Synthesis and Functionalization of Porphyrins through Organometallic

Oct 6, 2016 - Department of Applied Chemistry, Graduate School of Engineering, Nagoya University, Nagoya 464-8603, Japan. Chem. ... This review focuse...
2 downloads 9 Views 27MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Review pubs.acs.org/CR

Synthesis and Functionalization of Porphyrins through Organometallic Methodologies Satoru Hiroto, Yoshihiro Miyake, and Hiroshi Shinokubo* Department of Applied Chemistry, Graduate School of Engineering, Nagoya University, Nagoya 464-8603, Japan ABSTRACT: This review focuses on the postfunctionalization of porphyrins and related compounds through catalytic and stoichiometric organometallic methodologies. The employment of organometallic reactions has become common in porphyrin synthesis. Palladium-catalyzed cross-coupling reactions are now standard techniques for constructing carbon−carbon bonds in porphyrin synthesis. In addition, iridium- or palladium-catalyzed direct C−H functionalization of porphyrins is emerging as an efficient way to install various substituents onto porphyrins. Furthermore, the coppermediated Huisgen cycloaddition reaction has become a frequent strategy to incorporate porphyrin units into functional molecules. The use of these organometallic techniques, along with the traditional porphyrin synthesis, now allows chemists to construct a wide range of highly elaborated and complex porphyrin architectures.

CONTENTS 1. Introduction 1.1. Background: Porphyrin and Organometallic Chemistry 1.2. Classification of Porphyrin Functionalization 1.3. Overview of Organometallic Methodologies Used for Porphyrin Synthesis 2. Stoichiometric and Catalytic C−H Functionalization of Porphyrins 2.1. Organolithiums as Nucleophiles 2.2. Addition of Grignard Reagents and Heteroatom Nucleophiles 2.3. Stoichiometric Metalation of C−H Bonds 2.4. Catalytic C−H Functionalization of Porphyrins 2.4.1. Iridium-Catalyzed C−H Borylation 2.4.2. Catalytic C−H Arylation and Imidation 3. Reactions of Halogenated and Metalated Porphyrins 3.1. Core-Functionalization 3.1.1. Metalation of Haloporphyrins 3.1.2. Suzuki−Miyaura Coupling 3.1.3. Sonogashira Coupling 3.1.4. Migita−Kosugi−Stille, Negishi, and Kumada−Tamao−Corriu Couplings 3.1.5. Mizoroki−Heck Reaction 3.1.6. Other Coupling Methodologies for C−C Bond Formation 3.1.7. Homocoupling 3.1.8. Introduction of Heteroatoms 3.1.9. Copper- and Nickel-Catalyzed Introduction of Heteroatoms 3.1.10. Rhodium-Catalyzed Reactions 3.2. Peripheral-Functionalization © XXXX American Chemical Society

3.2.1. 3.2.2. 3.2.3. 3.2.4. 3.2.5.

Suzuki−Miyaura Coupling Sonogashira Coupling Migita−Kosugi−Stille Coupling The Mizoroki−Heck Reaction Amination, Amidation, and Etherification 4. Reactions at Alkenes and Alkynes Bound to Porphyrins 4.1. Homocoupling of Terminal Alkynes 4.2. Alkene, Alkyne, and Enyne Metatheses 4.3. Cyclotrimerization 4.4. Copper-Catalyzed Huisgen Cycloaddition 4.4.1. Click Reactions for the Synthesis of Porphyrin Oligomers and Multichromic Systems 4.4.2. Construction of Porphyrin−Nanocarbon Conjugates through Click Chemistry 4.4.3. Self-Assembly of Triazolylporphyrins 4.4.4. Click Reaction for Porphyrin Functionalization 4.4.5. Porphyrins Decorated with Sugars, Peptides, and DNA 4.4.6. Click Reactions for Surface Chemistry 4.5. Other Reactions Involving Alkenes and Alkynes 5. Construction of Porphyrin Macrocycles through Organometallic Methodologies 5.1. Porphyrin Analogues Containing Carbazole and Indole Units

B B B B D D F G I I M T T T Y AF AI AM AM AN AP AV AY AY

AY BB BH BI BI BK BK BM BP BP

BQ BW CA CC CD CE CG CH CH

Special Issue: Expanded, Contracted, and Isomeric Porphyrins Received: July 2, 2016

A

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 1. Post- and prefunctionalization of porphyrins.

5.2. Porphyrin Analogues from Dipyrrin Precursors 6. Summary and Outlook Author Information Corresponding Author Notes Biographies Acknowledgments Abbreviations References

materials, information storage, molecular wires, metal ligands, supramolecules, and so forth.1,2 To achieve these fascinating functions, design and synthesis of structurally diverse porphyrin molecules is essential. For example, hydrophilic substituents are often installed to increase solubility in aqueous media and enhance membrane permeability for cancer therapy applications. Porphyrins for solar cell applications require donor and acceptor moieties at certain positions. For energy and electron transfer studies, porphyrin oligomers with adequate potential gradients have to be prepared. These complex and elaborate structures are difficult to prepare using only conventional porphyrin synthesis. Undoubtedly, stoichiometric organometallic reagents and catalytic transition metal complexes are powerful tools in organic synthesis.3−5 One finds many pages describing novel organometallic methodologies in every issue of the major general chemistry journals. Transition metal catalysis not only enables high selectivity, high efficiency, and environmentally benign processes but also new types of direct transformations to short-cut lengthy multistep syntheses, which are not attainable by conventional methodologies. One of the most convincing examples is the olefin metathesis reaction catalyzed by ruthenium and molybdenum carbene complexes.6 Olefin metathesis has completely changed the organic synthesis of natural products and polymers.

CM CP CR CR CR CR CR CS CU

Figure 2. Classification of porphyrin functionalization by position.

1.2. Classification of Porphyrin Functionalization

1. INTRODUCTION

This review predominantly covers research into postfunctionalization methodologies of porphyrins and related compounds through catalytic and stoichiometric organometallic reactions (Figure 1). Namely, functional groups and substituents are introduced after the construction of the porphyrin macrocycles in these types of reactions. Porphyrin syntheses using functionalized precursors such as aldehydes, pyrroles, and dipyrromethanes are outside the scope of this review article. Postfunctionalization of porphyrins can be further classified into two categories (Figure 2): one is core-functionalization and the other is peripheral-functionalization. In the case of core-functionalization, the core-skeleton of the porphyrin macrocycle is directly functionalized at the meso- and/or βpositions. On the other hand, functional groups can be introduced on the peripheral substituents. The most of these functionalizations are performed using various types of porphyrins bearing reactive substituents such as halogens, alkynes, and metals on the porphyrin core or the substituents.

1.1. Background: Porphyrin and Organometallic Chemistry

Porphyrin is an 18π aromatic macrocyclic compound that consists of four pyrrole units and four bridging carbon atoms in a planar conformation. One can find the structure of porphyrin in nature, such as in various types of chlorophylls and hemes. Chlorophylls play pivotal roles in photosynthesis as both light harvesting antennae and charge separation reaction systems. Hemes are one of the key components for biocatalysts and oxygen carriers in the blood. Without porphyrins, no life can exist on earth. Particularly in chemistry, porphyrins and metalated porphyrins have attracted organic and inorganic chemists simply because of their beautiful molecular shapes. The highly symmetrical D4h planar structure of metalloporphyrin has inspired a number of synthetic chemists and brought out their molecular creativity. Moreover, porphyrin is one of the archetypal functional molecules, playing an important role in diverse areas of scientific research owing to its unique electronic and optical properties. Porphyrin research has a long history, covering a wide variety of disciplines of natural sciences, including photosynthesis, P450-related biocatalysis, organic photovoltaic cells, photodynamic therapeutic agents, bioimaging probes, chemosensors, conductive organic materials, lightemitting materials, near-infrared dyes, nonlinear optical

1.3. Overview of Organometallic Methodologies Used for Porphyrin Synthesis

The employment of organometallic methodologies in organic synthesis, such as the total synthesis of highly complex natural products, has become commonplace. This is also true in the field of porphyrin chemistry, and the use of organometallic transformations in porphyrin synthesis has been reviewed a B

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 1. Addition of Alkyllithiums to Tetraphenylporphyrin

Scheme 2. Trapping of an Anionic Intermediate with Alkyl Halides

Scheme 3. Addition of tert-Butyllithium under Palladium Catalysis

Scheme 4. Formal Direct Alkylation of Porphyrins through an Addition−Oxidation Sequence

number of times.7−12 The use of palladium-catalyzed crosscoupling reactions in porphyrin synthesis expands every year. Palladium-catalyzed coupling techniques include Suzuki− Miyaura coupling,13−16 Migita−Kosugi−Stille coupling,17−20 Negishi coupling,21−24 Sonogashira coupling,25−30 the Mizoroki−Heck reaction,31−34 and Buchwald−Hartwig amination.35−40 These methodologies have been successfully applied to derivatize halogenated porphyrins at various positions of the macrocyclic core and peripheral substituents. Copper-mediated transformations have a long history and have been important tools in organic synthesis. These methodologies include Glaser-type dimerization of terminal alkynes41,42 and the Ullmann coupling reaction,43,44 which are

also useful in the synthesis of butadiyne-linked porphyrins and heteroatom-substituted porphyrins. However, the most useful and most frequently employed copper-catalyzed reaction in porphyrin synthesis is clearly the Huisgen cycloaddition reaction between organic azides and terminal alkynes, the socalled “click reaction” or copper-catalyzed azide−alkyne cycloaddition.45−50 This methodology is a powerful and general way to connect two units into one molecule in a highly efficient manner. Organic synthesis has been largely and rapidly expanded in its efficacy by the use of direct C−H functionalization strategies.51−56 This is also true for porphyrin synthesis. Such direct C−H functionalization methodologies represent a major C

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 5. Sequential Functionalization of 5,15-Diphenylporphyrin

Scheme 6. Introduction of a Formyl Group Using Dithianyllithium

part of this review. Iridium-catalyzed direct C−H borylation was developed by Smith, Hartwig, Ishiyama, and Miyaura,57−69 and this methodology has been employed for postfunctionalization to modify relatively simple and unfunctionalized porphyrins. Palladium-catalyzed C−H arylation70−72 of porphyrins is also emerging as a direct method to install various aromatic substituents onto porphyrins. Several “organometallic” porphyrins with carbon−metal (Li, Mg, Zn, Pd, Pt, and Hg) and carbon−metalloid (Si and B) σbonds at the peripheral position of the porphyrin skeleton have been reported. Organometallic methodologies using such types of stoichiometric metallic reagents have also been effective.73

Scheme 7. Introduction of a Formyl Group Using (Pyridyldimethylsilyl)methyllithium

2. STOICHIOMETRIC AND CATALYTIC C−H FUNCTIONALIZATION OF PORPHYRINS

Scheme 8. Alkylation of Core-Modified Porphyrins

2.1. Organolithiums as Nucleophiles

Organolithium reagents work as effective nucleophiles to attack at meso- and β-positions of porphyrins, enabling direct functionalization of porphyrins.74−77 The addition of nucleophiles to meso- and β-positions of tetraphenylporphyrin 1.1 yielded phlorin 1.2 and chlorin 1.3, respectively (Scheme 1).78,79 The addition of butyllithium to 2.1 generates the anionic intermediate 2.2 (Scheme 2).80 In fact, quenching of 2.2 with hexyl iodide afforded porphodimethene 2.3 as expected. The reaction of tetraphenylporphyrin 1.1 with tert-butyllithium in the presence of a palladium catalyst and CuI resulted in a D

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 9. Alkylation of Diazaporphyrin

Scheme 10. Preparation of meso−meso-Linked Diporphyrin

Scheme 11. Reaction of β-Nitroporphyrins with Grignard Reagents

Scheme 12. Aromatic Nucleophilic Substitution of mesoBromoporphyrin with Sodium Azide

double addition of the nucleophile to furnish 5,10-porphodimethene 3.1 and chlorin 3.2 (Scheme 3).81,82 Senge and co-workers extensively investigated the sequential addition of nucleophiles to the porphyrin core, followed by oxidation. The addition of butyllithium to octaethylporphyrin 2.1 yielded meso-alkylated product 4.1 after hydrolysis. Oxidation of 4.1 with DDQ afforded the corresponding porphyrin 4.2 in high yield (Scheme 4).83,84 The methodology was applicable to the introduction of alkyl and aryl groups into all four meso-positions by the use of corresponding alkyllithium and aryllithium reagents.85−92 Stepwise addition and oxidation reactions of 5.1 with two different lithium reagents furnished A2BC-type porphyrin 5.3 having three different groups at the meso-positions (Scheme 5).93 The sequential introduction of substituents at the meso-positions enabled the synthesis of a E

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 13. Mercuration of a Porphyrin

Scheme 14. Mercuration of 5,15-Diarylporphyrin

variety of porphyrins such as push−pull type porphyrins94 and a number of unique porphyrins that were applicable to photodynamic therapy95,96 and in multichromophores.97 The sequential methodology using organolithium reagents can be further expanded to the direct introduction of useful functionality. A dithianyl group serves as an acyl anion equivalent. The reaction of 5,15-diphenylporphyrin 5.1 with dithianyllithium 6.1 provided 6.2, which was efficiently converted to formylporphyrin 6.3 after deprotection (Scheme 6).98,99 Takanami and co-workers used 2(pyridyldimethylsilyl)methyllithium 7.1 as an acyl anion equivalent and succeeded in the introduction of the formyl group at the meso-position of 6.3 (Scheme 7).100−102 The addition reaction of organolithium reagents is also applicable for the functionalization of core-modified porphyrins. A similar reaction of thiaporphyrin 8.1 afforded alkylated thiaporphyrin 8.2 (Scheme 8).103 The reaction of 5,15diazaporphyrin 9.1 with organolithium reagents exhibited intriguing selectivity (Scheme 9).104 The regioselective addition of lithium reagents at the 3-position afforded diazachlorin 9.2 and the subsequent DDQ oxidation provided 3-alkylated diazaporphyrins 9.3. The regioselectivity can be accounted for by the directing effect of the meso-nitrogen atom and the

intermediate 9.4 plays an important role in the addition reaction. Senge and co-worker reported the synthesis of meso−mesolinked diporphyrins on the basis of the addition reaction of organolithium reagents (Scheme 10).85,105 The reaction of 5.1 with butyllithium followed by DDQ oxidation resulted in the formation of meso−meso-linked diporphyrin 10.1. Oxidation of the anionic intermediate likely generates a π-delocalized radical species, which then undergoes dimerization to provide 10.1. 2.2. Addition of Grignard Reagents and Heteroatom Nucleophiles

Crossley and co-workers examined the reaction of β-nitroporphyrins with Grignard reagents (Scheme 11).106 The addition of Grignard reagents to nitroporphyrin 11.1 and subsequent elimination of the nitro group proceeded smoothly to afford the corresponding β-alkylated porphyrins 11.3. The reactivity of β-nitroporphyrins can also be applied to the introduction of various heteroatom nucleophiles. β-Alkoxy-, hydroxy-, and aminoporphyrins were obtained.107−111 Yamashita, Sugiura, and co-workers succeeded in the development of SNAr reactions at the meso-position (Scheme 12).112 Treatment of meso-bromoporphyrin 12.1Ni with sodium azide afforded the corresponding meso-azidoporphyrin F

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 15. Directed C−H Metalation of Pyridylporphyrins

afforded trimercurated porphyrin 13.2, functionalized at two βpositions and one meso-position (Scheme 13). Sugiura, Arnold, and co-workers succeeded in the regioselective mercuration of porphyrins (Scheme 14).126 The reaction of porphyrins 14.1 with mercury(II) trifluoroacetate afforded mercurated porphyrins 14.2 at the more sterically hindered β-position. The obtained mercurated porphyrin 14.2Ni was converted into the corresponding iodide 14.3, which was employed as a coupling partner in Suzuki−Miyaura coupling reactions. Directed metalation of C−H bonds is a useful method for the introduction of a transition metal into an aromatic skeleton, leading to cyclometalated compounds. This chemistry has been

12.2 in good yield. This process is applicable for the reaction of dibromoporphyrins. Similar SNAr reactions at the meso-position with heteroatom- and carbon-centered nucleophiles under catalyst-free conditions were also reported.113 2.3. Stoichiometric Metalation of C−H Bonds

Peripherally metalated porphyrin complexes have been actively investigated.114−117 Among them, porphyrin organometallic complexes with carbon−metal σ-bonds are useful for functionalization of the porphyrin macrocycle. In the beginning of work on directly metalated porphyrins, the mercuration of porphyrins was extensively examined.118−125 The reaction of porphyrin 13.1 with mercury(II) diacetate G

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 16. Formation of Pt(IV) and Pt(II) Complexes of Pyridylporphyrin

Scheme 17. Directed C−H Metalation of Phosphanylporphyrins

successfully applied to functionalization of porphyrins. Directing groups play a crucial role in the C−H metalation step, in which the position and the coordination of heteroatoms affect the structures and the properties of peripherally cyclometalated porphyrins. Pyridine is a useful directing group in this regard. The reaction of porphyrin 15.1 having two pyridyl groups at β-positions with palladium and platinum salts furnished pincer-type porphyrin complexes 15.2 and 15.3 through meso-C−H metalation (Scheme 15).127−130 Platinum complex 15.3 was transformed into Pt(IV) complex 15.4 by oxidation with CuCl2. The obtained 15.2, 15.3, and 15.4 promoted catalytic and stoichiometric carbon−carbon bond forming reactions.131 Porphyrin dimer 15.5 and trimer 15.6 were also synthesized by a similar methodology and these ladder-shaped oligomers with largely bent structures exhibited large two-photon absorption cross-section values.132

Pyridylporphyrin 16.1 also showed unique coordination behavior (Scheme 16).133 The reaction of 16.1 with (Bu4N)2PtCl6 smoothly afforded Pt(IV)-bridged cofacial porphyrin dimer 16.2. The reduction of 16.2 with MeNHNH2 yielded Pt(II) dimer 16.3. The Pt(IV) center in 16.2 adopted a distorted octahedral structure, while the coordination of Pt(II) in 16.3 was square planar. The interactions between the two porphyrin cores in 16.2 and 16.3 are dependent on the coordination geometries of the Pt centers. Platinum(II) metalation of 16.1 in DMF induced double cleavage of sp2 C−H and sp3 C−H bonds on one Pt center.134 Matano and co-workers designed meso-phosphanylporphyrin 17.1 which underwent selective β-C−H metalation (Scheme 17).135,136 Treatment of 17.1 with Pd(II) and Pt(II) salts afforded the metal-bridged porphyrin dimer 17.2 through cyclometalation at the β-position. The UV−vis absorption spectrum of 17.2 clearly showed the presence of electronic H

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 18. Iridium-Catalyzed Direct Borylation of Porphyrins

communication between the coplanar porphyrin π-systems through the peripheral β-C−M bonds. Yorimitsu, Osuka, and co-workers reported meso-C−H metalation of diphosphanylporphyrins 17.3M with nickel, palladium, and platinum salts, furnishing the corresponding PCP pincer complexes 17.4MM′.137 Interestingly, the catalytic activity of palladium pincer complexes 17.4MPd in the 1,4-reduction of chalcone was affected by the central metals.

active intermediate, the regioselectivity of the borylation is affected by the steric demand rather than the electronic properties of the substrates. The β-regioselectivity of the reaction enables novel functionalization of porphyrins. In addition, the resulting borylated products are employed as versatile building blocks for porphyrin derivatives owing to the rich chemistry of organoboron compounds. Osuka and Shinokubo reported the regioselective functionalization of porphyrins at the β-positions through iridiumcatalyzed direct borylation.140 Treatment of 5,15-diarylporphyrin 14.1H with bis(pinacolato)diboron ((Bpin)2) under the catalysis of [Ir(OMe) (cod)]2 and 4,4′-di(tert-butyl)-2,2′bipyridyl (dtbpy) afforded mono- and diborylated products 18.1H and 18.2 (Scheme 18). With an excess of (Bpin)2, tetraborylated product 18.3H was obtained in 73% yield from 14.1H. The use of 1,4-dioxane as a solvent was effective due to the relatively high solubility of porphyrin substrates as compared with hydrocarbon solvents, which are typically employed in the borylation of small molecules. 1H NMR spectral analysis revealed that borylation proceeded not at the meso-positions but at the β-positions. The regioselectivity was confirmed by X-ray diffraction analysis of 18.3H. Because the βpositions of porphyrins are sterically less demanding than the meso-positions, the borylation reaction exhibited perfect regioselectivity. This reaction also worked well for Ni(II) and Cu(II) complexes 14.1Ni and 14.1Cu, affording the corresponding monosubstituted products 18.1Ni and 18.1Cu in 47 and 44% yield, respectively. On the other hand, the reaction with Zn(II) complex 14.1Zn was sluggish because of its low solubility in 1,4-dioxane. The borylated product of Zn(II) complex 14.1Zn was obtained by the use of mesitylene as the solvent. The compatibility of functional groups was also examined (Scheme 19).141 Borylated product 19.2 was obtained in 37% yield in the reaction with 5,15-dihexylporphyrin 19.1. Meso−

2.4. Catalytic C−H Functionalization of Porphyrins

In porphyrin chemistry, cross-coupling reactions with halogenated porphyrins have often been used for introduction of various substituents and functionalities as well as the construction of nanoscale and mesoscopic structures. However, brominated porphyrins are not usually very soluble in common organic solvents, resulting in difficult handling of the substrates. To avoid the tedious synthesis and purification of halogenated porphyrins, direct C−H functionalization under transitionmetal catalysis is growing in popularity. In addition, the mild reaction conditions of these methodologies enable control of the regioselectivity of the functionalization. To date, both iridium- and palladium-catalyzed β-selective direct functionalization methodologies of porphyrin derivatives have been reported. 2.4.1. Iridium-Catalyzed C−H Borylation. As introduced in section 1, iridium-catalyzed direct borylation was developed by Smith, Hartwig, Ishiyama, and Miyaura. This reaction requires a catalytic amount of an iridium complex and a ligand as well as a stoichiometric amount of a boron source. The facile preparation of the iridium catalyst and the easy handling of these reagents make these reactions well-suited for their use in porphyrin chemistry.138 The reaction mechanism was also investigated and the insertion of the iridium(III) intermediate into a C−H bond of an aromatic skeleton is considered to be the rate-determining step.139 Because of the bulkiness of the I

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 19. Iridium-Catalyzed Direct Borylation of Porphyrins with Various meso-Substituents

reaction of meso-borylporphyrin 19.7Zn. Unfortunately, bromo substituents did not survive under the standard reaction conditions. Extensive debromination occurred in a refluxing 1,4-dioxane solution. In this case, the use of octane as a solvent was effective to afford borylated product 19.10 in 32% yield.

meso-linked diporphyrin 19.3 was also borylated under standard conditions, affording tetraborylated product 19.4 in 60% yield. Borylation of meso-alkynylporphyrin 19.5 provided borylated product 19.6 in 70% yield without suffering from the presence of the alkynyl moiety. Partial deborylation occurred during the J

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 20. Iridium-Catalyzed Direct Borylation at meso-Aryl Substituents

yield. On the other hand, borylation of β-alkylporphyrin 18.5 predominantly produced para-diborylated product 18.6a in 77% yield. The intrinsic reactivity at the β-position of a porphyrin was compared with that at a meso-aryl substituent (Scheme 21). Treatment of meso-(3-methoxyphenyl)porphyrin 21.1 provided β-borylated porphyrin 21.2a as a major product. This result indicated the higher reactivity of the β-positions of porphyrins for iridium-catalyzed direct borylation. After installation, the boron substituent can be transformed into various functionalities. Oxidation of 2,18-borylporphyrin 22.1H with Oxone produced 2,18-dihydroxyporphyrin 22.2H in good yield (Scheme 22).142 Porphyrin 22.2H was trans-

Although there are several limitations in this reaction, the borylation strategy is useful to introduce various functionalities only at the β-positions without modification at the mesopositions. In the case of fully meso-substituted porphyrins, borylation occurred at meso-aryl substituents (Scheme 20). For example, borylation of 2,6-dimethoxyphenylporphyrin 20.1 proceeded exclusively at the para-positions of meso-aryl groups, furnishing diborylated product 20.2 in 84% yield. The borylation at the aryl substituents is significantly influenced by the adjacent βsubstituents of the porphyrin core. For example, borylation of 5,10,15-tris(3,5-dioctyloxyphenyl)-20-phenylporphyrin 20.3 mainly provided meta,meta-diborylated product 18.4 in 82% K

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 21. Comparison of the Reactivity of C−H Bonds at the β-Position Compared to That of the Aromatic Substituent

Scheme 22. Transformation of β-Boryl Substituents into Various Functionalities

Scheme 23. Iridium-Catalyzed Direct Borylation of Tris(pentafluorophenyl)corrole

into iodo, bromo, and chloro substituents (22.4Ni−22.6Ni) by treatment with the corresponding N-halosuccinimide and copper(I) salt.143 These halogenated porphyrins are useful for

formed into stable bistriflate 22.3H by treatment with PhNTf2 under basic conditions. Yorimitsu, Osuka, and co-workers reported the excellent conversion of the boryl group in 22.1Ni L

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 24. Iridium-Catalyzed Direct Borylation of a meso-Free Subporphyrin

Scheme 25. Iridium-Catalyzed Direct Borylation of Hexaphyrins

because of the reductive nature of the reaction conditions. On the other hand, borylation of 25.2 proceeded smoothly to afford hexaborylated product 25.3 in excellent yield. 2.4.2. Catalytic C−H Arylation and Imidation. Introduction of aryl groups at the porphyrin periphery has been a useful method to control physical properties of porphyrins. The nucleophilic addition of aryllithium or alkyllithium reagents is one of the strategies to introduce these functionalities at the meso-positions (section 2.1). However, the high reactivity of organolithium reagents is problematic in terms of functional group compatibility. The most popular methodologies for the introduction of aryl groups at the porphyrin periphery are catalytic cross-coupling reactions. In this case, a multistep synthesis involving the preparation of bromide or metalated porphyrins is usually required. Palladium-catalyzed C−H arylation reactions of aromatic arenes with aryl halides are rapidly growing in popularity, and this methodology is also applicable for functionalization of porphyrins.147 Yorimitsu, Osuka, and co-workers reported direct C−H arylation of porphyrins.148 In this case, the

introduction of various functionalities through cross-coupling reactions. The borylation protocol can also be applied to other porphy rin analog ues. Boryl at ion of 5,10,15-t ris(pentafluorophenyl)corrole 23.1 proceeded at the 2-position of the corrole with high regioselectively (Scheme 23).144 The Osuka group reported regioselective diborylation of meso-free subporphyrin (Scheme 24).145 Treatment of mesofree subporphyrin 24.1 with bis(pinacolato)diboron in THF under the catalysis of [Ir(OMe) (cod)]2 and dtbpy produced diborylated product 24.2 bearing an O-Bpin group at the axial position. Due to the instability of 24.2 on silica gel columns, 24.2 was isolated as diiodosubporphyrin 24.3 by treatment with CuI and N-iodosuccinimide. In the case of expanded porphyrins, direct borylation of the meso-aryl substituents of [26]hexaphyrin(1.1.1.1.1.1) 25.1 and [28]hexaphyrin(1.1.1.1.1.1) 25.2 was investigated by the Osuka and Shinokubo groups (Scheme 25).146 In this case, the use of [26]hexaphyrin(1.1.1.1.1.1) 25.1 as a substrate resulted in quantitative formation of the reduced [28]hexaphyrin 25.2 M

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 26. Palladium-Catalyzed Direct Arylation of Porphyrins

a

A solution of Pd(OAc)2 (20 mol %) and DavePhos (40 mol %) in DMA was added every 20 h.

reaction between 5,10,15-tris(3,5-di-tert-butylphenyl)porphyrin Ni(II) 26.1Ni and 2-bromonaphthalene, in the presence of pivalic acid, K2CO3, and a catalytic amount of Pd(OAc)2, afforded diarylated product 28.1a in excellent yield (Scheme 28). Notably, the product yield (97%) was higher than that under the previous conditions with DavePhos (84%). This reaction worked well for other π-extended aryl bromides. The installation of a 1-naphthyl group was accomplished in spite of its steric demand. More bulky groups, such as the 9-anthryl substituent, could be introduced in moderate yield. The structures of the products 28.1a and 28.2c were clearly elucidated by X-ray diffraction analysis. The dihedral angle between the aryl group and the porphyrin plane in 28.1c was larger than that in 28.1a. However, due to the perpendicular conformation of the aryl groups to the porphyrin plane, bathochromically shifted electronic absorption spectra were observed, indicating the presence of π-electronic interactions. Direct C−H arylation also worked well for mesoformylporphyrin 29.1 (Scheme 29).151 However, the reaction proceeded more slowly under the same conditions as mesounsubstituted porphyrins, affording the corresponding monoarylated product 29.2 in a moderate yield. The second addition of the catalyst into the reaction media was effective for further arylation, yielding diarylated product 29.3. The identity of the products was unambiguously elucidated by X-ray diffraction analysis. Diarylated product 29.3 shows the largest tilting angle of the β-aryl groups. The formyl group can be transformed into various functional groups. The reduction of 29.2 followed by its treatment with trifluoroacetic acid provided a ring-fused

pivalate-assisted conditions developed by Fagnou were effective.149 Treatment of Ni(II) 5,10,15-tris(3,5-di-tertbutylphenyl)porphyrin 26.1Ni with various aryl bromides and pivalic acid under the catalysis of Pd(OAc)2 and DavePhos furnished the corresponding diarylated products 26.2 in good yields (Scheme 26). The 1H NMR spectra and X-ray diffraction analysis of the products revealed that the reaction proceeded with perfect regioselectivity at the β-positions. The β-aryl groups were twisted relative to the porphyrin plane by about 50°. The perfect regioselectivity at the β-positions can be explained by steric effects. According to Fagnou and coworkers, the mechanism of C−H arylation assisted by pivalate consists of four steps described in Scheme 27. Among these steps, the concerted-metalation/deprotonation (CMD) process is presumably rate-determining. In this CMD step, the substituted β-protons are the most accessible among the peripheral protons, including the meso-protons. This reaction could not be applied to Zn(II) porphyrins. Regardless of its limitation to Ni(II) porphyrin complexes, this direct arylation methodology is applicable for porphyrins with various meso-aryl substituents (Scheme 26). In particular, when the 5,15diarylporphyrin Ni(II) complex 14.1Ni was subjected to the C−H arylation conditions, various tetraarylated products 26.3 were obtained. However, products were formed in low yields considering the use of electron-rich aryl bromides. In these cases, the yields were improved by the addition of extra Pd(OAc)2 and DavePhos into the reaction mixture every 20 h. The direct C−H arylation reaction proceeded for π-extended aryl bromides under phosphine-ligand-free conditions.150 The N

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 27. Plausible Mechanism of the Direct Arylation of Porphyrins

Scheme 28. Direct Arylation of a Porphyrin with Bulky Aryl Bromides

K3PO4 provided fused product 30.2Ni with a direct linkage between the phenyl ortho-carbon and the adjacent β-position of the porphyrin (Scheme 30).152,153 This reaction can also be applied to Cu(II) complex 30.1Cu. Double cyclization of 5,15di(2-iodophenyl)porphyrin 30.3 furnished two regioisomers, 30.4 and 30.5. The group of Cammidge also reported that

product. McMurry coupling of 29.2 resulted in formation of a vinylene-bridged twisted diporphyrin in 55% yield. Boyle reported palladium-catalyzed intramolecular cyclization between meso-aryl substituents and the porphyrin periphery. Treatment of meso-(2-iodophenyl)porphyrin Ni(II) 30.1Ni with a catalytic amount of Pd(PPh3)4 and an excess of O

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 29. Direct Arylation of meso-Formylporphyrin

Scheme 30. Palladium-Catalyzed Intramolecular Cyclization of meso-(2-Iodophenyl)porphyrins

P

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 31. Palladium-Catalyzed Intramolecular Cyclization of meso-Naphthylporphyrin

Scheme 32. Synthesis of meso−β Thieno-Fused Porphyrins through Intramolecular C−H/C−Br Cyclization

oxidative addition of the C−Br bond to Pd(0), carbopalladation at the triple bond of the alkyne substrates affords a palladacycle through intramolecular C−H cleavage at the adjacent βposition. Finally, reductive elimination provided the cyclization products. According to the same protocol, silyldehydropurpurins 34.2 were synthesized through a reaction with trialkylsilylacetylenes, which were then converted to unsubstituted 7,8dehydropurpurin (Scheme 34).158 This procedure can be applied to 1,4-diphenylbutadiyne. Under similar reaction conditions, 2-fold annulation of meso-bromoporphyrin 34.1Ni furnished 8,8-linked 7,8-dehydropurpurin dimers 35.3, 35.4, and 35.5 in good yields (Scheme 35).159 This [3 + 2] intramolecular annulation protocol can be applied to cyclic olefins. The reaction of 34.1Ni with norbornene proceeded smoothly, affording 36.1 in 79% yield (Scheme 36).160 X-ray diffraction analysis of 36.1 revealed that the addition of norbornene to porphyrin proceeded with perfect regioselectivity in an exo-fused manner. Using other cyclic olefins such as benzoaza- or benzooxanorbornenes also produced the corresponding annulated products 36.2 and 36.3 in good yields. Notably, these products are chiral. Treatment of 34.1Ni with norbornadiene provided norbornane-bridged diporphyrins through double [3 + 2] annulation.161

palladium-catalyzed intramolecular cyclization of porphyrin triflate 31.1 afforded 31.2 in 14% yield (Scheme 31).154 This intramolecular fusion reaction was utilized for the creation of antiaromatic molecules based on porphyrins. Matsuo and coworkers developed the synthesis of thieno-fused porphyrins 32.3 and 32.5 with meso−β fused structures.155 Fused porphyrins 32.3 and 32.5 were prepared by palladium-catalyzed intramolecular C−H/C−Cl or C−H/C−Br cross-coupling reactions of 32.2 and 32.4, respectively (Scheme 32). The contribution of the antiaromatic character of 32.3 and 32.5 was investigated by theoretical analysis and optical characterizations. An intermolecular meso−β cyclization reaction at the porphyrin periphery was developed by Osuka, Shinokubo, and co-workers. The authors reported the novel synthesis of dehydropurpurins through palladium-catalyzed [3 + 2] annulation with alkynes (Scheme 33).156,157 Treatment of meso-bromo-5,15-(3,5-di-tert-butylphenyl)porphyrin Ni(II) 33.1Ni with various alkynes under the catalysis of Pd2dba3 and (o-tol)3P furnished meso−β cyclized products 33.2Ni in good yields. Cu(II) and Zn(II) complexes 33.1Cu and 33.1Zn could be used as the substrates, providing the corresponding 7,8-dehydropurpurins 33.2Cu and 33.2Zn in good yields. Scheme 33b illustrates a plausible reaction pathway. After Q

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 33. Synthesis of 7,8-Dehydropurpurins through Palladium-Catalyzed Intramolecular [3 + 2] Annulation with Alkynes

Scheme 34. Synthesis of Silyldehydropurpurins by Intramolecular [3 + 2] Annulation

reaction proceeded with perfect regioselectivity without any byproducts, except for the debrominated porphyrin. A crossover experiment suggested the plausible reaction mechanism described in Scheme 38. Aratani and Osuka also reported the synthesis of intermolecularly cyclized product 39.3 through palladiumcatalyzed cross-coupling reaction of meso-borylporphyrin with 1,2-diiodobenzene.163 In this reaction, 1,2-phenylene-bridged porphyrin dimer 39.2 was also isolated in 5% yield as the side product. The yield of 39.3 was improved under strongly basic conditions, namely the use of Ba(OH)2 as a base (Scheme 39).

Another palladium-catalyzed direct functionalization of porphyrin skeletons was reported by the group of Shinokubo and Osuka. When meso-bromoporphyrin 37.1Zn was subjected to the Mizoroki−Heck reaction conditions using Herrmann’s catalyst (Scheme 37), the meso−β singly linked dimer 37.2 was unexpectedly obtained in the presence of zirconium tetrachloride.162 The structure of the dimer was clearly determined by X-ray diffraction analysis. Finally, the use of indium(III) chloride as a Lewis acid improved the yield. After several experiments, the presence of a small amount of water was found to be important for reliable and reproducible results. The R

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 35. Two-Fold [3 + 2] Annulation to 8,8-Linked 7,8-Dehydropurpurin Dimers

Scheme 36. Palladium-Catalyzed [3 + 2] Annulation of meso-Bromoporphyrin with Norbornene Derivatives

S

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 37. Synthesis of meso−β Linked Diporphyrin

Scheme 38. Proposed Reaction Mechanism

various aromatic compounds under copper-catalyzed con-

3. REACTIONS OF HALOGENATED AND METALATED PORPHYRINS

ditions.164 They demonstrated the application of their reaction

3.1. Core-Functionalization

to porphyrin substrates. Treatment of 5-bromo-10,20-diary-

3.1.1. Metalation of Haloporphyrins. Organolithium and organomagnesium reagents have a long history, but they are still highly useful in organic synthesis.165,166 Introduction of magnesium and lithium at the peripheral positions of porphyrins is a fascinating strategy, which enables the creation of porphyrin-based organometallic reagents having high nucleophilicity and the potential to participate in a variety of bond-forming processes. However, the generation of such

Itami and co-workers reported the direct C−H imidation of

lporphyrin 33.1H with N-fluorobenzenesulfonimide in the presence of catalytic CuBr and 6,6′-dimethylbipyridyl provided meso-imidated product 40.1 in 59% yield (Scheme 40). This reaction also worked for meso-tetraarylporphyrin Ni(II) 40.2, affording β-imidated product 40.3 in 23% yield. T

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 39. Intermolecular meso−β Fusion Reaction with meso-Borylporphyrins

Scheme 40. Cu-Catalyzed Direct Amination of Porphyrins

important intermediates in palladium-catalyzed cross-coupling reactions. When chiral diphosphine ligands such as BINAP were used, chiral porphyrins were obtained.175 The induced asymmetry in the porphyrin was detected by 1H NMR and CD spectroscopy. Under Sonogashira coupling conditions, meso-platinoporphyrin 44.1 can be coupled with terminal alkynes. The use of meso-ethynylporphyrin 44.2 furnished platinum-bridged diporphyrin 44.3 (Scheme 44).176 The use of pyridine coordination to the platinum σ-complexes enabled rapid construction of a variety of multiporphyrin assemblies such as 44.6.177 Selective monobromination of 5,15-disubstituted porphyrins is difficult due to the similar reactivity of two unsubstituted meso positions. The reaction of 5,15-disubstituted porphyrins with NBS usually provides a mixture of mono- and dibrominated products and tedious purification is needed. Arnold and co-workers employed the reactivity of σ-complexes to achieve selective bromination (Scheme 45).178 Treatment of dibromoporphyrin 45.1H with 1.0 equiv of Pd2dba3/DPPE selectively provided monopalladium complex 45.2, which was converted to bromoporphyrin 12.1H through hydrodepalladation. In addition, the reaction of meso-palladioporphyrin 43.1 with iodine afforded meso-iodoporphyrin 45.3 in a good yield,

organometallic porphyrins is difficult because porphyrin cores are reactive to nucleophilic species. Fujimoto, Yorimitsu, and Osuka reported the generation of porphyrinyl Grignard reagents 41.2 and 41.5 through magnesium−iodine exchange reactions (Scheme 41).167 The obtained porphyrinyl Grignard reagents were nucleophilic enough to react with a variety of electrophiles. Porphyrinylmagnesium reagents also underwent transmetalation to yield porphyrinyl copper and zinc species. They also succeeded in the preparation of porphyrinyllithium reagents 42.1 and 42.3 as more reactive nucleophiles through a lithium−iodine exchange reaction. Transformation of these nucleophilic organometallic porphyrins into boryl, silyl, and 1hydroxyalkylporphyrins was also demonstrated (Scheme 42).168−170 meso-Bromoporphyrin 12.1H underwent facile oxidative addition to Pd(0) and Pt(0) species such as Pd(PPh3)4, Pd2dba3/PPh3, Pd2dba3/DPPE, and Pt(PPh3)3 in degassed hot toluene to afford air-stable meso-palladio- and meso-platinoporphyrin complexes such as 43.1 (Scheme 43).171−173 The structures of these palladium and platinum σ-complexes were clearly elucidated by X-ray crystallographic analysis. The same reactivity was confirmed for porphyrin metal complexes with various central metals.174 These meso-metalated species are U

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 41. Magnesiation of Iodoporphyrins

Scheme 42. Lithium−Iodine Exchange Reaction of Iodoporphyrins

Scheme 43. Formation of meso-Palladioporphyrin through Oxidative Addition to Pd(0)

The chemistry of porphyrin metal σ-complexes was merged with that of directly linked porphyrins (Scheme 46).180

thus allowing conversion of bromoporphyrins to iodoporphyrins.179 V

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 44. Multi-Porphyrin Assemblies Prepared Based on meso-Platinoporphyrins

Scheme 45. Synthetic Use of meso-Palladioporphyrins

W

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 46. Oxidative Dimerization of meso-Platinoporphyrin

Scheme 47. Palladium-Catalyzed Borylation of Bromoporphyrins

Treatment of meso-platinoporphyrin Zn(II) complex 46.1 with AgPF6 induced a meso−meso coupling reaction to afford

diporphynyl bisplatinum complex 46.2. Furthermore, DDQ/ Sc(OTf)3 oxidation of bisplatinum complex 46.2 furnished X

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 48. Suzuki−Miyaura Cross-Coupling Reaction of a meso-Borylporphyrin

alities. The most popular method is the Suzuki−Miyaura crosscoupling reaction. Via this method, borylated porphyrins have been widely used as an alternative to brominated porphyrins. Therien and co-workers reported a synthetic route to mesoarylporphyrins from meso-borylated porphyrins.186 According to this approach, carbazole-linked porphyrin dimer 48.1 could be synthesized in a 76% yield from meso-boryl-5,15diphenylporphyrin 47.1Zn. Owing to the mild reaction conditions, phenylalanine-appended porphyrin 48.2 was synthesized via cross-coupling with iodophenylalanine, which includes functionalities that are reactive toward nucleophiles (Scheme 48). The group of de Lera reported the synthesis of A2B2-type porphyrins through Suzuki−Miyaura cross-coupling between meso-bromoporphyrin 49.1 and arylboronic acids (Scheme 49).192 They developed milder reaction conditions for coupling by using catalytic amount of Pd(PPh3)4 and aqueous Na2CO3 as a base in DMF, which ensured good coupling rates for porphyrins functionalized with electrophilic groups. Under these conditions, an excess of boronic acid is required to obtain isolable products due to adventitious hydrolytic protiodeborylation. Shi and Boyle reported the preparation of A2B2-type porphyrins through Suzuki−Miyaura cross-coupling reactions.193 They described that the use of neopentyl glycolates for electron-deficient aryl substrates suppressed deborylation, allowing isolation of the coupling products 49.4 in 73−78% yield. They also demonstrated the synthesis of AA′B2-type porphyrin 49.6 by using dibromoporphyrin 49.5 as a substrate. The cross-coupling of sterically hindered substrates has often been problematic. Aratani and Osuka reported the synthesis of a meso−meso-linked hybrid porphyrin array via Suzuki−Miyaura cross-coupling.194 During the initial attempt of the reaction between meso-borylporphyrin 50.1Zn and meso-bromoporphyr-

triply linked diporphyrin 46.3 without breaking the carbon− platinum σ-bonds. Aryl halides can be converted to the corresponding arylboronic esters through palladium-catalyzed borylation with pinacolborane (pinBH) or bis(pinacolato)diboron (pinB)2 as boron sources.181−185 This methodology was applied to the synthesis of peripherally borylated porphyrin 47.1Zn for the first time by Therien and co-workers (Scheme 47).186 βBorylporphyrin 47.3 was then prepared by Chang and Nocera for the synthesis of a β−β-linked diporphyrin.187 Zhang and Suslick also reported the synthesis of TPP-type β-borylporphyrins.188 Treatment of Zn(II) 2-bromotetraphenylporphyrin 47.4Zn with bis(pinacolato)diboron, in the presence of Pd(PtBu3)2 as a catalyst and KOAc as a base, provided Zn(II) βborylporphyrin 47.5 in almost quantitative yield. Bringmann and co-workers improved this protocol, and the use of PdCl2(dppf) as a catalyst and a toluene/H2O two-phase solvent system led to the formation of 2-borylporphyrin in good yield.189 These boryl porphyrins are a convenient platform for the synthesis of functionalized porphyrin derivatives. The position of boryl substituents is predetermined by the position of the halide substituents. Consequently, regioselective introduction of bromide is the key for this strategy. 3.1.2. Suzuki−Miyaura Coupling. 3.1.2.1. Suzuki− Miyaura Coupling at meso-Positions. Peripherally halogenated porphyrins are useful precursors for the introduction of various substrates, such as aryl groups. Halogenation of porphyrins at the meso-positions is usually achieved through electrophilic substitution with NBS or I2 and Ag salts.190,191 Among the various transformations, palladium-mediated coupling reactions are versatile and straightforward approaches to making carbon−carbon linkages with a range of functionY

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 49. Suzuki−Miyaura Cross-Coupling at meso-Positions

Pd(PPh3)4 and Cs2CO3 worked well for the coupling of iodoferrocene with 5,15-diphenyl-10-borylporphyrin 47.1H (Scheme 53).197 However, the use of 9-anthranylboronic ester as a coupling partner resulted in only a low yield of the product. Anderson et al. reported the synthesis of bisanthracene-fused porphyrins 54.3a and 54.3b through the crosscoupling of meso-borylated porphyrin 54.1Zn with 9-anthryl triflate or 9-bromoanthracene in the presence of a catalytic amount of Pd2dba3 and SPhos (Scheme 54).198 Locos et al. found that the use of AsPh3 as a ligand with PdCl2(PPh3)2 improved the yield of the reaction between dibromoporphyrin 52.1Ni and 9-borylanthracene to 78%.199 In this case, Ni(II) complexes afforded superior product yields to free-base and Zn(II) derivatives. The authors suggested that this was possibly due to the more ruffled conformation of Ni(II) porphyrins, which allowed more facile access of the bulky anthryl group to the meso-positions. Other than arylboronic acids or esters, cross-coupling with allyl- and allenylboronic pinacol esters was reported by Senge et al. (Scheme 55).200 By using Pd(PPh3)4 and K3PO4 in refluxing toluene, meso-allylporphyrins 55.2 were obtained in good yields (50−95%). The reaction rate was enhanced by the use of

in 50.2, they obtained the desired coupling product 50.3 only in low yield using Pd(PPh3)4 as a catalyst and K2CO3 as a base in toluene. They then found that coupling in a DMF/toluene solution in the presence of a catalytic amount of Pd(PPh3)4 and 1.5 equiv of Cs2CO3 was effective, affording 50.3 in 62% yield (Scheme 50). Under these conditions, they succeeded in the synthesis of heterometallic hybrid porphyrin arrays including meso−meso-linked porphyrin octamer 50.5. The group of Osuka also reported the synthesis of L-shaped meso−mesolinked Zn(II) triporphyrin 51.3 through the Suzuki−Miyaura cross-coupling reaction. In this case, the use of tri-2furylphosphine as the ligand was effective to produce the trimer 51.3 in 24% yield (Scheme 51).195 The use of Cs2CO3 as a base worked well for coupling reactions with other bulky boronic acids, such as polycyclic aromatic hydrocarbons (PAHs).196 The reaction with 1naphthyl, 1-pyrenyl, 1-perylenyl, and 1-coronenyl boronic pinacol ester with 5,15-bis(3,5-di-tert-butylphenyl)-10,20-dibromoporphyrin Zn(II) 52.1Zn afforded the corresponding disubstituted products 52.2 in good yields (30−85%) through the use of Pd(PPh3)4/Cs2CO3 in toluene (Scheme 52). Senge and co-workers disclosed that a catalyst combination of Z

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 50. Synthesis of meso−meso-Linked Porphyrin Oligomers through Suzuki−Miyaura Cross-Coupling

acids or esters. Senge and co-workers demonstrated the straightforward synthesis of meso-alkylporphyrins via Suzuki− Miyaura cross coupling with alkyltrifluoroborates (Scheme 56). The highest yield was achieved under PdCl2(dppf)/Cs2CO3 conditions in a THF/H2O mixture. 3.1.2.2. Suzuki−Miyaura Coupling at β-Positions. Suzuki− Miyaura coupling at the β-positions has been utilized for the synthesis of highly congested porphyrins. The group of Chan reported the coupling of β-bromoporphyrins with arylboronic acids and the synthesis of dodecaarylporphyrins.204 The reactions were performed in the presence of Pd(PPh3)4 and K2CO3 under anhydrous conditions, producing the correspond-

PdCl2(dppf) as a palladium catalyst. They also investigated Suzuki−Miyaura cross-coupling between meso-bromotriphenylporphyrin 37.1 with allenylboronic acid pinacol ester.201 As alternatives to boronic acids or esters, potassium organotrifluoroborates have been found to be stable but reactive reagents for the Suzuki−Miyaura cross-coupling reaction.202,203 The use of potassium organotrifluoroborates suppresses β-hydride elimination in the case of coupling reactions involving alkylboronic acid derivatives. Their most beneficial aspect is their high stability in air, which has led to the commercial availability of a number of potassium organotrifluoroborates that cannot be obtained as boronic AA

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 51. Synthesis of L-Shaped meso−meso-Linked Triporphyrin

ing β-arylated porphyrins from β-mono, tetra-, and octabromoporphyrins 57.1, 57.3, and 57.5 in high yields (Scheme 57). Smith and co-workers screened and optimized this reaction to achieve the synthesis of a series of dodecaarylporphyrins (57.6). A wide range of aryl substituents was successfully introduced at porphyrin peripheries.205 However, reactions with arylboronic acids bearing electron-deficient groups at their ortho-positions resulted in no fully substituted porphyrins. In addition, porphyrins bearing heteroatoms at the ortho-position of their aryl rings, or with strongly electron-withdrawing heterocycles, did not lead to fully substituted products. The group of Sankar reported the synthesis of tri-β-substituted unsymmetrical porphyrin 58.2 via cross-coupling of 2-formyl12,13-dibromotetraphenylporphyrin 58.1 with phenylboronic acid (Scheme 58).206,207 Another Suzuki−Miyaura cross-coupling reaction was reported with β-borylated porphyrin 59.1 by Osuka et al. (Scheme 59).208 Bringmann developed the synthesis of β−β directly linked porphyrin dimer 60.3 through the Suzuki− Miyaura cross coupling reaction (Scheme 60).189 They prepared β-borylated tetra(p-tolyl)porphyrin 60.1 as a coupling partner from free-base 2-bromotetra(p-tolyl)porphyrin 60.2 via palladium-catalyzed borylation (described in section 3.1.1).

Cross-coupling between 60.1 and 60.2 was achieved with Pd(PPh3)4 and using K3PO4 as a weak base, producing β−β linked porphyrin dimer 60.3 in 26% yield. The yield was increased to 59% by using Ba(OH)2 in a toluene/water mixture. The central metal influenced the yield and hydrodebromination decreased the yield to 33% in the case of Cu(II) complex. Under the same conditions, the authors also synthesized β−meso,meso−β linked porphyrin trimer 61.4, described as a “basket handle” porphyrin array, through the reaction of meso-dibromoporphyrin 61.2 with β-borylated tetraarylporphyrin 61.1 (Scheme 61).209 For coupling with corannulene, the Pd(II) precatalyst complex 62.1, developed by Buchwald,210 was effective. Suzuki−Miyaura cross-coupling of β-bromoporphyrin 62.2Zn or meso-bromoporphyrin 34.1Ni with 7-boryldibenzo[a,g]corannulene furnished the corresponding singly linked products 62.3 or 62.4 in 59−75% yields (Scheme 62).211 meso,β Double cross-coupling was reported by Yorimitsu and Osuka.212 The authors prepared Ni(II) meso,β-dichloroporphyrin 63.1 through chlorination of β-chloroporphyrin with Palau’ Chlor. Treatment of Ni(II) meso,β-dichloroporphyrin 63.1 with 2-thienylboronic acid in the presence of a catalytic amount of Pd(OAc)2 and SPhos furnished meso,β-bis(2AB

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 52. Cross-Coupling Reactions of Zn(II) 5,15-Dibromoporphyrin with PAH Boronic Acids

Scheme 53. Cross-Coupling of a Borylated Porphyrin with Iodoferrocene

thienyl)porphyrin 63.2 in 82% yield (Scheme 63). SPhos ligand was also effective for the coupling of Zn(II) β-bromotetrapheAC

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 54. Cross-Coupling Reactions with 9-Haloanthracenes

bridged diporphyrin,219 pyrrole- and azobenzene-bridged porphyrin nanorings,220,221 and a doubly bridged porphyrinperylene-porphyrin triad.222 This protocol was further extended to the construction of directly meso−β doubly linked porphyrin ring 67.1 through the coupling of 22.1H with mesodibromoporphyrin 52.1H.223−225 This “Lego block” strategy furnished a variety of directly linked cyclic oligoporphyrins. The direct β−β-linked porphyrin array 67.3 was synthesized via cross-coupling of tetraborylporphyrin 18.3Ni and dibromodiporphyrin 67.2 (Scheme 67). Song and co-workers reported the synthesis of “earring” porphyrin 68.3 through the coupling between Ni(II) βtetrabromoporphyrin 68.1 and α,α′-diboryltripyrrane 68.2 (Scheme 68).226 In this case, the authors used an air-stable phosphonium salt t-Bu3P•HBF4 to generate tri(tert-butyl)phosphine in situ under basic conditions. Suzuki−Miyaura coupling also worked well for other porphyrinoids such as diazaporphyrins and subporphyrins. Matano and co-workers reported the introduction of aryl groups to 5,15-diazaporphyrins through Suzuki−Miyaura crosscoupling reactions (Scheme 69).227 The authors found the use of CyJohnphos was effective to afford the coupling products 69.2 in good yields. They also synthesized porphyrin− diazaporphyrin dyads via cross-coupling of 2-bromo-5,15diazaporphyrin 69.1Zn with meso-borylporphyrin 69.3 or β-

nylporphyrin 47.4Zn with borylated Zn(II) phthalocyanine 64.1, affording porphyrin−phthalocyanine hybrid 64.2 (Scheme 64).213 Ethynyl substituents can be introduced through Suzuki− Miyaura coupling with β-borylated porphyrins. In the presence of Pd2dba3 and Xantphos as precatalysts and K3PO4 as base, treatment of Ni(II) β,β′-diborylporphyrin 22.1Ni with bromoethynyl-tert-butyldimethylsilane in a 1,4-dioxane/H2O mixture provided diethynylporphyrin 65.1Ni in good yield (Scheme 65).214 This protocol is superior to the Sonogashira coupling reaction, which requires halogenated porphyrins prepared from β-borylporphyrins. The Suzuki−Miyaura cross-coupling reaction has been used extensively for the synthesis of cyclic oligoporphyrins. The group of Osuka reported the synthesis of 2,6-pyridylenebridged β-to-β porphyrin nanorings including a “porphyrin nanobarrel” through the Suzuki−Miyaura cross-coupling reaction of Ni(II) tetra(6-bromopyridyl)porphyrin 66.1 and Ni(II) β-tetraborylporphyrin 18.3Ni.133,215 The coupling reaction using Pd2dba3, PPh3, Cs2CO3, and CsF in a toluene/ DMF mixed solvent afforded porphyrin nanobarrel 66.2 in 10% yield (Scheme 66). This protocol worked effectively to afford various arylene- and heteroarylene-bridged cyclic oligoporphyrins, such as 2,5-thienylene-bridged cyclic porphyrins,216,217 2,6pyridylene-bridged porphyrin nanorings,218 a 3,6-carbazolylAD

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 55. Cross-Coupling with Allyl- and Allenylboronic Esters

Scheme 56. Cross-Coupling with Potassium Organotrifluoroborates

borylporphyrin 69.5Zn.228 In these cases, they used Pd(PPh3)4 as the catalyst in a toluene/DMSO mixed solvent system, and the corresponding products 69.4 and 69.6 were obtained in 87 and 64% yields, respectively. For subporphyrins, palladium-catalyzed Suzuki−Miyaura cross-coupling of meso-bromosubporphyrin 70.1 with 2-ethyl-

4-methoxyphenylboronic acid afforded the cross-coupling product 70.2 in 74% yield in the presence of a Pd2dba3/ SPhos catalytic system (Scheme 70). The coupling with styrylboronic acid was achieved with a catalytic amount of Pd(PPh3)4 in THF, furnishing styrylsubporphyrin 70.3 in 71% yield.229 AE

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 57. Suzuki−Miyaura Cross-Coupling of β-Brominated Porphyrins

Scheme 58. Synthesis of Unsymmetrically-Substituted Porphyrins via Suzuki−Miyaura Cross-Coupling Reactions

3.1.3. Sonogashira Coupling. Sonogashira coupling reactions of meso-halogenated porphyrins with alkynes are also a popular methodology for the introduction of functionalities at porphyrin peripheries. Dolphin and coworkers reported the introduction of an ethynyl substituent at the meso position of 5,15-diphenylporphyrin.230 Treatment of 10-iodo-5,15-diphenylporphyrin Zn(II) 45.3Zn with a range of terminal alkynes in THF/Et3N in the presence of PdCl2(PPh3)2 and CuI furnished the corresponding alkynylporphyrins 71.1 in 50−90% yields (Scheme 71). However,

hydroalkynylation of the introduced silylethynyl group in 71.2a afforded meso-enynylporphyrins 71.2b.231 The formation of an enyne adduct was suppressed by using Pd2dba3 or Pd(PPh3)4 as a palladium (pre)catalyst. The group of Therien reported the Sonogashira coupling of meso-bromoporphyrins Zn(II) with phenylacetylene.232 The group of Anderson utilized Sonogashira coupling reactions for preparation of π-conjugated porphyrin oligomers.233 The authors reported the synthesis of an acetylene-linked porphyrin dimer in 70% yield through a reaction between mesoAF

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 59. Synthesis of meso−β Linked Diporphyrin through Suzuki−Miyaura Cross-Coupling

Scheme 60. Synthesis of meso−β Linked Diporphyrin through Suzuki−Miyaura Cross-Coupling

Scheme 61. Synthesis of “Basket Handle” Porphyrins

ethynylporphyrin Zn(II) and meso-bromoporphyrin. The different reactivity between bromo and iodo groups in the

Sonogashira cross-coupling reaction allowed the expeditious preparation of unsymmetrically substituted porphyrins. The AG

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 62. Synthesis of Corannulene−Porphyrin Dyads

Scheme 63. meso,β Double Cross-Coupling of Ni(II) meso,β-Dichloroporphyrin

group of Boyle demonstrated the synthesis of 5-alkenyl-15alkynylporphyrins 72.3. Zn(II) bromoiodoporphyrin 72.1 first reacted with terminal alkynes in the presence of a catalytic amount of PdCl2(PPh3)2 and CuI. The remaining bromo group in 72.2 was then transformed into an alkenyl group through Migita−Kosugi−Stille cross-coupling with the corresponding alkenylstannane reagents (Scheme 72).234 Sonogashira coupling reactions at the β-positions using βhaloporphyrins as substrates were reported by the group of van Lier.235 Osuka and co-workers also prepared porphyrin bistriflate 22.3Ni through iridium-catalyzed direct borylation.142 The reaction of 22.3Ni with trimethylsilylacetylene in the presence of PdCl 2 (PPh 3 ) 2 and CuI afforded the corresponding ethynylated products 73.1 in good yield (Scheme 73). In this case, the use of a Et3N/THF mixed solvent system was effective. They also reported that Sonogashira coupling of β,β-diiodoporphyrin 22.4Ni with

triisopropylsilylacetylene successfully furnished the corresponding diethynylated product 73.2 in 92% yield.143 The groups of Kim and Sessler reported a direct palladiumcatalyzed alkynylation reaction from Zn(II) 2-borylporphyrin (Scheme 74).236 Treatment of 74.1 with terminal alkynes under aerobic oxidative conditions in the presence of a catalytic amount of Pd(OAc)2 and 10 equiv of sodium hydride afforded the corresponding 2-ethynylporphyrin 74.2 in good yield. Although electron-deficient alkynes were less effective than electron-rich alkynes, this procedure provides a variety of ethynylporphyrins. Sonogashira coupling was also effective for introduction of alkynyl substituents to porphyrinoids. The group of Matano reported the synthesis of ethynylene-bridged 5,15-diazaporphyrin dimer 75.2 through coupling of Ni(II) 3-bromo-5,15diazaporphyrin 69.1Ni with Ni(II) 3-ethynyl-5,15-diazaporphyrin 75.1 (Scheme 75).237 The group of Osuka reported the introduction of alkynyl substituents at the meso-position of AH

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 64. Synthesis of a Porphyrin−Phthalocyanine Hybrid

Scheme 65. Introduction of Ethynyl Groups at β-Positions through Suzuki−Miyaura Cross-Coupling

catalytic amount of Pd(PPh3)4, the reaction was carried out at 60 °C for 12 or 48 h. On the other hand, reactions in the presence of PdCl2(dppf) required milder reaction conditions. Complete conversion of the starting material was achieved within 12 h at room temperature. Coupling reactions with alkyl, aryl, vinyl, and benzylic organometallic reagents depicted in Scheme 77 were successfully carried out to yield novel porphyrins. This methodology allows introduction of various types of substrates, including those with sterically demanding, electron-donating, and electron-withdrawing groups, at the meso-positions. In addition, by using these coupling reactions, the preparation of divinylporphyrin 77.4, which is difficult to obtain from acrolein and pyrrole derivatives, was simplified. In particular, bipyridylmethylporphyrin 77.5 was prepared in good yield. This porphyrin was difficult to be prepared by Adler− Longo or Lindsey protocols due to the presence of the keto− enol equilibrium of the corresponding aldehyde. They also succeeded in the introduction of meso-trimethylsilylethynyl

subporphyrins through Sonogashira cross-coupling (Scheme 76).238 A bromo-N-confused porphyrin also underwent Sonogashira cross-coupling to afford the corresponding ethynyl-substituted product. In this case, the reaction was performed in the presence of Bu4NOAc without the addition of CuI and amine.239 3.1.4. Migita−Kosugi−Stille, Negishi, and Kumada− Tamao−Corriu Couplings. Other catalytic cross-coupling methodologies using organometallic species for C−C bond formation are the Migita−Kosugi−Stille coupling (with organotin reagents), the Negishi coupling (with organozinc reagents), and the Kumada−Tamao−Corriu coupling (with Grignard reagents). Therien et al. reported the first application of the metal-mediated cross-coupling reaction with halogenated porphyrins (Scheme 77).240 The authors used Zn(II) 5,15dibromo-10,20-diphenylporphyrin 45.1Zn as a substrate and organozinc chlorides or organotin reagents as cross-coupling partners. The reaction conditions required were found to be dependent on the choice of catalyst. In the presence of a AI

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 66. Synthesis of a Porphyrin Nanobarrel

cyclobutenedione-substituted porphyrin 81.1 via cross-coupling of meso-bromoporphyrin 12.1Zn with tributylstannylcyclobutenedione (Scheme 81).248 In this case, the use of a Pd2dba3/ AsPh3 catalytic system allowed efficient coupling to obtain the desired products in good yields. Odobel reported the coupling of meso-iodoporphyrin 41.4Zn with 5-tributylstannyl-2,4pentadienal acetal to afford highly π-conjugated porphyrin 82.1 in 86% yield (Scheme 82).249 The group of Anderson developed an efficient synthetic route to meso-to-meso (E)vinylene-linked porphyrin dimer 83.1 through palladiumcatalyzed Migita−Kosugi−Stille cross-coupling of bromoporphyrin 41.4Zn with bis(tributylstannyl)ethene (Scheme 83).250 In this case, the use of Baldwin’s protocol using CuI and CsF furnished the desired product 83.1 in 58% yield. The absence of either CuI or CsF decreased the yields of products. The Migita−Kosugi−Stille coupling reaction of β-brominated porphyrins is a useful strategy for installation of alkenyl and alkynyl groups at the β positions. Smith and co-workers reported the preparation of Ni(II) 2,3-dialkynyl-5,10,15,20tetraphenylporphyrins 84.2 through Pd(0)-catalyzed crosscoupling of dibromoporphyrins 84.1 with alkynyltrimethylstannanes (Scheme 84).251 Further ethynylation was achieved by the use of octabromoporphyrin 85.1 as a substrate (Scheme 85).252 The Migita−Kosugi−Stille coupling was also applied to meso,β-dichloroporphyrin 63.1, as reported by Yorimitsu, Osuka, and co-workers.212 The coupling was carried out under the catalysis of Pd2dba3 and SPhos, providing the meso,βdiethynylated product 86.1 in 70% yield (Scheme 86). Senge

groups through Negishi cross-coupling reactions (Scheme 78).241 Takanami and co-workers developed palladium-catalyzed pentafluorophenylation, silylmethylation, and ethoxycarbonylmethylation of bromoporphyrins with organozinc reagents (Scheme 79).242−246 The efficient introduction of pentafluorophenyl groups at the meso- or β-positions of porphyrins was achieved through cross-coupling with (C6F5)2Zn in the presence of a Pd(OAc)2/t-Bu 3P catalytic system. The connection of the ethoxycarbonylmethyl group was conducted by Negishi coupling of meso-bromoporphyrin 79.2 with the corresponding zinc reagent in the presence of a catalytic amount of Pd(OAc)2 and PCy3. Silylmethyl-substituted porphyrins 79.4 and 79.5 were prepared through Kumada− Tamao−Corriu cross-coupling of bromoporphyrins 12.1H and 47.4H with silylmethyl Grignard reagents in the presence of catalytic Pd2dba3/Ph2P(O)H. Negishi coupling reactions were also utilized for introduction of alkyl substituents to mesobromosubporphyrin.238 The groups of Yorimitsu and Osuka reported the synthesis of triphenylsilane-fused porphyrins. To synthesize the precursors to these porphyrins, they employed Negishi coupling reactions of β-iodo-, β,β′-diiodo-, and meso-iodoporphyrins with 2(diphenylsilyl)phenylzinc using 2-dicyclohexylphosphino-2′,6′diisopropoxybiphenyl (RuPhos) as a ligand (Scheme 80).247 Migita−Kosugi−Stille couplings have been utilized for the introduction of vinylene substituents at the meso positions of porphyrins. Liebeskind et al. reported the preparation of AJ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 67. Direct β-to-β-Linked Cyclic Oligoporphyrins Prepared via a “Lego Block” Strategy

reported the synthesis of thiophene- and pyrrole-bridged 5,15-

and co-workers reported the Migita−Kosugi−Stille coupling reactions with arylstannane reagents. 2-Bromoporphyrin 47.4Zn was converted into 4-stannylphenylporphyrin 87.1 in 30% yield (Scheme 87).253 Migita−Kosugi−Stille coupling has been effective for the synthesis of a range of porphyrinoids. The group of Matano

diazaporphyrin dimers 88.1 through the cross-coupling of 3b r o m o - 5 , 1 5 - d i a z a p o r p h y r i n 6 9 .1 N i w i t h 2 ,5 - bis (tributylstannyl)heteroles (Scheme 88).254 Shinokubo and coworkers also reported the synthesis of β-pyridyldiazaporphyrin AK

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 68. Synthesis of “Earring” Porphyrins

Scheme 69. Suzuki−Miyaura Cross-Coupling Reactions of 5,15-Diazaporphyrins

AL

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 70. Cross-Coupling Reactions of a meso-Bromosubporphyrin

coupling of β-brominated porphyrins 92.1 (Scheme 92).258 The reaction of tetrabromotetraarylporphyrins 92.1 with an excess of methyl acrylate in the presence of Pd(OAc)2/PPh3 and K2CO3 furnished opp-dibenzoporphyrins 92.2. In this case, simple Mizoroki−Heck products were not obtained. A similar Mizoroki−Heck reaction also took place with Zn(II) porphyrin 92.1Zn to provide the corresponding free-base porphyrin 92.2H through demetalation during the reaction. According to this protocol, Kadish and Wang et al. reported the synthesis of push−pull opp-dibenzoporphyrin 93.5 (Scheme 93).259 The Mizoroki−Heck reaction of dibromotetraarylporphyrin 93.1 with acrylonitrile formed a benzene ring at the β,β′-positions, affording benzofused porphyrin 93.2. Bromination of 93.2 with NBS provided 12,13-dibromobenzoporphyrin 93.3 with high regioselectivity. Porphyrin 93.3 was then subjected to a Mizoroki−Heck reaction with p-methoxystyrene to furnish dibenzofused push−pull porphyrin 93.4. 3.1.6. Other Coupling Methodologies for C−C Bond Formation. This section describes coupling reactions with other carbon-based nucleophiles without the use of organometallic reagents. Suda and co-workers serendipitously discovered a procedure for the introduction of a cyano substituent at the mesoposition.260 The reaction of free-base bromodiphenylporphyrin 12.1H with 2-cyanoethylzinc bromide under the catalysis of Pd(PPh3)4 afforded Zn(II) meso-cyano-substituted porphyrin 94.1 in 45% yield (Scheme 94). The yield was improved to 94% with a Pd2dba3/PPh3 catalytic system in 1,4-dioxane at 95 °C. This cyanation protocol was applicable to a wide range of mono- and dibromoporphyrins. The meso-cyanated porphyrins were also synthesized with Zn(CN)2 as the coupling reagent (Scheme 95).261 Dibromoporphyrin 52.1H was subjected to the reaction conditions reported by Jin and Confalone,262 providing Zn(II) dicyanoporphyrin 95.1 in 68% yield along with cyanoporphyrin 19.11Zn as a byproduct (10%). The addition of a fresh portion of catalyst was required for a largescale reaction because an excess of the cyano anion was observed to poison the catalyst. Anderson and co-workers reported the synthesis of 5,15dialkylideneporphyrins via the Takahashi coupling reaction (Scheme 96).263 Treatment of Zn(II) 5,15-dibromoporphyrin

88.2 through Migita−Kosugi−Stille cross-coupling of 69.1Ni with 2-(tributylstannyl)pyridine.255 3.1.5. Mizoroki−Heck Reaction. For the introduction of alkenyl substituents, the Mizoroki−Heck reaction is a good choice. Arnold and co-workers developed the synthesis of a series of meso-alkenylporphyrins by using Mizoroki−Heck reactions.256 The palladium-catalyzed coupling of mesobromoporphyrins 37.1 with methyl acrylate, styrene, and acrylonitrile was successful, leading to the corresponding meso-alkenylporphyrins 89.1 (Scheme 89). The normal conditions for this reaction were found not to be effective because of the low solubility of bromoporphyrin substrates in DMF. Instead, a 50:50 mixture of DMF and toluene allowed the reactions to proceed. While in all other cases, only transalkenylporphyrins were isolated, a mixture of cis and trans isomers was obtained in the case of acrylonitrile, probably due to the small steric demand of the cyano group. In addition, the ratio of the two isomers was dependent on the central metal ion. The coupling between Ni(II) meso-vinylporphyrin 89.2 and meso-bromoporphyrin 37.1Ni was sluggish and debrominated porphyrin was the major product. Interestingly, the mesoto-β ethenylene-linked dimer 89.3 was obtained. In this case, the use of 2-(di-tert-butylphosphino)biphenyl as ligand improved the yield. The insertions of alkenyl groups at the β-positions of porphyrins through Pd-mediated reactions have been extensively studied. Smith and co-workers reported palladiumcatalyzed C−C bond formation with a Zn(II) 2,4-bismercurated porphyrin (Scheme 90).125 Bismercurated deuteroporphyrin 90.1 was prepared by mixing an excess of mercury(II) acetate with the zinc(II) deuteroporphyrin dimethyl ester. Compound 90.1 was reacted with acrolein in an acetonitrile/ DMF mixture in the presence of a trace amount of triethylamine and a stoichiometric amount of LiPdCl3 to afford the vinylated product 90.2 in 35% yield. The Mizoroki−Heck coupling of β-bromoporphyrins has been extensively investigated. The group of Risch succeeded in the cross-coupling of Zn(II) bromodeuteroporphyrin dimethyl ester 91.1 with a bifunctional α,β-unsaturated carbonyl compound under phase-transfer conditions (Scheme 91).257 Benzo-fused porphyrins were synthesized via Mizoroki−Heck AM

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 71. Sonogashira Coupling of meso-Iodoporphyrins

amount of PdCl2 and K2CO3 under phase-transfer conditions, intramolecular cyclization of 2-bromo-5,10,15,20-tetraphenylporphyrin 47.4Zn led to construction of a carbon−carbon bond between the β-position and ortho-carbon of the adjacent meso-aryl substituent. Ishizuka, Kojima, and co-workers developed an analogous multiple fusion reaction of a tetraphenylporphyrin.268,269 They found that [Pd(η3-C3H5)Cl]2 was effective to provide quadruply fused product 99.4 in 79% yield from tetrabromoporphyrin 99.1 (Scheme 99). Yorimitsu, Osuka, and co-workers reported 2-fold intramolecular direct C−H coupling of diphenylmethyl-substituted dichloroporphyrin 100.1, which unexpectedly provided stable radical species 100.2 (Scheme 100).270 The reaction proceeded smoothly in the presence of Pd(OAc)2 and PCy3·HBF4 as a catalyst, and K2CO3 as a base, in toluene. 3.1.7. Homocoupling. Kobuke and co-workers reported the synthesis of porphyrin dimer 101.2 bridged with 2,2′bipyridyl (bpy) via a Ni(0)-mediated homocoupling reaction (Scheme 101).271 The spatial geometry of the two porphyrin

52.1Zn with NaCH(CN)2 in the presence of Pd2dba3, CuI, and PPh3, followed by oxidation with oxygen, afforded 5,15dialkylideneporphyrin 96.1 in 53% yield. They also succeeded in the synthesis of cumulenic dimer 96.3 and quinoidal dimer 96.5 via a similar procedure.264 In these cases, the use of NIS as an oxidant gave good results. The group of Sugiura described the synthesis of tetracyano-5,10-porphyrinquinodimethane 96.6 by Takahashi coupling with Zn(II) 5,10-dibromoporphyrin 51.2Zn.265 The authors found that PdCl2(dppf) effectively afforded the corresponding coupling product in 47% yield. The palladium-catalyzed β-acetylation of a porphyrin was reported by Yorimitsu, Osuka, and co-workers (Scheme 97).266 The introduction of acetyl groups was achieved by treatment of Ni(II) β,β′-diborylporphyrin 22.1Ni with acetic anhydride in the presence of a catalytic amount of Pd(OAc)2, tri(4methoxyphenyl)phosphine, and 2 equiv of NaOAc in THF/ H2O, providing diacetylporphyrin 97.2 in 56% yield. Chen and co-workers reported the palladium-catalyzed intramolecular cyclization reaction of β-bromoporphyrin metal complexes (Scheme 98).267 In the presence of a catalytic AN

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 72. Synthesis of a meso-Ethynyl-Alkenyl-Substituted Porphyrin

Scheme 73. Alkynylation at the β-Positions of Porphyrins

under an oxygen atmosphere successfully furnished the β−β linked porphyrin dimer 103.1 in good yield. Similarly, borylated corrole 23.2 underwent a palladium-catalyzed oxidative dimerization reaction to provide singly linked corrole dimer 103.2 in good yield.273 Dimer 103.2 was further oxidized with DDQ to doubly linked corrole dimer 103.3. Interestingly, dimer 103.3 exhibited broad absorption bands reaching up to 1600 nm, as well as significant singlet biradical character. Matano and co-workers reported the palladium-catalyzed dimerization of β-bromodiazaporphyrin 69.1Ni to afford directly connected bisdiazaporphyrin 104.1 (Scheme 104).237 Interestingly, the two diazaporphyrin units in 104.1 adopted an

units in 101.2 can be regulated by the reversible coordination of the bpy moiety to metals. In fact, the addition of PdCl2(CH3CN)2 converted 101.2 to the cofacial porphyrin dimer 101.3. The addition of 4,4′-dimethyl-2,2′-bipyridyl to 101.3 resulted in demetalation of 101.3 to regenerate the initial dimer 101.2. Homocoupling of bromoporphyrin 37.1Ni mediated by the nickel complex also proceeded smoothly to yield the meso−meso-linked dimer 102.1 (Scheme 102).272 A β−β directly linked porphyrin dimer was prepared via palladium-catalyzed homocoupling of a β-borylated porphyrin (Scheme 103).140 Treatment of 2-boryl-5,15-diarylporphyrin 18.2H with Pd(OAc)2/DPPP as the catalyst in DMSO/toluene AO

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 74. Ethynylation of β-Borylporphyrin

Scheme 75. Synthesis of an Ethynyl-Bridged 5,15-Diazaporphyrin Dimer

Scheme 76. Sonogashira Coupling of meso-Bromosubporphyrin

almost coplanar conformation, enabling effective expansion of the π-conjugation. 3.1.8. Introduction of Heteroatoms. The introduction of heteroatoms has also been achieved through palladiumcatalyzed reactions.274 Suda and co-workers reported palladium-catalyzed meso-amination and amidation reactions.275 The authors found that a combination of Pd(OAc)2, rac-BINAP, NaOt-Bu, and 18-crown-6 in THF was effective for the coupling of meso-brominated Ni(II) porphyrin 12.1Ni with hexylamine, providing meso-aminoporphryin 105.1Ni in good yield (Scheme 105). Interestingly, the reaction was found to be significantly dependent on the central metal. The reaction with Zn(II) complex 12.1Zn resulted in a lower yield. The reaction with free-base porphyrin 12.1H afforded an inseparable complex mixture, in addition to the recovered starting material. These reaction conditions were found to be also applicable to various amines. In particular, the reaction with m-phenylenediamine proceeded smoothly without any protection, affording the corresponding diporphyrin in 96% yield. The authors then expanded this protocol to meso-amidation. In the case of the

coupling of 12.1Ni with benzamide, the use of Cs2CO3 as base instead of NaOt-Bu was effective to produce the meso-amidated porphyrin 105.2Ni in good yield. This procedure was used for the synthesis of meso-dinaphthylporphyrin reported by Gryko et al.276 These reaction conditions were further optimized by Zhang et al.,277 who selected DPEphos as ligand due to its high catalytic reactivity when combined with palladium and its low cost (Scheme 106). They also found that the use of Cs2CO3 as a base in THF at elevated temperatures afforded the best results for Buchwald−Hartwig amination of Zn(II) meso-bromoporphyrin 45.1Zn with various alkyl and aryl amines. However, secondary aliphatic amines and cyclic aliphatic amines resulted in no reaction under these conditions. Imines are also suitable substrates for the coupling reaction under similar conditions, affording diiminoporphyrins in 84−94% yield. These optimized reaction conditions have been widely used for the synthesis of meso-aminoporphyrins. For example, Coutsolelos et al. reported the synthesis of a porphyrin triad connected by an amino group at the meso-positions.278 Moore, Gust, and co-workers reported AP

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 77. Synthesis of Various meso-Substituted Porphyrins through Negishi and Migita−Kosugi−Stille cross-Coupling Reactions

Scheme 78. Introduction of Ethynyl Substituents through Negishi Cross-Coupling

meso-amidation of meso-bromoporphyrins (Scheme 106).282 The use of Xantphos as a ligand was effective for installation of a variety of amide functionalities. Yorimitsu, Osuka, and co-workers reported the coupling reactions of meso-bromoporphyrins with bulky and less reactive arylamines (Scheme 107).283 In this case, the Pd/DPEphos system was not effective. The authors disclosed that the use of

the synthesis of aminophenyl-linked porphyrin polymers through the coupling reaction of Zn(II) 5-bromo-15-(4aminophenyl)porphyrin.279 The groups of Pawlicki and Latos-Grażyński reported a fused meso-aminoporphyrin.280 This system was effective for introduction of a variety of alkylamino and arylamino substituents at the meso-position of subporphyrins.281 Zhang and co-workers also reported the AQ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 79. Introduction of Alkyl Substituents at Porphyrin Peripheries through Negishi and Kumada−Tamao−Corriu Coupling Reactions

coupling reactions of meso-bromoporphyrins with alkyl carbazates.285 They found that products 108.1 and 108.2 were formed via auto-oxidation of 108.3. Porphyrin carbazate 108.3 was selectively obtained in 84% yield by reducing the reaction time under oxygen-free conditions. Arnold and co-workers extended their study to coupling with hydrazine (Scheme 109).286 The reaction of Ni(II) mesobromoporphyrin 37.1Zn with hydrazine sulfate with Pd(OAc)2/rac-BINAP as catalyst and Cs2CO3 as base in dry THF afforded meso-aminoporphyrin 109.1, hydroxyporphyrin 109.2, and bis(porphyrinyl)amine 109.3. When 2 equiv of meso-bromoporphyrin 37.1Zn and 1 equiv of hydrazine sulfate were used for this reaction, meso-aminoporphyrin 109.1 was obtained in 51% yield. On the other hand, 2 equiv of hydrazine

Pd-PEPPSI-IPent as catalyst worked effectively for the coupling of meso-bromoporphyrin 34.1Ni with phenoxazine, furnishing the corresponding coupling product 107.1 in 93% yield. This catalytic system was also effective for other cyclic diarylamines such as phenothiazine and diarylamines. Arnold and co-workers reported the synthesis of azocarboxylates and diiminoporphodimethanes.284 An excess of tertbutylcarbazate reacted with 5-bromo-10,20-diphenylporphyrin Ni(II) 12.1Ni under the conditions reported by Suda, yielding three isolated products (Scheme 108). The products were assigned as azocarboxylate 108.1, diiminoporphodimethenes 108.2, and carbazate 108.3. Longer reaction times and stirring the reaction mixture in air provided the highest yield of 108.2. The authors further investigated the Pd-catalyzed crossAR

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 80. Cross-Coupling of a Porphyrin with 2-(Diphenylsilyl)phenylzinc

Scheme 81. Synthesis of a Cyclobutanedione-Substituted Porphyrin

cross-coupling of bromoporphyrins.288 The initial attempt to induce reaction between the meso-bromoporphyrin and (trimethylsilyl)diphenylphosphine resulted in a low yield due to the formation of the debrominated porphyrin as the major product. After several trials, the authors succeeded in suppressing the debromination using meso-η1-palladioporphyrin 43.1 (Scheme 110). Treatment of 43.1 with a stoichiometric amount of diphenylphosphonic acid in the presence of Cs2CO3

sulfate produced bis(porphyrinyl)amine 109.3 in 25% yield. An alternative route to bis(porphyrinyl)amine 109.3 was reported by Ruppert et al.287 The reaction of Ni(II) meso-aminoporphyrin with meso-iodoporphyrin, in the presence of Pd(OAc)2/rac-BINAP as catalyst and NaOt-Bu as base in dioxane, led to formation of the coupling product in 56% yield. For the construction of C−P bonds at the meso-position, Arnold and co-workers reported the palladium-catalyzed C−P AS

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 82. Synthesis of a 2,4-Pentadienal-Substituted Porphyrin

Scheme 83. Synthesis of a meso-to-meso (E)-Vinylene-Linked Porphyrin Dimer

Scheme 84. Alkynylation of a Dibromoporphyrin through a Migita−Kosugi−Stille Coupling Reaction

elemental sulfur. The product 112.2 was easily converted back to 17.1 with P(NMe2)3 in refluxing toluene. Palladium-catalyzed C−O bond formation at the mesopositions were reported by Zhang et al. (Scheme 113).289,290 The reaction of meso-bromoporphyrin Zn(II) 45.1Zn with alcohols and phenols in the presence of Pd2dba3, DPEphos, and Cs2CO3 in toluene afforded the corresponding meso-alkoxy and meso-aryloxy products 113.1 under mild reaction conditions. This protocol can be applied to a variety of alcohols including sterically hindered 2-isopropylphenol. Double etherification with alcohols was also achieved. The reaction of mesobromoporphyrin 12.1H with ethylene glycol or hydroxyphenylporphyrin 113.3 furnished porphyrin dimers 113.2 or 113.4.291

led to reductive elimination, affording the corresponding phosphine oxide 110.1 in 85% yield. Based on this result, the authors developed a catalytic system for C−P bond formation at the meso-position (Scheme 111). The C−P coupling reaction of bromoporphyrins 111.1 with diphenylphosphonic acid was achieved using Pd(dppe)2 as catalyst and Cs2CO3 as base. Imahori, Matano, and co-workers reported another palladium-catalyzed methodology for the C−P cross-coupling of meso-iodoporphyrins 41.4Zn with diphenylphosphane (Scheme 112).135 The reaction was conducted in the presence of Pd(OAc)2 and Et3N in an MeCN/THF mixture, providing the corresponding meso-phosphanylporphyrin 17.1 in excellent yield. The product was rapidly oxidized in air, converting to meso-phosphonylporphyrin 112.1. The product was isolated as stable meso-thiophophorylporphyrin 112.2 by treatment with AT

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 85. Synthesis of an Octaalkynylporphyrin

Scheme 86. Synthesis of a Ni(II) meso,β-Diethynylporphyrin

Scheme 87. Synthesis of a 4-Stannylphenylporphyrin

The group of Senge reported the synthesis of a sulfur-linked porphyrin dimer through palladium-catalyzed sulfanylation of a meso-bromoporphyrin (Scheme 116).295 Treatment of mesobromoporphyrin 37.1Zn with isooctyl 3-mercaptopropionate in the presence of Pd2dba3/Xantphos and DIPEA in toluene at 80 °C provided the corresponding thiolated porphyrin 116.1. The coupling of meso-bromoporphyrin 12.1H with triisopropylsilanethiol was conducted by de Lera, Marsal, Alvarez-Puebla, and co-workers (Scheme 117). The authors obtained the thiolated product 117.1 in 36% yield using Pd(OAc)2/PPh3 as catalyst and Cs2CO3 as base.296 The introduction of heteroatoms at the β-position was reported by Zhang et al. β-Bromotetraphenylporphyrin 47.4H was converted into β-amino-, β-amido-, β-alkoxy-, and βthioporphyrins (118.1, 118.2, 118.3, and 118.4) via palladiumcatalyzed reactions with chelating diphosphine ligands (Scheme 118).297 The use of rac-BINAP as a ligand was effective for

Introduction of a hydroxy substituent was reported by Arnold et al.286,292 meso-Hydroxyporphyrin 109.2 was obtained as a side product (as shown in Scheme 114); however, the yield of 109.2 was improved to 79% in the absence of hydrazine sulfate. The group of Osuka also employed this procedure to obtain a meso-hydroxyporphyrin with different meso-substituents.293 Palladium catalyzed C−S and C−Se bond formations were reported by Zhang et al. The authors developed a method for sulfanylation of meso-bromoporphyrin 45.1H with a variety of thiols, in the presence of Pd2dba3 as catalyst and Cs2CO3 as base in toluene (Scheme 115).294 A wide range of ligands, including biaryl-based electron-rich bulky phosphines and Nheterocyclic carbenes, provided the desired meso-arylsulfanylporphyrins 115.1 in good yields. meso-Selenoporphyrin 115.2 was also prepared similarly, using selenols as coupling partners. AU

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 88. Synthesis of Heteroaromatic-Substituted 5,15-Diazaporphyrins

Scheme 89. Mizoroki−Heck Reactions at the meso-Position

amination and sulfanylation, affording the corresponding products in good yields. Alternatively, Xantphos and DPEphos were found to be effective for amidation and etherification. These procedures were found to be effective for a wide range of substrates. The groups of Yorimitsu and Osuka reported the crosscoupling of Zn(II) diiodoporphyrin 22.4Zn with amines and thiols (Scheme 119).143 The palladium-PEPPSI-IPr catalyst worked effectively for the introduction of amine groups, producing the corresponding β,β′-diaminoporphyrins 119.1 in good yields. For sulfanylation, the use of 1,1′-bis(diisopropylphosphino)ferrocene (DiPPF), which inhibits catalyst poisoning due to its sterically demanding nature, was effective. This ligand also worked well for the cyanation of

Zn(II) diiodoporphyrin 22.4Zn with Zn(CN)2, furnishing the corresponding dicyanoporphyrin in 67% yield. The groups of Yorimitsu and Osuka also reported the introduction of phosphonyl and phosphanyl groups with the corresponding phosphine-containing reagents, affording the products 120.1 and 120.2 in good yields (Scheme 120).143 Diphosphanylporphyrin 120.1Ni underwent cyclometalation at the meso-position to yield PCP pincer complexes upon treatment with nickel, palladium, and platinum salts. Similar procedures for phosphonylation, sulfanylation, and phenoxylation can be applied to subporphyrins.298 3.1.9. Copper- and Nickel-Catalyzed Introduction of Heteroatoms. The palladium-catalyzed amination of sp2carbon centers has proven to be highly effective in the AV

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 90. Vinylation of a Mercurated Porphyrin

Scheme 91. Mizoroki−Heck Reactions under Phase-Transfer Conditions

functionalization of porphyrins. Because this methodology generally provides satisfactory results, it is generally the first choice for the introduction of amino functionality into porphyrins. However, palladium complexes and phosphine ligands are far more expensive. In this regard, Ullmann-type copper-catalyzed amination represents a potentially useful secondary choice.

Ruppert and co-workers successfully introduced imidazole groups at the meso-positions through Ullmann amination using CuI as the catalyst (Scheme 121).299,300 The use of Nphenylbenzohydrazide as ligand was quite effective for the synthesis of meso-bromoporphyrin 121.2Zn.301 The imidazole moiety of 121.2 was butylated to yield an imidazolium porphyrin, which then acted as a carbene ligand to Pd(II) after deprotonation. The procedure could be extended to the AW

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 92. Synthesis of opp-Dibenzoporphyrins through Mizoroki−Heck Reactions of Tetrabromoporphyrins

Scheme 93. Synthesis of Push−Pull-Type opp-Dibenzoporphyrins

Scheme 94. Palladium-Catalyzed Cyanation of a meso-Bromoporphyrin

AX

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 95. Cyanation of meso-Bromoporphyrin with Zn(CN)2

ester functions into porphyrins was achieved through rhodiumcatalyzed addition of β-borylporphyrin 69.5H to acrylate and 2,4-pentadienoate esters followed by β-hydride elimination. In sharp contrast, the reaction of meso-borylporphyrin 19.7H with hexyl acrylate afforded saturated esters via 1,4-conjugate addition. The different reactivity was explained by the different steric environment of the meso- and β-positions, on the basis of DFT calculations. Installation of unsaturated moieties induces a substantial red-shift and broadening of the absorption spectra due to effective conjugation at the β-positions. The introduced unsaturated ester groups can be hydrolyzed to the corresponding carboxylic acids, which serve as anchoring groups to the surface of nanocrystalline titanium oxide.316 A range of porphyrins with unsaturated acid functions at their β-positions were employed as sensitizers for dye-sensitized solar cells. The reaction of meso- and β-borylated porphyrins with vinyl ketones resulted in the formation of saturated products. meso−meso-linked diporphyrin 127.1 can be efficiently functionalized through sequential use of iridium-catalyzed C− H borylation and rhodium-catalyzed acrylate addition (Scheme 127).317 On the basis of this strategy, meso−meso-porphyrin dimers were functionalized with multiple unsaturated carboxylic acid groups. The photovoltaic performance of these porphyrin dimers in dye-sensitized solar cells (DSSCs) was also tested.318

introduction of carbazole and phenoxazine units to furnish meso-substituted porphyrins 122.1 and 122.2 (Scheme 122).302 Ullmann-type copper-catalyzed amination successfully furnished bis(porphyrinyl)amine 122.5 in good yield under the same reaction conditions.303 A pyrazine-fused porphyrin dimer 123.2 was synthesized by copper-mediated dimerization of β,β′-aminobromoporphyrin 123.1 (Scheme 123).304 Although the pyrazine-fused porphyrin dimer 123.2 has a highly congested structure due to four phenyl groups around the pyrazine moiety, the coupling reaction was nevertheless successful. Similar pyrazine-fused porphyrin dimers were also prepared through different synthetic routes.305−308 Matano and co-workers reported effective copper-catalyzed sulfanylation and phosphorylation of meso-bromoporphyrins (Scheme 124). Treatment of meso-bromoporphyrin 41.1Zn with thiophenol or octanethiol in the presence of CuI and neocuproine as ligand furnished Zn(II) sulfanylporphyrins 124.1 in good yields.309 Sulfanylporphyrins 124.1 were then oxidized to the corresponding Zn(II) sulfinylporphyrins, which formed cofacial dimeric assemblies in solution. To introduce phosphoryl groups at the meso-positions, a catalyst combination of CuI and N,N′-ethylenediamine was employed.310 Dibutoxyphosphoryl- and diphenylphosphorylporphyrins 124.2 were prepared in excellent yields. Nickel catalysis has not been extensively explored in porphyrin synthesis. One study of note is by Chen and coworkers, who developed a procedure for aryloxylation and amination of meso-bromoporphyrins (Scheme 125).311 The nature of the central metal of the porphyrin substrate was found to be important in this reaction, with Ni(II) bromoporphyrins affording satisfactory results. 3.1.10. Rhodium-Catalyzed Reactions. Rhodium complexes can catalyze 1,4-addition reaction of organoboron compounds to α,β-unsaturated ketones and esters.312−314 This reactivity was applied to the reaction of meso- and βborylated porphyrins with unsaturated ketones and esters (Scheme 126).315 Introduction of α,β- and α,β,γ,δ-unsaturated

3.2. Peripheral-Functionalization

3.2.1. Suzuki−Miyaura Coupling. Lindsey and co-workers used Suzuki−Miyaura conditions to couple bromoporphyrin 128.3 and (p-borylphenyl)porphyrin Zn(II) complex 128.2, the latter of which was prepared through Miyaura borylation of iodophenylporphyrin 128.1 (Scheme 128).319 Several porphyrin arrays with different linker groups (which change the interporphyrin distances) were systematically prepared for photophysical studies. Ali and van Lier synthesized porphyrin−phthalocyanine conjugate 129.3 through Suzuki−Miyaura coupling of borylated phthalocyanine 129.2 with iodophenyl porphyrins 129.1 (Scheme 129).213 The electronic characteristics of these conjugates were also investigated. AY

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 96. Takahashi Coupling with meso-Bromoporphyrins

Scheme 97. Palladium-Catalyzed Acetylation of a β-Borylporphyrin

porphyrin 131.1 can be coupled with fluorenylboronic acid under Suzuki−Miyaura coupling conditions.321 In one study, a 4-hydroxyphenyl group was introduced through Suzuki−Miyaura coupling with β-(4-bromophenyl)porphyrin 131.1 (Scheme 131).322 Porphyrin triflate 131.3 was also used as a coupling partner in Suzuki−Miyaura coupling

Tetrakis(4-borylphenyl)porphyrin 130.1 could be coupled with bromobenzothiadiazoles under palladium catalysis (Scheme 130).320 The products 130.2 exhibited large twophoton absorbance (TPA) cross sections and effective singlet oxygen sensitization at 800 nm. Tetra(4-bromophenyl)AZ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 98. Intramolecular Cyclization of Zn(II) 2-Bromo-5,10,15,20-Tetraphenylporphyrin

Scheme 99. Multiple Fusion Reactions of a Tetraphenylporphyrin

Scheme 100. Two-Fold Palladium-Catalyzed Intramolecular C−H Arylation

reactions. Chan and co-workers prepared porphyrin−quinone conjugate 131.4 through Suzuki−Miyaura coupling followed by

deprotection and oxidation of the resultant hydroquinone porphyrins.323,324 In a similar manner, polyazamacrocycle units BA

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 101. Ni(0)-Mediated Homocoupling

Scheme 102. Reductive Dimerization of a Bromoporphyrin

coupling with para- and ortho-diiodobenzenes in the presence of palladium chloride and triphenylphosphine in triethylamine, providing the corresponding porphyrin dimers 134.2 and 134.3 in good yields under copper-free conditions (Scheme 134). They also prepared unsymmetrical diporphyrins using a mesobromovinyloctaethylporphyrin Ni(II) complex as the coupling partner in a Sonogashira-type reaction. In one report, propargyl alcohol was coupled with iodophenylporphyrin 129.1 in good yield (Scheme 135).235 Furthermore, it was found that Sonogashira coupling involving porphyrins can be conducted in aqueous reaction media (Scheme 136). 330 The Sonogashira coupling of tri(methylpyridinium)bromophenylporphyrin 136.1 in water/ acetonitrile was achieved with the water-soluble phosphine ligand TPPTS. Interestingly, the addition of a Cu(I) cocatalyst was not needed in this case. Lindsey and co-workers extensively used Sonogashira coupling to construct several porphyrin arrays for photophysical studies.331 Ethynylene-bridged porphyrin trimer 137.3 was efficiently constructed by the Sonogashira coupling of freebase bis(iodophenyl)porphyrin 137.2 with meso-alkynylpor-

can be installed onto porphyrin scaffolds through Suzuki− Miyaura coupling (Scheme 132).325 Kimura and co-workers systematically prepared various dendritic porphyrins through Suzuki−Miyaura coupling between tetrakis(3,5-dibromophenyl)porphyrin 133.1 and polyphenylene boronic acid 133.2 (Scheme 133).326 The free-base porphyrins were converted to the corresponding iron porphyrin 133.3, which served as a catalyst for alkene epoxidation. The dendritic porphyrin catalyst exhibited intriguing space-dependent reactivity. In a similar manner, dendritic porphyrins with benzoquinone termini were constructed through Suzuki−Miyaura coupling between tetra(4bromophenyl)porphyrin and dendritic boronic acids with 2,5dimethoxyphenyl groups at the terminal positions, which were finally converted to benzoquinones.327 The transfer of electrons within the porphyrin dendrimers was also examined. 3.2.2. Sonogashira Coupling. Sonogashira coupling is also quite effective for peripheral functionalization of porphyrins. Arnold and co-workers pioneered the use of Sonogashira coupling in porphyrin synthesis.328,329 mesoEthynyloctaethylporphyrin Ni(II) complex 134.1 underwent BB

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 103. Palladium-Catalyzed Oxidative Dimerization of Borylporphyrin and Borylcorrole

Scheme 104. Palladium-Catalyzed Dimerization of β-Bromodiazaporphyrin

phyrin Zn(II) 137.1 (Scheme 137).332 To avoid the metalation of the free-base porphyrin unit by copper, the copper-free Sonogashira reaction was used. In this case, triphenylarsine was the ligand of choice. The free-base porphyrin−Au(III) porphyrin conjugate 138.4, linked by a phenylene ethynylene unit, was synthesized by Sonogashira coupling (Scheme 138).333 Limberg and Hecht constructed diporphyrin 139.1, connected to a 2,6-bis(1,2,3-triazo-4-yl)pyridine unit, through a Sonogashira coupling reaction (Scheme 139).334 Conforma-

tional switching of this compound was achieved either by changing the pH or by metal complexation. Sonogashira coupling was also useful for the postmodification of highly elaborate structures. Lindsey and co-workers functionalized triple-decker sandwich complex 140.1 via Sonogashira coupling without decomposition of the europium complex (Scheme 140).335 The thiol-functionalized tripledecker porphyrin 140.2 formed a self-assembled monolayer (SAM) on a gold surface, allowing the examination of its electrochemical properties. BC

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 105. Palladium-Catalyzed Amination and Amidation of meso-Bromoporphyrins

Scheme 106. Palladium-Catalyzed Amination and Amidation of meso-Bromoporphyrins

A series of snowflake-shaped dendritic porphyrins such as 141.1 were effectively constructed through Sonogashira coupling reactions (Scheme 141).336−340 Because of the anthraquinone moieties at the termini of the dendrimers, photoinduced electron transfer was observed and comparative

studies indicated that the dendritic structures enhanced the electron transfer process. C60-terminated porphyrin-centered dendrimers were also constructed. Lindsey and co-workers extensively utilized the Sonogashira coupling to access various types of porphyrin−perylene imide BD

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 107. Synthesis of a Phenoxazine-Substituted Porphyrin

Scheme 108. Coupling with tert-Butyl Carbazate

Scheme 109. Coupling with Hydrazine Sulfate

conjugates, such as 142.1 (Scheme 142).341−349 The copperfree variant of the Sonogashira protocol was used in these

reactions. The authors explored the photophysical properties of these conjugates as light-harvesting arrays. The energy and hole BE

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 110. Reductive Elimination to Form a Phosphonylporphyrin

Scheme 111. Palladium-Catalyzed Phosphonylation

Scheme 112. Palladium-Catalyzed meso-Phosphanylation

BF

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 113. Palladium-Catalyzed meso-Alkoxylation

Scheme 114. Synthesis of meso-Hydroxyporphyrin

Sonogashira coupling is also useful for preparing porphyrinbased polymers under standard coupling conditions (Schemes 143 and 144).350,351 The field effect hole mobility of the

transfer processes of these porphyrin−perylene imide dyads were thoroughly investigated. BG

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 115. meso-Chalcogenation of Porphyrins

Scheme 116. Sulfanylation with Isooctyl 3-Mercaptopropionate

Scheme 117. Synthesis of a meso-Triisopropylsilylthioporphyrin

3.2.3. Migita−Kosugi−Stille Coupling. Senge and co-

porphyrin-dithienothiophene polymer was determined to be −4

2.1 × 10

2

−1 −1

s . Furthermore, polymer solar cells were

workers reported the synthesis of triphenylene-linked dipor-

fabricated using the porphyrin-dithienothiophene polymer

cm V

phyrin 145.1 through Migita−Kosugi−Stille coupling (Scheme

144.1 along with [6,6]-phenyl-C61-butyric acid methyl ester

145).253 The photophysical properties of 145.1, including

(PCBM) (Scheme 144). A power conversion efficiency of 0.3%

absorption, emission, and fluorescence lifetimes, were inves-

was observed.

tigated. BH

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 118. Palladium-Catalyzed β-Amination, Amidation, Alkoxylation, and Sulfanylation

3.2.4. The Mizoroki−Heck Reaction. Bromophenylporphyrin 146.1H can undergo Mizoroki−Heck reaction with styrenes and acrylates to furnish various alkenylated products, such as 146.2, in good yields (Scheme 146).352 Although the electronic effect of the alkenyl groups on the absorption spectra was small, the introduction of alkenyl groups enhanced the fluorescence quantum yields of tetraphenylporphyrin. Zn(II) protoporphyrin-IX dimethyl ester 147.1 was functionalized via a Mizoroki−Heck reaction (Scheme 147).353 In this example, a dinuclear palladium complex, which was obtained from the reaction of dibenzylphenylphosphine with palladium acetate, was effective. Unfortunately, however, the coupling reaction afforded a mixture of four regioisomers resulting from

nonselective addition of the phenyl group to the vinyl substituents. The Mizoroki−Heck reaction between di(vinylpheny)porphyrin 148.1 and 1,4-diiodobenzene 148.2 in the presence of palladium acetate afforded porphyrin-based polymer 148.3 (Scheme 148).354−356 The molecular weight of the resulting polymer ranged from 8300 to 46 000 Da on the basis of GPC analysis using polystyrene standards. The porphyrin polymer exhibited high levels of photoconductivity, imparting photorefractive properties on the material. 3.2.5. Amination, Amidation, and Etherification. Palladium-catalyzed amination (and related reactions) are also powerful techniques for the peripheral functionalization of porphyrins. Zhang and co-workers reported that the BI

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 119. Palladium-Catalyzed Amination and Sulfanylation of Diiodoporphyrin

Aza-crown ether-appended porphyrin 154.1 was prepared through palladium-catalyzed amination of bromophenylporphyrin 153.1 (Scheme 154).362,363 Addition of K+ ions to a solution of 154.1 resulted in the formation of a dimeric complex, while a monomeric complex was obtained with Na+ ions. The coordination behavior of DABCO to the Zn(II) center of 154.1 was also examined. Tetra(4-bromophenyl)tetrabenzoporphyrin can be also coupled with primary and secondary amines as well as aza-crown ethers under similar reaction conditions.364 Porphyrin-based chemosensors 155.2 and 155.3 were prepared by the Ullmann coupling of bromophenylporphyrins 155.1 and 146.1Zn with dipyridylamine in the presence of Cu powder and potassium carbonate (Scheme 155). The reaction required rather high temperature. Zn(II) porphyrin 155.3 served as a selective “turn-off” type sensor for a Cu(II) ion, while its fluorescence was quenched with a range of other transition metal ions.365,366 Zn(II) porphyrin-cored dendrimers such as 156.2 with triarylamine dendrons were constructed through coppercatalyzed amination (Scheme 156).367 Because of the antenna effect of the arylamine dendrons, the fluorescence quantum yields of the dendrimers were significantly enhanced. The redox properties of the dendrimers were also investigated. Beletskaya and co-workers investigated amination of 3iodophenylporphyrin 157.1 with various diamines under copper catalysis (Scheme 157).368 They also employed Buchwald−Hartwig amination, and the palladium-catalyzed route was observed to be superior to the copper-catalyzed route. Ullmann coupling was used for the synthesis of an etherbridged diporphyrin (Scheme 158).369 meso-Iodophenylpor-

bromophenyl groups of porphyrin 149.1 could be effectively coupled with a variety of amines, including primary and secondary alkyl- and arylamines (Scheme 149).357 The authors also extensively optimized the coupling conditions with respect to the ligands, bases, and solvents used. Zhang and co-workers prepared porphyrins functionalized with a variety of chiral amides through palladium-catalyzed amidation of the 2,5-dibromophenyl groups at the mesopositions of porphyrin 150.1 (Scheme 150).358 The resultant chiral porphyrins 150.2 were employed as chiral ligands for the asymmetric catalytic cyclopropanation of alkenes with diazo compounds. Palladium-catalyzed amination was used for the synthesis of various porphyrin architectures. Thereby, Cavaleiro and coworkers constructed the amino-bridged porphyrin−phthalocyanine conjugate 151.3 (Scheme 151).359 The interactions between two chromophores in the excited state were investigated and energy transfer from the porphyrin to the phthalocyanine units was observed. Porphyrin-bound amino groups can be arylated through Buchwald−Hartwig amination reactions (Scheme 152).360 Heating the resultant N-arylated porphyrins in refluxing nitrobenzene led to formation of quinolino-fused porphyrin 152.3. Six-membered cyclic secondary amines such as piperidine and morpholine could be introduced through palladiumcatalyzed amination of bromophenylporphyrin 153.1 (Scheme 153).361 Beletskaya and co-workers extensively optimized the conditions of this reaction and found that the use of DavePhos (2-dicyclohexylphosphino-2′-(N,N-dimethylamino)biphenyl) was optimal. BJ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 120. Palladium-Catalyzed β-Phosphanylation

Scheme 121. Copper-Catalyzed Introduction of Imidazole at the meso-Position

phyrin 158.1 could be coupled with meso-hydroxyphenylporphyrin 158.2 by using CuI as the catalyst.

butadiyne-bridged porphyrin oligomers, such as diporphyrin 159.1, as shown in a seminal report by Arnold and Johnson (Scheme 159).370−374 Anderson also synthesized butadiynebridged porphyrin dimer 159.3 and oligomers via Glaser−Hay coupling and investigated their photophysical properties and self-assembling behavior.375−378 Therien and co-workers then employed meso-arylporphyrins to prepare butadiyne-bridged porphyrin dyads such as 159.5 through Eglinton coupling.241,379 Lindsey and co-workers reported that a simple

4. REACTIONS AT ALKENES AND ALKYNES BOUND TO PORPHYRINS 4.1. Homocoupling of Terminal Alkynes

Homocoupling reactions of terminal alkynes bound to porphyrin cores are useful for the efficient synthesis of BK

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 122. Copper-Catalyzed Amination at the meso-Position

Scheme 123. Copper-Mediated Dimerization of an Aminobromoporphyrin

CuI promoted the selective trimerization of 65.1Zn to furnish cyclic trimer 161.2.214 A similar method can be applied to the effective synthesis of a meso-to-meso, β-to-β doubly butadiynebridged diporphyrin.212 The addition of polypyridyl templates can promote cyclic porphyrin oligomers selectively. Sanders and co-workers succeeded in the ligand-templated synthesis of cyclic porphyrin oligomers (Scheme 162).385−388 When 4,4′-bipyridyl was added as a template to the reaction mixture, the cyclic porphyrin dimer 162.1 was obtained selectively. On the other hand, the addition of tri(4-pyridyl)triazine led predominantly to trimerization to the cyclic trimer 162.2.

homocoupling under Sonogashira coupling conditions with iodine as an oxidant afforded linear butadiyne-bridged porphyrin dimer 159.6.380 Sugiura and co-workers extensively explored the homocoupling-based synthesis of porphyrin dimers and oligomers bridged by alkyne spacers at the meso-positions.381−384 Copper-mediated homocoupling was found to be an effective methodology for the synthesis of cyclic tetramer 160.1 and dodecamer 160.2 (Scheme 160). The structure of 160.2 was visualized by STM imaging. Homocoupling of β,β′-diethynylporphyrin 65.1Zn mediated by a copper salt afforded doubly bridged dimer 161.1 (Scheme 161).142 Interestingly, the combination of PdCl2(PPh3)2 with BL

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 124. Copper-Catalyzed Sulfanylation and Phosphorylation at the meso-Position

Scheme 125. Nickel-Catalyzed Aryloxylation and Amination at the meso-Position

the removal of the template, the 12-porphyrin nanoring 164.2 was formed efficiently. This methodology is applicable to the synthesis of a variety of porphyrin nanorings.392−398

Anderson and co-workers actively explored the synthesis and properties of porphyrin nanorings through the templatedirected homocoupling of terminal alkynes. The use of hexadentate template 163.1 with a porphyrin dimer precursor furnished the corresponding porphyrin[6] nanoring 163.2 with high selectivity (Scheme 163).389,390 Interestingly, when the reaction of the porphyrin tetramer with 163.1 was carried out, the figure-of-eight 12-porphyrin complex 164.1 was obtained through the Vernier templating strategy (Scheme 164).391 After

4.2. Alkene, Alkyne, and Enyne Metatheses

Alkene, alkyne, and enyne metatheses are powerful methods for the introduction of alkenyl and alkynyl groups on porphyrins through the scission and regeneration of carbon−carbon double and triple bonds. Alkene metathesis enables the efficient introduction of versatile functional groups to porphyrin BM

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 126. Rhodium-Catalyzed Reaction of Borylporphyrins with α,β-Unsaturated Esters

Scheme 127. Sequential Borylation and Introduction of Unsaturated Esters

compounds using alkenyl linkers.399,400 Dolphin and coworkers reported the first application of alkene cross metathesis reactions of vinyl porphyrins.401,402 This report clearly showed that porphyrins having alkenyl moieties can be connected with functional groups and other molecules through cross metathesis. Grin and co-workers applied alkene cross metathesis for the synthesis of chlorophyll a glycoconjugate 165.1 (Scheme 165).403 Dimeric porphyrin-RGD peptide conjugates can be also prepared through cross metathesis.404 These porphyrin conjugates were found to be promising candidates in PDT. Combination of cross metathesis with the Prato reaction (1,3-dipolar cycloaddition)405 successfully afforded C60-por-

phyrin dyad 166.1 (Scheme 166).406 Enyne metathesis is effective for the construction of a diene skeleton. Sequential enyne metathesis and [4 + 2] cycloaddition with C60 efficiently provided C60-porphyrin dyad 167.1 (Scheme 167).407−409 Ring-closing metathesis is useful for the π-extension and the construction of fused rings around porphyrin cores. Smith and co-workers reported a unique route to benzoporphyrins on the basis of ring-closing metathesis (Scheme 168).410 β,β′Diallylporphyrin 168.1 is a readily available precursor to benzoporphyrins and is prepared from regioselective bromination of tetraphenylporphyrin and Suzuki−Miyaura coupling. Both ring-closing metathesis and oxidation provided successful BN

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 128. Synthesis of a Diporphyrin Bridged by a Terphenyl Group

that the porphyrin trimer 172.2 was predominantly formed with C60 as a template (Scheme 172).427 Templates having pyridyl moieties preorganize the monomers into the assembled precursors through coordination to the metal centers of the porphyrins. Subsequent metathesis afforded the macrocycles efficiently.428 A porphyrin containing four pyridyl groups (173.1) served as a template for the synthesis of tetrameric porphyrin box 173.2 (Scheme 173).429 DABCO acted as a template, coordinating to metal centers in porphyrins at both sides, and subsequent metathesis afforded the macrocyclic porphyrin dimers.430,431 Zimmerman and co-workers succeeded in the construction of synthetic hosts via metathesis of dendritic porphyrin 174.1 (Scheme 174).432 After hydrolysis of the ester groups, the molecular imprinting polymer 174.2 containing one binding site was obtained, which functioned as a shape-selective host. Zimmerman and co-workers also explored the synthesis of similar nanosized cored polymers.433−435 Alkyne metathesis is effective in constructing linear and rigid cross-linking units between porphyrin moieties. Polymerization of alkynylated porphyrins afforded ethynylene-bridged porphyrin polymers through alkyne metathesis.436 The alkynyl linkers were also found to be useful motifs for the synthesis of porphyrin boxes.437 Rectangular prism 175.1 was constructed via alkyne metathesis in one step, and the molecule was observed to capture C70 selectively (Scheme 175).438

access to mono-, di-, and tribenzoporphyrins (e.g., 168.2). Ring-closing metathesis of porphyrin 169.1 bearing alkenyl moieties with long alkyl tethers enabled the formation of porphyrin 169.2 with macrocyclic skeletons (Scheme 169).411,412 This strategy in macrocyclization can be applied to the formation of [2] and [3]catenanes with porphyrin cores.413,414 Alkene metathesis is a reliable strategy for dimerization, oligomerization, and cyclooligomerization. Linear dimer, trimer, and star-shaped pentamers (of which 170.1 is an example) can be obtained through alkene metathesis (Scheme 170).415,416 Kobuke and Satake et al. systematically studied a covalent-linking strategy for the construction of macrocycles from coordination-organized porphyrin aggregates.417−421 A porphyrin-based supramolecular macrocycle was constructed by the use of complementary coordination of imidazolyl Zn(II) porphyrin 171.1, which was transformed to the corresponding macrocycle 171.2 with covalent-linking through metathesis (Scheme 171).422 The obtained macrocycles can be recognized as mimics of the photosynthetic light-harvesting system. A similar strategy was applied to the synthesis of the covalently bound nanostructure from two porphyrin wires leading to molecular tubes.423 Template-directed metathesis is a powerful method for the synthesis of macrocyclic porphyrin oligomers.424−426 Langford and co-workers reported the synthesis of the covalently linked porphyrin dimer 172.1 through alkene metathesis. They found BO

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 129. Synthesis of a Porphyrin−Phthalocyanine Conjugate

Scheme 130. Introduction of Benzothiadiazoles via Suzuki−Miyaura Coupling

4.3. Cyclotrimerization

structural motifs containing cofacial porphyrins were prepared. A similar methodology through palladium-catalyzed cyclotrimerization was also reported.440

The cobalt-mediated [2 + 2+2] cycloaddition of alkynes is an effective methodology for the construction of benzene cores. Fletcher and Therien applied this methodology to the preparation of porphyrins having alkynyl groups at the mesoand β-positions. The authors succeeded in the construction of dimeric and trimeric porphyrins 176.1, 176.2, and 176.3 bridged with phenylene linkers (Scheme 176).439 A variety of

4.4. Copper-Catalyzed Huisgen Cycloaddition

1,3-Dipolar cycloaddition has been established as an efficient way to combine two molecules into one. Huisgen cycloaddition between terminal alkynes and azides, which is one of the socalled “click” reactions, is often accelerated by the addition of BP

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 131. Introduction of a Hydroxyphenyl Group through Suzuki−Miyaura Coupling

Scheme 132. Introduction of Polyazamacrocycles through Suzuki−Miyaura Coupling

Cu(I) catalysts. This cycloaddition is sometimes referred to as the copper-catalyzed alkyne−azide cycloaddition (CuAAC). The CuAAC reaction has been used extensively to fabricate highly elaborated porphyrin architectures such as photosynthetic models, artificial enzymatic catalysts, solar cell organic dyes, and photodynamic therapeutic agents. The use of click chemistry in porphyrin synthesis has been reviewed.441,442 Copper(I) salts such as CuBr, CuI, and Cu(NCCH3)4PF6 are often employed as the catalyst in this reaction. As an alternative, an in situ-generated Cu(I) species prepared via reduction of CuSO4 with sodium ascorbate or ascorbic acid can also act as a catalyst. In the case of Cu(I)-mediated cycloaddition of porphyrin substrates, copper-insertion into the cavity of freebase porphyrins was in some cases problematic. To avoid Cumetalation of free-base porphyrins, ligands such as tris[(1-

benzyl-1H-1,2,3-triazol-4-yl)methyl]amine (TBTA) are often employed. Otherwise, metalated porphyrins, such as Zn(II) porphyrins, should be employed as substrates to minimize the potential copper metalation. The Zn ion can be used as a temporary protecting group for the central cavity of porphyrins because of its easy demetalation by acid. Ruthenium is also effective for this transformation. To overcome the issue of copper−metalation of free-base porphyrins, ruthenium-catalyzed reactions can be used (Scheme 177).443,444 In this case, no insertion of a ruthenium ion to the central cavity of 177.2 was observed. 4.4.1. Click Reactions for the Synthesis of Porphyrin Oligomers and Multichromic Systems. Porphyrin oligomers and porphyrin arrays have been targets of active research based on their photophysical and electrochemical BQ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 133. Synthesis of a Dendritic Porphyrin through Suzuki−Miyaura Coupling

these dyads, a Zn(II) porphyrin and expanded porphyrins served as electron donors and acceptors, respectively. Gryko et al. developed a protocol for the synthesis of azido and propargyl-functionalized corroles, which were employed for the fabrication of a heterobimetallic dicorrole through a click reaction.450 A number of porphyrin derivatives have been tested as organic sensitizers for dye-sensitized solar cells (DSSCs). In particular, porphyrins with a donor−π−acceptor (D−π−A) structure have attracted much attention, but the synthesis of such elaborate porphyrins requires lengthy multistep synthetic efforts. Click chemistry is a powerful tool to access such porphyrin architectures. In this regard, covalently linked porphyrin dimer functionalized with a carboxylic acid moiety was prepared through the CuAAC reaction.451 Barbe, Fukuzumi, and co-workers efficiently synthesized triporphyrin 180.2 tethered by a tris(triazoylmethyl)amine linker via a click reaction with azido porphyrin 180.1 (Scheme 180).452 The electrochemical properties of trimer 180.2 were explored. Interestingly, triporphyrin 180.2 led to a host−guest complex upon the addition of a pyridine-appended C60 derivative. Photoinduced electron transfer in the supramolecule was investigated. A similar triporphyrin could be obtained through a CuAAC reaction of p-azidophenylporphyrin with

properties, particularly to understand the energy transfer and electron transfer processes in photosynthesis. The construction of covalently connected oligoporphyrins generally requires sophisticated synthetic techniques. However, the CuAAC reaction may greatly simplify preparative routes to porphyrin oligomers. Chen developed synthetic routes to various β-azidotetraarylporphyrins, such as 178.1, from β-nitrotetraarylporphyrins.445 Azidoporphyrin 178.1 was converted to porphyrin dimer 178.3 through copper-catalyzed Huisgen cycloaddition with mesoethynylporphyrin 178.2 (Scheme 178). Odobel and co-workers also established a route to mesoazidoporphyrins and meso-azidophenyl porphyrins. A click reaction of azidoporphyrin 179.1 with ethynylporphyrin 179.2 furnished diporphyrin 179.3 linked by a 1,2,3-triazole linker at the meso-positions (Scheme 179).446 The reaction conditions (including copper catalysts) were optimized, and copper bromide ligated with an N-heterocyclic carbene ligand was found to provide good results in this case. Ravikanth and co-workers applied click chemistry to link core-modified porphyrin units and porphyrins, thereby constructing dyad systems.447,448 They also connected a Zn(II) porphyrin unit with core-modified expanded porphyrins such as thiarubyrin and oxasmaragdyrin through click reactions.449 In BR

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 134. Synthesis of Porphyrin Dimers Bridged by Diethynylphenylene Linkers

Scheme 135. Introduction of a Propargyl Alcohol Group through Sonogashira Coupling

Scheme 136. Sonogashira Coupling in Aqueous Reaction Media

1,3,5-triethynylbenzene.453 Beletskaya and co-workers expedi-

Harvey and co-workers constructed model compounds for Photosystems I and II by introducing two dendritic porphyrin units (181.2) to an artificial special pair model, which is a faceto-face porphyrin dimer (185.1; Scheme 181).455 The energy transfer in the dendrimers 181.3 was examined by transient

tiously employed the CuAAC reaction of di(ethynylphenyl)porphyrin and azidophenylporphyrin to fabricate a porphyrin trimer.454 BS

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 137. Synthesis of a Ethynylene-Bridged Porphyrin Trimer

Scheme 138. Copper-Free Sonogashira Coupling

absorption spectra. Click chemistry allowed the efficient synthesis of several different model compounds with various dendrons starting from the same central diporphyrin motif (181.1). Various Zn(II) porphyrin-cored dendrimers with polycarbazole dendrons were also systematically prepared through CuAAC chemistry.456 A dendritic multiporphyrin cluster 182.5 with a logical energetic gradient was constructed by the expeditious use of sequential click reactions (Scheme 182).457 A CuAAC reaction of 182.1 with 182.2 afforded tetraporphyrin 182.3 after removal of the trimethylsilyl group. Then 182.3 was further reacted with tetraazido porphyrin Au(III) complex 182.4 to

furnish dendritic multiporphyrin 182.5. The effective use of the trimethylsilyl group as a protection group for the terminal alkyne in 182.2 enables this highly efficient synthesis. The photophysical abilities of this gigantic molecule as a light harvester were investigated. A porphyrin dimer cage 183.3 was prepared in reasonably high yield through a CuAAC reaction between tetrapropargylporphyrin 183.1 and tetraazidoporphyrin 183.2 under high dilution conditions (Scheme 183).458 This double-decker porphyrin dimer 183.3 recognizes an azide ion though its binding inside the cage. BT

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 139. Synthesis of Porphyrin Dimers Connected via a 2,6-Bis(1,2,3-triazo-4-yl)pyridine Group

Scheme 140. Post-Modification of Triple-Decker Sandwich Complexes

changes in absorption and fluorescence quenching. The bisindole unit of 184.3, along with two 1,2,3-triazole rings, worked cooperatively as a ligand selective for Cu(II) ion to change the direction of energy flow in 184.3.

Two Zn(II) porphyrin units were introduced on diethynylbisindole linker 184.1 through a CuAAC reaction in good yield to furnish diporphyrin 184.3 (Scheme 184).459 The addition of Cu(II) ions to the solution of 184.3 resulted in BU

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 141. Synthesis of a Snowflake-Shaped Dendritic Porphyrin

on the electron transfer from the porphyrin unit to the POM was also investigated. Multiporphyrin-appended polymers were synthesized efficiently through click chemistry between alkynylporphyrins and azide-containing polymers, the latter of which were obtained

Two porphyrin units were covalently attached at the adjacent positions of a Dawson-type polyoxometalate (POM) through efficient CuAAC reaction between azidoporphyrin 207.1 and 207.2 (Scheme 185).460 The effect of flexible and rigid linkers BV

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 142. Synthesis of Porphyrin−Perylene Imide Conjugates through Sonogashira Coupling

Scheme 143. Polymerizarion of an Ethynylporphyrin through Sonogashira Coupling

Scheme 144. Synthesis of a Porphyrin-Dithienothiophene Polymer

through ring-opening metathesis polymerization (ROMP).461 CuAAC reaction of an alkynylated Ga(III) porphyrin with a hydrophilic polymeric diazide provided fluorescence hydrogel nanoparticles in the presence of a nonionic surfactant.462 A click reaction was effectively used to cap both ends of pseudorotaxane 186.1 with bulky 1,2,3-triazole capping groups 186.2, to build up [3]rotaxane 186.3 (Scheme 186).463 After removing two Cu(I) template cations, the two porphyrin units in 186.3 can move freely along the thread and bind guest molecules cooperatively.

Click chemistry was also used to construct corrole−BODIPY dyad 187.3 and core-modified porphyrin−BODIPY dyad 187.6 (Scheme 187). The CuAAC reaction of Ga(III) p-azidophenyl corrole 187.1 with alkynyl BODIPY 187.2 afforded the corrole−BODIPY conjugate 187.3 efficiently, the energy transfer properties of which were investigated.464 p-Ethynylphenyl dithaiporphyrin 187.5 underwent efficient Huisgen cycloaddition with α-azido BODIPY 187.4 to yield 187.6.465 4.4.2. Construction of Porphyrin−Nanocarbon Conjugates through Click Chemistry. Porphyrin−fullerene conjugates are often target molecules in the field of BW

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 145. Synthesis of Triphenylene-Linked Diporphyrin 145.1

Scheme 146. Installation of Slkene Moieties through a Mizoroki−Heck Reaction

Scheme 147. Introduction of Aryl Groups via Mizoroki−Heck Reaction

photoinduced electron transfer. Owing to its highly electronaccepting nature and uniquely small reorganization energy, C60 is one of the most effective acceptors, often achieving long excited life times. The spatial relationship between the porphyrin and fullerene moieties has a significant impact on

the rate of electron transfer and reverse electron transfer. By using a diverse range of porphyrin−fullerene conjugates, chemists have investigated the mechanism of the photoinduced electron transfer. In the synthesis of these conjugates, “click BX

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 148. Polymerization of a Di(vinylphenyl)porphyrin through the Mizoroki−Heck Reaction

Scheme 149. Palladium-Catalyzed Amination of Bromophenylporphyrin

Scheme 150. Palladium-Catalyzed Amidation of a Dibromophenylporphyrin

chemistry” has become a powerful tool to combine porphyrin and fullerene moieties. Fazio and Schuster employed the copper-catalyzed click reaction to fabricate porphyrin−fullerene conjugate 188.3 (Scheme 188).466 A click reaction between 185.1 and pethynylbenzaldehyde provided formylporphyrin 188.1. Formylporphyrin 188.1 was then coupled with C60 in the presence of sarcosine to afford 188.3 through 1,3-dipolar cycloaddition. The strong fluorescence quenching of zinc porphyrin units indicated effective electronic interaction between them in the excited state.467 Nierengarten and co-workers used the click reaction to prepare a Th-symmetric C60 derivative functionalized with 12 porphyrin units.468 This strategy is based on the selective 6-fold Bingel−Hirsch addition reaction of C60 to afford hexakis adducts.469−474 The C60 bisadduct and trisadduct were also used to functionalize Zn(II) porphyrins with C60 units.475 Hirsch developed a synthetic procedure for selective preparation of a C3-symmetric trisadduct of C60, which was also applied to furnish a functionalized fullerene with three porphyrin rings.476

The click reaction is also effective to connect several different chromophores into one molecule. Ng, Fukuzumi, and coworkers successfully synthesized BODIPY−porphyrin−C60 pentad 189.3 (Scheme 189).477 The kinetics of the electron transfer process and the lifetime of the charge-separated state were determined by transient absorption spectroscopy. In a similar manner, a porphyrin appended with four BODIPY units was constructed through a CuAAC reaction.478 A dendritic porphyrin−C60 conjugate 190.3 containing two porphyrin units on one fullerene was efficiently prepared via click reaction of ethynylated C60 190.1 (Scheme 190).479 The precursor diporphyrin unit 190.2 containing the azide function was also prepared by a CuAAC reaction. This adduct 190.3 forms a supramolecular polymer by encapsulation of the C60 unit into the jaws of a dimeric porphyrin moiety of another molecule. The CuAAC reaction usually occurs smoothly under very mild conditions at room temperature, thus the reaction usually does not interfere with noncovalent bonds such as coordination bonds. The click reaction enabled efficient construction of rotaxane- and catenane-type porphyrin−fullerene conjugates. BY

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 151. Synthesis of an Amino-Bridged Porphyrin−Phthalocyanine Conjugate

Scheme 152. Synthesis of a Quinolino-Fused Porphyrin

Scheme 153. Palladium-Catalyzed Amination with Secondary Amines

191.2 from copper complex 191.1 (Scheme 191). 480 Sulfonated bathophenanthroline (SBP) was effective in

Megiatto and Schuster applied their own click methodology to fabricate catenanes for the synthesis of porphyrin−C60 rotaxane BZ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 154. Synthesis of an aza-Crown Ether-Appended Porphyrin

Scheme 155. Introduction of Dipyridylamino Groups through Ullmann Coupling

were investigated and the lifetimes of the charge-separated states were determined. A SWNT functionalized with a Zn(II) porphyrin−Zn(II) phthalocyanine dendron was also obtained through a click reaction.486 Graft-polymer-like carbon nanotubes functionalized with porphyrin polymers were also prepared.487 A similar click strategy was also employed to functionalize carbon nano-onions, which are multishell fullerenes with graphitic multilayer structures.488 4.4.3. Self-Assembly of Triazolylporphyrins. The click reaction allows easy installation of triazole moieties at the peripheral positions of porphyrins from alkynyl or azide precursors. The triazole units function as coordinating groups to the central metal of the porphyrin to provide a selfaggregated architecture. Osuka and co-workers prepared Zn(II) triazolylporphyrin 193.1, which assembled into dimeric complex 193.2 by the coordination of a nitrogen atom to the Zn(II) center (Scheme 193).489 This strategy was further extended to the construction of a self-assembled porphyrin

improving the reaction rate and preventing dissociation of the precursor. Porphyrin−C60 dyad 191.2 exhibited strong electronic communication between porphyrin and C60 units. The same strategy was also found to be suitable for the synthesis of a porphyrin−C60 catenane.481−483 The combination of carbon nanotubes (CNT) with porphyrins is also an active goal for researchers. CNTs are expected to be next-generation materials owing to their spectacular electronic, mechanical, and optical properties. Click chemistry turned out to be quite effective to functionalize carbon nanotubes with various organic molecules. Single-walled nanotubes (SWNTs) can be decorated with Zn(II) porphyrin moieties through CuAAC reactions, such as that between ethynylphenylated SWNTs 192.1 and Zn(II) azidoporphyrin 192.2, which led to electron donor−acceptor conjugates 192.3 (Scheme 192).484 Tris(hydroxypropyltriazolylmethyl)amine (THPTA) was used as a ligand for Cu(I).485 The photophysics of the resulting SWNT−Zn(II) porphyrin conjugates 192.3 CA

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 156. Synthesis of a Zn(II) Porphyrin-Cored Dendrimer with Triarylamine Dendrons

Scheme 157. Copper-Catalyzed Amination of a (Halophenyl)porphyrin with a Diamine

Scheme 158. Synthesis of an Ether-Bridged Diporphyrin

box.195 A similar dimeric assembly was observed in a Zn(II) porphyrin−cyclic paraquat conjugate.490 Furuta and co-workers prepared Zn(II) triazolylporphyrin 194.2 in good yield through a CuAAC reaction of (ethynylphenyl)porphyrin 194.1 (Scheme 194).491 Interestingly, 189.2 afforded the self-assembled dimer 194.3 both in crystals and in solution. The two porphyrin units were bound together with two molecules of methanol, the oxygen atoms of

which coordinated to the zinc center, and the hydroxy groups formed hydrogen bonds to the triazole units. A click reaction of ethynylated chlorophyll 195.1 with benzyl azide efficiently provided self-assembled triazoylchlorin dimer 195.2 (Scheme 195).492,493 A triazole-linked chlorophyll dyad was also obtained from CuAAC reaction of the ethynylated chlorophyll with azidomethylchlorin. The coordination of triazole units was also confirmed in polymeric materials. Crossley and co-workers prepared porphyrin−polymer conCB

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 159. Homocoupling of Ethynylporphyrins

cells. In a similar manner, the CuAAC reaction efficiently provided a porphyrin−β-carboline conjugate, which also showed photocytotoxicity.501 A porphyrin−quinolone conjugate was prepared similarly.502 A porphyrin-tethered calix[6]arene was synthesized through a click reaction of triazido calix[6]arene with (ethynylphenyl)porphyrin.503 Four poly(ethylene oxide) chains were introduced to tetra(4propargyloxyphenyl) Zn(II) porphyrin 199.1 by CuAAC reaction (Scheme 199).504 The terminal hydroxyl group of 199.2 was then connected to Zn(II) 4-hydroxyphenyltri(4propargyloxyphenyl)porphyrin. A further click reaction with azido-terminated poly(ethylene oxide) furnished a watersoluble porphyrin-cored poly(ethylene oxide) dendrimer. The resulting dendrimer acts as a photosensitizer to generate singlet oxygen in an aqueous medium. Polyethylene and aminoalkyl chains were introduced to porphyrins similarly.505,506 The poly(ethylene glycol)-supported immobilized iron(II) porphyrin catalyst 200.3 was prepared through a click reaction of azido porphyrin 200.2 with propargyl-terminated poly(ethylene glycol) (PEG) 200.1 (Scheme 200).507 The PEG unit serves as a soluble polymer support, which enables both high catalytic performance under homogeneous reaction conditions and the efficient recovery of the catalysts by precipitation and filtration. In this case, the PEG-supported iron

jugates using CuAAC reactions, in which two or four linear polymer arms were attached to a Zn(II) porphyrin core with 1,2,3-triazole units.494 The conjugate exhibited self-assembly behavior through coordination of a triazole nitrogen to the zinc center, resulting in an increase of the glass transition temperatures of the bulk materials. 4.4.4. Click Reaction for Porphyrin Functionalization. A “capped” Zn(II) porphyrin 196.3 was prepared through a 4fold CuAAC reaction of tetraazido porphyrin 196.2 with tetrapropargyl ester 196.1 (Scheme 196).495 Various “capped” porphyrins were constructed by Baldwin et al.,496,497 but the use of click chemistry enabled a more efficient route to this type of elaborated porphyrin. To create cytochrome c oxidase model 197.3, a click reaction was employed to install a distal hydrogen-bonding pocket on a porphyrin ligand (Scheme 197).498 The zinc center, which prevented Cu metalation of the porphyrin cavity, was then replaced with iron. The corresponding iron porphyrin complex served as an O2 reduction electrocatalyst.499 Click reaction of propargyloxyporphyrin 198.1 was employed to prepare porphyrin−psoralen conjugate 198.3 (Scheme 198).500 The binding of the water-soluble porphyrin conjugate 198.3 to DNA by intercalation was investigated. The conjugate 198.3 exhibited high photocytotoxicity toward cancer CC

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 160. Porphyrin Squares Prepared by Homocoupling

ciently.511 The time-dependent uptake and subcellular distribution of the carbohydrate conjugates was investigated in human cancer cells. Scanlan et al. reported a microwave-mediated click reaction for the efficient synthesis of glycoporphyrins.512 In this procedure, the use of porphyrin diazide 203.1 allowed sequential “double-click” reactions with propargylated sugars 203.2 and 203.3 to fabricate diglycoporphyrin 203.4 with two different sugar moieties (Scheme 203). The reaction also allowed the use of unprotected propargyl glucosides. A library of glycoporphyrins were prepared and evaluated for their PDT activity against cancer cells in vitro.513 The effect of the sugar moiety on the PDT activity, cell uptake, and subcellular localization of the glycoporphyrins was investigated. It was also reported that a water-soluble porphyrin containing pyridinium groups underwent a CuAAC reaction with unprotected propargylated sugars under more mild reaction conditions at room temperature.514 Maillard and co-workers prepared several glycoporphyrin derivatives such as 204.3 through microwave-assisted click reactions of azidoporphyrin 204.1 with propargylated sugars such as 204.2 (Scheme 204).515 These glycoporphyrin derivatives were found to exhibit amphiphilic character. The cytotoxicity and photocytotoxicity of these Zn(II) porphyrin− glycoside conjugates in vitro were investigated. Snyder and coworkers also optimized the reaction conditions for the CuAAC reaction of Zn(II) 5,15-di(4-ethynylphenyl)porphyrin with various acetylated glycosyl azides.516 An antibacterial material can be prepared by grafting Zn(II) porphyrin units onto cotton fabrics through cellulose azidation and a CuAAC reaction with propargyloxyporphyrin 205.2 (Scheme 205).517 The porphyrin-appended polymer was found to kill Gram-positive and Gram-negative bacteria. Cellulose

porphyrin 200.3 effectively catalyzed the olefination of various aldehydes with ethyl diazoacetate in the presence of triphenylphosphine. Porphyrin-cored dendrimers 201.4, bearing polyglycerol dendrons, were efficiently prepared through CuAAC reactions between octapropargylporphyrins 201.1 and dendritic azide 201.2 (Scheme 201).508 Click chemistry enabled construction of a high molecular weight material (ca. 16000 Da) in a single step. In a similar manner, porphyrin dendrimers 201.5 with triazole-linked dendrons were also synthesized.509 The effect of the structure of the dendrons on the efficiency of the photoinduced electron transfer from the dendritic Zn(II) porphyrin 184.5 to electron acceptors was investigated. 4.4.5. Porphyrins Decorated with Sugars, Peptides, and DNA. The introduction of sugar moieties is an effective strategy to create hydrophilic porphyrins, which are important for bioimaging and medical applications of porphyrin derivatives such as photodynamic therapy (PDT). Careful selection of carbohydrate residues allows selective binding of the molecules to specific proteins. However, installation of such hydrophilic carbohydrate substituents into the porphyrin precursors would presumably result in low yields of the desired products because of side reactions caused by protecting groups. The highly polar nature of the compounds is also problematic in separating the products from the unwanted byproducts. In this regard, late-stage introduction of sugar moieties is desirable, and click chemistry is ideal for this purpose. Shinkai and co-workers successfully introduced porphyrin moieties into polysaccharides through the click reaction of alkynylporphyrin 202.1 with azido sugar 202.2 (Scheme 202).510 Vicente and co-workers applied the CuAAC reaction of propargylated porphyrin 202.4 with azido sugar 202.5 to synthesize porphyrin−carbohydrate conjugate 202.6 effiCD

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 161. Preparation of Butadiyne-Linked Cyclic Oligoporphyrins through Homocoupling

peptides and the limitations of the usable peptide sequence. The CuAAC reaction was reported to be effective to prepare porphyrin−peptide conjugate 207.3 (Scheme 207).523 Tris(benzyltriazolylmethyl)amine (TBTA) was employed as a multidentate ligand to enhance the catalytic activity and prevent the metalation of the free-base porphyrin by copper. This methodology was applied to the synthesis of peptideappended manganese porphyrins, the catalase- and peroxidaselike activities of which were examined. A Zn(II) porphyrin containing eight peptide substituents was also prepared through a click reaction of an alkynylporphyrin and azido peptide.524 A dendritic porphyrin was prepared through a click reaction between an azido-functionalized porphyrin and poly(L-lysine) dendrons.525 The toxicity and phototoxicity of the conjugate were examined in vitro. Gene transfection by the porphyrin− poly(L-lysine) conjugate was also investigated.526 Functionalized porphyrins with oligonucleotides were also prepared through click reactions.527,528 DNA−porphyrin conjugate 208.3 was efficiently fabricated with β-functionalized porphyrin 208.1 through a microwave-assisted CuAAC reaction with 208.2 (Scheme 208).529 The porphyrin moiety attached to complementary strands of DNA, resulting in the formation of H-aggregates in the minor groove, enhancing the thermal stability of the duplex. 4.4.6. Click Reactions for Surface Chemistry. In comparison to solution-phase reactions, the access of reagents

nanocrystals could be also functionalized through a similar strategy using a water-soluble alkynylporphyrin.518 Porphyrinbased glycoclusters were prepared in a similar CuAAC reaction.519 The inhibition activity of the glycoclusters toward plant and human lectins was investigated. A series of watersoluble porphyrin-cored dendrimers were fabricated through the click reaction of tetrapropargyloxyporphyrin with a dendritic azide.520 Porphyrins functionalized with β-cyclodextrin and permethylated β-cyclodextrin units were also prepared by CuAAC reactions of tetrapropargylporphyrin 186.1 with azido cyclodextrins 206.1 (Scheme 206).521 Intermolecular inclusion complexation of these cyclodextrin-decorated porphyrins with tetrasodium tetraphenylporphyrintetrasulfonate led to the formation of nanoarchitectures. A similar water-soluble porphyrin functionalized with eight permethylated β-cyclodextrin units was also synthesized.522 The cyclodextrin− porphyrin conjugate was self-assembled with pristine C60 to construct a nanorod structure. Porphyrin−peptide conjugates are being actively investigated for their use in the areas of catalytic oxidation, ion channels, protein recognition, DNA binding, and so forth. However, the synthesis of these molecules is not trivial as in the case of porphyrin−peptide conjugates. The conventional synthesis of these materials using standard amide coupling reagents has serious drawbacks, such as the necessity to use protected CE

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 162. Templated Synthesis of Porphyrin Dimers and Trimers

to heterogeneous surfaces is limited, resulting in lower efficiency of reactions at the surface. Consequently, high reactivity is required for surface modification reactions. Because of its high efficiency, click chemistry enables efficient surface modification. Using the CuAAC reaction, various surfaces can be efficiently functionalized with porphyrin units. Collman and co-workers employed click chemistry to immobilize a manganese porphyrin complex onto the surface of silica gel (Scheme 209).530 Azide moieties were anchored onto the silica surface through a silane-coupling reaction to obtain 209.2. Then porphyrin unit 209.1 was connected to the terminal azide groups through a CuAAC reaction. The catalytic activity of the immobilized manganese catalyst 209.3 was examined in epoxidation reactions. van Koten and co-workers also prepared silica-immobilized epoxidation catalysts through click reactions.531 Collman and co-workers also reported the click reaction of an alkynylporphyrin onto a gold surface functionalized by azidothiol.532

Hydrogenated silicon surfaces can react with alkenes and alkynes through hydrosilylation reactions. This unique reactivity can be used for the surface functionalization of hydrogenated silicon. Azide moieties were installed on the silicon surface through hydrogenation with a chloroalkene followed by nucleophilic azidation (Scheme 210). The terminal azide groups on 210.2 can be converted to triazole units through CuAAC reaction with alkynylporphyrin 137.1.533 Alkynyl groups can also be introduced via surface hydrosilylation reaction with diynes, which can then undergo click reaction with an azido porphyrin.534 Mesoporous silica is an attractive material as a catalyst support because of its high surface area, large pore size, and stability. The surface of mesoporous silica can be effectively functionalized through click chemistry (Scheme 211).535 Iron porphyrin complex 211.2 was anchored on the surface of 211.1 through a CuAAC reaction. The catalytic activity of this immobilized iron catalyst 211.3 was tested in a carbeneinsertion reaction with ethyl diazoacetate. The surface of a CF

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 163. Synthesis of Porphyrin Nanorings through Templated Homocoupling

conductive fluorine-doped tin oxide (FTO) can be similarly modified.536 Porphyrin moieties can be installed onto the surface of magnetic iron oxide nanoparticles through click chemistry.537 Click chemistry has also been applied to the formation of multilayered structures on surfaces that had been fabricated through the layer-by-layer method.538,539 An azide layer was prepared on a surface using the self-assembled monolayer (SAM) technique and the terminal azide groups were functionalized by tetra(ethynylphenyl)porphyrin. The alkyne moieties of the top surface then underwent another click reaction to form an azide layer. This strategy was also applicable for the modification of ITO (indium tin oxide) surfaces.540

azaindole group at the meso-position of a (4-ethylaminopyrid-3yl)ethynylporphyrin was catalyzed by Cu(OAc)2 (Scheme 213).544 This methodology was applicable to the synthesis of meso−meso linked diporphyrin 213.2 functionalized by azaindole moieties, which formed the self-assembled trimer 213.3. A pyrrole moiety was formed through copper-catalyzed addition−cyclization of 1,3-butadiynes with amines. By using this methodology, butadiyne-linked diporphyrin 214.1 was converted into pyrrole-bridged diporphyrin 214.2 upon reaction with aniline in the presence of CuCl (Scheme 214).545 Chan and co-workers expeditiously employed the Dötz reaction546,547 to fabricate various porphyrin−quinone conjugates (Scheme 215).323,324 meso-Alkynylporphyrin 215.1 underwent annulation with Fischer carbene complex 215.2 to provide substituted naphthoquinone derivative 215.3 after oxidation of the resultant hydroxyl group with PbO2. The Pauson−Khand reaction is a useful methodology to build up cyclopentenone rings from alkenes, alkynes, and carbon monoxide (or metal carbonyl complexes).548,549 Horn and Senge employed meso-allylporphyrins and ethynylphenylporphyrins as alkenyl and alkynyl substrates in the Pauson− Khand reaction.550 Treatment of ethynylphenylporphyrin 216.1 with norbornene in the presence of a stoichiometric amount of Co2(CO) 8 afforded a porphyrin bearing a cyclopentenone ring 216.2 (Scheme 216).

4.5. Other Reactions Involving Alkenes and Alkynes

An alkynyl group at the meso position of porphyrins was transformed into indole and benzofuran units though palladium-catalyzed cyclization with iodoaniline and iodophenol derivatives (Scheme 212).541 This type of [3 + 2] annulation of alkynes was originally developed by Larock and co-workers.542,543 The reaction of alkynylporphyrin 212.1 with aminoiodopyridine 212.2 provided azaindolylporphyrin 212.3 in good yield under palladium catalysis. The resultant Zn(II) porphyrin 212.3, containing an azaindole group, self-assembled into a dimeric structure both in solution and solid states. The cyclization of alkynes with heteroatomic nucleophiles can often be catalyzed by copper complexes. Formation of an CG

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 164. Synthesis of Gigantic Porphyrin Nanorings through the Vernier Templating Strategy

macrocycles has been explored.551,552 Such new methodologies would be beneficial in porphyrin chemistry to create exotic porphyrins with interesting structures and fascinating properties.

5. CONSTRUCTION OF PORPHYRIN MACROCYCLES THROUGH ORGANOMETALLIC METHODOLOGIES In general, porphyrin synthesis relies on the acid-catalyzed condensation of aldehydes, pyrroles, and oligopyrrolic building blocks such as dipyrromethanes followed by dehydrogenative oxidation. This standard method of porphyrin synthesis is convenient but generally low-yielding. Moreover, this strategy is not versatile enough to afford novel porphyrin-like skeletons. To access diverse types of novel porphyrinic compounds, the use of organometallic methodologies in the construction of

5.1. Porphyrin Analogues Containing Carbazole and Indole Units

Carbazole and indole constitute partial structures of meso,βfused porphyrins (Figure 3). Because there are a number of transformations of carbazoles and indoles available, macrocycles based on these heterocycles have been actively investigated. CH

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 165. Synthesis of a Chlorophyll a Glycoconjugate

Scheme 166. Synthesis of a C60−Porphyrin Dyad

Scheme 167. Enyne Metathesis for the Construction of a 1,3-Diene Moiety

217.1 and 2,6-dibromopyridine (Scheme 217).553 This molecule (217.2) exhibited metal coordination ability in the central cavity, similar to that of porphyrins. The effect of

Norouzi-Arasi and co-workers prepared a porphyrin-like macrocyclic biscarbazole 217.2 linked by two pyridine units through Suzuki−Miyaura coupling between diborylcarbazole CI

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 168. Ring-Closing Metathesis for Benzoannulation

Scheme 169. Ring-Closing Metathesis for Macrocyclization

metalation on the optical properties of this molecule was investigated. Assembly of Co(II) complex 217.3 on a Ni(001) surface was observed by scanning tunneling microscopy.554 Pyrrole-bridged cyclic biscarbazole 217.6 was also constructed through Suzuki−Miyaura coupling of diiodocarbazole 217.4 with 2,5-diborylated pyrrole 217.5.555 Oxidation of the

resultant macrocycle 217.6 with MnO2 induced a vivid color change due to the change in the conjugation pathway. The oxidized macrocycle 217.7 has a porphyrin-like macrocyclic conjugation, which results in distinct 18π aromaticity and nearIR absorption bands. CJ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 170. Synthesis of a Star-Shaped Porphyrin Pentamer through Alkene Metathesis

Maeda and co-workers prepared 1,3-butadiyne-bridged carbazole cyclic oligomers through Glaser coupling of diethynylcarbazole in the presence of Cu(OAc)2 and pyridine (Scheme 218).556 The cyclization of diethynylcarbazole precursor 218.1 afforded cyclic dimer 218.2, trimer 218.3, and a tetramer. These cyclic carbazoles were employed as precursors for porphyrin-like macrocycles on the basis of the reactivity of 1,3-butadiyne moieties toward heteroatom nucleophiles. Cu(I)-catalyzed amination of 218.2 with aniline afforded pyrrole-linked cyclic carbazoles 218.4. The reaction of

218.2 with sodium sulfide gave cyclic carbazoles 218.5 containing thiophene units. Dehydrogenative oxidation of the thiophene-bridged biscarbazole 218.5 with MnO2 furnished core-modified porphyrin analogues 218.6. Interestingly, the oxidized macrocycle 218.6 exhibited aromatic character due to macrocyclic conjugation, and intense NIR absorption bands from 800 to 1100 nm. This strategy was further extended to the synthesis of π-extended porphyrin analogues with alkynyl groups.557,558 CK

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 171. Synthesis of a Porphyrin-Based Supramolecular Macrocycle

Scheme 172. Template-Directed Synthesis of Porphyrin Dimers and Trimers through Alkene Metathesis

CL

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 173. Synthesis of a Porphyrin Box through Template-Directed Metathesis

dipyrrins should provide macrocycles that exhibit porphyrinlike photophysical and electronic properties. Boron dipyrrins (BODIPY) have been extensively investigated due to their high thermal and photochemical stability as well as strong absorption and emission in the visible region.567−570 As macrocyclic BODIPYs, butadiyne-linked BODIPY dimer 222.2 and trimer 222.3 were constructed through Glaser-type direct homocoupling of alkynylsilane 222.1 (Scheme 222).571 The rigid butadiyne and BF2 units in the macrocycles maintained planar molecular structures to allow effective extension of the π-conjugation. These cyclic BODIPY oligomers 222.2 and 222.3 exhibited antiaromatic character due to their 24π and 34π conjugated systems, which were indicated by their 1H and 19F NMR spectra. Interestingly, the cyclic BODIPY dimer 222.2 and trimer 222.3 have no emission due to their antiaromaticity. Bisdipyrrin metal complexes 223.1 and 223.4 were utilized as precursors to access macrocyclic dipyrrin dimers efficiently. In this strategy, the central metal of the complexes acted as a template to facilitate the cyclization reaction. In addition, this strategy should suppress the formation of linear or cyclic higher dipyrrin oligomers. In fact, palladium-catalyzed Buchwald− Hartwig amination of bisdipyrrin Ni(II) complex 223.1 with benzylamine afforded azacorroles 223.2 and 223.3572 as major products (Scheme 223). A trace amount of 5,15-diazaporphyrin 9.1 was also detected.573 The 1H NMR spectra of these azaporphyrinoids confirmed their 18π aromatic character. Interestingly, the use of a different catalyst combination based on Pd(OAc)2−XPhos afforded diazaporphyrin 9.1 as a major product along with azacorrole 223.2 in a decreased yield. Concomitantly, Matano and co-workers disclosed a metaltemplated strategy using bisdipyrrin precursor 223.4 to prepare diazaporphyrin 9.1.574,575 Treatment of bisdipyrrin 223.4 with sodium azide in the presence of copper(I) iodide efficiently afforded 5,15-diazaporphyrin 9.1 in excellent yield. Matano et al. further extended this strategy to construct stable 20π antiaromatic diazaporphyrins.576 Bisdipyrrin metal complexes 223.1 and 223.4 were also convenient precursors to 5,15dithiaporphyrins and 10-thiacorroles.577,578 Azacorroles were also prepared from nitrogen-bridged bisdipyrrin 224.2, which was obtained through palladium-

Maeda et al. also synthesized porphyrin analogue 219.4 containing both carbazole and indole units in a stepwise manner from 219.1 through Sonogashira coupling, indiumcatalyzed indole cyclization, and Migita−Kosugi−Stille coupling with 2,5-distannylthiophene (Scheme 219).559 The oxidized macrocycle 219.5 accommodated a Pd(II) ion in the cavity to exhibit NIR absorption. Various carbazole-based expanded thiaporphyrins were synthesized via Migita−Kosugi− Stille coupling between dibromocarbazole 220.1 and bis(tributylstannyl)bithiophene 220.2 or intramolecular Glaser coupling of diyne 220.4 followed by thiophene formation (Scheme 220).560 On the basis of cyclization of the 1,3butadiyne moiety of 220.5, various porphyrin analogues containing pyrrole, furan, thiophene, and selenophene units were systematically prepared and their properties were investigated.561,562 As examples of indole-based porphyrinoids, cyclic tetraindole 221.3 and its oxidized form 221.4 were synthesized through iridium-catalyzed direct borylation of bisindole 221.1 followed by Suzuki−Miyaura coupling of the resultant boronic ester species 221.2 (Scheme 221).563 Treatment of nonplanar cyclic tetraindole 221.3 with bromine in aqueous THF resulted in formation of a planar macrocycle 221.4 consisting of two indole and two indolinone units. Owing to the porphyrin-like fournitrogen central cavity, this macrocycle 221.4 can capture Cu(II), Zn(II), and Cu(II) to afford the corresponding metal complexes 221.5, in which substantial red-shifts of the absorption spectra were observed. Interestingly, the Ni(II) complex exhibited a black color in solution due to broad absorption bands that cover most of the visible region. Analogous indole-based macrocycles containing benzofuran and benzothiophene were prepared similarly through iridiumcatalyzed direct borylation and Suzuki−Miyaura coupling.564 Macrocyclic tetraindoles were also obtained though oxidative coupling of bisindoles without the use of organometallic reagents.565 5.2. Porphyrin Analogues from Dipyrrin Precursors

Dipyrrins can coordinate to various metal ions as bidentate ligands to afford metal complexes that often exhibit excellent photophysical and electrochemical properties.566 Because the dipyrrin unit is almost half of a porphyrin, dimerization of CM

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 174. Synthesis of a Molecular Imprinting Polymer

bromo substituent to the adjacent β-position. Al(III)-metalation of nitrogen-bridged bisdipyrrin 224.2 also induced formation of a 10-azacorrole, suggesting that the size of the metal ions is a key factor in this cyclization. Nitrogen-bridged bisdipyrrin 224.2 was also an effective precursor to free-base

catalyzed amination of dibromodipyrrin 224.1 (Scheme 224).579 Bisdipyrrin 224.2 acted as a tetradentate ligand to Zn(II) to afford helical complex 224.4. In contrast, metalation with Ni(II) directly provided 10-azacorrole Ni(II) complex 224.3 in good yield. This cyclization involved migration of the CN

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 175. Synthesis of a Porphyrin Box via Alkyne Metathesis

Scheme 176. Synthesis of Porphyrin Dimers and Trimers through [2 + 2+2] Cycloaddition

10-azacorrole 224.6.580 Ni(0)-mediated homocoupling of nitrogen-bridged bisdipyrrin Zn(II) complex 224.4 furnished 10-azacorrole Zn(II) complex 224.5 in good yield. Acid

treatment of Zn(II) complex 224.5 afforded free-base azacorrole 224.6, which was used as a precursor to other metalated azacorroles. CO

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 177. Ruthenium-Catalyzed Click Reaction

Scheme 178. Click Reaction of a β-Azidotetraarylporphyrin

the formation of doubly ethynylene-bridged bisdipyrrin 226.4 (Scheme 226).591 Tetrabromoporphoquinodimethane 226.3 was prepared from meso-unsubstituted porphyrin 226.1 in two steps. Treatment of 226.3 with t-BuLi at low temperatures furnished tetradehydro[20]porphyrin(2.1.2.1) 226.4. Interestingly, 226.4 exhibited distinct antiaromatic character.

Norcorrole is a ring-contracted porphyrin lacking two of the meso-carbons of a regular porphyrin. Isolation of norcorroles has been challenging due to their unstable nature.581 Ni(II) norcorrole complex 225.1 was isolated as a stable molecule by the use of bulky meso-mesityl substituents as steric protection.582 Ni(0)-mediated homocoupling583 of bisdipyrrin Ni(II) complex 223.4 provided Ni(II) dimesitylnorcorrole 225.1 in excellent yield (Scheme 225). Ni(II) dimesitylnorcorrole 225.1 is the smallest isolable antiaromatic porphyrinoid prepared to date. Because of its simple synthetic procedure, a gram-scale synthesis of 225.1 was achieved. Due to its distinct antiaromatic character, 225.1 is highly redox active, enabling its use as an electrode-active material in rechargeable batteries.584 Ni(II) norcorrole complex 225.1 also exhibited unique reactivities that are not typically observed for aromatic porphyrinoids. Regioselective silylene insertion, oxidative nucleophilic substitution, and regioselective hydrogenation of 225.1 were reported.585−589 Interestingly, a similar Ni(0)mediated coupling of Pd(II) dibromodipyrrin complex 225.2 exclusively provided the octaphyrin(1.0.1.0.1.0.1.0) bispalladium complex 225.3.590 Complex 225.3 adopted a nonplanar figure-of-eight conformation but its weak antiaromatic character was supported by its 1H NMR spectrum and DFT calculations. Osuka and co-workers reported that a Fritsch−Buttenberg− Wiechell rearrangement initiated by lithiation of 226.3 led to

6. SUMMARY AND OUTLOOK As we have described in this review, catalytic and stoichiometric organometallic methodologies have been actively investigated and widely used in the synthesis of porphyrins and related compounds. For example, various types of cross-coupling methodologies have already become standard synthetic procedures in porphyrin synthesis. The use of such modern techniques combined with traditional, well-established porphyrin synthetic methods have enabled chemists to access a wide range of highly elaborate and complex porphyrin architectures. In the field of organometallic chemistry, numerous novel methodologies have been reported. However, many of these new reactions are not particularly useful and some remain unused in organic synthesis, even though they employ novel reagents and new reaction mechanisms. In the flood of such newly developed reactions, chemists should explore and find the specific utility of them to achieve the ideal synthesis of desired materials. CP

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 179. Click Reaction of a meso-Azidotetraarylporphyrin

Scheme 180. Preparation of a Tris(triazoylmethyl)amine-Linked Triporphyrin

Owing to the fruitful development of the synthesis of porphyrins, porphyrin chemists have gained a variety of potent tools to create various types of porphyrin-based materials that are difficult to prepare through conventional organic chemistry without organometallic methodologies. Consequently, porphyr-

in chemists can now focus on exploring the properties and functions of sophisticated porphyrins without much synthetic effort. The important goal for chemists is to achieve intriguing and useful properties and functions of materials not the CQ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 181. Preparation of a Face-to-Face Porphyrin Dimer with Dendrons through Click Reaction

synthesis itself. To approach this goal, new design concepts of functional porphyrins are strongly awaited.

with Prof. Yoshiaki Nishibayashi on the development of novel transition metal-catalyzed and photochemical reactions. Since 2013, he has been an Associate Professor at Nagoya University, working with Prof. Hiroshi Shinokubo. He received the Japanese Incentive Award in Synthetic Organic Chemistry in 2014. His research interests include organic synthesis, organometallic chemistry, and the synthesis of unique π-conjugated molecules and functional materials.

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. Notes

Hiroshi Shinokubo was born in Kyoto in 1969. He graduated from Kyoto University in 1992 and obtained his doctorate there in 1998 under the guidance of Profs. Kiitiro Utimoto and Koichiro Oshima. He became Assistant Professor in 1995 at the Department of Engineering of Kyoto University. He then worked with Prof. Rick L. Danheiser at the Massachusetts Institute of Technology from 1999 to 2000 as a visiting scientist, and later collaborated with Prof. Atsuhiro Osuka at the Department of Science, Kyoto University, as Associate Professor from 2003 to 2008. He was selected to be a PRESTO researcher at the Japan Science and Technology Agency from 2003 to 2007. In 2008, he became Professor at the Department of Engineering, Nagoya University. He received the Chemical Society of Japan Award for Young Chemists in 2004, the Minister Award for Distinguished Young Scientists from MEXT in 2008, the JSPS Prize in 2012, and the SSOCJ DIC Award for Functional Material Chemistry 2012. His research has centered around developing efficient syntheses of novel organic molecules that have fascinating structures, properties and functions. His current targets include new porphyrin analogues and large polyaromatic molecules.

The authors declare no competing financial interest. Biographies Satoru Hiroto was born in Shimane Prefecture, Japan. He received his B.S. in Science from Kyoto University in 2004 and then entered the Graduate School of Science at Kyoto University and worked with Profs. Atsuhiro Osuka and Hiroshi Shinokubo to develop catalytic methods to functionalize porphyrin derivatives. He received his Ph.D. in Science from Kyoto University in 2009. Dr. Hiroto started his professional career at Nagoya University as an Assistant Professor in 2009. From 2015 to 2016, he joined Prof. C. L. Fraser’s group at the University of Virginia, USA, as a visiting researcher working on the development of novel stimuli-responsive materials. At Nagoya University, Dr. Hiroto works on the application of organic synthesis to the production of molecular devices, and developing synthetic methodology towards distorted π-conjugated molecules and dyes with chiroptical and electron-conducting properties. Yoshihiro Miyake was born in Hiroshima, Japan, in 1974. He received his B.Sc. degree in 1997 and Ph.D. in 2002 from Kyoto University under the supervision of Prof. Sakae Uemura. After postdoctoral research with Profs. Michinori Suginome and Masahiro Murakami, he became an Assistant Professor at Tokyo Metropolitan University in 2002, working with Prof. Masahiko Iyoda. He then moved to the University of Tokyo as an Assistant Professor in 2005 and collaborated

ACKNOWLEDGMENTS Financial support from Grant-in-Aid for Scientific Research on Innovative Areas (2601): π-System Figuration (JSPS KAKENHI Grant Number JP26102003) is gratefully acknowledged. CR

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 182. Construction of a Dendritic Multi-Porphyrin Cluster

ABBREVIATIONS Ac Ar BINAP BODIPY

Bpin bpy Bu t-Bu CMD

acetyl aryl 2,2′-bis(diphenylphosphino)-1,1′-binaphthyl 4,4-difluoro-4-bora-3a,4a-diaza-s-indacene CS

4,4,5,5-tetramethyl-1,3,2-dioxabororanyl 2,2′-bipyridyl butyl tert-butyl concerted-metalation/deprotonation DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 183. Preparation of a Dimeric Porphyrin Cage

cod CuAAC Cy CyJohnphos DABCO DavePhos dba DBU DDQ DIPEA DiPPF DMA DMF DMSO DPEphos DPPF DPPP DSSC dtbpy FTO IPent IPr

IR ITO NBS NIR PCBM PDT PEG PEPPSI

1,5-cyclooctadiene copper-catalyzed azide−alkyne cycloaddition cyclohexyl 2-(dicyclohexylphosphino)biphenyl 1,4-diazabicyclo[2.2.2]octane 2-dicyclohexylphosphino-2′-(N,Ndimethylamino)biphenyl dibenzylideneacetone 1,8-diazabicyclo[5.4.0]undec-7-ene 2,3-dichloro-5,6-dicyano-p-benzoquinone N,N-diisopropylethylamine 1,1′-dis(diisopropylphosphino)ferrocene N,N-dimethylacetamide N,N-dimethylformamide dimethyl sulfoxide 2,2′-bis(diphenylphosphino)diphenyl ether 1,1′-bis(diphenylphosphino)ferrocene 1,3-bis(diphenylphosphino)propane dye sensitized solar cell 4,4′-di-tert-butyl-2,2′-bipyridyl fluorine-doped tin oxide 1,3-bis(2,6-diisopentylphenyl)imidazol-2-ylidene 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene

Ph PMDETA POM i-Pr racRCM RGD ROMP rt RuPhos SAM SPhos

CT

infrared indium tin oxide N-bromosuccinimide near-infrared [6,6]-phenyl-C61-butyric acid methyl ester photodynamic therapy poly(ethylene glycol) pyridine-enhanced precatalyst preparation stabilization and initiation phenyl N,N,N′,N″,N″-pentamethyldiethylenetriamine polyoxometalate isopropyl racemic ring closing metathesis arginine-glycine-aspartic acid ring opening metathesis polymerization room temperature 2-dicyclohexylphosphino-2′,6′-diisopropoxybiphenyl self-assembled monolayer 2-dicyclohexylphosphino-2′,6′-dimethoxybiphenyl DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 184. Preparation of a Diethynylbisindole-Linked Diporphyrin through a Click Reaction

Tf TFA THF Tol TPA TPP TPPTS Mes Xantphos

(10) Hiroto, S.; Yamaguchi, S.; Shinokubo, H.; Osuka, A. Porphyrin Derivatives with Carbon-Metal Bonds. Yuki Gosei Kagaku Kyokaishi 2009, 67, 688−700. (11) Shinokubo, H.; Osuka, A. Marriage of Porphyrin Chemistry with Metal-Catalysed Reactions. Chem. Commun. 2009, 1011−1021. (12) Ren, T. Peripheral Covalent Modification of Inorganic and Organometallic Compounds through C−C Bond Formation Reactions. Chem. Rev. 2008, 108, 4185−4207. (13) Suzuki, A. Cross-Coupling Reactions of Organoboron Compounds with Organic Halides. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 1998; pp 48−97. (14) Miyaura, N. Metal-Catalyzed Cross-Coupling Reactions of Organoboron Compounds with Organic Halides. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 2004; pp 41−123. (15) Miyaura, N. Organoboron Compounds. In Cross-Coupling Reactions: A Practical Guide; Miyaura, N., Ed.; Springer: Heidelberg, 2002; pp 11−59. (16) Miyaura, N.; Suzuki, A. Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem. Rev. 1995, 95, 2457− 2483. (17) Mitchell, T. N. Palladium-Catalysed Reactions of Organotin Compounds. Synthesis 1992, 1992, 803−815. (18) Fugami, K.; Kosugi, M., Organotin Compounds. In CrossCoupling Reactions: A Practical Guide; Miyaura, N., Ed.; Springer: Heidelberg, 2002; pp 87−130. (19) Mitchell, T. N. Organotin Reagents in Cross-Coupling. MetalCatalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 1998; pp 167−202. (20) Mitchell, T. N. Organotin Reagents in Cross-Coupling Reactions. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 2004; pp 125−161. (21) Negishi, E.; Liu, F. Palladium- or Nickel-Catalyzed CrossCoupling with Organometals Containing Zinc, Magnesium, Aluminum, and Zirconium. Metal-Catalyzed Cross-Coupling Reactions; WileyVCH: Weinheim, 1998; pp 1−47.

trifluoromethanesulfonyl trifluoroacetic acid tetrahydrofuran tolyl two-photon absorption tetraphenylporphyrin 3,3′,3″-phosphanetriyltris(benzenesulfonic acid) trisodium salt mesityl 4,5-bis(diphenylphosphino)-9,9-dimethylxanthene

REFERENCES (1) The Porphyrin Handbook; Kadish, K. M., Smith, K. M., Guilard, R., Eds.; Academic Press: San Diego, 2000; Vol. 1−20. (2) Dolphin, D. The Porphyrins; Academic Press: New York, 1979. (3) Comprehensive Organometallic Chemistry III; Crabtree, R., Mingos, M. P., Eds, Elsevier: Oxford, 2007. (4) Tsuji, J. Transition Metal Reagents and Catalysts; Wiley: Chichester, U.K., 2000. (5) Hegedus, L. S. Transition Metals in the Synthesis of Complex Organic Molecules; University Science Books: Mill Valley, 1999. (6) Handbook of Metathesis; Grubbs, R. H., Ed.; Wiley-VCH: Weinheim, 2003. (7) Sergeeva, N. N.; Senge, M. O.; Ryan, A. Organometallic C−C Coupling Reactions for Porphyrins. In Handbook of Porphyrin Science; Kadish, K. M., Smith, K. M., Guilard, R., Eds.; World Scientific: Singapore, 2010; Vol. 3, pp 325−365. (8) Beletskaya, I. P.; Tyurin, V. S.; Uglov, A.; Stern, C.; Guilard, R. In Handbook of Porphyrin Science; Kadish, K. M., Smith, K. M., Guilard, R., Eds.; World Scientific: Singapore, 2012; Vol. 23, pp 81−279. (9) Satake, A. Transition Metal-Mediated Functionalization and Multiplication of Porphyrins toward Giant Functional Systems. Yuki Gosei Kagaku Kyokaishi 2007, 65, 298−307. CU

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 185. Introduction of Porphyrin Units onto a POM

(22) Negishi, E.; Zeng, X.; Tan, Z.; Qian, M.; Hu, Q.; Huang, Z. Palladium- or Nickel-Catalyzed Cross-Coupling with Organometals Containing Zinc, Aluminum, and Zirconium: The Negishi Coupling. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 2004; pp 815−889. (23) Knochel, P. Carbon−Carbon Bond Formation Reactions Mediated by Organozinc Reagents. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 1998; pp 386−419. (24) Knochel, P.; Calaza, M. I.; Hupe, E. Carbon−Carbon BondForming Reactions Mediated by Organozinc Reagents. MetalCatalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 2004; pp 619−670. (25) Sonogashira, K. Development of Pd−Cu Catalyzed CrossCoupling of Terminal Acetylenes with sp2-Carbon Halides. J. Organomet. Chem. 2002, 653, 46−49. (26) Negishi, E.; Anastasia, L. Palladium-Catalyzed Alkynylation. Chem. Rev. 2003, 103, 1979−2018. (27) Chinchilla, R.; Nájera, C. The Sonogashira Reaction: A Booming Methodology in Synthetic Organic Chemistry. Chem. Rev. 2007, 107, 874−922. (28) Sonogashira, K. Cross-Coupling Reactions to sp Carbon Atoms. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 1998; pp 203−229. (29) Marsden, J. A.; Haley, M. M. Cross-Coupling Reactions to sp Carbon Atoms. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 2004; pp 317−394. (30) Chinchilla, R.; Najera, C. Recent Advances in Sonogashira Reactions. Chem. Soc. Rev. 2011, 40, 5084−5121.

(31) de Meijere, A.; Meyer, F. E. Fine Feathers Make Fine Birds: The Heck Reaction in Modern Garb. Angew. Chem., Int. Ed. Engl. 1995, 33, 2379−2411. (32) Beletskaya, I. P.; Cheprakov, A. V. The Heck Reaction as a Sharpening Stone of Palladium Catalysis. Chem. Rev. 2000, 100, 3009− 3066. (33) Bräse, S.; De Meijere, A. Palladium-Catalyzed Coupling of Organyl Halides to Alkenes  the Heck Reaction. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 1998; pp 98−166. (34) Bräse, S.; Meijere, A. D. Cross-Coupling of Organyl Halides with Alkenes: the Heck Reaction. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, 2004; pp 217−315. (35) Muci, A. R.; Buchwald, S. L. Practical Palladium Catalysts for C−N and C−O Bond Formation. In Cross-Coupling Reactions: A Practical Guide; Miyaura, N., Ed.; Springer: Heidelberg, 2002; pp 131− 209. (36) Jiang, L.; Buchwald, S. L. Palladium-Catalyzed Aromatic Carbon−Nitrogen Bond Formation. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH, 2004; pp 699−760. (37) Wolfe, J. P.; Wagaw, S.; Marcoux, J.-F.; Buchwald, S. L. Rational Development of Practical Catalysts for Aromatic Carbon−Nitrogen Bond Formation. Acc. Chem. Res. 1998, 31, 805−818. (38) Hartwig, J. F. Transition Metal Catalyzed Synthesis of Arylamines and Aryl Ethers from Aryl Halides and Triflates: Scope and Mechanism. Angew. Chem., Int. Ed. 1998, 37, 2046−2067. (39) Hartwig, J. F. Evolution of a Fourth Generation Catalyst for the Amination and Thioetherification of Aryl Halides. Acc. Chem. Res. 2008, 41, 1534−1544. CV

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 186. Preparation of a Porphyrin [3]rotaxane through a Click Reaction

(47) Kolb, H. C.; Finn, M. G.; Sharpless, K. B. Click Chemistry: Diverse Chemical Function from a Few Good Reactions. Angew. Chem., Int. Ed. 2001, 40, 2004. (48) Tornoe, C. W.; Christensen, C.; Meldal, M. Peptidotriazoles on Solid Phase: [1,2,3]-Triazoles by Regiospecific Copper(I)-Catalyzed 1,3-Dipolar Cycloadditions of Terminal Alkynes to Azides. J. Org. Chem. 2002, 67, 3057−3064. (49) Rodionov, V. O.; Presolski, S. I.; Diaz Diaz, D.; Fokin, V. V.; Finn, M. G. Ligand-Accelerated Cu-Catalyzed Azide−Alkyne Cycloaddition: A Mechanistic Report. J. Am. Chem. Soc. 2007, 129, 12705− 12712. (50) Meldal, M.; Tornøe, C. W. Cu-Catalyzed Azide−Alkyne Cycloaddition. Chem. Rev. 2008, 108, 2952−3015. (51) Yamaguchi, J.; Yamaguchi, A. D.; Itami, K. C−H Bond Functionalization: Emerging Synthetic Tools for Natural Products and Pharmaceuticals. Angew. Chem., Int. Ed. 2012, 51, 8960−9009. (52) Segawa, Y.; Maekawa, T.; Itami, K. Synthesis of Extended πSystems through C−H Activation. Angew. Chem., Int. Ed. 2015, 54, 66−81. (53) Handbook of C−H Transformations; Dyker, G., Ed.; Wiley-VCH: Weinheim, 2005.

(40) Surry, D. S.; Buchwald, S. L. Biaryl Phosphane Ligands in Palladium-Catalyzed Amination. Angew. Chem., Int. Ed. 2008, 47, 6338−6361. (41) Siemsen, P.; Livingston, R. C.; Diederich, F. Acetylenic Coupling: A Powerful Tool in Molecular Construction. Angew. Chem., Int. Ed. 2000, 39, 2632−2657. (42) Ritter, T.; Carreira, E. M. C−H Transformation at Terminal Alkynes. Handbook of C−H Transformations; Wiley-VCH: Weinheim, 2005; pp 29−85. (43) Ley, S. V.; Thomas, A. W. Modern Synthetic Methods for Copper-Mediated C(aryl)−O, C(aryl)−N, and C(aryl)−S Bond Formation. Angew. Chem., Int. Ed. 2003, 42, 5400−5449. (44) Beletskaya, I. P.; Cheprakov, A. V. Copper in Cross-Coupling Reactions: The Post-Ullmann Chemistry. Coord. Chem. Rev. 2004, 248, 2337−2364. (45) Huisgen, R. 1,3-Dipolar Cycloadditions. Past and Future. Angew. Chem., Int. Ed. Engl. 1963, 2, 565−598. (46) Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. A Stepwise Huisgen Cycloaddition Process: Copper(I)-Catalyzed Regioselective “Ligation” of Azides and Terminal Alkynes. Angew. Chem., Int. Ed. 2002, 41, 2596−2599. CW

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 187. Preparation of Corrole−BODIPY and Core-Modified Porphyrin−BODIPY Dyads

2,2′-Bipyridine Complexes at Room Temperature. Angew. Chem., Int. Ed. 2002, 41, 3056−3058. (62) Hartwig, J. F. Borylation and Silylation of C−H Bonds: A Platform for Diverse C−H Bond Functionalizations. Acc. Chem. Res. 2012, 45, 864−873. (63) Mkhalid, I. A. I.; Barnard, J. H.; Marder, T. B.; Murphy, J. M.; Hartwig, J. F. C−H Activation for the Construction of C−B Bonds. Chem. Rev. 2010, 110, 890−931. (64) Ishiyama, T.; Miyaura, N. Metal-Catalyzed Reactions of Diborons for Synthesis of Organoboron Compounds. Chem. Rec. 2004, 3, 271−280. (65) Ishiyama, T.; Miyaura, N. Transition Metal-Catalyzed Borylation of Alkanes and Arenes via C−H Activation. J. Organomet. Chem. 2003, 680, 3−11. (66) Ishiyama, T. Transition Metal-Catalyzed Direct Borylation and Silylation Based on C−H Activation. Yuki Gosei Kagaku Kyokaishi 2005, 63, 440−452. (67) Ishiyama, T.; Miyaura, N. Metal-Catalyzed Borylation of C−H And C−Halogen Bonds of Alkanes, Alkenes, and Arenes for the Synthesis of Boronic Esters. In Boronic Acids: Preparation and Applications in Organic Synthesis, Medicine and Materials; Hall, D., Ed.; Wiley-VCH: Weinheim, 2011; Vol. 1 pp 135−169.

(54) Kakiuchi, F.; Chatani, N. Catalytic Methods for C−H Bond Functionalization: Application in Organic Synthesis. Adv. Synth. Catal. 2003, 345, 1077−1101. (55) Godula, K.; Sames, D. C−H Bond Functionalization in Complex Organic Synthesis. Science 2006, 312, 67−72. (56) Activation of Unreactive Bonds and Organic Synthesis; Murai, S., Ed.; Springer: Berlin, 1999. (57) Cho, J.-Y.; Tse, M. K.; Holmes, D.; Maleczka, R. E.; Smith, M. R. Remarkably Selective Iridium Catalysts for the Elaboration of Aromatic C−H Bonds. Science 2002, 295, 305−308. (58) Cho, J.-Y.; Iverson, C. N.; Smith, M. R. Steric and Chelate Directing Effects in Aromatic Borylation. J. Am. Chem. Soc. 2000, 122, 12868−12869. (59) Ishiyama, T.; Takagi, J.; Ishida, K.; Miyaura, N.; Anastasi, N. R.; Hartwig, J. F. Mild Iridium-Catalyzed Borylation of Arenes. High Turnover Numbers, Room Temperature Reactions, and Isolation of a Potential Intermediate. J. Am. Chem. Soc. 2002, 124, 390−391. (60) Ishiyama, T.; Nobuta, Y.; Hartwig, J. F.; Miyaura, N. Room Temperature Borylation of Arenes and Heteroarenes Using Stoichiometric Amounts of Pinacolborane Catalyzed by Iridium Complexes in an Inert Solvent. Chem. Commun. 2003, 2924−2925. (61) Ishiyama, T.; Takagi, J.; Hartwig, J. F.; Miyaura, N. A Stoichiometric Aromatic C−H Borylation Catalyzed by Iridium(I)/ CX

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 188. Preparation of Zn(II) Porphyrin−C60 Conjugates

(80) Bischoff, I.; Feng, X.; Senge, M. O. One-pot Synthesis of Functionalized, Highly Substituted Porphodimethenes. Tetrahedron 2001, 57, 5573−5583. (81) Sergeeva, N. N.; Senge, M. O. Palladium-Catalyzed Reactions for the Synthesis of Chlorins and 5,10-Porphodimethenes. Tetrahedron Lett. 2006, 47, 6169−6172. (82) Sergeeva, N. N.; Shaker, Y. M.; Finnigan, E. M.; McCabe, T.; Senge, M. O. Synthesis of Hydroporphyrins Based on Comparative Studies of Palladium-Catalyzed and Non-Catalyzed Approaches. Tetrahedron 2007, 63, 12454−12464. (83) Kalisch, W. W.; Senge, M. O. Facile meso Functionalization of Porphyrins by Nucleophilic Substitution with Organolithium Reagents. Angew. Chem., Int. Ed. 1998, 37, 1107−1109. (84) Feng, X.; Bischoff, I.; Senge, M. O. Mechanistic Studies on the Nucleophilic Reaction of Porphyrins with Organolithium Reagents. J. Org. Chem. 2001, 66, 8693−8700. (85) Feng, X.; Senge, M. O. Regioselective Reaction of 5,15Disubstituted Porphyrins with Organolithium Reagents−Synthetic Access to 5,10,15-Trisubstituted Porphyrins and Directly meso−mesoLinked Bisporphyrins. J. Chem. Soc., Perkin Trans. 1 2000, 3615−3621. (86) Feng, X.; Senge, M. O. One-pot Synthesis of Functionalized Asymmetric 5,10,15,20-Substituted Porphyrins from 5,15-Diaryl- or -Dialkylporphyrins. Tetrahedron 2000, 56, 587−590. (87) Senge, M. O.; Bischoff, I. Regioselective Synthesis of Conformationally Designed Porphyrins with Mixed meso-Substituent Types and Distortion Modes. Eur. J. Org. Chem. 2001, 2001, 1735− 1751. (88) Hatscher, S. S.; Senge, M. O. Synthetic Access to 5,15Disubstituted Porphyrins. Tetrahedron Lett. 2003, 44, 157−160. (89) Senge, M. O.; Bischoff, I. SNAr Reactions of β-Substituted Porphyrins and the Synthesis of meso Substituted Tetrabenzoporphyrins. Tetrahedron Lett. 2004, 45, 1647−1650. (90) Ryppa, C.; Senge, M. O.; Hatscher, S. S.; Kleinpeter, E.; Wacker, P.; Schilde, U.; Wiehe, A. Synthesis of Mono- and Disubstituted Porphyrins: A- and 5,10-A2-Type Systems. Chem. - Eur. J. 2005, 11, 3427−3442. (91) Senge, M. O.; Bischoff, I. Synthesis of Benzoporphyrins with One or Two meso-Substituents via Substitution Reactions. Heterocycles 2005, 65, 879−886.

(68) Ishiyama, T.; Takagi, J.; Nobuta, Y.; Miyaura, N. IridiumCatalyzed C−H Borylation of Arenes and Heteroarenes: 1-Chloro-3Iodo-5-(4,4,5,5-Tetramethyl-1,3,2-Dioxaborolan-2-yl)benzene and 2(4,4,5,5,- Tetramethyl-1,3,2-Dioxaborolan-2-yl)indole. Org. Synth. 2005, 82, 126. (69) Ishiyama, T.; Miyaura, N. Iridium-Catalyzed Borylation of Arenes and Heteroarenes via C−H Activation. Pure Appl. Chem. 2006, 78, 1369−1375. (70) Rostovtsev, V. V. C−H Transformation at Arenes. In Handbook of C−H Transformations; Wiley-VCH: Weinheim, 2005; pp 99−256. (71) Miura, M.; Nomura, M. Direct Arylation via Cleavage of Activated and Unactivated C−H Bonds. In Cross-Coupling Reactions: A Practical Guide; Miyaura, N., Ed.; Springer: Heidelberg, 2002; pp 211− 241. (72) Wencel-Delord, J.; Glorius, F. C−H Bond Activation Enables the Rapid Construction and Late-Stage Diversification of Functional Molecules. Nat. Chem. 2013, 5, 369−375. (73) Schlosser, M. Organometallics in Synthesis: A Manual; Wiley: Chichester, 2002. (74) Senge, M. O. Nucleophilic Substitution as a Tool for the Synthesis of Unsymmetrical Porphyrins. Acc. Chem. Res. 2005, 38, 733−743. (75) Krattinger, B.; Callot, H. J. New Routes from Porphyrins to Stable Phlorins. meso-Alkylation and Reduction of meso-Tetraphenyland Octaethylporphyrins. Tetrahedron Lett. 1996, 37, 7699−7702. (76) LeSaulnier, T. D.; Graham, B. W.; Geier, R., III Enhancement of Phlorin Stability by the Incorporation of meso-Mesityl Substituents. Tetrahedron Lett. 2005, 46, 5633−5637. (77) Senge, M. O.; Kalisch, W. W.; Bischoff, I. The Reaction of Porphyrins with Organolithium Reagents. Chem. - Eur. J. 2000, 6, 2721−2738. (78) Krattinger, B.; Callot, H. J. Addition of a Sterically Hindered Organolithium Compounds to meso-Tetraphenylporphyrin. Tetrahedron Lett. 1998, 39, 1165−1168. (79) Krattinger, B.; Callot, H. J. Alkylation and Reduction of Porphyrins and N-Substituted Porphyrins: New Routes to Chlorins and Phlorins. Eur. J. Org. Chem. 1999, 1999, 1857−1867. CY

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 189. Preparation of a BODIPY−Porphyrin−C60 Pentad

(92) Senge, M. O.; Richter, J.; Bischoff, I.; Ryan, A. Highly

(94) Notaras, E. G. A.; Fazekas, M.; Doyle, J. J.; Blau, W. J.; Senge,

Substituted 2,3,7,8,12,13,17,18-Octaethylporphyrins with meso Aryl

M. O. A2B2-Type Push−Pull Porphyrins as Reverse Saturable and

Residues. Tetrahedron 2010, 66, 3508−3524. (93) Feng, X.; Senge, M. O. An Efficient Synthesis of Highly

Saturable Absorbers. Chem. Commun. 2007, 2166−2168. (95) Wiehe, A.; Shaker, Y. M.; Brandt, J. C.; Mebs, S.; Senge, M. O.

Functionalized Asymmetric Porphyrins with Organilithium Reagents.

Lead Structures for Applications in Photodynamic Therapy. Part 1:

J. Chem. Soc., Perkin Trans. 1 2001, 1030−1038.

Synthesis and Variation of m-THPC (Temoporfin) Related CZ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 190. Preparation of a Dendritic Porphyrin−C60 Conjugate

Amphiphilic A2BC-Type Porphyrins. Tetrahedron 2005, 61, 5535− 5564. (96) Bakar, M. B.; Oelgemöller, M.; Senge, M. O. Lead Structures for Applications in Photodynamic Therapy. Part 2: Synthetic Studies for Photo-Triggered Release Systems of Bioconjugate Porphyrin Photosensitizers. Tetrahedron 2009, 65, 7064−7078. (97) Fazekas, M.; Pintea, M.; Senge, M. O.; Zawadzka, M. Synthetic Strategies and Porphyrin Building Blocks for Unsymmetrica; Multichromophores. Tetrahedron Lett. 2008, 49, 2236−2239. (98) Senge, M. O.; Hatscher, S. S.; Wiehe, A.; Dahms, K.; Kelling, A. The Dithianyl Group as a Synthon in Porphyrin Chemistry: Condensation Reactions and Preparation of Formylporphyrins under Basic Conditions. J. Am. Chem. Soc. 2004, 126, 13634−13635. (99) Dahms, K.; Senge, M. O.; Bakar, M. B. Exploration of mesoSubstituted Formylporphyrins and Their Grignard and Wittig Reactions. Eur. J. Org. Chem. 2007, 2007, 3833−3848.

(100) Takanami, T.; Wakita, A.; Sawaizumi, A.; Iso, K.; Onodera, H.; Suda, K. One-Pot Synthesis of meso-Formylporphyrins by SNAr Reaction of 5,15-Disubstituted Porphyrins with (2Pyridyldimethylsilyl)methyllithium. Org. Lett. 2008, 10, 685−687. (101) Takanami, T.; Matsumoto, J.; Kumagai, Y.; Sawaizumi, A.; Suda, K. A Facile One-Pot Preparation of meso-Hydroxymethylporphyrins via a Sequential SNAr Reaction with (2-Pyridyldimethylsilyl)methyllithium followed by Hydrolysis and Aerobic Oxidation. Tetrahedron Lett. 2009, 50, 68−70. (102) Takanami, T.; Hayashi, S.; Iso, K.; Matsumoto, J.; Hino, F. An Efficient One-Pot Protocol for Asymmetric Bifunctionalization of 5,15Disubstituted Porphyrins: Direct Access to meso Activated AlkenylSubstituted meso-Formylporphyrins. Tetrahedron Lett. 2011, 52, 5345− 5348. DA

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 191. Preparation of a Porphyrin−C60 Rotaxane

phyrins by Displacement of a Nitro Group. J. Chem. Soc., Perkin Trans. 1 1997, 3087−3096. (111) Crossley, M. J.; King, L. G.; Pyke, S. M.; Tansey, C. W. Reaction of 5-Nitro-Octaethylporphyrins with Nucleophiles. J. Porphyrins Phthalocyanines 2002, 06, 685−694. (112) Yamashita, K.; Kataoka, K.; Asano, M. S.; Sugiura, K. CatalystFree Aromatic Nucleophilic Substitution of meso-Bromoporphyrins with Azide Anion: Efficient Synthesis and Structural Analyses of mesoAzidoporphyrins. Org. Lett. 2012, 14, 190−193. (113) Chen, Q.; Zhu, Y.-Z.; Fan, Q.-J.; Zhang, S.-C.; Zheng, J.-Y. Simple and Catalyst-Free Synthesis of meso-O-, -S-, and -C-Substituted Porphyrins. Org. Lett. 2014, 16, 1590−1593. (114) Atefi, F.; Arnold, D. P. Porphyrins with Metal, Metalloid or Phosphorus Atoms Directly Bonded to the Carbon Periphery. J. Porphyrins Phthalocyanines 2008, 12, 801−831. (115) Suijkerbuijk, B. M. J. M.; Klein Gebbink, R. J. M. Merging Porphyrins with Organometallics: Synthesis and Applications. Angew. Chem., Int. Ed. 2008, 47, 7396−7421. (116) Richeter, S.; Jeandon, C.; Gisselbrecht, J. P.; Ruppert, R. Peripherally Metalated Porphyrin Derivatives: Synthetic Approaches and Properties. In Handbook of Porphyrin Science; Kadish, K. M., Smith, K. M., Guilard, R., Eds.; World Scientific Publishing Company: Singapore, 2010; Vol. 3, pp 429−483. (117) Mironov, A. Transition Metal Complexes of Porphyrins and Porphyrinoids. Handbook of Porphyrin Science; World Scientific Publishing Company: Singapore, 2012; Vol. 18, pp 303−413. (118) Smith, K. M.; Langry, K. C. Mercuration Reactions of Porphyrins: New Efficient Syntheses of Harderoporphyrin and Isoharderoporphyrin. J. Chem. Soc., Chem. Commun. 1980, 217−218.

(103) Agarwal, N.; Hung, C.-H.; Ravikanth, M. Thiaporphyrins with One, Two and Four Unsubstituted meso-Carbons: Synthesis and Functionalization. Eur. J. Org. Chem. 2003, 2003, 3730−3734. (104) Yamaji, A.; Hiroto, S.; Shin, J.-Y.; Shinokubo, H. Carbolithiation of meso-Aryl-Substituted 5,15-Diazaporphyrin Selectively Provides 3-Alkylated Diazachlorins. Chem. Commun. 2013, 49, 5064−5055. (105) Senge, M. O.; Feng, X. Synthesis of Directly meso−meso Linked Bisporphyrins Using Organolithium Reagents. Tetrahedron Lett. 1999, 40, 4165−4168. (106) Crossley, M. J.; Harding, M. M.; Tansey, C. W. A Convenient Synthesis of 2-Alkyl-5,10,15,20-Tetraphenylporphyrins: Reaction of Metallo-2-Nitro-5,10,15,20-Tetrapheynlporphyrins with Grignard and Organolithium Reagents. J. Org. Chem. 1994, 59, 4433−4437. (107) Catalano, M. M.; Crossley, M. J.; King, L. G. Efficient Synthesis of 2-Oxy-5,10,15,20-Tetraphenylporphyrins from a Nitroporphyrin by a Novel Multi-Step Cine-Substitution Sequence. J. Chem. Soc., Chem. Commun. 1984, 1537−1538. (108) Crossley, M. J.; King, L. G.; Pyke, S. M. A New and Highly Efficient Synthesis of Hydroxyporphyrins. Tetrahedron 1987, 43, 4569−4577. (109) Crossley, M. J.; King, L. G. Reaction of Metallo-2-Nitro5,10,15,20-Tetrapheynlporphyrins with Oxyanions. TemperatureDependent Competition between Nucleophilic Addition and SingleElectron Transfer Processes. J. Chem. Soc., Perkin Trans. 1 1996, 1251− 1260. (110) Crossley, M. J.; King, L. G.; Simpson, J. L. Solvent-Dependent Ambient Nucleophilicity of Phenoxide Ion Towards Nitroporphyrins: Synthesis of 2-Hydroxyaryl- and 2-Aryloxy-5,10,15,20-TetraphenylporDB

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 192. Preparation of Porphyrin-Functionalized SWNTs

Scheme 193. Dimerization of meso-Triazolylporphyrin Zn(II) Complexes

411 Porphyrin (Dehydrocoproporphyrin). J. Org. Chem. 1983, 48, 500−506. (121) Smith, K. M.; Langry, K. C.; Minnetian, O. M. Peripheral Mercuration of Metalloporphyrins: Novel Syntheses of Deoxophylloerythroetioporphyrin and Deoxophylloerythrin Methyl Ester. J. Org. Chem. 1984, 49, 4602−4609.

(119) Smith, K. M.; Langry, K. C. Protiodeacetylation of Porphyrins and Pyrroles: A New Partial Synthesis of Dehydrocoproporphyrin (S411-Porphyrin). J. Chem. Soc., Chem. Commun. 1981, 283−284. (120) Smith, K. M.; Langry, K. C. Electrophilic Mercuration Reactions of Derivatives of Deuteroporphyrin IX: New Syntheses of Coproporphyrin III, Harderoporphyrin, Isoharderoporphyrin, and SDC

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 194. Preparation of a Triazolylphenylporphyrin through a Click Reaction

Scheme 195. Preparation of a Self-Assembled Triazolylchlorin

(128) Yoshida, K.; Yamaguchi, S.; Osuka, A.; Shinokubo, H. Pt(II)and Pt(IV)-Porphyrin Pincer Complexes: Synthesis, Structures, and Reactivity. Organometallics 2010, 29, 3997−4000. (129) Yamamoto, J.; Shimizu, T.; Yamaguchi, S.; Aratani, N.; Shinokubo, H.; Osuka, A. Synthesis of a Diimidazolylporphyrin Pincer Palladium Complex. J. Porphyrins Phthalocyanines 2011, 15, 534−538. (130) Yoshida, K.; Nakashima, T.; Yamaguchi, S.; Osuka, A.; Shinokubo, H. Peripherally Cyclometalated Iridium Complexes of Dipyridylporphyrin. Dalton Trans. 2011, 40, 8773−8775. (131) Anabuki, S.; Shinokubo, H.; Aratani, N.; Osuka, A. A mesoSpiro[Cyclopentadiene-Isoporphyrin] from a Phenylethynyl Porphyrin Platinum(II) Pincer Complex. Angew. Chem., Int. Ed. 2012, 51, 3174− 3177. (132) Song, J.; Aratani, N.; Heo, J. H.; Kim, D.; Shinokubo, H.; Osuka, A. Directly Pd(II)-Bridged Porphyrin Belts with Remarkable Curvature. J. Am. Chem. Soc. 2010, 132, 11868−11869. (133) Yamaguchi, S.; Katoh, T.; Shinokubo, H.; Osuka, A. Pt(II)- and Pt(IV)-Bridged Cofacial Diporphyrins via Carbon-Transition Metal σBonds. J. Am. Chem. Soc. 2008, 130, 14440−14441. (134) Yamaguchi, S.; Shinokubo, H.; Osuka, A. Double Cleavage of sp2 C−H and sp3 C−H bonds on One Metal Center: DMF-Appended Cyclometalated Pt(II)- and Pt(IV)-Porphyrins. Inorg. Chem. 2009, 48, 795−797.

(122) Minnetian, O. M.; Morris, I. K.; Snow, K. M.; Smith, K. M. New Syntheses and Reactions of Some Halogenated Porphyrins. J. Org. Chem. 1989, 54, 5567−5574. (123) Smith, K. M.; Miura, M.; Morris, I. K. Syntheses of Derivatives of Protoporphyrin IX Bearing Deuteriated Methyls on the Propionate (C and D) Rings. J. Org. Chem. 1986, 51, 4660−4667. (124) Morris, I. K.; Snow, K. M.; Smith, N. W.; Smith, K. M. Syntheses of Novel Substituted Porphyrins by the Mercuration and Palladium/Olefin Methodology. J. Org. Chem. 1990, 55, 1231−1236. (125) Buchler, J. W.; Herget, G. Metal Complexes with Tetrapyrrole Ligands, XLV The Mercuration of Nickel(II), Palladium(II), and Platinum(II) Porphyrins. Z. Naturforsch., B: J. Chem. Sci. 1987, 42, 1003−1008. (126) Sugiura, K.; Kato, A.; Iwasaki, K.; Miyasaka, H.; Yamashita, M.; Hino, S.; Arnold, D. P. Unusual Regioselective Mercuration of Metalloporphyrins and Its Potential Applications. Chem. Commun. 2007, 2046−2047. (127) Yamaguchi, S.; Katoh, T.; Shinokubo, H.; Osuka, A. Porphyrin Pincer Complexes: Peripherally Cyclometalated Porphyrins and Their Catalytic Activities Controlled by Central Metals. J. Am. Chem. Soc. 2007, 129, 6392−6393. DD

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 196. Preparation of a “Capped” Porphyrin through a Click Reaction

Scheme 197. Preparation of a Cytochrome c Oxidase Model

(135) Matano, Y.; Matsumoto, K.; Nakao, Y.; Uno, H.; Sakaki, H.; Imahori, H. Regioselective β-Metalation of meso-Phosphanylporphyrins. Structure and Optical Properties of Porphyrin Dimers Linked by Peripherally Fused Phosphametallacycles. J. Am. Chem. Soc. 2008, 130, 4588−4589. (136) Matano, Y.; Matsumoto, K.; Hayashi, H.; Nakao, Y.; Kumpulainen, T.; Chukharev, V.; Tkachenko, N. V.; Lemmetyinen, H.; Shimizu, S.; Kobayashi, N.; et al. Effects of Carbon−Metal− Carbon Linkages on the Optical, Photophysical, and Electrochemical Properties of Phosphametallacycle-Linked Coplanar Porphyrin Dimers. J. Am. Chem. Soc. 2012, 134, 1825−1839. (137) Fujimoto, K.; Yoneda, T.; Yorimitsu, H.; Osuka, A. Synthesis and Catalytic Activities of Porphyrin-Based PCP Pincer Complexes. Angew. Chem., Int. Ed. 2014, 53, 1127−1130. (138) Shinokubo, H. Transition Metal Catalyzed Borylation of Functional π-Systems. Proc. Jpn. Acad., Ser. B 2014, 90, 1−11. (139) Boller, T. M.; Murphy, J. M.; Hapke, M.; Ishiyama, T.; Miyaura, N.; Hartwig, J. F. Mechanism of the Mild Functionalization

of Arenes by Diboron Reagents Catalyzed by Iridium Complexes. Intermediacy and Chemistry of Bipyridine-Ligated Iridium Tris Boryl Complexes. J. Am. Chem. Soc. 2005, 127, 14263−14278. (140) Hata, H.; Shinokubo, H.; Osuka, A. Highly Regioselective IrCatalyzed β-Borylation of Porphyrins via C−H Bond Activation and Construction of β−β-Linked Diporphyrin. J. Am. Chem. Soc. 2005, 127, 8264−8265. (141) Hata, H.; Yamaguchi, S.; Mori, G.; Nakazono, S.; Katoh, T.; Takatsu, K.; Hiroto, S.; Shinokubo, H.; Osuka, A. Regioselective Borylation of Porphyrins by C−H Bond Activation under Iridium Catalysis to Afford Useful Building Blocks for Porphyrin Assemblies. Chem. - Asian J. 2007, 2, 849−859. (142) Hisaki, I.; Hiroto, S.; Kim, K. S.; Noh, S. B.; Kim, D.; Shinokubo, H.; Osuka, A. Synthesis of Doubly β-to-β 1,3-ButadiyneBridged Diporphyrins: Enforced Planar Structures and Large Twophoton Absorption Cross Sections. Angew. Chem., Int. Ed. 2007, 46, 5125−5128. DE

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 198. Preparation of a Porphyrin−Psoralen Conjugate through a Click Reaction

Scheme 199. Preparation of Apoly(ethylene oxide)-Modified Porphyrin through a Click Reaction

(143) Fujimoto, K.; Yorimitsu, H.; Osuka, A. Facile Preparation of βHaloporphyrins as Useful Precursors of β-Substituted Porphyrins. Org. Lett. 2014, 16, 972−975. (144) Hiroto, S.; Hisaki, I.; Shinokubo, H.; Osuka, A. Synthesis of Corrole Derivatives through Regioselective Ir-Catalyzed Direct Borylation. Angew. Chem., Int. Ed. 2005, 44, 6763−6766. (145) Kitano, M.; Okuda, Y.; Tsurumaki, E.; Tanaka, T.; Yorimitsu, H.; Osuka, A. β,β-Diborylated Subporphyrinato Boron(III) Complexes as Useful Synthetic Precursors. Angew. Chem., Int. Ed. 2015, 54, 9275− 9279. (146) Mori, G.; Shinokubo, H.; Osuka, A. Highly Selective IrCatalyzed Direct Sixfold Borylation of Peripheral Aromatic Substituents on Hexakisaryl-Substituted [28]Hexaphyrin(1.1.1.1.1.1). Tetrahedron Lett. 2008, 49, 2170−2172. (147) Yorimitsu, H.; Osuka, A. Organometallic Approaches for Direct Modification of Peripheral C−H Bonds in Porphyrin Cores. Asian J. Org. Chem. 2013, 2, 356−373. (148) Kawamata, Y.; Tokuji, S.; Yorimitsu, H.; Osuka, A. PalladiumCatalyzed β-Selective Direct Arylation of Porphyrins. Angew. Chem., Int. Ed. 2011, 50, 8867−8870.

(149) Lafrance, M.; Fagnou, K. Palladium-Catalyzed Benzene Arylation: Incorporation of Catalytic Pivalic Acid as a Proton Shuttle and a Key Element in Catalyst Design. J. Am. Chem. Soc. 2006, 128, 16496−16497. (150) Yamamoto, Y.; Tokuji, S.; Tanaka, T.; Yorimitsu, H.; Osuka, A. Direct Arylation of Porphyrins with π-Extended Aryl Bromides under Ligand-Free Fagnou−Hartwig Conditions. Asian J. Org. Chem. 2013, 2, 320−324. (151) Tokuji, S.; Awane, H.; Yorimitsu, H.; Osuka, A. Direct Arylation of meso-Formyl Porphyrin. Chem. - Eur. J. 2013, 19, 64−68. (152) Boyle, R. W.; Fox, S. First Examples of Intramolecular Pd(0) Catalysed Couplings on ortho-Iodinated meso-Phenyl Porphyrins. Chem. Commun. 2004, 1322−1323. (153) Fox, S.; Boyle, R. W. Synthetic Routes to Porphyrins Bearing Fused Rings. Tetrahedron 2006, 62, 10039−10054. (154) Cammidge, A. N.; Scaife, P. J.; Berber, G.; Hughes, D. L. Cofacial Porphyrin−Ferrocene Dyads and a New Class of Conjugated Porphyrin. Org. Lett. 2005, 7, 3413−3416. (155) Mitsushige, Y.; Yamaguchi, S.; Lee, B. S.; Sung, Y. M.; Kuhri, S.; Schierl, C. A.; Guldi, D. M.; Kim, D.; Matsuo, Y. Synthesis of DF

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 200. Preparation of a Polymer-Supported Iron Porphyrin Catalyst through a Click Reaction

Thieno-Bridged Porphyrins: Changing the Antiaromatic Contribution by the Direction of the Thiophene Ring. J. Am. Chem. Soc. 2012, 134, 16540−16543. (156) Sahoo, A. K.; Mori, S.; Shinokubo, H.; Osuka, A. Facile Peripheral Functionalization of Porphyrins by Pd-Catalyzed [3 + 2] Annulation with Alkynes. Angew. Chem., Int. Ed. 2006, 45, 7972−7975. (157) Tokuji, S.; Takahashi, Y.; Shinmori, H.; Shinokubo, H.; Osuka, A. Synthesis of a Pyridine-Fused Porphyrinoid: Oxopyridochlorin. Chem. Commun. 2009, 1028−1030. (158) Fukui, N.; Arai, S.; Shinokubo, H.; Osuka, A. PalladiumCatalyzed [3 + 2] Annulation of meso-Bromoporphyrin with Silylacetylenes and Desilylation of 8a-Silyl-7,8-dehydropurpurin. Heterocycles 2015, 90, 252−259. (159) Fukui, N.; Yorimitsu, H.; Lim, J. M.; Kim, D.; Osuka, A. Synthesis of 7,8-Dehydropurpurin Dimers and Their Conversion into Conformationally Constrained β-to-β Vinylene-Bridged Porphyrin Dimers. Angew. Chem., Int. Ed. 2014, 53, 4395−4398. (160) Mizumura, M.; Shinokubo, H.; Osuka, A. Synthesis of Chiral Porphyrins through Pd-Catalyzed [3 + 2] Annulation and Heterochiral Self-Assembly. Angew. Chem., Int. Ed. 2008, 47, 5378−5381. (161) Mizumura, M.; Shinokubo, H.; Osuka, A. Synthesis of Norbornane-Bridged Diporphyrins via Palladium-Catalyzed [3 + 2] Annulation Strategy. Synthesis 2009, 2009, 59−61. (162) Tokuji, S.; Yurino, T.; Aratani, N.; Shinokubo, H.; Osuka, A. Palladium-Catalyzed Dimerization of meso-Bromoporphyrins: Highly Regioselective meso−β Coupling through Unprecedented Remote C− H Bond Cleavage. Chem. - Eur. J. 2009, 15, 12208−12211. (163) Osawa, K.; Aratani, N.; Osuka, A. Facile Synthesis and Photophysical Properties of 1,2-Phenylene-Bridged Porphyrin Dimers. Tetrahedron Lett. 2009, 50, 3333−3337.

(164) Kawakami, T.; Murakami, K.; Itami, K. Catalytic C−H Imidation of Aromatic Cores of Functional Molecules: LigandAccelerated Cu Catalysis and Application to Materials- and BiologyOriented Aromatics. J. Am. Chem. Soc. 2015, 137, 2460−2463. (165) Tomooka, K.; Ito, M. Lithium in Organic Synthesis. In Main Group Metals in Organic Synthesis; Yamamoto, H., Oshima, K., Eds.; Wiley-VCH: Weinheim, 2004; pp 1−34. (166) Inoue, A.; Oshima, K. Magnesium in Organic Synthesis. In Main Group Metals in Organic Synthesis; Yamamoto, H., Oshima, K., Eds.; Wiley-VCH: Weinheim, 2004; pp 51−154. (167) Fujimoto, K.; Yorimitsu, H.; Osuka, A. Efficient Synthesis and Versatile Reactivity of Porphyrinyl Grignard Reagents. Eur. J. Org. Chem. 2014, 2014, 4327−4334. (168) Fujimoto, K.; Yorimitsu, H.; Osuka, A. Porphyrinylboranes Synthesized via Porphyrinyllithiums. Chem. - Eur. J. 2015, 21, 11311− 11314. (169) Kato, K.; Fujimoto, K.; Yorimitsu, H.; Osuka, A. Peripherally Silylated Porphyrins. Chem. - Eur. J. 2015, 21, 13522−13525. (170) Kato, K.; Kim, W.; Kim, D.; Yorimitsu, H.; Osuka, A. Porphyrin Analogues of a Trityl Cation and Anion. Chem. - Eur. J. 2016, 22, 7041−7045. (171) Arnold, D. P.; Sakata, Y.; Sugiura, K.; Worthington, E. I. mesoη1-Metalloporphyrins: Preparation of Palladio- and Platinioporphyrins and the Crystal Structure of 5-[Bromo-1,2-Bis(diphenylphosphino)ethanepalladio(II)-10,20-Diphenylporphyrin. Chem. Commun. 1998, 2331−2332. (172) Arnold, D. P.; Healy, P. C.; Hodgson, M. J.; Williams, M. L. Peripherally-Metalated Porphyrins: Synthesis and Spectra of meso-η1Palladio- and Platinioporphyrins and the Crystal Structures of cis{PtBr[10,20-Diphenylporphyrinatonickel(II)-5-yl](PPh 3)2} and transDG

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 201. Preparation of Dendritic Porphyrins through Click Reactions

lporphyrins via Organopalladium Porphyrins. J. Porphyrins Phthalocyanines 2004, 08, 1222−1227. (179) Atefi, F.; Locos, O. B.; Senge, M. O.; Arnold, D. P. meso-Iodoand meso-Iodovinylporphyrins via Organopalladium Porphyrins and the Crystal Structure of 5-Iodo-10,20-Diphenylporphyrin. J. Porphyrins Phthalocyanines 2006, 10, 176−185. (180) Hartnell, R. D.; Yoneda, T.; Mori, H.; Osuka, A.; Arnold, D. P. The Marriage of Peripherally Metallated and Directly Linked Porphyrins: Bromidobis(phosphine)platinum(II) as a Cation-Stabilizing Substituent on Directly Linked and Fused Triply Linked Diporphyrins. Chem. - Asian J. 2013, 8, 2670−2679. (181) Ishiyama, T.; Itoh, Y.; Kitano, T.; Miyaura, N. Synthesis of Arylboronates via the Palladium(0)-Catalyzed Cross-Coupling Reaction of Tetra(Alkoxo)Diborons with Aryl Triflates. Tetrahedron Lett. 1997, 38, 3447−3450. (182) Ishiyama, T.; Ishida, K.; Miyaura, N. Synthesis of Pinacol Arylboronates via Cross-Coupling Reaction of Bis(pinacolato)diboron with Chloroarenes Catalyzed by Palladium(0)−Tricyclohexylphosphine Complexes. Tetrahedron 2001, 57, 9813−9816. (183) Billingsley, K. L.; Barder, T. E.; Buchwald, S. L. PalladiumCatalyzed Borylation of Aryl Chlorides: Scope, Applications, and Computational Studies. Angew. Chem., Int. Ed. 2007, 46, 5359−5363. (184) Fürstner, A.; Seidel, G. Microwave-Assisted Synthesis of Pinacol Boronates from Aryl Chlorides Catalyzed by a Palladium/ Imidazolium Salt System. Org. Lett. 2002, 4, 541−543.

{PtBr[10,20-Diphenylporphyrin-5-yl](PPh3)2}·0.25CH2Cl2. J. Organomet. Chem. 2000, 607, 41−50. (173) Hartnell, R. D.; Arnold, D. P. Peripherally Metallated Porphyrins: The First Examples of meso-η1-Palladio(II) and -Platinio(II) Complexes with Chelating Diamine Ligands. Eur. J. Inorg. Chem. 2004, 2004, 1262−1269. (174) Hodgson, M. J.; Healy, P. C.; Williams, M. L.; Arnold, D. P. Peripherally-Metalated Porphyrins: Preparations, Spectroscopic Properties and Structural Studies of trans-[PtBr(MDPP) (PPh3)2] (DPP = Dianion of 5,15-Diphenylporphyrin, M = MnCl, Co, Ni, Zn) and Related meso-η1-Organoplatinum Porphyrins. J. Chem. Soc., Dalton Trans. 2002, 4497−4504. (175) Hodgson, M. J.; Borovkov, V. V.; Inoue, Y.; Arnold, D. P. A New Type of Chiral Porphyrin: Organopalladium Porphyrins with Chiral Chelating Diphosphine Ligands. J. Organomet. Chem. 2006, 691, 2162−2170. (176) Hartnell, R. D.; Edwards, A. J.; Arnold, D. P. PeripherallyMetalated Porphyrins: meso-η1-Porphyrinylplatinum(II) Complexes of 5,15-Diaryl- and 5,10,15-Triarylporphyrins. J. Porphyrins Phthalocyanines 2002, 6, 695−707. (177) Hartnell, R. D.; Arnold, D. P. Peripherally η1-Platinated Organometallic Porphyrins as Building Blocks for Multiporphyrin Arrays. Organometallics 2004, 23, 391−399. (178) Kato, A.; Hartnell, R. D.; Yamashita, M.; Miyasaka, H.; Sugiura, K.; Arnold, D. P. Selective meso-Monobromination of 5,15-DiaryDH

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 202. Preparation of Porphyrin-Functionalized Polysaccharides

Scheme 203. Preparation of Sugar-Appended Porphyrins

(187) Deng, Y.; Chang, C. K.; Nocera, D. G. Facile Synthesis of βDerivatized Porphyrins−Structural Characterization of a β−β-BisPorphyrin. Angew. Chem., Int. Ed. 2000, 39, 1066−1068. (188) Zhang, C.; Suslick, K. S. Syntheses of Boronic-Acid-Appended Metalloporphyrins as Potential Colorimetric Sensors for Sugars and Carbohydrates. J. Porphyrins Phthalocyanines 2005, 09, 659−666. (189) Bringmann, G.; Götz, D. C. G.; Gulder, T. A. M.; Gehrke, T. H.; Bruhn, T.; Kupfer, T.; Radacki, K.; Braunschweig, H.; Heckmann,

(185) Murata, M.; Sambommatsu, T.; Watanabe, S.; Masuda, Y. An Efficient Catalyst System for Palladium-Catalyzed Borylation of Aryl Halides with Pinacolborane. Synlett 2006, 2006, 1867−1870. (186) Hyslop, A. G.; Kellett, M. A.; Iovine, P. M.; Therien, M. J. Suzuki Porphyrins: New Synthons for the Fabrication of PorphyrinContaining Supramolecular Assemblies. J. Am. Chem. Soc. 1998, 120, 12676−12677. DI

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 204. Preparation of Sugar-Appended Porphyrins

Scheme 205. Preparation of Porphyrin-Appended Cotton Fabric

A.; Lambert, C. Axially Chiral β,β′-Bisporphyrins: Synthesis and Configurational Stability Tuned by the Central Metals. J. Am. Chem. Soc. 2008, 130, 17812−17825. (190) Fuhrhop, J. H. Irreversible Reactions on the Porphyrin Periphery (Excluding Oxidations, Reductions, and Photochemical Reactions). In The Porphyrins; Dolpin, D., Ed.; Academic Press: San Diego, 1978; Vol. 2, pp 131−159. (191) Nakano, A.; Shimidzu, H.; Osuka, A. Facile Regioselective meso-Iodination of Porphyrins. Tetrahedron Lett. 1998, 39, 9489− 9492. (192) Vaz, B.; Alvarez, R.; Nieto, M.; Paniello, A. I.; de Lera, A. R. Suzuki Cross-Coupling of meso-Dibromoporphyrins for the Synthesis of Functionalized A2B2 Porphyrins. Tetrahedron Lett. 2001, 42, 7409− 7412. (193) Shi, B.; Boyle, R. W. Synthesis of Unsymmetrically Substituted meso-Phenylporphyrins by Suzuki Cross Coupling Reactions. J. Chem. Soc., Perkin Trans. 1 2002, 1397−1400.

(194) Aratani, N.; Osuka, A. Synthesis of meso-meso Linked Hybrid Porphyrin Arrays by Pd-Catalyzed Cross-Coupling Reaction. Org. Lett. 2001, 3, 4213−4216. (195) Maeda, C.; Kim, P.; Cho, S.; Park, J. K.; Lim, J. M.; Kim, D.; Vura-Weis, J.; Wasielewski, M. R.; Shinokubo, H.; Osuka, A. Large Porphyrin Squares from the Self-Assembly of meso-Triazole-Appended L-Shaped meso−meso-Linked Zn(II)−Triporphyrins: Synthesis and Efficient Energy Transfer. Chem. - Eur. J. 2010, 16, 5052−5061. (196) Diev, V. V.; Schlenker, C. W.; Hanson, K.; Zhong, Q.; Zimmerman, J. D.; Forrest, S. R.; Thompson, M. E. Porphyrins Fused with Unactivated Polycyclic Aromatic Hydrocarbons. J. Org. Chem. 2012, 77, 143−159. (197) Bakar, M. A.; Sergeeva, N. N.; Juillard, T.; Senge, M. O. Synthesis of Ferrocenyl Porphyrins via Suzuki Coupling and Their Photophysical Properties. Organometallics 2011, 30, 3225−3228. DJ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 206. Preparation of Porphyrins Functionalized with β-Cyclodextrins

Scheme 207. Preparation of Porphyrin−Peptide Conjugates

(198) Davis, N. K. S.; Thompson, A. L.; Anderson, H. L. Bis-

(200) Locos, O. B.; Dahms, K.; Senge, M. O. Allenylporphyrins: A

Anthracene Fused Porphyrins: Synthesis, Crystal Structure, and Near-

New Motif on the Porphyrin Periphery. Tetrahedron Lett. 2009, 50,

IR Absorption. Org. Lett. 2010, 12, 2124−2127. (199) Davis, M.; Senge, M. O.; Locos, O. B. Anthracenylporphyrins.

2566−2569. (201) Plunkett, S.; Dahms, K.; Senge, M. O. Synthesis and Reactivity

Z. Naturforsch., B: J. Chem. Sci. 2010, 65, 1472−1484.

of Allenylporphyrins. Eur. J. Org. Chem. 2013, 2013, 1566−1579.

DK

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 208. Preparation of DNA−Porphyrin Conjugates

Scheme 209. Catalyst immobilization on surfaces through click chemistry

(202) Horn, S.; Cundell, B.; Senge, M. O. Exploration of the Reaction of Potassium Organotrifluoroborates with Porphyrins. Tetrahedron Lett. 2009, 50, 2562−2565. (203) Senge, M. O. Stirring The Porphyrin Alphabet Soup Functionalization Reactions for Porphyrins. Chem. Commun. 2011, 47, 1943−1960. (204) Chan, K. S.; Zhou, X.; Luo, B.; Mak, T. C. W. Synthesis of βAryl Substituted Porphyrins by Palladium-catalysed Cross-coupling Reactions. J. Chem. Soc., Chem. Commun. 1994, 271−272. (205) Muzzi, C. M.; Medforth, C. J.; Voss, L.; Cancilla, M.; Lebrilla, C.; Ma, J.-G.; Shelnutt, J. A.; Smith, K. M. Novel Dodecaarylporphyrins: Synthesis and Dynamic Properties. Tetrahedron Lett. 1999, 40, 6159−6162.

(206) Kumar, R.; Sankar, M. Synthesis, Spectral, and Electrochemical Studies of Electronically Tunable β-Substituted Porphyrins with Mixed Substituent Pattern. Inorg. Chem. 2014, 53, 12706−12719. (207) Prakash, K.; Kumar, R.; Sankar, M. Mono- and Tri-βSubstituted Unsymmetrical Metalloporphyrins: Synthesis, Structural, Spectral and Electrochemical Properties. RSC Adv. 2015, 5, 66824− 66832. (208) Ikeda, T.; Aratani, N.; Easwaramoorthi, S.; Kim, D.; Osuka, A. meso−β Doubly Linked Zn(II) Porphyrin Trimers: Distinct antiversus-syn Effects on Their Photophysical Properties. Org. Lett. 2009, 11, 3080−3083. DL

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 210. Introduction of Porphyrins onto a Silicon Surface through a Click Reaction

Scheme 211. Functionalization of the Surface of Mesoporous Silica through a CuAAC Reaction

Scheme 212. Preparation of Azaindolylporphyrin through Palladium-Catalyzed [3 + 2] Annulation

(209) Gehrold, A. C.; Bruhn, T.; Bringmann, G. Axial, Helical, and

(210) Kinzel, T.; Zhang, Y.; Buchwald, S. L. A New Palladium Precatalyst Allows for the Fast Suzuki−Miyaura Coupling Reactions of Unstable Polyfluorophenyl and 2-Heteroaryl Boronic Acids. J. Am. Chem. Soc. 2010, 132, 14073−14075.

Planar Chirality in Directly Linked Basket-Handle Porphyrin Arrays. J. Org. Chem. 2016, 81, 1075−1088. DM

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 213. Preparation of Azaindolyl Functions through Copper-Catalyzed Intramolecular Cyclization

Scheme 214. Copper-Catalyzed Conversion of a 1,3-Butadiyne Moiety to a Pyrrole Ring

DN

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 215. Dötz Reaction of an Alkynylporphyrin

Scheme 216. Pauson−Khand Reaction of Alkynylporphyrin

Nanorings through Cross-Coupling. Chem. - Eur. J. 2010, 16, 3009− 3012. (219) Hao, Q.; Zeng, Y.; Yu, T.; Chen, J.; Yang, G.; Li, Y. Synthesis and Photophysical Properties of Doubly β-to-β Bridged Cyclic Zn(II) Porphyrin Arrays. Chem. - Asian J. 2013, 8, 1015−1022. (220) Rao, Y.; Kim, J. O.; Kim, W.; Zhong, G.; Yin, B.; Zhou, M.; Shinokubo, H.; Aratani, N.; Tanaka, T.; Liu, S.; et al. β-to-β 2,5Pyrrolylene-Linked Cyclic Porphyrin Oligomers. Chem. - Eur. J. 2016, 22, 8801−8804. (221) Huang, W.; Lee, S.-K.; Sung, Y. M.; Peng, F.; Yin, B.; Ma, M.; Chen, B.; Liu, S.; Kirk, S. R.; Kim, D.; et al. Azobenzene-Bridged Porphyrin Nanorings: Syntheses, Structures, and Photophysical Properties. Chem. - Eur. J. 2015, 21, 15328−15338. (222) Ikeda, S.; Aratani, N.; Osuka, A. A Doubly 2,6-PyridyleneBridged Porphyrin−Perylene−Porphyrin Triad. Chem. Commun. 2012, 48, 4317−4319. (223) Song, J.; Aratani, N.; Kim, P.; Kim, D.; Shinokubo, H.; Osuka, A. Porphyrin “Lego Block” Strategy To Construct Directly meso−β Doubly Linked Porphyrin Rings. Angew. Chem., Int. Ed. 2010, 49, 3617−3620. (224) Song, J.; Anabuki, S.; Aratani, N.; Shinokubo, H.; Osuka, A. A Hexameric Porphyrin Triangle Constructed by Suzuki−Miyaura Cross-Coupling Reaction. Chem. Lett. 2011, 40, 902−903. (225) Cai, H.; Fujimoto, K.; Lim, J. M.; Wang, C.; Huang, W.; Rao, Y.; Zhang, S.; Shi, H.; Yin, B.; Chen, B.; et al. Synthesis of Direct β-toβ Linked Porphyrin Arrays with Large Electronic Interactions: Branched and Cyclic Oligomers. Angew. Chem., Int. Ed. 2014, 53, 11088−11091. (226) Rao, Y.; Kim, T.; Park, K. H.; Peng, F.; Liu, L.; Liu, Y.; Wen, B.; Liu, S.; Kirk, S. R.; Wu, L.; et al. π-Extended “Earring” Porphyrins with Multiple Cavities and Near-Infrared Absorption. Angew. Chem., Int. Ed. 2016, 55, 6438−6442. (227) Omomo, S.; Furukawa, K.; Nakano, H.; Matano, Y. Comparison of Electronic Effects of β-Aryl Substituents on Optical and Electrochemical Properties of 5,15-Diazaporphyrin π-Systems. J. Porphyrins Phthalocyanines 2015, 19, 775−785. (228) Abou-Chahine, F.; Fujii, D.; Imahori, H.; Nakano, H.; Tkachenko, N. V.; Matano, Y.; Lemmetyinen, H. Synthesis and Photophysical Properties of Two Diazaporphyrin-Porphyrin Hetero Dimers in Polar and Nonpolar Solutions. J. Phys. Chem. B 2015, 119, 7328−7337.

Figure 3. Carbazole- and indole-containing porphyrin analogues. (211) Ota, K.; Tanaka, T.; Osuka, A. meso−β Dibenzo[a,g]corannulene-Fused Porphyrins. Org. Lett. 2014, 16, 2974−2977. (212) Fukui, N.; Yorimitsu, H.; Osuka, A. meso, β-Oligohaloporphyrins as Useful Synthetic Intermediates of Diphenylamine-Fused Porphyrin and meso-to-meso β-to-β Doubly Butadiyne-Bridged Diporphyrin. Angew. Chem., Int. Ed. 2015, 54, 6311−6314. (213) Ali, H.; Lier, J. E. An Efficient Method for the Synthesis of C− C Connected Phthalocyanine−Porphyrin Oligomers. Tetrahedron Lett. 2009, 50, 1113−1116. (214) Tokuji, S.; Yorimitsu, H.; Osuka, A. Preferential Formation of Cyclic Trimers by Palladium-Catalyzed Oxidative Coupling Reactions of 2,18-Diethynylporphyrins. Angew. Chem., Int. Ed. 2012, 51, 12357− 12361. (215) Song, J.; Aratani, N.; Shinokubo, H.; Osuka, A. A Porphyrin Nanobarrel That Encapsulates C60. J. Am. Chem. Soc. 2010, 132, 16356−16357. (216) Song, J.; Jang, S. Y.; Yamaguchi, S.; Sankar, J.; Hiroto, S.; Aratani, N.; Shin, J.-Y.; Easwaramoorthi, S.; Kim, K. S.; Kim, D.; et al. 2,5-Thienylene-Bridged Triangular and Linear Porphyrin Trimers. Angew. Chem., Int. Ed. 2008, 47, 6004−6007. (217) Song, J.; Aratani, N.; Shinokubo, H.; Osuka, A. A β-to-β 2,5Thienylene-Bridged Cyclic Porphyrin Tetramer: Its Rational Synthesis and 1:2 Binding Mode with C60. Chem. Sci. 2011, 2, 748−751. (218) Song, J.; Kim, P.; Aratani, N.; Kim, D.; Shinokubo, H.; Osuka, A. Strategic Synthesis of 2,6-Pyridylene-Bridged β-to-β Porphyrin DO

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 217. Preparation of Carbazole-Based Porphyrinoids

(229) Yoshida, K.; Copley, G.; Mori, H.; Osuka, A. Probing the Rotational Dynamics of meso-(2-Substituted)aryl Substituents in A2BType Subporphyrins. Chem. - Eur. J. 2014, 20, 10065−10072. (230) Boyle, R. W.; Johnson, C. K.; Dolphin, D. Iodination and Heck Alkynylation of 5,15-Diphenylporphyrin. A Convenient Entry to Asymmetrically meso-Substituted Porphyrins. J. Chem. Soc., Chem. Commun. 1995, 527−528. (231) Chen, Y.-J.; Lee, G.-H.; Peng, S.-M.; Yeh, C.-Y. Unexpected Formation of Porphyrinic Enyne under Sonogashira Conditions. Tetrahedron Lett. 2005, 46, 1541−1544. (232) LeCours, S. M.; DiMagno, S. G.; Therien, M. J. Exceptional Electronic Modulation of Porphyrins through meso-Arylethynyl Groups. Electronic Spectroscopy, Electronic Structure, and Electrochemistry of [5,15-Bis[(aryl)ethynyl]-10,20-diphenylporphinato]zinc(II) Complexes. X-ray Crystal Structures of [5,15-Bis[(4″fluorophenyl)ethynyl]-10,20-diphenylporphinato]zinc(II) and 5,15Bis[(4″-methoxyphenyl)ethynyl]-10,20-diphenylporphyrin. J. Am. Chem. Soc. 1996, 118, 11854−11864. (233) Screen, T. E. O.; Blake, I. M.; Rees, L. H.; Clegg, W.; Borwick, S. J.; Anderson, L. A. Making Conjugated Connections to Porphyrins: A Comparison of Alkyne, Alkene, Imine and Azo Links. J. Chem. Soc., Perkin Trans. 1 2002, 320−329. (234) Shanmugathasan, S.; Johnson, C. K.; Edwards, C.; Matthews, E. K.; Dolphin, D.; Boyle, R. W. Regioselective Halogenation and Palladium-Catalysed Couplings on 5,15-Diphenylporphyrin. J. Porphyrins Phthalocyanines 2000, 4, 228−232. (235) Ali, H.; van Lier, J. E. Synthesis of β-Substituted Porphyrins Using Palladium Catalysed Reactions. Tetrahedron 1994, 50, 11933− 11944. (236) Ishida, M.; Hwang, D.; Koo, Y. B.; Sung, J.; Kim, D. Y.; Sessler, J. L.; Kim, D. β-(Ethynylbenzoic Acid)-Substituted Push−Pull Porphyrins: DSSC Dyes Prepared by a Direct Palladium-Catalyzed Alkynylation Reaction. Chem. Commun. 2013, 49, 9164−9166.

(237) Matano, Y.; Fujii, D.; Shibano, T.; Furukawa, K.; Higashino, T.; Nakano, H.; Imahori, H. Covalently Linked 5,15-Diazaporphyrin Dimers: Promising Scaffolds for a Highly Conjugated Azaporphyrin π System. Chem. - Eur. J. 2014, 20, 3342−3349. (238) Kitano, M.; Hayashi, S.; Tanaka, T.; Yorimitsu, H.; Aratani, N.; Osuka, A. Effective meso Fabrications of Subporphyrins. Angew. Chem., Int. Ed. 2012, 51, 5593−5597. (239) Toganoh, M.; Kimura, T.; Furuta, H. Endocyclic Extension of Porphyrin π-System by Interior Functionalization of N-Confused Porphyrins. Chem. - Eur. J. 2008, 14, 10585−10594. (240) DiMagno, S. G.; Lin, S.-Y.; Therien, M. J. Facile Elaboration of Porphyrins via Metal-Mediated Cross-Coupling. J. Org. Chem. 1993, 58, 5983−5993. (241) Lin, V. S.-Y.; Dimagno, S. G.; Therien, M. J. Highly Conjugated, Acetylenyl Bridged Porphyrins: New Models for LightHarvesting Antenna Systems. Science 1994, 264, 1105−1111. (242) Hayashi, S.; Yotsukura, M.; Noji, M.; Takanami, T. Bis(Zinc Porphyrin) as a CD-Sensitive Bidentate Host Molecule: Direct Determination of Absolute Configuration of Mono-Alcohols. Chem. Commun. 2015, 51, 11068−11071. (243) Sugita, N.; Hayashi, S.; Ishii, S.; Takanami, T. PalladiumCatalyzed Polyfluorophenylation of Porphyrins with Bis(polyfluorophenyl)zinc Reagents. Catalysts 2013, 3, 839−852. (244) Sugita, N.; Hayashi, S.; Hino, F.; Takanami, T. Palladiumcatalyzed Kumada Coupling Reaction of Bromoporphyrins with Silylmethyl Grignard Reagents: Preparation of Silylmethyl-substituted Porphyrins as a Multipurpose Synthon for Fabrication of Porphyrin Systems. J. Org. Chem. 2012, 77, 10488−10497. (245) Takanami, T.; Yotsukura, M.; Inoue, W.; Inoue, N.; Hino, F.; Suda, K. A Facil and Efficient Synthesis of Mono- and Bisfunctionalized meso-Substituted Porphyrins via Palladium-Catalyzed Negishi Cross-Coupling. Heterocycles 2008, 76, 439−453. DP

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 218. Preparation of Butadiyne-Linked Carbazoles and Carbazole-Based Porphyrinoids

and Porphyrin Dimers via Stille Coupling and Their 119Sn NMR and Fluorescence Properties. J. Org. Chem. 2009, 74, 7140−7147. (254) Omomo, S.; Maruyama, Y.; Furukawa, K.; Furuyama, T.; Nakano, H.; Kobayashi, K.; Matano, Y. Optical, Electrochemical, and Magnetic Properties of Pyrrole- and Thiophene-Bridged 5,15Diazaporphyrin Dimers. Chem. - Eur. J. 2015, 21, 2003−2010. (255) Yamaji, A.; Shin, J.-Y.; Miyake, Y.; Shinokubo, H. A 3-Pyridyl5,15-Diazaporphyrin Nickel(II) Complex as a Bidentate Metalloligand for Transition Metals. Angew. Chem., Int. Ed. 2014, 53, 13924−13927. (256) Locos, O. B.; Arnold, D. P. The Heck Reaction for Porphyrin Functionalisation: Synthesis of meso-Alkenyl Monoporphyrins and Palladium-Catalysed Formation of Unprecedented meso−β EtheneLinked Diporphyrins. Org. Biomol. Chem. 2006, 4, 902−916. (257) Risch, N.; Gauler, R.; Keuper, R. Synthesis of Porphyrin Dimers Using a Heck-Type Coupling Reaction with Bisacrylates. Tetrahedron Lett. 1999, 40, 2925−2926. (258) Deshpande, R.; Jiang, L.; Schmidt, G.; Rakovan, J.; Wang, X.; Wheeler, K.; Wang, H. A Concise Approach to the Synthesis of oppDibenzoporphyrins through the Heck Reaction. Org. Lett. 2009, 11, 4251−4253. (259) Jinadasa, R. G. W.; Fang, Y.; Kumar, S.; Osinski, A. J.; Jiang, X.; Ziegler, C. J.; Kadish, K. M.; Wang, H. β-Functionalized Push−Pull opp-Dibenzoporphyrins. J. Org. Chem. 2015, 80, 12076−12087. (260) Takanami, T.; Hayashi, M.; Chijimatsu, H.; Inoue, W.; Suda, K. Palladium-Catalyzed Cyanation of Porphyrins Utilizing Cyanoethylzinc Bromide as an Efficient Cyanide Ion Source. Org. Lett. 2005, 7, 3937−3940. (261) Balaban, M. C.; Eichhöfer, A.; Buth, G.; Hauschild, R.; Szmytkowski, J.; Kalt, H.; Balaban, T. S. Programmed Metal-

(246) Takanami, T. Functionalization of Porphyrins through C−C Bond Formation Reactions with Functional Group-Bearing Organometallic Reagents. Heterocycles 2013, 87, 1659−1689. (247) Kato, K.; Kim, J. O.; Yorimitsu, H.; Kim, D.; Osuka, A. Triphenylsilane-Fused Porphyrins. Chem. - Asian J. 2016, 11, 1738− 1746. (248) Shi, X.; Amin, S. R.; Liebeskind, L. S. 3-Cyclobutenyl-1,2Dione-Substituted Porphyrins. A General and Efficient Entry to Porphyrin-Quinone and Quinone-Porphyrin-Quinone Architectures. J. Org. Chem. 2000, 65, 1650−1664. (249) Odobel, F.; Suzenet, F.; Blart, E.; Quintard, J.-P. An Efficient Synthetic Approach to Highly Conjugated Porphyrin-Based Assemblies Containing a Bipyridine Moiety. Org. Lett. 2000, 2, 131−133. (250) Frampton, M. J.; Akdas, H.; Cowley, A. R.; Rogers, J. E.; Slagle, J. E.; Fleitz, P. A.; Drobizhev, M.; Rebane, A.; Anderson, H. L. Synthesis, Crystal Structure, and Nonlinear Optical Behavior of βUnsubstituted meso−meso E-Vinylene-Linked Porphyrin Dimers. Org. Lett. 2005, 7, 5365−5368. (251) Aihara, H.; Jaquinod, L.; Nurco, D. J.; Smith, K. M. Multicarbocycle Formation Mediated by Arenoporphyrin 1,4Diradicals: Synthesis of Picenoporphyrins. Angew. Chem., Int. Ed. 2001, 40, 3439−3441. (252) Chandra, T.; Kraft, B. J.; Huffman, J. C.; Zaleski, J. M. Synthesis and Structural Characterization of Porphyrinic Enediynes: Geometric and Electronic Effects on Thermal and Photochemical Reactivity. Inorg. Chem. 2003, 42, 5158−5172. (253) Sergeeva, N. N.; Scala, A.; Bakar, M. A.; O’Riordan, G.; O’Brien, J.; Grassi, G.; Senge, M. O. Synthesis of Stannyl Porphyrins DQ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 219. Preparation of Porphyrinoids Containing Both Carbazole and Indole Units

(267) Shen, D.-M.; Liu, C.; Chen, Q.-Y. A General and Efficient Palladium-Catalyzed Intramolecular Cyclization Reaction of βBrominated Porphyrins. J. Org. Chem. 2006, 71, 6508−6511. (268) Ishizuka, T.; Saegusa, Y.; Shiota, Y.; Ohtake, K.; Yoshizawa, K.; Kojima, T. Multiply-Fused PorphyrinsEffects of Extended πConjugation on the Optical and Electrochemical Properties. Chem. Commun. 2013, 49, 5939−5941. (269) Saegusa, Y.; Ishizuka, T.; Komamura, K.; Shimizu, S.; Kotani, H.; Kobayashi, N.; Kojima, T. Ring-Fused Porphyrins: Extension of πConjugation Significantly Affects the Aromaticity and Optical Properties of the Porphyrin π-Systems and the Lewis Acidity of the Central Metal Ions. Phys. Chem. Chem. Phys. 2015, 17, 15001−15011. (270) Kato, K.; Cha, W.; Oh, J.; Furukawa, K.; Yorimitsu, H.; Kim, D.; Osuka, A. Spontaneous Formation of an Air-Stable Radical upon the Direct Fusion of Diphenylmethane to a Triarylporphyrin. Angew. Chem., Int. Ed. 2016, 55, 8711−8714. (271) Tomohiro, Y.; Satake, A.; Kobuke, Y. Synthesis of Bipyridylene-Bridged Bisporphyrin by Nickel-Mediated Coupling Reaction: ON-OFF Control of Cofacial Porphyrin Unit by Reversible Complexation. J. Org. Chem. 2001, 66, 8442−8446.

loporphyrins for Self-Assembly within Light-Harvesting Stacks: (5,15Dicyano-10,20-bis(3,5-di-tert-butylphenyl)porphyrinato)zinc(II) and Its Push−Pull 15-N,N-Dialkylamino-5-cyano Congeners Obtained by a Facile Direct Amination. J. Phys. Chem. B 2008, 112, 5512−5521. (262) Jin, F.; Confalone, P. N. Palladium-Catalyzed Cyanation Reactions of Aryl Chlorides. Tetrahedron Lett. 2000, 41, 3271−3273. (263) Blake, I. M.; Anderson, H. L.; Beljonne, D.; Brédas, J.-L.; Clegg, W. Synthesis, Crystal Structure, and Electronic Structure of a 5,15-Dialkylideneporphyrin: A TCNQ/Porphyrin Hybrid. J. Am. Chem. Soc. 1998, 120, 10764−10765. (264) Blake, I. M.; Rees, L. H.; Claridge, T. D. W.; Anderson, H. L. Synthesis and Crystal Structure of a Cumulenic Quinoidal Porphyrin Dimer with Strong Electronic Absorption in the Infrared. Angew. Chem., Int. Ed. 2000, 39, 1818−1821. (265) Yamashita, K.; Sakamoto, S.; Suzuki, A.; Sugiura, K. Molecular Design and Syntheses of Tetracyano-5,10-porphyrinquinodimethane Showing Stabilized LUMO. Chem. - Asian J. 2016, 11, 1004−1007. (266) Tokuji, S.; Maeda, C.; Yorimitsu, H.; Osuka, A. New Synthetic Strategy for Diporphyrins: Pinacol Coupling−Rearrangement. Chem. Eur. J. 2011, 17, 7154−7157. DR

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 220. Bithiophene-Bridged Cyclic Carbazoles

Scheme 221. Preparation of Cyclic Tetraindoles

(272) Lu, X.-Q.; Guo, Y.; Chen, Q.-Y. Efficient Synthesis of of meso−

(273) Hiroto, S.; Furukawa, K.; Shinokubo, H.; Osuka, A. Synthesis

meso-Linked Diporphyrins by Nickel(0)-Mediated Ullmann Homo-

and Biradicaloid Character of Doubly Linked Corrole Dimers. J. Am.

coupling. Synlett 2011, 2011, 77−80.

Chem. Soc. 2006, 128, 12380−12381. DS

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 222. Preparation of Butadiyne-Bridged Cyclic BODIPYs

Scheme 223. Preparation of Azaporphyrinoids from Bisdipyrrin Metal Complexes

Visible Absorption: Synthesis and Photophysical Studies. Inorg. Chem. 2012, 51, 10548−10556. (279) Schmitz, R. A.; Liddell, P. A.; Kodis, G.; Kenney, M. J.; Brennan, B. J.; Oster, N. V.; Moore, T. A.; Moore, A. L.; Gust, D. Synthesis and Spectroscopic Properties of a Soluble Semiconducting Porphyrin Polymer. Phys. Chem. Chem. Phys. 2014, 16, 17569−17579. (280) Pawlicki, M.; Hurej, K.; Kwiecińska, K.; Szterenberg, L.; LatosGrażyński, L. A Fused meso-Aminoporphyrin: A Switchable Near-IR Chromophore. Chem. Commun. 2015, 51, 11362−11365. (281) Tsurumaki, E.; Osuka, A. Synthesis of Peripherally Nitrated, Aminated, and Arylaminated Subporphyrins. Chem. - Asian J. 2013, 8, 3042−3050. (282) Gao, G.-Y.; Chen, Y.; Zhang, X. P. General Synthesis of mesoAmidoporphyrins via Palladium-Catalyzed Amidation. Org. Lett. 2004, 6, 1837−1840.

(274) Kimberly, B. F.; Joshua, V. R.; Nicole, L. S.; Zhang, X. P. Porphyrin Functionalization via Palladium-Catalyzed Carbon−Heteroatom Cross-Coupling Reactions. Handbook of Porphyrin Science; World Scientific: Singapore, 2010; Vol. 3, pp 367−427. (275) Takanami, T.; Hayashi, M.; Hino, F.; Suda, K. PalladiumCatalyzed meso-Amination and Amidation of Porphyrins: Marked Acceleration with the Ni(II) Central Metal Ion. Tetrahedron Lett. 2003, 44, 7353−7357. (276) Nowak-Król, A.; Gryko, D. T. Oxidative Aromatic Coupling of meso-Arylamino-porphyrins. Org. Lett. 2013, 15, 5618−5621. (277) Chen, Y.; Zhang, X. P. Facile and Efficient Synthesis of mesoArylamino- and Alkylamino-Substituted Porphyrins via PalladiumCatalyzed Amination. J. Org. Chem. 2003, 68, 4432−4438. (278) Ladomenou, K.; Lazarides, T.; Panda, M. K.; Charalambidis, G.; Daphnomili, D.; Coutsolelos, A. G. Meso-substituted Porphyrin Derivatives via Palladium-Catalyzed Amination Showing Wide Range DT

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 224. Preparation of 10-Azacorroles from Nitrogen-Bridged Bisdipyrrin 224.2

Scheme 225. Preparation of a Norcorrole Ni(II) Complex and an Octaphyrin Pd(II) Complex

Dearomatisation of Porphyrins Using Palladium-Catalysed Hydrazination: the 5,15-Diiminoporphodimethenes and Their Redox Products. Tetrahedron 2014, 70, 517−532. (286) Esdaile, L. J.; Senge, M. O.; Arnold, D. P. New Palladium Catalysed Reactions of Bromoporphyrins: Synthesis and Crystal Structures of Nickel(II) Complexes of Primary 5-Aminoporphyrin, 5,5′-Bis(porphyrinyl) Secondary Amine, and 5-Hydroxyporphyrin. Chem. Commun. 2006, 4192−4194.

(283) Fukui, N.; Cha, W.-Y.; Lee, S.; Tokuji, S.; Kim, D.; Yorimitsu, H.; Osuka, A. Oxidative Fusion Reactions of meso-(Diarylamino)porphyrins. Angew. Chem., Int. Ed. 2013, 52, 9728−9732. (284) Esdaile, L. J.; McMurtrie, J. C.; Turnerb, P.; Arnold, D. P. Modified Porphyrinoids from Carbazates and Hydrazones and the First Crystal Structure of a Di-iminoporphodimethene. Tetrahedron Lett. 2005, 46, 6931−6935. (285) Esdaile, L. J.; Simpkins, G. L.; Locos, O. B.; Rintoul, L.; Duriska, M. P.; Turner, P.; McMurtrie, J. C.; Arnold, D. P. Facile DU

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Scheme 226. Preparation of an Ethynylene-Bridged Bisdipyrrin

(287) Pereira, A. M. V. M; Neves, M. G. P. M. S.; Cavaleiro, J. A. S.; Jeandon, C.; Gisselbrecht, J.-P.; Choua, S.; Ruppert, R. Diporphyrinylamines: Synthesis and Electrochemistry. Org. Lett. 2011, 13, 4742− 4745. (288) Atefi, F.; McMurtrie, J. C.; Turner, P.; Duriska, M.; Arnold, D. P. meso-Porphyrinylphosphine Oxides: Mono- and Bidentate Ligands for Supramolecular Chemistry and the Crystal Structures of Monomeric {[10,20-Diphenylporphyrinatonickel(II)-5,15-diyl]-bis-[P(O)Ph2] and Polymeric Self-Coordinated {[10,20Diphenylporphyrinatozinc(II)-5,15-diyl]-bis-[P(O)Ph 2 ]}. Inorg. Chem. 2006, 45, 6479−6489. (289) Chen, Y.; Gao, G.-Y.; Zhang, X. P. Palladium-Mediated Synthesis of Novel meso-Chiral Porphyrins: Cobalt-Catalyzed Cyclopropanation. Tetrahedron Lett. 2005, 46, 4965−4969. (290) Gao, G.-Y.; Colvin, A. J.; Chen, Y.; Zhang, X. P. Versatile Synthesis of meso-Aryloxy- and Alkoxy-Substituted Porphyrins via Palladium-Catalyzed C−O Cross-Coupling Reactions. Org. Lett. 2003, 5, 3261−3264. (291) Gao, G.-Y.; Ruppel, J. V.; Fields, K. B.; Xu, X.; Chen, Y.; Zhang, X. P. Synthesis of Diporphyrins via Palladium-Catalyzed C−O Bond Formation: Effective Access to Chiral Diporphyrins. J. Org. Chem. 2008, 73, 4855−4858. (292) Esdaile, L. J.; Rintoul, L.; Goh, M. S.; Merahi, K.; Parizel, N.; Wellard, R. M.; Choua, S.; Arnold, D. P. Nickel(II) mesoHydroxyporphyrin Complexes Revisited: Palladium-Catalysed Synthesis, Electronic Structures of Derived Oxy Radicals, and Oxidative Coupling to a Dioxoporphodimethene Dyad. Chem. - Eur. J. 2016, 22, 3430−3446. (293) Shimizu, D.; Oh, J.; Furukawa, K.; Kim, D.; Osuka, A. Triarylporphyrin meso-Oxy Radicals: Remarkable Chemical Stabilities and Oxidation to Oxophlorin π-Cations. J. Am. Chem. Soc. 2015, 137, 15584−15594. (294) Gao, G.-Y.; Colvin, A. J.; Chen, Y.; Zhang, X. P. Synthesis of meso-Arylsulfanyl- and Alkylsulfanyl-Substituted Porphyrins via Palladium-Mediated C−S Bond Formation. J. Org. Chem. 2004, 69, 8886− 8892. (295) Ryan, A. A.; Plunkett, S.; Casey, A.; McCabe, T.; Senge, M. O. From Thioether Substituted Porphyrins to Sulfur Linked Porphyrin Dimers: An Unusual SNAr via Thiolate Displacement? Chem. Commun. 2014, 50, 353−355. (296) Alba, M.; Pazos-Perez, N.; Vaz, B.; Formentin, P.; Tebbe, M.; Correa-Duarte, M. A.; Granero, P.; Ferrø-Borrull, J.; Alvarez, R.; Pallares, J.; et al. Macroscale Plasmonic Substrates for Highly Sensitive Surface- Enhanced Raman Scattering. Angew. Chem., Int. Ed. 2013, 52, 6459−6463.

(297) Gao, G.-Y.; Ruppel, J. V.; Allen, D. B.; Chen, Y.; Zhang, X. P. Synthesis of β-Functionalized Porphyrins via Palladium-Catalyzed Carbon−Heteroatom Bond Formations: Expedient Entry into βChiral Porphyrins. J. Org. Chem. 2007, 72, 9060−9066. (298) Kitano, M.; Shimizu, D.; Tanaka, T.; Yorimitsu, H.; Osuka, A. Synthesis of meso-Heteroatom-substituted Subporphyrins. J. Porphyrins Phthalocyanines 2014, 18, 659−665. (299) Haumesser, J.; Gisselbrecht, J.-P.; Weiss, J.; Ruppert, R. Carbene Spacers in Bis-porphyrinic Scaffolds. Chem. Commun. 2012, 48, 11653−11655. (300) Haumesser, J.; Gisselbrecht, J.-P.; Karmazin-Brelot, L.; Bailly, C.; Weiss, J.; Ruppert, R. Synthesis and Electrochemical Studies of Porphyrin Dimers Linked by Metallocarbenes. Organometallics 2014, 33, 4923−4930. (301) Li, L.; Zhu, L.; Chen, D.; Hu, X.; Wang, R. Use of Acylhydrazine- and Acylhydrazone-Type Ligands to Promote CuICatalyzed C−N Cross-Coupling Reactions of Aryl Bromides with NHeterocycles. Eur. J. Org. Chem. 2011, 2011, 2692−2696. (302) Haumesser, J.; Pereira, A. M. V. M; Gisselbrecht, J.-P.; Merahi, K.; Choua, S.; Weiss, J.; Cavaleiro, J. A. S.; Ruppert, R. Inexpensive and Efficient Ullmann Methodology To Prepare Donor-Substituted Porphyrins. Org. Lett. 2013, 15, 6282−6285. (303) Pereira, A.; Jeandon, C.; Cavaleiro, J. A. S.; Neves, M.; Ruppert, R. A New Porphyrin Dimer as an Unexpected Side-Product. J. Porphyrins Phthalocyanines 2014, 18, 727−734. (304) Bruhn, T.; Witterauf, F.; Götz, D. C. G.; Grimmer, C. T.; Würtemberger, M.; Radius, U.; Bringmann, G. C,C- and N,C-Coupled Dimers of 2-Aminotetraphenylporphyrins: Regiocontrolled Synthesis, Spectroscopic Properties, and Quantum-Chemical Calculations. Chem. - Eur. J. 2014, 20, 3998−4006. (305) Ito, S.; Hiroto, S.; Lee, S.; Son, M.; Hisaki, I.; Yoshida, T.; Kim, D.; Kobayashi, N.; Shinokubo, H. Synthesis of Highly Twisted and Fully π-Conjugated Porphyrinic Oligomers. J. Am. Chem. Soc. 2015, 137, 142−145. (306) Akita, M.; Hiroto, S.; Shinokubo, H. Oxidative Annulation of β-Aminoporphyrins into Pyrazine-Fused Diporphyrins. Angew. Chem., Int. Ed. 2012, 51, 2894−2897. (307) Ito, S.; Hiroto, S.; Ousaka, N.; Yashima, E.; Shinokubo, H. Control of Conformation and Chirality of Nonplanar π-Conjugated Diporphyrins Using Substituents and Axial Ligands. Chem. - Asian J. 2016, 11, 936−942. (308) Mandoj, F.; Nardis, S.; Pudi, R.; Lvova, L.; Fronczek, F. R.; Smith, K. M.; Prodi, L.; Genovese, D.; Paolesse, R. β-Pyrazino-Fused Tetrarylporphyrins. Dyes Pigm. 2013, 99, 136−143. DV

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(309) Matano, Y.; Shinokura, T.; Matsumoto, K.; Imahori, H.; Nakano, H. Synthesis and Aggregation Behavior of meso-Sulfinylporphyrins: Evaluation of S-Chirality Effects on the Self-Organization to S−Oxo-Tethered Cofacial Porphyrin Dimers. Chem. - Asian J. 2007, 2, 1417−1429. (310) Matano, Y.; Matsumoto, K.; Terasaka, Y.; Hotta, H.; Araki, Y.; Ito, O.; Shiro, M.; Sasamori, T.; Tokitoh, N.; Imahori, H. Synthesis, Structures, and Properties of meso-Phosphorylporphyrins: SelfOrganization through P−Oxo−Zinc Coordination. Chem. - Eur. J. 2007, 13, 891−901. (311) Liu, C.; Shen, D.-M.; Chen, Q.-Y. Practical and Efficient Synthesis of Various meso-Functionalized Porphyrins via Simple Ligand-Free Nickel-Catalyzed C−O, C−N, and C−C Cross-Coupling Reactions. J. Org. Chem. 2007, 72, 2732−2736. (312) Fagnou, K.; Lautens, M. Rhodium-Catalyzed Carbon−Carbon Bond Forming Reactions of Organometallic Compounds. Chem. Rev. 2003, 103, 169−196. (313) Hayashi, T.; Yamasaki, K. Rhodium-Catalyzed Asymmetric 1,4Addition and Its Related Asymmetric Reactions. Chem. Rev. 2003, 103, 2829−2844. (314) Miura, T.; Murakami, M. Formation of Carbocycles through Sequential Carborhodation Triggered by Addition of Organoborons. Chem. Commun. 2007, 217−224. (315) Baba, H.; Chen, J.; Shinokubo, H.; Osuka, A. Efficient Rhodium-Catalyzed Installation of Unsaturated Ester Functions onto Porphyrins: Site-Specific Heck-Type Addition versus Conjugate Addition. Chem. - Eur. J. 2008, 14, 4256−4262. (316) Park, J. K.; Lee, H. R.; Chen, J.; Shinokubo, H.; Osuka, A.; Kim, D. Photoelectrochemical Properties of Doubly β-Functionalized Porphyrin Sensitizers for Dye-Sensitized Nanocrystalline-TiO2 Solar Cells. J. Phys. Chem. C 2008, 112, 16691−16699. (317) Chen, J.; Aratani, N.; Shinokubo, H.; Osuka, A. PostModification of meso−meso-Linked Porphyrin Arrays by Iridium and Rhodium Catalyses for Tuning of Energy Gap. Chem. - Asian J. 2009, 4, 1126−1133. (318) Park, J. K.; Chen, J.; Lee, H. R.; Park, S. W.; Shinokubo, H.; Osuka, A.; Kim, D. Doubly β-Functionalized meso−meso Directly Linked Porphyrin Dimer Sensitizers for Photovoltaics. J. Phys. Chem. C 2009, 113, 21956−21963. (319) Taniguchi, M.; Lindsey, J. S. Synthesis of Oligo(p-Phenylene)Linked Dyads Containing Free Base, Zinc(II) or Thallium(III) Porphyrins for Studies in Artificial Photosynthesis. Tetrahedron 2010, 66, 5549−5565. (320) Ishi-i, T.; Taguri, Y.; Kato, S.; Shigeiwa, M.; Gorohmaru, H.; Maeda, S.; Mataka, S. Singlet Oxygen Generation by Two-Photon Excitation of Porphyrin Derivatives Having Two-Photon-Absorbing Benzothiadiazole Chromophores. J. Mater. Chem. 2007, 17, 3341− 3346. (321) Barker, C.; Zeng, X.; Bettington, S.; Batsanov, A. S.; Bryce, M. R.; Beeby, A. Porphyrin, Phthalocyanine and Porphyrazine Derivatives with Multifluorenyl Substituents as Efficient Deep-Red Emitters. Chem. - Eur. J. 2007, 13, 6710−6717. (322) Cunha, A. C.; Gomes, A. T. P. C.; Ferreira, V. F.; de Souza, M. C. B. V.; Neves, M. G. P. M. S.; Tomé, A. C.; Silva, A. M. S.; Cavaleiro, J. A. S. Synthesis of β-Arylporphyrins and Oligophenylenediporphyrins by the Suzuki−Miyaura Reaction. Synthesis 2010, 2010, 510−514. (323) Chan, C.-S.; Mak, C. C.; Chan, K. S. Synthesis of Monoquinonyl Porphyrins via Benzannulation of Fischer Carbene Complexes and Palladium Catalysed Cross Coupling Reactions. Tetrahedron Lett. 1993, 34, 5125−5126. (324) Chan, C.-S.; Tse, A. K. S.; Chan, K. S. Highly Versatile Methods for the Synthesis of Quinonylporphyrins via Benzannulation of Fischer Carbene Complexes and Palladium-Catalyzed CrossCoupling Reactions. J. Org. Chem. 1994, 59, 6084−6089. (325) Michalak, J.; Birin, K. P.; Muniappan, S.; Ranyuk, E.; Enakieva, Y. Y.; Gorbunova, Y. G.; Stern, C.; Bessmertnykh-Lemeune, A.; Guilard, R. Synthesis of Porphyrin-Bis(Polyazamacrocycle) Triads via Suzuki Coupling Reaction. J. Porphyrins Phthalocyanines 2014, 18, 35− 48.

(326) Kimura, M.; Shiba, T.; Yamazaki, M.; Hanabusa, K.; Shirai, H.; Kobayashi, N. Construction of Regulated Nanospace around a Porphyrin Core. J. Am. Chem. Soc. 2001, 123, 5636−5642. (327) Capitosti, G. J.; Guerrero, C. D.; Binkley, D. E.; Rajesh, C. S.; Modarelli, D. A. Efficient Synthesis of Porphyrin-Containing, Benzoquinone-Terminated, Rigid Polyphenylene Dendrimers. J. Org. Chem. 2003, 68, 247−261. (328) Arnold, D. P.; Nitschinsk, L. J. The Preparation of Novel Porphyrins and Bis(Porphyrins) Using Palladium Catalysed Coupling Reactions. Tetrahedron Lett. 1993, 34, 693−696. (329) Arnold, D. P.; James, D. A. Dimers and Model Monomers of Nickel(II) Octaethylporphyrin Substituted by Conjugated Groups Comprising Combinations of Triple Bonds with Double Bonds and Arenes. Synthesis and Electronic Spectra. J. Org. Chem. 1997, 62, 3460−3469. (330) Tremblay-Morin, J.-P.; Ali, H.; van Lier, J. E. First Example of a Palladium Catalyzed Coupling Reaction between Cationic Porphyrins and Alkynyls in Aqueous Medium. Tetrahedron Lett. 2005, 46, 6999− 7002. (331) Clausen, C.; Gryko, D. T.; Dabke, R. B.; Dontha, N.; Bocian, D. F.; Kuhr, W. G.; Lindsey, J. S. Synthesis of Thiol-Derivatized Porphyrin Dimers and Trimers for Studies of Architectural Effects on Multibit Information Storage. J. Org. Chem. 2000, 65, 7363−7370. (332) Wagner, R. W.; Johnson, T. E.; Li, F.; Lindsey, J. S. Synthesis of Ethyne-Linked or Butadiyne-Linked Porphyrin Arrays Using Mild, Copper-Free, Pd-Mediated Coupling Reactions. J. Org. Chem. 1995, 60, 5266−5273. (333) Ljungdahl, T.; Pettersson, K.; Albinsson, B.; Mårtensson, J. Solvent and Base Dependence of Copper-Free Palladium-Catalyzed Cross-Couplings between Terminal Alkynes and Arylic Iodides: Development of Efficient Conditions for the Construction of Gold(III)/Free-Base Porphyrin Dimers. J. Org. Chem. 2006, 71, 1677−1687. (334) Meudtner, R. M.; Ostermeier, M.; Goddard, R.; Limberg, C.; Hecht, S. Multifunctional “Clickates” as Versatile Extended Heteroaromatic Building Blocks: Efficient Synthesis via Click Chemistry, Conformational Preferences, and Metal Coordination. Chem. - Eur. J. 2007, 13, 9834−9840. (335) Li, J.; Gryko, D.; Dabke, R. B.; Diers, J. R.; Bocian, D. F.; Kuhr, W. G.; Lindsey, J. S. Synthesis of Thiol-Derivatized Europium Porphyrinic Triple-Decker Sandwich Complexes for Multibit Molecular Information Storage. J. Org. Chem. 2000, 65, 7379−7390. (336) Kozaki, M.; Akita, K.; Okada, K. Enhanced Electron Transfer by Dendritic Architecture: Energy Transfer and Electron Transfer in Snowflake-Shaped Zn Porphyrin Dendrimers. Org. Lett. 2007, 9, 1509−1512. (337) Kozaki, M.; Akita, K.; Suzuki, S.; Okada, K. Construction of a Rigid Zn Porphyrin−C60 Dyad within Dendritic Structure: Dendrimer Effect on Singlet Energy Transfer. Org. Lett. 2007, 9, 3315−3318. (338) Kozaki, M.; Uetomo, A.; Suzuki, S.; Okada, K. A LightHarvesting Array Composed of Porphyrins and Rigid Backbones. Org. Lett. 2008, 10, 4477−4480. (339) Kozaki, M.; Tujimura, H.; Suzuki, S.; Okada, K. Stepwise Construction of a Cross-Shaped Covalent Assembly of Dendrimers. Tetrahedron Lett. 2008, 49, 2931−2934. (340) Kozaki, M.; Morita, S.; Suzuki, S.; Okada, K. Construction of Snowflake-Shaped Dendritic Covalent Assemblies with Rigid Conjugated Networks. J. Org. Chem. 2012, 77, 9447−9457. (341) Tomizaki, K.; Loewe, R. S.; Kirmaier, C.; Schwartz, J. K.; Retsek, J. L.; Bocian, D. F.; Holten, D.; Lindsey, J. S. Synthesis and Photophysical Properties of Light-Harvesting Arrays Comprised of a Porphyrin Bearing Multiple Perylene-Monoimide Accessory Pigments. J. Org. Chem. 2002, 67, 6519−6534. (342) Prathapan, S.; Yang, S. I.; Seth, J.; Miller, M. A.; Bocian, D. F.; Holten, D.; Lindsey, J. S. Synthesis and Excited-State Photodynamics of Perylene−Porphyrin Dyads. 1. Parallel Energy and Charge Transfer via a Diphenylethyne Linker. J. Phys. Chem. B 2001, 105, 8237−8248. (343) Kirmaier, C.; Song, H.; Yang, E.; Schwartz, J. K.; Hindin, E.; Diers, J. R.; Loewe, R. S.; Tomizaki, K.; Chevalier, F.; Ramos, L.; et al. DW

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Synthetic Approach to N-Arylquinolino[2,3,4-at]porphyrins from βArylaminoporphyrins. J. Org. Chem. 2008, 73, 7353−7356. (361) Artamkina, G. A.; Sazonov, P. K.; Shtern, M. M.; Grishina, G. V.; Veselov, I. S.; Semeikin, A. S.; Syrbu, S. A.; Koifman, O. I.; Beletskaya, I. P. Palladium-Catalyzed N-Arylation of Hydroxypiperidines with meso-Bromophenyl(polyalkyl)porphyrins. Synlett 2008, 2008, 45−48. (362) Mikhalitsyna, E. A.; Tyurin, V. S.; Zamylatskov, I. A.; Khrustalev, V. N.; Beletskaya, I. P. Synthesis, Characterization and Cation-Induced Dimerization of New Aza-Crown Ether-Appended Metalloporphyrins. Dalton Trans. 2012, 41, 7624−7636. (363) Mikhalitsyna, E. A.; Tyurin, V. S.; Khrustalev, V. N.; Lonin, I. S.; Beletskaya, I. P. Palladium-Catalyzed Amination of meso(Bromophenyl)porphyrins with Diamines and Azamacrocycles. Dalton Trans. 2014, 43, 3563−3575. (364) Ranyuk, E. R.; Filatov, M. A.; Averin, A. D.; Cheprakov, A. V.; Beletskaya, I. P. The Synthesis of Highly Basic π-Extended Porphyrins by Palladium-Catalyzed Amination. Synthesis 2012, 44, 393−398. (365) Weng, Y.-Q.; Teng, Y.-L.; Yue, F.; Zhong, Y.-R.; Ye, B.-H. A New Selective Fluorescent Chemosensor for Cu(II) Ion Based on Zinc Porphyrin-Dipyridylamino. Inorg. Chem. Commun. 2007, 10, 443−446. (366) Weng, Y.-Q.; Yue, F.; Zhong, Y.-R.; Ye, B.-H. A Copper(II) Ion-Selective On−Off-Type Fluoroionophore Based on Zinc Porphyrin−Dipyridylamino. Inorg. Chem. 2007, 46, 7749−7755. (367) Huang, C.-Y.; Su, Y. O. Spectral and Redox Properties of Zinc Porphyrin Core Dendrimers with Triarylamines as Dendron. Dalton Trans. 2010, 39, 8306−8312. (368) Yakushev, A. A.; Averin, A. D.; Maloshitskaya, O. A.; Syrbu, S. A.; Koifman, O. I.; Beletskaya, I. P. Palladium- and Copper-Catalyzed Amination of Halogenophenyl Substituted Porphyrins for the Synthesis of Porphyrin-Azacrown Ethers Conjugates and Evaluation of Their Sensing Properties. Makrogeterotsikly 2016, 9, 65−72. (369) Wu, J.; Jiang, X.-L.; Zhang, H.-L. Synthesis of Diaryl EtherLinked Porphyrin Dimers via Ullmann Coupling Reaction. Heterocycles 2006, 68, 2153−2160. (370) Arnold, D. P.; Johnson, A. W.; Mahendran, M. Some Reactions of meso-Formyloctaethylporphyrin. J. Chem. Soc., Perkin Trans. 1 1978, 366−370. (371) Arnold, D. P. Two Rings Are Better Than One: Adventures in Porphyrin Chemistry. Synlett 2000, 296−305. (372) Arnold, D. P.; Nitschinsk, L. J. Porphyrin Dimers Linked by Conjugated Butadiynes. Tetrahedron 1992, 48, 8781−8792. (373) Stranger, R.; McGrady, J. E.; Arnold, D. P.; Lane, I.; Heath, G. A. Communication between Porphyrin Rings in the Butadiyne-Bridged Dimer Ni(OEP)(μ-C4)Ni(OEP): A Density Functional Study. Inorg. Chem. 1996, 35, 7791−7797. (374) Arnold, D. P.; Heath, G. A.; James, D. A. Conjugated Dimers of Nickel(II) Octaethylporphyrin Linked by Extended meso,mesoAlkynyl Bridges. II. Redox Properties and Electronic Spectra of Electrogenerated Anions and Dianions. New J. Chem. 1998, 22, 1377− 1387. (375) Anderson, H. L. Building Molecular Wires from the Colours of Life: Conjugated Porphyrin Oligomers. Chem. Commun. 1999, 2323− 2330. (376) Anderson, H. L. Conjugated Porphyrin Ladders. Inorg. Chem. 1994, 33, 972−981. (377) Anderson, H. L.; Martin, S. J.; Bradley, D. D. C. Synthesis and Third-Order Nonlinear Optical Properties of a Conjugated Porphyrin Polymer. Angew. Chem., Int. Ed. Engl. 1994, 33, 655−657. (378) Drobizhev, M.; Stepanenko, Y.; Dzenis, Y.; Karotki, A.; Rebane, A.; Taylor, P. N.; Anderson, H. L. Understanding Strong Two-Photon Absorption in π-Conjugated Porphyrin Dimers via Double-Resonance Enhancement in a Three-Level Model. J. Am. Chem. Soc. 2004, 126, 15352−15353. (379) Lin, V. S.; Therien, M. J. The Role of Porphyrin-to-Porphyrin Linkage Topology in the Extensive Modulation of the Absorptive and Emissive Properties of a Series of Ethynyl- and Butadiynyl-Bridged Bis-

Excited-State Photodynamics of Perylene−Porphyrin Dyads. 5. Tuning Light-Harvesting Characteristics via Perylene Substituents, Connection Motif, and Three-Dimensional Architecture. J. Phys. Chem. B 2010, 114, 14249−14264. (344) Tomizaki, K.; Thamyongkit, P.; Loewe, R. S.; Lindsey, J. S. Practical Synthesis of Perylene-Monoimide Building Blocks That Possess Features Appropriate for Use in Porphyrin-Based LightHarvesting Arrays. Tetrahedron 2003, 59, 1191−1207. (345) Muthukumaran, K.; Loewe, R. S.; Kirmaier, C.; Hindin, E.; Schwartz, J. K.; Sazanovich, I. V.; Diers, J. R.; Bocian, D. F.; Holten, D.; Lindsey, J. S. Synthesis and Excited-State Photodynamics of A Perylene-Monoimide-Oxochlorin Dyad. A Light-Harvesting Array. J. Phys. Chem. B 2003, 107, 3431−3442. (346) Kirmaier, C.; Hindin, E.; Schwartz, J. K.; Sazanovich, I. V.; Diers, J. R.; Muthukumaran, K.; Taniguchi, M.; Bocian, D. F.; Lindsey, J. S.; Holten, D. Synthesis and Excited-State Photodynamics of Perylene-Bis(imide)-Oxochlorin Dyads. A Charge-Separation Motif. J. Phys. Chem. B 2003, 107, 3443−3454. (347) Loewe, R. S.; Tomizaki, K.; Chevalier, F.; Lindsey, J. S. Synthesis of Perylene-Porphyrin Dyads for Light-Harvesting Studies. J. Porphyrins Phthalocyanines 2002, 06, 626−642. (348) Loewe, R. S.; Tomizaki, K.; Youngblood, W. J.; Bo, Z.; Lindsey, J. S. Synthesis of Perylene-Porphyrin Building Blocks and Rod-Like Oligomers for Light-Harvesting Applications. J. Mater. Chem. 2002, 12, 3438−3451. (349) Miller, M. A.; Lammi, R. K.; Prathapan, S.; Holten, D.; Lindsey, J. S. A Tightly Coupled Linear Array of Perylene, Bis(Porphyrin), and Phthalocyanine Units that Functions as a Photoinduced Energy-Transfer Cascade. J. Org. Chem. 2000, 65, 6634−6649. (350) Jiang, B.; Yang, S.-W.; Barbini, D. C.; Jones, W. E., Jr Synthesis of Soluble Conjugated Metalloporphyrin Polymers with Tunable Electronic Properties. Chem. Commun. 1998, 213−214. (351) Huang, X.; Zhu, C.; Zhang, S.; Li, W.; Guo, Y.; Zhan, X.; Liu, Y.; Bo, Z. Porphyrin−Dithienothiophene π-Conjugated Copolymers: Synthesis and Their Applications in Field-Effect Transistors and Solar Cells. Macromolecules 2008, 41, 6895−6902. (352) Pereira, M. M.; Muller, G.; Ordinas, J. I.; Azenha, M. E.; Arnaut, L. G. Synthesis of Vinylated 5,10,15,20-Tetraphenylporphyrins via Heck-Type Coupling Reaction and Their Photophysical Properties. J. Chem. Soc. Perkin Trans. 2 2002, 1583−1588. (353) Castella, M.; Calahorra, F.; Sainz, D.; Velasco, D. New HeckType Reaction Applied to the Synthesis of Protoporphyrin-IX Derivatives. Org. Lett. 2001, 3, 541−544. (354) Lee, Y.; Liang, Y.; Yu, L. The Heck Polycondensation for Functional Polymers. Synlett 2006, 2006, 2879−2893. (355) You, W.; Wang, L.; Wang, Q.; Yu, L. Synthesis and Structure/ Property Correlation of Fully Functionalized Photorefractive Polymers. Macromolecules 2002, 35, 4636−4645. (356) Wang, Q.; Wang, L.; Yu, J.; Yu, L. Fully Functionalized Photorefractive Polymers Incorporating Transition Metal Phthalocyanine and Porphyrin Complexes as Photosensitizers. Adv. Mater. 2000, 12, 974−978. (357) Gao, G. Y.; Chen, Y.; Zhang, X. P. General and Efficient Synthesis of Arylamino- and Alkylamino-Substituted Diphenylporphyrins and Tetraphenylporphyrins via Palladium-Catalyzed Multiple Amination Reactions. J. Org. Chem. 2003, 68, 6215−6221. (358) Zhu, S.; Ruppel, J. V.; Lu, H.; Wojtas, L.; Zhang, X. P. CobaltCatalyzed Asymmetric Cyclopropanation with Diazosulfones: Rigidification and Polarization of Ligand Chiral Environment via Hydrogen Bonding and Cyclization. J. Am. Chem. Soc. 2008, 130, 5042−5043. (359) Soares, A. R. M.; Martínez-Díaz, M. V.; Bruckner, A.; Pereira, A. M. V. M; Tomé, J. P. C.; Alonso, C. M. A.; Faustino, M. A. F.; Neves, M. G. P. M. S.; Tomé, A. C.; Silva, A. M. S.; et al. Synthesis of Novel N-Linked Porphyrin−Phthalocyanine Dyads. Org. Lett. 2007, 9, 1557−1560. (360) Pereira, A. M. V. M; Alonso, C. M. A.; Neves, M. G. P. M. S.; Tomé, A. C.; Silva, A. M. S.; Paz, F. A. A.; Cavaleiro, J. A. S. A New DX

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

and Tris(porphinato)zinc Chromophores. Chem. - Eur. J. 1995, 1, 645−651. (380) Youngblood, W. J.; Gryko, D. T.; Lammi, R. K.; Bocian, D. F.; Holten, D.; Lindsey, J. S. Glaser-Mediated Synthesis and Photophysical Characterization of Diphenylbutadiyne-Linked Porphyrin Dyads. J. Org. Chem. 2002, 67, 2111−2117. (381) Nakamura, K.; Fujimoto, T.; Takara, S.; Sugiura, K.; Miyasaka, H.; Ishii, T.; Yamashita, M.; Sakata, Y. Systematic Synthesis of Porphyrin Dimers Linked by Conjugated Oligoacetylene Bridges. Chem. Lett. 2003, 32, 694−695. (382) Rana, S.; Yamashita, K.; Akita, Y.; Sugiura, K. Stepwise Synthesis of Structurally Well-Defined Porphyrin Wires Connected by Platinum Acetylides. Chem. Lett. 2015, 44, 1226−1228. (383) Sugiura, K.; Fujimoto, Y.; Sakata, Y. A porphyrin Square: Synthesis of A Square-Shaped π-Conjugated Porphyrin Tetramer connected by Diacetylene Linkages. Chem. Commun. 2000, 1105− 1106. (384) Kato, A.; Sugiura, K.; Miyasaka, H.; Tanaka, H.; Kawai, T.; Sugimoto, M.; Yamashita, M. A Square Cyclic Porphyrin Dodecamer: Synthesis and Single-Molecule Characterization. Chem. Lett. 2004, 33, 578−579. (385) Wang, S.-P.; Shen, Y.-F.; Zhu, B.-Y.; Wu, J.; Li, S. Recent Advances in Template-Directed Synthesis of Porphyrin Nanorings. Chem. Commun. 2016, 52, 10205−10216. (386) Anderson, H. L.; Sanders, J. K. M. Amine-Template-Directed Synthesis of Cyclic Porphyrin Oligomers. Angew. Chem., Int. Ed. Engl. 1990, 29, 1400−1403. (387) Anderson, S.; Anderson, H. L.; Sanders, J. K. M. TemplateDirected Synthesis of Linear and Cyclic Butadiyne-Linked Porphyrin Oligomers up to a Linear Octamer. J. Chem. Soc., Perkin Trans. 1 1995, 2247−2254. (388) Marvaud, V.; Vidal-Ferran, A.; Webb, S. J.; Sanders, J. K. M. Stereospecific Templated Synthesis of a Trisubstituted ButadiyneLinked Cyclic Porphyrin Trimer. J. Chem. Soc., Dalton Trans. 1997, 985−990. (389) Hoffmann, M.; Kärnbratt, J.; Chang, M.-H.; Herz, L. M.; Albinsson, B.; Anderson, H. L. Enhanced π Conjugation around a Porphyrin[6] Nanoring. Angew. Chem., Int. Ed. 2008, 47, 4993−4996. (390) Sprafke, J. K.; Kondratuk, D. V.; Wykes, M.; Thompson, A. L.; Hoffmann, M.; Drevinskas, R.; Chen, W.-H.; Yong, C. K.; Kärnbratt, J.; Bullock, J. E.; et al. Belt-Shaped π-Systems: Relating Geometry to Electronic Structure in a Six-Porphyrin Nanoring. J. Am. Chem. Soc. 2011, 133, 17262−17273. (391) O’Sullivan, M. C.; Sprafke, J. K.; Kondratuk, D. V.; Rinfray, C.; Claridge, T. D. W.; Saywell, A.; Blunt, M. O.; O’Shea, J. N.; Beton, P. H.; Malfois, M.; et al. Vernier Templating and Synthesis of a 12Porphyrin Nano-Ring. Nature 2011, 469, 72−75. (392) Hoffmann, M.; Wilson, C. J.; Odell, B.; Anderson, H. L. Template-Directed Synthesis of a π-Conjugated Porphyrin Nanoring. Angew. Chem., Int. Ed. 2007, 46, 3122−3125. (393) Sprafke, J. K.; Odell, B.; Claridge, T. D. W.; Anderson, H. L. All-or-Nothing Cooperative Self-Assembly of an Annulene Sandwich. Angew. Chem., Int. Ed. 2011, 50, 5572−5575. (394) Langton, M. J.; Matichak, J. D.; Thompson, A. L.; Anderson, H. L. Template-Directed Synthesis of π-Conjugated Porphyrin [2]Rotaxanes and a [4]Catenane Based on a Six-Porphyrin Nanoring. Chem. Sci. 2011, 2, 1897−1901. (395) Liu, P.; Neuhaus, P.; Kondratuk, D. V.; Balaban, T. S.; Anderson, H. L. Cyclodextrin-Templated Porphyrin Nanorings. Angew. Chem., Int. Ed. 2014, 53, 7770−7773. (396) Neuhaus, P.; Cnossen, A.; Gong, J. Q.; Herz, L. M.; Anderson, H. L. A Molecular Nanotube with Three-Dimensional π-Conjugation. Angew. Chem., Int. Ed. 2015, 54, 7344−7348. (397) Rousseaux, S. A. L.; Gong, J. Q.; Haver, R.; Odell, B.; Claridge, T. D. W.; Herz, L. M.; Anderson, H. L. Self-Assembly of Russian Doll Concentric Porphyrin Nanorings. J. Am. Chem. Soc. 2015, 137, 12713−12718. (398) Favereau, L.; Cnossen, A.; Kelber, J. B.; Gong, J. Q.; Oetterli, R. M.; Cremers, J.; Herz, L. M.; Anderson, H. L. Six-Coordinate Zinc

Porphyrins for Template-Directed Synthesis of Spiro-Fused Nanorings. J. Am. Chem. Soc. 2015, 137, 14256−14259. (399) Sergeeva, N. N.; Bakar, M. B.; Senge, M. O. Synthesis, Transformations, and Comparative Studies of Porphyryl Acrylic Acids and Their Homologues. J. Org. Chem. 2009, 74, 1488−1497. (400) Lo, M.; Mahajan, D.; Wytko, J. A.; Boudon, C.; Weiss, J. Toward Generic Models of Hemoproteins. Org. Lett. 2009, 11, 2487− 2490. (401) Liu, X.; Sternberg, E.; Dolphin, D. Cross-Metathesis Reactions of Vinyl-Chlorins and -Porphyrins Catalyzed by a “Second Generation” Grubbs’ Catalyst. Chem. Commun. 2004, 852−853. (402) Liu, X.; Sternberg, E.; Dolphin, D. Cross-Metathesis of the Vinyl Group on Tetrapyrrolic Macrocycles: Reactivity, Selectivity, and Mechanism. J. Org. Chem. 2008, 73, 6542−6550. (403) Lonin, I. S.; Grin, M. A.; Lakhina, A. A.; Mironov, A. F. Synthesis of Chlorophyll a Glycoconjugates using Olefin CrossMetathesis. Mendeleev Commun. 2012, 22, 157−158. (404) Sol, V.; Chaleix, V.; Granet, R.; Krausz, P. An Efficient Route to Dimeric Porphyrin−RGD Peptide Conjugates via Olefin Metathesis. Tetrahedron 2008, 64, 364−371. (405) Maggini, M.; Scorrano, G.; Prato, M. Addition of Azomethine Ylides to C60: Synthesis, Characterization, and Functionalization of Fullerene Pyrrolidines. J. Am. Chem. Soc. 1993, 115, 9798−9799. (406) Bell, T. D. M.; Ghiggino, K. P.; Haynes, A.; Langford, S. J.; Woodward, C. P. Charge Separation in a Conformationally-Flexible Porphyrin−Fullerene Dyad Synthesised Using Cross-Metathesis. J. Porphyrins Phthalocyanines 2007, 11, 455−462. (407) Zheng, G.; Dougherty, T. J.; Pandey, R. K. Novel Chlorin− Diene Building Block by Enyne Metathesis: Synthesis of Chlorin− Fullerene Dyads. Chem. Commun. 1999, 2469−2470. (408) Ostrowski, S.; Mikus, A. Synthesis of Porphyrin-Acetylene and Porphyrin-Diene Building Blocks for New Dyads Preparation. Heterocycles 2005, 65, 2339−2346. (409) Ostrowski, S.; Mikus, A. A New Approach to the Synthesis of Porphyrin-Fullerene Dyads. Mol. Diversity 2000, 6, 315−321. (410) Jiao, L.; Hao, E.; Fronczek, F. R.; Vicente, M. G. H.; Smith, K. M. Benzoporphyrins via an Olefin Ring-Closure Metathesis Methodology. Chem. Commun. 2006, 3900−3902. (411) Guo, Y.-G.; Mele, G.; Martina, F.; Margapoti, E.; Vasapollo, G.; Xiao, W.-J. An Efficient Route to Biscardanol Derivatives and Cardanol-Based Porphyrins via Olefin Metathesis. J. Organomet. Chem. 2006, 691, 5383−5390. (412) Mele, G.; Li, J.; Margapoti, E.; Martina, F.; Vasapollo, G. Synthesis of Novel Porphyrins Cardanol Based via Cross Metathesis. Catal. Today 2009, 140, 37−43. (413) Coumans, R. G. E.; Elemans, J. A. A. W.; Thordarson, P.; Nolte, R. J. M.; Rowan, A. E. Synthesis of Porphyrin-Containing [3]Rotaxanes by Olefin Metathesis. Angew. Chem., Int. Ed. 2003, 42, 650−654. (414) Coumans, R. G. E.; Elemans, J. A. A. W.; Rowan, A. E.; Nolte, R. J. M. Interlocked Porphyrin Switches. Chem. - Eur. J. 2013, 19, 7758−7770. (415) Langford, S. J.; Latter, M. J.; Woodward, C. P. Construction of Multiporphyrin Arrays via Selective Cross-Metathesis. Org. Lett. 2006, 8, 2595−2598. (416) Bell, T. D. M.; Bhosale, S. V.; Ghiggino, K. P.; Langford, S. J.; Woodward, C. P. Synthesis and Photophysical Properties of a Conformationally Flexible Mixed Porphyrin Star-Pentamer. Aust. J. Chem. 2009, 62, 692−699. (417) Ikeda, C.; Satake, A.; Kobuke, Y. Proofs of Macrocyclization of Gable Porphyrins as Mimics of Photosynthetic Light-Harvesting Complexes. Org. Lett. 2003, 5, 4935−4938. (418) Ohashi, A.; Satake, A.; Kobuke, Y. Covalent Linking of Coordination-Organized Slipped Cofacial Porphyrin Dimers. Bull. Chem. Soc. Jpn. 2004, 77, 365−374. (419) Shoji, O.; Okada, S.; Satake, A.; Kobuke, Y. Coordination Assembled Rings of Ferrocene-Bridged Trisporphyrin with Flexible Hinge-like Motion: Selective Dimer Ring Formation, Its TransDY

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

formation to Larger Rings, and Vice Versa. J. Am. Chem. Soc. 2005, 127, 2201−2210. (420) Dy, J. T.; Ogawa, K.; Satake, A.; Ishizumi, A.; Kobuke, Y. Water-Soluble Self-Assembled Butadiyne-Bridged Bisporphyrin: A Potential Two-Photon-Absorbing Photosensitizer for Photodynamic Therapy. Chem. - Eur. J. 2007, 13, 3491−3500. (421) Dy, J.; Ogawa, K.; Kamada, K.; Ohta, K.; Kobuke, Y. Stepwise Elongation Effect on the Two-Photon Absorption of Self-Assembled Butadiyne Porphyrins. Chem. Commun. 2008, 3411−3413. (422) Satake, A.; Yamamura, M.; Oda, M.; Kobuke, Y. Transmembrane Nanopores from Porphyrin Supramolecules. J. Am. Chem. Soc. 2008, 130, 6314−6315. (423) Ishida, T.; Morisaki, Y.; Chujo, Y. Synthesis of Covalently Bonded Nanostructure from Two Porphyrin Molecular Wires Leading to a Molecular Tube. Tetrahedron Lett. 2006, 47, 5265−5268. (424) van Gerven, P. C. M.; Elemans, J. A. A. W.; Gerritsen, J. W.; Speller, S.; Nolte, R. J. M.; Rowan, A. E. Dynamic Combinatorial Olefin Metathesis: Templated Synthesis of Porphyrin Boxes. Chem. Commun. 2005, 3535−3537. (425) Bakker, J. M.; Langford, S. J.; Latter, M. J.; Lee, K. A.; Woodward, C. P. Template-Directed Assembly of a Macrocyclic Porphyrin Tetramer Using Olefin Metathesis. Aust. J. Chem. 2005, 58, 757−761. (426) Mullholland, A. R.; Woodward, C. P.; Langford, S. J. FullereneTemplated Synthesis of a Cyclic Porphyrin Trimer Using Olefin Metathesis. Chem. Commun. 2011, 47, 1494−1496. (427) Totten, R. K.; Ryan, P.; Kang, B.; Lee, S. J.; Broadbelt, L. J.; Snurr, R. Q.; Hupp, J. T.; Nguyen, S. T. Enhanced Catalytic Decomposition of a Phosphate Triester by Modularly Accessible Bimetallic Porphyrin Dyads and Dimers. Chem. Commun. 2012, 48, 4178−4180. (428) Youm, K.-T.; Nguyen, S. T.; Hupp, J. T. Hollow Porphyrin Prisms: Modular Formation of Permanent, Torsionally Rigid Nanostructures via Templated Olefin Metathesis. Chem. Commun. 2008, 3375−3377. (429) Kang, B.; Kurutz, J. W.; Youm, K.-T.; Totten, R. K.; Hupp, J. T.; Nguyen, S. T. Catalytically Active Supramolecular Porphyrin Boxes: Acceleration of the Methanolysis of Phosphate Triesters via a Combination of Increased Local Nucleophilicity and Reactant Encapsulation. Chem. Sci. 2012, 3, 1938−1944. (430) Taesch, J.; Heiz, V.; Topic, F.; Rissanen, K. Templated Synthesis of a Large and Flexible Covalent Porphyrinic Cage Bearing Orthogonal Recognition Sites. Chem. Commun. 2012, 48, 5118−5120. (431) Zhu, B.; Chen, H.; Lin, W.; Ye, Y.; Wu, J.; Li, S. TemplateDirected Synthesis of Flexible Porphyrin Nanocage and Nanorings via One-Step Olefin Metathesis. J. Am. Chem. Soc. 2014, 136, 15126− 15129. (432) Zimmerman, S. C.; Wendland, M. S.; Rakow, N. A.; Zharov, I.; Suslick, K. S. Synthetic Hosts by Monomolecular Imprinting Inside Dendrimers. Nature 2002, 418, 399−403. (433) Zimmerman, S. C.; Zharov, I.; Wendland, M. S.; Rakow, N. A.; Suslick, K. S. Molecular Imprinting Inside Dendrimers. J. Am. Chem. Soc. 2003, 125, 13504−13518. (434) Kim, Y.; Mayer, M. F.; Zimmerman, S. A New Route to Organic Nanotubes from Porphyrin Dendrimers. Angew. Chem., Int. Ed. 2003, 42, 1121−1126. (435) Beil, J. B.; Zimmerman, S. C. Synthesis of Nanosized “Cored” Star Polymers. Macromolecules 2004, 37, 778−787. (436) Jyothish, K.; Wang, Q.; Zhang, W. Highly Active Multidentate Alkyne Metathesis Catalysts: Ligand- Activity Relationship and Their Applications in Efficient Synthesis of Porphyrin-Based Aryleneethynylene Polymers. Adv. Synth. Catal. 2012, 354, 2073−2078. (437) Zhang, C.; Long, H.; Zhang, W. A C84 Selective Porphyrin Macrocycle with an Adaptable Cavity Constructed Through Alkyne Metathesis. Chem. Commun. 2012, 48, 6172−6174. (438) Zhang, C.; Wang, Q.; Long, H.; Zhang, W. A Highly C70 Selective Shape-Persistent Rectangular Prism Constructed through One-Step Alkyne Metathesis. J. Am. Chem. Soc. 2011, 133, 20995− 21001.

(439) Fletcher, J. T.; Therien, M. J. Transition-Metal-Mediated [2 + 2 + 2] Cycloaddition Reactions with Ethyne-Containing Porphyrin Templates: New Routes to Cofacial Porphyrin Structures and FaciallyFunctionalized (Porphyrinato)metal Species. J. Am. Chem. Soc. 2000, 122, 12393−12394. (440) Fletcher, J. T.; Therien, M. J. Extreme Electronic Modulation of the Cofacial Porphyrin Structural Motif. J. Am. Chem. Soc. 2002, 124, 4298−4311. (441) Ladomenou, K.; Nikolaou, V.; Charalambidis, G.; Coutsolelos, A. G. Click”-Reaction: An Alternative Tool for New Architectures of Porphyrin Based Derivatives. Coord. Chem. Rev. 2016, 306, 1−42. (442) Dumoulin, F.; Ahsen, V. Click Chemistry: The Emerging Role of the Azide-Alkyne Huisgen Dipolar Addition in the Preparation of Substituted Tetrapyrrolic Derivatives. J. Porphyrins Phthalocyanines 2011, 15, 481−504. (443) Zhang, L.; Chen, X.; Xue, P.; Sun, H. H. Y.; Williams, I. D.; Sharpless, K. B.; Fokin, V. V.; Jia, G. Ruthenium-Catalyzed Cycloaddition of Alkynes and Organic Azides. J. Am. Chem. Soc. 2005, 127, 15998−15999. (444) Chromiński, M.; Zieleniewska, A.; Karczewski, M.; Gryko, D. Porphyrins as Substrates in CuAAC  Exclusion of Unwanted Copper Insertion into the Macrocyclic Core. J. Porphyrins Phthalocyanines 2014, 18, 267−281. (445) Shen, D.-M.; Liu, C.; Chen, Q.-Y. Synthesis and Versatile Reactions of β-Azidotetraarylporphyrins. Eur. J. Org. Chem. 2007, 2007, 1419−1422. (446) Séverac, M.; Pleux, L. L.; Scarpaci, A.; Blart, E.; Odobel, F. Synthesis of New Azido Porphyrins and Their Reactivity in Copper(I)-Catalyzed Huisgen 1,3-Dipolar Cycloaddition Reaction with Alkynes. Tetrahedron Lett. 2007, 48, 6518−6522. (447) Punidha, S.; Sinha, J.; Kumar, A.; Ravikanth, M. First TriazoleBridged Unsymmetrical Porphyrin Dyad via Click Chemistry. J. Org. Chem. 2008, 73, 323−326. (448) Shetti, V. S.; Ravikanth, M. Synthesis of Triazole-Bridged Unsymmetrical Porphyrin Dyads and Porphyrin−Ferrocene Conjugates. Eur. J. Org. Chem. 2010, 2010, 494−508. (449) Pareek, Y.; Ravikanth, M. Synthesis and Properties of Porphyrin-Expanded Porphyrin Click Dyads. Tetrahedron 2013, 69, 7943−7949. (450) Buckley, H. L.; Rubin, L. K.; Chromiński, M.; McNicholas, B. J.; Tsen, K. H. Y.; Gryko, D. T.; Arnold, J. Corroles That “Click”: Modular Synthesis of Azido- and Propargyl-Functionalized Metallocorrole Complexes and Convergent Synthesis of a Bis-corrole Scaffold. Inorg. Chem. 2014, 53, 7941−7950. (451) Nikolaou, V.; Angaridis, P. A.; Charalambidis, G.; Sharma, G. D.; Coutsolelos, A. G. A ″Click-Chemistry″ Approach for the Synthesis of Porphyrin Dyads as Sensitizers for Dye-Sensitized Solar Cells. Dalton Trans. 2015, 44, 1734−1747. (452) Takai, A.; Chkounda, M.; Eggenspiller, A.; Gros, C. P.; Lachkar, M.; Barbe, J.-M.; Fukuzumi, S. Efficient Photoinduced Electron Transfer in a Porphyrin Tripod−Fullerene Supramolecular Complex via π−π Interactions in Nonpolar Media. J. Am. Chem. Soc. 2010, 132, 4477−4489. (453) Yaschuk, Y. P.; Tyurin, V. S.; Beletskaya, I. P. Trimer Porphyrin Star. Makrogeterotsikly 2012, 5, 302−307. (454) Polevaya, Y. P.; Tyurin, V. S.; Beletskaya, I. P. Linear Conjuncted Porphyrin Trimer Synthesis via ″Click″ Reaction. J. Porphyrins Phthalocyanines 2014, 18, 20−34. (455) Brégier, F.; Aly, S. M.; Gros, C. P.; Barbe, J.-M.; Rousselin, Y.; Harvey, P. D. Design and Photophysical Properties of Zinc(II) Porphyrin-Containing Dendrons Linked to a Central Artificial Special Pair. Chem. - Eur. J. 2011, 17, 14643−14662. (456) Nguyen, N. T.; Hofkens, J.; Scheblykin, I. G.; Kruk, M.; Dehaen, W. Click Reaction Synthesis and Photophysical Studies of Dendritic Metalloporphyrins. Eur. J. Org. Chem. 2014, 2014, 1766− 1777. (457) Le Pleux, L.; Pellegrin, Y.; Blart, E.; Odobel, F.; Harriman, A. Long-Lived, Charge-Shift States in Heterometallic, Porphyrin-Based DZ

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Dendrimers Formed via Click Chemistry. J. Phys. Chem. A 2011, 115, 5069−5080. (458) Zhang, J.; Li, Y.; Yang, W.; Lai, S.-W.; Zhou, C.; Liu, H.; Che, C.-M.; Li, Y. A Smart Porphyrin Cage For Recognizing Azide Anions. Chem. Commun. 2012, 48, 3602−3604. (459) Yoon, H.; Lim, J. M.; Gee, H.-C.; Lee, C.-H.; Jeong, Y.-H.; Kim, D.; Jang, W.-D. A Porphyrin-Based Molecular Tweezer: GuestInduced Switching of Forward and Backward Photoinduced Energy Transfer. J. Am. Chem. Soc. 2014, 136, 1672−1679. (460) Harriman, A.; Elliott, K. J.; Alamiry, M. A. H.; Pleux, L. L.; Séverac, M.; Pellegrin, Y.; Blart, E.; Fosse, C.; Cannizzo, C.; Mayer, C. R.; et al. Intramolecular Electron Transfer Reactions Observed for Dawson-Type Polyoxometalates Covalently Linked to Porphyrin Residues. J. Phys. Chem. C 2009, 113, 5834−5842. (461) Fiset, E.; Morin, J.-F. Synthesis, Characterization and Modification of Azide-Containing Dendronized Diblock Copolymers. Polymer 2009, 50, 1369−1377. (462) Fu, G.-D.; Jiang, H.; Yao, F.; Xu, L.-Q.; Ling, J.; Kang, E.-T. Preparation of Fluorescent Organometallic Porphyrin Complex Nanogels of Controlled Molecular Structure via Reverse-Emulsion Click Chemistry. Macromol. Rapid Commun. 2012, 33, 1523−1527. (463) Collin, J.-P.; Frey, J.; Heitz, V.; Sauvage, J.-P.; Tock, C.; Allouche, L. Adjustable Receptor Based on a [3]Rotaxane Whose Two Threaded Rings Are Rigidly Attached to Two Porphyrinic Plates: Synthesis and Complexation Studies. J. Am. Chem. Soc. 2009, 131, 5609−5620. (464) Brizet, B.; Desbois, N.; Bonnot, A.; Langlois, A.; Dubois, A.; Barbe, J.-M.; Gros, C. P.; Goze, C.; Denat, F.; Harvey, P. D. Slow and Fast Singlet Energy Transfers in BODIPY−Gallium(III) Corrole Dyads Linked by Flexible Chains. Inorg. Chem. 2014, 53, 3392−3403. (465) Ganapathi, E.; Madhu, S.; Ravikanth, M. Synthesis and Properties of Triazole Bridged BODIPY-Conjugates. Tetrahedron 2014, 70, 664−671. (466) Fazio, M. A.; Lee, O. P.; Schuster, D. I. First Triazole-Linked Porphyrin−Fullerene Dyads. Org. Lett. 2008, 10, 4979−4982. (467) de Miguel, G.; Wielopolski, M.; Schuster, D. I.; Fazio, M. A.; Lee, O. P.; Haley, C. K.; Ortiz, A. L.; Echegoyen, L.; Clark, T.; Guldi, D. M. Triazole Bridges as Versatile Linkers in Electron Donor− Acceptor Conjugates. J. Am. Chem. Soc. 2011, 133, 13036−13054. (468) Iehl, J.; Pereira de Freitas, R.; Delavaux-Nicot, B.; Nierengarten, J.-F. Click Chemistry for the Efficient Preparation of Functionalized [60]Fullerene Hexakis-Adducts. Chem. Commun. 2008, 2450−2452. (469) Hirsch, A.; Lamparth, I.; Groesser, T.; Karfunkel, H. R. Regiochemistry of Multiple Additions to the Fullerene Core: Synthesis of a Th-Symmetric Hexakis Adduct of C60 with Bis(ethoxycarbonyl)methylene. J. Am. Chem. Soc. 1994, 116, 9385−9386. (470) Helmreich, M.; Ermilov, E. A.; Meyer, M.; Jux, N.; Hirsch, A.; Röder, B. Dissipation of Electronic Excitation Energy within a C60 [6:0]-Hexaadduct Carrying 12 Pyropheophorbide a Moieties. J. Am. Chem. Soc. 2005, 127, 8376−8385. (471) Diekers, M.; Luo, C.; Guldi, D. M.; Hirsch, A. Th-Symmetrical Hexakisadducts of C60 with a Densely Packed π-Donor Shell Can Act as Energy- or Electron-Transducing Systems. Chem. - Eur. J. 2002, 8, 979−991. (472) Klumpp, C.; Lacerda, L.; Chaloin, O.; Ros, T. D.; Kostarelos, K.; Prato, M.; Bianco, A. Multifunctionalised Cationic Fullerene Adducts for Gene Transfer: Design, Synthesis and DNA Complexation. Chem. Commun. 2007, 3762−3764. (473) Li, H.; Kitaygorodskiy, A.; Carino, R. A.; Sun, Y.-P. Simple Modification in Hexakis-Addition for Efficient Synthesis of C60Centered Dendritic Molecules Bearing Multiple Aromatic Chromophores. Org. Lett. 2005, 7, 859−861. (474) Li, H.; Haque, S. A.; Kitaygorodskiy, A.; Meziani, M. J.; TorresCastillo, M.; Sun, Y.-P. Alternatively Modified Bingel Reaction for Efficient Syntheses of C60 Hexakis-Adducts. Org. Lett. 2006, 8, 5641− 5643. (475) Iehl, J.; Osinska, I.; Louis, R.; Holler, M.; Nierengarten, J.-F. A Stable Fullerene-Azide Building Block for the Construction of a

Fullerene−Porphyrin Conjugate. Tetrahedron Lett. 2009, 50, 2245− 2248. (476) Gmehling, A.; Hirsch, A. Facile Access to Functional Building Blocks of C60 Involving C3-Symmetrical Addition Patterns. Eur. J. Org. Chem. 2013, 2013, 5093−5105. (477) Liu, J.-Y.; El-Khouly, M. E.; Fukuzumi, S.; Ng, D. K. P. Mimicking Photosynthetic Antenna-Reaction-Center Complexes with a (Boron Dipyrromethene)3−Porphyrin−C60 Pentad. Chem. - Eur. J. 2011, 17, 1605−1613. (478) Liu, J.-Y.; Ermilov, E. A.; Röder, B.; Ng, D. K. P. A LightHarvesting Porphyrin-Boron Dipyrromethene Conjugate. J. Porphyrins Phthalocyanines 2013, 17, 831−835. (479) Ho, K.-H. L.; Hijazi, I.; Rivier, L.; Gautier, C.; Jousselme, B.; de Miguel, G.; Romero-Nieto, C.; Guldi, D. M.; Heinrich, B.; Donnio, B.; Campidelli, S. Host−Guest Complexation of [60]Fullerenes and Porphyrins Enabled by “Click Chemistry. Chem. - Eur. J. 2013, 19, 11374−11381. (480) Megiatto, J. D.; Spencer, R.; Schuster, D. I. Efficient One-Pot Synthesis of Rotaxanes Bearing Electron Donors and [60]Fullerene. Org. Lett. 2009, 11, 4152−4155. (481) Megiatto, J. D.; Schuster, D. I.; Abwandner, S.; de Miguel, G.; Guldi, D. M. [2]Catenanes Decorated with Porphyrin and [60]Fullerene Groups: Design, Convergent Synthesis, and Photoinduced Processes. J. Am. Chem. Soc. 2010, 132, 3847−3861. (482) Megiatto, J. D.; Schuster, D. I.; de Miguel, G.; Wolfrum, S.; Guldi, D. M. Topological and Conformational Effects on Electron Transfer Dynamics in Porphyrin−[60]Fullerene Interlocked Systems. Chem. Mater. 2012, 24, 2472−2485. (483) Kirner, S. V.; Guldi, D. M.; Megiatto, J. J. D.; Schuster, D. I. Synthesis and Photophysical Properties of New Catenated Electron Donor−Acceptor Materials with Magnesium and Free Base Porphyrins as Donors and C60 as the Acceptor. Nanoscale 2015, 7, 1145−1160. (484) Palacin, T.; Khanh, H. L.; Jousselme, B.; Jegou, P.; Filoramo, A.; Ehli, C.; Guldi, D. M.; Campidelli, S. Efficient Functionalization of Carbon Nanotubes with Porphyrin Dendrons via Click Chemistry. J. Am. Chem. Soc. 2009, 131, 15394−15402. (485) Liu, X.-M.; Thakur, A.; Wang, D. Efficient Synthesis of Linear Multifunctional Poly(ethylene glycol) by Copper(I)-Catalyzed Huisgen 1,3-Dipolar Cycloaddition. Biomacromolecules 2007, 8, 2653− 2658. (486) Le Ho, K. H.; Rivier, L.; Jousselme, B.; Jegou, P.; Filoramo, A.; Campidelli, S. Zn-porphyrin/Zn-phthalocyanine dendron for SWNT functionalisation. Chem. Commun. 2009, 46, 8731−8733. (487) Hijazi, I.; Jousselme, B.; Jegou, P.; Filoramo, A.; Campidelli, S. Formation of Linear and Hyperbranched Porphyrin Polymers on Carbon Nanotubes via a CuAAC “Grafting From” Approach. J. Mater. Chem. 2012, 22, 20936−20942. (488) Flavin, K.; Chaur, M. N.; Echegoyen, L.; Giordani, S. Functionalization of Multilayer Fullerenes (Carbon Nano-Onions) using Diazonium Compounds and “Click” Chemistry. Org. Lett. 2010, 12, 840−843. (489) Maeda, C.; Yamaguchi, S.; Ikeda, C.; Shinokubo, H.; Osuka, A. Dimeric Assemblies from 1,2,3-Triazole-Appended Zn(II) Porphyrins with Control of NH-Tautomerism in 1,2,3-Triazole. Org. Lett. 2008, 10, 549−552. (490) Fathalla, M.; Barnes, J. C.; Young, R. M.; Hartlieb, K. J.; Dyar, S. M.; Eaton, S. W.; Sarjeant, A. A.; Co, D. T.; Wasielewski, M. R.; Stoddart, J. F. Photoinduced Electron Transfer within a Zinc Porphyrin−Cyclobis(paraquat-p-phenylene) Donor−Acceptor Dyad. Chem. - Eur. J. 2014, 20, 14690−14697. (491) Toganoh, M.; Harada, H.; Ikawa, Y.; Furuta, H. Self-Assembly of Zn(II) Porphyrin−1,2,3-Triazole Conjugate with Alcohol Glue. Chem. Lett. 2010, 39, 252−253. (492) Sasaki, S.; Mizutani, K.; Kunieda, M.; Tamiaki, H. Construction of Chlorophyll Assemblies Based on Zinc Complexes of Triazole Chlorin Conjugates. Tetrahedron 2012, 68, 7133−7139. (493) Sasaki, S.; Mizutani, K.; Kunieda, M.; Tamiaki, H. Cycloaddition to a C3-Ethynylated Chlorophyll Derivative and SelfEA

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Aggregation of Zinc Chlorin−Pyrazole/Triazole Conjugates. Tetrahedron 2013, 69, 9772−9778. (494) Roberts, D. A.; Schmidt, T. W.; Crossley, M. J.; Perrier, S. Tunable Self-Assembly of Triazole-Linked Porphyrin−Polymer Conjugates. Chem. - Eur. J. 2013, 19, 12759−12770. (495) Webb, J. E. A.; Maharaj, F.; Blake, I. M.; Crossley, M. J. A ‘Click’ Chemistry Route to ‘Capped’ Porphyrins. Synlett 2008, 2008, 2147−2149. (496) Almog, J.; Baldwin, J. E.; Dyer, R. L.; Peters, M. Condensation of tetraaldehydes with pyrrole. Direct synthesis of capped porphyrins. J. Am. Chem. Soc. 1975, 97, 226−227. (497) Almog, J.; Baldwin, J. E.; Crossley, M. J.; Debernardis, J. F.; Dyer, R. L.; Huff, J. R.; Peters, M. K. Synthesis of “Capped Porphyrins. Tetrahedron 1981, 37, 3589−3601. (498) Samanta, S.; Sengupta, K.; Mittra, K.; Bandyopadhyay, S.; Dey, A. Selective Four Electron Reduction of O2 by an Iron Porphyrin Electrocatalyst under Fast and Slow Electron Fluxes. Chem. Commun. 2012, 48, 7631−7633. (499) Samanta, S.; Mittra, K.; Sengupta, K.; Chatterjee, S.; Dey, A. Second Sphere Control of Redox Catalysis: Selective Reduction of O2 to O2− or H2O by an Iron Porphyrin Catalyst. Inorg. Chem. 2013, 52, 1443−1453. (500) Kumar, D.; Mishra, B. A.; Chandra Shekar, K. P.; Kumar, A.; Akamatsu, K.; Kurihara, R.; Ito, T. Novel Porphyrin-Psoralen Conjugates: Synthesis, DNA Interaction and Cytotoxicity Studies. Org. Biomol. Chem. 2013, 11, 6675−6679. (501) Kumar, D.; Mishra, B. A.; Chandra Shekar, K. P.; Kumar, A.; Akamatsu, K.; Kusaka, E.; Ito, T. Remarkable Photocytotoxicity of a Novel Triazole-Linked Cationic Porphyrin−β-Carboline Conjugate. Chem. Commun. 2013, 49, 683−685. (502) Santos, F. d. C.; Cunha, A. C.; de Souza, M. C. B. V.; Tomé, A. C.; Neves, M. G. P. M. S.; Ferreira, V. F.; Cavaleiro, J. A. S. Synthesis of Porphyrin−Quinolone Conjugates. Tetrahedron Lett. 2008, 49, 7268−7270. (503) Monnereau, C.; Rebilly, J.-N.; Reinaud, O. Synthesis and First Studies of the Host−Guest and Substrate Recognition Properties of a Porphyrin-Tethered Calix[6]arene Ditopic Ligand. Eur. J. Org. Chem. 2011, 2011, 166−175. (504) Wirotius, A.-L.; Ibarboure, E.; Scarpantonio, L.; Schappacher, M.; McClenaghan, N. D.; Deffieux, A. Hydrosoluble dendritic poly(ethylene oxide)s with zinc tetraphenylporphyrin branching points as photosensitizers. Polym. Chem. 2013, 4, 1903−1912. (505) Unterlass, M. M.; Espinosa, E.; Boisson, F.; D’Agosto, F.; Boisson, C.; Ariga, K.; Khalakhan, I.; Charvet, R.; Hill, J. P. Polyethylenes bearing a terminal porphyrin group. Chem. Commun. 2011, 47, 7057−7059. (506) Bakleh, M. E.; Sol, V.; Estieu-Gionnet, K.; Granet, R.; Déléris, G.; Krausz, P. An Efficient Route to VEGF-Like Peptide Porphyrin Conjugates via Microwave-Assisted ‘Click-Chemistry’. Tetrahedron 2009, 65, 7385−7392. (507) Chinnusamy, T.; Rodionov, V.; Kühn, F. E.; Reiser, O. Update: An Efficient Synthesis of Poly(ethylene glycol)-Supported Iron(II) Porphyrin Using a Click Reaction and its Application for the Catalytic Olefination of Aldehydes. Adv. Synth. Catal. 2012, 354, 1827−1831. (508) Elmer, S. L.; Man, S.; Zimmerman, S. C. Synthesis of Polyglycerol, Porphyrin-Cored Dendrimers Using Click Chemistry. Eur. J. Org. Chem. 2008, 2008, 3845−3851. (509) Kimura, M.; Nakano, Y.; Adachi, N.; Tatewaki, Y.; Shirai, H.; Kobayashi, N. Intramolecular Axial Ligation of Zinc Porphyrin Cores with Triazole Links within Dendrimers. Chem. - Eur. J. 2009, 15, 2617−2624. (510) Hasegawa, T.; Umeda, M.; Numata, M.; Li, C.; Bae, A.-H.; Fujisawa, T.; Haraguchi, S.; Sakurai, K.; Shinkai, S. ‘Click Chemistry’ on Polysaccharides: a Convenient, General, and Monitorable Approach to Develop (1→3)-β-D-Glucans with Various Functional Appendages. Carbohydr. Res. 2006, 341, 35−40. (511) Hao, E.; Jensen, T. J.; Vicente, M. G. H. Synthesis of Porphyrin-Carbohydrate Conjugates Using ″Click″ Chemistry and

Their Preliminary Evaluation in Human Hep2 Cells. J. Porphyrins Phthalocyanines 2009, 13, 51−59. (512) Locos, O. B.; Heindl, C. C.; Corral, A.; Senge, M. O.; Scanlan, E. M. Efficient Synthesis of Glycoporphyrins by Microwave-Mediated “Click” Reactions. Eur. J. Org. Chem. 2010, 2010, 1026−1028. (513) Daly, R.; Vaz, G.; Davies, A. M.; Senge, M. O.; Scanlan, E. M. Synthesis and Biological Evaluation of a Library of Glycoporphyrin Compounds. Chem. - Eur. J. 2012, 18, 14671−14679. (514) Giuntini, F.; Bryden, F.; Daly, R.; Scanlan, E. M.; Boyle, R. W. Huisgen-Based Conjugation of Water-Soluble Porphyrins to Deprotected Sugars: Towards Mild Strategies for the Labelling of Glycans. Org. Biomol. Chem. 2014, 12, 1203−1206. (515) Garcia, G.; Naud-Martin, D.; Carrez, D.; Croisy, A.; Maillard, P. Microwave-Mediated ‘Click-Chemistry’ Synthesis of Glycoporphyrin Derivatives and In Vitro Photocytotoxicity for Application in Photodynamic Therapy. Tetrahedron 2011, 67, 4924−4932. (516) Mukosera, G. T.; Adams, T. P.; Rothbarth, R. F.; Langat, H.; Akanda, S.; Barkley, R. G.; Dolewski, R. D.; Ruppel, J. V.; Snyder, N. L. Synthesis of Glycosylated Zinc (II) 5,15-Diphenylporphyrin and Zinc (II) 5,10,15,20-Tetraphenylporphyrin Analogs Using Cu-Catalyzed Azide-Alkyne 1,3-Dipolar Cycloaddition Reactions. Tetrahedron Lett. 2015, 56, 73−77. (517) Ringot, C.; Sol, V.; Granet, R.; Krausz, P. Porphyrin-Grafted Cellulose Fabric: New Photobactericidal Material Obtained by “ClickChemistry” Reaction. Mater. Lett. 2009, 63, 1889−1891. (518) Feese, E.; Sadeghifar, H.; Gracz, H. S.; Argyropoulos, D. S.; Ghiladi, R. A. Photobactericidal Porphyrin-Cellulose Nanocrystals: Synthesis, Characterization, and Antimicrobial Properties. Biomacromolecules 2011, 12, 3528−3539. (519) Cecioni, S.; Matthews, S. E.; Blanchard, H.; Praly, J.-P.; Imberty, A.; Vidal, S. Synthesis of Lactosylated Glycoclusters and Inhibition Studies with Plant and Human Lectins. Carbohydr. Res. 2012, 356, 132−141. (520) Kushwaha, D.; Tiwari, V. K. Click Chemistry Inspired Synthesis of Glycoporphyrin Dendrimers. J. Org. Chem. 2013, 78, 8184−8190. (521) Liu, Y.; Ke, C.-F.; Zhang, H.-Y.; Cui, J.; Ding, F. Complexation-Induced Transition of Nanorod to Network Aggregates: Alternate Porphyrin and Cyclodextrin Arrays. J. Am. Chem. Soc. 2008, 130, 600−605. (522) Fathalla, M.; Li, S.-C.; Diebold, U.; Alb, A.; Jayawickramarajah, J. Water-Soluble Nanorods Self-Assembled via Pristine C60 and Porphyrin Moieties. Chem. Commun. 2009, 4209−4211. (523) Umezawa, N.; Matsumoto, N.; Iwama, S.; Kato, N.; Higuchi, T. Facile Synthesis of Peptide−Porphyrin Conjugates: Towards Artificial Catalase. Bioorg. Med. Chem. 2010, 18, 6340−6350. (524) Aly, S. M.; Guernon, H.; Guérin, B.; Harvey, P. D. Polyoligopeptides Functionalized Zinc(II) Porphyrins: Step towards Artificial Hemes. J. Porphyrins Phthalocyanines 2011, 15, 871−882. (525) Ma, D.; Liu, Z.-H.; Zheng, Q.-Q.; Zhou, X.-Y.; Zhang, Y.; Shi, Y.-F.; Lin, J.-T.; Xue, W. Star-Shaped Polymer Consisting of a Porphyrin Core and Poly(L-lysine) Dendron Arms: Synthesis, Drug Delivery, and In Vitro Chemo/Photodynamic Therapy. Macromol. Rapid Commun. 2013, 34, 548−552. (526) Ma, D.; Zhao, Y.; Zhou, X.-Y.; Lin, Q.-M.; Zhang, Y.; Lin, J.-T.; Xue, W. Photoenhanced Gene Transfection by a Star-Shaped Polymer Consisting of a Porphyrin Core and Poly(L-lysine) Dendron Arms. Macromol. Biosci. 2013, 13, 1221−1227. (527) Stephenson, A. W. I.; Partridge, A. C.; Filichev, V. V. Synthesis of β-Pyrrolic-Modified Porphyrins and Their Incorporation into DNA. Chem. - Eur. J. 2011, 17, 6227−6238. (528) Clave, G.; Chatelain, G.; Filoramo, A.; Gasparutto, D.; SaintPierre, C.; Le Cam, E.; Pietrement, O.; Guerineau, V.; Campidelli, S. Synthesis of A Multibranched Porphyrin-Oligonucleotide Scaffold for the Construction of DNA-Based Nano-Architectures. Org. Biomol. Chem. 2014, 12, 2778−2783. (529) Stephenson, A. W. I.; Bomholt, N.; Partridge, A. C.; Filichev, V. V. Significantly Enhanced DNA Thermal Stability Resulting from EB

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Porphyrin H-Aggregate Formation in the Minor Groove of the Duplex. ChemBioChem 2010, 11, 1833−1839. (530) Collman, J. P.; Zeng, L.; Wang, H. J. H.; Lei, A.; Brauman, J. I. Kinetics of (Porphyrin)manganese(III)-Catalyzed Olefin Epoxidation with a Soluble Iodosylbenzene Derivative. Eur. J. Org. Chem. 2006, 2006, 2707−2714. (531) McDonald, A. R.; Franssen, N.; van Klink, G. P. M.; van Koten, G. ‘Click’ Silica Immobilisation of Metallo-Porphyrin Complexes and Their Application in Epoxidation Catalysis. J. Organomet. Chem. 2009, 694, 2153−2162. (532) Devaraj, N. K.; Decreau, R. A.; Ebina, W.; Collman, J. P.; Chidsey, C. E. D. Rate of Interfacial Electron Transfer through the 1,2,3-Triazole Linkage. J. Phys. Chem. B 2006, 110, 15955−15962. (533) Huang, K.; Duclairoir, F.; Pro, T.; Buckley, J.; Marchand, G.; Martinez, E.; Marchon, J.-C.; De Salvo, B.; Delapierre, G.; Vinet, F. Ferrocene and Porphyrin Monolayers on Si(100) Surfaces: Preparation and Effect of Linker Length on Electron Transfer. ChemPhysChem 2009, 10, 963−971. (534) Liu, H.; Duclairoir, F.; Fleury, B.; Dubois, L.; Chenavier, Y.; Marchon, J.-C. Porphyrin Anchoring on Si(100) Using a β-Pyrrolic Position. Dalton Trans. 2009, 3793−3799. (535) Prasad, S.; Bhadbhade, M.; Thordarson, P. Tin(IV) Porphyrin Functionalization of Electrochemically Active Fluoride-Doped TinOxide (FTO) via Huisgen [3 + 2] Click Chemistry. J. Porphyrins Phthalocyanines 2011, 15, 75−82. (536) Nakazawa, J.; Smith, B. J.; Stack, T. D. P. Discrete Complexes Immobilized onto Click-SBA-15 Silica: Controllable Loadings and the Impact of Surface Coverage on Catalysis. J. Am. Chem. Soc. 2012, 134, 2750−2759. (537) Thandu, M.; Rapozzi, V.; Xodo, L.; Albericio, F.; Comuzzi, C.; Cavalli, S. ″Clicking″ Porphyrins to Magnetic Nanoparticles for Photodynamic Therapy. ChemPlusChem 2014, 79, 90−98. (538) Krawicz, A.; Palazzo, J.; Wang, G.-C.; Dinolfo, P. H. Layer-ByLayer Assembly of Zn(II) And Ni(II) 5,10,15,20-Tetra(4Ethynylphenyl)Porphyrin Multilayers on Au Using Copper Catalyzed Azide-Alkyne Cycloaddition. RSC Adv. 2012, 2, 7513−7522. (539) Palomaki, P. K. B.; Krawicz, A.; Dinolfo, P. H. Thickness, Surface Morphology, and Optical Properties of Porphyrin Multilayer Thin Films Assembled on Si(100) Using Copper(I)-Catalyzed Azide− Alkyne Cycloaddition. Langmuir 2011, 27, 4613−4622. (540) Palomaki, P. K. B.; Dinolfo, P. H. Structural Analysis of Porphyrin Multilayer Films on ITO Assembled Using Copper(I)Catalyzed Azide−Alkyne Cycloaddition by ATR IR. ACS Appl. Mater. Interfaces 2011, 3, 4703−4713. (541) Maeda, C.; Shinokubo, H.; Osuka, A. Synthesis of meso-5Azaindolyl-Appended Zn(II) Porphyrins via Pd-Catalyzed Annulation. Org. Lett. 2007, 9, 2493−2496. (542) Larock, R. C.; Yum, E. K. Synthesis of Indoles via PalladiumCatalyzed Heteroannulation of Internal Alkynes. J. Am. Chem. Soc. 1991, 113, 6689−6690. (543) Larock, R. C.; Yum, E. K.; Refvik, M. D. Synthesis of 2,3Disubstituted Indoles via Palladium-Catalyzed Annulation of Internal Alkynes. J. Org. Chem. 1998, 63, 7652−7662. (544) Maeda, C.; Shinokubo, H.; Osuka, A. meso,meso′-Bis(5azaindol-2-yl)-Appended meso−meso-Linked Zn(II) Diporphyrin: A Discrete Fluorescent Assembly. Org. Lett. 2009, 11, 5322−5325. (545) Maeda, C.; Shinokubo, H.; Osuka, A. Synthesis of meso,meso′Pyrrole-Bridged Diporphyrins by Cu(I)-Mediated Annulation. Org. Lett. 2010, 12, 1820−1823. (546) Dötz, K. H. Synthesis of the Naphthol Skeleton from Pentacarbonyl[methoxy(phenyl)carbene]chromium(0) and Tolan. Angew. Chem., Int. Ed. Engl. 1975, 14, 644−645. (547) Wulff, W. D.; Tang, P. C.; McCallum, J. S. Regiochemistry of the Reaction of Chromium-Carbene Complexes with Acetylenes. J. Am. Chem. Soc. 1981, 103, 7677−7678. (548) Geis, O.; Schmalz, H.-G. New Developments in the PausonKhand Reaction. Angew. Chem., Int. Ed. 1998, 37, 911−914. (549) Gibson, S. E.; Stevenazzi, A. The Pauson-Khand Reaction: the Catalytic Age Is Here! Angew. Chem., Int. Ed. 2003, 42, 1800−1810.

(550) Horn, S.; Senge, M. O. The Intermolecular Pauson−Khand Reaction of meso-Substituted Porphyrins. Eur. J. Org. Chem. 2008, 2008, 4881−4890. (551) Shinokubo, H. Synthesis of Novel Porphyrinoids through Organometallic Means. Yuki Gosei Kagaku Kyokaishi 2014, 72, 149− 157. (552) Shinokubo, H. Synthesis of Novel Porphyrinoids from Dipyrrins. In Chemical Science of π-Electron Systems; Akasaka, T., Osuka, A., Fukuzumi, S., Kandori, H., Aso, Y., Eds.; Springer: Tokyo, 2015; pp 243−255. (553) Arnold, L.; Norouzi-Arasi, H.; Wagner, M.; Enkelmann, V.; Müllen, K. A Porphyrin-Related Macrocycle from Carbazole and Pyridine Building Blocks: Synthesis and Metal Coordination. Chem. Commun. 2011, 47, 970−972. (554) Wackerlin, C.; Maldonado, P.; Arnold, L.; Shchyrba, A.; Girovsky, J.; Nowakowski, J.; Ali, M. E.; Hahlen, T.; Baljozovic, M.; Siewert, D.; et al. Magnetic Exchange Coupling of a Synthetic Co(II)Complex to a Ferromagnetic Ni Substrate. Chem. Commun. 2013, 49, 10736−10738. (555) Arnold, L.; Baumgarten, M.; Müllen, K. A CarbazoleContaining Porphyrinoid: Synthesis and Oxidation to the PorphyrinState. Chem. Commun. 2012, 48, 9640−9642. (556) Maeda, C.; Yoneda, T.; Aratani, N.; Yoon, M.-C.; Lim, J. M.; Kim, D.; Yoshioka, N.; Osuka, A. Synthesis of Carbazole-Containing Porphyrinoids by a Multiple Annulation Strategy: A Core-Modified and π-Expanded Porphyrin. Angew. Chem., Int. Ed. 2011, 50, 5691− 5694. (557) Maeda, C.; Yoshioka, N. Peripherally Ethynylated CarbazoleBased Core-Modified Porphyrins. Org. Biomol. Chem. 2012, 10, 5182− 5185. (558) Maeda, C.; Masuda, M.; Yoshioka, N. Effective π-Extension of Carbazole-Based Thiaporphyrins by Peripheral Phenylethynyl Substituents. Org. Lett. 2013, 15, 3566−3569. (559) Maeda, C.; Yoshioka, N. Synthesis and Characterization of Novel Fused Porphyrinoids Based on Cyclic Carbazole[2]indolones. Org. Lett. 2012, 14, 2122−2125. (560) Masuda, M.; Maeda, C. Synthesis and Characterization of Carbazole-Based Expanded Thiaporphyrins. Chem. - Eur. J. 2013, 19, 2971−2975. (561) Masuda, M.; Maeda, C.; Yoshioka, N. Synthesis of CarbazoleBased Selenaporphyrin via Annulation. Org. Lett. 2013, 15, 578−581. (562) Maeda, C.; Masuda, M.; Yoshioka, N. Synthesis of CarbazoleBased Hetero-Core-Modified Porphyrins. Org. Biomol. Chem. 2014, 12, 2656−2662. (563) Nakamura, S.; Hiroto, S.; Shinokubo, H. Synthesis and Oxidation of Cyclic Tetraindole. Chem. Sci. 2012, 3, 524−527. (564) Nakamura, S.; Kondo, T.; Hiroto, S.; Shinokubo, H. Porphyrin Analogues That Consist of Indole, Benzofuran, and Benzothiophene Subunits. Asian J. Org. Chem. 2013, 2, 312−319. (565) Chen, R.; Bhadbhade, M.; Kumar, N.; Black, D. S. Synthesis of Macrocyclic Tetraindolyls via Oxidative Coupling Reactions. Tetrahedron Lett. 2012, 53, 3337−3341. (566) Wood, T. E.; Thompson, A. Advances in the Chemistry of Dipyrrins and Their Complexes. Chem. Rev. 2007, 107, 1831−1861. (567) Loudet, A.; Burgess, K. BODIPY Dyes and Their Derivatives: Syntheses and Spectroscopic Properties. Chem. Rev. 2007, 107, 4891− 4932. (568) Ulrich, G.; Ziessel, R.; Harriman, A. The Chemistry of Fluorescent Bodipy Dyes: Versatility Unsurpassed. Angew. Chem., Int. Ed. 2008, 47, 1184−1201. (569) Ni, Y.; Wu, J. Far-Red and Near Infrared BODIPY Dyes: Synthesis and Applications for Fluorescent pH Probes and BioImaging. Org. Biomol. Chem. 2014, 12, 3774−3791. (570) Lu, H.; Mack, J.; Yang, Y.; Shen, Z. Structural Modification Strategies for the Rational Design of Red/NIR Region BODIPYs. Chem. Soc. Rev. 2014, 43, 4778−4823. (571) Sakida, T.; Yamaguchi, S.; Shinokubo, H. Metal-Mediated Synthesis of Antiaromatic Porphyrinoids from a BODIPY Precursor. Angew. Chem., Int. Ed. 2011, 50, 2280−2283. EC

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(572) Johnson, A. W.; Kay, I. T.; Rodrigo, R. 2,2′-Bipyrrolic Macrocyclic Ring Systems. J. Chem. Soc. 1963, 0, 2336−2342. (573) 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, 5919−5923. (574) Matano, Y.; Shibano, T.; Nakano, H.; Imahori, H. Nickel(II) and Copper(II) Complexes of β-Unsubstituted 5,15-Diazaporphyrins and Pyridazine-Fused Diazacorrinoids: Metal-Template Syntheses and Peripheral Functionalizations. Chem. - Eur. J. 2012, 18, 6208−6216. (575) Matano, Y.; Shibano, T.; Nakano, H.; Kimura, Y.; Imahori, H. Free Base and Metal Complexes of 5,15-Diaza-10,20-Dimesitylporphyrins: Synthesis, Structures, Optical and Electrochemical Properties, and Aromaticities. Inorg. Chem. 2012, 51, 12879−12890. (576) Satoh, T.; Minoura, M.; Nakano, H.; Furukawa, K.; Matano, Y. Redox-Switchable 20π-, 19π-, and 18π-Electron 5,10,15,20-Tetraaryl5,15-Diazaporphyrinoid Nickel(II) Complexes. Angew. Chem., Int. Ed. 2016, 55, 2235−2238. (577) Kamiya, H.; Kondo, T.; Sakida, T.; Yamaguchi, S.; Shinokubo, H. meso-Thiaporphyrinoids Revisited: Missing of Sulfur by Small Metals. Chem. - Eur. J. 2012, 18, 16129−16135. (578) Wachi, N.; Kondo, T.; Ito, S.; Hiroto, S.; Shin, J. Y.; Shinokubo, H. Synthesis, Reactivity and Property of 5,15-Dithiaporphyrin Copper(II) Complex. J. Porphyrins Phthalocyanines 2014, 18, 675−678. (579) 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, 3540−3543. (580) Omori, H.; Hiroto, S.; Shinokubo, H. Synthesis of Free-Base 10-Azacorroles. Org. Lett. 2016, 18, 2978−2981. (581) Bröring, M.; Köhler, S.; Kleeberg, C. Norcorrole: Observation of the Smallest Porphyrin Variant with a N4 Core. Angew. Chem., Int. Ed. 2008, 47, 5658−5660. (582) 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, 8542−8545. (583) Kende, A. S.; Liebeskind, L. S.; Braitsch, D. M. In Situ Generation of a Solvated Zerovalent Nickel Reagent. Biaryl Formation. Tetrahedron Lett. 1975, 16, 3375−3378. (584) Shin, J.-Y.; Yamada, T.; Yoshikawa, H.; Awaga, K.; Shinokubo, H. An Antiaromatic Electrode-Active Material Enabling High Capacity and Stable Performance of Rechargeable Batteries. Angew. Chem., Int. Ed. 2014, 53, 3096−3101. (585) Fukuoka, T.; Uchida, K.; Sung, Y. M.; Shin, J.-Y.; Ishida, S.; Lim, J. M.; Hiroto, S.; Furukawa, K.; Kim, D.; Iwamoto, T.; Shinokubo, H. Near-IR Absorbing Nickel(II) Porphyrinoids Prepared by Regioselective Insertion of Silylenes into Antiaromatic Nickel(II) Norcorrole. Angew. Chem., Int. Ed. 2014, 53, 1506−1509. (586) Nozawa, R.; Yamamoto, K.; Shin, J.-Y.; Hiroto, S.; Shinokubo, H. Regioselective Nucleophilic Functionalization of Antiaromatic Nickel(II) Norcorroles. Angew. Chem., Int. Ed. 2015, 54, 8454−8457. (587) Nozawa, R.; Yamamoto, K.; Hisaki, I.; Shin, J.-Y.; Shinokubo, H. Ni(II) Tetrahydronorcorroles: Antiaromatic Porphyrinoids with Saturated Pyrrole Units. Chem. Commun. 2016, 52, 7106−7109. (588) Liu, B.; Li, X.; Stępień, M.; Chmielewski, P. J. Towards Norcorrin: Hydrogenation Chemistry and the Heterodimerization of Nickel(II) Norcorrole. Chem. - Eur. J. 2015, 21, 7790−7797. (589) Deng, Z.; Li, X.; Stępień, M.; Chmielewski, P. J. Nitration of Norcorrolatonickel(II): First Observation of a Diatropic Current in a System Comprising a Norcorrole Ring. Chem. - Eur. J. 2016, 22, 4231−4246. (590) Kido, H.; Shin, J.-Y.; Shinokubo, H. Selective Synthesis of a [32]Octaphyrin(1.0.1.0.1.0.1.0) Bis(palladium) Complex by a MetalTemplated Strategy. Angew. Chem., Int. Ed. 2013, 52, 13727−13730. (591) Umetani, M.; Tanaka, T.; Kim, T.; Kim, D.; Osuka, A. Double Ring Expansion from an Aromatic [18]Porphyrin(1.1.1.1) to an Antiaromatic [20]Porphyrin(2.1.2.1). Angew. Chem., Int. Ed. 2016, 55, 8095−8099. ED

DOI: 10.1021/acs.chemrev.6b00427 Chem. Rev. XXXX, XXX, XXX−XXX