem. 2016;55:9461–9470. doi: 10.1021/acs.inorgchem.6b01702. [DOI] [PubMed] [Google Scholar]
  45. Gouterman M.. Spectra of Porphyrins. J. Mol. Spectrosc. 1961;6:138–163. doi: 10.1016/0022-2852(61)90236-3. [DOI] [Google Scholar]
  46. Gouterman M., Wagniére G. H., Snyder L. C.. Spectra of Porphyrins. Part II. Four-Orbital Model. J. Mol. Spectrosc. 1963;11:108–115. doi: 10.1016/0022-2852(63)90011-0. [DOI] [Google Scholar]
  47. Gouterman M.. Optical Spectra and Electronic Structure of Porphyrins and Related Rings. Porphyrins. 1978;3:1–165. doi: 10.1016/B978-0-12-220103-5.50008-8. [DOI] [Google Scholar]
  48. Ghosh A.. An Exemplary Gay Scientist and Mentor: Martin Gouterman (1931–2020) Angew. Chem., Int. Ed. 2021;60:9760–9770. doi: 10.1002/anie.202012840. [DOI] [PubMed] [Google Scholar]
  49. Ghosh A., Wondimagegn T., Parusel A. B.. 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. doi: 10.1021/ja9943243. [DOI] [Google Scholar]
  50. Ghosh A.. Electronic Structure of Corrole Derivatives: Insights from Molecular Structures, Spectroscopy, Electrochemistry, and Quantum Chemical Calculations. Chem. Rev. 2017;117:3798–3881. doi: 10.1021/acs.chemrev.6b00590. [DOI] [PubMed] [Google Scholar]
  51. Koren K., Borisov S. M., Saf R., Klimant I.. Strongly phosphorescent iridium (III)–porphyrins – New Oxygen Indicators with Tuneable Photophysical Properties and Functionalities. Eur. J. Inorg. Chem. 2011;2011:1531–1534. doi: 10.1002/ejic.201100089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Qin L., Guan X., Yang C., Huang J. S., Che C.-M.. Near-Infrared Phosphorescent Supramolecular Alkyl/Aryl-Iridium Porphyrin Assemblies by Axial Coordination. Chem.-Eur. J. 2018;24:14400–14408. doi: 10.1002/chem.201803238. [DOI] [PubMed] [Google Scholar]
  53. Hua L., Zhang K. Y., Liu H. W., Chan K. S., Lo K. K. W.. Luminescent iridium­(III) porphyrin complexes as near-infrared-emissive biological probes. Dalton Trans. 2023;52:12444–12453. doi: 10.1039/D3DT02104A. [DOI] [PubMed] [Google Scholar]
  54. Palmer J. H., Durrell A. C., Gross Z., Winkler J. R., Gray H. B.. Near-IR phosphorescence of iridium (III) corroles at ambient temperature. J. Am. Chem. Soc. 2010;132:9230–9231. doi: 10.1021/ja101647t. [DOI] [PubMed] [Google Scholar]
  55. Sinha W., Ravotto L., Ceroni P., Kar S.. NIR-Emissive Iridium­(III) Corrole Complexes as Efficient Singlet Oxygen Sensitizers. Dalton Trans. 2015;44:17767–17773. doi: 10.1039/C5DT03041B. [DOI] [PubMed] [Google Scholar]
  56. Thomassen I. K., McCormick-McPherson L. J., Borisov S. M., Ghosh A.. Iridium Corroles Exhibit Weak Near-Infrared Phosphorescence but Efficiently Sensitize Singlet Oxygen Formation. Sci. Rep. 2020;10(1):7551. doi: 10.1038/s41598-020-64389-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Eastwood D., Gouterman M.. Porphyrins XVIII. Luminescence of (Co), (Ni), Pd, Pt Complexes. J. Mol. Spectrosc. 1970;35:359–375. doi: 10.1016/0022-2852(70)90179-7. [DOI] [Google Scholar]
  58. Harriman A.. Luminescence of porphyrins and metalloporphyrins. Part 3.Heavy-atom effects. J. Chem. Soc., Faraday Trans. 1981;77:1281–1291. doi: 10.1039/F29817701281. [DOI] [Google Scholar]
  59. Drouet S., Paul-Roth C. O., Fattori V., Cocchi M., Williams J. G.. Platinum and palladium complexes of fluorenyl porphyrins as red phosphors for light-emitting devices. New J. Chem. 2011;35:438–444. doi: 10.1039/C0NJ00561D. [DOI] [Google Scholar]
  60. Sommer J. R., Shelton A. H., Parthasarathy A., Ghiviriga I., Reynolds J. R., Schanze K. S.. Photophysical Properties of Near-Infrared Phosphorescent π-Extended Platinum Porphyrins. Chem. Mater. 2011;23:5296–5304. doi: 10.1021/cm202241e. [DOI] [Google Scholar]
  61. Subedi D. R., Reid R., D’Souza P. F., Nesterov V. N., D’Souza F.. Singlet Oxygen Generation in Peripherally Modified Platinum and Palladium Porphyrins: Effect of Triplet Excited State Lifetimes and meso-Substituents on 1O2 Quantum Yields. ChemPluschem. 2022;87(4):e202200010. doi: 10.1002/cplu.202200010. [DOI] [PubMed] [Google Scholar]
  62. Lai S. W., Hou Y. J., Che C. M., Pang H. L., Wong K. Y., Chang C. K., Zhu N.. Electronic Spectroscopy, Photophysical Properties, and Emission Quenching Studies of An Oxidatively Robust Perfluorinated Platinum Porphyrin. Inorg. Chem. 2004;43:3724–3732. doi: 10.1021/ic049902h. [DOI] [PubMed] [Google Scholar]
  63. Borisov S. M., Alemayehu A., Ghosh A.. Osmium-Nitrido Corroles as NIR Indicators for Oxygen Sensors and Triplet Sensitizers for Organic Upconversion and Singlet Oxygen Generation. J. Mater. Chem. C. 2016;4:5822–5828. doi: 10.1039/C6TC01126H. [DOI] [Google Scholar]
  64. Alemayehu A. B., McCormick L. J., Gagnon K. J., Borisov S. M., Ghosh A.. Stable platinum (IV) corroles: synthesis, molecular structure, and room-temperature near-IR phosphorescence. ACS Omega. 2018;3(8):9360–9368. doi: 10.1021/acsomega.8b01149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Alemayehu A. B., Day N. U., Mani T., Rudine A. B., Thomas K. E., Gederaas O. A., Vinogradov S. A., Wamser C. C., Ghosh A.. Gold Tris­(carboxyphenyl)­corroles as Multifunctional Materials: Room Temperature Near-Ir Phosphorescence and Applications to Photodynamic Therapy and Dye-Sensitized Solar Cells. ACS Appl. Mater. Interfaces. 2016;8:18935–18942. doi: 10.1021/acsami.6b04269. [DOI] [PubMed] [Google Scholar]
  66. Soll M., Sudhakar K., Fridman N., Müller A., Röder B., Gross Z.. One-Pot Conversion of Fluorophores to Phosphorophores. Org. Lett. 2016;18:5840–5843. doi: 10.1021/acs.orglett.6b02877. [DOI] [PubMed] [Google Scholar]
  67. Lemon C. M., Powers D. C., Brothers P. J., Nocera D. G.. Gold corroles as near-IR phosphors for oxygen sensing. Inorg. Chem. 2017;56:10991–10997. doi: 10.1021/acs.inorgchem.7b01302. [DOI] [PubMed] [Google Scholar]
  68. Sahu K., Angeloni S., Conradie J., Villa M., Nayak M., Ghosh A., Ceroni P., Kar S.. NIR-Emissive, Singlet-Oxygen-Sensitizing Gold Tetra­(thiocyano)­corroles. Dalton Trans. 2022;51:13236–13245. doi: 10.1039/D2DT01959K. [DOI] [PubMed] [Google Scholar]
  69. Reiher M., Salomon O., Hess B. A.. Reparameterization of Hybrid Functionals Based on Energy Differences of States of Different Multiplicity. Theor. Chem. Acc. 2001;107(1):48–55. doi: 10.1007/s00214-001-0300-3. [DOI] [Google Scholar]
  70. Rabinovich D., Parkin G.. Synthesis and Structure of W­(PMe3)4(Te)2: The First Transition-Metal Complex with a Terminal Tellurido Ligand. J. Am. Chem. Soc. 1991;113:9421–9422. doi: 10.1021/ja00024a088. [DOI] [Google Scholar]
  71. Murphy V. J., Parkin G.. Syntheses of Mo­(PMe3)6 and trans-Mo­(PMe3)4(E)2 (E = S, Se, Te): the First Series of Terminal Sulfido, Selenido, and Tellurido Complexes of Molybdenum. J. Am. Chem. Soc. 1995;117:3522–3528. doi: 10.1021/ja00117a021. [DOI] [Google Scholar]
  72. Kisko J. L., Hascall T., Parkin G.. Multiple Bonding of Titanium and Vanadium to the Heavier Chalcogens: Syntheses and Structures of the Terminal Selenido and Tellurido Complexes [η4-Me8taa]­ME (M = Ti, V; E= Se, Te) J. Am. Chem. Soc. 1997;119:7609–7610. doi: 10.1021/ja971043r. [DOI] [Google Scholar]
  73. Trnka T. M., Parkin G.. A Survey of Terminal Chalcogenido Complexes of the Transition Metals: Trends in Their Distribution and the Variation of their M = E Bond Lengths. Polyhedron. 1997;16:1031–1045. doi: 10.1016/S0277-5387(96)00411-1. [DOI] [Google Scholar]
  74. Chivers T., Laitinen R. S.. Tellurium: A Maverick Among the Chalcogens. Chem. Soc. Rev. 2015;44:1725–1739. doi: 10.1039/C4CS00434E. [DOI] [PubMed] [Google Scholar]
  75. Sandblom N., Ziegler T., Chivers T.. A Density Functional Study of The Bonding in Tertiary Phosphine Chalcogenides and Related Molecules. Can. J. Chem. 1996;74:2363–2371. doi: 10.1139/v96-263. [DOI] [Google Scholar]
  76. Alvarado S. R., Shortt I. A., Fan H. J., Vela J.. Assessing Phosphine–Chalcogen bond Energetics from Calculations. Organometallics. 2015;34:4023–4031. doi: 10.1021/acs.organomet.5b00428. [DOI] [Google Scholar]
  77. Zingaro R. A., Steeves B. H., Irgolic K.. Phosphine tellurides. J. Organomet. Chem. 1965;4:320–323. doi: 10.1016/S0022-328X(00)88840-3. [DOI] [Google Scholar]
  78. Austad T., Rød T., Åse K., Songstad J., Norbury A. H., Swahn C.-G.. The Adduct between Triphenyl Phosphine and Triphenyl Phosphine Telluride. A Tellurium(0) Compound. Acta Chem. Scand. 1973;27:1939–1949. doi: 10.3891/acta.chem.scand.27-1939. [DOI] [Google Scholar]
  79. Nordheider A., Woollins J. D., Chivers T.. Organophosphorus–Tellurium Chemistry: From Fundamentals to Applications. Chem. Rev. 2015;115:10378–10406. doi: 10.1021/acs.chemrev.5b00279. [DOI] [PubMed] [Google Scholar]
  80. Sun H., Wang F., Buhro W. E.. Tellurium Precursor for Nanocrystal Synthesis: Tris­(dimethylamino)­phosphine Telluride. ACS Nano. 2018;12:12393–12400. doi: 10.1021/acsnano.8b06468. [DOI] [PubMed] [Google Scholar]
  81. Wasbotten I. H., Wondimagegn T., Ghosh A.. Electronic Absorption, Resonance Raman, and Electrochemical Studies of Planar and Saddled Copper­(III) Meso-Triarylcorroles. Highly Substituent-Sensitive Soret Bands as a Distinctive Feature of High-Valent Transition Metal Corroles. J. Am. Chem. Soc. 2002;124:8104–8116. doi: 10.1021/ja0113697. [DOI] [PubMed] [Google Scholar]
  82. Koszarna B., Gryko D. T.. Efficient Synthesis of Meso-Substituted Corroles in a H2O–MeOH mixture. J. Org. Chem. 2006;71:3707–3717. doi: 10.1021/jo060007k. [DOI] [PubMed] [Google Scholar]
  83. Buchler J. W., Kruppa S. B.. Metal Complexes with Tetrapyrrole Ligands, LV [1] Improved Syntheses of Oxorhenium­(V) Porphyrins and Novel Trichlororhenium­(V) Porphyrins. J. Nat. Res. B. 1990;45:518–530. doi: 10.1515/znb-1990-0418. [DOI] [Google Scholar]
  84. Sheldrick G. M. S.. Integrated space-group and crystal-structure determination. Acta Crystallogr. 2015;A71:3–8. doi: 10.1107/S2053273314026370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Sheldrick G. M.. Crystal structure refinement with ShelXL. Acta Crystallogr. 2015;C71:3–8. doi: 10.1107/S2053229614024218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. a Maluendes S. A., McLean A. D., Yamashita K., Herbst E.. Relativistic regular two-component Hamiltonians. J. Chem. Phys. 1993;99(4):2812–2820. doi: 10.1063/1.465190. [DOI] [PubMed] [Google Scholar]; b van Lenthe E., Baerends E. J., Snijders J. G.. Relativistic total energy using regular approximations. J. Chem. Phys. 1994;101(11):9783–9792. doi: 10.1063/1.467943. [DOI] [Google Scholar]; c van Lenthe E., Ehlers A., Baerends E-J.. Geometry optimizations in the zero order regular approximation for relativistic effects. J. Chem. Phys. 1999;110:8943–8953. doi: 10.1063/1.478813. [DOI] [Google Scholar]
  87. Handy N. C., Cohen A.. Left-Right Correlation Energy. J. Mol. Phys. 2001;99:403–412. doi: 10.1080/00268970010018431. [DOI] [Google Scholar]
  88. Lee 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. doi: 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
  89. Grimme S., Anthony 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(15):154104. doi: 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
  90. 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(9):931–967. doi: 10.1002/jcc.1056. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ic5c01593_si_001.pdf (7.6MB, pdf)

Data Availability Statement

All data generated or analyzed in this study are included in this published article and its Supporting Information.


Articles from Inorganic Chemistry are provided here courtesy of American Chemical Society

RESOURCES