A Porphyrin Analogue Containing Phosphorus at the Meso Position

electrons, its oxide exhibited weak antiaromaticity, which was confirmed experimentally and ... rin-like skeleton that offer attractive optical and el...
0 downloads 0 Views 4MB Size
Subscriber access provided by ECU Libraries

Communication

Ni(II) 10-Phosphacorrole: A Porphyrin Analogue Containing Phosphorus at the Meso Position Hiroto Omori, Satoru Hiroto, Youhei Takeda, Heike Fliegl, Satoshi Minakata, and Hiroshi Shinokubo J. Am. Chem. Soc., Just Accepted Manuscript • Publication Date (Web): 07 Mar 2019 Downloaded from http://pubs.acs.org on March 7, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Ni(II) 10-Phosphacorrole: A Porphyrin Analogue Containing Phosphorus at the Meso Position Hiroto Omori,† Satoru Hiroto,† Youhei Takeda,‡ Heike Fliegl,§* Satoshi Minakata,‡ and Hiroshi Shinokubo†* †Department of Molecular and Macromolecular Chemistry, Graduate School of Engineering, Nagoya University, Furo-cho, Chikusaku, Nagoya, Aichi 464-8603, Japan ‡Department of Applied Chemistry, Graduate School of Engineering, Osaka University, Yamadaoka 2-1, Suita, Osaka 565-0871, Japan § Karlsruhe Institute of Technology, Institute of Nanotechnology, Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany Supporting Information Placeholder ABSTRACT: Ni(II) 10-Phosphacorrole, a porphyrinoid con-

taining phosphorus at the meso position, was synthesized from a bis(α,α’-dibromodipyrrin) Ni(II) complex and a phosphine anion via the palladium-catalyzed formation of a C–C and two C–P bonds. The optoelectronic properties of Ni(II) 10-phosphacorrole can be modulated effectively by oxidation or coordination of a metal to the phosphorus center. While Ni(II) 10-phosphacorrole exhibits a distinctly aromatic character due to the cyclic conjugation of 18 πelectrons, its oxide exhibited weak antiaromaticity, which was confirmed experimentally and theoretically.

Phosphorus-containing π-conjugated molecules exhibit unique structures that are different from the corresponding carbon analogues due to the trigonal pyramidal geometry of the phosphorus atom.1 Another important feature of phosphorus is its electron-donating nature in π-systems owing to the presence of a lone pair of electrons. Furthermore, various modifications on phosphorus such as oxidation, alkylation, and metal-coordination effectively alter its structural and electronic characteristics. These features allow fine-tuning the optoelectronic properties of phosphorus-containing πsystems.2 Porphyrinoids are macrocyclic compounds with a porphyrin-like skeleton that offer attractive optical and electrochemical characteristics.3 These properties can be tuned substantially via their skeleton and structural factors. Considering the fascinating properties of phosphorus-containing πsystems, the introduction of phosphorus in the porphyrin skeleton should uncover novel aspects of porphyrin chemistry. Matano and co-workers have created phospholecontaining porphyrins4 as core-modified porphyrins.5 In such core-modified porphyrins, which result from the replacement of the NH moiety in pyrrole with other heteroatoms, the heteroatom behaves as a substituent to the macrocyclic π-conjugation circuit. In contrast, the replacement of meso

carbon atoms with heteroatoms should result in a more drastic electronic perturbation as the heteroatoms are directly involved in the π-conjugation circuit.6 Alas, such mesophosphaporphyrinoids have not yet been reported. Recently, Bröring and co-workers have reported the effective synthesis of 10-heterocorroles such as 10-oxa-, 10-thia-, and 10-selenacorroles.7 Independently, our group has synthesized 10-aza-, 10-oxa-, 10-thia-, and 10-silacorroles using a metal-template strategy.8 On the basis of the aforementioned attractive features of phosphorus, 10-phosphacorroles should also represent promising targets. Here, we report the synthesis of Ni(II) 10-phosphacorroles, i.e., phosphaporphyrinoids that contain phosphorus at the meso position of the porphyrin skeleton, and its derivatives. The synthesis of Ni(II) 10-phosphacorrole 2 was achieved by nucleophilic substitution of bis(α,α’-dibromodipyrrin) Ni(II) complex 1 with a phosphine anion. The reaction of 1 with bis(trimethylsilyl)mesitylphosphine9 induced the simultaneous formation of one C–C and two C–P bonds, providing Ni(II) 10-phosphacorrole 2 in 13% yield. We eventually found that the use of Pd-catalyzed C–P coupling enhanced the yield of 2 (Scheme 1).10 Treatment of 1 with mesitylphosphine in the presence of PdCl2(dppf)×CH2Cl2 and potassium tert-butoxide afforded 2 in 52% yield. Scheme 1. Synthesis of Ni(II) 10-phosphacorrole 2. Mes MesPH2 (10 equiv) PdCl2(dppf)•CH2Cl2 (30 mol%) N N t-BuOK (6.0 equiv) Br Br Ni Br Br toluene N N 90 °C, 15 h

Mes

1

Mes N

N

P Mes

Ni N Mes

N

2

52%

Compound 2 was characterized by NMR spectroscopy and HR-MS spectrometry. The presence of the phosphorus atom was confirmed by a peak at –42.0 ppm in its 31P NMR

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

spectrum. The 1H NMR spectrum of 2 exhibited four doublets for the pyrrole protons in the aromatic region from 7.35 to 7.69 ppm. These chemical shifts indicate a distinctly aromatic nature of Ni(II) 10-phosphacorrole 2.11 While the para-methyl protons of the meso-mesityl groups appeared at 2.47 ppm, the ortho-methyl protons were upfield-shifted to 2.00 ppm because of the diatropic ring current effect.12 The parent mass ion peak of 2 was observed at m/z = 727.2473 (calcd for (C45H42N4PNi)+ = 727.2495), and the isotope pattern matched theoretical expectations. The solid-state structure of 2 was unambiguously determined by X-ray diffraction analysis (Figure 1a, 1b). Crystals of 2 contain two crystallographically inequivalent molecules per asymmetric unit (Figure S11). The macrocyclic skeleton of 2 is planar except for the phosphorus atom, which is deviated from the molecular mean plane. The endocyclic C–P bonds of 2 (1.7653(39), 1.7693(37), 1.7695(37), and 1.7715(39) Å) are slightly shorter than the exocyclic C–P bond (1.7836(102) and 1.823(34)Å) in 2 (Figure S13). This suggests a double bond character of the endocyclic C–P bonds in 2, which is indicative of π-conjugation between the pyrrole α-carbon and phosphorus to induce aromaticity in 2. The planarity of the phosphorus center was evaluated on the basis of the sum of the three C–P–C angles. Phosphacorrole 2 shows larger values (323.1° and 316.5°) than that in mesitylphosphole (312.8°).13 The bottom depth of 2 (0.640 and 0.704 Å) is smaller than that in PPh3 (0.81 Å), suggesting a relatively planar structure for phosphorus in 2. In addition, the C–C bond lengths around the meso-carbon atoms are almost identical, supporting the aromatic nature of 2 (Figure S13). (a)

(b)

Page 2 of 7

ellipsoids are shown at 50% probability; one of the individual molecules of 2 and 2-O are shown.

Subsequently, we attempted a chemical modification of phosphorus in 2 (Scheme 2). Oxidation of 2 with mchloroperoxybenzoic acid (mCPBA) furnished the corresponding oxide 2-O in 95% yield. The 1H NMR spectrum of 2-O showed four doublets for the pyrrole protons at 5.71, 5.61, 5.60, and 5.56 ppm. Compared to nonaromatic porphyrinoids, these signals are substantially upfield shifted due to weak antiaromaticity (vide infra).14 The structure of 2-O was unambiguously determined by X-ray diffraction analysis (Figure 1d, 1e, and Figure S12). The endocyclic C–P bonds (1.784(3), 1.790(3), and 1.788(3) Å) are longer than those in 2, but still shorter than the exocyclic C–P bond (1.813(3) and 1.821(3) Å) (Figure S14). The reaction of 2 with [Au(SMe2)Cl] provided gold(I) complex 2-AuCl in 68% yield. Its 1H NMR spectrum exhibited pyrrole proton peaks at 6.93–6.70 ppm, indicating its nonaromatic or weak aromatic character (vide infra). An X-ray diffraction analysis revealed the formation of a 1:1 complex (Figure 1e, 1f, and Figure S15). In comparison to 2, 2-O and 2-AuCl exhibited distinct bond length alternation around the meso-carbon atoms, indicating loss of the aromaticity in 2 (Figure S14, S15). Moreover, these derivatives can be easily converted back into 2. Oxide 2-O was reduced to 2 in 68% yield using trichlorosilane, while a treatment of 2-AuCl with triphenylphosphine regenerated 2 quantitatively. However, the demetallation of Ni(II) from these derivatives remained unsuccessful under a variety of conditions. Scheme 2. Chemical modifications of the phosphorus atom in 2. Mes

2

(c)

mCPBA (2.0 eq.)

N

CH2Cl2 r.t., 5 min

N

(d)

O

N P

Ni N

Mes

Mes

95%

2-O

Mes

2 (e)

Au(SMe2)Cl (2.0 eq.)

N

CH2Cl2 r.t., 5 min

N

(f)

Mes

Figure 1. X-ray crystal structures of 2 (a) top and (b) side view; 2-O (c) top and (d) side view; and 2-AuCl (e) top and (f) side view; meso-Mesityl groups are omitted for clarity and thermal

AuCl

N P

Ni N

Mes

2-AuCl

68%

Then, we examined the dynamic behavior of the pyramidal phosphorus center in 2. The 1H NMR spectrum of 2 exhibited the meta-proton signals on the P-mesityl group as a broad singlet at 25 °C and this peak was split into two peaks at low temperatures (Figure S10). Similar temperature-dependent behaviors were also observed for the ortho-methyl protons on the P-mesityl and meso-mesityl groups. This result indi-

ACS Paragon Plus Environment

Page 3 of 7

cates that the pyramidal conformation of phosphorus inverts quickly at 25 °C.15 The inversion barrier was measured to be 9.8 kcal mol–1 at –55 °C on the basis of the coalescence temperature of the meta protons.16 The inversion dynamics were also calculated by DFT methods at the B3LYP/6311+G(2d,p)+SDD//B3LYP/6-31G(d)+SDD level of theory, affording an activation barrier of 10.2 kcal mol–1 (Figure S20), which is significantly lower than those of phosphines (30–35 kcal mol–1) and phosphole (~17 kcal mol–1).17 This facile inversion behavior can be rationalized in terms of an enhanced aromaticity in the transition state as well as relatively planar conformation of the phosphorus center in the ground state. The DFT-calculated transition-state structure suggests a planar geometry around phosphorus, in which the aromatic character is improved.18 In fact, the NICS values of the planar transition state were more negative than those of the pyramidal ground state (Figure S22). Figure 2 displays the UV/Vis/NIR absorption spectra of 2, 2-O, and 2-AuCl in CH2Cl2. The lowest-energy absorption band of 2 was bathochromically shifted compared to that of normal porphyrins and 10-azacorroles. Furthermore, this absorption band is even further bathochromically shifted relative to that of a core-modified phosphaporphyrin (701 nm),4b demonstrating the profound effect of the modification at the meso position on the electronic character of the porphyrin. This band further shifted to lower energy upon oxidation or coordination to gold. In particular, the band of 2-O reached into the NIR region (1054 nm). Thus, it can be concluded that the electronic structure of Ni(II) 10phosphacorrole is dramatically modulated by modification of the phosphorus atom.

in CH2Cl2 using [Bu4N][PF6] as the supporting electrolyte (Table 1 and Figure S16). The first oxidation wave of 2 was observed at 0.239 V, which is higher than that of Ni(II) azacorrole (0.181 V), but much lower than that of Ni(II) tetraphenylporphyrin (0.55 V) and core-modified phosphaporphyrin (0.38 V)4b owing to the electron donation from phosphorus. On the other hand, the first reduction wave of 2 was observed at a much higher potential (–1.61 V) than that of Ni(II) azacorrole (–2.01 V). Oxidation and coordination to gold increased both the oxidation and reduction potentials, albeit that the reduction potential was affected more profoundly. The difference between the first oxidation and the first reduction potentials (ΔE = Eox1–Ered1) decreases in the order 2>2-AuCl>2-O, which is in good agreement with the absorption spectra. Table 1. Summarized redox potentials of 2, 2-O, and 2-AuCl.a

Compound 2 2-O 2-AuCl

Eox (V) 0.239 0.474 0.495

Ered (V) –1.61 –1.05 –1.17

ΔE (V)b 1.85 1.52 1.67

a: values versus Fc/Fc+, b: DE = DEox – DEred a)

b)

c)

d)

7.0

6.0

5.0

ε [104 cm–1M–1]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

e)

f)

4.0

3.0

2.0

Figure 3. a) HOMO and b) LUMO of 2, c) HOMO and d) LUMO of 2-O, and e) HOMO and f) LUMO of 2-AuCl.

1.0

0.0 300

600

λ [nm]

900

1200

Figure 2. UV/vis/NIR absorption spectra of 2 (black), 2-O (red), and 2-AuCl (blue) in CH2Cl2.

The electrochemical properties of 2, 2-O, and 2-AuCl were investigated by cyclic voltammetry. The oxidation and reduction potentials of 2, 2-O, and 2-AuCl were determined

In order to gain further insights on the electronic structures, we performed DFT calculations at the B3LYP/631G(d)+SDD level of theory. In the LUMO of 2-O, the π* orbital is extended over the phosphorus atom and the adjacent carbon atoms (Figure 3 and Figure S17). This result suggests an interaction between the σ* orbitals of the C–P and O–P bonds and the π* orbital of the macrocycle, which would significantly lower the energy level of the LUMO of 2O. This notion is consistent with the results of the electro-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

chemical analysis. This interpretation is also supported by natural bond orbital (NBO) analysis (Table S2, Figure S18).19 The TD-DFT calculation suggests that the lowest energy absorption band of 2-O is originated from its HOMO–LUMO transition (Figure S19a). The features of HOMO and LUMO of 2-O are similar to those of isocorrole in terms of the orbital symmetry and nodal positions except the presence of the σ*–π* conjugation in LUMO of 2-O (Figure S19b). This fact would explain the similar absorption features of 2-O to isocorrole.20 To obtain more insights on the ring current effect in 2, 2O, and 2-AuCl, we performed a ring-current analysis using the gauge-including magnetically induced current (GIMIC) method (Figure 4).21 The GIMIC method is a reliable approach to quantitatively determine the degree of aromaticity. In line with the 1H NMR analysis, 2 shows a distinct diatropic current stream around the periphery of the molecule (Figure 4a). Although the cyclic π-conjugation circuit through the phosphorus atom is not obvious in 2-O given the lack of p-orbitals on phosphorus, the σ* orbitals of the C–P and O– P bonds could be involved to construct a current path via the C–P–C linkage. In fact, the current strength analysis indicates a weak but non-negligible paratropic ring current (–5.0 nA T–1) through the phosphorus atom, supporting the weak anticipated antiaromaticity of 2-O (Figure 4b). On the other hand, a diatropic ring current (6.5 nA T–1) flows through the C–P–C linkage of 2-AuCl, suggesting its weak aromaticity compared to 13.0 nA T–1 as seen for 2. This could be due to the involvement of the P–Au σ orbital to the conjugation pathway, supported by NBO analysis (Table S2, Figure S18). The diatropic ring current 2 and 2-AuCl as well as the paratropic ring current in 2-O are also evident from the anisotropy of the induced current density (ACID) plots (Figure S2527) and NICS calculations (Figure S23, S24).22

Page 4 of 7

In conclusion, we have achieved the synthesis of a 10phosphacorrole Ni(II) complex from a bis(α,α’dibromodipyrrin) Ni(II) complex and a phosphine anion via the palladium-catalyzed formation of a C–C and two C–P bonds. The introduction of phosphorus directly into the porphyrin π-system effectively induces bathochromic shifts in the absorption spectrum without significant changing the molecular size. Chemical modifications on the phosphorus atom stabilize the LUMO energy level and thus change the absorption spectrum. The successful introduction of phosphorus at the meso position and the ability to fine-tune should deliver a means to develop the chemistry of mesomodified heteroporphyrinoids. The coordination ability of 2 at phosphorus should allow constructing porphyrin-based transition-metal catalysts and supramolecular metal complexes. ASSOCIATED CONTENT Supporting Information. Experimental details and spectral data for all new compounds; crystallographic data (CIF files) for 2, 2O, and 2-AuCl; computational details; this material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author

[email protected]; [email protected] Funding Sources

No competing financial interests have been declared.

ACKNOWLEDGMENT This work was supported by JSPS KAKENHI grants JP26102003 (H.S.), JP15K21721 (H.S.), JP17H01190 (H.S.), and JP17H05155 (Y.T.). H.O. expresses his gratitude for a JSPS Research Fellowship for Young Scientists (JP17J10877). H.S. acknowledges the Murata Science Foundation for financial support. H.F. thanks the Norwegian Research Council for support via the CoE Hylleraas Centre for Quantum Molecular Sciences (grants 262695 and 231571/F20). This work has received support from the Norwegian Supercomputing Program (NOTUR) via a grant of computer time (NN4654K).

REFERENCES (1)

Figure 4. Calculated integrated ring current strength susceptibilities in nA T-1 of a) 2, b) 2-O, and c) 2-AuCl for selected bonds obtained from the GIMIC method at the BHLYP/def2TZVP level. The red and black arrows indicate diatropic (clockwise; positive values) and paratropic (counter-clockwise; negative values) currents, respectively.

(a) Mathey, F. The Organic Chemistry of Phospholes. Chem. Rev. 1988, 88, 429. (b) Hissler, M.; Dyer, P. W.; Réau, R. Linear Organic π-Conjugated Systems Featuring the Heavy Group 14 and 15 Elements. Coord. Chem. Rev. 2003, 244, 1. (c) Hobbs, M. G.; Baumgartner, T. Recent Developments in Phosphole-Containing Oligoand Polythiophene Materials. Eur. J. Inorg. Chem. 2007, 23, 3611. (d) Matano, Y.; Imahori, H. Design and Synthesis of PhospholeBased π Systems for Novel Organic Materials. Org. Biomol. Chem. 2009, 7, 1258. (e) Fukazawa, A.; Yamaguchi, S. Ladder πConjugated Materials Containing Main-Group Elements. Chem. – Asian J. 2009, 4, 1386. (f) Matano, Y.; Saito, A.; Fukushima, T.; Tokudome, Y.; Suzuki, F.; Sakamaki, D.; Kaji, H.; Ito, A.; Tanaka, K.; Imahori, H. Fusion of Phosphole and 1,1’-Biacenaphthene: Phosphorus(V)-Containing Extended π-Systems with High Electron Affinity and Electron Mobility. Angew. Chem., Int. Ed. 2011, 50,

ACS Paragon Plus Environment

Page 5 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

(2)

(3)

(4)

(5)

8016. (g) Märkl, G.; Schubert, H. Phosphepine – Eine Neue Verbindungsklasse. Tetrahedron Lett. 1970, 11, 1273. (h) Märkl, G.; Burger, W. 3-Phenyl-3-Benzphospherin. Tetrahedron Lett. 1983, 24, 2545. (i) Märkl, G.; Burger, W. 2,7-Di-tert-Butyl-1-phenyl-1Hphosphepine–The First Stable, Monocyclic Phosphepine. Angew. Chem., Int. Ed. 1984, 23, 894. (j) Kurita, J.; Shiratori, S.; Yasuike, S.; Tuchiya, T. Synthesis of the First Examples of 1Benzophosphepines. J. Chem. Soc., Chem. Commun. 1991, 1227. (k) Segall, Y.; Shirin, E.; Granoth, I. Phosphorus Sulfur Silicon Relat. Elem. 1980, 8, 243. (a) Baumgartner, T.; Bergmans, W.; Kárpáti, T.; Neumann, T.; Nieger, M.; Nyulászi, L. From Model Compounds to Extended πConjugated Systems: Synthesis and Properties of Dithieno[3,2b:2’,3’-d]phospholes. Chem.-Eur. J. 2005, 11, 4687. (b) Fukazawa, A.; Hara, M.; Okamoto, T.; Son, E.-C.; Xu, C.; Tamao, K.; Yamaguchi, S. Bis-Phosphoryl-Bridged Stilbenes Synthesized by an Intramolecular Cascade Cyclization. Org. Lett. 2008, 10, 913. (c) Bouit, P. A.; Escande, A.; Szűcs, R.; Szieberth, D.; Lescop, C.; Nyulászi, L.; Hissler, M.; Réau, R. Dibenzophosphapentaphenes: Exploiting P Chemistry for Gap Fine-Tuning and Coordination-Driven Assembly of Planar Polycyclic Aromatic Hydrocarbons. J. Am. Chem. Soc. 2012, 134, 6524. (d) Yamaguchi, E.; Wang, C.; Fukazawa, A.; Taki, M.; Sato, Y.; Sasaki, T.; Ueda, M.; Sasaki, N.; Higashiyama, T.; Yamaguchi, S. Environment-Sensitive Fluorescent Probe: A Benzophosphole Oxide with an Electron-Donating Substituent. Angew. Chem., Int. Ed. 2015, 54, 4539. (a) Kadish, K.; Smith, K. M.; Guilard, R. The Porphyrin Handbook, Vol. 11-20, Academic Press, Boston, 2003. (b) Kadish, K.; Smith, K. M.; Guilard, R. Handbook of Porphyrin Science, Vol. 1-30, World Scientific Publishing Co. Pte. Ltd. Singapore, 2010. (a) Matano, Y.; Nakabuchi, T.; Miyajima, T.; Imahori, H.; Nakano, H. Synthesis of a Phosphorus-Containing Hybrid Porphyrin. Org. Lett. 2006, 8, 5713. (b) Matano, Y.; Nakashima, M.; Nakabuchi, T.; Imahori, H.; Fujishige, S.; Nakano, H. Monophosphaporphyrins:  Oxidative π-Extension at the Peripherally Fused Carbocycle of the Phosphaporphyrin Ring. Org. Lett. 2008, 10, 553. (c) Matano, Y.; Nakabuchi, T.; Fujishige, S.; Nakano, H.; Imahori, H. Redox-Coupled Complexation of 23-Phospha-21-thiaporphyrin with Group 10 Metals: A Convenient Access to Stable Core-Modified Isophlorin−Metal Complexes. J. Am. Chem. Soc. 2008, 130, 16446. (d) Nakabuchi, T.; Nakashima, M.; Fujishige, S.; Nakano, H.; Matano, Y.; Imahori, H. Synthesis and Reactions of Phosphaporphyrins: Reconstruction of πSkeleton Triggered by Oxygenation of a Core Phosphorus Atom. J. Org. Chem. 2010, 75, 375. (a) Chmielewski, P. J.; Latos-Grażyński, L. Core Modified Porphyrins – A Macrocyclic Platform for Organometallic Chemistry. Coord. Chem. Rev. 2005, 249, 2510. (b) Gupta, I.; Ravikanth, Recent Developments in Heteroporphyrins and Their Analogues. M. Coord. Chem. Rev. 2006, 250, 468. (c) Ulman, A.; Manassen, J. Synthesis of New Tetraphenylporphyrin Molecules Containing Heteroatoms Other than Nitrogen. I. Tetraphenyl-21,23-Dithiaporphyrin. J. Am. Chem. Soc. 1975, 97, 6540. (d) Ulman, A.; Manassen, J. Synthesis of Tetraphenylporphyrin Molecules Containing Heteroatoms Other than Nitrogen. Part 4. Symmetrically and Unsymmetrically Substituted Tetraphenyl-21,23-Dithiaporphyrins. J. Chem. Soc., Perkin Trans. 1 1979, 1066. (e) Latos-Grażyński, L.; Lisowski, J.; Olmstead, M. M.; Balch, A. L. 21-Thiatetra-p-tolylporphyrin and Its Copper(II) Bicarbonate Complex. Structural Effects of Copper– Thiophene Binding. J. Am. Chem. Soc., 1987, 109, 4428. (f) Cho, W.-S.; Kim, H.-J.; Litter, B. J.; Miller, M. A.; Lee, C.-H.; Lindsey, J. S. Rational Synthesis of Trans-Substituted Porphyrin Building Blocks Containing One Sulfur or Oxygen Atom in Place of Nitrogen at a Designated Site. J. Org. Chem. 1999, 64, 7890. (g) Latos-Grażyński, L.; Lisowski, J.; Chmielewski, P.; Grzeszczuk, M.; Olmstead, M. M.; Balch, A. L. Palladium Complexes of 21-Thiaporphyrin: Syntheses and Characterization. Inorg. Chem. 1994, 33, 192. (h) Chuang, C.H.; Ou, C.-K.; Liu, S.-T.; Kumar, A.; Chiang, W.-M.; Chiang, P.-C.

(6)

(7)

(8)

(9)

(10)

De La Rosa, M. A. C.; Hung, C.-H. Ruthenium Complexes of Thiaporphyrin and Dithiaporphyrin. Inorg. Chem. 2011, 50, 11947. (a) Orłowski, R.; Gryko, T.; Gryko, D. T. Synthesis of Corroles and Their Heteroanalogs. Chem. Rev. 2017, 117, 3102. (b) Matano, Y. Synthesis of Aza-, Oxa-, and Thiaporphyrins and Related Compounds. Chem. Rev. 2017, 117, 3138. (c) Chatterjee, T.; Shetti, V. S.; Sharma, R.; Ravikanth, M. Heteroatom-Containing Porphyrin Analogues. Chem. Rev. 2017, 117, 3254. (a) Bröring, M.; Brégier, F.; Tejero, E. C.; Hell, C.; Holthausen, M. C. Revisiting the Electronic Ground State of Copper Corroles. Angew. Chem., Int. Ed. 2007, 46, 445. (b) Sakow, D.; Böker, B.; Brandhorst, K.; Burghaus, O.; Bröring, M. 10-Heterocorroles: RingContracted Porphyrinoids with Fine-Tuned Aromatic and MetalBinding Properties. Angew. Chem., Int. Ed. 2013, 52, 4912. (c) Sakow, D.; Baabe, D.; Böker, B.; Burghaus, O.; Funk, M.; Kleeberg, C.; Menzel, D.; Pietzonka, C.; Bröring, M. Iron 10-Thiacorroles: Bioinspired Iron(III) Complexes with an Intermediate Spin (S=3/2) Ground State. Chem.-Eur. J. 2014, 20, 2913. (a) Kamiya, H.; Kondo, T.; Sakida, T.; Yamaguchi, S.; Shinokubo, H. meso-Thiaporphyrinoids Revisited: Missing of Sulfur by Small Metals. Chem.-Eur. J. 2012, 18, 16129. (b) Ito, T.; Hayashi, Y.; Shimizu, S.; Shin, J.-Y.; Kobayashi, N.; Shinokubo, H. Gram-Scale Synthesis of Nickel(II) Norcorrole: The Smallest Antiaromatic Porphyrinoid. Angew. Chem., Int. Ed. 2012, 51, 8542. (c) Horie, M.; Hayashi, Y.; Yamaguchi, S.; Shinokubo, H. Synthesis of Nickel(II) Azacorroles by Pd-Catalyzed Amination of α,α’-Dichlorodipyrrin Ni(II) Complex and Their Properties. Chem.-Eur. J. 2012, 18, 5919. (d) Omori, H.; Hiroto, S.; Shinokubo, H. The Synthesis of Ni(II) and Al(III) 10-Azacorroles through Coordination-Induced Cyclisation Involving 1,2-Migration. Chem. Commun. 2016, 52, 3540. (e) Omori, H.; Hiroto, S.; Shinokubo, H. Synthesis of Free-Base 10Azacorroles. Org. Lett. 2016, 18, 2978. (f) Omori, H.; Hiroto, S.; Shinokubo, H. 10-Silacorroles Exhibiting Near-Infrared Absorption and Emission. Chem.-Eur. J. 2017, 23, 7866. (a) Becker, V. G.; Mundt, O.; Rössler, M.; Schneider, E. Bildung und Eigenschaften von Acylphosphanen. VI. Synthese von Alkyl- und Arylbis(trimethylsilyl)- sowie Alkyl und Aryltrimethylsilyphosphanen. Z. Anorg. Allg. Chem. 1978, 443, 42. (b) Dugal-Tessier, J.; Kuhn, P.-S.; Dake, G. R.; Gates, D. P. Synthesis of Functional Phosphines with ortho-Substituted Aryl Groups: 2-RC6H4PH2 and 2-RC6H4 P(SiMe3)2 (R = i-Pr- or t-Bu). Heteroatom Chem. 2010, 21, 265. (c) Takeda, Y.; Nishida, T.; Minakata, S. 2,6-Diphospha-s-indacene-1,3,5,7(2H,6H)tetraone: A Phosphorus Analogue of Aromatic Diimides with the Minimal Core Exhibiting High Electron-Accepting Ability. Chem.-Eur. J. 2014, 20, 10266. (d) Takeda, Y.; Nishida, T.; Hatanaka, K.; Minakata, S. Revisiting Phosphorus Analogues of Phthalimides and Naphthalimides: Syntheses and Comparative Studies. Chem.-Eur. J. 2015, 21, 1666. (e) Takeda, Y.; Hatanaka, K.; Nishida, T.; Minakata, S. Thieno[3,4-c]phosphole-4,6-dione: A Versatile Building Block for Phosphorus-Containing Functional π-Conjugated Systems. Chem.Eur. J. 2016, 22, 10360. (a) Tappe, F. M. J.; Trepohl, V. T.; Oestreich, M. Transition-MetalCatalyzed C–P Cross-Coupling Reactions. Synthesis 2010, 18, 3037. (b) Wauters, I.; Debrouwer, W.; Stevens, C. V. Preparation of Phosphines through C–P Bond Formation. Beilstein J. Org. Chem. 2014, 10, 1064. (c) Herd, O.; Heßler, A.; Hingst, M.; Tepper, M.; Stelzer, O. Water Soluble Phosphines VII. Palladium-Catalyzed P– C Cross Coupling Reactions between Primary or Secondary Phosphines and Functional Aryliodides – A Novel Synthetic Route to Water Soluble Phosphines. J. Organomet. Chem. 1996, 522, 69. (d) Herd, O.; Heßler, A.; Hingst, M.; Machnitzki, P.; Tepper, M.; Stelzer, O. Palladium Catalyzed P–C Coupling – A Powerful Tool for the Syntheses of Hydrophilic Phosphines. Catal. Today 1998, 42, 413. (e) Machnitzki, P.; Nickel, T.; Stelzer, O.; Landgrafe, C. Novel Syntheses of Mono- and Bisphosphonated Aromatic Phosphanes by Consecutive Pd-Catalyzed P–C Coupling Reactions and Nucleophilic Phosphanylation – X-ray Structure of Ph2P-C6H4-mPO3Na2·5.5H2O·iPrOH. Eur. J. Inorg. Chem. 1998, 1029.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(11)

(12) (13)

(14)

(15) (16) (17) (18)

(19)

(20)

Recently, the aromaticity of 10-phosphacorrole has been predicted by DFT calculations; for details, see: Foroutan-Nejad, C.; Ghosh, A. Magnetic Diversity in Heteroisocorroles: Aromatic Pathways in 10Heteroatom-Substituted Isocorroles. ACS Omega 2018, 3, 15865. These chemical shift values of ortho- and para-methyl protons are in line with other aromatic 10-heterocorroles.8 Miesel, D.; Hildebrandt, A.; Korb, M.; Wild, D. A.; Low, P. J.; Lang, H. Influence of P-Bonded Bulky Substituents on Electronic Interactions in Ferrocenyl-Substituted Phospholes. Chem.-Eur. J. 2015, 21, 11545. Phosphole oxides also exhibit weak antiaromaticity; see: (a) Nyulászi, L.; Hollóczki, O.; Lescop, C.; Hissler, M.; Réau, R. An Aromatic–Antiaromatic Switch in P-Heteroles. A Small Change in Delocalisation Makes a Big Reactivity Difference. Org. Biomol. Chem. 2006, 4, 996. (b) Mucsi, Z.; Keglevich, G. Why are Phosphole Oxides Unstable? The Phenomenon of Antiaromaticity as a Destabilizing Factor. Eur. J. Org. Chem. 2007, 4765. We excluded the possibility of the rotation of the mesityl group due to its calculated high energy barrier (19.7 kcal mol–1)(Figure S21). Sandström, J. Dynamic NMR Spectroscopy, Academic Press, London, 1982, p.96. Rauk, A.; Allen, L. C.; Mislow, K. Pyramidal Inversion. Angew. Chem. Int. Ed. Engl. 1970, 9, 400. Fujimoto, K.; Osuka, A. Effective Stabilization of a Planar Phosphorus(III) Center Embedded in a Porphyrin-Based Fused Aromatic Skeleton. Chem. Sci. 2017, 8, 8231. Reed, A. E.; Curtiss, L. A.; Weinhold, F. Intermolecular interactions from a natural bond orbital, donor-acceptor viewpoint. Chem. Rev. 1988, 88, 899. (a) Pomarico, G.; Xiao, X.; Nardis, S.; Paolesse, R.; Fronczek, F. R.; Smith, K. M.; Fang, Y.; Ou, Z.; Kadish, K. M. Synthesis and Charac-

(21)

(22)

terization of Free-Base, Copper, and Nickel Isocorroles. Inorg. Chem. 2010, 49, 5766. (b) Hoffmann, M.; Cordes, B.; Kleeberg, C.; Schweyen, P.; Wolfram, B.; Bröring, M. Template Synthesis of Alkyl-Substituted Metal Isocorroles. Eur. J. Inorg. Chem. 2016, 3076. (a) Valiev, R. R.; Benkyi, I.; Konyshev, Y. V.; Fliegl, H.; Sundholm, D. Computational Studies of Aromatic and Photophysical Properties of Expanded Porphyrins. J. Phys. Chem. A 2018, 122, 4756. (b) Fliegl, H.; Valiev, R. R.; Pichierri, F.; Sundholm, D. Theoretical Studies as a Tool for Understanding the Aromatic Character of Porphyrinoid Compounds. Chem. Modell. 2018, 14, 1. (c) Sundholm, D.; Fliegl, H.; Berger, R. J. F. Calculations of Magnetically Induced Current Densities: Theory and Applications. WIREs Comput. Mol. Sci. 2016, 6, 639. (d) Benkyi, I.; Fliegl, H.; Valiev, R. R.; Sundholm, D. New Insights into Aromatic Pathways of Carbachlorins and Carbaporphyrins Based on Calculations of Magnetically Induced Current Densities. Phys. Chem. Chem. Phys. 2016, 18, 11932. (e) Fliegl, H.; Jusélius, J.; Sundholm, D. Gauge-Origin Independent Calculations of the Anisotropy of the Magnetically Induced Current Densities. J. Phys. Chem. A 2016, 120, 5658. (f) Fliegl, H.; Taubert, S.; Lehtonen, O.; Sundholm, D. The Gauge Including Magnetically Induced Current Method. Phys. Chem. Chem. Phys. 2011, 13, 20500. (g) Jusélius, J.; Sundholm, D.; Gauss, J. Calculation of Current Densities Using Gauge-Including Atomic Orbitals. J. Chem. Phys. 2004, 121, 3952. (a) Geuenich, D.; Hess, K.; Köhler, F.; Herges, R. Anisotropy of the Induced Current Density (ACID), a General Method To Quantify and Visualize Electronic Delocalization. Chem. Rev. 2005, 105, 3758. (b) Herges, R.; Geuenich, D. Delocalization of Electrons in Molecules. J. Phys. Chem. A 2001, 105, 3214.

ACS Paragon Plus Environment

Page 6 of 7

Page 7 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Ar

Br Br

C-P & C–C coupling

N

N

Br

Ni

Br N

N Ar

Ar

Ar N N Ni N N

X P R

Ar

Ar

18π aromatic

N N Ni N N

X P

R

X = O or Au

tuning of the electronic property

ACS Paragon Plus Environment