02.03.2013 Views

cpgp3.SmithVicente.QAY 1081..12

cpgp3.SmithVicente.QAY 1081..12

cpgp3.SmithVicente.QAY 1081..12

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

a17.8 Product Class 8:<br />

Porphyrins and Related Compounds<br />

K. M. Smith and M. G. H. Vicente<br />

General Introduction<br />

Previously published information regarding this product class can be found in Houben–<br />

Weyl, Vol. 4/1b, p 828 and Vol. E 9d, pp 577–833.<br />

Porphyrins and related compounds, such as chlorins, are tetrapyrrole systems which<br />

provide the basic chromophore for a host of biologically important natural products.<br />

Principal examples are the hemes (iron complexes found in hemoglobins, myoglobins,<br />

cytochromes, catalases, and peroxidases) and the chlorophylls and bacteriochlorophylls<br />

(dihydro- and tetrahydroporphyrin–magnesium complexes found in photosynthetic organisms).<br />

There also exist a number of nonnatural contracted and expanded porphyrin<br />

systems, which have been invented as vehicles for comparative investigation of the effects<br />

of ring size on the chemistry and spectroscopy of porphyrins. Examples of contracted<br />

systems are the corroles, which have one of the porphyrin interpyrrolic carbon atoms<br />

missing, and sapphyrins and pentaphyrins, both of which have no less than five pyrrole<br />

subunits within their chromophores.<br />

The field has been extensively reviewed in the very recent past; The Porphyrin Handbook<br />

(published in 2000) is a 10-volume, 3500-page, 61-chapter treatise dealing with all aspects<br />

of porphyrins and related compounds; [1] volumes 11 through 20 were published in<br />

2003. Prior to these, two versions of Porphyrins and Metalloporphyrins were published (in<br />

1975 [2] and 1964 [3] ) and a multivolume series, The Porphyrins, appeared between 1977 and<br />

1979. [4] From time to time, reviews of specific areas of the porphyrin research field have<br />

appeared, and four which are relevant to the present chapter can be found in Rodd s Chemistry<br />

of the Carbon Compounds, [5,6] and Total Syntheses of Natural Products. [7,8]<br />

There are two systems currently in use for nomenclature of porphyrins. These are<br />

the IUPAC-approved system [9] and the so-called Fischer system; both are shown in<br />

Scheme 1.<br />

Scheme 1 The IUPAC and Fischer Systems of Nomenclature of Porphyrins<br />

23<br />

2 2<br />

2 1<br />

2<br />

3<br />

A<br />

4<br />

5<br />

6<br />

7<br />

B 8<br />

1 N<br />

21<br />

N<br />

22<br />

9<br />

20<br />

10<br />

19<br />

24<br />

N<br />

23<br />

N<br />

11<br />

18 D C 12<br />

17<br />

16<br />

15<br />

14<br />

13<br />

IUPAC<br />

N N<br />

N N<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

1<br />

δ<br />

2 α 3<br />

The IUPAC nomenclature system assigns every carbon atom in the cyclic chromophore<br />

with a unique number from 1–20. The nitrogen atoms are numbered 21–24. The Fischer<br />

system, which was used in almost all of the rich historical literature before 1960, and is<br />

still used to a certain extent in the North American literature, numbers the carbon at the<br />

pyrrole b-positions 1–8, and labels the four interpyrrolic carbons (the so-called meso-carbons)<br />

as , b, g, d. While one can see that the assignment in the IUPAC system of a number<br />

to every single carbon atom in the chromophore would be advantageous, for example in<br />

8<br />

7<br />

γ<br />

Fischer<br />

6<br />

4<br />

β<br />

5<br />

1081<br />

for references see p 1223


1082 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

abiosynthetic studies, the earlier Fischer system carried with it a very convenient subnomenclature<br />

of trivial names; whether or not the IUPAC system is being used by an investigator,<br />

these trivial names are still used in abundance! The examples protoheme IX<br />

(“heme”, 1), protoporphyrin IX (2), coproporphyrin III (3), uroporphyrin I (4), chlorophyll<br />

a (5, R 2 = Me), chlorophyll b (5, R 2 = CHO), chlorin e 6 (6, R 1 = Me), and rhodin g 7 (6,<br />

R 1 = CHO) are shown in Scheme 2.<br />

Scheme 2 Typical Natural Porphyrin Derivatives<br />

HO 2C<br />

HO 2C<br />

HO 2C<br />

5 R 1 =<br />

N N<br />

Fe<br />

N N<br />

1<br />

NH N<br />

N HN<br />

R 1 O 2C<br />

R 2 = Me, CHO<br />

3<br />

CO2H<br />

CO2H<br />

N N<br />

Mg<br />

N N<br />

13 2<br />

MeO 2C<br />

CO 2H<br />

R 2<br />

E 13 1<br />

O<br />

Et<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

HO2C<br />

HO 2C<br />

HO 2C<br />

HO 2C<br />

HO 2C<br />

HO2C<br />

2<br />

NH N<br />

N HN<br />

4<br />

CO 2H<br />

CO 2H<br />

NH N<br />

N HN<br />

CO 2H<br />

CO2H<br />

R 1<br />

CO 2H<br />

CO2H<br />

CO2H 6 R 1 = Me, CHO<br />

Though Scheme 2 might appear to indicate otherwise, there is in fact a system hidden<br />

within Fischer s trivial nomenclature. Porphyrins that have the same two different substituents<br />

on each pyrrole subunit, for example, methyl and ethyl, can have a maximum<br />

of four so-called “primary type-isomers” (structures 7–10) as indicated in Roman<br />

numerals in Scheme 3.<br />

Et


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1083<br />

aScheme 3 The Four Primary Type-Isomers of the Etioporphyrins<br />

Et<br />

Et<br />

Et<br />

Et<br />

NH N<br />

I<br />

N HN<br />

7<br />

NH N<br />

III<br />

N HN<br />

9<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et Et<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

II<br />

8<br />

Et Et<br />

NH N<br />

IV<br />

N HN<br />

When the substituents on each pyrrole subunit are methyl and ethyl, the series is called<br />

“etioporphyrin”. If instead, the substituents are methyl and propanoic acid the series is<br />

“coproporphyrin” or if they are acetic and propanoic acids they are “uroporphyrins”.<br />

These names are historically derived. For example, coproporphyrins are not very watersoluble<br />

and are therefore excreted in the feces, while uroporphyrins are extremely water-soluble<br />

and are excreted in the urine. The normal metabolites of heme are of the “primary<br />

type-isomer III” orientation, though “type I” is occasionally found in some disease<br />

states. If one has three (rather than two) types of substituent spread around the chromophore,<br />

with one identical substituent (usually methyl) common to each pyrrole subunit,<br />

there are 15 possible type-isomers. Protoporphyrin IX is the natural “secondary type-isomer”<br />

in heme metabolism and this can be seen to be related to “primary type-isomer III”<br />

by further modification of the non-methyl substituents of rings A and B in coproporphyrin<br />

III in the latter (Scheme 2). “Proto” simply refers to the importance of this molecule; its<br />

protiodevinylated analogue (3,8-di-H in place of 3,8-divinyl) is named “deuteroporphyrin<br />

IX” because it was regarded as the second most important porphyrin, and contrary to the<br />

assumption of many journal editors, deuteroporphyrin contains no deuterium! The trivial<br />

nomenclature employed for chlorophyll a and b derivatives, such as chlorin e 6 and rhodin<br />

g 7, is unfathomable, and is best left alone (though it is worth knowing that the subscript<br />

Arabic numeral refers to the number of oxygen atoms in the molecule).<br />

Scheme 4 Structures of Porphyrin, Chlorin, Bacteriochlorin, and Isobacteriochlorin<br />

R 1<br />

R 8<br />

R 2 R 3<br />

R 7<br />

NH N<br />

N HN<br />

11<br />

R 6<br />

R 4<br />

R 5<br />

R 2<br />

R 1<br />

R 10<br />

10<br />

R4 R3 R 9<br />

N HN<br />

NH N<br />

12<br />

Et<br />

R 5<br />

R 8<br />

Et<br />

R 6<br />

R 7<br />

for references see p 1223


1084 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

R 2<br />

aN<br />

R<br />

HN<br />

1<br />

R 12<br />

R4 R3 R 11<br />

NH N<br />

13<br />

R 5<br />

R 6<br />

R 7<br />

R10 R9 R8 N N<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 2<br />

R 1<br />

R 12<br />

R4 R3 R 11<br />

14<br />

R6 R5<br />

Porphyrins 11 are aromatic macrocycles containing a total of 22 conjugated p-electrons.<br />

At any one time only 18 p-electrons participate in the delocalization pathway, two other<br />

double bonds being cross conjugated. Porphyrins conform with Hückel s [4n+2] rule for<br />

aromaticity (n = 4). Therefore, one or two of the peripheral double bonds of porphyrins<br />

can undergo addition reactions to form chlorins 12, bacteriochlorins 13, or isobacteriochlorins<br />

14 (shown in Scheme 4), without interfering with the delocalization pathway<br />

or introducing any major loss of aromatic character. Chlorins and bacteriochlorins are,<br />

however, progressively less aromatic than porphyrins qualitatively because fewer resonance<br />

structures can be drawn for the more reduced chromophores. The positions<br />

around the porphyrin periphery available, for example for electrophilic aromatic substitution,<br />

are the unsubstituted b-pyrrolic positions (at 2, 3, 7, 8, 12, 13, 17, and 18 in 11),<br />

and the four meso positions at 5, 10, 15, and 20. X-ray studies [10,11] indicate that the numerous<br />

C—C and the C—N bonds in 1 are almost equivalent. NMR spectroscopic, [12] X-ray, and<br />

theoretical studies [13,14] show that the thermodynamically most-favored tautomers for<br />

symmetrically substituted porphyrins are the degenerate trans-NH-tautomers depicted<br />

in 15 and 17 (Scheme 5). The mechanism for N–H migration between porphyrin trans-tautomers<br />

appears to proceed in a stepwise manner, via the less-favored cis-tautomers, such<br />

as 16. [15]<br />

The favored pathways for delocalization of the p-electrons in free-base porphyrins, in<br />

metal porphyrinates, and in porphyrin dications are also shown in Scheme 5. The peripheral<br />

b-positions of metal porphyrinates and porphyrin dications display similar reactivities,<br />

whereas the two opposite b–b¢ double bonds in free-base porphyrins experience<br />

more double-bond character. In the case of free-base chlorins, the most stable trans-tautomers<br />

(e.g., 18) appear to be those in which the imine-type nitrogens are on the reduced<br />

pyrrole subunit D and its opposite ring B. Though both forms of chlorin trans-tautomers<br />

18 and 19 have been observed experimentally, X-ray studies [10,11] and ab initio calculations<br />

[16] indicate that 18 is the major tautomer. The predominant p-electron delocalization<br />

pathway for free-base chlorins 18 causes the double bond of ring B to be the preferred<br />

site for electrophilic aromatic substitutions and addition reactions. In metal chlorinates,<br />

based on experimental observations, the preferred b-pyrrolic reaction sites are<br />

those in rings A and C, either side of the reduced pyrrole subunit. In metal porphyrinates,<br />

consideration of the various resonance structures leads to structure 20 as the resonance<br />

hybrid.<br />

Scheme 5 Tautomeric and Resonance Forms of Porphyrin, Metal Complex, Dication, and<br />

Chlorin<br />

NH N<br />

N HN<br />

15<br />

NH HN<br />

N N<br />

16<br />

R 10<br />

R 7<br />

R 8<br />

R 9<br />

N HN<br />

NH N<br />

17


A B<br />

aNH N<br />

17.8.1 Porphyrins 1085<br />

R 1<br />

R 2<br />

R 3 R 4<br />

N HN<br />

D C<br />

N N<br />

M<br />

N N<br />

N N<br />

M<br />

N N<br />

R 1<br />

R 2<br />

N N<br />

M<br />

N N<br />

N N<br />

M<br />

N N<br />

R 3 R 4<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

18<br />

N N<br />

M<br />

N N<br />

N N<br />

M<br />

N N<br />

R 1<br />

R 2<br />

R 3 R 4<br />

N HN<br />

NH N<br />

19<br />

N N<br />

M<br />

N N<br />

20 M = metal, 2H +<br />

Although they have been shown to be more electron-rich than the b-positions, the meso<br />

positions of the porphyrin system 11 are often sterically hindered, especially when two of<br />

the abutting b-positions are occupied by substituents. The b-pyrrolic positions, therefore,<br />

are sterically favored and tend to undergo electrophilic substitution and addition reactions.<br />

In metal-free porphyrins, the central nitrogen atoms readily react with electrophilic<br />

reagents and are easily protonated and metalated. In metal porphyrinates the central<br />

nitrogen atoms play an important role (both chemically and biologically) in the transmission<br />

of electronic information from the metal ion and its axial ligands to the porphyrin psystem,<br />

and vice versa. In chlorins, the meso positions adjacent to the reduced (nonaromatic)<br />

pyrrole unit are more electron rich and are correspondingly more reactive toward<br />

electrophilic reagents. [17]<br />

17.8.1 Product Subclass 1:<br />

Porphyrins<br />

There are many methods available for the synthesis of porphyrins, and these have been<br />

reviewed in the recent past. [18–21] The synthetic route to be used should depend upon the<br />

symmetry and the complexity of the various peripheral substituents on the target porphyrin.<br />

For example, it would be wasteful in time and resources to attempt the synthesis<br />

of 2,3,7,8,12,13,17,18-octaethylporphyrin (OEP, 21; Scheme 6) by way of a laborious multistep<br />

fabrication of an open-chain tetrapyrrolic intermediate, followed by cyclization to<br />

produce the porphyrin macrocycle; symmetrically substituted compounds such as 21 are<br />

most efficiently synthesized by polymerization of a suitable monopyrrole. On the other<br />

hand, without help from enzymes which participate in the biosynthesis of heme and<br />

chlorophylls, there is no possibility that protoporphyrin IX (2) could be synthesized by a<br />

monopyrrole self-condensation route; with such complex asymmetric target molecules,<br />

highly sophisticated multistep chemical approaches are the only way that this can be accomplished.<br />

for references see p 1223


1086 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 6 Structures of Octaethylporphyrin and meso-Tetraphenylporphyrin<br />

Et<br />

Et<br />

Et Et<br />

Et<br />

NH N<br />

N HN<br />

21<br />

Et<br />

Et<br />

Et<br />

Ph<br />

NH N<br />

N HN<br />

The porphyrin field has a very rich history, which has produced numerous Nobel laureates.<br />

Hans Fischer [22–24] was unquestionably the major personality in porphyrin synthesis<br />

and chemistry in the 20th century. Fischer s group in Munich was responsible for<br />

much of the fundamental porphyrin synthetic chemistry that has been used since. New<br />

principles of contemporary organic chemistry and new reagents have been developed<br />

(benzyl esters, tert-butyl esters, diborane, catalytic hydrogenation, etc.), but Fischer must<br />

be credited with the basic philosophy that we all still employ. For porphyrin synthesis,<br />

Fischer s choice of intermediates were dipyrromethenes (see Section 17.8.1.1.1.1), and in<br />

the 1920s and 1930s the Munich school developed many approaches based on these dipyrrolic<br />

intermediates. However, methodologies for differential protection of reactive functional<br />

groups had not yet been developed, so syntheses of unique porphyrins without<br />

contamination by other isomers and congeners were usually not possible in Fischer s<br />

time. The problems of symmetry inherent when two dipyrromethenes are used to construct<br />

a porphyrin were fully understood; when syntheses were ambiguous and mixtures<br />

of porphyrins resulted, methods were developed to enable the separation of these mixtures.<br />

Ingenious use of substituent polarity differences usually allowed the mixtures to<br />

be separated. Moreover, in Fischer s time it was a straightforward matter to synthesize<br />

porphyrins (and mixtures of porphyrins) on the gram scale, thereby permitting development<br />

of separation methodology such as countercurrent distribution, solvent extraction,<br />

based on Willstätter s acid numbers, crystallization, and later, chromatography.<br />

Modern practitioners of the art of porphyrin synthesis have preferred to design their<br />

routes so that one unique porphyrin is the product; the scale of a research reaction also<br />

rarely exceeds 100 milligrams of product porphyrin. Separation of synthetic porphyrin<br />

mixtures, these days, is regarded as inelegant, even though chromatographic methods<br />

for separation (including high-performance liquid chromatography) have improved dramatically.<br />

Hence, very complex porphyrins with diverse substituents have, in the recent<br />

past, been approached via routes involving characterizable open-chain tetrapyrrolic intermediates.<br />

The task then reduces to a search for conditions, which will allow relatively<br />

labile side chains to be carried through reaction sequences intact, and will not scramble<br />

the often acid-labile pyrrole rings in oligopyrrole intermediates.<br />

Scheme 7 depicts the various bond-forming strategies of syntheses of porphyrins. Approaches<br />

A and B involve the tetramerization of monopyrroles. Case A utilizes a pyrrole<br />

(which must have identical substituents on its 3- and 4-positions to avoid production of<br />

mixtures) and an interpyrrolic one-carbon unit; the classic example here is the reaction<br />

of pyrrole with benzaldehyde to give 5,10,15,20-tetraphenylporphyrin (TPP, 22; Scheme<br />

6). In Scheme 7, case B, the interpyrrolic carbon is already affixed to the 2-position of the<br />

pyrrole; strangely enough, because of acid-catalyzed pyrrole ring redistribution reactions,<br />

the 3- and 4-positions must also usually possess identical substituents, and such an approach<br />

is the method of choice for synthesis of 2,3,7,8,12,13,17,18-octaethylporphyrin<br />

(OEP, 21). Mode C represents a new development in which two different pyrroles (one<br />

with two interpyrrolic carbons, and one with none) are condensed together to give one<br />

porphyrin only. Methods D–F in Scheme 7 employ dipyrrolic precursors. In case D, a sin-<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Ph<br />

Ph<br />

22<br />

Ph


17.8.1 Porphyrins 1087<br />

agle dipyrromethane is treated with a reagent carrying a one-carbon linker (e.g., a formaldehyde<br />

equivalent, or benzaldehyde); in this case the dipyrromethane self-condenses, so<br />

use of two dipyrromethanes will yield a mixture of three porphyrins. Example E involves<br />

reaction of two different dipyrromethanes or dipyrromethenes with each other; in this<br />

case the two required interpyrrolic carbons are attached to one or other of the dipyrromethanes,<br />

and for synthesis of a unique porphyrin product, one or other of the dipyrromethanes<br />

must be symmetrically substituted about the dipyrromethane 5-(meso)-carbon.<br />

When dipyrromethanes are used in case E, the approach is one of the variations of the socalled<br />

MacDonald [2+2] route; use of dipyrromethenes in this approach was pioneered by<br />

Fischer. Scheme 7, case F also depicts the MacDonald route [25] when a dipyrromethane is<br />

used, or a variation of the Fischer method when a dipyrromethene is used. In this case, a<br />

single dipyrromethane or dipyrromethene bearing one interpyrrolic carbon is self-condensed<br />

to give a centrosymmetric porphyrin; use of two dipyrromethanes or dipyrromethenes<br />

will afford a mixture of three porphyrin products.<br />

Scheme 7 Common Bond Formations in Porphyrin Syntheses<br />

N N<br />

N N<br />

A<br />

N N<br />

N N<br />

D<br />

N N<br />

N N<br />

G<br />

N N<br />

N N<br />

B<br />

N N<br />

N N<br />

E<br />

N N<br />

N N<br />

H<br />

N N<br />

N N<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

C<br />

N N<br />

N N<br />

F<br />

N N<br />

N N<br />

Mode G in Scheme 7 depicts the so-called [3+1] route to porphyrins, and methods H and I<br />

illustrate two examples of the open-chain tetrapyrrole route (using bilenes or a,c-biladienes,<br />

see Sections 17.8.1.1.3.2 and 17.8.1.1.3.3). Case H shows the approach when the<br />

final interpyrrolic carbon is added separately (as a formaldehyde or orthoformate equivalent,<br />

or simply as benzaldehyde), and in mode I the interpyrrolic carbon is pre-attached to<br />

the open-chain tetrapyrrole (seco-porphyrin).<br />

17.8.1.1 Syntheses of Intermediates Used in Porphyrin Syntheses<br />

No matter how simple or complex the target may be, porphyrin syntheses require the<br />

ready availability of functionalized monopyrrole building blocks. Syntheses and functionalization<br />

of such monopyrroles are outside the scope of this chapter, but these issues have<br />

been reviewed comprehensively by Black in Section 9.13, Science of Synthesis, Vol. 9 (Fully<br />

I<br />

for references see p 1223


1088 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aUnsaturated Small-Ring Heterocycles and Monocyclic Five-Membered Hetarenes with<br />

One Heteroatom). The functionalization reactions of monopyrroles are also of key importance<br />

because approaches to porphyrins via oligopyrroles require that monopyrroles be<br />

selectively linked to other pyrroles in a rational and unambiguous way.<br />

17.8.1.1.1 Dipyrroles<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9b, pp 585–588.<br />

There are four types of dipyrrole unit that have been used extensively in syntheses of<br />

porphyrins and their ring-contracted and -expanded derivatives. These are dipyrromethenes<br />

(usually handled as the highly crystalline hydrobromide salts 23), dipyrromethanes<br />

24, dipyrroketones 25, and bipyrroles 26 (Scheme 8). An added level of complexity<br />

is related to whether or not the dipyrroles are symmetrically substituted about<br />

some central point in the molecule (e.g., the 5-position in 23–25). For the purposes of generality<br />

in the syntheses of porphyrins, the unsymmetrically functionalized dipyrroles are<br />

clearly the most useful; it also follows that they are the most difficult to prepare. Unsymmetrical<br />

dipyrromethenes and dipyrromethanes are invariably approached using methodology<br />

that employs differential protection of substituents on the two constituent pyrrole<br />

rings. Symmetrical examples, on the other hand, can often be obtained by some kind<br />

of self-condensation of a monopyrrole, or by treatment of two moles of a monopyrrole<br />

with a one-carbon reagent which will provide the 5-carbon. Both types of approach will<br />

be discussed here. In the case of bipyrroles 26, which are not actually synthetic precursors<br />

of normal porphyrin systems, only symmetrical routes have been developed to date.<br />

Scheme 8 Common Dipyrrolic Intermediates<br />

R 2<br />

R 2<br />

R 3 R 4<br />

R<br />

NH HN<br />

5<br />

R1 R6 +<br />

Br −<br />

R 3 R 4<br />

NH HN<br />

R 1 R 6<br />

24<br />

17.8.1.1.1.1 Method 1:<br />

Dipyrromethenes<br />

R 5<br />

R 2<br />

23<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 2<br />

R3 R4 O<br />

NH HN<br />

R 1 R 6<br />

25<br />

R 3 R 4<br />

R<br />

NH HN<br />

5<br />

R1 R6 +<br />

Br −<br />

For historical reasons, dipyrromethene synthesis will be dealt with first, though they are<br />

little used in contemporary porphyrin synthesis.<br />

17.8.1.1.1.1.1 Variation 1:<br />

Condensation of 2-Formyl-1H-pyrroles with 2-Unsubstituted 1H-Pyrroles<br />

The best and most commonly used method for the preparation of unsymmetrically substituted<br />

dipyrromethenes (e.g., 29) is the condensation of a 2-formyl-1H-pyrrole 27 with a<br />

2-unsubstituted 1H-pyrrole 28 in the presence of acid; the product is often obtained in virtually<br />

quantitative yield, provided it is isolated as the salt 29 (Scheme 9). [26] Because of<br />

R 5<br />

R2 4<br />

R 1<br />

5<br />

R3 3<br />

N<br />

H<br />

2 2'<br />

26<br />

R 4<br />

3'<br />

N<br />

H<br />

5'<br />

R5 4'<br />

R 6


17.8.1 Porphyrins 1089<br />

atheir acid–base chemistry, dipyrromethenes tend to spread widely on chromatography<br />

columns, so crystallization is usually the best method for isolation, and recrystallization<br />

for purification.<br />

Scheme 9 Condensation of 2-Formyl-1H-pyrroles with 2-Unsubstituted 1H-Pyrroles [26,27]<br />

HO2C<br />

N<br />

H<br />

27<br />

Et<br />

CHO<br />

+<br />

N<br />

H<br />

28<br />

NH HN<br />

+<br />

Br −<br />

29<br />

3-Ethyl-2,8,9-trimethyldipyrromethene-1-carbocylic Acid Hydrobromide (29); Typical<br />

Procedure: [27]<br />

4-Ethyl-5-formyl-3-methyl-1H-pyrrole-2-carboxylic acid (27; 2.4 g, 13.25 mmol) and 2,3-dimethyl-1H-pyrrole<br />

(28; 1.2 g, 12.61 mmol) were suspended in AcOH (6 mL) and treated, after<br />

cooling, with 48% HBr (2 mL). Scratching the container with a glass rod caused the dipyrromethene<br />

hydrobromide to crystallize from soln. After 1 h the solid was collected by<br />

filtration and washed with AcOH/Et 2O. The dipyrromethene salt 29 was obtained by recrystallization<br />

(AcOH); yield: 3.3 g (77%); mp 1708C.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

HBr<br />

77%<br />

Et<br />

HO 2C<br />

17.8.1.1.1.1.2 Variation 2:<br />

Reaction of 2-(Bromomethyl)-1H-pyrroles with 2-Bromo-1H-pyrroles<br />

Heating of 2-(bromomethyl)-1H-pyrroles, e.g. 31 (obtained in situ by bromination of the<br />

corresponding 2-methyl-1H-pyrrole 30), with the same 2-bromo-1H-pyrrole 30 (synthesized<br />

by bromination of 2-unsubstituted 1H-pyrrole 32) and bromine affords good yields<br />

of dipyrromethene hydrobromides 33 (Scheme 10). [28]<br />

Scheme 10 Reaction of 2-(Bromomethyl)-1H-pyrroles with 2-Bromo-1H-pyrroles [28]<br />

Br<br />

Br<br />

N<br />

H<br />

30<br />

N<br />

H<br />

31<br />

Br 2<br />

CO2Et<br />

CO2Et<br />

Br<br />

+<br />

Br<br />

N<br />

H<br />

32<br />

N<br />

H<br />

30<br />

Br 2<br />

CO2Et<br />

CO2Et<br />

Br 2, AcOH<br />

37%<br />

EtO2C<br />

Br<br />

NH HN<br />

+<br />

Br −<br />

33<br />

CO 2Et<br />

Diethyl 1-Bromo-2,7,9-trimethyldipyrromethene-3,8-dicarboxylate Hydrobromide (33);<br />

Typical Procedure: [28]<br />

Ethyl 5-bromo-2,4-dimethyl-1H-pyrrole-3-carboxylate (30; 1.2 g, 4.88 mmol) was dissolved<br />

in AcOH and treated with Br 2 (0.8 g, 5.00 mmol). The soln turned red and HBr gas was given<br />

off. The solid product was collected by filtration and washed with EtOH then Et 2O. Recrystallization<br />

(acetone/H 2O) gave 33; yield: 0.66 g (37%); mp >2608C.<br />

for references see p 1223


1090 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a17.8.1.1.1.1.3 Variation 3:<br />

Self-Condensation of Monopyrroles<br />

Very useful dipyrromethenes, the so-called “brominated kryptodipyrromethenes” 35 and<br />

36, can be obtained by treatment of kryptopyrrole 34 with bromine in hot acetic acid<br />

(Scheme 11). [29] If the 2-tert-butyl ester 1H-pyrrole 37 is used, [30] dipyrromethene 36 can<br />

be obtained pure. These dipyrromethene salts can be separated by crystallization (based<br />

upon the relative insolubility of the perbromide salt 35), or used as a mixture in formic<br />

acid for the synthesis of etioporphyrin I (see Section 17.8.1.2).<br />

Dipyrromethenes (e.g., 39) which are symmetrically substituted about the 5-(meso)carbon<br />

are prepared by self-condensation of 2-unsubstituted 1H-pyrroles 38 (R 1 =H) or<br />

1H-pyrrole-2-carboxylic acids 38 (R 1 =CO 2H) in boiling formic acid containing hydrobromic<br />

acid. [31]<br />

Scheme 11 Self-Condensation of Monopyrroles [29–33]<br />

Et<br />

Et<br />

Et<br />

N<br />

H<br />

34<br />

N<br />

H<br />

37<br />

N<br />

H<br />

CO2Bu t<br />

R 1<br />

38 R 1 = H, CO2H<br />

Br 2, AcOH, heat<br />

Br 2, Et 2O<br />

56−70%<br />

1. HCO2H, heat<br />

2. HBr<br />

R1 = H 37%<br />

Et<br />

NH HN<br />

+<br />

Br −<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Br<br />

35<br />

Et<br />

Br<br />

NH HN<br />

+<br />

Br −<br />

36<br />

Et<br />

Br<br />

Et Et<br />

NH HN<br />

+<br />

Br −<br />

39<br />

+<br />

Et<br />

Br<br />

NH HN<br />

+<br />

Br −<br />

1-Bromo-9-(bromomethyl)-3,8-diethyl-2,7-dimethyldipyrromethene Hydrobromide (36);<br />

Typical Procedure: [30]<br />

A rapidly stirred soln of tert-butyl 4-ethyl-3,5-dimethyl-1H-pyrrole-2-carboxylate (37; 20g,<br />

89.6 mmol) in Et 2O (600 mL) was treated with Br 2 (4.8 mL, 93.4 mmol) in Et 2O (400 mL),<br />

adding the Br 2 soln over a period of 2 or 3 min. The soln was stirred for a further 30 min<br />

and then added dropwise over a period of 1 h to a stirred soln of Br 2 (9.6 mL, 187.1 mmol)<br />

in Et 2O (800 mL). After 30 min the mixture was placed in a refrigerator and left overnight.<br />

The product was collected by filtration, washed with ice-cold Et 2O, and then dried in vacuo<br />

to give 36; yield: 12–15 g (56–70%); mp >300 8C.<br />

2,8-Diethyl-3,7-dimethyldipyrromethene Hydrobromide (39); Typical Procedure: [32,33]<br />

3-Ethyl-4-methyl-1H-pyrrole (38, R 1 = H; 3 g, 27.5 mmol) was treated with dry HCO 2H<br />

(30 mL) and heated for 5 min on a boiling water bath. After cooling, the mixture was treated<br />

with aq HBr (6 mL) and left to stand for 12 h. The product 39 was collected by filtration;<br />

yield: 1.57 g (37%); mp 1708C<br />

36<br />

Et<br />

Br


17.8.1 Porphyrins 1091<br />

a17.8.1.1.1.1.4 Variation 4:<br />

Oxidation of Dipyrromethanes<br />

Controlled oxidation of dipyrromethanes 40, for example with iron(III) chloride or bromine,<br />

[26,34] also provides a convenient route to dipyrromethenes 41 (Scheme 12). The bromine-promoted<br />

oxidation method has also been widely used for the identification of dipyrromethanes<br />

on analytical thin-layer chromatography plates; exposure of a developed<br />

plate to bromine vapor causes colorless dipyrromethanes to turn red-pink (dipyrromethene<br />

salt) thereby enabling their identification in mixtures of products.<br />

Scheme 12 Oxidation of Dipyrromethanes [26,34]<br />

R 2<br />

R 3<br />

NH HN<br />

40<br />

R 4<br />

R 1 R 6<br />

17.8.1.1.1.2 Method 2:<br />

Dipyrromethanes<br />

R 5<br />

FeCl3 or Br2<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 2<br />

R 3<br />

R<br />

NH HN<br />

5<br />

+<br />

X −<br />

R1 R6 17.8.1.1.1.2.1 Variation 1:<br />

Unsymmetrically Substituted Dipyrromethanes by Reaction of 2-Substituted<br />

1H-Pyrroles with 2-Unsubstituted 1H-Pyrroles<br />

Unsymmetrically substituted dipyrromethanes, e.g. 44, can be prepared by condensation<br />

of 2-(acetoxymethyl)-1H-pyrroles 42 with 2-unsubstituted 1H-pyrroles 43 in methanol or<br />

acetic acid containing a catalytic amount (


1092 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 13 Unsymmetrically Substituted Dipyrromethanes by Reaction of 2-Substituted<br />

1H-Pyrroles with 2-Unsubstituted 1H-Pyrroles [35,38–43]<br />

BnO 2C<br />

BnO2C<br />

R 1<br />

R 1<br />

R 1<br />

R 2<br />

R 2<br />

R 2<br />

Et<br />

N<br />

H<br />

47<br />

N<br />

H<br />

50<br />

N<br />

H<br />

42<br />

N<br />

Me<br />

45<br />

R 3<br />

R 3<br />

CO 2Me<br />

Et<br />

SePh<br />

Br<br />

OAc<br />

OAc<br />

+<br />

+<br />

+<br />

Et<br />

Et<br />

N<br />

H<br />

43<br />

N<br />

H<br />

43<br />

R 4 R 5<br />

N<br />

H<br />

48<br />

R 6<br />

CO 2Bn<br />

CO2Bn<br />

R 4 R 5<br />

A B<br />

HO2C A<br />

N<br />

H<br />

51<br />

py<br />

R 3<br />

+<br />

N<br />

Br −<br />

+<br />

LiO 2C<br />

N<br />

H<br />

52<br />

R 4 R 5<br />

B<br />

N<br />

H<br />

53<br />

LiOMe<br />

MeOH<br />

TsOH, MeOH<br />

MeO 2C<br />

BnO2C<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

93%<br />

K-10 clay<br />

CH2Cl2 90%<br />

A: CuOTf•benzene<br />

CH2Cl2, −78 oC B: benzene, heat<br />

R 6<br />

R 6<br />

heat<br />

Et<br />

BnO 2C<br />

R 2<br />

R 2<br />

R 1<br />

R 1<br />

Et<br />

R 3<br />

R 3<br />

44<br />

NMe HN<br />

46<br />

NH HN<br />

49<br />

A B<br />

NH HN<br />

54<br />

Et<br />

Et<br />

R 4<br />

R 4<br />

CO2Bn<br />

CO 2Bn<br />

Dibenzyl 7-Ethyl-3-[2-(methoxycarbonyl)ethyl]-2,8-dimethyldipyrromethane-1,9-dicarboxylate<br />

(44); Typical Procedure: [35]<br />

A suspension of benzyl 5-(acetoxymethyl)-4-[2-(methoxycarbonyl)ethyl]-3-methyl-1H-pyrrole-2-carboxylate<br />

(42; 373 mg, 1.0 mmol) and benzyl 4-ethyl-3-methyl-1H-pyrrole-2-carboxylate<br />

(43; 243 mg, 0.95 mmol) in MeOH (5 mL) was treated with TsOH (10 mg) and heated<br />

at 408C under N 2 with stirring for 5 h. The soln was diluted with H 2O (0.5 mL) and the<br />

dipyrromethane was collected by filtration. Recrystallization (CH 2Cl 2/hexanes) gave 44;<br />

yield: 520 mg (93%); mp 89–90 8C.<br />

Dibenzyl 2,3,7-Triethyl-N A,8-dimethyldipyrromethane-1,9-dicarboxylate (46); Typical<br />

Procedure: [43]<br />

Benzyl 5-(acetoxymethyl)-3,4-diethyl-1-methyl-1H-pyrrole-2-carboxylate (45; 1.70 g,<br />

4.95 mmol) was added to a well-stirred soln of benzyl 4-ethyl-3-methyl-1H-pyrrole-2-carboxylate<br />

(43; 1.20 g, 4.93 mmol) in CH 2Cl 2 (100 mL) containing a suspension of montmorillonite<br />

K-10 clay (20 g). After being stirred for 30 min the mixture was filtered free from<br />

R 6<br />

R 6<br />

R 5<br />

R 5


17.8.1 Porphyrins 1093<br />

athe clay and concentrated to dryness to afford the dipyrromethane 46 as a viscous oil;<br />

yield: 2.34 g (90%).<br />

Dipyrromethanes from 2-[(Phenylselenyl)methyl]-1H-pyrroles and 2-Unsubstituted 1H-<br />

Pyrroles; General Procedures: [40]<br />

Using Copper(I) Triflate: To a soln of 2-[(phenylselenyl)methyl]-1H-pyrrole (47; 0.2 mmol)<br />

and 2-unsubstituted 1H-pyrrole (48; 0.21 mmol) in dry degassed CH 2Cl 2 (5 mL) containing<br />

powdered CaCO 3 (0.24 mmol) at –78 8C under argon was added the benzene complex of<br />

copper(I) trifluoromethanesulfonate (0.12 mmol) in dry degassed benzene (0.5 mL) (CAU-<br />

TION: carcinogen). The mixture was stirred for 1–2 min at –788C, then poured into H 2O<br />

(20 mL), and extracted with CH 2Cl 2 (4 ” 10 mL). The combined organic extracts were washed<br />

with H 2O (20 mL). The solvent was removed and the product 49 was purified by preparative<br />

TLC (silica gel, hexanes/Et 2O 1:9).<br />

Thermal Method: The 2-[(phenylselenyl)methyl]-1H-pyrrole (47; 0.2 mmol) and 2-unsubstituted<br />

1H-pyrrole (48; 0.21 mmol) were heated in dry degassed benzene (10 mL)<br />

(CAUTION: carcinogen) under argon for 5 h. The workup was then as above for the copper(I)<br />

trifluoromethanesulfonate method.<br />

Dipyrromethanes from Pyrrole Pyridinium Salts and Lithium Carboxylate; General Procedure:<br />

[42]<br />

A soln of the 1H-pyrrole-2-carboxylic acid (52; 0.01 mol) and LiOMe (0.38 g, 0.01 mol) in<br />

warm aq MeOH (100 mL, 50%) was added to the 2-(bromomethyl)-1H-pyrrole (50;<br />

0.01 mol) in pyridine (3.2 mL, 0.04 mol). The mixture was refluxed on a steam bath for several<br />

hours under N 2. On cooling to rt the dipyrromethane crystallized and was collected<br />

by filtration, washed with H 2O, and dried. When ethylene glycol or HCONH 2 was used as<br />

solvent, the mixture was worked up by dilution with H 2O (ca. 1 L) and extracted with Et 2O<br />

(4 ” 150 mL). The Et 2O was dried (MgSO 4), concentrated to dryness, and the residue was<br />

chromatographed [alumina, petroleum ether (bp 60–80 8C)/toluene] to give 54.<br />

17.8.1.1.1.2.2 Variation 2:<br />

Unsymmetrically Substituted Dipyrromethanes by Reduction of<br />

Dipyrromethenes<br />

Reduction of dipyrromethenes, e.g. 55, with sodium borohydride also furnishes dipyrromethanes<br />

56 (Scheme 14). [44] Both symmetrical and unsymmetrical dipyrromethenes are<br />

available; this method is applicable to both.<br />

Scheme 14 Unsymmetrically Substituted Dipyrromethanes by Reduction of<br />

Dipyrromethenes [44]<br />

+<br />

H3N CO2Me<br />

NH HN<br />

+<br />

2Br −<br />

55<br />

NaBH4, H2O<br />

H2N CO2Me<br />

NH HN<br />

56<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1094 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a2-(2-Aminoethyl)-8-[2-(methoxycarbonyl)ethyl]-3,7-dimethyldipyrromethane (56); Typical<br />

Procedure: [44]<br />

2-(2-Aminoethyl)-8-[2-(methoxycarbonyl)ethyl]-3,7-dimethyldipyrromethene dihydrobromide<br />

(55; 1.0 g, 2.08 mmol) in deoxygenated distilled H 2O (60 mL) was treated with<br />

NaBH 4 (1.7 g, 44.9 mmol) in H 2O (30 mL) while a stream of N 2 was being passed through<br />

the soln. Effervescence took place and the red dipyrromethene color was immediately discharged.<br />

CH 2Cl 2 (100 mL) and 10% aq NaHCO 3 (30 mL) were added rapidly and, after separation,<br />

the organic phase was dried (Na 2SO 4), concentrated to dryness, and used immediately<br />

in the next reaction (to form the porphyrin in Woodward s chlorophyll a synthesis).<br />

17.8.1.1.1.2.3 Variation 3:<br />

Symmetrically Substituted Dipyrromethanes by Self-Condensation of<br />

2-Substituted 1H-Pyrroles<br />

Dipyrromethanes 59 which are symmetrically substituted about the 5-carbon are obtained<br />

in good yield by self-condensation of 2-(bromomethyl)-1H-pyrroles 58 (obtained<br />

from the corresponding 2-methyl-1H-pyrrole 57 by bromination in diethyl ether) in hot<br />

methanol, [45] or by heating 2-(acetoxymethyl)-1H-pyrroles 42 [obtained by treatment of<br />

the corresponding 2-methyl-1H-pyrrole 60 with lead(IV) acetate in acetic acid] in methanol/hydrochloric<br />

acid to give, for example, 61 [46] (Scheme 15); the mechanism for formation<br />

of 59 and 61 involves loss of one 2-CH 2X group.<br />

Scheme 15 Symmetrically Substituted Dipyrromethanes by Self-Condensation of<br />

2-Substituted 1H-Pyrroles [45–48]<br />

N<br />

H<br />

57<br />

CO 2Bn<br />

Br 2, Et 2O<br />

Br<br />

MeOH<br />

heat<br />

BnO 2C<br />

NH HN<br />

59 56%<br />

CO 2Bn<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

H<br />

58<br />

CO 2Bn<br />

MeO2C MeO2C<br />

N<br />

H<br />

60<br />

CO2Bn<br />

Pb(OAc) 4<br />

AcOH<br />

AcO<br />

N<br />

H<br />

42<br />

CO2Bn<br />

BnO 2C<br />

MeOH, HCl, heat<br />

88%<br />

MeO 2C CO2Me<br />

NH HN<br />

61<br />

CO2Bn<br />

Dibenzyl 2,3,7,8-Tetramethyldipyrromethane-1,9-dicarboxylate (59); Typical<br />

Procedure: [47]<br />

Benzyl 3,4,5-trimethyl-1H-pyrrole-2-carboxylate (57; 2.0 g, 8.22 mmol) in rapidly stirred<br />

anhyd Et 2O (45 mL) was treated with Br 2 (0.5 mL, 9.73 mmol) in Et 2O (10 mL). The 2-(bromomethyl)-1H-pyrrole<br />

58 immediately precipitated from the soln, but the mixture was stirred<br />

for an additional 1.5 h before evaporation of the Et 2O(CAUTION: HBr gas) to give an<br />

orange residue of benzyl 5-(bromomethyl)-3,4-dimethyl-1H-pyrrole-2-carboxylate (58).


17.8.1 Porphyrins 1095<br />

aThe solid was dissolved in MeOH (25 mL) and refluxed for 4 h, then set aside at rt overnight.<br />

The product 59 was collected by filtration and recrystallized (CH 2Cl 2/petroleum<br />

ether); yield: 1.1 g (56%); mp 178–1798C.<br />

Dibenzyl 3,7-Bis[2-(methoxycarbonyl)ethyl]-2,8-dimethyldipyrromethane-1,9-dicarboxylate<br />

(61); Typical Procedure: [48]<br />

Benzyl 5-(acetoxymethyl)-4-[2-(methoxycarbonyl)ethyl]-3-methyl-1H-pyrrole-2-carboxylate<br />

(42; 320 mg, 0.857 mmol) in MeOH (5 mL) and HCl (d 1.18; 0.3 mL) was heated on a<br />

water bath for 4 h. After cooling the dipyrromethane 61 was collected by filtration; yield:<br />

230 mg (88%); mp 99–1008C.<br />

17.8.1.1.1.2.4 Variation 4:<br />

Symmetrically Substituted Dipyrromethanes from 2-Unsubstituted<br />

1H-Pyrroles<br />

2-Unsubstituted 1H-pyrroles such as 62, when treated with formaldehyde, give dipyrromethanes,<br />

e.g. 63. [49,50] A general method for 5-alkyldipyrromethane synthesis is available;<br />

[51] high yields of dipyrromethanes 64 and 65 can be obtained by using a 2-unsubstituted<br />

monopyrrole (e.g., 43 or 62, respectively) with dimethylacetals of aliphatic aldehydes<br />

and an acid catalyst (Scheme 16). 5-Substituted dipyrromethanes (usually 5-aryl derivatives,<br />

e.g. 66) can also be synthesized by treatment of an aldehyde (e.g., benzaldehyde)<br />

with an excess of 2-unsubstituted 1H-pyrrole (e.g., pyrrole) and an acid catalyst. [52]<br />

Scheme 16 Symmetrically Substituted Dipyrromethanes from 2-Unsubstituted 1H-<br />

Pyrroles [39,49–52]<br />

Et Et<br />

Et<br />

N<br />

H<br />

62<br />

N<br />

H<br />

43<br />

Et Et<br />

N<br />

H<br />

N<br />

H<br />

62<br />

CO 2Et<br />

CO2Bn<br />

CO2Et<br />

HCHO<br />

MeO<br />

PhCHO, TFA (cat.)<br />

49%<br />

OMe<br />

77%<br />

Et<br />

CO 2Me<br />

TsOH, toluene, heat<br />

OMe<br />

MeO CO 2Me<br />

EtO 2C<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

TsOH, K-10 clay, 1,4-dioxane, 70 o C<br />

66%<br />

Ph<br />

NH HN<br />

66<br />

63<br />

BnO 2C<br />

Et<br />

Et<br />

Et<br />

CO 2Et<br />

NH HN<br />

Et<br />

64<br />

EtO 2C<br />

CO2Me<br />

Et<br />

Et<br />

CO 2Bn<br />

CO2Me<br />

Et<br />

NH HN<br />

65<br />

Et<br />

CO 2Et<br />

for references see p 1223


1096 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aDibenzyl 3,7-Diethyl-5-[(methoxycarbonyl)methyl]-2,8-dimethyldipyrromethane-1,9-dicarboxylate<br />

(64); Typical Procedure: [51]<br />

Benzyl 4-ethyl-3-methyl-1H-pyrrole-2-carboxylate (43; 3.31g, 13.6 mmol) in toluene<br />

(50 mL) was treated with TsOH (326 mg) and methyl 3,3-dimethoxypropanoate (1.0 mL,<br />

7.05 mmol). The mixture was heated at 120 8C for 5 h then CH 2Cl 2 was added and the organic<br />

layer washed with H 2O, brine, and then dried (Na 2SO 4). The solvent was removed<br />

under reduced pressure and the residue was flash chromatographed (silica gel, EtOAc/cyclohexane<br />

first 1:9, increasing to 1:5) to give 64 as a yellow foam; yield 2.98 g (77%).<br />

Diethyl 5-(Methoxycarbonyl)-2,3,7,8-tetraethyldipyrromethane-1,9-dicarboxylate (65);<br />

Typical Procedure: [39]<br />

Ethyl 3,4-diethyl-1H-pyrrole-2-carboxylate (62; 632 mg, 3.24 mmol) in 1,4-dioxane (7 mL)<br />

was treated with methyl dimethoxyacetate (0.15 mL, 3.24 mmol) and TsOH/K-10 clay (see<br />

below; 3 g). The mixture was stirred at 708C for 5 h before filtering off the clay. Evaporation<br />

of the filtrate gave a residue which was dissolved in CH 2Cl 2 and washed with aq NaH-<br />

CO 3 and then with brine. After drying (Na 2SO 4) the solvent was removed to give an oil<br />

which was then chromatographed (silica gel, CH 2Cl 2 then increasing amounts of THF).<br />

The appropriate eluates were collected and concentrated to give the product, which was<br />

crystallized (CHCl 3) to give 65; yield: 493 mg (66%); mp (slowly melts) 169–2118C.<br />

TsOH/K-10 clay: Dioxane (1 L) was saturated with TsOH. Montmorillonite K-10 clay<br />

was added until the mixture almost became too stiff to stir with a stirrer bar. The mixture<br />

was stirred overnight and then concentrated to dryness. The residual mixture was suspended<br />

in CH 2Cl 2 and then filtered and rinsed with THF and 1,4-dioxane to remove the<br />

excess of non-adsorbed TsOH from the clay. The clay was dried overnight in a vacuum<br />

oven at 528C.<br />

5-Phenyldipyrromethane (66); Typical Procedure: [52]<br />

A soln of benzaldehyde (0.1 mL, 1 mmol) and 1H-pyrrole (2.8 mL, 40 mmol) was degassed<br />

by bubbling argon through it for 10 min and then treated with TFA (0.008 mL, 0.1 mmol).<br />

The soln was stirred for 15 min at rt (TLC monitoring), before being diluted with CH 2Cl 2<br />

(50 mL), washed with 0.1 M aq NaOH and H 2O, and then dried (Na 2SO 4). The solvent was<br />

removed under reduced pressure, and unreacted 1H-pyrrole was removed by vacuum distillation<br />

at rt. The resulting yellow amorphous solid was dissolved in a small amount of<br />

the eluent and then purified by flash chromatography [silica gel (230–400 mesh), cyclohexane/EtOAc/Et<br />

3N 80:20:1). The dipyrromethane-containing fractions were collected<br />

and evaporated to give 66; yield: 110 mg (49%); mp 100–1018C.<br />

17.8.1.1.1.3 Method 3:<br />

Dipyrroketones<br />

The most efficient method for the synthesis of unsymmetrical dipyrroketones (e.g., 68,<br />

R 1 = O) involves a modification of the Vilsmeier–Haack reaction. Condensation of the<br />

phosphoryl chloride complex of a pyrrole N,N-dimethylcarboxamide 67 with a nucleophilic<br />

2-unsubstituted 1H-pyrrole (e.g., 34) followed by hydrolysis of the intermediate<br />

imine salt 68 (R 1 =NMe 2 + Cl – ) gives the dipyrroketone 68 (R 1 = O) in very good yield<br />

(Scheme 17). [53] Alternatively, treatment of a pyrrole acid chloride 69 with a pyrrole Grignard<br />

derivative 70 (or with the 2-unsubstituted 1H-pyrrole in the presence of a Friedel–<br />

Crafts catalyst) gives dipyrroketones 71 [53,54] (Scheme 17).<br />

Symmetrically substituted dipyrroketones (e.g., 73) are generally obtained by treatment<br />

of pyrrole Grignard reagents 72 with phosgene, [55] or, for example in the preparation<br />

of 76, by oxidative hydrolysis of dipyrrothiones 75, which can in turn be obtained<br />

directly from 2-unsubstituted 1H-pyrroles 74 by treatment with thiophosgene (Scheme<br />

17). [56]<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


17.8.1 Porphyrins 1097<br />

aOxidation of dipyrromethanes (e.g., 77) with lead(IV) oxide, lead(IV) acetate, [57] sulfuryl<br />

chloride, or bromine/sulfuryl chloride [58] also gives good yields of dipyrroketones 78.<br />

Reduction of dipyrroketones 78 with sodium amalgam [57] or diborane yields the corresponding<br />

dipyrromethanes 77.<br />

Scheme 17 Dipyrroketone Syntheses [53–57]<br />

Et<br />

BnO2C CONMe2<br />

N<br />

H<br />

67<br />

EtO2C<br />

N<br />

COCl<br />

H<br />

Et<br />

N<br />

H<br />

74<br />

N<br />

H<br />

72<br />

69<br />

Et<br />

MgX<br />

S<br />

Cl Cl<br />

benzene<br />

EtOH<br />

64%<br />

+<br />

O<br />

XMg<br />

Cl Cl<br />

EtO2C CO 2Et<br />

EtO2C CO 2Et<br />

NH HN<br />

EtO 2C CO2Et<br />

77<br />

1. POCl3<br />

Et<br />

2.<br />

, CH2Cl2<br />

N<br />

H<br />

34<br />

N<br />

H<br />

70<br />

80%<br />

BnO2C<br />

NH HN<br />

68 R 1 = O, NMe 2 + Cl −<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

O<br />

Et<br />

EtO 2C<br />

Et Et<br />

NH HN<br />

S<br />

NH HN<br />

75<br />

Pb(OAc) 4, PbO 2<br />

62%<br />

Zn/Hg or B 2H 6<br />

Zn/Hg: 20%<br />

73<br />

KOH, H 2O 2<br />

EtOH<br />

71%<br />

Et<br />

R 1<br />

NH HN<br />

78<br />

O<br />

NH HN<br />

71<br />

EtO2C CO2Et<br />

O<br />

Et<br />

NH HN<br />

76<br />

EtO2C CO 2Et<br />

O<br />

EtO2C CO2Et Benzyl 3,8-Diethyl-2,7,9-trimethyl-5-dipyrroketone-1-carboxylate (68,R 1 = O); Typical Procedure:<br />

[53]<br />

Benzyl 5-[(dimethylamino)carbonyl]-4-ethyl-3-methyl-1H-pyrrole-2-carboxylate (67; 10g,<br />

31.81 mmol) was dissolved in POCl 3 (15 mL) and stirred until the absorption at 374 nm<br />

was maximized (e 15,000). The excess of POCl 3 was removed by concentration under reduced<br />

pressure, then 1,2-dibromoethane (10 mL) was added and the soln concentrated to<br />

Et<br />

for references see p 1223


1098 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aremove the last traces of POCl 3. The oily product was dissolved in CH 2Cl 2 (15 mL) and 3ethyl-2,4-dimethyl-1H-pyrrole<br />

(34; 4.35 g, 35.31 mmol) in CH 2Cl 2 (15 mL) was added. After<br />

stirring until the absorption at 402 nm was maximized (e 7900), NaOAc (60 g) in H 2O<br />

(100 mL) was added and the mixture was refluxed with vigorous stirring for 2 h. The mixture<br />

was cooled, the organic layer was separated, and the aqueous layer was extracted<br />

with CHCl 3. The combined organic phases were washed with aq Na 2CO 3 (10%) and H 2O,<br />

dried (MgSO 4), and concentrated to dryness. The dark oil crystallized when triturated<br />

with MeOH, and the product was recrystallized (CHCl 3/petroleum ether) to give the product;<br />

yield: 9.7 g (80%); mp 1828C.<br />

1,2,3,7,8,9-Hexamethyl-5-dipyrroketone (76); Typical Procedure: [56]<br />

CAUTION: Thiophosgene is a severe respiratory irritant and very toxic by inhalation.<br />

A stirred soln of 2,3,4-trimethyl-1H-pyrrole (74; 5.0 g, 45.8 mmol) in Et 2O (100 mL) was<br />

treated dropwise at 0 8C with a soln of thiophosgene (1.78 mL) in benzene (20 mL) (CAU-<br />

TION: carcinogen). The mixture was stirred for 10 min before addition of aq EtOH (90%,<br />

100 mL) and stirring for another 30 min at rt. The mixture was concentrated to give a crystalline<br />

residue, which was recrystallized (aq EtOH) to give the dipyrrothione 75; yield:<br />

3.8 g (64%); mp 205–2078C. The thione 75 (50 mg, 0.19 mmol) in EtOH (95%, 10 mL) containing<br />

KOH (0.1 g) was treated with aq H 2O 2 (5%, 1 mL) and then heated on a steam bath<br />

for 5 min. H 2O (30 mL) was added and the mixture was extracted with CHCl 3 (3 ” 5 mL).<br />

Evaporation of the solvent gave a residue, which was purified by sublimation (2008C/<br />

1 Torr) to give the dipyrroketone 76; yield: 33 mg (71%); mp 219–2218C.<br />

Diethyl 3,7-Bis[2-(ethoxycarbonyl)ethyl]-2,8-bis[(ethoxycarbonyl)methyl]-5-dipyrroketone-1,9-dicarboxylate<br />

(78); Typical Procedure: [57]<br />

A soln of diethyl 3,7-bis[2-(ethoxycarbonyl)ethyl]-2,8-bis[(ethoxycarbonyl)methyl]dipyrromethane-1,9-dicarboxylate<br />

(77; 2.0 g, 3.02 mmol) in AcOH (75 mL) was treated with<br />

Pb(OAc) 4 (2.9 g, 6.54 mmol) and stirred at rt for 4 d. PbO 2 (90%, 2.3 g, 8.65 mmol) was then<br />

added and stirring was continued for another 2 d. The mixture was centrifuged and the<br />

supernatant was poured into ice water (500 mL). The colorless precipitate was separated,<br />

washed with H 2O, dissolved in Et 2O, and then washed successively with H 2O, 5% aq NaH-<br />

CO 3, and H 2O, and then dried (Na 2SO 4). After concentration, crystals separated; these were<br />

then recrystallized (Et 2O) to give 78; yield: 1.27 g (62%); mp 152.5–153.5 8C.<br />

Diethyl 3,7-Bis[2-(ethoxycarbonyl)ethyl]-2,8-bis(ethoxycarbonylmethyl)dipyrromethane-<br />

1,9-dicarboxylate (77); Typical Procedure: [57]<br />

Diethyl 3,7-bis[2-(ethoxycarbonyl)ethyl]-2,8-bis(ethoxycarbonylmethyl)-5-dipyrroketone-<br />

1,9-dicarboxylate (78; 668 mg, 0.99 mmol) in EtOH (5 mL) was added to H 2O (1 mL) and<br />

concd HCl (1 mL) and Zn amalgam (780 mg). The mixture was refluxed for 3 h, cooled,<br />

and then filtered. The soln was concentrated and then refrigerated to give crystals of the<br />

dipyrromethane 77; yield: 146 mg (20%), mp 948C.<br />

17.8.1.1.1.4 Method 4:<br />

Bipyrroles<br />

Bipyrroles 26 are not used in porphyrin syntheses because they contain a direct pyrrole–<br />

pyrrole link. However, this direct link is present in corroles, which will be discussed later<br />

in the chapter. Bipyrroles are prepared using a variation [59–61] of the Ullmann reaction, in<br />

which treatment of iodopyrrole 79 with copper affords the symmetrical bipyrrole 80 in<br />

ca. 60% yield (Scheme 18). The yields in this Ullmann approach have been improved [62] by<br />

the use of iodopyrroles (e.g., 81), which are functionalized with electron-withdrawing<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


17.8.1 Porphyrins 1099<br />

agroups on the ring nitrogen (to give 82). Thus, treatment of 82 with copper powder in dimethylformamide<br />

at 1108C affords bipyrroles 83, which then can be pyrolyzed to give a<br />

good overall yield of bipyrroles 84.<br />

Scheme 18 Bipyrrole Synthesis [59–62]<br />

EtO 2C<br />

EtO 2C<br />

N<br />

H<br />

79<br />

R 1 R 2<br />

N<br />

H<br />

81<br />

CO2Et<br />

I<br />

I<br />

Cu powder, DMF, 110 o C<br />

Cu bronze, DMF, rt, 17 h<br />

63%<br />

DMAP, CH 2Cl 2, (t-BuO 2C) 2O<br />

EtO 2C<br />

EtO2C<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 1<br />

R 2<br />

EtO 2C<br />

R 2<br />

N<br />

H<br />

N<br />

CO2Et<br />

N<br />

H<br />

80<br />

R 1 R 2<br />

R 1<br />

82<br />

R 1<br />

EtO 2C<br />

CO 2Bu t<br />

EtO<br />

N N<br />

2C<br />

CO2Et Bu<br />

83<br />

tO2C CO2But I<br />

N<br />

H<br />

R 2<br />

CO 2Et<br />

heat<br />

R 2<br />

N<br />

H<br />

84 65−83%<br />

R 1<br />

CO 2Et<br />

Tetraethyl 4,4¢-Dimethyl-2,2¢-bipyrrole-3,3¢,5,5¢-tetracarboxylate (80); Typical Procedure:<br />

[59]<br />

Diethyl 5-iodo-3-methyl-1H-pyrrole-2,4-dicarboxylate (79; 10.0 g, 28.48 mmol) in DMF<br />

(50 mL) was treated with powdered Cu bronze (10 g) and stirred at rt for 17 h. The Cu<br />

bronze was separated off and washed with hot CHCl 3 (4 ” 50 mL). The filtrate and washings<br />

were then washed with aq 1 M HCl (2 ” 100 mL), and H 2O (2 ” 100 mL), then dried.<br />

The solvent was removed under reduced pressure and the crude product was triturated<br />

with petroleum ether (bp 60–80 8C; 30 mL). The product was collected by filtration, washed<br />

with petroleum ether (bp 60–80 8C; 2 ” 15 mL), and then dried at 80 8C, to give 80; yield:<br />

4.04 g (63%); mp 178–1808C.<br />

Bipyrrole Synthesis Using the Modified Ullmann Reaction; General Procedure: [62]<br />

The 2-iodo-1H-pyrrole (81; 0.01 mol) and DMAP (0.10 equiv) were mixed with di-tert-butyl<br />

dicarbonate (1.2 equiv) in dry CH 2Cl 2 (50 mL). The mixture was then passed through a<br />

short plug of silica gel, eluting with CH 2Cl 2. After evaporation, the residual oil 82 was dissolved<br />

in DMF (20 mL) under N 2 and then treated with Cu powder (4.7 equiv); after heating<br />

at 110 8C in an oil bath until iodopyrrole could no longer be detected (TLC monitoring),<br />

the mixture was cooled to rt, and filtered through Celite (to remove Cu residues). The Celite<br />

was washed with hot CHCl 3 until eluates were clear, and the combined organic phases<br />

were washed with 10% aq HCl (” 3), 10% aq Na 2S 2O 3 (” 3), sat. aq NaHCO 3, and then with<br />

sat. aq NaCl. The organic phase was dried (Na 2SO 4) and concentrated to give the bipyrrole<br />

for references see p 1223


1100 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a83. This solid was heated under an inert atmosphere at 1808C until gas evolution ceased.<br />

The cooled solidified glass was recrystallized (hot EtOH or hexanes) to give 84; yield: 65–<br />

83%.<br />

17.8.1.1.2 Tripyrroles<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, pp 588–590.<br />

The tripyrroles most often used in porphyrin syntheses are the so-called tripyrranes<br />

86, which possess three pyrrole rings joined by two methylene groups (Scheme 19); tripyrrenes<br />

85 are non-nucleophilic on the dipyrromethene side. Tripyrranes have been obtained<br />

from tripyrrenes by reduction.<br />

Scheme 19 Unsymmetrical Tripyrranes<br />

R 5<br />

R 6<br />

R 4 R 3<br />

R 7<br />

NH HN<br />

X −<br />

+<br />

NH<br />

R 8<br />

85<br />

R 1<br />

R 2<br />

reduction<br />

17.8.1.1.2.1 Method 1:<br />

The [2+1] Approach to Tripyrranes<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 5<br />

R 6<br />

R 4 R 3<br />

Tripyrrane 88 is obtained in 35% yield by treatment of a dipyrromethane 87 (R 1 = H) with a<br />

2-(acetoxymethyl)-1H-pyrrole 42 (Scheme 20). [35,63] The tripyrrene 91 was employed as an<br />

intermediate in the synthesis of tripyrrane 92; this was subsequently used in the synthesis<br />

of a pentapyrrolic sapphyrin. [64] Treatment of the dipyrromethane 89 [35] with the 2formyl-1H-pyrrole<br />

90 afforded the tripyrrene dibenzyl diester 91 in excellent yield. Catalytic<br />

hydrogenation to tripyrrane caused debenzylation of the esters as well as reduction<br />

of the dipyrromethene link. The unstable tripyrrane 92, however, was not characterized<br />

(Scheme 20). [64] Reduction of the methene link has also been achieved by MacDonald and<br />

co-workers using sodium borohydride. [37] The route used to prepare tripyrrane 88 is also<br />

capable of yielding tripyrranes with an unsymmetrical array of substituents, as is the approach<br />

to 92.<br />

R 7<br />

NH<br />

R 8<br />

86<br />

R 1<br />

R 2


17.8.1 Porphyrins 1101<br />

aScheme 20 Unsymmetrical Tripyrranes by the [2 +1] Approach [35,63,64]<br />

MeO 2C<br />

MeO2C<br />

NH<br />

NH<br />

87 R 1 = H, CO 2t-Bu<br />

NH<br />

NH<br />

89<br />

CO2Bn<br />

+<br />

R 1<br />

CO 2Bn<br />

OHC<br />

+<br />

N<br />

H<br />

90<br />

MeO 2C<br />

AcO<br />

CO2Bn<br />

MeOH, TsOH (cat.)<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

H<br />

42<br />

CO 2Bn<br />

MeO2C<br />

MeO2C<br />

HCl<br />

H 2, Pd/C<br />

35%<br />

NH<br />

88<br />

CO2Bn<br />

NH HN<br />

Cl − +<br />

NH<br />

CO2Bn<br />

91<br />

NH HN<br />

NH<br />

CO 2H<br />

92<br />

CO 2Me<br />

CO 2Bn<br />

CO 2Bn<br />

Dibenzyl 3,7,12-Tris[2-(methoxycarbonyl)ethyl]-2,8,13-trimethyltripyrrane-1,14-dicarboxylate<br />

(88); Typical Procedure: [35]<br />

Benzyl 9-tert-(butoxycarbonyl)-3,7-bis[2-(methoxycarbonyl)ethyl]-2,8-dimethyldipyrromethane-1-carboxylate<br />

(87, R 1 =CO 2t-Bu; 290 mg, 0.498 mmol) was dissolved in TFA<br />

(4 mL) and the soln was set aside at rt for 30 min while a stream of N 2 was passed through<br />

it. The solvent was removed under reduced pressure and the residual oil was dissolved in<br />

CH 2Cl 2, washed with sat. aq NaHCO 3, and dried (Na 2SO 4). The solvent was removed under<br />

reduced pressure to leave 87 (R 1 = H) as an oil. This was treated with benzyl 5-(acetoxymethyl)-4-[2-(methoxycarbonyl)ethyl]-3-methyl-1H-pyrrole-2-carboxylate<br />

(42; 186 mg,<br />

0.498 mmol) and TsOH (9 mg) in MeOH (4 mL). This mixture was warmed at 358C with stirring<br />

for 5 h before being cooled to rt and then neutralized with a few crystals of<br />

CO2H<br />

for references see p 1223


1102 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aNaOAc•3H 2O. The tripyrrane was collected by filtration and recrystallized (Et 2O/hexanes)<br />

to give 88; yield: 140 mg (35%); mp 97–99 8C.<br />

17.8.1.1.2.2 Method 2:<br />

The [1]+[1] 2 Approach to Symmetrical Tripyrranes<br />

The classical route to tripyrranes is that discovered by Sessler and co-workers. [65,66] In its<br />

present state of development it only yields symmetrically substituted tripyrranes, but the<br />

ease with which tripyrranes can be synthesized using this method far outweighs the apparent<br />

symmetry limitation. In a typical example, treatment of electron-rich 3,4-diethyl-<br />

1H-pyrrole with 2 equivalents of a 2-(acetoxymethyl)-1H-pyrrole 93 affords tripyrrane 94<br />

in excellent yield, presumably via an intermediate dipyrromethane (Scheme 21). Catalytic<br />

debenzylation then affords the tripyrrane-1,14-dicarboxylic acid 95 (R 1 =CO 2H), which is<br />

chemically equivalent to the corresponding 1,14-di-unsubstituted derivative 95 (R 1 =H).<br />

Scheme 21 Unsymmetrical Tripyrranes by the [1]+[1] 2 Approach [65,66]<br />

Et Et<br />

N<br />

H<br />

93<br />

82%<br />

+<br />

AcO<br />

Et<br />

Et<br />

Et<br />

N<br />

H<br />

93<br />

NH HN<br />

NH<br />

CO 2Bn<br />

Et Et<br />

CO 2Bn<br />

94<br />

CO2Bn<br />

TsOH, EtOH<br />

60 oC, 8 h<br />

H 2, Pd/C, THF<br />

R 1 = CO2H 100%<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

Et<br />

Et Et<br />

NH HN<br />

NH<br />

CO 2Bn<br />

Et Et<br />

R 1<br />

95 R 1 = H, CO 2H<br />

Dibenzyl 3,7,8,12-Tetraethyl-2,13-dimethyltripyrrane-1,14-dicarboxylate (94); Typical<br />

Procedure: [65]<br />

3,4-Diethyl-1H-pyrrole (0.6 g, 4.9 mmol), benzyl 5-(acetoxymethyl)-4-ethyl-3-methyl-1Hpyrrole-2-carboxylate<br />

(93; 2.5 g, 7.9 mmol), and TsOH (0.15 g) were dissolved in abs EtOH<br />

(60 mL) and heated at 608C for 8 h under N 2. The resulting suspension was reduced in volume<br />

to 30 mL and then placed in a freezer for several hours. The product was collected by<br />

filtration, washed with a small amount of cold EtOH, and recrystallized (CH 2Cl 2/EtOH) to<br />

give 94 as a white powder; yield: 2.07 g (82%); mp 211 8C.<br />

3,7,8,12-Tetraethyl-2,13-dimethyltripyrrane-1,14-dicarboxylic Acid (95,R 1 =CO 2H); Typical<br />

Procedure: [65]<br />

Dibenzyl 3,7,8,12-tetraethyl-2,13-dimethyltripyrrane-1,14-dicarboxylate (94; 4.5g,<br />

7.1 mmol) in THF (500 mL) containing Et 3N (1 drop) was hydrogenated over 5% Pd/C<br />

(250 mg) at 1 atm of H 2 until TLC indicated completion of the reaction. The catalyst was<br />

separated by filtration and the filtrate was concentrated to dryness on a rotary evaporator.<br />

The residue was crystallized (CH 2Cl 2/hexanes) to give 95 (R 1 =CO 2H) as an unstable<br />

white powder; yield: 3.2 g (100%); mp 111–1158C.<br />

R 1


17.8.1 Porphyrins 1103<br />

a17.8.1.1.3 Open-Chain Tetrapyrrolic Intermediates<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, pp 590–596.<br />

Open-chain tetrapyrrole intermediates for use in porphyrin syntheses can potentially<br />

exist in a variety of different forms; these are bilanes 96, a-bilenes 97 and b-bilenes 98,<br />

a,b-biladienes 99 and a,c-biladienes 100, and a,b,c-bilatrienes 101 (Scheme 22). However,<br />

not all open-chain tetrapyrroles can be employed effectively. Among the bilane group,<br />

only the a-oxo- and b-oxobilanes (102 and 103, respectively) are useful. The only types of<br />

bilene to be used for syntheses of pure porphyrins are the b-bilenes 98; a,c-biladienes 100<br />

have been shown to be the only useful biladienes (and in fact have been the most useful of<br />

all open-chain tetrapyrroles). a,b,c-Bilatrienes 101, though proposed to be intermediates<br />

in the syntheses of porphyrins via b-bilenes and a,c-biladienes, have not been used routinely<br />

as intermediates in porphyrin synthesis.<br />

Scheme 22 Open-Chain Tetrapyrroles<br />

R 2<br />

R 3<br />

R 2<br />

R 3<br />

R 2<br />

O<br />

R 3<br />

R 1 R 8<br />

R 4<br />

NH HN<br />

a c<br />

NH HN<br />

b<br />

96<br />

R 5<br />

R 1 R 8<br />

R 4<br />

N HN<br />

NH N<br />

99<br />

R 5<br />

R 1 R 8<br />

R 4<br />

NH HN<br />

NH HN<br />

102<br />

R 5<br />

R 7<br />

R 6<br />

R 7<br />

R 6<br />

R 7<br />

R 6<br />

R 2<br />

R 3<br />

R 2<br />

R 3<br />

R 2<br />

R 3<br />

R 1 R 8<br />

R 4<br />

NH HN<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

97<br />

N HN<br />

NH N<br />

R 5<br />

R 1 R 8<br />

R 4<br />

100<br />

R 5<br />

R 1 R 8<br />

R 4<br />

NH HN<br />

NH HN<br />

O<br />

103<br />

R 5<br />

R 7<br />

R 2<br />

R 1 R 8<br />

NH HN<br />

R 7<br />

N HN<br />

R3 R6 R6 R 7<br />

R 2<br />

R 4<br />

98<br />

NH N<br />

R 5<br />

R 1 R 8<br />

R 7<br />

N N<br />

R3 R6 R6 R 7<br />

R 6<br />

R 4<br />

101<br />

R 5<br />

for references see p 1223


1104 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a17.8.1.1.3.1 Method 1:<br />

Oxobilanes<br />

17.8.1.1.3.1.1 Variation 1:<br />

a-Oxobilanes<br />

Making use of one of the dipyrromethane syntheses mentioned earlier (Section<br />

17.8.1.1.1.2.1), reaction between the pyridinium salt 104 of a 1-(chloromethyl)dipyrroketone<br />

and the lithium salt 105 of a dipyrromethane-5-carboxylic acid gives a good yield of<br />

a-oxobilane-1,19-dibenzyl ester 106 (Scheme 23). [67,68] These compounds are easily isolated,<br />

crystallized, and characterized.<br />

Scheme 23 Synthesis of a-Oxobilanes [67,68]<br />

MeO 2C<br />

N + Cl −<br />

O<br />

NH HN<br />

104<br />

CO 2Me<br />

CO2Bn<br />

MeO 2C<br />

NH HN<br />

HCONH2, py, heat CO2Bn 35%<br />

CO2Bn NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

+<br />

−<br />

O2C CO2Bn Li NH HN<br />

+<br />

MeO 2C CO 2Me<br />

105<br />

O<br />

106<br />

CO 2Me<br />

MeO2C CO2Me<br />

Dibenzyl 3,8,13,17-Tetrakis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethyl-b-oxobilane-1,19-dicarboxylate<br />

(106); Typical Procedure: [68]<br />

Benzyl 9-(chloromethyl)-3,8-bis[2-(methoxycarbonyl)ethyl]-2,7-dimethyl-5-dipyrroketone-<br />

1-carboxylate (6.4 g, 11.79 mmol) was dissolved in pyridine (12 mL) with gentle warming<br />

on a water bath to give 104. 9-(Benzyloxycarbonyl)-3,7-bis[2-(methoxycarbonyl)ethyl]-2,8dimethyldipyrromethane-2-carboxylic<br />

acid (6.92 g, 13.19 mmol) was suspended in<br />

HCONH 2 (200 mL), LiOMe (502 mg, 13.21 mmol) was added, and the suspension was shaken<br />

until the solid had dissolved to give 105. The reactants were combined and the resulting<br />

clear amber soln was heated at 50 8C under N 2 for 18 h. An oil slowly deposited during<br />

heating to form a viscous lower layer. After a further 17 h at rt under N 2, the oil had partially<br />

solidified and the supernatant liquid was decanted. The gummy deposit was washed<br />

with H 2O and dissolved in CH 2Cl 2. The soln was washed with H 2O, dried (MgSO 4), and concentrated<br />

in vacuo. Reconcentration of the residue with Et 2O (50 mL) under reduced pressure<br />

yielded a brown foam, which was redissolved in Et 2O (100 mL) and set aside overnight<br />

under N 2. The crystallized material was collected by filtration, and washed with a<br />

little ice-cold Et 2O to give 106; yield: 4.07 g (35%); mp 123–1248C.


a17.8.1.1.3.1.2 Variation 2:<br />

b-Oxobilanes<br />

17.8.1 Porphyrins 1105<br />

b-Oxobilanes can be readily synthesized from dipyrromethane precursors using the Vilsmeier–Haack<br />

procedure, as was described earlier for the synthesis of simple dipyrroketones<br />

(Section 17.8.1.2.1). Thus, the phosphoryl chloride complex of a 1-[(dimethylamino)carbonyl]dipyrromethane<br />

107 reacts with a 1-unsubstituted dipyrromethane 87 (R 1 =H)<br />

to give a good yield of tetrapyrrolic imine salt 108; column chromatography at this stage<br />

enables the polar imine salt to be obtained in fairly pure form. Hydrolysis then gives the<br />

b-oxobilane-1,19-dibenzyl ester 109 (Scheme 24). [69] Catalytic debenzylation of 109 then<br />

gives a quantitative yield of the corresponding 1,19-dicarboxylic acid 110.<br />

Scheme 24 Synthesis of b-Oxobilanes [69]<br />

Et<br />

Me2N<br />

POCl 3<br />

CH 2Cl 2<br />

O<br />

NH HN<br />

107<br />

Et<br />

Me2N +<br />

Cl −<br />

Et<br />

CO2Bn<br />

+<br />

NH HN<br />

NH HN<br />

108<br />

NH HN<br />

MeO 2C CO 2Me<br />

Et<br />

CO 2Bn<br />

CO 2Bn<br />

MeO2C CO 2Me<br />

87 R 1 = H<br />

aq Na 2CO 3<br />

H2, Pd/C<br />

CO 2Bn<br />

NH HN<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

CH 2Cl 2<br />

86%<br />

Et<br />

O<br />

Et<br />

O<br />

NH HN<br />

NH HN<br />

109 45%<br />

110<br />

Et<br />

MeO 2C CO 2Me<br />

Et<br />

CO 2Bn<br />

CO 2Bn<br />

CO 2H<br />

CO 2H<br />

MeO2C CO2Me<br />

Dibenzyl 3,8-Diethyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethyl-b-oxobilane-1,19-dicarboxylate<br />

(109); Typical Procedure: [69]<br />

Benzyl 3,8-diethyl-9-[(dimethylamino)carbonyl]-2,7-dimethyldipyrromethane-1-carboxylate<br />

(107; 4.35 g, 9.99 mmol) was dissolved in POCl 3 (40 mL) and kept at 50 8C for 15 min.<br />

The excess of POCl 3 was removed by distillation under reduced pressure and dry 1,2-dibromoethane<br />

(10 mL) was added and distilled off to remove the POCl 3. The residual orange-brown<br />

oil was dissolved in CH 2Cl 2 and mixed with a soln of benzyl 3,7-bis[2-(meth-<br />

for references see p 1223


1106 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aoxycarbonyl)ethyl]-2,8-dimethyldipyrromethane-1-carboxylate [87, R 1 = H (obtained by<br />

decarboxylation of the corresponding 1-carboxylic acid); 5.5 g, 11.44 mmol]. The mixture<br />

was refluxed and a slow stream of N 2 gas was bubbled through it until spectrophotometry<br />

showed maximal intensity of a peak at 412 nm (e ca. 17,000 after 20–24 h). The mixture<br />

was washed with H 2O, dried (MgSO 4), and concentrated to dryness, and the residual<br />

brown oil (containing 108) was chromatographed (neutral alumina, 0–100% EtOAc in toluene).<br />

The column was finally stripped with MeOH and evaporation of the MeOH eluates<br />

gave 108 as a yellow oil. The oil was taken up in CH 2Cl 2 (50 mL) and hydrolyzed by vigorous<br />

refluxing (with stirring) with aq Na 2CO 3 (10%, 50 mL) for 90 min. The organic layer<br />

was separated and the aqueous phase was washed with CH 2Cl 2 (30 mL). The combined organic<br />

phases were washed with H 2O and dried (MgSO 4) before evaporation to dryness in<br />

vacuo. The residual oil, after chromatography [neutral alumina (Brockmann Grade III),<br />

toluene/EtOAc 4:1] was taken up in boiling MeOH (ca. 20 mL) and on cooling and standing<br />

overnight at rt, the b-oxobilane 109 was obtained; yield: 3.88 g (45%); mp 167–1698C.<br />

3,8-Diethyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethyl-b-oxobilane-1,19dicarboxylic<br />

Acid (110); Typical Procedure: [69]<br />

Dibenzyl 3,8-diethyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethyl-b-oxobilane-1,19-dicarboxylate<br />

(109; 1.0 g, 1.14 mmol) in THF (250 mL) containing Et 3N (0.05 mL)<br />

was hydrogenated at rt and atmospheric pressure over 10% Pd/C (250 mg) until uptake of<br />

H 2 had ceased. The catalyst was filtered off and after evaporation of the solvent, the residual<br />

yellow oil was taken up in warm THF (5 mL), and Et 2O (50 mL) was added. The dicarboxylic<br />

acid 110 separated as small needles; yield: 700 mg (86%); mp 178–1808C.<br />

17.8.1.1.3.2 Method 2:<br />

b-Bilenes<br />

17.8.1.1.3.2.1 Variation 1:<br />

1,19-Dimethyl-b-bilenes<br />

1,19-Dimethyl-b-bilenes can be prepared by two general methods (Scheme 25). The first is<br />

suited to the synthesis of fully alkylated bilenes, but gives low yields when bilenes bearing<br />

electron-withdrawing substituents are required; condensation of 2 equivalents of a<br />

2,3,4-trialkyl-1H-pyrrole (e.g., 74) with one of a 1,9-bis(methoxymethyl)dipyrromethene<br />

111 affords a b-bilene salt 112, which is usually symmetrical about the 10-carbon<br />

atom. [70–73] Better is the route in which a 1-formyldipyrromethane 113 is treated with a 1unsubstituted<br />

dipyrromethane (or with the corresponding 1-carboxylic acid) 114; [74] in<br />

this way unsymmetrically substituted b-bilenes 115 can be obtained in high yield, but it<br />

is advisable [75] to have an electronegative group on the terminal rings of the b-bilene in<br />

order to prevent acid-promoted side reactions which result in mixtures of porphyrins.<br />

Scheme 25 Syntheses of 1,19-Dimethyl-b-bilenes [70–74]<br />

Et<br />

Et<br />

NH<br />

+<br />

NH<br />

111<br />

OMe<br />

Br −<br />

OMe<br />

N<br />

H<br />

74<br />

(2 equiv), HBr, benzene<br />

84%<br />

NH HN<br />

+<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

112<br />

Br −


17.8.1 Porphyrins 1107<br />

OHC<br />

NH HN<br />

113<br />

a1. TFA, MeOH<br />

2. HBr, AcOH<br />

3. Et3N, CHCl3 Ac<br />

+<br />

R 1<br />

NH HN<br />

114 R 1 = H, CO2H<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Ac<br />

NH HN<br />

N HN<br />

7,13-Diethyl-1,2,3,8,12,17,18,19-octamethyl-b-bilene Hydrobromide (112); Typical<br />

Procedure: [71]<br />

2,8-Diethyl-1,9-bis(methoxymethyl)-3,7-dimethyldipyrromethene hydrobromide (111;<br />

3.0 g, 7.55 mmol) in benzene (30 mL) (CAUTION: carcinogen) and 2,3,4-trimethyl-1H-pyrrole<br />

(74; 1.9 g, 17.40 mmol) were refluxed for 30 min. Hot petroleum ether (10 mL) was<br />

then added and the soln was cooled. The hydrobromide precipitated, was washed with<br />

petroleum ether, and was then recrystallized (CHCl 3/petroleum ether) to give 112; yield:<br />

3.5 g (84%); mp 187 8C (dec).<br />

2,18-Diacetyl-1,3,7,8,12,13,17,19-octamethyl-b-bilene (115); Typical Procedure: [74]<br />

A soln of 8-acetyl-2,3,7,9-tetramethyldipyrromethane-1-carboxylic acid (114 R 1 =CO 2H;<br />

1 g, 3.47 mmol) and 2-acetyl-9-formyl-1,3,7,8-tetramethyldipyrromethane (113; 1g,<br />

3.67 mmol) in TFA (5 mL) was added to MeOH (50 mL). The mixture was stirred for<br />

10 min, a soln of HBr in AcOH (40%, 3 mL) was added dropwise, and the precipitate<br />

(115•HBr) was collected by filtration after 2 h and washed with MeOH. A portion of the<br />

product was dissolved in CHCl 3 and treated with an excess of Et 3N. The residue obtained<br />

by concentration of the soln was crystallized (pyridine/CHCl 3) to give the b-bilene 115 as<br />

yellow needles; mp 2018C (dec).<br />

17.8.1.1.3.2.2 Variation 2:<br />

b-Bilene-1,19-dicarboxylates<br />

A very useful approach [76,77] to b-bilene-1,19-dicarboxylates involves the condensation of a<br />

1-unsubstituted dipyrromethane 116 (R 1 = H) [or the corresponding dipyrromethane-1carboxylic<br />

acid, 116 (R 1 =CO 2H)] with a 1-formyldipyrromethane, e.g. 117, and provides<br />

high yields of b-bilene-1,19-di-tert-butyl esters 118 as highly crystalline hydrochloride<br />

salts (Scheme 26). b-Bilene-1,19-dicarboxylic acids are also transient intermediates in the<br />

synthesis of porphyrins from a-oxobilanes.<br />

115<br />

Ac<br />

Ac<br />

for references see p 1223


1108 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 26 Synthesis of b-Bilene-1,19-dicarboxylates [76]<br />

Et<br />

NH HN<br />

R 1 CO 2Bu t<br />

116 R 1 = H, CO 2H<br />

CO 2Me<br />

+<br />

NH HN<br />

NH HN<br />

+<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

OHC<br />

Et<br />

1. TsOH, MeOH, rt, 30 min<br />

2. HCl (g), CH2Cl2 85%<br />

117<br />

CO2Bu t<br />

Et<br />

Et<br />

CO2Me<br />

118<br />

CO 2Bu t<br />

CO 2Bu t<br />

CO 2Me<br />

CO2Me<br />

Di-tert-butyl 7,13-Diethyl-2,18-bis[2-(methoxycarbonyl)ethyl]-3,8,12,17-tetramethyl-b-bilene-1,19-dicarboxylate<br />

Hydrochloride (118); Typical Procedure: [76]<br />

tert-Butyl 7-ethyl-9-formyl-2-[2-(methoxycarbonyl)ethyl]-3,8-dimethyldipyrromethane-1carboxylate<br />

(117; 312 mg, 0.75 mmol) and 9-(tert-butoxycarbonyl)-3-ethyl-8-[2-(methoxycarbonyl)ethyl]-2,7-dimethyldipyrromethane-1-carboxylic<br />

acid (116, R 1 =CO 2H; 324 mg,<br />

0.75 mmol) in CH 2Cl 2 (80 mL) were treated with TsOH (600 mg) in MeOH (20 mL). After stirring<br />

at rt for 30 min the orange-red soln was washed with 2% aq Na 2CO 3 (100 mL) and then<br />

with H 2O. The soln was dried (MgSO 4) and concentrated to dryness to give a yellow-brown<br />

oil. Dry benzene (CAUTION: carcinogen) was added and removed under reduced pressure<br />

and then CH 2Cl 2 (30 mL) was added. Dry HCl gas was passed through the soln for a few<br />

seconds and dry benzene was quickly added to the deep red soln and then removed under<br />

reduced pressure. More dry benzene was added and the soln was concentrated to give a<br />

foam which was taken up in a small amount of Et 2O and refrigerated overnight. The b-bilene<br />

hydrochloride 118 was collected by filtration; yield: 512 mg (85%); mp 188–1898C.<br />

17.8.1.1.3.3 Method 3:<br />

a,c-Biladienes<br />

17.8.1.1.3.3.1 Variation 1:<br />

Symmetrical 1,19-Dimethyl-a,c-biladiene Salts<br />

Using the dipyrromethene synthetic methodology developed by Fischer (see earlier), [26] it<br />

has been shown that reaction of 1 equivalent of a dipyrromethane 119 with 2 equivalents<br />

of a 2-formyl-5-methyl-1H-pyrrole 120, in the presence of hydrogen bromide, gives a very<br />

high yield of the corresponding symmetrical a,c-biladiene dihydrobromide 121 (Scheme<br />

27). [71] Such salts crystallize readily from the reaction mixtures. Because of the stoichiometry<br />

used, it follows that rings A and D on the a,c-biladiene must be identical, and this introduces<br />

a symmetry restriction to the generalization of this approach. In fact, these a,cbiladienes<br />

are often completely symmetrical (e.g., 123) about the 10- (i.e., b) carbon atom<br />

because the starting dipyrromethane (e.g., 59) is also symmetrically substituted.<br />

Cl −


17.8.1 Porphyrins 1109<br />

aScheme 27 Synthesis of Symmetrical 1,19-Dimethyl-a,c-biladiene Salts [47,71]<br />

A<br />

B B<br />

N<br />

H<br />

57<br />

NH HN<br />

119<br />

CO2Bn<br />

NH HN<br />

BnO2C CO2Bn<br />

59<br />

A<br />

N<br />

H<br />

122<br />

C D<br />

N<br />

H<br />

120<br />

H2, Pd/C, THF, Et 3N<br />

85%<br />

H2, Pd/C<br />

THF, Et3N 100%<br />

(2 equiv), HBr<br />

CHO<br />

(2 equiv), HBr, TFA, MeOH<br />

CHO<br />

76%<br />

NH HN<br />

+<br />

+<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

H<br />

A<br />

D<br />

CO2H<br />

B B<br />

C<br />

121<br />

NH HN<br />

HO2C CO2H<br />

BzCl, DMF<br />

82%<br />

NH HN<br />

+<br />

+<br />

NH HN<br />

2-Formyl-3,4,5-trimethyl-1H-pyrrole (122); Typical Procedure: [47]<br />

Benzyl 3,4,5-trimethyl-1H-pyrrole-2-carboxylate (57; 5.0 g, 20.55 mmol) in THF (200 mL)<br />

containing Pd/C (500 mg) and Et 3N (0.1 mL) was hydrogenated at rt and atmospheric pressure<br />

until uptake of H 2 ceased (3 h). The soln was filtered through a bed of Celite, which<br />

was washed with THF (50 mL) and the combined filtrates were concentrated to dryness to<br />

give 3,4,5-trimethyl-1H-pyrrole-2-carboxylic acid; yield: 2.66 g (85%). This material (1.92 g,<br />

123<br />

C<br />

N<br />

H<br />

122<br />

A<br />

D<br />

2Br −<br />

CHO<br />

2Br −<br />

for references see p 1223


1110 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a12.53 mmol) was dissolved under N 2 in DMF (6 mL) and refluxed for 3 h before cooling to<br />

08C and addition of BzCl (3 mL). The mixture was stirred for an additional 2 h, after which<br />

toluene (15 mL) was added and the soln was stirred for 30 min at rt. A brown-yellow precipitate<br />

was collected by filtration, air dried, and suspended in sat. Na 2CO 3 (100 mL). The<br />

mixture was stirred for 3 h at 608C, and a yellow precipitate was collected. The product<br />

was recrystallized (CH 2Cl 2/hexanes) to give the formylpyrrole 122; yield: 1.42 g (82%); mp<br />

102–1048C.<br />

1,2,3,7,8,12,13,17,18,19-Decamethyl-a,c-biladiene Dihydrobromide (123); Typical<br />

Procedure: [47]<br />

Dibenzyl 2,3,7,8-tetramethyldipyrromethane-1,9-dicarboxylate (59; 480 mg, 1.02 mmol)<br />

and 10% Pd/C (50 mg) were added to THF (100 mL) containing Et 3N (0.1 mL) and the resulting<br />

soln was hydrogenated at rt and atmospheric pressure for 6 h. Filtration through a 2cm<br />

pad of Celite, followed by evaporation of the solvent, afforded 2,3,7,8-tetramethyldipyrromethane-1,9-dicarboxylic<br />

acid as a white solid which was used directly without further<br />

purification; yield: 297 mg. To the dicarboxylic acid (297 mg, 1.02 mmol) was added<br />

TFA (4 mL) and the mixture was stirred for 5 min before addition of a soln of 2-formyl-<br />

3,4,5-trimethylpyrrole (122; 297 mg, 2.16 mmol) in MeOH (20 mL). Upon addition of 122,<br />

the soln immediately became orange and to this was added 33% HBr in AcOH (1 mL). Et 2O<br />

(50 mL) was slowly added to the soln and the mixture was left in a refrigerator overnight.<br />

Dark orange crystals of 123 were collected by filtration; yield: 473 mg (76%), mp >300 8C.<br />

17.8.1.1.3.3.2 Variation 2:<br />

Unsymmetrical 1,19-Dimethyl-a,c-biladiene Salts<br />

The symmetry restrictions in the above a,c-biladiene approach were solved by the development<br />

of stepwise methods for the synthesis of unsymmetrically substituted a,c-biladiene<br />

salts. [78–80] Two complementary methods are now available, known as the “clockwise”<br />

and “counterclockwise” routes (Scheme 28). In the counterclockwise route, [78,79] catalytic<br />

debenzylation of a benzyl tert-butyl dipyrromethane-1,9-dicarboxylate 124 provides<br />

the corresponding monocarboxylic acid 116 (R 1 =CO 2H), which can then be treated<br />

under mild conditions (4-toluenesulfonic acid/methanol) with a 2-formyl-5-methyl-1Hpyrrole<br />

125 to provide a high yield of a moderately stable tripyrrene salt 126 after anion<br />

exchange with hydrogen bromide (which, if not done very carefully, might cause cleavage<br />

of the tert-butyl ester on the tripyrrene). After treatment with trifluoroacetic acid (to<br />

cleave the tert-butyl ester) and then with a second 2-formyl-1H-pyrrole 127, a high yield<br />

of a fully unsymmetrical 1,19-dimethyl-a,c-biladiene salt 128 is obtained.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


17.8.1 Porphyrins 1111<br />

aScheme 28 Synthesis of Unsymmetrical 1,19-Dimethyl-a,c-biladiene Salts via “Clockwise”<br />

and “Counterclockwise” Routes [78–80]<br />

counterclockwise<br />

BnO2C<br />

Et<br />

NH HN<br />

124<br />

MeO2C<br />

CO2Bu t<br />

N<br />

H<br />

125<br />

1. TFA<br />

MeO2C 2.<br />

OHC<br />

CO 2Me<br />

80%<br />

H2, Pd/C<br />

, TsOH, MeOH, CH2Cl2 CHO<br />

N<br />

H<br />

127<br />

then HBr, AcOH<br />

89%<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

95%<br />

CO2Me<br />

, MeOH<br />

HO2C<br />

Et<br />

Et<br />

Et<br />

MeO2C<br />

CO 2Me<br />

NH<br />

+<br />

NH<br />

CO2Bu t<br />

116 R 1 = CO 2H<br />

HN<br />

NH HN<br />

+<br />

+<br />

NH HN<br />

OTs −<br />

126<br />

128<br />

CO 2Bu t<br />

CO 2Me<br />

CO2Me<br />

CO 2Me<br />

CO 2Me<br />

CO 2Me<br />

2Br −<br />

for references see p 1223


1112 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

clockwise<br />

aNH HN<br />

BnO2C<br />

Et<br />

124<br />

MeO2C<br />

OHC<br />

CO 2Bu t<br />

N<br />

H<br />

127<br />

CO 2Me<br />

CO 2Me<br />

1. H2SO4, HBr<br />

MeO2C 2.<br />

, HBr<br />

N<br />

H<br />

125<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

TFA<br />

CHO<br />

BnO 2C<br />

BnO2C<br />

Et<br />

Et<br />

NH HN<br />

Et<br />

CO2Me<br />

NH HN<br />

+<br />

HN<br />

MeO2C<br />

NH HN<br />

+<br />

+<br />

NH HN<br />

128<br />

CO 2Me<br />

Br −<br />

CO 2Me<br />

CO 2Me<br />

CO 2Me<br />

CO 2Me<br />

CO 2Me<br />

In the alternative (more recent) “clockwise” approach (shown in Scheme 29), [80] the benzyl<br />

tert-butyl dipyrromethane-1,9-dicarboxylate 129 is treated with trifluoroacetic acid<br />

and a 2-formyl-1H-pyrrole 125 to give the very stable tripyrrene benzyl ester 130, after<br />

anion exchange using hydrogen bromide gas. Catalytic hydrogenation could not be used<br />

to cleave the benzyl ester because this would concomitantly reduce the tripyrrene to a<br />

tripyrrane, thereby causing it to be sensitive to pyrrole ring scrambling in acid. This problem<br />

is avoided by effecting acidic cleavage (TFA/HBr in AcOH) of the benzyl ester. [80,81] Subsequent<br />

treatment with pyrrole 131 gave the a,c-biladiene dihydrobromide 132 in good<br />

yield.<br />

2Br −


17.8.1 Porphyrins 1113<br />

aScheme 29 Synthesis of Unsymmetrical 1,19-Dimethyl-a,c-biladiene Salts via a<br />

“Clockwise” Route [80,81]<br />

BnO 2C<br />

Et<br />

NH HN<br />

129<br />

MeO 2C<br />

Et<br />

CO2Bu t<br />

1. AcOH, HBr, TFA<br />

Cl<br />

2.<br />

TFA<br />

1. , MeOH, 90 min<br />

N<br />

H<br />

CHO<br />

125<br />

2. HBr, AcOH, Et2O, 15 min<br />

N<br />

H<br />

131<br />

BnO2C<br />

NH HN<br />

NH HN<br />

+<br />

+<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

73%<br />

CHO<br />

BnO2C<br />

, MeOH<br />

Et<br />

NH HN<br />

Et<br />

Cl<br />

Et<br />

+<br />

HN<br />

130 83%<br />

Et<br />

132<br />

Et<br />

Br −<br />

CO 2Me<br />

Et<br />

2Br −<br />

CO2Me<br />

Engel and Gossauer have also reported the synthesis of an a,c-biladiene by way of a tertbutyl<br />

tripyrrenecarboxylate; this is then converted into a porphyrin, as well as into metal<br />

tetradehydrocorrinate salts. [82]<br />

Tripyrrenes can apparently be obtained without the need for differential protection<br />

of the two (1,9) ends of the initial dipyrromethane (Scheme 30). [83] Thus, condensation of a<br />

1,9-di-unsubstituted dipyrromethane 133 with a 2-formyl-5-methyl-1H-pyrrole, such as<br />

125, gives very high yields of the corresponding tripyrrene salt 134; quite unexpectedly,<br />

these are apparently uncontaminated with byproducts due to reaction of the formylpyrrole<br />

125 at both ends of 133. The tripyrrene intermediates can then be treated with a different<br />

formylpyrrole 135 to give high yields of the appropriate a,c-biladiene 136 in a manner<br />

which accomplishes the same objectives as the “clockwise” and “counterclockwise”<br />

a,c-biladiene approaches reported previously, but with symmetry restrictions which the<br />

other methods do not have.<br />

for references see p 1223


1114 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 30 Symmetry-Restricted Unsymmetrical a,c-Biladiene Synthesis [83]<br />

A<br />

B B<br />

R 1 = I, H, CO 2H<br />

NH HN<br />

133<br />

A<br />

MeO 2C<br />

R 1<br />

MeO2C<br />

N<br />

H<br />

135<br />

NH HN<br />

+<br />

+<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

H<br />

125<br />

, HBr<br />

CHO<br />

, HBr<br />

CHO<br />

A<br />

A<br />

MeO 2C<br />

B B<br />

NH HN<br />

+<br />

HN<br />

134<br />

B B<br />

R 1<br />

136<br />

A<br />

Br −<br />

CO 2Me<br />

A<br />

2Br −<br />

CO2Me<br />

tert-Butyl 8-Ethyl-2,13-bis[2-(methoxycarbonyl)ethyl]-1,3,7,12-tetramethyl-5-tripyrrene-<br />

14-carboxylate 4-Toluenesulfonate (126) and Hydrobromide Salts; Typical Procedure: [79]<br />

Benzyl 9-(tert-butoxycarbonyl)-3-ethyl-8-[2-(methoxycarbonyl)ethyl]-2,7-dimethyldipyrromethane-1-carboxylate<br />

(124; 2.0 g, 3.83 mmol) in THF (70 mL) containing Et 3N (0.2 mL)<br />

and 10% Pd/C (270 mg) was hydrogenated at rt and atmospheric pressure until uptake of<br />

H 2 ceased. After filtration through Celite the solvent was removed under reduced pressure<br />

to give, after crystallization (THF/hexanes), the dipyrromethane-1-carboxylic acid<br />

116; yield: 1.6 g (95%); mp dec. Compound 116 (788 mg, 1.82 mmol) and 2-formyl-4-[2-<br />

(methoxycarbonyl)ethyl]-3,5-dimethyl-1H-pyrrole (125; 381 mg, 2.10 mmol) in CH 2Cl 2<br />

(100 mL) were stirred with a soln of TsOH (800 mg) in MeOH (10 mL) over a period of<br />

30 min (monitoring by spectrophotometry, observing maximization of the 492-nm absorption<br />

band). The soln was washed with aq Na 2CO 3 (150 mL) and then dried (Na 2SO 4)<br />

and concentrated to dryness. Crystallization (CH 2Cl 2/Et 2O) gave the tripyrrene 4-toluenesulfonate<br />

126; yield: 1.09 g (80%); mp >1358C (dec). Tripyrrene salts are best isolated as<br />

the hydrobromide salts by brief treatment of the 4-toluenesulfonate in CH 2Cl 2 with sat.<br />

aq Na 2CO 3, evaporation, and then brief treatment of a CH 2Cl 2 soln of the tripyrrene freebase<br />

with dry HBr gas, followed by rapid evaporation, azeotropic removal of HBr gas and<br />

moisture with benzene (CAUTION: carcinogen), and crystallization from CH 2Cl 2/Et 2O.<br />

8-Ethyl-2,13,17-tris[2-(methoxycarbonyl)ethyl]-18-[(methoxycarbonyl)methyl]-1,3,7,12,19pentamethyl-a,c-biladiene<br />

Dihydrobromide (128); Typical Procedure: [81]<br />

tert-Butyl 8-ethyl-2,13-bis[2-(methoxycarbonyl)ethyl]-1,3,7,12-tetramethyl-5-tripyrrene-14carboxylate<br />

4-toluenesulfonate (126; 500 mg, 0.665 mmol) in TFA (5 mL) was stirred before<br />

addition of 2-formyl-3-[2-(methoxycarbonyl)ethyl]-4-[(methoxycarbonyl)methyl]-5methyl-1H-pyrrole<br />

(127; 217 mg, 0.812 mmol) in MeOH (7 mL) and then 45% HBr/AcOH<br />

(2 mL). After stirring for 30 min, Et 2O (100 mL) was added dropwise with continued stir-


17.8.1 Porphyrins 1115<br />

aring. The a,c-biladiene dihydrobromide was collected by filtration and washed thoroughly<br />

with Et 2O to give 128 as red-brown crystals; yield: 525 mg (89%); mp >150 8C (dec).<br />

Benzyl 7,12-Diethyl-2-[2-(methoxycarbonyl)ethyl]-1,3,8,13-tetramethyl-5-tripyrrene-14carboxylate<br />

Hydrobromide (130); Typical Procedure: [80]<br />

Benzyl 9-(tert-butoxycarbonyl)-3,8-diethyl-2,7-dimethyldipyrromethane-1-carboxylate<br />

(129; 400 mg, 0.86 mmol) was treated with TFA (3 mL) with stirring under N 2 at rt for<br />

5 min. Then, 2-formyl-4-[2-(methoxycarbonyl)ethyl]-3,5-dimethyl-1H-pyrrole (125;<br />

180 mg, 0.87 mmol) in MeOH (20 mL) was added all at once. The red soln was stirred for<br />

90 min then 30% HBr/AcOH (3 drops) and Et 2O (25 mL) were added. After stirring for another<br />

15 min reddish-orange crystals appeared. These were collected by filtration and washed<br />

thoroughly with Et 2O to give the tripyrrene hydrobromide 130; yield: 461 mg (83%); mp<br />

115–1168C.<br />

2-(2-Chloroethyl)-8,13-diethyl-18-[2-(methoxycarbonyl)ethyl]-1,3,7,12,17,19-hexamethyla,c-biladiene<br />

Dihydrobromide (132); Typical Procedure: [80]<br />

Benzyl 7,12-diethyl-2-[2-(methoxycarbonyl)ethyl]-1,3,8,13-tetramethyl-5-tripyrrene-14carboxylate<br />

hydrobromide (130; 107 mg, 0.17 mmol) was treated with a mixture of 30%<br />

HBr/AcOH (0.5 mL) and TFA (2.5 mL) with stirring under N 2 at rt for 6 h. Then, 4-(2-chloroethyl)-2-formyl-3,5-dimethyl-1H-pyrrole<br />

(131; 33 mg, 0.18 mmol) in MeOH (10 mL) was<br />

added all at once. The orange soln immediately turned dark red in color. After stirring<br />

for 30 min, Et 2O (30 mL) was added rapidly but dropwise to form red crystals. Collection<br />

by filtration and washing thoroughly with Et 2O gave the a,c-biladiene dihydrobromide<br />

132; yield: 92 mg (73%); mp 195–1968C.<br />

17.8.1.1.3.3.3 Variation 3:<br />

1-Bromo-19-methyl-a,c-biladiene Salts<br />

1-Bromo-19-methyl-a,c-biladienes 139 have been shown by Johnson and co-workers to be<br />

very useful precursors in porphyrin syntheses. [84] They are prepared, as shown in Scheme<br />

31, by treatment of 1-bromo-9-(bromomethyl)dipyrromethene salts (e.g., 137) with 9-unsubstituted<br />

1-methyldipyrromethenes 138 in the presence of a Friedel–Crafts catalyst<br />

[usually tin(IV) chloride]; this reaction gives very high yields of the biladiene after removal<br />

of a templating tin ion [inserted by the tin(IV) chloride] with hydrobromic acid.<br />

Scheme 31 Unsymmetrical 1-Bromo-19-methyl-a,c-biladiene Synthesis [84]<br />

Et<br />

Br<br />

NH HN<br />

+<br />

137<br />

Br<br />

CO 2Me<br />

Br −<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

+<br />

Et<br />

+<br />

NH HN<br />

138<br />

Et<br />

1. SnCl 4, CH 2Cl 2, rt, 2 h<br />

2. HBr, MeOH Br<br />

81%<br />

Et<br />

NH<br />

+<br />

NH<br />

139<br />

HN<br />

+<br />

HN<br />

CO2Me<br />

Br −<br />

CO2Me<br />

2Br −<br />

CO 2Me<br />

for references see p 1223


1116 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a1-Bromo-2,17-diethyl-8,12-bis[2-(methoxycarbonyl)ethyl]-3,7,13,18,19-pentamethyl-a,cbiladiene<br />

Dihydrobromide (139); General Procedure: [84]<br />

Millimolar proportions of the 1-bromo-9-(bromomethyl)dipyrromethene hydrobromide<br />

{1-bromo-9-(bromomethyl)-2-ethyl-8-[2-(methoxycarbonyl)ethyl]-3,7-dimethyldipyrromethene<br />

hydrobromide (137)} and 1-unsubstituted 9-methyldipyrromethene hydrobromide<br />

{7-ethyl-2-[2-(methoxycarbonyl)ethyl]-3,8,9-trimethyldipyrromethene hydrobromide<br />

(138)} were suspended in CH 2Cl 2 (50 mL) and SnCl 4 (1 mL) was added. The dry orange<br />

soln was kept at rt for 2 h before the solvent was removed under reduced pressure. The<br />

residue was treated with 20% HBr/MeOH (50 mL) and after standing for 15 min the a,c-biladiene<br />

dihydrobromide was separated and washed with a little MeOH containing a few<br />

drops of HBr. It was then washed with Et 2O and dried to give the a,c-biladiene dihydrobromide<br />

139; yield: 81%; mp >300 8C.<br />

17.8.1.2 Syntheses of Porphyrins<br />

17.8.1.2.1 Method 1:<br />

From Monopyrrole Tetramerization<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, p 581.<br />

The stoichiometry of this reaction is as shown in Scheme 7, case A. The best known<br />

and most used monopyrrole polymerization route to porphyrins yields 5,10,15,20-tetraphenylporphyrin<br />

(TPP, 22; Scheme 32) as the product. The original method, due to Rothemund,<br />

[85,86] employed a high-temperature reaction between pyrrole and benzaldehyde in<br />

pyridine, in a sealed tube. This reaction was modified by Adler, Longo, and colleagues, [87]<br />

by using hot propanoic acid as a solvent for the pyrrole and benzaldehyde (at atmospheric<br />

pressure). The reaction was finally optimized by Lindsey s group, [19,88] and in their hands<br />

involves a two-step reaction, (i) an acid-catalyzed macrocyclization to give tetraarylporphyrinogen,<br />

followed by (ii) an oxidation step usually using a high-potential quinone<br />

such as 2,3-dichloro-5,6-dicyanobenzo-1,4-quinone (DDQ).<br />

The crude product from the Rothemund and Adler/Longo approach usually contains<br />

[89] between 2 and 5% of meso-tetraphenylchlorin 140, which is best dealt with by<br />

brief treatment [90,91] of the crude product with 2,3-dichloro-5,6-dicyanobenzo-1,4-quinone;<br />

this accomplishes transformation of the chlorin into 5,10,15,20-tetraphenylporphyrin<br />

(22) and is much easier than attempting to chromatographically separate one<br />

from the other. The newer modifications to the Rothemund reaction enable a large variety<br />

of 5,10,15,20-tetraarylporphyrins to be prepared, [19] and also permit syntheses of highly<br />

non-planar dodecasubstituted porphyrins. [92–96]<br />

Scheme 32 Synthesis of Chlorin-Free 5,10,15,20-Tetraphenylporphyrin [91]<br />

N<br />

H<br />

PhCHO, EtCO2H<br />

heat<br />

26%<br />

Ph<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Ph<br />

Ph<br />

22<br />

Ph<br />

+<br />

Ph<br />

DDQ, CHCl 3, benzene<br />

94−96%<br />

Ph<br />

NH N<br />

N HN<br />

Ph<br />

140<br />

Ph


17.8.1 Porphyrins 1117<br />

aSince it involves some demanding pyrrole chemistry, the approach to octaalkyl-type porphyrins<br />

by tetramerization (or similar chemistry) of monopyrroles is more complicated.<br />

To ensure that only one porphyrin product is obtained, the substituents at the 3- and 4positions<br />

of the monopyrrole must be identical. In this way, symmetrical porphyrins<br />

such as 2,3,7,8,12,13,17,18-octaethylporphyrin (OEP, 21) can be prepared.<br />

Two primary approaches have been developed. The first (Scheme 33) involves polymerization<br />

of 2,5-di-unsubstituted 1H-pyrroles in the presence of agents which will provide<br />

the four carbons required for the meso-methine carbons of the product; this stoichiometry<br />

is as shown in Scheme 7, case A. Macrocyclization of such -free pyrroles with formic<br />

acid [97] is a viable method; an improved synthesis of 3,4-diethyl-1H-pyrrole has been<br />

reported, and this has been coupled with a formic acid method for the synthesis of<br />

2,3,7,8,12,13,17,18-octaethylporphyrin (21) to give yields between 55–75%, depending<br />

upon the scale. [98]<br />

The second approach (Scheme 33) is the tetramerization of pyrroles bearing 2-CH 2R 1<br />

substituents, the methylene carbon of which will eventually be the source of the 5,10,15<br />

and 20-carbons of the desired porphyrin; this is an example of the mode shown in<br />

Scheme 7, case B. Tetramerization of monopyrroles bearing an appropriately modified<br />

2-methyl substituent (which can provide the porphyrin interpyrrolic carbons), followed<br />

by aerial oxidation, often affords good yields of symmetrical porphyrins. This approach<br />

is also best illustrated by syntheses of 2,3,7,8,12,13,17,18-octaethylporphyrin (21). A typical<br />

example is the Mannich reaction of 3,4-diethyl-1H-pyrrole with dimethylamine and<br />

formaldehyde [or with commercially available (N,N-dimethylmethylene)ammonium iodide<br />

(Eschenmoser s reagent) [99,100] ] to give the 2-[(dimethylamino)methyl]-1H-pyrrole<br />

141; when this is heated in refluxing acetic acid a 52% yield of 21 is obtained. [101] Alternatively,<br />

hydrolysis of the pyrrole 142 gives pyrrole 143 which can be converted into<br />

2,3,7,8,12,13,17,18-octaethylporphyrin (21) in 44% yield by heating in acetic acid in the<br />

presence of potassium ferricyanide. [102] The introduction of the Barton–Zard pyrrole synthesis<br />

[103,104] has enabled ready access to monopyrroles of the type 62 in high yield; reduction,<br />

followed by acid-catalyzed cyclotetramerization of the resulting 1H-pyrrole-2-carbinol<br />

144 affords excellent yields of 2,3,7,8,12,13,17,18-octaethylporphyrin (21); [105,106] a potentially<br />

troublesome cleavage of the 2-CH 2OH group is circumvented by addition of dimethoxymethane<br />

(“methalal”).<br />

Scheme 33 Syntheses of 2,3,7,8,12,13,17,18-Octaethylporphyrin [98,101,102,104,105]<br />

Et Et<br />

N<br />

H<br />

HCO 2H<br />

55−75%<br />

Et<br />

Et<br />

Et Et<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

21<br />

Et<br />

Et<br />

Et<br />

for references see p 1223


1118 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

Et Et<br />

Et<br />

aNH N<br />

H2C NMe2 I<br />

N HN<br />

Et<br />

−<br />

Et Et<br />

Et Et<br />

+<br />

AcOH, heat<br />

NMe2 100%<br />

52%<br />

N<br />

N<br />

H<br />

H<br />

AcO<br />

Et Et<br />

N<br />

H<br />

62<br />

Et Et<br />

N<br />

H<br />

142<br />

CO 2Et<br />

CO2Et<br />

LiAlH4<br />

THF<br />

141<br />

KOH, MeOH<br />

reflux, 4 h<br />

96%<br />

Et Et<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

H<br />

144<br />

HO<br />

Et Et<br />

44%<br />

N<br />

H<br />

AcOH<br />

K3[Fe(CN) 6]<br />

heat<br />

OH<br />

143<br />

CO 2H<br />

Et<br />

Et<br />

Et<br />

Et<br />

21<br />

NH N<br />

N HN<br />

Et<br />

Et Et<br />

Et<br />

MeO OMe<br />

TsOH, CH2Cl2, rt, 12 h<br />

Et<br />

21<br />

NH N<br />

N HN<br />

Et<br />

Et Et<br />

Et<br />

21 55%<br />

If the substituents at the pyrrole 3- and 4-positions are not identical, then mixtures can<br />

result. It is not immediately obvious why this should happen, since one might expect<br />

that a stepwise polymerization of four pyrroles should maintain the ordering of the peripheral<br />

substituents. Indeed, self-condensations of pyrroles such as 145 proceed through<br />

dipyrromethanes (e.g., 146), tripyrranes 147, bilanes 148, even higher oligomers, and<br />

porphyrinogens 149 (Scheme 34). Unfortunately, all these species, under the acid conditions<br />

essential for the reaction, can scramble to give statistical mixtures of porphyrins<br />

such as (for the specific case of pyrrole 145) the four “primary type-isomers” (7–10,<br />

Scheme 3) of the etioporphyrins. The porphyrinogen scrambling process was investigated<br />

in an early paper by Mauzerall. [107] However, it should be pointed out that the enzymemediated<br />

tetramerization of porphobilinogen (150) to give uroporphyrinogen III (151)<br />

(Scheme 34) is uniquely regiospecific under normal circumstances. [108–112]<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et


17.8.1 Porphyrins 1119<br />

aScheme 34 Monopyrrole Oligomerization<br />

2<br />

HO 2C<br />

N<br />

H<br />

145<br />

Et<br />

N<br />

H<br />

150<br />

X<br />

Et<br />

CO 2H<br />

− HX<br />

NH HN<br />

X<br />

NH 2<br />

147<br />

HN<br />

Et<br />

enzymes<br />

Et<br />

Et<br />

NH HN<br />

NH HN<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

146<br />

N<br />

H<br />

145<br />

Et<br />

X<br />

HO2C HO2C<br />

X<br />

, − HX<br />

− HX<br />

Et<br />

NH HN<br />

NH HN<br />

Et<br />

X<br />

Et<br />

N<br />

H<br />

145<br />

Et<br />

Et<br />

Et<br />

X<br />

CO 2H<br />

, − HX<br />

NH HN<br />

NH HN<br />

148<br />

CO 2H<br />

149<br />

HO2C<br />

CO2H HO2C CO2H 151<br />

Some ingenious methods have been developed to enable the type-I isomer of a porphyrin<br />

to be obtained by rational acid-catalyzed tetramerization of a monopyrrole. For example,<br />

when perfluoroalkyl-1H-pyrrole-2-carbinols are tetramerized, exclusively the type-I isomer<br />

is formed. [106] Large steric hindrance by one of the two b-substituents on a pyrrole<br />

(e.g., 152) has also been shown to result in the formation of only the type-I porphyrin<br />

153 (Scheme 35). [113]<br />

A more general solution to the acid-catalyzed pyrrole redistribution reaction can be<br />

achieved simply by avoidance of the acid catalysts usually used in the monopyrrole selfcondensation.<br />

For example, treatment of 2-[(dialkylamino)methyl]-1H-pyrroles (such as<br />

155, obtained from pyrroles 154 by catalytic hydrogenation) with iodomethane (to afford<br />

156) gives a 25% yield of pure etioporphyrin I (7) (Scheme 35); the iodomethane quaterni-<br />

Et<br />

Et<br />

Et<br />

Et<br />

for references see p 1223


1120 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

azes the amino group to produce an excellent “benzylic” leaving group which can react at<br />

the unsubstituted pyrrole 5-position to give porphyrinogen 149 under neutral conditions.<br />

Porphyrin 7 can then be obtained by way of in situ oxidation using potassium ferricyanide.<br />

[114]<br />

Scheme 35 Unsymmetrical Monopyrrole Oligomerizations [113,114]<br />

Pr i<br />

N<br />

H<br />

BnO 2C<br />

OH<br />

152<br />

MeOH, Et3N<br />

R 1 = Me, Et<br />

N<br />

H<br />

Pr i<br />

154<br />

Pr i<br />

Et<br />

NR 1 2<br />

Et<br />

H2 Pd/C<br />

THF<br />

Et<br />

Pr i<br />

NH HN<br />

NH HN<br />

149<br />

HO 2C<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

N<br />

H<br />

Pr i<br />

155<br />

Et<br />

Pr i<br />

Pr i<br />

Et<br />

Pr i<br />

NR 1 2<br />

Pr i<br />

Pr i<br />

153<br />

MeI<br />

CH2Cl2 K3[Fe(CN) 6]<br />

MeOH, Et3N<br />

Et<br />

Pr i<br />

Pr i<br />

Pr i<br />

Pr i<br />

HO 2C<br />

Et<br />

N<br />

H<br />

156<br />

Pr i<br />

NH N<br />

N HN<br />

7 36%<br />

Et<br />

NR1 +<br />

2Me<br />

An interesting one-pot procedure which uses the mode of synthesis illustrated in Scheme<br />

7, case C, has been reported. [115,116] This induces two different pyrroles to react together, in<br />

one flask, but to produce only one regiochemically pure porphyrin. For example, treatment<br />

of the 2,5-bis[(dimethylamino)methyl]-1H-pyrrole (157) with 3,4-dimethyl-1H-pyrrole<br />

(158) in methanol containing potassium ferricyanide affords the porphyrin 160<br />

with two pairs of identical pyrrole rings opposite each other (Scheme 36). The avoidance<br />

of acid catalysis and use of an in situ oxidant [potassium ferricyanide] serves to circumvent<br />

any pyrrole ring redistribution reactions in the intermediate species 159 prior to formation<br />

of porphyrin 160.<br />

Et<br />

I −<br />

Et


17.8.1 Porphyrins 1121<br />

aScheme 36 One-Pot Porphyrin Synthesis with Two Pyrroles [115,116]<br />

Et Et<br />

N<br />

H<br />

+<br />

H2C NMe2 MeNO2<br />

86%<br />

Et<br />

Et<br />

I −<br />

NH HN<br />

NH HN<br />

159<br />

Me 2N<br />

Et<br />

Et Et<br />

N<br />

H NMe 2<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

157<br />

Et<br />

N<br />

H<br />

158<br />

Et<br />

, K 3[Fe(CN) 6], MeOH, heat<br />

Et<br />

NH N<br />

N HN<br />

160 19.5%<br />

Chlorin-Free 5,10,15,20-Tetraphenylporphyrin (TPP, 22); Typical Procedure: [91]<br />

Benzaldehyde (66.5 mL, 0.63 mol) and 1H-pyrrole (46.5 mL, 0.69 mol) were added simultaneously<br />

to refluxing propanoic acid (2.5 L) and the mixture was refluxed for a further<br />

30 min before being set aside at rt. The product was collected by filtration, washed with<br />

hot H 2O and then with MeOH to give crystals of crude 5,10,15,20-tetraphenylporphyrin<br />

(22); yield: 20.4 g (19.8%). Concentration of the propanoic acid soln afforded another<br />

crop of crude 5,10,15,20-tetraphenylporphyrin (22); yield: 6.0 g (26% overall). The crude<br />

5,10,15,20-tetraphenylporphyrin (20 g, 32.5 mmol) in EtOH-free CHCl 3 (2.5 L) was treated<br />

with DDQ (5 g) in dry benzene (150 mL) (CAUTION: carcinogen). The mixture was refluxed<br />

for 3 h before filtration of the yellow-red soln under suction through a sintered glass funnel<br />

(6.4 cm diameter ” 20.4 cm) containing alumina (100–300 g, Brockmann Grade I). The<br />

alumina was washed with CH 2Cl 2 (200 mL) and the combined filtrates were concentrated<br />

to ca. 200 mL before addition of MeOH (200 mL). Filtration afforded chlorin-free<br />

5,10,15,20-tetraphenylporphyrin (22); yield: 19.2 g (96%); mp >300 8C. In a smaller scale<br />

reaction [5,10,15,20-tetraphenylporphyrin (22; 1g),CH 2Cl 2 (250 mL), DDQ (250 mg), benzene<br />

(15 mL)], only 30 min reflux was required, to give pure 5,10,15,20-tetraphenylporphyrin<br />

(22); yield: 940 mg (94%).<br />

2-[(Dimethylamino)methyl]-3,4-diethyl-1H-pyrrole (141); Typical Procedure: [101]<br />

3,4-Diethyl-1H-pyrrole (41.7 g, 0.338 mol) in MeOH (325 mL) under N 2 was cooled to –158C<br />

in an acetone/dry ice bath. A soln of Me 2N•HCl (28.1 g, 0.344 mol), KOAc (33.8 g,<br />

0.344 mol), and 37% aq HCHO (27.8 g, 0.343 mol) in H 2O (130 mL) was added dropwise<br />

over 1.75 h, the soln temperature being kept between –15 and –108C. After complete addition,<br />

the soln was stirred at –10 8C for 30 min and then kept at 08C for 12 h. Cold 5% aq<br />

HCl (400 mL) was slowly added with stirring and the cold soln was extracted with Et 2O.<br />

The aqueous layer was made basic by slow addition of 2 N NaOH (500 mL) while stirring<br />

and cooling in an ice bath. The basic soln was extracted with Et 2O (3 ” 500 mL) and the<br />

combined Et 2O layers were dried (K 2CO 3) and concentrated under reduced pressure to afford<br />

141 as a brown oil, pure by 1 H NMR spectroscopy; yield: 61.0 g (100%).<br />

Et<br />

Et<br />

for references see p 1223


1122 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a2,3,7,8,12,13,17,18-Octaethylporphyrin (21) from 2-[(Dimethylamino)methyl]-3,4-diethyl-<br />

1H-pyrrole (141); Typical Procedure: [101]<br />

2-[(Dimethylamino)methyl]-3,4-diethyl-1H-pyrrole (141; 61.0 g, 0.338 mol) was taken up in<br />

pure AcOH (500 mL). The soln was refluxed for 1 h with rapid stirring and with a strong<br />

stream of O 2 bubbling through it. Within minutes a dark precipitate appeared. The soln<br />

was cooled, the AcOH was removed under reduced pressure, and MeOH (500 mL) was added<br />

to the residue. The slurry was stirred and collected by filtration, and the precipitate<br />

was washed with MeOH until it was no longer brown. After drying, crude 21 was recrystallized<br />

(toluene); yield: 23.51 g (52%); mp 324–3258C.<br />

3,4-Diethyl-5-hydroxymethyl-1H-pyrrole-2-carboxylic Acid (143); Typical Procedure: [102]<br />

A mixture of ethyl 5-(acetoxymethyl)-3,4-diethyl-1H-pyrrole-2-carboxylate (142; 15.2 g,<br />

56.86 mmol) and KOH (30 g) in MeOH (210 mL) was refluxed for 4 h. The soln was concentrated<br />

under reduced pressure and H 2O (150 mL) was added. The soln was extracted with<br />

Et 2O, and the aqueous phase was then acidified with dil aq HCl (350 mL, dropwise with<br />

cooling and stirring). The product was collected by filtration, washed with aq NaOAc,<br />

and H 2O, then dried to give the pyrrole 143; yield: 10.7 g (96%); mp 115–1208C.<br />

2,3,7,8,12,13,17,18-Octaethylporphyrin (21) from 3,4-Diethyl-5-hydroxymethyl-1H-pyrrole-2-carboxylic<br />

Acid (143); Typical Procedure: [102]<br />

3,4-Diethyl-5-hydroxymethyl-1H-pyrrole-2-carboxylic acid (143; 10 g, 50.7 mmol) in AcOH<br />

(40 mL) containing K 3[Fe(CN) 6] (1 g, 3.03 mmol) was heated at 1008C with stirring for 1 h.<br />

After standing overnight at rt, the product was collected by filtration to give 21 (2 g). The<br />

filtrate was concentrated under vacuum, and purified by chromatography [alumina<br />

(Brockmann Grade III), CH 2Cl 2] to give a second crop (1.0 g) of 21; total yield: 3.0 g (44%).<br />

Ethyl 3,4-Diethyl-1H-pyrrole-2-carboxylate (62); Typical Procedure: [105]<br />

A mixture of 3-acetoxy-4-nitrohexane (16.3 g, 0.083 mol), ethyl isocyanoacetate (9.8 g,<br />

0.087 mol), and DBU (26.4 g, 0.17 mole) in THF (100 mL) was stirred at 208C for 12 h. The<br />

mixture was poured into H 2O containing 1 M HCl and extracted with EtOAc. The extracts<br />

were washed with H 2O and dried (MgSO 4). Concentration of the solvents under reduced<br />

pressure gave a residue, which was chromatographed (silica gel, hexanes/CH 2Cl 2). Concentration<br />

of the eluates gave 62 as an oil; yield: 14.2 g (86%).<br />

2,3,7,8,12,13,17,18-Octaethylporphyrin (21) from Ethyl 3,4-Diethyl-1H-pyrrole-2-carboxylate<br />

(62); Typical Procedure: [105]<br />

Ethyl 3,4-diethyl-1H-pyrrole-2-carboxylate (62; 657 mg, 3.2 mmol) was added dropwise at<br />

0–5 8C to a stirred soln of LiAlH 4 (320 mg, 8.0 mmol) in dry THF (15 mL). The mixture was<br />

stirred for 2 h at 0–5 8C before the excess of LiAlH 4 was destroyed by addition of EtOAc. It<br />

was then poured into sat. aq NH 4Cl, extracted with EtOAc (3 ” 10 mL), washed with sat.<br />

brine, and dried (MgSO 4). The soln was concentrated to dryness in vacuo before addition<br />

of CH 2Cl 2 (15 mL). To this soln was added dimethoxymethane (methalal; 0.7 mL,<br />

9.6 mmol) and TsOH (110 mg, 0.65 mmol), and the mixture was stirred for 12 h at rt. Aerial<br />

oxidation took place under these conditions, but p-chloranil could also be used (without<br />

any improvement in yield). The mixture was washed with aq NaHCO 3 and the organic layer<br />

was dried (MgSO 4). Evaporation gave a residue which was chromatographed (silica gel,<br />

CH 2Cl 2) to give, after evaporation of the eluates and recrystallization of the residue<br />

(CH 2Cl 2/MeOH), 21; yield: 240 mg (55%).<br />

Benzyl 4-Ethyl-5-[(dimethylamino)methyl]-3-methyl-1H-pyrrole-2-carboxylate (154,<br />

R 1 = Me); Typical Procedure: [114]<br />

DMF (5 mL) was cooled an in an ice bath and then POCl 3 (5 mL) was added. The resulting<br />

soln was stirred at 0 8C for 20 min before a soln of benzyl 4-ethyl-3-methyl-1H-pyrrole-2-<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


17.8.1 Porphyrins 1123<br />

acarboxylate (3.0 g, 12.3 mmol) in CH 2Cl 2 (80 mL) was added. The light brown soln was stirred<br />

at rt under N 2 for 12 h before the solvent was removed to yield a thick brown oil. The<br />

oil was redissolved in CH 2Cl 2 (50 mL), cooled in an ice bath, and then NaBH 4 (3.32 g,<br />

87.6 mmol) in MeOH (5 mL) was added slowly. The resulting mixture was stirred at rt for<br />

5 h. The excess of NaBH 4 was quenched with AcOH, and the mixture was diluted with<br />

CH 2Cl 2 (50 mL), washed with H 2O, aq NaHCO 3, and brine, and dried (Na 2SO 4). Removal of<br />

the solvent gave 154 (R 1 = Me) as a light brown oil; yield: 2.6 g (54%).<br />

Benzyl 4-Ethyl-5-[(diethylamino)methyl]-3-methyl-1H-pyrrole-2-carboxylate (154,R 1 = Et);<br />

Typical Procedure: [114]<br />

A soln of Br 2 (6.65 g, 42 mmol) in CH 2Cl 2 (25 mL) was added dropwise to a soln of benzyl 4ethyl-3,5-dimethyl-1H-pyrrole-2-carboxylate<br />

(10.7 g, 42 mmol) and K 2CO 3 (100 mg) in dry<br />

Et 2O (150 mL). The mixture was stirred at rt for 20 min after the addition of Br 2. A soln of<br />

Et 2NH (16 mL) in Et 2O (50 mL) was then added (color change; deep red to pale yellow), and<br />

the mixture was stirred at rt for 30 min before H 2O (200 mL) was added. The organic phase<br />

was separated and washed with H 2O and then extracted with ice-cold 3.7% HCl (100 mL).<br />

The aqueous layer was washed rapidly with Et 2O before a soln of 15% NH 4OH (200 mL) was<br />

added and the resulting mixture was extracted with Et 2O. The combined organic phases<br />

were washed with H 2O, dried (Na 2SO 4), and concentrated to give 154 (R 1 = Et) as a yellow<br />

oil, which turned to an amorphous solid upon standing; yield: 11.5 g (86%).<br />

3,8,13,18-Tetraethyl-2,7,12,17-tetramethylporphyrin (Etioporphyrin I, 7); Typical<br />

Procedure: [114]<br />

Benzyl 4-ethyl-5-[(dialkylamino)methyl]-3-methyl-1H-pyrrole-2-carboxylate (154;<br />

6.6 mmol) was dissolved in THF (600 mL), and 10% Pd/C (500 mg) was added. The resulting<br />

mixture was stirred under H 2 at rt for 12 h before the catalyst was removed and the solvent<br />

was removed under reduced pressure. Recrystallization from CH 2Cl 2/hexanes afforded<br />

carboxylic acid 155 as an off-white powder in quantitative yield. Because of spontaneous<br />

decarboxylation at rt, this was used immediately: The pyrrole carboxylic acid 155<br />

(6.6 mmol) was stirred briefly with an excess of MeI (1 mL) in CH 2Cl 2 (10 mL) to give 156.<br />

Solvents were evaporated at rt and the residue was dissolved in a soln of MeOH (200 mL)<br />

and Et 3N (2 mL) and refluxed for 15 min. K 3[Fe(CN) 6] (3.8 g, 11.6 mmol) was added and the<br />

mixture was then refluxed for 10 h. The solvent was removed and the residue was redissolved<br />

in CHCl 3. Some insoluble material was filtered off and the red soln was passed<br />

through a plug of silica gel (CHCl 3). The solvent was removed under reduced pressure<br />

and the residual porphyrin was crystallized (CH 2Cl 2/MeOH) to afford etioporphyrin I (7);<br />

yield: 284 mg (36%); mp >300 8C.<br />

2,5-Bis[(dimethylamino)methyl]-3,4-diethyl-1H-pyrrole (157); Typical Procedure: [116]<br />

3,4-Diethyl-1H-pyrrole (2.22 g, 18.02 mmol) and Eschenmoser s reagent [99] (4.48 g,<br />

46.90 mmol) were dissolved in dry MeNO 2 (200 mL) and stirred at rt under N 2 for 12 h before<br />

the solvent was removed. The residue was redissolved in CH 2Cl 2 and the soln washed<br />

with sat. Na 2CO 3, dried (Na 2SO 4), and the solvent was removed under reduced pressure to<br />

afford 157 as a crude dark brown solid, pure enough (by NMR analysis) for subsequent reactions<br />

without further purification; yield: 3.67 g (86%).<br />

2,3,12,13-Tetraethyl-7,8,17,18-tetramethylporphyrin (160); Typical Procedure: [116]<br />

Pyrrole 157 (0.84 g, 3.55 mmol) and 3,4-dimethyl-1H-pyrrole (158; 0.34 g, 3.56 mmol) were<br />

added to a soln of K 3[Fe(CN) 6] (3.80 g, 11.54 mmol) in MeOH (75 mL). The mixture was refluxed<br />

for 4 h before the solvent was removed. The residue was dissolved in CH 2Cl 2 and<br />

washed with H 2O, 5% NH 4OH (” 2), H 2O, and brine, and then dried (Na 2SO 4). The solvent<br />

was removed under reduced pressure and the crude product was chromatographed (silica<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1124 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

agel, 10% petroleum ether/CH 2Cl 2). Recrystallization (CH 2Cl 2/cyclohexane) afforded the<br />

porphyrin 160; yield: 162 mg (19.5%); mp >3008C.<br />

17.8.1.2.2 Method 2:<br />

From Dipyrrolic Intermediates: The [2+2] Routes<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, p 585.<br />

The dipyrrolic intermediates most often used for porphyrin synthesis are dipyrromethenes<br />

and dipyrromethanes. Because dipyrromethanes are sensitive to cleavage by<br />

acidic reagents if they are not substituted with electron-withdrawing groups, most of<br />

the Fischer-era developments in porphyrin synthesis depended upon dipyrromethenes<br />

as the intermediates. In those times, Fischer did not always concern himself with avoidance<br />

of mixtures due to ambiguities related to symmetry limitations; there were occasions<br />

when access to a particular porphyrin was so important that preparation of a mixture<br />

followed by chromatographic separation was perfectly acceptable. However, this still<br />

required an extra layer of tactical planning; for example, if a hypothetical [2 +2] dipyrromethene<br />

condensation reaction was to be used (Scheme 37), the two halves would be<br />

chosen so that one half, 162, possessed polar groups while the other, 161, did not. If the<br />

required product was mesoporphyrin IX, then this molecule (163, bearing two carboxylate<br />

groups) could be readily separated from the two other porphyrin products, etioporphyrin<br />

I (7, bearing no polar carboxylate groups) and coproporphyrin II (164, bearing<br />

four carboxylates).<br />

Scheme 37 Strategy for Porphyrin Synthesis Using Side-Chain Polarity of Dipyrromethenes<br />

Br<br />

Et<br />

Et<br />

NH HN<br />

+<br />

161<br />

NH N<br />

N HN<br />

163<br />

Et<br />

Br −<br />

Et<br />

CO2H CO2H<br />

+<br />

+<br />

Br −<br />

Et<br />

Br Br<br />

+<br />

NH HN<br />

CO 2H<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

7<br />

162<br />

Et<br />

Et<br />

CO 2H<br />

+<br />

heat<br />

CO 2H CO2H<br />

CO2H<br />

NH N<br />

N HN<br />

Obvious symmetry problems will be associated with all [2 +2] approaches to porphyrins<br />

whether dipyrromethenes or dipyrromethanes are used. In the absence of methodology<br />

to prevent equal reactivity at both ends of each dipyrrole component, mixtures will result<br />

if two different dipyrroles are used. For example, condensation of dipyrromethenes 165<br />

and 166, presumably intended to give 167 will also result in the production of porphyrin<br />

168; there is no simple way that the second porphyrin can be avoided (Scheme 38). One<br />

must therefore carefully strategize symmetry considerations if a single pure porphyrin is<br />

164<br />

CO2H


17.8.1 Porphyrins 1125<br />

arequired, and a mixture is to be avoided. This is not difficult to do; in simple terms, as long<br />

as one of the two dipyrroles is symmetrically substituted about its 5-carbon, a mixture<br />

will not result. This analysis applies to Scheme 7, mode E. Thus, porphyrins (e.g., 170)<br />

which possess symmetry in one or both precursor halves (165, 169) of the molecule can<br />

be targeted through this approach. As it happens this is not a significant hardship because<br />

many natural porphyrins, for example protoporphyrin IX (2), coproporphyrin III (3), and<br />

uroporphyrin III (171) do have “lower-hemisphere” symmetry either side of meso-carbon-<br />

15. These complex natural molecules can therefore be synthesized by the [2 +2] approach<br />

provided that the strategy involves condensation of an A–B dipyrrole unit with a C–D dipyrrole<br />

in which either the A–B or C–D hemisphere is symmetrical about its interpyrrolic<br />

carbon atom. Alternatively, as in the case shown in Scheme 7, mode F, the route can be<br />

limited to the synthesis of porphyrins (e.g., 173) which are centrosymmetrically substituted<br />

(i.e., produced by self-condensation of a dipyrrolic compound 172).<br />

Scheme 38 Strategy for Porphyrin Synthesis Using Side-Chain Symmetry<br />

A<br />

A<br />

B C<br />

NH HN<br />

+<br />

D<br />

X<br />

Br Br<br />

−<br />

165<br />

+<br />

X<br />

−<br />

H<br />

+<br />

NH HN<br />

G<br />

A<br />

H<br />

B C<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

166<br />

G<br />

F<br />

167<br />

B<br />

NH<br />

C<br />

HN<br />

+<br />

D<br />

X<br />

+<br />

X<br />

−<br />

E<br />

+<br />

NH HN<br />

Br Br<br />

F<br />

F<br />

−<br />

A B D C<br />

165 169<br />

E<br />

F<br />

E<br />

D<br />

E<br />

heat<br />

+<br />

heat<br />

A<br />

E<br />

A<br />

E<br />

B C<br />

F<br />

NH N<br />

N HN<br />

168<br />

A B<br />

NH N<br />

N HN<br />

D C<br />

G<br />

B C<br />

F<br />

170<br />

F<br />

D<br />

H<br />

D<br />

E<br />

for references see p 1223


1126 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

HO 2C<br />

aNH N<br />

HO2C<br />

2<br />

HO 2C<br />

HO 2C<br />

A<br />

Br<br />

N HN<br />

171<br />

B C<br />

NH HN<br />

+<br />

172<br />

X −<br />

CO2H<br />

CO2H D<br />

CO2H<br />

CO2H<br />

heat<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

A<br />

D<br />

B C<br />

Application of mode D shown in Scheme 7 almost invariably involves the use of dipyrromethanes.<br />

Dipyrromethenes are simply not nucleophilic enough (particularly in the presence<br />

of acid, which produces the non-nucleophilic cationic salts) to react with the necessary<br />

one-carbon linking units (formaldehyde or orthoformate equivalents, or benzaldehyde).<br />

17.8.1.2.2.1 Variation 1:<br />

Using Dipyrromethenes<br />

Most porphyrin syntheses from dipyrromethenes were developed by Hans Fischer s<br />

group in Munich. The case illustrated in Scheme 7 mode F is achieved by self-condensation<br />

of 1-bromo-9-methyldipyrromethenes [e.g., 174 R 1 = Et, (CH 2) 2CO 2H; R 2 = Me] in boiling<br />

formic acid or in organic acid melts (succinic, tartaric, etc.) at temperatures up to<br />

and exceeding 200 8C to give, for example, etioporphyrin I (176, R 1 = Et) or coproporphyrin<br />

I (176)[R 1 = (CH 2) 2CO 2H, usually isolated as the tetramethyl ester 178] (Scheme 39). [117–<br />

119] Fischer s collaborators determined which organic acid to use by trial and error, the<br />

choice simply relating to the temperature at which the best yield of porphyrin was obtained.<br />

It is also possible to heat 1-bromo-9-(bromomethyl)dipyrromethene hydrobromides<br />

[e.g., 174 (R 2 =CH 2Br)] [120] or 1-bromo-9-methyldipyrromethene perbromides, (e.g.,<br />

175) [30] or even a mixture of both [121,122] in formic acid to give very good yields of centrosymmetrically<br />

substituted porphyrins, such as etioporphyrin I (176, R 1 = Et), coproporphyrin<br />

I tetramethyl ester (178) [via acid 177], 2,3,7,8,12,13,17,18-octaethylporphyrin<br />

(21), or more recently the opp-dibenzoporphyrin 180 (from the dipyrromethene 179)<br />

(Scheme 39). [123] (See also Scheme 7, case F.)<br />

C<br />

173<br />

B<br />

D<br />

A


17.8.1 Porphyrins 1127<br />

aScheme 39 Synthesis of Porphyrins from Dipyrromethenes [30,117–122]<br />

R 1<br />

R<br />

NH HN<br />

1<br />

Br R2 +<br />

Br −<br />

174<br />

R 1 = Et, (CH 2) 2CO 2H; R 2 = Me, CH 2Br<br />

HO 2C<br />

2<br />

Br<br />

Cl<br />

NH HN<br />

+<br />

Br −<br />

177<br />

NH HN<br />

+<br />

Br −<br />

179<br />

Et<br />

and/or<br />

CO2H<br />

NH HN<br />

+<br />

Br<br />

−<br />

3<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Br<br />

R 1<br />

1,2-Cl2C6H4<br />

heat<br />

175<br />

heat<br />

1. Br2, HCO2H heat<br />

2. MeOH, H2SO4 19%<br />

MeO 2C<br />

MeO 2C<br />

Et<br />

R 1<br />

R 1<br />

R 1<br />

NH N<br />

N HN<br />

178<br />

NH N<br />

N HN<br />

180<br />

NH N<br />

N HN<br />

Et<br />

176<br />

CO2Me<br />

R 1<br />

CO 2Me<br />

These kinds of dipyrromethene approach can be adapted to the synthesis of fairly complex<br />

porphyrins (e.g., natural type-III systems) by heating of 1,9-dibromodipyrromethenes<br />

182 with 1,9-dimethyl- or 1,9-bis(bromomethyl)dipyrromethenes 181 in organic acid<br />

melts; this is an example of the mode shown in Scheme 7, case E. A notable example of<br />

this approach is Fischer s synthesis of deuteroporphyrin IX (183), an intermediate in the<br />

Munich total synthesis of hemin (184) (Scheme 40). [124] The formation of a mixture of por-<br />

R 1<br />

for references see p 1223


1128 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aphyrins is avoided by selection of one dipyrromethene (i.e., 182) which is symmetrically<br />

substituted about its 5-(meso)-carbon atom.<br />

Scheme 40 Hemin Synthesis from Dipyrromethenes [124]<br />

NH HN<br />

R1 R1 Br<br />

+<br />

−<br />

+<br />

Br Br<br />

+<br />

NH HN<br />

181 R1 CO2H CO2H = Me, CH2Br 182<br />

CO 2H<br />

NH N<br />

N HN<br />

183<br />

N Cl N<br />

N N<br />

3,8,13,18-Tetraethyl-2,7,12,17-tetramethylporphyrin (Etioporphyrin I, 176,R 1 = Et); Typical<br />

Procedure: [120]<br />

Brominated kryptodipyrromethene hydrobromide (36; 10 g, 24.86 mmol) in HCO 2H<br />

(20 mL) was heated on a boiling water bath with regular shaking for 4 h. After cooling<br />

the soln was diluted with H 2O and treated with aq NaOH until alkaline. The precipitate<br />

was subjected to Soxhlet extraction, initially using MeOH, until the solvents were weakly<br />

brown. Then the extraction was performed with CHCl 3 (ca. 20 mL) until the percolating<br />

solvent was colorless. The CHCl 3 soln was separated, treated with MeOH, and the crystals<br />

were collected by filtration. Recrystallization (CHCl 3/MeOH) gave 176 (R 1 = Et); yield: 2 g;<br />

mp >300 8C.<br />

3,8,13,18-Tetrakis[2-(methoxycarbonyl)ethyl]-2,7,12,17-tetramethylporphyrin (Coproporphyrin<br />

I Tetramethyl Ester, 178); Typical Procedure: [30]<br />

1-Bromo-3,8-bis(2-carboxyethyl)-2,7,9-trimethyldipyrromethene hydrobromide (177; 1g,<br />

2.04 mmol) was suspended in HCO 2H (10 mL) and refluxed for 2 h before the solvent was<br />

removed by distillation at atmospheric pressure. The residue was dissolved in H 2SO 4/<br />

MeOH [5% (v/v), 100 mL] and left at rt in the dark for 12 h. The soln was poured into H 2O<br />

and extracted with CH 2Cl 2; this soln was then washed with H 2O, dried (MgSO 4), and concentrated<br />

to dryness. The purple residue was chromatographed [alumina (Brockmann<br />

Grade III), CH 2Cl 2], the red eluates were concentrated and the residue was crystallized<br />

(CH 2Cl 2/MeOH) to give the porphyrin 178; yield: 140 mg (19%); mp 255–2568C.<br />

In an improvement designed to simulate use of the perbromide 175 [R 1 =(CH 2)CO 2H]<br />

instead of the hydrobromide 174, compound 174 [R 1 =(CH 2)CO 2H, R 2 =Me; 2g,<br />

4.07 mmol] in HCO 2H (20 mL) was treated with Br 2 (0.65 g, 4.07 mmol) and refluxed for<br />

2 h. After a workup as above, coproporphyrin I tetramethyl ester (178) was isolated; yield:<br />

733 mg (50%).<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

CO 2H<br />

Br −<br />

CO 2H<br />

Fe<br />

184<br />

CO2H


17.8.1 Porphyrins 1129<br />

a17.8.1.2.2.2 Variation 2:<br />

Using Dipyrromethanes<br />

Fischer s almost exclusive use of dipyrromethenes overshadowed the development of<br />

porphyrin syntheses using other dipyrroles; even in those days it was known that dipyrromethanes<br />

were so unstable toward acidic reagents (and most certainly to those used by<br />

Fischer) that they might not be useful as porphyrin precursors. This is indeed true for<br />

one of Fischer s procedures, [125] for which it has been shown that the self-condensation<br />

of dipyrromethane-1,9-dicarboxylic acids (e.g., 185) in formic acid gives a mixture of<br />

porphyrin type-isomers rather than pure etioporphyrin II (186,R 1 = Et) or coproporphyrin<br />

II [186,R 1 =(CH 2) 2CO 2H] (Scheme 41). [126,127]<br />

Scheme 41 Synthesis of Porphyrins from Dipyrromethanes [25,48,126–128]<br />

R 1 R 1<br />

NH HN<br />

HO 2C CO 2H<br />

185<br />

R 1 = Et, (CH 2) 2CO 2H<br />

MeO 2C<br />

MeO2C<br />

HCO 2H, heat<br />

R 1 R 1<br />

NH N<br />

N HN<br />

OHC CHO<br />

NH HN<br />

+<br />

NH HN<br />

1 R R1<br />

187 R1 CO2Me CO2Me MeO2C CO2Me<br />

MeO2C CO2Me<br />

= H, CO2H 188<br />

HI, [O]<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 1<br />

MeO 2C<br />

MeO 2C<br />

MeO 2C<br />

186<br />

R 1<br />

NH N<br />

N HN<br />

MeO2C CO2Me<br />

189<br />

CO2Me<br />

CO2Me CO2Me<br />

for references see p 1223


1130 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

MeO 2C<br />

OHC CHO<br />

aNH HN<br />

CO2Me<br />

+<br />

NH HN<br />

1 R R2<br />

190 R 1 = H, CO 2H, CO 2Bn; R 2 = H, CO 2-t-Bu<br />

CO 2Me<br />

NH HN<br />

R1 CHO<br />

193 R 1 = H, CO 2H<br />

CO2Me<br />

TsOH, [O]<br />

MeO 2C CO 2Me<br />

MeO2C<br />

MeO 2C<br />

MeO 2C<br />

MeO 2C<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

191<br />

NH N<br />

N HN<br />

192 41%<br />

NH N<br />

N HN<br />

178<br />

CO 2Me<br />

CO2Me<br />

CO2Me<br />

CO 2Me<br />

A critically important development in the art and science of porphyrin synthesis occurred<br />

with the discovery by MacDonald and his group [25] that a 1,9-diformyldipyrromethane<br />

(e.g., 188) can be treated with a 1,9-di-unsubstituted dipyrromethane 187 (R 1 = H, or its<br />

chemically equivalent 1,9-dicarboxylic acid R 1 =CO 2H) in the presence of an acid catalyst<br />

to afford a pure porphyrin [e.g., uroporphyrin III octamethyl ester (189)] in yields often as<br />

high as 60% (Scheme 41). MacDonald s key publications employed hydriodic acid (which<br />

is not easy to handle or to purify), and the high yields achieved by his group appear to be<br />

limited to sterically encumbered porphyrins such as those of the uroporphyrin series. Recently,<br />

4-toluenesulfonic acid has been shown [48,128] to be a much more convenient alternative<br />

to hydriodic acid as the important acid catalyst; in this way, for example, coproporphyrin<br />

III tetramethyl ester (192) is obtained from the dipyrromethanes 190 (R 1 =CO 2Bn,<br />

R 2 =CO 2t-Bu, via R 1 =CO 2H and R 1 ,R 2 = H) and 191.<br />

The MacDonald [2 +2] route is probably the most widely used pathway to synthetic<br />

porphyrins, even though it suffers the same symmetry restrictions apparent in Fischer s<br />

syntheses from dipyrromethenes. Thus, one of the two dipyrromethanes (e.g., compounds<br />

188 and 191 in Scheme 41) must be symmetrical about its interpyrrolic 5-carbon<br />

atom; this is an example of mode E illustrated in Scheme 7. The approach can also be<br />

modified (Scheme 7, mode F) to include the preparation of centrosymmetrically substitut-


17.8.1 Porphyrins 1131<br />

aed porphyrins 178 by self-condensation of a 9-unsubstituted 1-formyldipyrromethane<br />

193 (Scheme 41).<br />

5,10,15,20-Tetraarylporphyrins, such as 5,10,15,20-tetraphenylporphyrin (22), are<br />

prepared by reaction of an arylaldehyde with pyrrole, but if a tetraarylporphyrin is required<br />

that has different aryl groups at the each of the meso positions some major synthetic<br />

problems need to be overcome. Such porphyrins are usually synthesized by the condensation<br />

of pyrrole with a mixture of arylaldehydes, [130,131] but are obtained in poor yields after<br />

lengthy chromatographic separation and purification. If pyrrole were to be condensed<br />

with a mixture of four arylaldehydes there is only a minor possibility that a porphyrin<br />

with one each of the aryl groups in it would be obtained, and even then one would not<br />

know the spatial relationship of each of the aryl groups to the others (i.e., which is at position<br />

5, 10, 15, 20, etc.). However, a synthetic route has been developed which, for example,<br />

can yield the totally unsymmetrical porphyrin 195. [132] It is based on the MacDonald [2+2]<br />

route, using 5-aryl-1,9-diaroyldipyrromethanes 194 rather than 1,9-diformyldipyrromethanes.<br />

Scheme 42 shows such a route to porphyrin 195. This is an example of mode E<br />

outlined in Scheme 7. A very similar approach was subsequently reported by Lindsey<br />

and co-workers. [133]<br />

Scheme 42 Unsymmetrical Tetraarylporphyrins by the MacDonald Approach [132]<br />

MeO<br />

MeO<br />

O<br />

O<br />

N<br />

H<br />

N<br />

H<br />

LiAlH4<br />

POCl3<br />

DMF<br />

4-TolMgBr<br />

44%<br />

O<br />

N<br />

H<br />

CHO<br />

71%<br />

OH<br />

4-Tol<br />

MeO<br />

N<br />

H O<br />

K-10 clay<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

45%<br />

MeO<br />

Ph<br />

MeO<br />

4-Tol<br />

NH HN<br />

OH HO<br />

MeO<br />

Ph<br />

4-Tol<br />

NH HN<br />

O O<br />

194<br />

4-Tol<br />

F<br />

195<br />

NH HN<br />

F<br />

EtCO2H, heat<br />

12%<br />

Ph<br />

Ph<br />

for references see p 1223


1132 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aAround the same time (1960) that MacDonald published his general [2 +2] approach using<br />

dipyrromethanes, R. B. Woodward published his synthesis of chlorophyll a. [44,134,135]<br />

Woodward also used a bisdipyrromethane [2 +2] mode, but characteristically, designed<br />

his approach so that the normal requirement that there be symmetry about one of the<br />

dipyrromethane 5-carbon atoms was circumvented; though mode E, Scheme 7, was employed,<br />

enhanced reactivity of one of the terminal positions of the 1,9-di(“carbonyl”)dipyrromethane<br />

196 (Scheme 43) allowed exquisitely regioselective condensation with 56,<br />

by way of the tethered tetrapyrrole (see arrow in 197) to give only one b-bilene 198 and<br />

finally only one porphyrin 199. The transient involvement of the b-bilene prompted later<br />

researchers to investigate approaches to unsymmetrical porphyrin syntheses via openchain<br />

tetrapyrrole precursors such as 198 and particularly of biladienes.<br />

Scheme 43 Woodward s Version of the [2 +2] Dipyrromethane Approach [44,134,135]<br />

CO2Me<br />

NH 2<br />

NH<br />

NH<br />

+<br />

S<br />

O<br />

H<br />

HN<br />

HN<br />

CO 2Me<br />

56 196<br />

+ NH3 Et<br />

CO2Me<br />

MeO 2C<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

AcOH<br />

CO2Me<br />

Et<br />

HCl, MeOH<br />

NH HN<br />

+<br />

I2/Ac2O 48%<br />

NH HN<br />

CO 2Me<br />

O<br />

198<br />

CO 2Me<br />

CO2Me<br />

2Cl −<br />

N<br />

NH N<br />

NH HN<br />

O<br />

197<br />

CO2Me<br />

NH 2<br />

Et<br />

CO 2Me<br />

CO2Me<br />

199<br />

Et<br />

CO2Me<br />

The MacDonald procedure, through dipyrromethanes, fully mirrors the Fischer routes<br />

through dipyrromethenes, each method having its relative advantages and disadvantages.<br />

An obvious disadvantage of the Fischer route is the vigorous thermal and acidic<br />

conditions required; these often cause destruction of the sensitive side chains which are<br />

available and essential in modern porphyrin syntheses. The advantage of the MacDonald<br />

approach is that dipyrromethanes with complex substituents tend to be more readily prepared<br />

than the corresponding dipyrromethene analogues, and once synthesized these intermediates<br />

can be transformed into porphyrins using very mild reaction conditions.<br />

Thus, in recent times, the MacDonald approach has even been modified to afford a very<br />

useful [3+1] analogue (see Section 17.8.1.2.3). [116,136–143]<br />

Scheme 7, mode D illustrates a procedure in which two 1,9-di-unsubstituted dipyrroles<br />

are joined together by use of two identical one-carbon (formaldehyde, ortho-


17.8.1 Porphyrins 1133<br />

aformate, or arylaldehyde) units. This approach only works well for dipyrromethanes due<br />

to the non-nucleophilicity of the corresponding dipyrromethene analogues. This mode of<br />

synthesis is demonstrated in the synthesis of coproporphyrin II tetramethyl ester (201)by<br />

treatment of the dipyrromethane 200 in dichloromethane with trimethyl orthoformate<br />

in the presence of trichloroacetic acid catalyst (Scheme 44). [68] A 24% yield of coproporphyrin<br />

II tetramethyl ester is obtained in this way.<br />

Methodology has also been developed which allows the synthesis of 5,15-diaryl- or<br />

5,15-dialkylporphyrins from a dipyrromethane and a one-carbon linker unit (formaldehyde,<br />

orthoformate equivalent, or alkyl/aryl aldehyde). [52] Conditions are similar to those<br />

employed in the MacDonald route. Scheme 44 shows typical examples, e.g. porphyrin 203<br />

produced from the 5-mesityldipyrromethane 202 and 4-iodobenzaldehyde.<br />

Scheme 44 Synthesis of Porphyrins from Dipyrromethanes [52,68]<br />

MeO 2C<br />

NH HN<br />

HO2C CO 2H<br />

Mes<br />

200<br />

NH HN<br />

202<br />

CO2Me<br />

CHO<br />

I<br />

BF3 OEt2, DDQ, CHCl3 32%<br />

Cl3CCO2H, HC(OMe)3, CH2Cl2, O2<br />

24%<br />

MeO2C<br />

MeO 2C<br />

NH N<br />

N HN<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Mes<br />

Mes<br />

203<br />

201<br />

CO 2Me<br />

CO2Me<br />

I I<br />

for references see p 1223


1134 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

Mes<br />

NH HN<br />

a202<br />

CHO<br />

TMS<br />

TMS<br />

, H +<br />

NH N<br />

N HN<br />

A mixed dipyrromethane/dipyrromethene synthesis has also been reported by MacDonald<br />

s group (Scheme 45). Condensation of a 1,9-bis(bromomethyl)dipyrromethene 205<br />

with a 1,9-di-unsubstituted dipyrromethane 204 affords a good yield of the isomerically<br />

pure porphyrin 206. [25]<br />

Scheme 45 Synthesis of Porphyrin from Dipyrromethane and Dipyrromethene [25]<br />

HO2C HO2C NH HN<br />

CO2H<br />

CO2H<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

+<br />

Mes<br />

Mes<br />

Br Br<br />

+<br />

NH HN<br />

204 205<br />

HO2C HO2C NH N<br />

N HN<br />

206<br />

Br −<br />

TMS<br />

HO2C CO 2H<br />

CO2H<br />

CO2H<br />

HO2C CO 2H<br />

3,8,13,17-Tetrakis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin (Coproporphyrin<br />

III Tetramethyl Ester, 192); Typical Procedure: [129]<br />

Benzyl 9-(tert-butoxycarbonyl)-3,8-bis[2-(methoxycarbonyl)ethyl]-2,7-dimethyldipyrromethane-1-carboxylate<br />

(190, R 1 =CO 2Bn; R 2 =CO 2t-Bu; 456 mg, 0.79 mmol) in THF<br />

(100 mL) containing Et 3N (2 drops) and 10% Pd/C (46 mg) was hydrogenated at rt and atmospheric<br />

pressure until uptake of H 2 ceased. The catalyst was filtered off through Celite<br />

and the filtrate was concentrated to dryness to give 190 (R 1 =CO 2H; R 2 =CO 2t-Bu); the residue<br />

was dissolved in TFA (25 mL) and kept under N 2 for 45 min before concentration under<br />

reduced pressure. CH 2Cl 2 and H 2O were added and the organic phase was washed with<br />

aq NaHCO 3 and then H 2O to give 190 (R 1 =R 2 = H) before being dried (Na 2SO 4) and made up<br />

to a total volume of 150 mL with CH 2Cl 2. This soln was added to a darkened flask (alumi-


17.8.1 Porphyrins 1135<br />

anum foil) containing 1,9-diformyl-3,7-bis[2-(methoxycarbonyl)ethyl]-2,8-dimethyldipyrromethane<br />

(191; 250 mg, 0.62 mmol) in CH 2Cl 2 (100 mL), and then treated with a soln of<br />

TsOH (900 mg, 4.7 mmol) in MeOH (12.5 mL). After stirring for 6 h in the dark, the soln<br />

was treated with a sat. soln of Zn(OAc) 2 in MeOH (12.5 mL) and set aside overnight. The<br />

mixture was washed with H 2O, aq NaHCO 3,H 2O, and then dried (Na 2SO 4). After concentration<br />

to dryness, the residue was dissolved in H 2SO 4/MeOH (5%, 20 mL) and set aside overnight<br />

at rt in the dark. The soln was poured into CH 2Cl 2/H 2O, and the organic phase was<br />

collected and washed with H 2O, aq NaHCO 3,H 2O, and then dried (Na 2SO 4). Concentration<br />

gave a red residue, which was chromatographed [neutral alumina (Brockmann Grade III),<br />

CH 2Cl 2]. The red eluates were concentrated and recrystallization of the residue (CH 2Cl 2/<br />

MeOH) gave 192; yield: 180 mg (41%); mp 150–1538C, remelting at 179–1828C.<br />

3,7,13,17-Tetrakis[2-(methoxycarbonyl)ethyl]-2,8,12,18-tetramethylporphyrin<br />

(Coproporphyrin II Tetramethyl Ester, 201); Typical Procedure: [68]<br />

3,7-Bis[2-(methoxycarbonyl)ethyl]-2,8-dimethyldipyrromethane-1,9-dicarboxylic acid<br />

(200; 550 mg, 1.27 mmol) [obtained by catalytic hydrogenation (in THF, over 5% Pd/C) of<br />

dibenzyl 3,7-bis[2-(methoxycarbonyl)ethyl]-2,8-dimethyldipyrromethane-1,9-dicarboxylate]<br />

was treated with 1 M trichloroacetic acid in CH 2Cl 2 (72 mL) and an excess of<br />

HC(OMe) 3 (2.5 g, 23.6 mmol) in CH 2Cl 2 (408 mL). The soln was stirred in the presence of<br />

O 2 overnight at rt in the dark, monitored by observation of the Soret absorption band<br />

(410 nm, dication salt). The soln was washed with 10% aq Na 2SO 4 (200 mL), H 2O<br />

(2 ” 200 mL), dried (MgSO 4), and concentrated to dryness. The residue was chromatographed<br />

[neutral alumina (Brockmann Grade III), CH 2Cl 2] and the red eluates were collected.<br />

Concentration and crystallization (CH 2Cl 2/MeOH) gave 201; yield: 104 mg (24%); mp 285–<br />

2888C.<br />

5,15-Bis(4-iodophenyl)-10,20-di(mesityl)porphyrin (203); Typical Procedure: [52]<br />

A soln of 4-iodobenzaldehyde (58 mg, 0.25 mmol) and 5-mesityldipyrromethane (202;<br />

66 mg, 0.25 mmol) in CHCl 3 (25 mL) was purged with argon for 10 min before addition of<br />

2.5 M BF 3•OEt 2 in CHCl 3 (33 mL, 3.3 mM). The soln was stirred for 1 h at rt and then DDQ<br />

(43 mg, 0.19 mmol) was added. The mixture was stirred at rt for 1 h and the solvent was<br />

removed. Column chromatography (silica gel, CH 2Cl 2) afforded the porphyrin 203 as the<br />

first moving band; yield: 38 mg (32%); mp >300 8C. Comparable yields were obtained using<br />

TFA (0.01–0.05 M) as the catalyst.<br />

17.8.1.2.2.3 Variation 3:<br />

Using Dipyrroketones<br />

The carbonyl group in dipyrroketones is a bis-vinylogous amide, and therefore does not<br />

chemically behave like a regular ketone; reagents such as diborane are required to accomplish<br />

the reduction to methylene. It cannot, therefore, be reduced simply by the addition<br />

of sodium borohydride to the dipyrroketone. Porphyrins per se cannot be synthesized directly<br />

from dipyrroketones using a [2 +2] protocol (e.g., Scheme 7, modes D, E, and F) because<br />

the 5-carbonyl function in the dipyrroketone will remain in the final cyclization<br />

product to yield species known as oxophlorins. However, oxophlorins (e.g., 209) which<br />

can be readily converted into the corresponding porphyrins, can be prepared in good<br />

yield [144] by a MacDonald-type condensation of 1,9-diformyldipyrroketones 207 with 1,9di-unsubstituted<br />

dipyrromethanes 208 (R 1 = H) or the corresponding 1,9-dicarboxylic<br />

acids 208 (R 1 =CO 2H) (Scheme 46). 1-Formyl-9-(hydroxymethyl)dipyrroketones (e.g., 210)<br />

can be used in place of 207. [145] The diformyldipyrroketones 207 are best obtained by direct<br />

oxidation of readily available 1,9-diformyldipyrromethanes 211 (Scheme 39; see also<br />

Section 17.8.1.1.1.3). [146] However, the formyl groups, which eventually form the bridging<br />

carbon atoms between the two dipyrrolic halves, must be placed on the dipyrroketone<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1136 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

amoiety because 1,9-di-unsubstituted dipyrroketones (e.g., 212) are not nucleophilic<br />

enough to react with 1,9-diformyldipyrromethanes (e.g., 211).<br />

As might be anticipated, the same symmetry considerations that restrict the Fischer<br />

syntheses from dipyrromethenes, and the standard MacDonald syntheses from dipyrromethanes,<br />

are also applicable to [2+2]-type syntheses of 5-oxophlorins.<br />

Scheme 46 Synthesis of Oxophlorin from Dipyrroketones [144,145]<br />

MeO 2C<br />

A<br />

OHC<br />

O<br />

NH HN<br />

OHC CHO<br />

207<br />

B O C<br />

NH HN<br />

210<br />

D<br />

OH<br />

CO 2Me<br />

A<br />

+<br />

MeO2C<br />

B C<br />

NH HN<br />

R 1 R 1<br />

NH HN<br />

208 R 1 = H, CO 2H<br />

MeO2C<br />

MeO2C<br />

CO 2Me<br />

NH HN<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

D<br />

OHC CHO<br />

211<br />

A<br />

O<br />

209<br />

1. TFA<br />

2. NH3, MeOH<br />

B O C<br />

NH HN<br />

212<br />

CO 2Me<br />

CO2Me<br />

Transformation of oxophlorins 213 into porphyrins 215 is possible using several different<br />

routes (Scheme 47). [147] The oxophlorin oxo-group can be removed by: (i) sodium amalgam<br />

reduction followed by re-oxidation of an intermediate porphyrinogen 214, usually<br />

with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone, to porphyrin 215; or (ii) catalytic hydrogenation<br />

of the oxophlorin to give the macrocyclic tetrapyrroketone 217, followed by reduction<br />

with diborane to give the porphyrinogen 214 and re-oxidation to porphyrin 215;<br />

or (iii) treatment of the oxophlorin with acetic anhydride in pyridine to give the enol-acetate<br />

5-acetoxyporphyrin 216, followed by catalytic hydrogenation to the porphyrinogen<br />

214 (or possibly intermediate 218) and re-oxidation with 2,3-dichloro-5,6-dicyanobenzo-<br />

1,4-quinone to give the meso-unsubstituted porphyrin, mesoporphyrin IX dimethyl ester<br />

(215). Method (iii), via the readily isolated 5-acetoxyporphyrin 216, is the method of<br />

choice, and yields of porphyrin 215 from 216 are often as high as 85%. The yield of 5-acetoxyporphyrin<br />

216 from oxophlorin 213 is usually excellent (see Section 17.8.1.2.4.2).<br />

D


17.8.1 Porphyrins 1137<br />

aScheme 47 Porphyrin Syntheses from Oxophlorins [147]<br />

Et<br />

CO 2Me<br />

MeO2C<br />

NH HN<br />

N HN<br />

Ac2O py<br />

213<br />

MeO2C<br />

Et<br />

Et<br />

NH HN<br />

NH HN<br />

217<br />

Et<br />

O<br />

CO 2Me<br />

Na/Hg<br />

NH N<br />

N HN<br />

216<br />

Et<br />

Et<br />

CO2Me<br />

Et<br />

OAc<br />

CO 2Me<br />

NH HN<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

214<br />

H 2<br />

Pd/C<br />

MeO2C<br />

Et<br />

Et<br />

DDQ<br />

CO 2Me<br />

NH HN<br />

NH HN<br />

218<br />

Et<br />

CO 2Me<br />

O OAc<br />

CO2Me<br />

Et<br />

NH N<br />

N HN<br />

CO2Me<br />

215<br />

Et<br />

CO 2Me<br />

3,7,13,17-Tetrakis[2-(methoxycarbonyl)ethyl]-2,8,12,18-tetramethyl-5-oxophlorin (209);<br />

Typical Procedure: [144]<br />

1,9-Diformyl-3,7-bis[2-(methoxycarbonyl)ethyl]-2,8-dimethyl-5-dipyrroketone (207;<br />

830 mg, 1.99 mmol) was added in portions over 15 min to a stirred soln of 3,7-bis[2-(methoxycarbonyl)ethyl]-2,8-dimethyldipyrromethane-1,9-dicarboxylic<br />

acid (208, R 1 =CO 2H;<br />

900 mg, 2.07 mmol) in TFA (8.0 mL). After 2 h, the violet soln was poured into cold MeOH<br />

(40 mL) and the stirred mixture was kept at 108C while 3 M aq NH 3 was added dropwise<br />

until the pH was slightly alkaline. The green precipitate was collected, washed with aq<br />

MeOH (90%), and recrystallized (CHCl 3/MeOH) to give the oxophlorin 209; yield: 1.2 g<br />

(83%); mp 232–2348C.<br />

3,8-Diethyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin (Mesoporphyrin<br />

IX Dimethyl Ester, 215); Typical Procedure: [69]<br />

5-Acetoxy-3,18-diethyl-8,12-bis[2-(methoxycarbonyl)ethyl]-2,7,13,17-tetramethylporphyrin<br />

(216; 50 mg, 0.078 mmol) in THF (40 mL) containing Et 3N (1 drop) was hydrogenated<br />

over 10% Pd/C (50 mg) until uptake of H 2 ceased (7 h). The catalyst was filtered off and<br />

the solvent was removed under vacuum to give a yellow gum, which was dissolved in<br />

for references see p 1223


1138 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aCH 2Cl 2 (300 mL) and oxidized by addition of portions (12 ” 100 mL) of 0.005% I 2 in 3% aq<br />

NaOAc. [Note: addition of an excess of DDQ in toluene is more convenient and gives a<br />

higher yield]. The mixture was shaken after each addition and the final organic layer<br />

was separated, washed with H 2O, and dried (MgSO 4). After concentration to dryness, the<br />

residual red gum was chromatographed [neutral alumina (Brockmann Grade III), eluting<br />

with CH 2Cl 2] to give, after recrystallization (CH 2Cl 2/MeOH), mesoporphyrin IX dimethyl<br />

ester (215); yield: 29 mg (64%); mp 216–2178C.<br />

5-Acetoxy-3,18-diethyl-8,12-bis[2-(methoxycarbonyl)ethyl]-2,7,13,17-tetramethylporphyrin<br />

(216); Typical Procedure: [69]<br />

Crude 3,18-diethyl-8,12-bis[2-(methoxycarbonyl)ethyl]-5-oxo-2,7,13,17-tetramethylphlorin<br />

[213; obtained from cyclization of 3,8-diethyl-13,17-bis[2-(methoxycarbonyl)ethyl]-<br />

2,7,12,18-tetramethyl-b-oxobilane-1,19-dicarboxylic acid (110; 345 mg, 0,45 mmol)] was<br />

taken up in pyridine (30 mL) and Ac 2O (8 mL). After stirring at rt for 10 min, the deep red<br />

soln was concentrated under reduced pressure to give an oil, which was chromatographed<br />

twice [neutral alumina (Brockmann Grade III), CH 2Cl 2]. The red eluates were concentrated<br />

to give a residue, which was crystallized from CH 2Cl 2/MeOH to give the 5-acetoxyporphyrin<br />

216; yield: 245 mg (70%); mp 236–2378C.<br />

17.8.1.2.3 Method 3:<br />

From Tripyrrolic Intermediates: The [3+1] Route<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, p 588.<br />

The protocol involving cyclization of an open-chain tripyrrole with a 2,5-difunctionalized<br />

monopyrrole to produce a porphyrin (Scheme 48) utilizes the mode of attachment<br />

G shown in Scheme 7, and has a number of useful applications. Porphyrin models are being<br />

increasingly used as mimics for natural systems and for commercial applications, the<br />

objective of a particular approach often being the attachment of some biological moiety<br />

to the porphyrin system; access to porphyrins in which one pyrrole subunit might be<br />

functionalized uniquely for that attachment is made possible by the [3+1] approach. For<br />

example, if a porphyrin attached at one point to a peptide was the synthetic target, using<br />

the [3 +1] protocol it would be possible to quickly synthesize a porphyrin bearing, for example,<br />

one carboxylic acid group (on the monopyrrole component). Less sophisticated<br />

monopyrrole polymerization methods would obviously yield a porphyrin with one carboxylic<br />

acid group on each of the four porphyrin subunits. For usefulness in such biological<br />

applications, the [3 +1] protocol must be simple, and almost as easy to operate as a<br />

monopyrrole tetramerization. Such simplicity has been facilitated by recent advances in<br />

tripyrrane syntheses (see Section 17.8.1.1.2.2) and in the accompanying monopyrrole activation<br />

[either by formylation (see Variation 1, 17.8.1.2.3.1) or (dimethylamino)methylation<br />

using Eschenmoser s reagent, (dimethylmethylidene)ammonium iodide (Variation 2,<br />

17.8.1.2.3.2)].<br />

Scheme 48 Porphyrin Syntheses from Tripyrranes<br />

R 4<br />

R 5<br />

R 3 R 2<br />

R 6<br />

NH HN<br />

NH<br />

HN<br />

R 7<br />

R 1<br />

R 8<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 4<br />

R 5<br />

R 3 R 2<br />

R 6<br />

R 7<br />

R 1<br />

R 8


17.8.1 Porphyrins 1139<br />

aFor the [3+1] approach to be used successfully, the tripyrrole component is always a tripyrrane<br />

(i.e., a tripyrrole linked by two methylene groups). To date, tripyrrenes have not<br />

been used in this approach because of the lack of nucleophilicity (in acid) of the “methene”<br />

end of the tripyrrole. The [3 +1] route is not without symmetry limitations. The Sessler<br />

approach used for the tripyrrane synthesis requires that the terminal rings of the tripyrrane<br />

be identical (i.e., R 1 =R 6 ,R 2 =R 5 , Scheme 48) , though it should not take much effort<br />

to develop a rational route to unsymmetrically substituted tripyrranes if one was required.<br />

However, even if unsymmetrical tripyrranes were readily available, one would<br />

need to use a 2,5-diformyl-1H-pyrrole in which both the 3- and 4-substituents are identical<br />

(i.e., R 7 =R 8 , Scheme 48); otherwise, two porphyrins would result.<br />

17.8.1.2.3.1 Variation 1:<br />

Using a Tripyrrane and a 2,5-Diformyl-1H-pyrrole<br />

Johnson and co-workers used a [3+1] method for synthesis of monoheteroporphyrins, but<br />

not for porphyrins themselves. [148] Boudif and Momenteau were the first to apply the<br />

[3+1] approach (Scheme 7, mode G) for the synthesis of porphyrins, using a 2,5-diformyl-1H-pyrrole<br />

and a tripyrrane; [136,137] using this route they chose to synthesize only bisacrylic<br />

or bis-propanoic porphyrins (e.g., 221) from the tripyrrane 219 and diformylpyrroles<br />

220. Lash and co-workers demonstrated the generality of the approach, and have<br />

completed a number of syntheses of porphyrins such as 224 and 225 (Scheme 49). [138–142]<br />

For the synthesis of 224, the tripyrrane-1,14-dicarboxylic acid 222 is synthesized from 2<br />

equivalents of the 2-(acetoxymethyl)-1H-pyrrole 92 and one of 4,5,6,7-tetrahydro-2H-isoindole,<br />

followed by catalytic debenzylation. The tripyrrane 222 is then treated with the 2,5diformyl-1H-pyrrole<br />

223 to give porphyrin 224. Sessler and co-workers have also shown<br />

the applicability of the approach and accomplished syntheses of porphyrins such as 226<br />

and 227 through similar methodology. [143] A one-step route for synthesis of 2,5-diformyl-<br />

1H-pyrroles is available, which makes the [3+1] approach even more attractive. [149]<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1140 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 49 Porphyrin Syntheses Using the [3+1] Approach [136–143]<br />

EtO2C<br />

EtO 2C<br />

Et<br />

NH<br />

219<br />

NH HN<br />

NH<br />

CO2H<br />

AcO<br />

1.<br />

Et<br />

Et<br />

2. H2, Pd/C<br />

CO 2H<br />

N<br />

H<br />

92<br />

1.<br />

OHC<br />

2. DDQ<br />

+<br />

EtO 2C<br />

EtO 2C<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

OHC<br />

HN<br />

CO (2 equiv), H<br />

2Bn<br />

+<br />

Et Et<br />

N<br />

H<br />

223<br />

47%<br />

CHO<br />

CHO<br />

220<br />

CO2Et<br />

Et<br />

, TFA, CH 2Cl2<br />

Et<br />

CO 2Et<br />

Et<br />

221<br />

NH HN<br />

NH<br />

CO2H<br />

222<br />

Et<br />

Et<br />

CO2Et<br />

NH N<br />

N HN<br />

224<br />

CO 2H<br />

Et<br />

Et<br />

CO2Et<br />

Et


aNH N<br />

17.8.1 Porphyrins 1141<br />

Et<br />

Et<br />

MeO 2C<br />

Et<br />

N HN<br />

225<br />

NH N<br />

N HN<br />

227<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

CO2Me<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

Et<br />

226<br />

Et<br />

CO 2Me<br />

2:3-Butano-7,12,13,18-tetraethyl-8,17-dimethylporphyrin (224); Typical Procedure: [141]<br />

7:8-Butano-3,12-diethyl-2,13-dimethyltripyrrane-1,14-dicarboxylic acid (222; 100 mg,<br />

0.22 mmol) was stirred in TFA under a N 2 atmosphere for 10 min before being diluted<br />

with CH 2Cl 2 (19 mL) and treated with 3,4-diethyl-2,5-diformyl-1H-pyrrole (223; 40mg,<br />

0.223 mmol). The mixture was stirred for 2 h under N 2, and then neutralized by dropwise<br />

addition of Et 3N. DDQ (53 mg, 0.23 mmol) was then added and the mixture was stirred at<br />

rt for 2 h under N 2. The soln was washed with H 2O, concentrated under reduced pressure,<br />

and the residue was chromatographed [neutral alumina (Brockmann Grade III), CH 2Cl 2].<br />

The porphyrin fraction was collected and concentrated under vacuum to give a residue<br />

which was crystallized (CHCl 3/MeOH) to give 224; yield: 47 mg (47%); mp >300 8C.<br />

17.8.1.2.3.2 Variation 2:<br />

Using a Tripyrrane and a 2,5-Bis[(dimethylamino)methyl]-1H-pyrrole<br />

While in the process of developing “acid-free” porphyrin macrocyclization techniques<br />

(e.g., Scheme 35), [114,115] a [3 +1] protocol for porphyrin synthesis was devised [116] which,<br />

though not a MacDonald approach, is nevertheless an example of mode G illustrated in<br />

Scheme 7. Treatment of the tripyrrane 229 with, for example, the 2,5-bis[(dimethylamino)methyl]-1H-pyrrole<br />

230 in methanol containing ferricyanide, gives the porphyrin<br />

231 in 16% yield (Scheme 50). The starting tripyrrane 228 is synthesized from 2 equivalents<br />

of the 2-(acetoxymethyl)-1H-pyrrole 42 and 3,4-diethyl-1H-pyrrole, and then subjected<br />

to catalytic debenzylation to give 229; the 2,5-bis[(dimethylamino)methyl]-1H-pyrrole<br />

230 is obtained by treatment of 3,4-diphenyl-1H-pyrrole with an excess of Eschenmoser s<br />

salt.<br />

for references see p 1223


1142 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 50 Porphyrin Syntheses Using the [3+1] Approach [116]<br />

Et Et N<br />

H<br />

42<br />

N<br />

H<br />

Ph Ph<br />

N<br />

H<br />

MeO2C<br />

AcO<br />

H2, Pd/C, THF<br />

100%<br />

(2 equiv), MeOH, TsOH<br />

CO2Bn MeO 2C<br />

84%<br />

Et<br />

H2C NMe2 I − +<br />

, MeNO2 83%<br />

MeO 2C<br />

MeO2C<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

NH HN<br />

NH<br />

CO2H<br />

229<br />

CO 2H<br />

Et<br />

CO 2Me<br />

Ph Ph<br />

Et<br />

Me2N NMe 2<br />

N<br />

H<br />

230<br />

Et<br />

Et<br />

NH HN<br />

NH<br />

CO2Bn<br />

228<br />

Ph Ph<br />

231<br />

N<br />

H<br />

Ph<br />

CO 2Me<br />

CO 2Bn<br />

Me 2N NMe2<br />

230<br />

MeOH, K3[Fe(CN)6]<br />

16%<br />

CO 2Me<br />

Dibenzyl 7,8-Diethyl-3,12-bis[2-(methoxycarbonyl)ethyl]-2,13-dimethyltripyrrane-1,14-dicarboxylate<br />

(228); Typical Procedure: [116]<br />

3,4-Diethyl-1H-pyrrole (0.37 g, 3.00 mmol) and benzyl 5-(acetoxymethyl)-4-[2-(methoxycarbonyl)ethyl]-3-methyl-1H-pyrrole-2-carboxylate<br />

(42; 2.02 g, 5.41 mmol) were dissolved in<br />

MeOH (35 mL) and TsOH (0.10 g, 0.5 mmol) was added. The mixture was heated at 608C<br />

under N 2 for 12 h before the volume was reduced to about 20 mL. The resulting suspension<br />

was stored at 0 8C for several hours, then the solid was collected by filtration and<br />

washed with cold MeOH to afford 228 as an off-white powder; yield: 1.72 g (84%); mp<br />

163–1648C.<br />

2,5-Bis[(dimethylamino)methyl]-3,4-diphenyl-1H-pyrrole (230); Typical Procedure: [116]<br />

3,4-Diphenyl-1H-pyrrole (2.60 g, 11.90 mmol) and Eschenmoser s reagent (6.61 g,<br />

35.62 mmol) were dissolved in dry MeNO 2 (250 mL) and stirred at rt under N 2 for 12 h. Further<br />

Eschenmoser s salt (another 1.52 g, 8.19 mmol) was added and the mixture was re-<br />

Ph


17.8.1 Porphyrins 1143<br />

afluxed for 15 min before the solvent was removed. Workup gave 230; yield: 3.97 g (83%);<br />

mp dec.<br />

7,8-Diethyl-3,12-bis[2-(methoxycarbonyl)ethyl]-2,13-dimethyl-17,18-diphenylporphyrin<br />

(231); Typical Procedure: [116]<br />

Tripyrrane 228 (0.33 g, 0.45 mmol) was dissolved in THF (80 mL), and 10% Pd/C (70 mg) and<br />

one drop of Et 3N were added. The resulting mixture was stirred under H 2 at rt for 10 h. The<br />

catalyst was removed by filtration and the solvent was removed from the filtrate to leave<br />

a residue. Crystallization (CH 2Cl 2/hexanes) afforded tripyrrane-1,14-dicarboxylic acid 229<br />

as a white powder in quantitative yield; because of spontaneous decarboxylation at room<br />

temperature, this was used immediately. Tripyrrane diacid 229 (0.26 g, 0.45 mmol) was<br />

dissolved in MeOH (100 mL) and refluxed for 15 min before pyrrole 230 (0.28 g,<br />

0.45 mmol) and K 3[Fe(CN) 6] (0.96 g, 2.93 mmol) were added. The resulting mixture was refluxed<br />

for 6 h. The MeOH was removed and the residue was redissolved in CH 2Cl 2. Some<br />

insoluble material was filtered off and the filtrate was washed with H 2O, 5% NH 4OH, H 2O,<br />

and brine, and then dried (Na 2SO 4). The solvent was removed and the crude mixture was<br />

chromatographed (silica gel, CH 2Cl 2). Recrystallization (CH 2Cl 2/cyclohexane) afforded the<br />

porphyrin 231 as a purple crystalline solid; yield: 52 mg (16%); mp 195–1968C.<br />

17.8.1.2.4 Method 4:<br />

From Open-Chain Tetrapyrrolic Intermediates<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, pp 590–596.<br />

Truly general porphyrin syntheses proceed in a stepwise manner from monopyrroles,<br />

eventually through unique open-chain tetrapyrrolic intermediates. In this way,<br />

the synthesis of a porphyrin with a completely asymmetric array of peripheral substituents<br />

is possible. Once constructed, with modern spectroscopic methods it is possible to<br />

verify that the intended array of substituents is present on the open-chain intermediate,<br />

and it must also be verified that the array remains intact during and after the cyclization<br />

step; this is particularly so with intermediates such as bilanes and bilenes, which have saturated<br />

methylene linkages, because cyclizations invariably require acidic reagents. The<br />

success and efficiency of a newly developed synthetic porphyrin approach is demonstrated<br />

definitively by choosing a target porphyrin which has been well-characterized by earlier<br />

workers; in this way, comparisons can be made with literature data and presumably<br />

by comparison with an authentic sample. Additionally, if the potential for acid-catalyzed<br />

pyrrole ring redistribution reactions is apparent, the target porphyrin should have pyrrole<br />

ring subunits that are differently substituted, for example, with pyrroles bearing<br />

methyl/ethyl and methyl/propanoate pairs of substituents. If acid-catalyzed scrambling<br />

of the pyrrole subunits in the open-chain tetrapyrrole occurs, the product will consist of<br />

a mixture of porphyrins containing, for example, from one to four propanoic ester<br />

groups. This complication is readily observable, even by simple thin-layer chromatography<br />

monitoring of the reaction mixture. If, on the other hand, the target molecule contains<br />

identical rings with only methyl/ethyl (or with only methyl/propanoate) substituent<br />

pairs, as in the etioporphyrins (or coproporphyrins), then more refined techniques are required<br />

to check for the presence of only one isomer. If conditions for cyclization of the<br />

open-chain tetrapyrrole are not mild, or if electron-withdrawing groups have not been<br />

strategically placed within the molecule, the synthesis may yield a mixture of several different<br />

porphyrins from a single, pure open-chain tetrapyrrole.<br />

The earliest attempt [150] to synthesize a porphyrin from a bilane targeted etioporphyrin<br />

II (8), a porphyrin in which each ring has one ethyl and one methyl substituent, the<br />

worst possible choice for a target. Only recently, for example, has it been possible to identify<br />

individual etioporphyrin type-isomers. [151] There is little doubt that this first attempt<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1144 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

ato use a bilane resulted in the formation of a mixture of etioporphyrins. [150] Only after the<br />

publication of this synthesis was the lability of bilanes [67] and porphyrinogens [107,152] toward<br />

s redistribution reactions in acidic solution, understood. The dipyrromethane methylene<br />

linkage, particularly when it lacks stabilizing electronegative substituents on adjacent<br />

pyrrole rings, is extremely unstable toward pyrrole ring redistribution reactions<br />

(Scheme 51). Protonation of a dipyrromethane 232 bearing two different rings, X and Y,<br />

at either the 4- or 6-positions, can give either 233 or 234. All steps shown in Scheme 51 are<br />

acid–base reactions, and are therefore reversible; deprotonation will reform 232, but fragmentation,<br />

as shown in structure 234, can give two species, 237 and 238. Similarly, protonated<br />

species 233 can afford 235 and 236; once again, these pairs of intermediates can<br />

react together to afford eventually 232. However, pyrrole cation 237 can react with 235 to<br />

afford the X–X dipyrromethane 239. Likewise, the Y–Y dipyrromethane 240 can be produced<br />

from a reaction between 238 and 236. Thus, treatment of the X–Y dipyrromethane<br />

232 with acid eventually results in formation of an equilibrium mixture of the X–X, X–Y,<br />

Y–X, and Y–Y dipyrromethanes. There are additional factors that can affect the position of<br />

the equilibrium; placement of electron-withdrawing groups on the X and Y pyrrole rings<br />

will inhibit protonation of that pyrrole subunit, and thus inhibit the overall redistribution<br />

reaction.<br />

Scheme 51 Acid Scrambling of Pyrroles in Dipyrromethanes<br />

H<br />

X<br />

NH<br />

+<br />

X<br />

N<br />

H<br />

233<br />

+<br />

HN<br />

Y<br />

N+<br />

H<br />

235 236<br />

X<br />

NH<br />

239<br />

X<br />

HN<br />

X Y<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

232<br />

X<br />

N<br />

H<br />

+<br />

NH<br />

234<br />

+<br />

H<br />

Y<br />

HN<br />

+<br />

Y<br />

N<br />

H<br />

237 238<br />

Y<br />

NH<br />

240<br />

Y<br />

HN


17.8.1 Porphyrins 1145<br />

aSyntheses of compounds which can be used as porphyrin intermediates have been outlined<br />

in earlier sections of this chapter. As was mentioned above, bilanes 96 without electron-withdrawing<br />

functionalities are so sensitive to acidic conditions that they cannot be<br />

used with success; with the exception of the transient intermediacy of a bilane in the aoxobilane<br />

approach (Variation 1; Section 17.8.1.2.4.1); attempts to use them have always<br />

failed to give pure porphyrins. [150] Likewise, a-bilenes, a,c-biladienes, and a,b,c-bilatrienes<br />

have not found any useful application in porphyrin syntheses, mostly because of acid<br />

scrambling at the dipyrromethane linkages and lack of nucleophilicity due to the dipyrromethene<br />

link(s).<br />

17.8.1.2.4.1 Variation 1:<br />

Using a-Oxobilanes<br />

This approach employs mode H shown in Scheme 7. The oxo function at the a-position<br />

exerts a stabilizing influence on the pyrrole rings on either side of it, but the highly electronegative<br />

nature of this carbonyl group reduces the nucleophilicity of the terminal ring<br />

such that it will not react with an electrophilic one-carbon linking unit. In order to effect<br />

macrocyclization, it is, therefore, necessary to remove the oxo function; diborane reduction<br />

of 106 (Scheme 23) gives the dibenzyl bilane-1,19-dicarboxylate, which is then catalytically<br />

debenzylated to give the bilane-1,19-dicarboxylic acid 241 (Scheme 52). As mentioned<br />

above, acid-catalyzed cyclization using a one-carbon equivalent produces a complex<br />

mixture of porphyrins due to the use of acid, so the bilane must be pre-stabilized by<br />

transformation into the b-bilene 242 by oxidation with tert-butyl hypochlorite; use of this<br />

reagent predominantly causes oxidation at the central (“b”) methylene rather than at “a”<br />

or “c”. Since bilenes are readily protonated in acid (similarly to dipyrromethenes), the bilene<br />

moiety has the effect of placing a conjugated, positively charged (and hence highly<br />

electronegative) function in the middle of the tetrapyrrole. Cyclization of the b-bilene<br />

242 in the presence of trichloroacetic acid and trimethyl orthoformate (as the one-carbon<br />

linking unit) affords pure porphyrin, coproporphyrin III tetramethyl ester (192), in 23%<br />

yield from 106 (Scheme 52). [67,68]<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1146 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 52 Porphyrin Synthesis through a-Oxobilanes [67,68]<br />

MeO 2C<br />

MeO2C<br />

O<br />

NH HN<br />

NH HN<br />

106<br />

241<br />

CO 2Me<br />

MeO2C CO 2Me<br />

NH HN<br />

NH HN<br />

CO 2Bn<br />

CO2Bn<br />

CO 2Me<br />

CO 2H<br />

CO2H<br />

MeO2C CO2Me<br />

H + , HC(OMe)3, CH2Cl2<br />

1. B2H6, THF, EtOAc<br />

2. H2, Pd/C, THF<br />

t-BuOCl<br />

THF, Et2O MeO 2C<br />

MeO2C<br />

Cl<br />

MeO2C CO2Me<br />

242<br />

−<br />

MeO 2C<br />

NH HN<br />

+<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

N<br />

N<br />

HN<br />

192 23%<br />

CO 2Me<br />

CO 2H<br />

CO2H<br />

CO2Me<br />

CO2Me<br />

The a-oxobilane porphyrin synthesis can be used to prepare a variety of unsymmetrically<br />

substituted porphyrins, and the intermediates are readily synthesized from easily accessible<br />

dipyrromethanes. Nevertheless, this route has been somewhat ignored because it is<br />

complex and lengthy, and requires excellent experimental technique. After some initial<br />

examples were published by the inventors [67,68] it moved into obscurity, being passed over<br />

in favor of the b-oxobilane and a,c-biladiene routes (discussed in Sections 17.8.1.2.4.2–<br />

17.8.1.2.4.8 inclusive).<br />

3,8,13,17-Tetrakis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin (Coproporphyrin<br />

III Tetramethyl Ester, 192); Typical Procedure: [68]<br />

Dibenzyl 3,8,13,17-tetrakis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethyl-a-oxobilane-<br />

1,19-dicarboxylate (106; 1.18 g, 1.2 mmol) in dry THF (35.2 mL) and dry EtOAc (35.2 mL)


17.8.1 Porphyrins 1147<br />

awas reduced with diborane [generated from NaBH 4 (580 mg) in bis(2-methoxyethyl) ether<br />

(21 mL) and BF 3•OEt 2 (5.8 mL) in bis(2-methoxyethyl) ether (16 mL)] over a period of 5 min.<br />

The diborane was swept into the reaction vessel with a slow stream of N 2. After about<br />

90 min the absorption due to the dipyrroketone (at 360 nm) had disappeared. The resulting<br />

colorless soln was concentrated to dryness under a reduced pressure of N 2, treated<br />

with MeOH (26 mL) and set aside under N 2 until the oil had dissolved. THF (26 mL) was<br />

then added, together with 10% Pd/C (520 mg) and a soln of Et 3N in THF (3%, 8 drops). The<br />

soln was then hydrogenated at rt and atmospheric pressure for 17 h. The mixture was filtered<br />

through Hiflosupercel under N 2 and then concentrated under reduced pressure. It<br />

was then reconcentrated from Et 2O to remove traces of THF and MeOH to give 241 as a<br />

buff-colored gummy solid. This material was dissolved in THF (140 mL) and the soln diluted<br />

to 280 mL with Et 2O. The vessel was flushed with N 2 and the soln was cooled, under N 2,<br />

to –15 8C. tert-Butyl hypochlorite (0.144 mL) in dry Et 2O (50 mL), cooled to 08C, was then<br />

added over a period of 1 h in the dark. After 15 min, a starch/KI test was negative and so<br />

the b-bilene 242 was allowed to warm to rt. The suspension was concentrated to dryness<br />

and the residue was triturated with Et 2O to give the solid b-bilene [l max (CH 2Cl 2) 505 nm].<br />

This bilene (assumed 1200 mmol), dissolved in CH 2Cl 2 (240 mL) containing HC(OMe) 3<br />

(2.56 mL), was added to dry trichloroacetic acid (12.3 g, 75.3 mmol) in CH 2Cl 2 (240 mL),<br />

and then stirred overnight under O 2 in the dark. The soln was then washed with aq 1 N<br />

Na 2CO 3 (4 ” 100 mL), H 2O (5 ” 100 mL), dried (MgSO 4), and concentrated. The residue, in a<br />

little CH 2Cl 2, was filtered through a plug of neutral alumina (Brockmann Grade III), and<br />

then chromatographed [neutral alumina (Brockmann Grade III), toluene/CH 2Cl 2 1:1]. The<br />

red eluates gave a purple solid (262 mg), which was triturated with Et 2O to give 192; yield:<br />

198 mg (23%); mp 150–155 8C and then again at 179–1828C.<br />

17.8.1.2.4.2 Variation 2:<br />

Using b-Oxobilanes<br />

This approach, through b-oxobilanes, is another example of the pyrrole ring connection<br />

mode H shown in Scheme 7. Unlike the a-oxobilane case, it is not necessary to remove the<br />

b-oxo function before proceeding to the porphyrin because it is too distant to affect the<br />

nucleophilicity of the bilane termini; in fact, the presence of the electronegative carbonyl<br />

group is a bonus because it provides protection against acid-promoted pyrrole subunit redistribution<br />

reactions in the B/C-portion of the bilane. The A and D rings, of course, are<br />

already protected from protonation by the electronegative benzyl esters. Thus, direct catalytic<br />

hydrogenation of 109 gives the dicarboxylic acid 110 (see Scheme 24), which can<br />

then be cyclized (as for a-oxobilanes) by treatment with trichloroacetic acid and trimethyl<br />

orthoformate, to give the oxophlorin 213 after air oxidation and chromatographic purification;<br />

yields from dipyrromethanes are usually quite high (Scheme 53). However, the<br />

blue oxophlorin product must first be transformed into a red porphyrin by reduction of<br />

the meso-carbonyl group. Methods for carrying out this transformation were discussed<br />

earlier (see Section 17.8.1.2.2.3, Scheme 47), and as mentioned then, the best method involves<br />

firstly the preparation of the enol-acetate 216 of the oxophlorin by treatment with<br />

acetic anhydride in pyridine. This is followed by catalytic hydrogenation and re-oxidation<br />

of the resulting porphyrinogen with 2,3-dichloro-5,6-dicyanobenzo-1,4-quinone, to give<br />

mesoporphyrin IX dimethyl ester (215) (see Scheme 47). [69]<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1148 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 53 Porphyrin Synthesis through b-Oxobilanes [69]<br />

Et<br />

O<br />

B A<br />

NH HN<br />

NH HN<br />

C D<br />

110<br />

Et<br />

CO 2H<br />

CO2H<br />

MeO2C CO 2Me<br />

H + , HC(OMe) 3, CH 2Cl 2<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

70%<br />

Et<br />

O<br />

MeO 2C<br />

NH<br />

NH<br />

213<br />

N<br />

HN<br />

Et<br />

CO2Me<br />

The b-oxobilane approach has experienced much more success than the a-oxobilane<br />

route, and has been used for the synthesis of several unsymmetrical porphyrins. [48,153,154]<br />

The route involves a large number of steps, and for this reason has been superceded by<br />

the various a,c-biladiene routes, and also by the MacDonald [2 +2] pathway. Nevertheless,<br />

it can be used for the synthesis of a wide variety of porphyrins, and has the significant<br />

advantage that it affords the biologically interesting oxophlorins as intermediates. The<br />

dipyrromethane starting materials are available in large amounts, and unlike the corresponding<br />

dipyrromethenes, are easy to purify by chromatography or by recrystallization.<br />

3,18-Diethyl-8,12-bis[2-(methoxycarbonyl)ethyl]-10-oxo-2,7,13,17-tetramethylphlorin<br />

(213); Typical Procedure: [69]<br />

3,8-Diethyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethyl-b-oxobilane-1,19dicarboxylic<br />

acid (110; 368 mg, 0.53 mmol) in CH 2Cl 2 (50 mL) was treated successively<br />

with 1 M trichloroacetic acid in CH 2Cl 2 (30 mL) and HC(OMe) 3 (1.16 mL) in CH 2Cl 2<br />

(130 mL). The deep red soln was stirred for 2 h in the dark before addition of pyridine<br />

(2.4 mL), and the mixture was stirred exposed to the air overnight. The resulting green<br />

soln was concentrated to dryness under reduced pressure, toluene (20 mL) was added<br />

and the pyridinium trichloroacetate was filtered off and washed with a little toluene.<br />

The combined organic phases were concentrated to give a green oil, which was chromatographed<br />

twice [neutral alumina (Brockmann Grade III), CH 2Cl 2]. The blue eluates were<br />

concentrated to a small volume and on being kept at 0 8C the oxophlorin 213 crystallized<br />

as deep blue prisms; yield: 227 mg (70%); mp 186–1878C.<br />

17.8.1.2.4.3 Variation 3:<br />

Using 1,19-Dimethyl-b-bilenes<br />

As mentioned earlier, cyclization reactions of a-bilenes do not give pure porphyrins. [155]<br />

However, use of b-bilenes has resulted in good success. The various synthetic uses of b-bilenes<br />

have been summarized by Clezy, [21] who has been the major proponent for the use<br />

of b-bilenes in porphyrin synthesis since A. W. Johnson s time. Clezy has pointed out that<br />

the b-bilene method works best when electron-withdrawing substituents are present on<br />

the b-bilene. [21]<br />

b-Bilenes 115 (see Scheme 25) bearing 1- and 19-methyl groups readily undergo oxidative<br />

cyclization to give copper(II) porphyrinates 243 by using copper(II) salts in pyridine<br />

or in dimethylformamide (Scheme 54). [71–73] The copper porphyrinates can be demetalated<br />

with concentrated sulfuric acid, or better, with concentrated sulfuric acid in trifluoroacetic<br />

acid, to give metal-free porphyrin 244. [79,80] This procedure has also been used successfully<br />

[21,74,156–159] for the cyclization of b-bilenes bearing electron-withdrawing groups. The<br />

mechanism of the cyclization of 1,19-dimethyl-b-bilenes and of a,c-biladienes to give porphyrins<br />

has been extensively investigated. It appears that both b-bilenes and a,c-bila-


17.8.1 Porphyrins 1149<br />

adienes are transformed into the same a,b,c-bilatriene species in the first steps of the<br />

mechanism. Since most of the definitive isotope labeling and other work was carried<br />

out using a,c-biladienes, the mechanism will be discussed in full in that section of this<br />

chapter (Section 17.8.1.2.4.6).<br />

Scheme 54 Porphyrin Synthesis Using 1,19-Dimethyl-b-bilenes [71–73]<br />

NH<br />

N<br />

115<br />

HN<br />

HN<br />

Ac<br />

Ac<br />

Cu(II), DMF<br />

heat or py<br />

H2SO4 or<br />

H2SO4, TFA<br />

N N<br />

N N<br />

The b-bilene general method for porphyrin synthesis has many advantages, mostly related<br />

to its simplicity and to the way it can be used to prepare relatively large quantities of<br />

porphyrins. Its disadvantages relate to substituent limitations associated with the bilene<br />

intermediates, and difficulty in purification of the b-bilenes if they do not happen to crystallize<br />

spontaneously.<br />

3,7-Diacetyl-2,8,12,13,17,18-hexamethylporphyrin (244); Typical Procedure: [74]<br />

2,18-Diacetyl-1,3,7,8,12,13,17,19-octamethyl-b-bilene (115; all of the material isolated<br />

from the preparation described in Section 17.8.1.1.3.2.1) in pyridine (40 mL) was treated<br />

with Cu(OAc) 2 (2 g) and the mixture was stirred for 10 min at 408C. AcOH (5 mL) and<br />

more Cu(OAc) 2 (3 g) were then added and the mixture was stirred for 2 h at 608C before<br />

being diluted with EtOH (50 mL). The precipitate was collected by filtration after 4 h, and<br />

washed with EtOH to give the copper(II) complex 243 (400 mg), mp >3008C. A soln of the<br />

copper(II) complex (200 mg) in concd H 2SO 4 (10 mL) was poured onto ice (100 g) and the<br />

mixture was stirred for 4 h. The precipitate was collected to give 244; yield: 150 mg; mp<br />

>350 8C.<br />

17.8.1.2.4.4 Variation 4:<br />

Using b-Bilene-1,19-diesters<br />

The synthesis of mesoporphyrin XIII dimethyl ester 245 is shown in Scheme 55. [76] Treatment<br />

of the di(tert-butyl) b-bilene-1,19-diester 117 with trifluoroacetic acid, followed by<br />

cyclization in the presence of trichloroacetic acid and trimethyl orthoformate (as the<br />

one-carbon linking unit) and then oxidation with air gives a 57% yield of porphyrin. The<br />

b-bilenes can be conveniently prepared from a tert-butyl 9-formyldipyrromethane-1-carboxylate<br />

and a tert-butyl 9-unsubstituted dipyrromethane-1-carboxylate (see Section<br />

17.8.1.1.1.2.1). This method is simple, direct, and yields crystalline intermediates at all<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Cu<br />

243<br />

Ac<br />

Ac<br />

NH<br />

N<br />

244<br />

N<br />

HN<br />

Ac<br />

Ac<br />

for references see p 1223


1150 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

astages. However, it can be erratic and occasionally yields mixtures, particularly when<br />

electron-withdrawing substituents are sited on the B–C portion of the b-bilene.<br />

Scheme 55 Porphyrin Synthesis Using b-Bilene-1,19-diesters [76]<br />

Et<br />

MeO 2C<br />

NH<br />

+<br />

NH<br />

HN<br />

HN<br />

Bu t O 2C CO 2Bu t<br />

Et<br />

Cl −<br />

1. TFA<br />

2. HC(OMe)3<br />

H + , O2<br />

N HN<br />

NH N<br />

CO2Me<br />

MeO2C CO2Me 118 245<br />

2,8-Diethyl-13,17-bis[2-(methoxycarbonyl)ethyl]-3,7,12,18-tetramethylporphyrin (Mesoporphyrin<br />

XIII Dimethyl Ester, 245); Typical Procedure: [76]<br />

Di-tert-butyl 7,13-diethyl-2,18-bis[2-(methoxycarbonyl)ethyl]-3,8,12,17-tetramethyl-b-bilene-1,19-dicarboxylate<br />

(118; 247 mg, 0.284 mmol) in TFA (15 mL) was purged with N 2 for<br />

20 min. The TFA was removed under reduced pressure, dry benzene (CAUTION: carcinogen)<br />

was added and removed under vacuum, and the resulting red oil was taken up in<br />

CH 2Cl 2 (50 mL). It was washed with H 2O, aq Na 2CO 3, and then H 2O again. The deep yellow<br />

soln was dried (MgSO 4), concentrated to dryness under reduced pressure, and the residue<br />

was dissolved in CH 2Cl 2 (57 mL) containing HC(OMe) 3 (0.65 mL) before addition of 1 M trichloroacetic<br />

acid in CH 2Cl 2 (42 mL). Spectrophotometry showed porphyrin formation to<br />

be almost instantaneous, but the soln was stirred in air overnight in the dark at rt. It was<br />

washed with aq Na 2CO 3 and then H 2O. After drying (MgSO 4), the soln was concentrated to<br />

dryness under reduced pressure to give a red-brown oil which was separated from the<br />

nonvolatile liquid by heating at 508C and 0.2 Torr for 30 min. Chromatography (2 ”) (alumina,<br />

CH 2Cl 2) separated a red band. These fractions were concentrated to dryness to give<br />

a red residue, which was crystallized (CH 2Cl 2/MeOH) to give mesoporphyrin XIII dimethyl<br />

ester (245); yield: 102 mg (57%); mp 2178C.<br />

17.8.1.2.4.5 Variation 5:<br />

Using Other b-Bilenes<br />

As mentioned earlier, a b-bilene 198 (Scheme 43) is a transient intermediate in Woodward<br />

s synthesis of chlorophyll a [44] and provides an elegant way of circumventing the<br />

symmetry limitations normally apparent using the MacDonald [2+2] route. Deoxophylloerythroetioporphyrin<br />

247 is also synthesized [160] in low yield from b-bilene 246 (Scheme<br />

56). tert-Butyl 19-methyl-b-bilene-1-carboxylates have been prepared, [161] but they are not<br />

easy to cyclize to porphyrins. A synthesis [156–159] of porphyrins using the condensation of 1formyl-9-methyldipyrromethanes<br />

with 1-unsubstituted 9-(N,N-dimethylformimino)dipyrromethane<br />

salts in methanol/acetic acid, proceeds by way of a b-bilene intermediate.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

57%<br />

Et<br />

Et


17.8.1 Porphyrins 1151<br />

aScheme 56 Porphyrin Synthesis Using Other b-Bilenes [160]<br />

Et<br />

Et<br />

NH<br />

NH<br />

R 1<br />

R 2<br />

+<br />

HN<br />

HN<br />

Et<br />

Cl −<br />

NH N<br />

N HN<br />

246 247<br />

R 1 = 4-MeOC6H4CH2CO2; R 2 = CHO<br />

17.8.1.2.4.6 Variation 6:<br />

Using 1,19-Dimethyl-a,c-biladienes<br />

The synthesis and subsequent oxidative cyclization of a,c-biladienes, and particularly<br />

1,19-dimethyl-a,c-biladienes 248, to give metal porphyrinates 249 has been thoroughly<br />

reviewed recently. [20]<br />

Johnson and Kay [71] have shown that oxidative cyclization of 1,19-dimethyl-a,c-biladiene<br />

and 1,19-dimethyl-b-bilene salts with copper(II) acetate in methanol gives porphyrins<br />

in overall yields as high as 30% (Scheme 57). Use of boiling dimethylformamide for<br />

brief periods, rather than extended periods of refluxing in methanol, [162] or cyclization at<br />

room temperature with copper(II) chloride, results in improved yields. Treatment of the<br />

zinc(II) complex of the a,c-biladiene with a variety of oxidizing agents [163] affords yields<br />

similar to those obtained by the copper(II) salt method in hot dimethylformamide. Cyclization<br />

of 1,19-dimethyl-a,c-biladienes 248 via metal porphyrinates [usually copper(II)<br />

complexes, 249] to give porphyrins (e.g., 8) must involve loss of at least one of the 1-/19methyl<br />

groups in the first step. It is possible that both methyl groups are lost, and that a<br />

one-carbon unit from the solvent (usually dimethylformamide) is inserted. However,<br />

much more likely, one methyl group is lost in a manner similar to the way in which one<br />

of the pyrrole -carbons is lost in the synthesis of symmetrical dipyrromethanes from 2-<br />

(bromomethyl)- or 2-(acetoxymethyl)-1H-pyrroles (see Section 17.8.1.1.1.2.3).<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

Et<br />

for references see p 1223


1152 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 57 Porphyrin Synthesis from 1,19-Dimethyl-a,c-biladienes [162]<br />

Et<br />

Et<br />

NH<br />

+<br />

NH<br />

248<br />

HN<br />

1<br />

+ 19<br />

HN<br />

Et<br />

Et<br />

2Br −<br />

CuCl2 2H2O<br />

DMF, heat<br />

H2SO4 or<br />

H2SO4, TFA<br />

N N<br />

N N<br />

Et Et<br />

Et Et<br />

NH N<br />

In the simplest terms, 1,19-dimethyl-a,c-biladiene salts can be oxidatively cyclized to give<br />

porphyrins using a copper(II) oxidant in hot dimethylformamide. The copper salt is usually<br />

copper(II) acetate; copper(II) chloride has been used but this can result in the production<br />

of chlorinated porphyrins as byproducts. [164] The product from the copper cyclization<br />

is invariably the copper complex of the porphyrin, but this can be demetalated to give<br />

metal-free porphyrin using acid (usually 5–20% sulfuric acid in trifluoroacetic acid). Interestingly,<br />

if a chromium(III) salt is used in the oxidative cyclization, metal-free porphyrins<br />

are produced, and this can be extremely advantageous if the product porphyrin is sensitive<br />

to strong acids. Metal-free porphyrins can also be obtained when 1,19-dimethyl-a,cbiladiene<br />

salts are subjected to cyclization using anodic oxidation. [47,165]<br />

The Oxidant: It was long believed that the metal salt used in the oxidative cyclization<br />

played a dual role: (a) as a one-electron oxidizing agent and therefore the reagent driving<br />

the chemistry of the reaction, and (b) as a chelating or templating agent which arranged<br />

the a,c-biladiene ligand into a cyclic conformation, thus ensuring that its terminal methyl<br />

groups were close enough to facilitate macrocyclization. However, it has been shown that<br />

copper(II) is not uniquely qualified to perform this reaction, and recent results appear to<br />

show that chelation is not a critically important factor; when a,c-biladienes were treated<br />

with a known non-chelating oxidant in place of, for example, a copper(II) salt, the lack of<br />

importance of templation became obvious. [166]<br />

In 1986, a study using 20 oxidants was reported. [163] In some of the reactions, zinc(II)<br />

was used as a separate chelating salt, but it was also hoped that the zinc would prevent<br />

other metals, more difficult to remove than zinc, from ending up in the center of the<br />

product porphyrin. Optimal yields of porphyrin were obtained with potassium dichromate<br />

or silver iodate, and other oxidants, such as lead(IV) oxide, potassium iodide, mercury(II)<br />

acetate, silver(I) oxide, and silver(I) acetate, gave only moderate yields of porphyrin;<br />

others gave only a trace of porphyrin, or none at all.<br />

Boschi and co-workers [167] demonstrated that chromium(III) salts promote very efficient<br />

cyclizations of a,c-biladienes 248 to directly afford metal-free porphyrins 8. The<br />

chromium methodology has also been extended to the use of chromium(III) hydroxy acetate<br />

[Cr 3(OAc) 7(OH) 2] instead of chromium(III) acetate. [168,169] This reagent also affords good<br />

yields of metal-free porphyrins, but dimethylformamide at 1008C is chosen as the prefer-<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

61%<br />

Et<br />

Et<br />

Cu<br />

249<br />

N<br />

8<br />

Et<br />

HN<br />

Et


17.8.1 Porphyrins 1153<br />

ared solvent, rather than hot ethanol buffered with sodium acetate (as preferred by Boschi<br />

and co-workers).<br />

The mechanism of the cyclization of a,c-biladienes to give porphyrins using the copper(II)<br />

salt method, or using chromium(III), or electrochemical oxidation, has been extensively<br />

investigated, and will be discussed in the mechanism section. It should be mentioned<br />

that since the first step in both the a,c-biladiene and b-bilene cyclizations has<br />

been postulated to be formation of an a,b,c-bilatriene, it is very likely that after that first<br />

step, the mechanisms of the a,c-biladiene and b-bilene cyclizations are identical.<br />

The Mechanism – Origin of the Bridging Meso-Carbon Atom: As mentioned earlier, at<br />

least one of the 1- or 19-methyl groups must be lost in the macrocyclization reaction. It<br />

is possible that both carbons are lost, since dimethylformamide has been shown to provide<br />

the meso-carbon in certain cyclizations of a,c-biladienes to give porphyrins. [162] As early<br />

as 1969, however, Grigg and co-workers [162] postulated a free-radical mechanism for the<br />

cyclization, in which one of the two a,c-biladiene terminal methyl groups is incorporated<br />

into the newly formed porphyrin at a meso position. More support for this idea was published<br />

in 1971, when it was shown [170] that oxidative cyclization of the 2,18-di-unsubstituted<br />

1,19-dimethyl-a,c-biladiene 250 using copper(II)/lead(IV) oxide gives the 13-formylporphyrin<br />

251, this suggesting that the “second” methyl group is somehow oxidized and has<br />

migrated to the b-position of the product porphyrin 251 as indicated in Scheme 58. 13C studies have eventually confirmed definitively that one of the a,c-biladiene terminal<br />

methyl groups finishes up as a bridging meso-carbon; the 13C enriched a,c-biladiene dihydrobromide<br />

252 is synthesized using the methodology described above. Cyclization using<br />

copper(II) chloride in dimethylformamide affords porphyrin 253 in which the new mesocarbon<br />

is enriched with 13C. [171]<br />

Scheme 58 Studies Showing the Origin of the Bridging meso-Carbon Atom [170,171]<br />

MeO2C<br />

MeO 2C<br />

+<br />

NH HN<br />

2Br N N<br />

−<br />

CuCl2/PbO2 250<br />

+<br />

NH<br />

+<br />

NH<br />

252<br />

HN<br />

+<br />

HN<br />

13CH3<br />

13 CH3<br />

2Br −<br />

Cu(II), DMF<br />

MeO 2C<br />

MeO2C<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

− Cu<br />

Cu<br />

251<br />

CHO<br />

NH<br />

N<br />

253<br />

N<br />

HN<br />

13 C<br />

for references see p 1223


1154 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aIt has been shown subsequently [172] that copper(II)-mediated cyclization of the 2-unsubstituted<br />

1,19-dimethyl-a,c-biladiene 254 (R 1 ,R 2 = Me) affords the expected porphyrin 259,<br />

along with the 15-methyl- (256), 15-dimethylamino- (255), 13-formyl- (257), and 15-formyl-<br />

(258) porphyrins (Scheme 59), depending upon the conditions used. Isolation of compound<br />

257 directly complements the earlier Russian results (Scheme 58), [170] and compounds<br />

256 and 258 also provide further information on the fate of the “missing” a,c-biladiene<br />

terminal carbon after it is cleaved during the cyclization.<br />

Scheme 59 Products of Cyclization of 2-Unsubstituted 1,19-Dimethyl-a,c-biladiene [172]<br />

Et<br />

Et<br />

Et<br />

Et<br />

N N<br />

Cu<br />

N N<br />

NMe2<br />

255<br />

N N<br />

Cu<br />

N N<br />

259<br />

Et<br />

Et<br />

CuCl2, PbO2 Cu(OAc) 2<br />

DMF<br />

DMF<br />

9.1% 9.3%<br />

CuSO4<br />

DMF<br />

23%<br />

Et<br />

Et<br />

Et<br />

Et<br />

N N<br />

N N<br />

N N<br />

N N<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

+<br />

NH<br />

R 1<br />

Cu<br />

HN<br />

+<br />

R 2<br />

CHO<br />

258<br />

HN<br />

254 R 1 = R 2 = Me, 13 CH 3<br />

3.7% Cu(NH4) 2Cl 4<br />

DMF<br />

Et<br />

Et<br />

2Br −<br />

Et<br />

Et<br />

CuCl 2, PbO 2<br />

DMF<br />

Et<br />

Et<br />

Cu<br />

256<br />

N N<br />

Cu<br />

N N<br />

The doubly and singly 13 C-labeled a,c-biladienes 254 (R 1 ,R 2 = 13 CH 3 and R 1 = 13 CH 3,R 2 =Me,<br />

respectively) have been prepared using standard methods and then cyclized. [172] Using either<br />

doubly or singly 13 C-labeled a,c-biladiene 254 (Scheme 59), the 15-dimethylaminoporphyrin<br />

255 contains no 13 C enrichment, confirming that the dimethylamino<br />

group and the bridging meso-carbon are derived from the dimethylformamide solvent.<br />

The labeled a,c-biladienes 254 afford porphyrins that show that terminal methyls migrate<br />

only across the pyrrole ring to which they were originally attached (in the a,c-biladiene),<br />

and confirm in all respects the mechanisms proposed [166,172] to account for the cyclization.<br />

The Mechanism – An Intermediate in the Cyclization Reaction: An important result was<br />

obtained upon anodic oxidation of 1,19-dimethyl-a,c-biladienes (e.g., 123, Scheme 60).<br />

This methodology was investigated as a means of promoting 1,19-dimethyl-a,c-biladiene<br />

9.2%<br />

257<br />

CHO<br />

Et<br />

Et


17.8.1 Porphyrins 1155<br />

acyclizations without the troublesome chelation of, for example, copper. It was argued<br />

that the cleanest of all oxidative cyclizations might simply involve an anode, from which<br />

the cyclized product could readily be retrieved without contamination. This approach<br />

was successful, and provided vital information on the mechanism of the cyclization; it<br />

also enables the isolation of macrocyclic intermediates, which allowed conclusions to be<br />

drawn for both the anodic oxidation approach and the standard copper(II) procedure. Cyclic<br />

voltammetry showed that a,c-biladienes, such as 123, are subject to clean and reversible<br />

one-electron oxidations. [47,165] Bulk electrolyses (using an “H” cell) were carried out at<br />

800 mV and gave spectrophotometric evidence (Soret absorption band at 400–410 nm) of<br />

porphyrin formation. Larger scale electro-oxidations at 800 mV produced a large quantity<br />

of a blue-green intermediate 260 with a very simple 1 H NMR spectrum. This demonstrated<br />

that the 1-methyl macrocycle 260 is an intermediate on the pathway from the a,c-biladiene<br />

123 to the porphyrin 262, and after spectroelectrochemistry, also provided definitive<br />

evidence for formation of a fully oxidized tetrapyrrole 261 prior to the loss of the 1methyl<br />

group to afford porphyrin 262. Metal-free 1-substituted macrocycles (such as 260)<br />

can be further transformed to produce porphyrins using electrochemistry. [47,165] The 1methyl<br />

intermediate 260 can also be obtained chemically by oxidative cyclization of the<br />

decamethyl-a,c-biladiene 123 with, for example, potassium ferricyanide. [166]<br />

Scheme 60 Intermediates in the a,c-Biladiene Cyclization [47,165]<br />

NH<br />

+<br />

HN<br />

NH N<br />

NH<br />

+<br />

HN<br />

2Br 63%<br />

N HN<br />

−<br />

K3[Fe(CN) 6]<br />

DMF<br />

123 260<br />

[O]<br />

NH N<br />

N N<br />

261<br />

NH N<br />

N HN<br />

a,c-Biladienes (e.g., 263) substituted at their 1- and 19-positions with groups other than<br />

methyl (e.g., acetic and propanoic ester, arylmethyl) also undergo ring cyclization using<br />

copper(II) or chromium(III) reagents (Scheme 61). However, since these bulky 1,19-substituents<br />

are not as easily eliminated as methyl, stable macrocycles, such as 264, are<br />

formed, and at high temperatures these can be induced to undergo intramolecular migrations<br />

to give chlorins (such as 265 and 266) and even novel meso-functionalized porphyrins<br />

(e.g., 267). [168,169,173]<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

262<br />

for references see p 1223


1156 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 61 Cyclizations of Novel a,c-Biladiene Salts [168,169,173]<br />

NH<br />

+<br />

NH<br />

EtO2C<br />

heat<br />

263<br />

HN<br />

+ Cu 2+<br />

2Br −<br />

HN<br />

4-Tol<br />

N<br />

N<br />

Cu<br />

EtO2C<br />

265<br />

N<br />

N<br />

4-Tol<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

heat<br />

heat<br />

N<br />

N<br />

Cu<br />

EtO2C<br />

264<br />

N<br />

N<br />

N<br />

N<br />

4-Tol<br />

Cu<br />

N<br />

N<br />

CO 2Et<br />

266<br />

N<br />

N<br />

Cu<br />

N<br />

N<br />

CO2Et 267<br />

The mechanism proposed [166] for formation of a porphyrin from 1,19-dimethyl-a,c-biladienes<br />

through oxidative cyclization is outlined in Scheme 62. Deprotonation of the a,cbiladiene<br />

268 affords the conjugated a,b,c-bilatriene-type molecule 269 (a step also proposed<br />

[21] as the first in the cyclization of b-bilenes, see Scheme 63). This is then oxidized<br />

(chemically or electrochemically) to give the 1-alkyl radical 270; loss of another electron<br />

and proton produces an intermediate 271 with an exocyclic double bond. This compound<br />

then undergoes electrocyclic ring closure to form the isolated 1-methyl intermediate 272<br />

(cf. 260), which then undergoes metalation and oxidation to give 273 (cf. macrocycle 261)<br />

followed by nucleophilic displacement of the 1-methyl group to give metal porphyrinate<br />

274.<br />

4-Tol<br />

4-Tol


17.8.1 Porphyrins 1157<br />

aScheme 62 Proposed Mechanism of the Cyclization of a,c-Biladiene Salts To Give<br />

Porphyrins [166]<br />

R 7<br />

R 6<br />

− e − , − H +<br />

R 5<br />

NH<br />

HN<br />

+ +<br />

NH HN<br />

R 4<br />

R 8 R 1<br />

electrocyclic<br />

ring closure<br />

R 7<br />

268<br />

R 6<br />

R 5<br />

NH<br />

NH<br />

R 3<br />

R 2<br />

2Br −<br />

HN<br />

N<br />

R 4<br />

R 8 R 1<br />

270<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 3<br />

− 2H +<br />

R 2<br />

R<br />

NH N<br />

6 R3 R 7<br />

R 5<br />

R 8<br />

N<br />

272<br />

HN<br />

R 4<br />

R 1<br />

R 2<br />

R 7<br />

− e − , − H +<br />

R 6<br />

1. metalation<br />

2. [O]<br />

R 5<br />

NH<br />

NH<br />

R 6<br />

R 7<br />

R 5 R 4<br />

R 8<br />

HN<br />

N<br />

N<br />

N<br />

R 4<br />

R 8 R 1<br />

R 7<br />

R 6<br />

269<br />

R 5<br />

N<br />

NH<br />

+<br />

M<br />

273<br />

N<br />

HN<br />

N<br />

N<br />

R 3<br />

R 2<br />

R 4<br />

R 8 R 1<br />

R 6<br />

271<br />

N<br />

N<br />

Nu −<br />

M<br />

N<br />

N<br />

R 1<br />

R 5 R 4<br />

As mentioned above, a mechanism has also been proposed for the cyclization of b-bilenes<br />

(e.g., 275), which is presented in Scheme 63. [21] It is similar in many respects to the mechanism<br />

for a,c-biladiene cyclizations, since the first proposed intermediate in Scheme 63 is<br />

the same compound that is proposed in Scheme 62 (cf. 269). [166] The main differences between<br />

the a,c-biladiene and b-bilene proposals are: (i) the proposed involvement of radical<br />

processes in the a,c-biladiene mechanism, and (ii) the intermediacy of a metalloporphodimethene<br />

276 in the latter pathway.<br />

R 7<br />

R 8<br />

274<br />

R 1<br />

R 3<br />

R 2<br />

R 3<br />

R 2<br />

R 3<br />

R 2<br />

for references see p 1223


1158 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 63 Proposed Mechanism of Cyclization of b-Bilene Salts To Give Porphyrins [21]<br />

R 7<br />

R 6<br />

R 5<br />

NH<br />

NH<br />

HN<br />

+<br />

HN<br />

R 4<br />

R 8 R 1<br />

R 7<br />

R 6<br />

275<br />

R 5<br />

NH<br />

NH<br />

R 3<br />

R 2<br />

HN<br />

HN<br />

R 4<br />

R 8 R 1<br />

R 6<br />

R 7<br />

[O]<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 3<br />

R 2<br />

R 5 R 4<br />

R 8<br />

N N<br />

Cu<br />

N<br />

276<br />

N<br />

R 7<br />

R 6<br />

R 5<br />

N<br />

NH<br />

1. electrocyclization<br />

2. oxidation<br />

3. copper insertion<br />

R 1<br />

R 3<br />

R 2<br />

HN<br />

HN<br />

R 4<br />

R 8 R 1<br />

[O]<br />

R 6<br />

R 7<br />

R 6<br />

R 3<br />

N<br />

R 2<br />

R5 R4 H +<br />

R 8<br />

Nu −<br />

N N<br />

Cu<br />

N<br />

N<br />

N N<br />

Cu<br />

N<br />

R 1<br />

R 5 R 4<br />

(2,8,12,18-Tetraethyl-3,7,13,17-tetramethylporphyrinato)copper(II) [Copper(II) Etioporphyrin<br />

II, 249]; Typical Procedure: [162]<br />

A soln of CuCl 2•2H 2O (6.2 g, 20 equiv) in DMF (62 mL) was added to 2,8,12,18-tetraethyl-<br />

1,3,7,13,17,19-hexamethyl-a,c-biladiene dihydrobromide (248; 1.2 g, 1 equiv) and the mixture<br />

was refluxed gently for 2 min. After cooling, the crystalline copper(II) porphyrinate<br />

was collected on Celite, washed with H 2O and MeOH (20 mL), and extracted with CHCl 3<br />

(400 mL) using a Soxhlet apparatus. The extract was reduced in volume to ca. 10 mL and<br />

hot MeOH was added. The copper(II) porphyrinate 249 crystallized; yield: 598 mg (61%);<br />

mp >300 8C. Demetalation was accomplished with concd H 2SO 4 to give etioporphyrin II<br />

(8).<br />

Cyclization of 1,2,3,7,8,12,13,17,18,19-Decamethyl-a,c-biladiene Dihydrobromide (123)<br />

To Give 1,2,3,7,8,12,13,17,18-Nonamethyl-20-dihydroporphyrin (260); Typical<br />

Procedure: [166]<br />

DMF (25 mL) was purged with dry argon for 15 min before a,c-biladiene dihydrobromide<br />

123 (56 mg, 0.093 mmol), K 3[Fe(CN) 6] (260 mg, 0.93 mmol), and Et 3N (1–2 mL) were added.<br />

The temperature was increased to 1008C for 60 min. The soln was poured into ice water<br />

and extracted with CH 2Cl 2 (3 ” 150 mL). The combined organic layers were then washed<br />

with deionized H 2O (2 ” 100 mL) and sat. NaCl soln (100 mL), dried (Na 2SO 4), filtered, and<br />

concentrated. Residual DMF was removed in vacuo. The residue was then chromatograph-<br />

R 7<br />

R 8<br />

R 1<br />

R 3<br />

R 2<br />

R 3<br />

R 2


17.8.1 Porphyrins 1159<br />

aed (alumina, CH 2Cl 2/petroleum ether 1:1), to afford the 1-methyl intermediate 260; yield:<br />

26 mg (63%); mp >300 8C.<br />

17.8.1.2.4.7 Variation 7:<br />

Using 1-Bromo-19-methyl-a,c-biladienes<br />

1-Bromo-19-methyl-a,c-biladiene dihydrobromides 139 can be cyclized to give the corresponding<br />

porphyrin, e.g. mesoporphyrin IX dimethyl ester (215) in very high yield by<br />

heating in 1,2-dichlorobenzene [84] (Scheme 64) or at room temperature in dimethyl sulfoxide<br />

and pyridine. [174] These a,c-biladienes 139 can also be cyclized to produce porphyrins<br />

by treatment with copper(II) salts and other one-electron oxidizing agents. [162] In a certain<br />

sense, this procedure developed by Johnson and co-workers can be considered to be an<br />

ingenious two-stage version of Fischer s dipyrromethene approach. Use of the two stages<br />

eliminates the symmetry restrictions of the [2 +2] dipyrromethene approach. The route<br />

affords good yields at all of its intermediate and final stages, is amenable to large-scale<br />

preparations and has been used to prepare a large variety of unsymmetrically substituted<br />

porphyrins. Its main disadvantage is that the dipyrromethene starting materials 137 and<br />

138 (Scheme 31) are often difficult to obtain and purify in more complex cases.<br />

Scheme 64 Preparation of Porphyrins from 1-Bromo-19-methyl-a,c-biladienes [84]<br />

Et<br />

Br<br />

Et<br />

NH<br />

+<br />

NH<br />

HN<br />

+<br />

HN<br />

CO 2Me<br />

2Br −<br />

CO 2Me<br />

1,2-Cl2C6H4 heat<br />

69−93%<br />

139 215<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

NH<br />

N<br />

N<br />

HN<br />

CO 2Me<br />

CO2Me<br />

Cyclization of 1-Bromo-19-methyl-a,c-biladiene Salts (e.g., 139) To Give Porphyrins (e.g.,<br />

215); General Procedure: [84]<br />

The 1-bromo-19-methyl-a,c-biladiene dihydrobromide (200 mg) was suspended in redistilled<br />

dry 1,2-dichlorobenzene (50 mL) and then refluxed for 15 min. The solvent was removed<br />

under reduced pressure. When the product contained free-acid substituents, these<br />

were (re)esterified by dissolving the crude residue in 5% H 2SO 4 in MeOH, and leaving for<br />

several hours at rt. The soln was then diluted with H 2O, the porphyrin ester was extracted<br />

into CHCl 3, and this soln was washed with dil aq NH 4OH, then with H 2O, and dried<br />

(MgSO 4). After concentration to ca. 15 mL the soln was chromatographed [alumina<br />

(Spence type H), CHCl 3]. Concentration of the eluates, followed by crystallization of the<br />

residue, gave the porphyrin; yield: 69–93%. When the porphyrin contained no free-acid<br />

substituents the crude residue was extracted with CHCl 3 and the extract was then chromatographed<br />

as above.<br />

17.8.1.2.4.8 Variation 8:<br />

Using Other a,c-Biladienes<br />

1,19-Di-unsubstituted a,c-biladiene salts are most often used for synthesis of corroles (see<br />

Section 17.8.3.2.3). [175,176] Heating of 1-unsubstituted 19-methyl-a,c-biladienes furnishes<br />

porphyrins, and these compounds are more often used as intermediates in the synthesis<br />

of tetradehydrocorrins, [176] but a number of such a,c-biladienes have been transformed<br />

into porphyrins, such as 277 (Scheme 65). [83] In related studies, 1-methyl-a,c-biladiene-19-<br />

for references see p 1223


1160 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

acarboxylic acids (e.g., 278) can be cyclized very efficiently to give porphyrins 279 by heating<br />

in 1,2-dichlorobenzene in the presence of iodine, [177,178] or iodine and bromine. [179]<br />

1,19-Di-unsubstituted a,c-biladienes and a,c-biladiene-1,19-dicarboxylic acids 280 have<br />

been efficiently converted into porphyrins 281 under acidic conditions by using aldehydes<br />

to provide the new meso-carbon atom (Scheme 65). [71,180–182]<br />

Scheme 65 Preparation of Porphyrins by Cyclization of Other a,c-Biladienes [71,83,177–182]<br />

MeO2C<br />

A<br />

R 1 = H, CO2H, I<br />

AcO<br />

R 5<br />

R 6<br />

R 4<br />

R 7<br />

R 1 = H, Ph<br />

B<br />

NH<br />

+<br />

NH<br />

NH<br />

+<br />

NH<br />

NH<br />

+<br />

NH<br />

R 1<br />

HN<br />

+<br />

HN<br />

CO2H<br />

HN<br />

+<br />

HN<br />

B<br />

A<br />

2Br −<br />

CO 2Me<br />

OAc<br />

[O], heat<br />

I2, 1,2-Cl2C6H4 heat<br />

38%<br />

CO2Me<br />

278 279<br />

280<br />

HN<br />

+<br />

HN<br />

R 3<br />

R 8<br />

R 2<br />

CO2H<br />

CO 2H<br />

R 9<br />

2Cl −<br />

2X −<br />

B B<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

R 1 CHO<br />

R 5<br />

R 6<br />

A<br />

MeO 2C CO2Me<br />

277<br />

AcO<br />

NH N<br />

N HN<br />

R 4 R 3<br />

NH N<br />

N HN<br />

R 7 R 8<br />

281<br />

R 2<br />

R 1<br />

R 9<br />

A<br />

CO2Me<br />

3,8-Bis(2-acetoxyethyl)-13-[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin<br />

(279); Typical Procedure: [178]<br />

8,13-Bis(2-acetoxyethyl)-18-[2-(methoxycarbonyl)ethyl]-1,3,7,12,15-pentamethyl-a,c-biladiene-19-carboxylic<br />

acid dihydrochloride (278; 539 mg, 0.72 mmol) in 1,2-dichlorobenzene<br />

(130 mL) was treated with I 2 (1.11 g, 4.37 mmol) and then stirred under reflux for<br />

20 min. After cooling to rt, Et 3N (0.5 mL) was added dropwise and the mixture was applied<br />

to a flash chromatography column [alumina (Brockmann Grade V), hexanes (3 column<br />

volumes) then 5% THF in CH 2Cl 2]. A second purification using flash column chromatogra-<br />

OAc


17.8.2 Reduced Porphyrins 1161<br />

aphy was carried out (silica gel, 5% THF in CH 2Cl 2) and the red eluates were collected and<br />

concentrated to dryness. The residue was crystallized (CH 2Cl 2/cyclohexane) to give 279;<br />

yield: 150 mg (38%); mp 211–2148C.<br />

17.8.2 Product Subclass 2:<br />

Reduced Porphyrins<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, p 600; further reactions: Vol. E 9d, pp 597–613.<br />

17.8.2.1 Chlorins (b,b¢-Dihydroporphyrins)<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, p 614.<br />

Chlorins, such as 12 (Scheme 4), with adjacent b,b¢-hydrogen atoms on a particular<br />

pyrrole subunit, are fairly sensitive to re-oxidation to porphyrin under conditions used<br />

for ring synthesis of porphyrins by most, if not all, of the methodologies discussed above.<br />

Thus, unless the b,b¢-bond is blocked against re-oxidation, for example with geminal<br />

methyl substituents, most approaches seem doomed to failure. Indeed, there appeared<br />

until fairly recently to be no good reason to perform rational syntheses from pyrroles to<br />

give direct access to chlorins; instead, chlorins were readily accessible simply by reduction<br />

of the corresponding porphyrins or metal porphyrinates. That situation has changed<br />

with reports of several direct approaches to chlorins. Since it appeared to be steric factors<br />

that affected whether or not a porphyrin would be obtained from an attempted chlorin<br />

synthesis, most new developments have been achieved by paying particular attention to<br />

steric issues (but not ignoring electronic factors).<br />

17.8.2.1.1 Method 1:<br />

By Ring Synthesis<br />

A stepwise preparation of a chlorin from an open-chain tetrapyrrole has been reported<br />

(Scheme 66). [183] MacDonald-type [2 +2] condensation of a dipyrromethane 282 with a<br />

1,9-diformyldipyrromethane 283 gives the porphodimethene intermediate 284, from<br />

which chlorin 285 is obtained after metalation with zinc(II) and thermolysis. The porphodimethene<br />

is the normal MacDonald intermediate before oxidation to porphyrin, and<br />

this procedure is based upon the known equilibria between metalated meso-dihydroporphyrins<br />

and metallochlorins. [184–186]<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1162 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 66 Chlorin Ring Synthesis by Oxidation of a Porphodimethene Intermediate [183]<br />

MeO 2C<br />

HO 2C<br />

NH<br />

MeO 2C<br />

HN<br />

+<br />

OHC<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

Et<br />

282 283<br />

NH<br />

N<br />

N<br />

HN<br />

Et Et<br />

284<br />

HN<br />

Et<br />

CHO<br />

Zn(OAc)2, heat<br />

H +<br />

MeO 2C<br />

N N<br />

Zn<br />

N N<br />

Et Et<br />

Numerous chlorins which, because of the geminal substitution mentioned above, cannot<br />

adventitiously re-oxidize to porphyrins have been prepared. Chlorin 287 is prepared in a<br />

stepwise fashion (Scheme 67) from the open-chain precursor 286 (by treatment with 1,8diazabicyclo[5.4.0]undec-7-ene);<br />

[187,188] oxidation to the porphyrin level during the process<br />

is prevented by the presence of the geminal dimethyl group in ring A.<br />

Scheme 67 Chlorin Ring Synthesis by Cyclization of a Geminally Disubstituted Open-Chain<br />

Precursor [187,188]<br />

MeO2C<br />

NC<br />

Br<br />

A<br />

N N<br />

CN<br />

Zn<br />

N N<br />

286<br />

base<br />

MeO2C<br />

NC<br />

285<br />

N N<br />

Zn<br />

N N<br />

Rational syntheses of b-blocked chlorins have also been reported. For example, treatment<br />

of the reduced dipyrrole 290 with the 3-aryldipyrromethane 289 (obtained from 288 by<br />

reduction with sodium borohydride) affords the dihydrobilene a 291, which undergoes<br />

metal-mediated oxidative cyclization [with silver periodate and zinc(II) acetate] to give<br />

the meso-substituted zinc(II) chlorinate 292 (Scheme 68). [189,190]<br />

287


17.8.2 Reduced Porphyrins 1163<br />

aScheme 68 Chlorin Ring Synthesis by Metal-Mediated Oxidative Cyclization of a<br />

Dihydrobilene Precursor [189,190]<br />

O<br />

4-Tol<br />

4-Tol<br />

Br<br />

N<br />

HN<br />

N<br />

N<br />

288<br />

4-Tol<br />

Br<br />

291<br />

HN<br />

N<br />

I<br />

NaBH 4<br />

THF/MeOH<br />

I<br />

N N<br />

N N<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

HO<br />

Br<br />

AgIO 3<br />

N<br />

HN<br />

Zn(OAc) 2<br />

toluene<br />

piperidine<br />

289<br />

I<br />

4-Tol<br />

Zn<br />

290<br />

292 18%<br />

NH<br />

N<br />

, TFA, MeCN<br />

Some stepwise synthetic approaches to bonellin (a natural chlorin involved in sex differentiation,<br />

which is obtained from the echurian worm), as its dimethyl ester 297 have<br />

been described. [191,192] For example, [191] condensation of the two compounds 293 and 294<br />

gives the open-chain tetrapyrrole 295 (Scheme 69); the selectively differentiated cyanoethyl<br />

and (methoxycarbonyl)ethyl substituents are used in order to provide access to<br />

natural amino acid derivatives of bonellin. Photocyclization of 295 then affords the cyano<br />

derivative 296 from which bonellin dimethyl ester (297) is produced.<br />

Scheme 69 Chlorin Ring Synthesis by Photocyclization of an Open-Chain Tetrapyrrole<br />

Precursor [191]<br />

NH<br />

N<br />

293<br />

hν (580 nm)<br />

+<br />

OHC<br />

MeO<br />

N<br />

HN<br />

NC CO 2Me<br />

294<br />

NH N<br />

N HN<br />

NC CO2Me 296<br />

H +<br />

NC<br />

N<br />

N<br />

MeO<br />

295<br />

HN<br />

N<br />

NH N<br />

N HN<br />

MeO2C CO2Me 297<br />

I<br />

CO 2Me<br />

for references see p 1223


1164 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aIn connection with studies of the intermediates in vitamin B 12 biosynthesis, and identification<br />

of siroheme (298) and sirohydrochlorin (299), a number of similar geminally Cmethylated<br />

chlorins [e.g., faktor-I octamethyl ester (300)] have been synthesized (Scheme<br />

70).<br />

Scheme 70 Geminally C-Methylated Chlorins<br />

HO2C<br />

HO2C<br />

MeO 2C<br />

H<br />

CO 2H<br />

N N<br />

Fe<br />

N N<br />

CO 2H<br />

CO 2H<br />

HO2C CO 2H<br />

MeO 2C<br />

H<br />

298<br />

CO2Me<br />

NH N<br />

N HN<br />

MeO 2C CO 2Me<br />

300<br />

H<br />

CO2H<br />

CO2Me<br />

CO 2Me<br />

CO2Me<br />

N N<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

HO2C<br />

HO2C<br />

H<br />

CO 2H<br />

CO 2H<br />

CO 2H<br />

HO2C CO 2H<br />

[17,18-Dihydro-18,18-dimethyl-5-(4-tolyl)-12-(4-iodophenyl)chlorinato]zinc(II) (292); Typical<br />

Procedure: [190]<br />

To a soln of 1-bromo-3-(4-iodophenyl)-9-(4-methylbenzoyl)dipyrromethane (288; 110 mg,<br />

0.20 mmol) in anhyd THF/MeOH (4:1; 7.5 mL) at rt was added an excess of NaBH 4<br />

(100 mg, 2.6 mmol) in small portions. The reaction was monitored by TLC and upon completion<br />

was quenched with cold H 2O (10 mL) and extracted with CH 2Cl 2 (3 ” 25 mL). The<br />

combined organic phases were washed with brine (50 mL), dried (K 2CO 3) for 2–3 min,<br />

and concentrated at rt under reduced pressure to give 289 in CH 2Cl 2 (ca. 1–2 mL). 1,3,3-<br />

Trimethyldihydrodipyrromethene (290; 45 mg, 0.24 mmol) in anhyd MeCN (ca. 2–3 mL)<br />

was combined with 289 (from above) and additional dry MeCN was added to give a total<br />

volume of 22 mL. The soln was stirred at rt under argon and TFA (20 mL, 0.26 mmol, ca. 11<br />

mM) was added. The reaction was monitored by TLC (observing disappearance of 290),<br />

and the mixture was quenched with aq NaHCO 3 (10%) then extracted with CH 2Cl 2<br />

(3 ” 25 mL). The combined organic phases were washed with H 2O and brine, dried<br />

(Na 2SO 4), and concentrated under vacuum at rt to give 291. The residue was dissolved in<br />

anhyd toluene (14 mL) under argon and AgIO 3 (848 mg, 3.0 mmol), piperidine (300 mL,<br />

3.0 mmol), and Zn(OAc) 2 (550 mg, 3.0 mmol) were added. This mixture was heated at<br />

808C for 2.5 h, monitoring by TLC and spectrophotometry (new bands at about 410 and<br />

610 nm). The mixture was cooled and passed through a short column (silica gel, hexanes/CH<br />

2Cl 2); a major fraction was collected, concentrated, and rechromatographed under<br />

the same conditions. Concentration of the eluates gave a greenish-blue solid, which<br />

299<br />

H<br />

CO2H


17.8.2 Reduced Porphyrins 1165<br />

awas taken up in CH 2Cl 2 and precipitated by addition of hexanes to give 292 as a solid;<br />

yield: 25 mg (18%).<br />

17.8.2.1.2 Method 2:<br />

By Reduction of Porphyrins or Metal Porphyrinates<br />

Porphyrins can be reduced to provide a wide variety of dihydroporphyrins (chlorins 301,<br />

phlorins 302, and porphodimethenes 303), tetrahydroporphyrins (bacteriochlorins 304,<br />

isobacteriochlorins 305, and porphomethenes 306), and hexahydroporphyrins (pyrrocorphins<br />

307 and porphyrinogens 308) (Scheme 71).<br />

Scheme 71 Hydroporphyrin Types<br />

R 1<br />

R 8<br />

R 1<br />

R 8<br />

R 1<br />

R 8<br />

R 2<br />

R 7<br />

R 2<br />

R 7<br />

R 2<br />

R 7<br />

N<br />

NH<br />

N<br />

NH<br />

N<br />

NH<br />

301<br />

304<br />

307<br />

HN<br />

N<br />

HN<br />

N<br />

HN<br />

N<br />

R 3<br />

R 6<br />

R 3<br />

R 6<br />

R 3<br />

R 6<br />

R 4<br />

R 5<br />

R 4<br />

R 5<br />

R 4<br />

R 5<br />

R 1<br />

R 8<br />

R 1<br />

R 8<br />

R 1<br />

R 8<br />

trans-Chlorins, such as trans-310 are most often obtained by the reduction of iron(III) bsubstituted<br />

porphyrins 309 with sodium metal in isopentyl alcohol (Scheme 72). [193–195]<br />

When the porphyrin is unsymmetrically substituted at the b-positions, mixtures of chlorin<br />

regioisomers are formed in ratios that are dependent upon electronic and steric factors,<br />

as pyrrole subunits bearing electron-withdrawing groups and/or an adjacent substituted<br />

meso position are preferentially reduced. Reduction to cis-chlorins (e.g., cis-310) is<br />

usually accomplished using the classical diimide approach, [196–199] the diimide being obtained<br />

from 4-toluenesulfonylhydrazine (Scheme 72). Porphyrins can also be photoreduced<br />

in the presence of amines, ascorbic acid, and other compounds. [200–205]<br />

R 2<br />

R 7<br />

R 2<br />

R 7<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

NH<br />

N<br />

R 2<br />

R 7<br />

NH<br />

302<br />

305<br />

NH<br />

NH<br />

HN<br />

N<br />

N<br />

HN<br />

308<br />

HN<br />

HN<br />

R 3<br />

R 6<br />

R 3<br />

R 6<br />

R 3<br />

R 6<br />

R 4<br />

R 5<br />

R 4<br />

R 5<br />

R 4<br />

R 5<br />

R 1<br />

R 8<br />

R 1<br />

R 8<br />

R 2<br />

R 7<br />

R 2<br />

R 7<br />

N<br />

NH<br />

NH<br />

NH<br />

303<br />

306<br />

HN<br />

N<br />

N<br />

HN<br />

R 3<br />

R 6<br />

R 3<br />

R 6<br />

R 4<br />

R 5<br />

R 4<br />

R 5<br />

for references see p 1223


1166 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 72 Reductions of Porphyrins to Chlorins [193–199]<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

N Cl N<br />

Fe<br />

N N<br />

Et<br />

Et Et<br />

Et<br />

309<br />

NH N<br />

N HN<br />

Et<br />

Et Et<br />

21<br />

Et<br />

Et<br />

Et<br />

Et<br />

Na/iPr(CH2)2OH<br />

− Fe<br />

N2H2<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et Et<br />

trans-310<br />

NH N<br />

N HN<br />

Et<br />

Et Et<br />

cis-310<br />

Regioselective photoreductions to provide cis-chlorins have been reported, using ascorbic<br />

acid in the presence of diazabicyclo[2.2.2]octane (Scheme 73). [204–207] Photoreduction of<br />

zinc(II) methyl pyropheophorbide a (311) in 8–10% ethanol in pyridine gives initially the<br />

(vinyl)isobacteriochlorin 312, which undergoes allylic shift of the vinyl double bond to<br />

afford the E-(ethylidene)isobacteriochlorin 313 (Scheme 73). This pathway has been proposed<br />

as a chemical model for the reduction of the 8-vinyl to ethyl in chlorophyll biosynthesis,<br />

and also for the biological formation of bacteriochlorophyll a derivatives from the<br />

corresponding vinylporphyrin or vinylchlorin. [208,209] Photoreduction to give predominantly<br />

(vinyl)isobacteriochlorin 312 is achieved by using benzene as the solvent and a<br />

smaller amount of the diazabicyclo[2.2.2]octane catalyst.<br />

Et<br />

Et<br />

Et<br />

Et


17.8.2 Reduced Porphyrins 1167<br />

aScheme 73 Diazabicyclo[2.2.2]octane Reduction of a Vinylchlorin [204–207]<br />

MeO2C<br />

N N<br />

Zn<br />

N N<br />

311<br />

O<br />

Et<br />

ascorbic acid<br />

DABCO<br />

EtOH, benzene or py, hν<br />

DABCO<br />

MeO2C<br />

MeO2C<br />

N N<br />

N N<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Zn<br />

312<br />

N N<br />

Zn<br />

N N<br />

Reduction of methyl nickel(II) mesopyropheophorbide a (314) with Raney nickel under<br />

hydrogen gives isobacteriochlorin 315 (as a mixture of diastereomers), the 13 1 -deoxochlorin<br />

316, and a minor amount of hexahydroporphyrin 317 (Scheme 74). [210] The isobacteriochlorin<br />

315 is preferentially formed when acetone is used as solvent, but 13 1 -deoxochlorin<br />

316 is the major product when methanol is used as solvent.<br />

Scheme 74 Raney Nickel Reduction of Methyl Nickel(II) Mesopyropheophorbide a [210]<br />

MeO2C<br />

Et<br />

N N<br />

Ni<br />

N N<br />

314<br />

O<br />

Et<br />

H 2<br />

Raney Ni<br />

N N<br />

MeO2C<br />

+ Ni<br />

+<br />

MeO 2C<br />

Et<br />

N N<br />

316<br />

Et<br />

Et<br />

N N<br />

Ni<br />

N N<br />

315<br />

MeO2C<br />

Et<br />

O<br />

313<br />

Et<br />

N N<br />

Ni<br />

N N<br />

317<br />

O<br />

O<br />

O<br />

Et<br />

Et<br />

Et<br />

for references see p 1223


1168 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aRaney nickel reduction of the nickel(II) “anhydro”-rhodoporphyrin 318 produces six<br />

products: deoxo derivative 319, isobacteriochlorins 320 and 321, hexahydroporphyrins<br />

322 and 323 (as major products), and the octahydroporphyrin 324 (Scheme 75). [211]<br />

Scheme 75 Raney Nickel Reduction of Nickel(II) “Anhydro”-Rhodoporphyrin [211]<br />

Et<br />

Et<br />

N<br />

N<br />

N<br />

N<br />

Ni<br />

318<br />

Ni<br />

320<br />

N<br />

N<br />

O<br />

N<br />

N<br />

O<br />

Et<br />

CO 2Me<br />

Et<br />

CO2Me<br />

Raney Ni<br />

Et<br />

O<br />

321<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

N<br />

N<br />

N<br />

Et<br />

N<br />

N<br />

Et<br />

Ni<br />

319<br />

+ Ni<br />

+<br />

Et<br />

N<br />

323<br />

CO 2Me<br />

+ Ni<br />

+<br />

N<br />

N<br />

N<br />

O<br />

Et<br />

CO2Me<br />

N<br />

N<br />

Et<br />

CO 2Me<br />

Et<br />

Et<br />

N<br />

N<br />

Ni<br />

322<br />

N<br />

N<br />

+<br />

N<br />

N<br />

O<br />

Ni<br />

324<br />

Et<br />

CO 2Me<br />

N<br />

N<br />

O<br />

Et<br />

CO 2Me<br />

Dihydroporphyrins can also be prepared by intramolecular cyclizations, such as in Woodward<br />

s chlorophyll a synthesis which involves the acid-catalyzed transformation of porphyrin<br />

325 into the purpurin-type chlorin 326. Intramolecular cyclizations of meso-acrolein-<br />

and meso-acrylateporphyrins 327 (R 1 = CHO and CO 2Et, respectively), produce the<br />

corresponding trans-purpurins 328 (Scheme 76). [212–214] As in Woodward s case, formation<br />

of purpurins 328 occurs in refluxing acetic acid, although base-catalyzed cyclizations (using<br />

Et 3N or KOH/MeOH in CH 2Cl 2) have been employed for the preparation of 5,15-disubstituted<br />

purpurins. [215–217] In this case the cyclization reaction is regiospecific, giving only<br />

one purpurin product from unsymmetrically substituted porphyrins; for example, purpurin<br />

330 is the preferred product from the cyclization of the etioporphyrin I 5-ethyl<br />

acrylic ester (329). Bis-cyclization of the 5,15-bis[2-(ethoxycarbonyl)vinyl]porphyrin 331<br />

produces regiospecifically the air-sensitive bacteriochlorin 332, containing two purpurin<br />

rings (Scheme 76). [218]


17.8.2 Reduced Porphyrins 1169<br />

aScheme 76 Formation of Purpurins [212–214,218]<br />

AcHN<br />

MeO 2C<br />

Et<br />

Et<br />

Et<br />

NH<br />

N<br />

NH<br />

N<br />

325<br />

N<br />

HN<br />

CO2Me<br />

327<br />

N<br />

HN<br />

Et Et<br />

R 1 = CHO, CO2Et<br />

Et<br />

EtO2C<br />

Et<br />

NH<br />

N<br />

Et<br />

Et<br />

329<br />

Et<br />

N<br />

HN<br />

Et<br />

Et<br />

NH<br />

N<br />

Et<br />

CO 2Me<br />

Et<br />

Et<br />

Et<br />

N<br />

HN<br />

Et<br />

Et Et<br />

331<br />

R 1<br />

AcOH, heat<br />

AcOH, heat<br />

R 1 = CO2Et 68%<br />

CO2Et AcOH, heat<br />

Et<br />

Et<br />

CO 2Et<br />

MeO 2C<br />

AcOH, heat<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

AcHN<br />

Et<br />

Et<br />

EtO2C Et<br />

Et<br />

Et<br />

MeO2C<br />

Et<br />

Et<br />

Et<br />

NH<br />

N<br />

NH<br />

N<br />

326<br />

328<br />

N<br />

HN<br />

N<br />

HN<br />

Et Et<br />

NH<br />

N<br />

NH<br />

N<br />

330<br />

N<br />

HN<br />

N<br />

HN<br />

Et<br />

Et<br />

Et Et<br />

332<br />

Et<br />

CO2Me<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

R 1<br />

CO 2Et<br />

CO2Et<br />

for references see p 1223


1170 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aMontforts and co-workers [219–221] have shown that hematoporphyrins [i.e., (1-hydroxyethyl)porphyrins],<br />

such as 333, react with N,N-dimethylacetamide dimethyl acetal at high<br />

temperatures, to give the corresponding (Z)-ethylidenechlorins (e.g., 334) (Scheme 77).<br />

When subjected to catalytic hydrogenation, these chlorins yield the corresponding alkylchlorins.<br />

[222] Oxidative cleavage of the exocyclic double bond in the zinc(II) or nickel(II)<br />

complexes of chlorin 334 gives the oxochlorin 335 after demetalation; oxochlorin 335<br />

can be hydrolyzed to the ester, and then stereoselectively reduced with lithium tri-tert-butoxyaluminohydride<br />

to give hydroxychlorin 336. [223,224] The corresponding dioxoisobacteriochlorins<br />

have also been prepared from the hematoporphyrin IX dimethyl ester 337<br />

using the same methodology. [225,226]<br />

In an interesting intramolecular cyclization to give chlorins, treatment of [(2-acetylamino)ethyl]porphyrins<br />

(e.g., 338) with phosphoryl chloride in pyridine results in the intramolecular<br />

cyclization of the Vilsmeier-complex substituent with formation of spirochlorin<br />

339 (Scheme 77). [227,228]<br />

Scheme 77 Intramolecular Cyclizations To Give Chlorins [219–228]<br />

HO<br />

NH<br />

N<br />

MeO2C CO2Me 333<br />

N<br />

HN<br />

1. metalation (Zn or Ni)<br />

2. oxidative cleavage<br />

3. demetalation<br />

O<br />

Me2N<br />

MeO OMe<br />

xylene, 160 o C<br />

MeO2C CO2Me 335<br />

MeO2C CO2Me 334<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

O<br />

N<br />

NMe 2<br />

NH<br />

N<br />

HN<br />

O<br />

Me 2N<br />

MeO 2C<br />

HO<br />

N<br />

NH<br />

N<br />

HN<br />

1. hydrolysis<br />

2. reduction<br />

N<br />

NH<br />

MeO2C CO2Me 336<br />

N<br />

HN


17.8.2 Reduced Porphyrins 1171<br />

HO<br />

NH<br />

MeO2C CO2Me a337<br />

Et<br />

NH<br />

N<br />

338<br />

N<br />

N<br />

HN<br />

N HN<br />

Et Et<br />

OH<br />

NHAc<br />

POCl 3, py<br />

MeO 2C<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

O<br />

NH<br />

N<br />

N<br />

HN<br />

CO2Me<br />

MeO2C CO 2Me<br />

Et<br />

NH<br />

339<br />

N<br />

O<br />

N HN<br />

Et Et<br />

Metal porphyrinates are easily reduced to the p-dianions with reductants such as sodium<br />

in tetrahydrofuran or with sodium anthracenide. [184,229–231] Protonation and alkylation of<br />

porphyrin p-anions usually occur at the 5-, 10-, 15-, or 20-positions, resulting inthe formation<br />

of phlorins and porphodimethenes, which can rearrange to chlorins or be oxidized<br />

to give porphyrins.<br />

Ethyl 2,3,7,8,12,13,17,18-Octaethylpurpurin-3 1 -carboxylate (328,R 1 =CO 2Et); Typical Procedure:<br />

[213]<br />

5-[2-(Ethoxycarbonyl)vinyl]-2,3,7,8,12,13,17,18-octaethylporphyrin (327, R 1 =CO 2Et;<br />

100 mg, 0.16 mmol) in glacial AcOH (20 mL) was refluxed under N 2 for 24 h. After cooling<br />

to rt, the solvent was removed under vacuum and replaced with CH 2Cl 2. Chromatography<br />

(silica gel, CH 2Cl 2) gave a green major fraction, which was concentrated to dryness. The<br />

residue was crystallized (CH 2Cl 2/MeOH) to give purple microcrystals of 328 (R 1 =CO 2Et);<br />

yield: 68 mg (68%); mp 135–1388C.<br />

17.8.2.1.3 Method 3:<br />

By Oxidation of Porphyrins or Metal Porphyrinates<br />

Strangely enough, the chlorin chromophore can be obtained by oxidation, as well as the<br />

more traditional reduction approach. Thus, osmiumtetroxide reacts selectively at the b—<br />

b¢ double bonds of porphyrins [e.g., 2,3,7,8,12,13,17,18-octaethylporphyrin (21)] to produce<br />

cis-dihydroxychlorins (e.g., 340). [232–234] Such b,b¢-disubstituted cis-chlorins readily<br />

undergo pinacol–pinacolone rearrangement in perchloric acid or fuming sulfuric acid to<br />

give b-oxochlorins (e.g., 341). The regioselectivity in the osmiumtetroxide oxidation reaction<br />

and the migratory aptitudes of different peripheral substituents in the subsequent<br />

pinacol–pinacolone rearrangement have been studied extensively. [235–238] The use of hydrogen<br />

peroxide in concentrated sulfuric acid also achieves direct oxidation of porphyrins<br />

to oxochlorins, but mixtures of mono-, di- and trioxo derivatives are often obtained.<br />

[239,240] Benzoyl peroxide reacts with 5,10,15,20-tetraphenylporphyrin (342, R 1 =H)<br />

to produce a b-benzoyloxyporphyrin, which can be converted into 343 by metalation, hydrolysis,<br />

and demetalation (Scheme 78). [241] Porphyrin 343 is, however, obtained in higher<br />

yield from 2-nitro-5,10,15,20-tetraphenylporphyrin 342 (R 1 =NO 2) by nucleophilic dis-<br />

N<br />

for references see p 1223


1172 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aplacement of the nitro group with the sodium salt of (E)-benzaldoxime. [242] Oxidation of<br />

343 to afford the 2,3-dioneporphyrin 344 can be accomplished under a variety of oxidative<br />

conditions [chromium(VI) oxide/acetic acid, [243] photooxidation or selenium dioxide/<br />

dioxane] (Scheme 78). [242]<br />

Scheme 78 Oxidations of Porphyrins To Give Chlorins [232–234,241,242]<br />

Et Et<br />

NH N<br />

Et<br />

Ph<br />

Et<br />

N<br />

21<br />

HN<br />

Et<br />

Et Et<br />

NH<br />

N<br />

Ph<br />

Ph<br />

N<br />

HN<br />

342 R 1 = H, NO 2<br />

Et<br />

R 1<br />

Ph<br />

OsO4, CH2Cl2, py<br />

56%<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

A: 70% HClO4, CH2Cl2<br />

B: H2SO4<br />

A: 72%<br />

B: 72%<br />

1. BzOOH, PhCl, 100 oC 2. Zn(OAc)2<br />

3. NaOH, THF, MeOH<br />

4. H +<br />

CrO 3, AcOH<br />

Et<br />

NH<br />

N<br />

340<br />

N<br />

HN<br />

Et<br />

Et Et<br />

Et<br />

Et<br />

Et<br />

OH<br />

OH<br />

Et<br />

NH<br />

N<br />

Et<br />

341<br />

N<br />

HN<br />

O<br />

Et Et<br />

2,3,7,8,12,13,17,18-Octaethyl-cis-2,3-dihydroxy-22H,24H-chlorin (340); Typical<br />

Procedure: [233]<br />

2,3,7,8,12,13,17,18-Octaethylporphyrin (21; 700 mg, 1.31 mmol) in CH 2Cl 2 (120 mL) and<br />

pyridine (0.5 mL) was stirred with OsO 4 (545 mg, 2.14 mmol) under N 2 in the dark for<br />

90 h. The solvent was removed under reduced pressure and the residue was dissolved in<br />

MeOH (100 mL) and CH 2Cl 2 (30 mL) and H 2S was passed through the soln for 30 min. The<br />

Ph<br />

Ph<br />

NH<br />

N<br />

Ph<br />

Ph<br />

343<br />

N<br />

HN<br />

NH<br />

N<br />

Ph<br />

OH<br />

Ph<br />

344<br />

N<br />

HN<br />

Ph<br />

O<br />

Et<br />

Et<br />

Et<br />

O<br />

Ph


17.8.2 Reduced Porphyrins 1173<br />

aprecipitate was filtered off (glass sinter) and the pad was washed with CHCl 3. The total filtrate<br />

was concentrated to dryness, the residue dissolved in CHCl 3 (5 mL), and MeOH<br />

(50 mL) was added to precipitate unchanged 21. The soln was filtered, the pad was washed<br />

with MeOH (100 mL) and the total filtrate was concentrated to dryness. The residue was<br />

chromatographed [silica gel (300 ” 25 mm diameter), CHCl 3/acetone 98:2]. After elution<br />

of unchanged 21 (total recovery 20%), the major band afforded diol 340; yield: 420 mg,<br />

(56%); mp 2138C (dec). The product was recrystallized (CH 2Cl 2/hexanes); mp 219–2208C<br />

(Et 2O); 218 8C (dec) (MeOH/CHCl 3).<br />

3,3,7,8,12,13,17,18-Octaethyl-2-oxochlorin (341); Typical Procedure: [233]<br />

Method A: Using perchloric acid: 2,3,7,8,12,13,17,18-Octaethyl-cis-2,3-dihydroxy-22H,24Hchlorin<br />

(340; 100 mg, 0.18 mmol) in CH 2Cl 2 (60 mL) was stirred with 70% HClO 4 (1 mL) for<br />

30 min at 208C. The mixture was washed in turn with H 2O, aq NaHCO 3, and H 2O. The organic<br />

layer was separated, dried (Na 2SO 4) and concentrated to dryness to give a residue<br />

which was then purified by chromatography [silica gel (250 ” 35 mm), CH 2Cl 2]. The faster-moving<br />

band was discarded and the major band was worked up to give 341; yield:<br />

71 mg (72%); mp 247–2508C.<br />

Method B: Using sulfuric acid: Concd H 2SO 4 (2.5 mL) and fuming H 2SO 4 (2 mL; 20% SO 3)<br />

were added to 2,3,7,8,12,13,17,18-octaethyl-cis-2,3-dihydroxy-22H,24H-chlorin (340;<br />

100 mg, 0.18 mmol) and the resulting green soln was kept at rt for 7 min, then poured<br />

into a large excess of ice and extracted with CHCl 3. The organic layer was washed with<br />

H 2O, dried (Na 2SO 4), and concentrated to dryness to give a residue, which was crystallized<br />

from CHCl 3/hexanes to give 341; yield: 71 mg (72%); mp 247–2508C.<br />

17.8.2.2 Bacteriochlorins and Isobacteriochlorins (b,b¢,b¢¢,b¢¢¢-Tetrahydroporphyrins)<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, pp 636 and 644.<br />

Bacteriochlorins 13 are tetrapyrroles in which opposite pyrrole subunits are reduced,<br />

whereas isobacteriochlorins 14 have adjacent pyrrole subunits modified (see<br />

Scheme 4). Until fairly recently, with the discovery of siroheme (298) and of precorrin intermediates<br />

on the biosynthetic pathway to vitamin B 12, and of heme d (345) (Scheme 79),<br />

the only naturally occurring tetrahydroporphyrin-type macrocycles were thought to be<br />

the bacteriochlorins found as bacteriochlorophyll a and bacteriochlorophyll b (and their<br />

demetalated/deesterified bacteriopheophorbide derivatives).<br />

Tetrahydroporphyrins are obtained by further reduction of chlorins under the conditions<br />

described for the preparation of chlorins (i.e., sodium in alcohols, or diimide); isobacteriochlorins<br />

(e.g., 346) are formed preferentially by reduction of metal porphyrinates,<br />

whereas bacteriochlorins (e.g., 347) are usually produced from metal-free porphyrins.<br />

Likewise, in the oxidation protocol for removal of a peripheral double bond from the<br />

porphyrin chromophore, metal-free oxochlorins 348 (M = 2 H) react further and regioselectively<br />

at the opposite pyrrole ring with osmiumtetroxide (see above) to give diol 349,<br />

whereas the metal derivatives 348 (M = Zn) react preferentially at an adjacent b—b¢ double<br />

bond to give diol 350 (Scheme 79). [244,245] Thus, the oxidative methodology has been<br />

extended for the preparation of vic-dihydroxy- and oxobacteriochlorins and isobacteriochlorins.<br />

[246–248]<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1174 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 79 Typical Structures and Syntheses of Bacteriochlorins and<br />

Isobacteriochlorins [244–248]<br />

HO2C<br />

Ph<br />

Et<br />

Et<br />

O<br />

N<br />

N<br />

Fe<br />

HO2C CO2H<br />

345<br />

NH<br />

N<br />

Ph<br />

Ph<br />

347<br />

N<br />

HN<br />

Et O<br />

N<br />

N<br />

M<br />

N<br />

N<br />

Et Et<br />

348 M = 2H, Zn<br />

N<br />

N<br />

Et<br />

Ph<br />

Et<br />

Et<br />

CO 2H<br />

O<br />

OsO4<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

M = 2H<br />

M = Zn<br />

− Zn<br />

Ph<br />

N<br />

N<br />

Et<br />

Ph<br />

Zn<br />

Ph<br />

346<br />

N<br />

N<br />

NH<br />

349<br />

Ph<br />

Et O<br />

Et<br />

HO<br />

N<br />

HO<br />

Et<br />

HO<br />

Et<br />

Et<br />

HO Et<br />

N<br />

NH<br />

350<br />

N<br />

HN<br />

N<br />

HN<br />

Et<br />

O<br />

Et Et<br />

Geminally C-methylated isobacteriochlorins have been synthesized in connection with<br />

the characterization of intermediates on the vitamin B 12 biosynthetic pathway. [249–256]<br />

17.8.2.3 Syntheses of Benzoporphyrins<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9b, pp 607, 621, 623 and Vol. E 9d, pp 717–833 (phthalocyanines).<br />

Benzoporphyrins have b,b¢-pyrrole-fused benzene rings and some have been found,<br />

in trace amounts, in various oil shales and petroleums. The class comprises the classical<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et


17.8.2 Reduced Porphyrins 1175<br />

atetrabenzoporphyrins (e.g., 351, 352), as well as monobenzoporphyrins (e.g., 353–355),<br />

dibenzoporphyrins (e.g., 356–358) and the so-called benzoporphyrin derivatives (e.g.,<br />

359) (Scheme 80).<br />

Scheme 80 Benzoporphyrin Types<br />

Et<br />

Et<br />

MeO 2C<br />

NH<br />

N<br />

NH<br />

N<br />

351<br />

354<br />

N<br />

HN<br />

N<br />

HN<br />

Et Et<br />

356<br />

MeO 2C<br />

NH<br />

N<br />

Ph<br />

N<br />

HN<br />

NH<br />

N<br />

MeO2C Ph<br />

Ph<br />

358<br />

N<br />

HN<br />

17.8.2.3.1 Method 1:<br />

Tetrabenzoporphyrins<br />

Ph<br />

CO 2Me<br />

Ph<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

N<br />

Ph<br />

Ph<br />

352<br />

MeO2C<br />

MeO2C<br />

N<br />

HN<br />

Ph<br />

N<br />

NH<br />

HN<br />

NH<br />

N<br />

N<br />

HN<br />

Et Et<br />

353<br />

MeO2C CO2Me 355<br />

Et<br />

NH<br />

N<br />

357<br />

N<br />

HN<br />

MeO 2C<br />

MeO2C<br />

Et<br />

R 1 O2C<br />

N<br />

N<br />

NH<br />

HN<br />

N<br />

359 R 1 = R 2 = H, Me<br />

CO2R 2<br />

Compounds of the type 351 (Scheme 80) are porphyrin analogues of the phthalocyanines<br />

360 (Scheme 81) and are prepared by cyclotetramerization of phthalimidine derivatives,<br />

isoindoles, or certain pyrrolic intermediates. Derivatives of tetrabenzoporphyrin contain-<br />

for references see p 1223


1176 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aing one, two, or three meso-nitrogen bridging atoms have also been synthesized, using<br />

acetophenone derivatives and metal salts. [257–262]<br />

When phthalimidine derivatives, such as methylphthalimidine, 3-carboxymethylphthalimidine,<br />

or potassium phthalimide, are heated at 350–4008C in the presence of metals,<br />

such as iron, zinc, magnesium, cadmium, or metallic acetates, the corresponding<br />

metal tetrabenzoporphyrinates are obtained in yields of up to 26%. [261–271] The metal tetrabenzoporphyrinates<br />

isolated using this method are usually very impure and require extensive<br />

purification. A modification of this synthesis involves the use of acetophenone-2carboxylic<br />

acid and aqueous ammonia in the presence of iron or zinc acetate and molecular<br />

sieves, and heating at 4008C; this yields the zinc(II) complex of 361 in 17% yield. [272–<br />

274] The free-base tetrabenzoporphyrin 361 is then obtained in good yield after demetalation<br />

using acid.<br />

Scheme 81 Typical Phthalocyanine and Tetrabenzoporphyrin Structures<br />

NH<br />

N<br />

N<br />

360<br />

N<br />

N N<br />

N<br />

HN<br />

Isoindoles are fairly unstable intermediates for use in tetrabenzoporphyrin syntheses.<br />

Though the tetrabenzoporphyrin 361 is symmetrical, its synthesis by tetramerization of<br />

isoindoles is still low-yielding. Nevertheless, several metal complexes of octamethyltetrabenzoporphyrin<br />

361 have been prepared by pyrolysis (at 4008C) of 1,3,4,7-tetramethylisoindole<br />

in the presence of metals or metal salts, under an inert atmosphere, or by refluxing<br />

in a high boiling solvent (1,2,4-trichlorobenzene or 1-chloronaphthalene) in the<br />

presence of a metal salt. [275–277] A recently introduced improvement involves the<br />

cyclotetramerization of 1,3-dimethyl- and 1,3,4,7-tetramethylisoindoles in the presence<br />

of metals or their acetate salts; this leads to the metal complexes of 351 and 361 in yields<br />

as high as 81%. [278,279] As might be expected, use of unsymmetrically substituted isoindoles<br />

gives the four possible metal tetrabenzoporphyrinate isomers.<br />

The tetrabenzoporphyrin 351 has been prepared by a variation of the Rothemund<br />

methodology (see Section 17.8.1.2.1), using isoindole and formaldehyde in the presence<br />

of a metal or a metal salt. [280] High temperatures (3758C) are required for this reaction,<br />

but a 53% yield is reported for the preparation of the zinc(II) complex of 351. If the reaction<br />

is performed in the absence of a metal or a salt, very low yields of product are obtained,<br />

so compound 351 is generally obtained by demetalation of its zinc(II) complex.<br />

Condensation reactions between isoindole and benzaldehyde in the presence of a metal<br />

or metal salt, or of 3-benzylidenephthalimidine in the presence of zinc acetate, at high<br />

temperatures, afford a mixture of metal 5,10,15,20-substituted tetrabenzoporphyrinates;<br />

the major product is usually the metal complex of tetraphenyltetrabenzoporphyrin<br />

352. [281–286] Hexadecafluoro derivatives of tetrabenzoporphyrins 351 and 352 are also prepared<br />

using the same approach. [282,283] Pure 5,10,15,20-tetraphenyltetrabenzoporphyrins<br />

are prepared by tetramerization of 3-benzylidenephthalimidine, in the presence of zinc<br />

benzoate, [285] and also by condensation of potassium phthalimide with phenylacetic acid<br />

in the presence of zinc chloride. [270,287] The instability of isoindoles mentioned above can<br />

be circumvented by the use of a 4,7-dihydro-4,7-ethano-2H-isoindole ethyl ester (362;<br />

Scheme 82); reduction (lithium aluminum hydride) and tetramerization, followed by oxi-<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

N<br />

361<br />

N<br />

HN


17.8.2 Reduced Porphyrins 1177<br />

adation, affords the porphyrin 363 which, upon heating at 2008C (retro-Diels–Alder reaction),<br />

gives a quantitative yield of the tetrabenzoporphyrin 351. [288,289]<br />

Scheme 82 Preparation of a Tetrabenzoporphyrin [288,289]<br />

N<br />

H<br />

362<br />

CO 2Et<br />

1. LiAlH 4<br />

2. H +<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

3. [O] 200 o C<br />

Butanoporphyrins, such as 364, can be transformed into the corresponding tetrabenzoporphyrins<br />

365 by metalation (e.g., with nickel, copper, or zinc) followed by treatment<br />

with 2,3-dichloro-5,6-dicyanobenzo-1,4-quinone (Scheme 83). [290] Tetrabenzoporphyrin<br />

351 has been prepared in good yield by cyclotetramerization of a phenylsulfonylcyclohexanepyrrole-2-carboxylate,<br />

followed by base-promoted elimination of the phenylsulfonyl<br />

groups and aromatization using 2,3-dichloro-5,6-dicyanobenzo-1,4-quinone. [291]<br />

N<br />

363<br />

N<br />

HN<br />

NH<br />

N<br />

351<br />

N<br />

HN<br />

for references see p 1223


1178 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 83 Preparation of a Tetrabenzoporphyrin from a Butanoporphyrin [290]<br />

MeO2C<br />

MeO 2C<br />

Ar 1<br />

NH<br />

N<br />

Ar 1<br />

364<br />

N<br />

HN<br />

CO 2Me<br />

Ar 1<br />

CO2Me<br />

Ar1 MeO2C<br />

CO2Me MeO2C CO2Me K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

1. M 2+<br />

MeO 2C<br />

2. DDQ, heat<br />

MeO 2C<br />

Ar 1<br />

NH<br />

N<br />

Ar 1<br />

365<br />

N<br />

HN<br />

CO 2Me<br />

Ar 1<br />

CO 2Me<br />

Ar1 MeO2C<br />

CO2Me MeO2C CO2Me 17.8.2.3.2 Method 2:<br />

Monobenzoporphyrins, Dibenzoporphyrins, and Benzoporphyrin<br />

Derivatives<br />

The main interest in these macrocycles with individual benzo-fused rings comes from<br />

their discovery in some petroleum and related deposits, although their origins in these<br />

deposits are still poorly understood. [292–297] In 1977, the first synthesis of a monobenzoporphyrin<br />

353 (Scheme 80) from a porphyrin precursor bearing a fused cyclohexanone ring<br />

was reported. [298] Sodium borohydride reduction of the ketone group and subsequent dehydration<br />

with benzoyl chloride/dimethylformamide gives the cyclohexene derivative,<br />

which is then oxidized with 2,3-dichloro-5,6-dicyanobenzo-1,4-quinone to give 353. Using<br />

similar methodology, the same group prepared a naturally occurring monobenzoporphyrin<br />

354 containing an isocyclic cyclopentane ring, which they obtained by acid-catalyzed<br />

cyclization of the corresponding b-vinylmonobenzoporphyrin. [299]<br />

The so-called adj- (adjacent) and opp- (opposite) dibenzoporphyrins 356 and 357<br />

(Scheme 80) are also prepared following the same methodology as above; [300] the fused cyclohexanone<br />

rings of the precursors are prepared by base-induced cyclization of acetic<br />

and propanoic acid side chains on adjacent b-pyrrolic positions, followed by hydrolysis<br />

of the resulting b-oxo esters with aqueous acid.<br />

Monobenzoporphyrins such as 353 can also be synthesized from the appropriate tetrahydrobenzoporphyrins<br />

by oxidation with 2,3-dichloro-5,6-dicyanobenzo-1,4-quinone.<br />

[301,302] opp-Dibenzoporphyrins, such as 357, can also be obtained from the corresponding<br />

saturated precursors by treatment with 2,3-dichloro-5,6-dicyanobenzo-1,4-quinone<br />

in refluxing toluene, [116] provided that the precursor porphyrin has been previously<br />

chelated with zinc(II). [303]<br />

Fischer-type self-condensation of benzopyrromethene hydrobromides (179, see Section<br />

17.8.1.2.2.1), prepared from haloformylisoindoles and 2-unsubstituted pyrroles,


17.8.2 Reduced Porphyrins 1179<br />

agives good yields of opp-dibenzoporphyrins, such as 180 (Scheme 39). [123,304] The starting<br />

haloformylisoindoles are prepared from phthalimidine, by a double Vilsmeier reaction.<br />

Johnson and co-workers [305] were the first to show that the external vinyl groups and<br />

the neighboring peripheral b,b¢-double bonds of protoporphyrin IX dimethyl ester (366)<br />

(Scheme 84) react with activated dienophiles in Diels–Alder-type reactions, producing<br />

chlorins and isobacteriochlorins. This new approach to superstructured porphyrin systems<br />

was extended by others and has been applied to numerous b-vinyl-substituted porphyrins<br />

and chlorins to furnish chlorins, bacteriochlorins, isobacteriochlorins, and<br />

monobenzoporphyrins. [306–312] Monobenzoporphyrins, such as 355, can be obtained from<br />

Diels–Alder [4 +2]-cycloaddition reactions on the dimethyl ester 366 of protoporphyrin<br />

IX, followed by elimination of the angular methyl group in the presence of base, or simply<br />

by oxidation with benzo-1,4-quinone. [313] Catalytic hydrogenation of the Diels–Alder adduct<br />

obtained from the reaction with b-phenylsulfonylacrylonitrile, followed by base-promoted<br />

elimination of the phenylsulfonyl group, affords the corresponding cyclohexene<br />

derivative, which after aerobic oxidation and aromatization, produces the monobenzoporphyrin.<br />

[314] Monobenzoporphyrins are also obtained directly from Diels–Alder cycloaddition<br />

reactions of b-unsubstituted b¢-monovinylporphyrins in the presence of an excess<br />

of dienophile. [314]<br />

The so-called “benzoporphyrin derivatives” [e.g., benzoporphyrin derivative monocarboxylic<br />

acid, “BPD-MA”, Visudyne (359)] (Scheme 80) have been shown to be useful in<br />

the photodynamic therapy treatment of the wet form of age-related macular degeneration.<br />

They are obtained by rearrangement of the Diels–Alder adducts from protoporphyrin<br />

IX dimethyl ester (366) and dimethyl acetylenedicarboxylate, at room temperature, in<br />

the presence of triethylamine (to produce the trans-isomer) or 1,8-diazabicyclo[5.4.0]undec-7-ene<br />

(to produce the cis-isomer). [315–319] Benzobacteriopurins, for example 367, [320]<br />

and benzobacteriochlorins such as 368, [315,321,322] have also been synthesized using the<br />

above methodology (Scheme 84).<br />

Scheme 84 Benzoporphyrin Derivatives<br />

MeO2C<br />

EtO 2C<br />

Et<br />

NH<br />

N<br />

NH<br />

N<br />

366<br />

CO 2Et<br />

N<br />

HN<br />

EtO2C<br />

368<br />

N<br />

HN<br />

Et<br />

CO 2Me<br />

CO2Et<br />

MeO 2C<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

NH<br />

N<br />

EtO2C<br />

367<br />

N<br />

HN<br />

O O O<br />

CO 2Et<br />

for references see p 1223


1180 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a17.8.3 Product Subclass 3:<br />

Isomeric, Contracted, and Expanded Porphyrin Systems<br />

This rapidly growing area has been extensively reviewed in the recent past. [323,324]<br />

Though the porphyrin ligand appears to be the one which dominates biology (being<br />

utilized in heme proteins and in chlorophyll systems), many scientific disciplines (e.g.,<br />

chemistry, medicine, and materials science) have become interested in modification of<br />

the tetrapyrrole ligand system. Therefore, a large number of related macrocycles have<br />

been synthesized in the last few decades in order to study structure/function relationships,<br />

and possibly to improve upon some characteristics of the natural tetrapyrrole chromophore.<br />

Examples include porphyrin isomers; if the porphyrin macrocycle is regarded<br />

as the parent [18]porphyrin (1:1:1:1) system, where the numerals refer to the number of<br />

carbons between each pyrrole subunit, then isomers might have the (2,0,2,0) 369,<br />

(2,1,0,1) 370, (1,0,3,0) 371, etc., arrangement of meso-carbons (Scheme 85). The bestknown<br />

example of a contracted porphyrin system is found in corrole 372, and two commonly<br />

prepared expanded porphyrin systems (among many) are the sapphyrins 373 and<br />

pentaphyrins 374. The synthetic work, summarized below, has indeed been rewarded<br />

with new porphyrin analogues possessing novel and often unique chemical, physical, biological,<br />

and spectroscopic properties.<br />

Scheme 85 Isomeric, Contracted, and Expanded Porphyrin Types<br />

NH N<br />

N HN<br />

369<br />

NH N<br />

NH HN<br />

372<br />

N<br />

H<br />

NH N<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

N HN N HN<br />

N<br />

H<br />

370 371<br />

N<br />

H<br />

N N<br />

17.8.3.1 Syntheses of [18]Porphyrin (1,1,1,1) Isomers<br />

H<br />

N<br />

373<br />

17.8.3.1.1 Method 1:<br />

Porphycene {[18]Porphyrin (2,0,2,0)}<br />

NH N<br />

N HN<br />

Porphycene was the first porphyrin isomer to be synthesized, and has enjoyed a great<br />

deal of celebrity because of its novel properties and ingenious synthesis as part of Vogel s<br />

investigations of annulene systems. Exposure of 5,5¢-diformyl-2,2¢-bipyrrole (375) to the<br />

reductive McMurry conditions [titanium(IV) chloride/zinc–copper] followed by autoxidation<br />

of the presumed intermediate 376 gives porphycene 369 (Scheme 86). [325] Since this<br />

first synthesis, a large number of peripherally substituted porphycenes have been synthesized.<br />

[326–330]<br />

N<br />

374


17.8.3 Isomeric, Contracted, and Expanded Porphyrin Systems 1181<br />

aScheme 86 Synthesis of Porphycene [325]<br />

2<br />

OHC<br />

N<br />

H<br />

375<br />

N<br />

H<br />

CHO<br />

TiCl4 NH HN<br />

Zn/Cu [O]<br />

NH HN<br />

17.8.3.1.2 Method 2:<br />

Corrphycene (Porphycerin) {[18]Porphyrin (2,1,0,1)}<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

376<br />

NH N<br />

N HN<br />

One other example of a porphyrin isomer synthesis will be discussed for completeness.<br />

This is the [2,1,0,1] isomer (named corrphycene [331] or porphycerin [332] ), and it has received<br />

attention from a number of groups. [331–333] The corrphycene 379 is obtained from the 1,18diformyltetrapyrrole<br />

377 using the low-valent titanium (McMurry) methodology to give<br />

378. [331] This is then oxidized with air or with iron(III) chloride to give 379 in low yield<br />

(Scheme 87).<br />

Scheme 87 Synthesis of Corrphycene (Porphycerin) [331]<br />

Et<br />

Et Et<br />

N<br />

H<br />

377<br />

N<br />

H<br />

NH HN<br />

CHO OHC<br />

Et<br />

TiCl 4<br />

Zn/Cu<br />

FeCl 3<br />

Et<br />

N<br />

H<br />

N<br />

H<br />

369<br />

Et NH HN Et<br />

Et<br />

378<br />

N<br />

H<br />

Et<br />

Et N HN Et<br />

379<br />

N<br />

Et<br />

for references see p 1223


1182 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a17.8.3.2 Contracted Porphyrins:<br />

Syntheses of Corroles<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, pp 665–672, 684 ff.<br />

Corroles 372 are analogues of porphyrins that are fully conjugated but lack one of<br />

the four meso-carbon atoms present in porphyrins. Because of this, metal-free corroles<br />

possess three NHs and, with their smaller-sized central core, can accommodate small<br />

metal ions or metal ions with high oxidation states. [334] The general and coordination<br />

chemistry of corroles, [175,176,323] and the various methods available for their synthesis, [335]<br />

have been reviewed recently.<br />

17.8.3.2.1 Method 1:<br />

From Monopyrroles<br />

17.8.3.2.1.1 Variation 1:<br />

Using 2-Substituted 1H-Pyrroles<br />

Treatment of the 5-[hydroxy(phenyl)methyl]-1H-pyrrole-2-carboxylic acid 380 with acidic<br />

ethanol, under conditions that would normally be expected to give the corresponding<br />

porphyrin (see Section 17.8.1.2.1), in the presence of cobalt(II) acetate and triphenylphosphine<br />

affords a 25% yield of the corresponding corrole cobalt complex 382, [336] possibly<br />

via the dipyrromethene 381 (Scheme 88). [337] If, for example, the copper(II) or nickel(II)<br />

salts are used, the corresponding metal octamethyltetraphenylporphyrinate is obtained.<br />

The cobalt pathway also yields corroles with a variety of substituents on the aryl rings if<br />

the appropriate pyrroles are used. Use of the 2-formyl-1H-pyrrole 27 under the same acidic<br />

conditions also results in the formation of corrole if cobalt(II) ions are used, but in this<br />

case a mixture of a porphyrin 383 and three corroles 384–386 are obtained (Scheme<br />

88); [338] it is again postulated that the corrole mixture is produced as a result of the intermediacy<br />

of dipyrromethenes. When other metals are used in the reaction, no corrole (and<br />

usually only porphyrin) is obtained.<br />

Scheme 88 Syntheses of Corroles from Monopyrroles [336–338]<br />

HO2C<br />

N<br />

H<br />

Ph<br />

OH<br />

H + , EtOH<br />

Co(OAc) 2, Ph3P Ph<br />

PPh3 N N<br />

N N<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

+<br />

380 381<br />

Ph<br />

N<br />

Ph<br />

Co<br />

Ph<br />

382 25%<br />

Ph


HO2C N<br />

CHO<br />

H<br />

a384<br />

17.8.3 Isomeric, Contracted, and Expanded Porphyrin Systems 1183<br />

Et<br />

Et<br />

27<br />

Et<br />

N N<br />

Co<br />

PPh 3<br />

N N<br />

Et<br />

H + , EtOH<br />

Co(OAc) 2, Ph3P<br />

Et<br />

N N<br />

N N<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Co<br />

385<br />

Et<br />

PPh3<br />

17.8.3.2.1.2 Variation 2:<br />

Using 2,5-Di-unsubstituted 1H-Pyrroles and Aldehydes<br />

Et<br />

+<br />

Et<br />

Et<br />

N<br />

N<br />

383<br />

Et<br />

Co<br />

+<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

+<br />

N N<br />

Co<br />

N N<br />

Two examples of serendipitous meso-substituted corrole formation during porphyrin syntheses<br />

using pyrrole and aldehydes have been reported. [339,340] It was subsequently shown,<br />

as part of a planned corrole synthesis, that reaction of pyrrole and benzaldehyde (4:3<br />

molar ratio) in refluxing acetic acid for four hours gives a 6% yield of 5,10,15-triphenylcorrole<br />

(387) (Scheme 89). [341] 5,10,15-Tris(pentafluorophenyl)corrole has also been shown to<br />

be produced from the analogous reaction using pentafluorobenzaldehyde. [342]<br />

Scheme 89 Syntheses of Corroles Using 2,5-Di-unsubstituted 1H-Pyrroles and<br />

Aldehydes [341]<br />

4<br />

N<br />

H<br />

17.8.3.2.2 Method 2:<br />

From Dipyrroles<br />

PhCHO (3 equiv)<br />

6%<br />

NH N<br />

NH HN<br />

Johnson and co-workers were the first to show that corroles can be synthesized by a Mac-<br />

Donald-type [2 +2] cyclization. [343] Thus, treatment of the 2,2¢-bipyrrole-5,5¢-dicarboxylic<br />

acid 388 with the 1,9-diformyldipyrromethane 283 in the presence of acid gives the corrole<br />

complex 389 after treatment with cobalt(II) acetate and triphenylphosphine; indeed,<br />

the reaction fails without the metal salt, and this has been shown to be due to the formation<br />

of a macrocyclic octapyrrole. [344] The reaction also works successfully if the bridging<br />

Ph<br />

Ph<br />

387<br />

Ph<br />

386<br />

PPh3<br />

Et<br />

Et<br />

for references see p 1223


1184 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aformyl carbons are switched to the other component in the [2 +2] reaction, i.e. using the<br />

5,5¢-diformyl-2,2¢-bipyrrole 390 and the dipyrromethane-1,9-dicarboxylic acid 391<br />

(Scheme 90).<br />

Scheme 90 Syntheses of Corroles from Bipyrroles and Dipyrromethanes [343,344]<br />

Et<br />

Et<br />

Et<br />

Et<br />

CO 2H<br />

NH<br />

NH<br />

CO2H<br />

CHO<br />

NH<br />

NH<br />

CHO<br />

+<br />

388 283<br />

+<br />

OHC<br />

OHC<br />

HO 2C<br />

HO2C<br />

HN<br />

HN<br />

HN<br />

HN<br />

390 391<br />

Et<br />

Et<br />

Et<br />

Et<br />

1. HBr, MeOH<br />

2. Co(OAc)2, MeOH<br />

Ph3P<br />

N N<br />

N N<br />

Et Et<br />

(2,8,12,18-Tetraethyl-3,7,13,17-tetramethylcorrolato)(triphenylphosphine)cobalt<br />

(389): [343]<br />

3,3¢-Diethyl-4,4¢-dimethyl-2,2¢-bipyrrole-5,5¢-dicarboxylic acid (388; 100 mg, 0.33 mmol)<br />

and 3,7-diethyl-1,9-diformyl-2,8-dimethyldipyrromethane (283; 100 mg, 0.35 mmol) were<br />

dissolved by warming to 608C in MeOH. The soln was cooled in ice and 49% aq HBr<br />

(0.5 mL) was added dropwise with stirring. A red precipitate formed and this was collected<br />

by filtration and washed with MeOH (5 mL) containing a few drops of HBr. The solid was<br />

dissolved in hot MeOH containing Co(OAc) 2 (200 mg) and Ph 3P (200 mg) and the soln was<br />

boiled on a steam bath for 10 min before being cooled at 08C overnight. The separated<br />

solid was collected by filtration, washed with MeOH, and then crystallized (CHCl 3/<br />

MeOH) to give 389; yield: 54 mg (21%).<br />

17.8.3.2.3 Method 3:<br />

From a,c-Biladiene Salts<br />

As is true for the corresponding porphyrins, the a,c-biladiene route to corroles is the most<br />

general method, and the 1,19-di-unsubstituted a,c-biladiene hydrobromides are usually<br />

obtained by treatment of a dipyrromethane with 2 equivalents of a 2-formylpyrrole in<br />

the presence of hydrogen bromide. There exist numerous ways to achieve cyclization of<br />

the a,c-biladiene. In the earliest described syntheses this is accomplished photochemically<br />

in basic methanol. [345] When ammonia is used as the base, small amounts of monoazaporphyrins<br />

are also obtained, so the procedure is modified by addition of oxidizing agents<br />

(e.g., potassium ferricyanide). [174] meso-Substituted a,c-biladienes also give corroles. For example,<br />

the 5-phenyldipyrromethane-1,9-dicarboxylic acid 392 is treated with 2 equivalents<br />

of the 2-formyl-1H-pyrrole 393 in the presence of hydrogen bromide to give the crystalline<br />

a,c-biladiene dihydrobromide 394 (Scheme 91). This is then cyclized in basic methanol<br />

to give the 10-substituted corrole in good yield; [346–348] the cyclizations are, as usual,<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

A<br />

B<br />

A: 21%<br />

Et<br />

Co<br />

389<br />

PPh 3<br />

Et


17.8.3 Isomeric, Contracted, and Expanded Porphyrin Systems 1185<br />

acarried out in the presence of cobalt(II) and triphenylphosphine, so the cobalt complexes<br />

395 are isolated.<br />

Scheme 91 Syntheses of Corroles from 1H-Pyrroles and Dipyrromethanes [346–348]<br />

HO2C<br />

HO2C<br />

HN<br />

HN<br />

392<br />

Ph<br />

(2 equiv)<br />

N<br />

CHO<br />

H<br />

393<br />

HBr, AcOH<br />

Co(OAc)2, NaOAc<br />

Ph3P, MeOH<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

+<br />

NH<br />

394<br />

HN<br />

+<br />

HN<br />

Ph<br />

N N<br />

Co<br />

N N<br />

1,19-Dibromo-a,c-biladiene dihydrobromides can also be cyclized to afford corroles, the<br />

cyclization step usually involving thermolysis. [84,349,350] 1,19-Dibromo-a,c-biladienes (e.g.,<br />

398) are typically prepared by condensation of two dipyrromethenes (396, 397) using<br />

the Friedel–Crafts alkylation developed by Johnson and co-workers for 1-bromo-19-methyl-a,c-biladiene<br />

and porphyrin syntheses (see Section 17.8.1.1.3.3.3). The metal-free corrole<br />

399 is obtained from 398 simply by refluxing in dimethylformamide followed by esterification<br />

with 5% methanol/sulfuric acid (Scheme 92). [84] Unsymmetrically substituted<br />

1,19-diiodo-a,c-biladienes have also been synthesized and similarly cyclized to give corroles.<br />

[351]<br />

Scheme 92 Syntheses of Corroles by Condensation of Two Dipyrromethenes [84]<br />

Et<br />

Br<br />

Br<br />

Br<br />

Et<br />

Et<br />

NH<br />

NH<br />

+<br />

NH<br />

HN<br />

+<br />

398<br />

HN<br />

+<br />

HN<br />

Br −<br />

CO 2Me<br />

CO 2Me<br />

2Br −<br />

CO2H<br />

+<br />

Br<br />

+<br />

−<br />

NH HN<br />

396 397<br />

Et<br />

Br<br />

1. DMF, heat<br />

2. MeOH, H2SO4 26%<br />

Et<br />

Et<br />

Br<br />

CO 2H<br />

NH N<br />

NH HN<br />

399<br />

395<br />

PPh 3<br />

2Br −<br />

1. SnCl 4<br />

2. HBr<br />

72%<br />

Ph<br />

CO2Me<br />

CO2Me<br />

for references see p 1223


1186 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a3,18-Diethyl-8,12-bis[2-(methoxycarbonyl)ethyl]-2,7,13,17-tetramethylcorrole (399); Typical<br />

Procedure: [84]<br />

9-Bromo-1-(bromomethyl)-8-ethyl-2-(2-carboxyethyl)-3,7-dimethyldipyrromethene hydrobromide<br />

(397) and 1-bromo-3-ethyl-8-[2-(methoxycarbonyl)ethyl]-2,7-dimethyldipyrromethene<br />

hydrobromide (396) were condensed together (see Section 17.8.1.1.3.3.3) to<br />

give 1,19-dibromo-8-(2-carboxyethyl)-3,18-diethyl-12-[2-(methoxycarbonyl)ethyl]-<br />

2,7,13,17-tetramethyl-a,c-biladiene dihydrobromide (398); yield: 72%. This a,c-biladiene<br />

(188 mg, 0.225 mmol) in DMF (35 mL) was heated on a steam bath for 30 min. The solvent<br />

was removed under vacuum and the residue was esterified with 5% MeOH/H 2SO 4 before<br />

being worked up and chromatographed [alumina (Spence type H), CH 2Cl 2]. The corrole<br />

399 was isolated after crystallization (acetone/hexanes); yield: 33 mg (26%); mp 168–<br />

1708C.<br />

17.8.3.2.4 Method 4:<br />

By Porphyrin Ring Contraction<br />

Corroles have also been synthesized by way of sulfur extrusion, with concomitant ring<br />

contraction, from a thiaphlorin 401. The thiaphlorin is synthesized by the MacDonald<br />

[2+2] approach from a bis(5-formyl-1H-pyrrol-2-yl) sulfide 400 and the dipyrromethane-<br />

1,9-dicarboxylic acid 391. Ring contraction to give corrole 402 is then accomplished by<br />

heating in 1,2-dichlorobenzene; [352] using triphenylphosphine in the reaction affords<br />

higher yields of corrole (Scheme 93).<br />

Scheme 93 Synthesis of Corroles by Ring Contraction [352]<br />

EtO 2C<br />

S<br />

EtO2C<br />

NH<br />

NH<br />

CHO<br />

CHO<br />

+<br />

HO 2C<br />

HO 2C<br />

17.8.3.3 Expanded Porphyrins<br />

HN<br />

HN<br />

EtO 2C<br />

EtO2C<br />

EtO2C<br />

NH N<br />

NH HN<br />

NH N<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

HCl(g)/MeOH<br />

52%<br />

Ph 3P<br />

S<br />

EtO2C<br />

400 391 401<br />

Expanded porphyrins comprise a group of pyrrole-based macrocycles with an aromatic<br />

delocalized system larger than that found in the porphyrins. The most investigated examples<br />

are the sapphyrins 373 and pentaphyrins 374 (Scheme 85); the relationship between<br />

sapphyrin and pentaphyrin is similar to that between corrole and porphyrin. The expanded<br />

p-system facilitates the study of aromaticity in large “annulene” systems, and also ex-<br />

15%<br />

402<br />

Et<br />

Et<br />

Et<br />

Et


17.8.3 Isomeric, Contracted, and Expanded Porphyrin Systems 1187<br />

aamination of the characteristics of coordination of large metal ions because the core size<br />

is larger than that found in porphyrins.<br />

17.8.3.3.1 Method 1:<br />

Syntheses of Sapphyrins<br />

Sapphyrin 373 was the first example of an expanded porphyrin to be described, when it<br />

was accidentally obtained by Woodward and co-workers in the early stages of the Harvard<br />

synthesis of vitamin B 12. [353] The macrocycle possesses five pyrrole rings linked by only<br />

four methine bridges and therefore has one direct pyrrole–pyrrole bond. The beautiful<br />

green chromophore features an aromatic 22-electron p-system.<br />

17.8.3.3.1.1 Variation 1:<br />

The [3+2] Approach<br />

A rational [3+2] synthesis of sapphyrin 405 from 403 and 404 (Scheme 94) was eventually<br />

reported by the Harvard group, [64] and this was followed by a number of related [3+2] approaches.<br />

[354,355] With the new advances in tripyrrane synthesis, [65,66] a more efficient approach<br />

to sapphyrins was developed. [61,66] However, the synthetic pathway to sapphyrins<br />

was still based upon the original [3+2] approach pioneered by Woodward and co-workers,<br />

[64] and possessed some symmetry restrictions with regard to the peripheral substituent<br />

array.<br />

More recently, Sessler and co-workers have reported the preparation of symmetrical<br />

meso-diaryl substituted sapphyrins (e.g., 407) by way of a Lindsey-type reaction of arylaldehyde<br />

406, pyrrole, and a bipyrrole dialdehyde 390 under Lewis acid catalysis. [356]<br />

Though the reaction appears to be a [2+1+1+1] method, a tripyrrane is presumably<br />

formed in situ by reaction of two moles of the arylaldehyde and three of pyrrole, and<br />

this subsequently condenses with the bipyrrole unit, once again in a [3 +2] mode (Scheme<br />

94).<br />

Scheme 94 Syntheses of Sapphyrins by the [3 +2] Approach [64,356]<br />

OHC<br />

N N<br />

H H CHO<br />

403<br />

+<br />

CO 2H CO 2H<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

H<br />

N<br />

404<br />

TsOH, O2<br />

44%<br />

N<br />

H<br />

405<br />

N<br />

H<br />

N N<br />

H<br />

N<br />

for references see p 1223


1188 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

OHC<br />

N<br />

H<br />

Et Et<br />

390<br />

N<br />

H<br />

CHO<br />

+ 2<br />

R1 R1<br />

a407<br />

17.8.3.3.1.2 Variation 2:<br />

The [3+(1) 2 ] Approach<br />

CHO<br />

R 1<br />

406<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

+ 3<br />

N<br />

H<br />

1. BF3 MeOH<br />

2. DDQ<br />

N<br />

H<br />

Et Et<br />

N<br />

H<br />

+ 2Cl<br />

NH HN<br />

−<br />

+<br />

Treatment of 1,14-diformyltripyrranes (e.g., 408) with 2 equivalents of a 2,5-di-unsubstituted<br />

3,4-diethyl-1H-pyrrole affords sapphyrins 409 in 28–34% yields (Scheme 95). [357]<br />

Scheme 95 Syntheses of Sapphyrins by the [3 +(1) 2 ] Approach [357]<br />

CHO OHC<br />

NH HN<br />

H<br />

N<br />

HO OH<br />

Et<br />

408<br />

Et<br />

1.<br />

Et<br />

N<br />

H<br />

Et<br />

TFA, CH2Cl2 2. DDQ, Et3N<br />

34%<br />

(2 equiv)<br />

HO<br />

Et<br />

N<br />

H<br />

H<br />

N<br />

409<br />

N<br />

H<br />

N N<br />

Et<br />

Et Et<br />

H<br />

N<br />

Et<br />

Et<br />

OH


17.8.3 Isomeric, Contracted, and Expanded Porphyrin Systems 1189<br />

a2,3,12,13,22,23-Hexaethyl-8,17-bis(2-hydroxyethyl)-7,18-dimethyl-1,24-sapphyrin (409);<br />

Typical Procedure: [357]<br />

5% TFA in CH 2Cl 2 (200 mL) was added to 1,14-diformyl-7,8-diethyl-3,12-bis(2-hydroxyethyl)-2,13-dimethyltripyrrane<br />

(408; 96 mg, 0.2 mmol) and 3,4-diethyl-1H-pyrrole (49 mg,<br />

0.4 mmol). The mixture was stirred at rt overnight before being neutralized with Et 3N,<br />

and then DDQ (45 mg, 0.2 mmol) was added. The mixture was concentrated to a volume<br />

of 100 mL on a rotary evaporator and then washed with aq NaHCO 3 (2 ” 100 mL) and H 2O<br />

(2 ” 100 mL). The organic phase was dried (MgSO 4), filtered, and concentrated, and the residue<br />

was chromatographed (silica gel, CH 2Cl 2/MeOH 96:4) to afford the sapphyrin 409;<br />

yield: 47 mg (34%); mp > 3008C.<br />

17.8.3.3.1.3 Variation 3:<br />

The [4+1] Approach<br />

Corroles are readily obtained by cyclization of 1,19-di-unsubstituted a,c-biladienes (Section<br />

17.8.3.2.3), a reaction in which the direct pyrrole–pyrrole link is formed as the last<br />

stage in the cyclization process; attempts have, therefore, been made to apply this approach<br />

to sapphyrin syntheses. Since 1,19-di-unsubstituted a,c-biladiene salts have been<br />

shown to react with aldehydes in acidic ethanol to afford the corresponding porphyrins<br />

in good yields, a new route involving condensation of an a,c-biladiene 410 with a 2-formyl-1H-pyrrole<br />

411 was investigated; this was based on the hypothesis that a transient<br />

pentapyrrolic intermediate 412 might subsequently cyclize to produce sapphyrin 413<br />

rather than a meso-pyrrolylporphyrin. In this way, the sapphyrin 413 was successfully obtained<br />

in good yield (presumably via the intermediate 412), along with small amounts of<br />

2,3,7,8,12,13,17,18-octaethylporphyrin (414, R 1 = Et), another porphyrin 414 (R 1 =Me),<br />

and corrole 415 (Scheme 96). [358]<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1190 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 96 Syntheses of Sapphyrins by the [4+1] Approach [358]<br />

Et<br />

NH HN<br />

+<br />

+<br />

NH HN<br />

410<br />

Et<br />

2Br −<br />

Et Et<br />

N<br />

CHO<br />

H<br />

411<br />

A: TsOH, abs EtOH<br />

B: AcOH, reflux, 1 h<br />

Et<br />

+<br />

Et<br />

Et<br />

N<br />

H<br />

+ N<br />

H<br />

+<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

H<br />

N<br />

H<br />

N N<br />

Et<br />

H<br />

N<br />

Et<br />

413 A: 20%<br />

B: 18%<br />

Et<br />

Et<br />

Et Et<br />

NH N<br />

N HN<br />

Et<br />

+<br />

+<br />

R 1<br />

R 1<br />

Et<br />

H<br />

N+<br />

412<br />

R 1<br />

R 1<br />

Et<br />

NH N<br />

N HN<br />

414 R 1 = Me, Et<br />

NH N<br />

NH HN<br />

415<br />

3OTs −<br />

2,3,13,17-Tetraethyl-7,8,12,18,22,23-hexamethylsapphyrin (413); Typical Procedure: [358]<br />

Method A: 2-Formyl-3,4-diethyl-1H-pyrrole (411; 0.5 g, 3.31 mmol), 8,12-diethyl-<br />

2,3,7,13,17,18-hexamethyl-a,c-biladiene dihydrobromide (410; 0.5 g, 0.83 mmol), and<br />

TsOH (0.25 g) were dissolved in abs EtOH (100 mL) and the mixture was refluxed in the<br />

dark (aluminum foil). Progress of the reaction was monitored spectrophotometrically;<br />

when absorbances attributable to 410 disappeared the solvent was removed under vacuum,<br />

the resulting solid was redissolved in CH 2Cl 2, then washed with H 2O (” 3) and dried<br />

(Na 2SO 4). The solvent was removed and the crude mixture was chromatographed (silica<br />

gel, CH 2Cl 2) to yield two bands. The first (red-brown) fraction to be eluted contained<br />

2,3,7,8,12,13,17,18-octaethylporphyrin (21); yield: 65 mg, together with traces of 8,12-diethyl-2,3,7,13,17,18-hexamethylporphyrin<br />

414 (R 1 = Me), while the second (red-violet)<br />

fraction contained 8,12-diethyl-2,3,7,13,17,18-hexamethylcorrole (415); yield: 48 mg.<br />

The column was then eluted with CH 2Cl 2 containing increasing amounts of methanol<br />

(2–10%). A green fraction was collected, the solvent was removed under vacuum, and the<br />

residue was crystallized (CH 2Cl 2/hexanes) to give 413 as its 4-toluenesulfonate salt; yield:<br />

158 mg (20%); mp > 3008C. The dihydrochloride derivative was obtained by washing a<br />

Et<br />

21<br />

Et<br />

Et<br />

Et<br />

R 1<br />

R 1<br />

Et<br />

Et<br />

Et<br />

Et


17.8.3 Isomeric, Contracted, and Expanded Porphyrin Systems 1191<br />

aCH 2Cl 2 soln of [2H•413] 2+ (OTs) 2 with sat. aq Na 2CO 3 until spectrophotometry showed the<br />

formation of the corresponding free base; subsequent treatment with dil HCl quantitatively<br />

afforded [2H•413] 2+ Cl 2; mp > 300 8C.<br />

Method B: Formylpyrrole 411 (0.5 g, 3.31 mmol) and a,c-biladiene 410 (0.5 g,<br />

0.83 mmol) were dissolved in AcOH (100 mL) and refluxed in the dark for 1 h. The solvent<br />

was removed under vacuum, the resulting solid was redissolved in CH 2Cl 2, washed with<br />

sat. Na 2CO 3, and with dil HCl, and dried (Na 2SO 4). The solvent was removed under vacuum<br />

and the crude mixture was chromatographed (silica gel, as outlined in Method A). The appropriate<br />

green fraction was collected, and the solvent was removed in vacuo to give<br />

[2H•413] 2+ Cl 2; yield: 98 mg (18%).<br />

17.8.3.3.1.4 Variation 4:<br />

The [1] 5 Approach<br />

5,10,15,20-Tetraphenylsapphyrin (416) has been shown to be formed in low yield during<br />

the Rothemund synthesis of meso-tetraphenylporphyrin (Section 17.8.1.2.1) (Scheme<br />

97). [359]<br />

Scheme 97 5,10,15,20-Tetraphenylsapphyrin [359]<br />

Ph N N Ph<br />

H H<br />

N N<br />

H<br />

N<br />

Ph Ph<br />

416<br />

17.8.3.3.2 Method 2:<br />

Syntheses of Pentaphyrins<br />

Previously published information regarding this product subclass can be found in Houben–Weyl,<br />

Vol. E 9d, pp 700–708.<br />

Pentaphyrins 374 are fully conjugated macrocycles with five pyrrole subunits linked<br />

by methine (sp 2 carbons); they differ from the more highly investigated sapphyrins in that<br />

pentaphyrins do not possess a direct pyrrole–pyrrole bond. The first pentaphyrin synthesis<br />

was reported by Rexhausen and Gossauer, [360] and employed a [3+2] MacDonald-type<br />

approach. [361–363] Thus, the 1,14-diformyltripyrrane 418 reacts under acid catalysis with<br />

the 1,9-di-unsubstituted dipyrromethane 417, presumably to give the macrocycle 419,<br />

which is subsequently oxidized with chloranil to give the pentaphyrin 420 in good yield<br />

(Scheme 98); several other versions of this same approach have subsequently been reported,<br />

along with the revelation that pentaphyrins cannot be obtained if the new mesobridge<br />

carbons (i.e., the 1- and 14-formyl groups in 418) are situated on the dipyrromethane.<br />

[364–367] Presumably, the unprotected tripyrrane is subject to acid-catalyzed ring<br />

redistribution reactions and, therefore, the more favorable porphyrin products are produced.<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1192 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 98 Syntheses of Pentaphyrins by the [3+2] Approach [360]<br />

NH HN<br />

417<br />

H +<br />

+<br />

N<br />

H<br />

+ N<br />

H<br />

+<br />

NH HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

H<br />

N<br />

419<br />

CHO OHC<br />

NH HN<br />

H<br />

N<br />

MeO 2C CO 2Me<br />

418<br />

2X −<br />

MeO2C CO 2Me<br />

17.8.4 Reactions around the Porphyrin Periphery<br />

N<br />

H<br />

N<br />

N HN<br />

N<br />

MeO2C CO2Me Very thorough reviews of reactivity in both the 2,3,7,8,12,13,17,18-octaalkyl- [368] and<br />

5,10,15,20-tetraalkyl/aryl- [369] porphyrin series are now available. Since porphyrins are aromatic<br />

molecules, this dominates their overall reactivity patterns and profiles.<br />

17.8.4.1 Method 1:<br />

Electrophilic Substitution Reactions<br />

Substitution and addition reactions can take place at the b—b¢ peripheral double bonds,<br />

with retention of the 18p-aromatic delocalization pathway, leading to functionalized porphyrins,<br />

chlorins, bacteriochlorins, or isobacteriochlorins. Substitutions at the meso positions<br />

can also occur, with formation of meso-substituted porphyrins, phlorins, porphodimethenes,<br />

or nonaromatic expanded “homo-type” macrocycles. The basic inner nitrogen<br />

atoms of porphyrins react with electrophilic reagents and are easily protonated (thereby<br />

deactivating the porphyrin nucleus toward electrophilic attack), unless previously protected<br />

by metalation; zinc(II) is often the ion of choice for metalation because it is moderately<br />

robust and does not draw electron density away from the porphyrin p-system. However,<br />

the zinc(II) ion can be removed under acidic conditions, [370] so nickel(II) or copper(II)<br />

420<br />

[O]


aare sometimes used instead; these metals are more electronegative within the porphyrin<br />

core, but at the same time they are also more resistant to demetalation by mild acids. Copper(II)<br />

and nickel(II) porphyrinates are usually demetalated using trifluoroacetic acid in<br />

sulfuric acid.<br />

17.8.4.1.1 Variation 1:<br />

Halogenation<br />

17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1193<br />

Porphyrins undergo halogenation at the peripheral unsubstituted b- and meso positions.<br />

Fluorination and chlorination usually take place at the more-reactive meso positions,<br />

whereas bromination and iodination occur mainly at the less sterically congested b-positions.<br />

N-Fluoropyridinium triflates [371] or cesium fluoroxysulfate [372] have been used for<br />

meso-polyfluorination of porphyrins, such as 2,3,7,8,12,13,17,18-octaethylporphyrin<br />

(21), which gives the mono-, di- (two isomers), tri-, and tetrafluoro-2,3,7,8,12,13,17,18-octaethylporphyrins<br />

421–425 shown in Scheme 99. b-Fluorination of zinc(II) 5,10,15,20-tetraphenylporphyrinate<br />

has been accomplished with cobalt(II) fluoride or silver(II) fluoride<br />

in dichloromethane. [373]<br />

Scheme 99 Fluorinated Porphyrin Isomers<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

F<br />

NH N<br />

N HN<br />

421<br />

F<br />

NH N<br />

N HN<br />

F<br />

424<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

F<br />

422<br />

Et<br />

Et<br />

NH N<br />

F F<br />

F<br />

Et<br />

Et<br />

N HN<br />

Et<br />

F<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

F<br />

NH N<br />

N HN<br />

Chlorination of porphyrins is achieved using N-chlorosuccinimide, [374–379] chlorine/iron(-<br />

III) chloride, [380] hydrogen chloride/hydrogen peroxide, [377,380] and phenylselenium trichloride.<br />

[381] Mono- and dichloro-2,3,7,8,12,13,17,18-octaethylporphyrin are usually obtained<br />

with N-chlorosuccinimide or with hydrogen chloride/hydrogen peroxide in aqueous<br />

tetrahydrofuran, but meso-tetrachlorination occurs with hydrogen chloride/hydrogen<br />

peroxide or chlorine in acetic acid.<br />

Porphyrins are brominated with N-bromosuccinimide, [375,376,379,382–387] hydrogen bromide/hydrogen<br />

peroxide, [380,385] bromine in various solvents, [376,385–389] or phenylselenium<br />

tribromide. [381] Electrophilic bromination of 5,15-diphenylporphyrin (426) with 2 equivalents<br />

of N-bromosuccinimide occurs selectively at the unsubstituted meso positions with<br />

formation of dibromide 427 (Scheme 100). [384] Regioselective tetrabromination of metalfree<br />

meso-tetraarylporphyrins, such as 5,10,15,20-tetraphenylporphyrin (22), is accomplished<br />

using an excess of N-bromosuccinimide, and regioselectively affords 2,3,12,13-tet-<br />

Et<br />

Et<br />

F<br />

F<br />

425<br />

Et<br />

Et<br />

Et<br />

Et<br />

F<br />

423<br />

Et<br />

Et<br />

Et<br />

Et<br />

for references see p 1223


1194 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

arabromoporphyrins, such as 428 (Scheme 100); [390] the initial b-bromo substituent perturbs<br />

the position of the tautomeric equilibrium, favoring the tautomer with the substituent<br />

on an isolated double bond and thereby causing bromination on the opposite pyrrolic<br />

unit. [390,391] 2-Nitro-5,10,15,20-tetraphenylporphyrin also undergoes vicinal dibromination<br />

with N-bromosuccinimide at the pyrrole subunit opposite the nitro group to give<br />

12,13-dibromo-2-nitro-5,10,15,20-tetraphenylporphyrin (429) (Scheme 100). The nitro<br />

group can be removed using sodium borohydride to give the dibromonitrochlorin 430<br />

which eliminates nitrous acid to afford 2,3-dibromo-5,10,15,20-tetraphenylporphyrin<br />

(431). The nitrochlorin 430, also reacts with copper(I) cyanide to give the copper(II) 2,3dicyano-5,10,15,20-tetraphenylporphyrinate<br />

(432) by nucleophilic displacement of the<br />

two bromines and metalation (Scheme 100). [390]<br />

Scheme 100 Bromination of Porphyrins [384,390]<br />

Ph<br />

NH N<br />

N HN<br />

Ph<br />

426<br />

Ph<br />

NH N<br />

Ph Ph<br />

N HN<br />

Ph<br />

22<br />

NBS (2 equiv)<br />

CHCl3 NBS (excess)<br />

CHCl3 NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Ph<br />

Br Br<br />

Ph<br />

Br<br />

Ph<br />

427<br />

Br<br />

Ph<br />

NH N<br />

N HN<br />

Ph<br />

428<br />

Br<br />

Br<br />

Ph


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1195<br />

Ph<br />

NH N<br />

Ph Ph<br />

N HN<br />

Ph<br />

a429<br />

NO 2<br />

Ph<br />

NH N<br />

82%<br />

Ph Ph<br />

Br<br />

Br<br />

N HN<br />

Ph<br />

430<br />

NO2<br />

NBS<br />

CHCl3<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Ph<br />

Ph Ph<br />

Br<br />

Br<br />

silica gel<br />

CHCl3, reflux<br />

− HNO 2<br />

98%<br />

CuCN<br />

quinoline, heat<br />

− HNO 2<br />

75%<br />

Ph<br />

NO 2<br />

Ph<br />

N N<br />

Ph Cu<br />

Ph<br />

NC<br />

NC<br />

Ph<br />

NH N<br />

Ph Ph<br />

Br<br />

Br<br />

N HN<br />

Ph<br />

431<br />

N N<br />

Although less favored, iodination of porphyrins has been accomplished using bis(trifluoroacetoxy)iodobenzene–iodine<br />

[392] and with iodine in 1,2-dichlorobenzene. [385] b-Iodination<br />

is favored for deuteroporphyrin IX dimethyl ester (433), which produces 434 by reaction<br />

with iodine at high temperature (Scheme 101).<br />

Scheme 101 Iodination of Porphyrins [385]<br />

MeO2C<br />

NH N<br />

N HN<br />

433<br />

CO2Me<br />

I2, 1,2-Cl2C6H4<br />

180 o C<br />

MeO2C<br />

I<br />

Ph<br />

NaBH4<br />

THF<br />

432<br />

NH N<br />

N HN<br />

434<br />

84%<br />

I<br />

CO2Me<br />

Iodine, bromine, and N-bromosuccinimide can produce the p-cation radicals of porphyrins,<br />

which are known to react with a variety of nucleophiles, or to participate in dimerization<br />

reactions (vide infra). The reaction of 2,3,7,8,12,13,17,18-octaethylporphyrin (21)<br />

with N-bromosuccinimide in the presence of 2,2¢-azobisisobutyronitrile produces (E)-(2-<br />

for references see p 1223


1196 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

abromovinyl)porphyrin 436 by way of a 2-(1-bromoethyl)porphyrin intermediate 435,<br />

which can be trapped as an ether derivative if alcohols are present (Scheme 102). [378,393]<br />

Scheme 102 Bromovinylporphyrin Synthesis [378,393]<br />

Et<br />

Et<br />

Et<br />

Et<br />

NH N<br />

N HN<br />

21<br />

Et<br />

Et<br />

Et<br />

Et<br />

NBS, AIBN<br />

1,2-dichloroethane<br />

NH N<br />

N HN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

Et<br />

Et<br />

435<br />

Et<br />

Et<br />

Br<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

NH N<br />

N HN<br />

436 92%<br />

Bromoporphyrins are important starting materials in various metal-catalyzed reactions.<br />

[394,395] Nucleophilic substitutions of bromo substituents by cyanide (e.g.,<br />

430fi432), [241,384,396] nitrite, [397] or by thiolates [398] are also known.<br />

Fluorination of 2,3,7,8,12,13,17,18-Octaethylporphyrin (OEP, 21); Typical Procedure: [372]<br />

2,3,7,8,12,13,17,18-Octaethylporphyrin (21; 100 mg, 0.187 mmol) was dissolved in dioxane<br />

(350 mL) with heating (908C, 2 h) and then treated at rt with an excess of freshly prepared<br />

aq cesium fluoroxysulfate (500 mg in 3 mL) with vigorous stirring. After 10 min the<br />

soln was poured into H 2O (800 mL) and extracted with CHCl 3 (2 ” 100 mL). The combined<br />

organic extracts were washed with H 2O (3 ” 200 mL), dried (Na 2SO 4), concentrated, and<br />

subjected to preparative TLC (silica gel, CHCl 3/petroleum ether 1:2) to give six bands: (1)<br />

recovered 2,3,7,8,12,13,17,18-octaethylporphyrin (21); 19.8 mg. (2) 2,3,7,8,12,13,17,18-octaethyl-5-fluoroporphyrin<br />

(421); yield: 29 mg (29%); mp 235–2378C. (3)<br />

2,3,7,8,12,13,17,18-octaethyl-5,10-difluoroporphyrin (422); yield: 12 mg (11%); mp 280–<br />

2828C. (4) 2,3,7,8,12,13,17,18-octaethyl-5,15-difluoroporphyrin (423); yield: 10 mg (9.5%);<br />

mp 290–2928C. (5) 2,3,7,8,12,13,17,18-octaethyl-5,10,15-trifluoroporphyrin (424); yield:<br />

5 mg (4%); mp 232–2348C. (6) 2,3,7,8,12,13,17,18-octaethyl-5,10,15,20-tetrafluoroporphyrin<br />

(425); yield: 3 mg (2%); mp 208–2098C.<br />

12,13-Dibromo-2-nitro-5,10,15,20-tetraphenylporphyrin (429); Typical Procedure: [390]<br />

A mixture of 2-nitro-5,10,15,20-tetraphenylporphyrin (1.0 g, 1.52 mmol) and NBS (0.68 g,<br />

2.42 equiv) in dry CHCl 3 (EtOH-free; 150 mL) was refluxed overnight. After cooling to rt,<br />

the mixture was filtered through a neutral alumina plug (Brockmann Grade III) eluting<br />

with CH 2Cl 2. The filtrate was concentrated to dryness and the resulting residue was crystallized<br />

(CH 2Cl 2/MeOH) to give 429; yield: 1.01 g (82%); mp > 3008C.<br />

Et<br />

Br<br />

Et<br />

Et


a12,13-Dibromo-2,3-dihydro-2-nitro-5,10,15,20-tetraphenylporphyrin (430); Typical Procedure:<br />

[390]<br />

To a cold soln (ice/NaCl) of dry THF (30 mL) under argon was added a mixture of 12,13-dibromo-2-nitro-5,10,15,20-tetraphenylporphyrin<br />

(429; 500 mg, 0.61 mmol) and NaBH 4<br />

(40 mg, 1.11 mmol). The resulting mixture was stirred for 2 h, the ice bath being removed<br />

after 1 h. The progress of the reaction was monitored by spectrophotometry; after 2 h the<br />

Soret band had shifted from 436 to 424 nm. CH 2Cl 2 (100 mL) was then added and the mixture<br />

was poured into H 2O. The organic phase was washed with H 2O (2 ”) and concentrated<br />

to dryness. The residue was redissolved in CH 2Cl 2 and filtered through a short neutral alumina<br />

plug (Brockmann Grade V, eluting with CH 2Cl 2). The filtrate was concentrated to<br />

dryness, and the residue was crystallized (CH 2Cl 2/MeOH) to yield 430; yield: 420 mg<br />

(84%); mp > 3008C (dec with elimination of HNO 2 >808C).<br />

2,3-Dibromo-5,10,15,20-tetraphenylporphyrin (431); Typical Procedure: [390]<br />

A mixture of nitrochlorin 430 (200 mg, 0.24 mmol), silica gel (20 g), and CHCl 3 (100 mL)<br />

was refluxed for 1 d under argon. The mixture was cooled to rt and the silica gel was removed<br />

by filtration and washed thoroughly with CH 2Cl 2. After concentration of the solvents<br />

from the combined filtrate and washings, the residue was crystallized (CH 2Cl 2/<br />

MeOH) to yield 431; yield: 185 mg (98%); mp > 300 8C.<br />

(2,3-Dicyano-5,10,15,20-tetraphenylporphyrinato)copper(II) (432); Typical Procedure: [390]<br />

A mixture of nitrochlorin 430 (500 mg, 0.61 mmol), CuCN (1.1 g, 12 mmol), and quinoline<br />

(15 mL) was heated at 2008C under argon for 2 h. The mixture was allowed to cool and<br />

CH 2Cl 2 (100 mL) was added. The excess of CuCN was removed by filtration and the organic<br />

phase was washed with 10% HCl (3 ” 100 mL), H 2O (” 2), dried (Na 2SO 4), and then concentrated<br />

to dryness. The residue was crystallized (CH 2Cl 2/cyclohexane) to give 432; yield:<br />

330 mg (75%); mp > 3008C.<br />

2-[(E)-2-Bromovinyl]-3,7,8,12,13,17,18-heptaethylporphyrin (436); Typical Procedure: [378]<br />

NBS (39.5 mg, 0.22 mmol) in 1,2-dichloroethane (3 mL) was added to a soln of<br />

2,3,7,8,12,13,17,18-octaethylporphyrin (21; 61.4 mg, 0.12 mmol) in 1,2-dichloroethane<br />

(40 mL). AIBN (2.7 mg, 0.02 mmol) in 1,2-dichloroethane (0.5 mL) was added with continuous<br />

stirring. The mixture was then refluxed for 5 h, concentrated under vacuum, and then<br />

chromatographed on preparative TLC plates (silica gel, 30% petroleum ether/CH 2Cl 2). The<br />

front-running band was collected and crystallized (CH 2Cl 2/MeOH) to give porphyrin 436;<br />

yield: 57.9 mg (83%); mp >350 8C. Recovery of 2,3,7,8,12,13,17,18-octaethylporphyrin (21)<br />

from a minor TLC band raised the yield to 92%, based on consumed starting material.<br />

17.8.4.1.2 Variation 2:<br />

Nitration<br />

17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1197<br />

Nitration of porphyrins is accomplished by aromatic electrophilic substitution or via pcation<br />

radical intermediates, and takes place preferentially at the meso positions, though<br />

b-nitrations occur when these positions are sterically accessible. [399] Electrophilic nitration<br />

is achieved with fuming nitric acid, [400,401] nitric acid in acetic or sulfuric acid, [402–405]<br />

copper(II) or zinc(II) nitrate in acetic anhydride, [406] or with nitronium tetrafluoroborate.<br />

[402,407] Reaction of nitrogen dioxide [408–411] or nitrite ion [397,412–414] with the p-cation radicals<br />

of porphyrins accomplishes efficient nitration of the porphyrin ring, even on a large<br />

scale (e.g., preparation of 438 from 437, Scheme 103). The electron-withdrawing effect of<br />

the nitro groups deactivates the macrocycle toward further electrophilic substitutions,<br />

therefore the mono-nitro derivatives can be obtained in good yields. This fact notwithstanding,<br />

heptanitration of 439 to produce 440 has been accomplished using fuming ni-<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1198 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

atric acid in nitromethane in the presence of acetic anhydride and montmorillonite K-10<br />

clay (Scheme 103). [415]<br />

Scheme 103 Nitration of 5,10,15,20-Tetraarylporphyrins [408–411,415]<br />

N<br />

N<br />

Ph<br />

437<br />

N<br />

Ph Cu Ph<br />

N<br />

Ph<br />

Ar Zn<br />

1 Ar1 N<br />

Ar 1 = 2,6-Cl 2C 6H 3<br />

Ar 1<br />

Ar 1<br />

439<br />

N<br />

N<br />

N<br />

N2O4, CH2Cl2<br />

88%<br />

1. HNO3, Ac2O, K-10 clay<br />

2. TfOH<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

N<br />

Ph<br />

438<br />

N<br />

Ph Cu Ph<br />

Ph<br />

O 2N<br />

N<br />

NO2<br />

Ar1 O2N NO2 NH<br />

N<br />

Ar<br />

440<br />

1<br />

N<br />

Ar 1 Ar 1<br />

Reduction of peripheral nitro groups to amino groups can be achieved with tin(II) chloride/hydrogen<br />

chloride, [416] or with sodium borohydride and 10% palladium on carbon in<br />

methanol. [412,417] b-Aminoporphyrins undergo a variety of reactions, such as acylation, [404]<br />

diazotization, [418,419] and reactions with aldehydes and ketones. [242,420,421] 5,10,15,20-Tetraaryl-2-nitroporphyrins<br />

and their metal derivatives can undergo ipso-substitution of the<br />

nitro group [243,412,422–424] and nucleophilic Michael additions [242,243,411,425–429] with several nucleophiles.<br />

b-Fused pyrroloporphyrins 442 are synthesized from metal 2-nitro-5,10,15,20tetraphenylporphyrinates<br />

(e.g., 441) by reaction with -isocyanoacetic esters in the presence<br />

of a base (Scheme 104); a transesterification of the ester takes place during the reaction<br />

unless tert-butyl alcohol is used as the solvent. [430,431]<br />

O2N<br />

Scheme 104 Reaction of 5,10,15,20-Tetraaryl-2-nitroporphyrins [430,431]<br />

Ph<br />

N<br />

N<br />

Ph<br />

Ni<br />

Ph<br />

441<br />

N<br />

N<br />

NO2<br />

Ph<br />

CNCH2CO2Me<br />

DBU, BnOH<br />

Ph<br />

N<br />

N<br />

Ph<br />

Ni<br />

N<br />

N<br />

Ph<br />

442<br />

HN<br />

NH<br />

Ph<br />

NO2<br />

CO 2Bn<br />

(2-Nitro-5,10,15,20-tetraphenylporphyrinato)copper(II) (438); Typical Procedure: [411]<br />

A suspension of (5,10,15,20-tetraphenylporphyrinato)copper(II) (437; 25.0 g, 37.0 mmol)<br />

in CH 2Cl 2 (2.0 L) was stirred vigorously. N 2O 4/petroleum ether soln (preparation given be-<br />

NO 2<br />

NO 2


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1199<br />

alow) was added dropwise over 2–3 h. The addition was performed rapidly at the beginning,<br />

but was slowed down near completion of the reaction to avoid over-nitration. Close<br />

reaction monitoring by TLC (alumina, CHCl 3/cyclohexane 1:2) was essential. In this way,<br />

all of the (5,10,15,20-tetraphenylporphyrinato)copper(II) was converted into the mononitro<br />

product with minimal dinitro product formation. The mixture was filtered to remove<br />

any unreacted (5,10,15,20-tetraphenylporphyrinato)copper(II) and then concentrated to a<br />

volume of about 100 mL. The product 438 was precipitated by addition of MeOH; yield:<br />

22.0 g (88%).<br />

N 2O 4/Petroleum Ether Solution: N 2O 4 gas was prepared by the dropwise addition of<br />

concd HNO 3 (150 mL) from a dropping funnel into a 500-mL three-necked flask (with an<br />

argon inlet) containing Zn metal (50 g). The gas was carried by a slow stream of argon toward<br />

a trap containing P 2O 5 to remove any H 2O or HNO 3. The P 2O 5 trap was connected to a<br />

hose leading to a container of petroleum ether (ca. 150 mL) cooled with liq N 2 and the<br />

N 2O 4 was bubbled through the cold petroleum ether. The petroleum ether quickly dissolved<br />

the N 2O 4, producing a soln that was blue at low temperatures and dark orange at rt.<br />

17.8.4.1.3 Variation 3:<br />

Formylation and Acylation<br />

The Vilsmeier formylation of porphyrins has long been one of the most effective and popular<br />

methods for introducing carbon substituents onto the porphyrin periphery. Because<br />

zinc(II) porphyrinates are usually demetalated by the acidic Vilsmeier reagent (POCl 3/<br />

DMF), the nickel(II) or copper(II) complexes are normally used; the initial product obtained<br />

from the Vilsmeier reaction is an imine, and this is subsequently hydrolyzed by<br />

treatment with base. Thus, metal b-octaalkylporphyrinates (e.g., 443) are converted into<br />

the corresponding meso-formyl derivatives 444 in high yields, [432,433] whereas metal mesotetraarylporphyrinates<br />

undergo b-formylation, giving, for example, 446 from 445<br />

(Scheme 105). [434–436] Under the Vilsmeier conditions, copper(II) deuteroporphyrin IX dimethyl<br />

ester (447) gives a complex mixture of b- and meso-formylated products. [437] As<br />

would be expected, sterically hindered Vilsmeier reagents (e.g., diisobutylformamide/<br />

POCl 3) favor b-substitution. [438] Alternatively, trimethyl orthoformate in trifluoroacetic<br />

acid selectively mono-b-formylates metallodeuteroporphyrin IX dimethyl esters (e.g.,<br />

447) to give the two regioisomeric b-monoformyl products 448 and 449 after demetalation<br />

(Scheme 105). [221,439]<br />

Scheme 105 Syntheses of Formylporphyrins [221,432–436,439]<br />

Et<br />

Et<br />

Et Et<br />

Et<br />

N<br />

N<br />

Ni<br />

443<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

1. POCl3, DMF<br />

2. H2O 92%<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

444<br />

N<br />

Ni CHO<br />

N<br />

Et<br />

Et Et<br />

Et<br />

Et<br />

for references see p 1223


1200 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a445<br />

N<br />

N<br />

Ph<br />

N<br />

Ph Cu Ph<br />

Ph<br />

N<br />

N<br />

N<br />

Cu<br />

MeO 2C CO 2Me<br />

447<br />

N<br />

N<br />

1. POCl3, DMF<br />

2. H2O 1. HC(OMe) 3, TFA<br />

2. H2SO4 3. CH2N2 K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

N<br />

N<br />

Ph<br />

Ph<br />

446<br />

N<br />

Ph Cu Ph<br />

+<br />

OHC<br />

N<br />

NH<br />

N<br />

CHO<br />

MeO2C CO2Me 448 39%<br />

NH<br />

N<br />

N<br />

HN<br />

N<br />

HN<br />

MeO2C CO 2Me<br />

449 50%<br />

Vinylporphyrins such as 450 react faster with Vilsmeier reagents to give (E)-(2-formylvinyl)porphyrins<br />

451 after hydrolysis of the intermediate imine (Scheme 106). [440] 2-Formylvinyl<br />

groups can be directly introduced at the meso positions by treatment of copper(II) or<br />

nickel(II) 2,3,7,8,12,13,17,18-octaethylporphyrinates 452 with a vinylogous Vilsmeier reagent<br />

prepared from 3-(dimethylamino)acrolein and phosphoryl chloride, followed by basic<br />

hydrolysis to afford 453 (Scheme 106). [441,442]<br />

Scheme 106 Syntheses of Formylvinylporphyrins [440–442]<br />

N<br />

N<br />

Ni<br />

MeO 2C CO2Me<br />

450<br />

N<br />

N<br />

1. POCl3, DMF<br />

2. H2O OHC<br />

N<br />

N<br />

Ni<br />

CHO<br />

MeO2C CO2Me<br />

451<br />

N<br />

N<br />

CHO


a452<br />

17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1201<br />

Et<br />

Et<br />

Et<br />

Et<br />

M = Cu, Ni<br />

N<br />

N<br />

M<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

1. POCl3, CH2Cl2 CHO<br />

Me2N<br />

2. H 2O<br />

M = Cu 57%<br />

The formyl group has a very rich organic chemistry, which makes formylporphyrins very<br />

versatile as synthetic intermediates. For example, meso- and b-formyl groups are transformed<br />

by standard chemical routes into oximes, which can be dehydrated to produce<br />

the corresponding cyanoporphyrins. [435,443–445] Reactions of formylporphyrins with Wittig-type<br />

reagents (Ph 3P=CHR 1 ,R 1 = H, Me, Ph, Br, Cl, CO 2Me, CO 2t-Bu) [434,446–453] yield a<br />

wide variety of E- and Z-(2-substituted vinyl)porphyrins. Reduction of acrylates to propanoates<br />

is achieved with Raney nickel or by catalytic hydrogenation in formic acid. [447,453]<br />

Metalloalkynylporphyrins are obtained by dehydrobromination of metal (2-bromovinyl)porphyrinates<br />

[454] and participate in palladium-catalyzed cross-coupling reactions with<br />

aryl iodides and with b-bromovinylporphyrins. [394,455] Metalloethynylporphyrins undergo<br />

oxidative couplings using tetra(triphenylphosphine)palladium, copper(I) iodide, and triethylamine,<br />

or copper(I) chloride–N,N,N¢,N¢-tetramethylethylenediamine complex in the<br />

presence of air, to give butadiyne-linked bisporphyrins. [456] Reactions of metal formylporphyrinates<br />

with Grignard and aryllithium reagents usually produce the corresponding alcohols.<br />

[457,458] Allylboronic acid reacts with 3- and 8-formyldeuteroporphyrin IX dimethyl<br />

ester to afford (1-hydroxybutenyl)porphyrins. [459] Borohydride reduction of the meso-formyl<br />

group of 444 gives the corresponding meso-hydroxymethylporphyrin 454, but use of<br />

sodium borohydride/acetic acid results in complete reduction to the methyl group, with<br />

formation of 455 (Scheme 107). [460,461]<br />

Scheme 107 Reductions of 5-Formylporphyrins [460,461]<br />

Et<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

Ni<br />

444<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

CHO<br />

Et<br />

NaBH 4, EtOH<br />

NaBH 4, AcOH<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

M<br />

453<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

Ni<br />

454<br />

Ni<br />

455<br />

Et<br />

Et<br />

N<br />

N<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

CHO<br />

OH<br />

for references see p 1223


1202 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aIt was discovered during attempts to demetalate the nickel(II) meso-hydroxymethylporphyrinate<br />

454 that these compounds undergo a self-condensation reaction under acidic<br />

conditions to give bisporphyrins 456 linked by ethane bridges (Scheme 108); [462] these<br />

can be transformed into trans-ethene bisporphyrins 457 by a strange dehydrogenation<br />

in acetic acid. [463–466] 1,2-Ethene–bisporphyrins (cis- and trans-isomers), such as 457, are accessible<br />

directly by reductive coupling of metal formylporphyrinates, such as 444, using<br />

low-valent titanium reagents. [394,441,442,467–469] Similar dimerization of metal (2-formylvinyl)porphyrinates<br />

458 yields hexatriene-bridged bisporphyrins 459 (Scheme 108).<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

N<br />

N<br />

Ni<br />

454<br />

Ni<br />

444<br />

N<br />

N<br />

Scheme 108 Self-Condensation Reactions of Formylporphyrins [394,441,442,462–469]<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

CHO<br />

Et<br />

OH<br />

H2SO4<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

TiCl3(DME) 1.5<br />

N N Et<br />

Zn/Cu, DME<br />

Ni<br />

N N<br />

N N<br />

Ni<br />

Et<br />

Et<br />

Et<br />

N N<br />

Et<br />

Et<br />

457<br />

Et<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

N<br />

N<br />

Ni<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

456<br />

Et<br />

Et<br />

AcOH<br />

N<br />

N<br />

Ni<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et


Et<br />

a458<br />

17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1203<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

Ni<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

CHO<br />

Et<br />

Et<br />

N<br />

N<br />

TiCl 3(DME) 1.5<br />

Zn/Cu, DME<br />

Hydroxymethylporphyrins (e.g., 461, obtained by borohydride reduction of 460) are converted<br />

into the “benzylic” acetoxymethyl derivatives with acetic anhydride/pyridine, and<br />

these react with a range of nucleophiles to give a variety of functionalized porphyrins. [469–<br />

472] Reaction of hydroxymethylporphyrins 461 with thionyl chloride gives chloromethylporphyrins<br />

462 that can be converted into phosphonium salts, such as 463, by treatment<br />

with triphenylphosphine (Scheme 109); [473] compound 463 subsequently undergoes Wittig-type<br />

reactions to yield a variety of functionalized compounds, as well as porphyrin<br />

oligomers linked by, e.g., butadiene and styryl moieties. [454]<br />

Scheme 109 Reactions of Formylporphyrins [470–473]<br />

Ph<br />

N<br />

N<br />

Ph<br />

Ph<br />

Ni<br />

Ph<br />

460<br />

N<br />

N<br />

CHO<br />

N<br />

N<br />

Ph<br />

Ph<br />

462<br />

N<br />

NaBH 4<br />

Ni Ph<br />

Ph<br />

N<br />

Cl<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Ni<br />

N<br />

N<br />

Ph<br />

Et<br />

Et<br />

Ph 3P<br />

Et<br />

Et<br />

N<br />

N<br />

459<br />

Ph<br />

461<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

Ni Ph<br />

Ph<br />

Ph<br />

OH<br />

N<br />

N<br />

Ph<br />

N<br />

N<br />

463<br />

N<br />

N<br />

Ni<br />

N<br />

N<br />

SOCl2<br />

Et<br />

Et<br />

Ni Ph<br />

Ph<br />

Et<br />

Et<br />

PPh 3 + Cl −<br />

for references see p 1223


1204 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aElectrophilic acetylation of metal porphyrinates can be achieved with acetic anhydride in<br />

the presence of Lewis acid catalysts, and occurs preferentially at the more accessible b-positions.<br />

[474] Diacetylation (at positions 3- and 8-) of iron(III) deuteroporphyrin IX was a key<br />

step in Fischer s synthesis of hemin. [124] Acetylation of copper(II) deuteroporphyrin IX dimethyl<br />

ester (447) is accomplished using acetic anhydride and tin(IV) chloride to give 3,8diacetyldeuteroporphyrin<br />

IX dimethyl ester (464) after demetalation; the reaction conditions<br />

can be controlled so that the individual 3- and 8-monoacetyl derivatives 465 and<br />

466, respectively, can be isolated in high yield (Scheme 110). [437,475]<br />

Both b- and meso-acetylation are accomplished on nickel(II) heptaalkylporphyrinate<br />

467, using a large excess of acetic anhydride and tin(IV) chloride, to afford 468 and 469<br />

(Scheme 110); b-acetylation occurs preferentially with the vanadium(IV) and palladium(II)<br />

heptaalkylporphyrinate complexes. [476]<br />

Scheme 110 Friedel–Crafts Acetylations of Metal Porphyrinates [437,475,476]<br />

N<br />

N<br />

Cu<br />

MeO2C CO2Me 447<br />

Ac<br />

NH<br />

N<br />

MeO2C CO2Me 465 48%<br />

N<br />

N<br />

N<br />

HN<br />

1. Ac2O, SnCl4, CH2Cl2<br />

2. TFA, H2SO4<br />

3. CH2N2<br />

MeO2C CO2Me 464 27%<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

NH<br />

N<br />

Ac<br />

NH<br />

N<br />

N<br />

HN<br />

MeO2C CO2Me 466 44%<br />

N<br />

HN<br />

Ac<br />

Ac


a467<br />

17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1205<br />

Et<br />

Et<br />

N<br />

N<br />

Ni<br />

N<br />

N<br />

Et<br />

Ac2O, SnCl4, CH2Cl2<br />

468 82%<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

N<br />

N<br />

Ni<br />

N<br />

N<br />

Et<br />

Et<br />

Ac<br />

N<br />

N<br />

Et<br />

Ni<br />

N<br />

N<br />

469 43%<br />

Acetylporphyrins can be converted into vinylporphyrins by reduction with sodium borohydride<br />

followed by dehydration using 4-toluenesulfonic acid in hot 1,2-dichlorobenzene,<br />

or using benzoyl chloride/dimethylformamide. [477] The enantioselective reduction<br />

of a monoacetyl- and of diacetyldeuteroporphyrin IX dimethyl ester 464 to the corresponding<br />

(R)-(1-hydroxyethyl) (i.e., hematoporphyrin) derivatives is accomplished with<br />

borane dimethyl sulfide in the presence of methyloxazaborolidine as catalyst. [478–480] Acetylporphyrins<br />

464 react with the Vilsmeier reagent producing novel (1-chloro-2-formyl)vinyl<br />

derivatives 470, which can be converted into ethynylporphyrins 471 upon treatment<br />

with base (Scheme 111). [481,482]<br />

+<br />

Ac<br />

Ac<br />

Et<br />

for references see p 1223


1206 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 111 Synthesis of 3,8-Ethynyldeuteroporphyrin [481,482]<br />

MeO2C<br />

Ac<br />

NH<br />

N<br />

464<br />

N<br />

HN<br />

Ac<br />

CO 2Me<br />

POCl 3, DMF<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

base<br />

OHC<br />

MeO2C<br />

MeO 2C<br />

Cl<br />

NH<br />

N<br />

470<br />

N<br />

HN<br />

NH<br />

N<br />

471<br />

N<br />

HN<br />

Cl<br />

CO 2Me<br />

CHO<br />

CO2Me<br />

(2,3,7,8,12,13,17,18-Octaethyl-5-formylporphyrinato)nickel(II) (444); Typical Procedure:<br />

[433]<br />

POCl 3 (1.20 mL, 12.0 mmol) was added dropwise to DMF (2.0 mL, 25.8 mmol) and the mixture<br />

was stirred under N 2 at 08C for 30 min. It was then added to a soln of<br />

(2,3,7,8,12,13,17,18-octaethylporphyrinato)nickel(II) (443; 196 mg, 0.33 mmol) in CH 2Cl 2<br />

(100 mL) at rt and stirred at 388C for 7 h. Sat. aq Na 2CO 3 (100 mL) was added and the soln<br />

was stirred overnight at rt. The mixture was extracted with CH 2Cl 2, the combined organic<br />

layers were washed with H 2O (3 ” 200 mL), dried (Na 2SO 4), and the solvent was removed<br />

under vacuum. The resulting residue was chromatographed (silica gel, petroleum ether/<br />

CH 2Cl 2 2:5). The product was collected and crystallized (CH 2Cl 2/MeOH) to give 444; yield:<br />

190 mg (92%); mp 277–278 8C.<br />

3-Formyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin (3-Formyldeuteroporphyrin<br />

IX Dimethyl Ester, 448) and 8-Formyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin<br />

(8-Formyldeuteroporphyrin IX Dimethyl Ester,<br />

449); Typical Procedure: [439]<br />

TFA (3.0 mL) was added slowly to a suspension of copper(II) deuteroporphyrin IX dimethyl<br />

ester (447; 860 mg, 1.43 mmol) in HC(OMe) 3 (3.0 mL) at rt. During the addition, the mixture<br />

became homogeneous and its color changed from red to green. After further stirring<br />

at rt for 1 h, H 2O (100 mL) was added and the product was extracted with CH 2Cl 2. The organic<br />

layer was washed with aq NaHCO 3, and brine, and then dried by filtration through<br />

cotton wool. After removal of the solvent, the product was chromatographed on a short<br />

column (silica gel, CH 2Cl 2/EtOAc 10:1); concentration of the appropriate eluates yielded<br />

recovered starting material (447; 267 mg) and a mixture of Cu•448 and Cu•449; yield:<br />

601 mg (97% based on recovered starting material). The mixture of Cu•448 and Cu•449<br />

was separated by preparative MPLC. For demetalation of the separate isomers, each was


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1207<br />

atreated with concd H 2SO 4 (5 mL) for 1 h. Subsequent addition of H 2O, extraction with<br />

CH 2Cl 2, re-esterification with CH 2N 2, and crystallization afforded 448; yield: 218 mg (39%<br />

based on consumed 447); mp 231 8C; and 449; yield: 278 mg (50% based on consumed<br />

447); mp 2668C.<br />

[2,3,7,8,12,13,17,18-Octaethyl-5-(2-formylvinyl)porphyrinato]copper(II) (453, M = Cu): [442]<br />

POCl 3 (0.40 mL, 4.0 mmol) was added dropwise to a soln of 3-(dimethylamino)acrolein<br />

(0.40 mL, 4.0 mmol) in dry CH 2Cl 2 (4.0 mL) and the mixture was kept at 08C for 15 min.<br />

This mixture was then added to a soln of (2,3,7,8,12,13,17,18-octaethylporphyrinato)copper(II)<br />

(452, M = Cu; 80.7 mg, 0.135 mmol) in dry CH 2Cl 2 (20.0 mL) with continuous stirring<br />

at 08C. The mixture was then allowed to warm to rt and stirred for 18 h. Sat. aq Na 2CO 3<br />

(100 mL) was then added and the soln was stirred overnight. The mixture was extracted<br />

with CH 2Cl 2, the combined organic layers were washed with H 2O (3 ” 200 mL), dried<br />

(Na 2SO 4), and the solvent was removed under vacuum. The residue was chromatographed<br />

(silica gel, 30% petroleum ether in CH 2Cl 2) and the product was collected and crystallized<br />

(CH 2Cl 2/MeOH) to give 453 (M = Cu); yield: 50.2 mg (57%); mp 230–2318C.<br />

3,8-Diacetyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin (464);<br />

Typical Procedure: [437]<br />

Copper(II) deuteroporphyrin IX dimethyl ester (447; 200 mg, 0.33 mmol) in ice-cold<br />

CH 2Cl 2 (100 mL) was treated with Ac 2O (20 mL) and then with SnCl 4 (1.5 mL). After stirring<br />

for 15 min at 20 8C, the mixture was diluted with ice water and extracted with CH 2Cl 2. The<br />

organic phase was concentrated and the residue was dissolved in concd H 2SO 4 (25 mL) and<br />

left to stand for 1 h at 20 8C. The mixture was worked up once more with CH 2Cl 2 and ice<br />

water and the residue from the organic phase was treated with an excess of ethereal<br />

CH 2N 2, then chromatographed (alumina, CHCl 3/EtOAc 9:1). The red eluates were collected,<br />

concentrated to dryness, and the residue was crystallized (CHCl 3/MeOH) to give 464;<br />

yield: 120 mg (58%); mp 2448C.<br />

3- and 8-Acetyl-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin (465<br />

and 466); Typical Procedure: [475]<br />

Acetylation: Copper(II) deuteroporphyrin IX dimethyl ester (447; 1.92 g, 3.2 mmol) in<br />

CH 2Cl 2 (150 mL) was treated with Ac 2O (40 mL). The mixture was cooled to 0 8C (which<br />

caused the porphyrin to precipitate) before SnCl 4 (4 mL) was added. The resulting homogeneous<br />

green soln was stirred for 20 s before being poured onto ice (200 g). The mixture<br />

was diluted with CH 2Cl 2 (200 mL) and the organic phase was collected and then concentrated<br />

to dryness to give a residue. This was chromatographed [neutral alumina (Brockmann<br />

Grade III), CH 2Cl 2 then 0.5% MeOH in CH 2Cl 2] to recover diacetylated material. The<br />

separated fractions were concentrated to dryness and the residues were crystallized<br />

(CH 2Cl 2/heptane) to give a mixture of Cu•465 and Cu•466; yield: 804 mg (60%). Copper(II)<br />

3,8-diacetyldeuteroporphyrin IX dimethyl ester (Cu•464) was also obtained; yield: 364 mg<br />

(27%).<br />

Separation of isomers: Small quantities of the copper(II) monoacetyl isomers (ca. 30–<br />

50 mg per plate) could be separated on thick LC plates (20 ” 40 cm ” 1 mm, silica gel, 2%<br />

THF in CH 2Cl 2). More convenient large-scale separation (1–2 g) was achieved by means of<br />

MPLC (6.8–10.2 atm). With the aid of an infusion pump, the monoacetyl mixture (ca. 1.5 g)<br />

in CH 2Cl 2 (30 mL) was transferred from a 35-mL plastic syringe at a rate of 2.5 mL•min –1<br />

onto a column of silica gel (2.5 ” 100 cm) which had previously been equilibrated with<br />

25% THF in CH 2Cl 2. Once all the soln had been applied to the column the compounds<br />

were eluted with 2% THF in CH 2Cl 2, a process usually requiring 6–10 h before the first fraction<br />

began to exit the system. An additional 3–6 h was required for the compounds to<br />

completely pass through the column. The effluent was collected in 15-mL aliquots using<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1208 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aa Gilson microfractionator, and each aliquot was assayed by means of analytical HPLC<br />

(microporasil column, 25 ” 0.4 cm, 1% THF in CH 2Cl 2).<br />

Demetalation: To the concentrated fractions containing each isomer was added TFA<br />

(30 mL) and concd H 2SO 4 (3.0 mL). The mixtures were allowed to stand for 1 h at rt before<br />

each was diluted with CH 2Cl 2 (100 mL) and washed with H 2O. The organic phases were<br />

treated briefly with an excess of ethereal CH 2N 2, and concentrated to dryness. After this,<br />

the residues were taken up in CH 2Cl 2 and individually passed through a plug of neutral<br />

alumina (Brockmann Grade III) eluting with CH 2Cl 2. Evaporation of the solvent and crystallization<br />

gave separated 465 and 466. In a typical separation, copper(II) complex mixture<br />

(1.538 g) gave 465 (the more mobile fraction); yield: 658 mg (48%); mp 238–2398C,<br />

and 466 (the less mobile fraction); yield: 608 mg (44%); mp 214–2168C. Total recovery 91%.<br />

(3,5-Diacetyl-7,13,18-triethyl-2,8,12,17-tetramethylporphyrinato)nickel(II) (469); Typical<br />

Procedure: [476]<br />

(7,13,18-Triethyl-2,8,12,17-tetramethylporphyrinato)nickel(II) (467; 20 mg, 0.04 mmol) in<br />

CH 2Cl 2 (20 mL) and Ac 2O (4.5 mL) was cooled to 0 8C and treated with SnCl 4 (1.2 mL) (molar<br />

ratio of 467/Ac 2O/SnCl 4 1:1200:260). After 15 min, the green soln was diluted with CH 2Cl 2<br />

(15 mL), hydrolyzed with ice for 1.5 h, and washed with H 2O. The organic phase was dried<br />

(Na 2SO 4) and concentrated to dryness. Chromatography (silica gel, CH 2Cl 2/hexanes 2:1)<br />

gave the nickel(II) diacetylporphyrinate 469, which was crystallized (CH 2Cl 2/MeOH); yield:<br />

10 mg (43%).<br />

(3-Acetyl-7,13,18-triethyl-2,8,12,17-tetramethylporphyrinato)nickel(II) (468); Typical Procedure:<br />

[476]<br />

(7,13,18-Triethyl-2,8,12,17-tetramethylporphyrinato)nickel(II) (467; 18 mg, 0.035 mmol)<br />

in CH 2Cl 2 (20 mL) and Ac 2O (0.3 mL) was cooled to 0 8C and treated with SnCl 4 (0.08 mL)<br />

(molar ratio of 467/Ac 2O/SnCl 4 1:80:17). After 2.25 min the green soln was hydrolyzed<br />

with ice for 1.5 h, and washed with H 2O. The organic phase was dried (Na 2SO 4) and concentrated<br />

to dryness. Chromatography (silica gel, CH 2Cl 2) gave the nickel(II) acetylporphyrinate<br />

468, which was crystallized (CH 2Cl 2/MeOH); yield: 16 mg (82%).<br />

17.8.4.1.4 Variation 4:<br />

Peripheral Metalation<br />

Mercuration [e.g., of zinc(II) deuteroporphyrin IX dimethyl ester (472)] occurs preferentially<br />

at the least sterically hindered b-positions to produce metalated derivatives (e.g.,<br />

473), which undergo transmetalation in the presence of lithium trichloropalladate and<br />

alkenes (such as methyl acrylate, acrolein, styrene, vinylnaphthalene and vinylferrocene)<br />

to provide a variety of interesting substituted porphyrins, such as 474, after demetalation<br />

(Scheme 112). [483–486] Mercurated porphyrins also react with bromine/trichloromethane<br />

and iodine/sodium iodide, producing the corresponding halogenated derivatives in good<br />

yields. [485] In a similar way, vinylporphyrins react with arylmercurials in the presence of<br />

lithium trichloropalladate to produce the corresponding styrene derivatives. [487,488]<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1209<br />

aScheme 112 Synthesis of Porphyrin Acrylates via Mercuration [483–486]<br />

N<br />

N<br />

Zn<br />

MeO2C CO2Me 472<br />

N<br />

N<br />

1.<br />

1. Hg(OAc) 2, MeOH, THF<br />

2. NaCl<br />

CO2Me<br />

LiPdCl 3, MeCN<br />

Et3N, DMSO, THF<br />

2. TFA<br />

37%<br />

MeO2C CO2Me 473<br />

MeO2C<br />

MeO2C CO2Me 474<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

ClHg<br />

NH<br />

N<br />

N<br />

N<br />

N<br />

HN<br />

Zn<br />

N<br />

N<br />

HgCl<br />

CO2Me<br />

{3,8-Bis(chloromercurio)-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrinato}zinc(II)<br />

(473); Typical Procedure: [483]<br />

Under a dry N 2 atmosphere, Hg(OAc) 2 (1.8 g) in MeOH (25 mL) was added rapidly but dropwise<br />

to zinc(II) deuteroporphyrin IX dimethyl ester (472; 860 mg, 1.43 mmol) in dry THF<br />

(100 mL) with stirring at 60 8C. After 5 h, sat. aq NaCl (100 mL) was added and the biphasic<br />

mixture was vigorously stirred for 10 min while the soln cooled. The mixture was diluted<br />

with CH 2Cl 2 (100 mL) and the organic phase containing a large amount of precipitated<br />

porphyrin was collected and washed with H 2O (4 ” 125 mL). The organic layer was separated<br />

and concentrated to give a residue, which was triturated with EtOH (95%, 50 mL) and<br />

brought to a gentle boil. The dark metal porphyrinate was scraped from the side of the<br />

flask and collected by filtration. The product was dried at rt under high vacuum to give<br />

473; yield: 1.543 g (contaminated with some trimercurated material) which did not melt.<br />

3,8-Bis[2-(methoxycarbonyl)vinyl]-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin<br />

(474); Typical Procedure: [483]<br />

3,8-Bis(chloromercurio)-13,17-bis[2-(methoxycarbonyl)ethyl]-2,7,12,18-tetramethylporphyrin<br />

473 (609 mg, 0.70 mmol) in dry DMSO (25 mL) was treated with dry THF (25 mL)<br />

and freshly distilled methyl acrylate (10 mL). The mixture was stirred under N 2 at 508C.<br />

Following addition of Et 3N (0.5 mL), a soln of LiPdCl 3 in MeCN (15 mL) [prepared by refluxing<br />

for 30 min PdCl 2 (250 mg) in MeCN (15 mL) containing LiCl (20 mg)] was added dropwise<br />

but rapidly to the porphyrin soln. As the reaction proceeded the mixture became<br />

dark and reduced Pd metal was observed. After 40 min the mixture was removed from<br />

the heat, allowed to cool for 15 min, and then filtered through a 4-cm thick pad of Celite.<br />

The pad was then rinsed with THF/CH 2Cl 2 and the combined filtrates were washed with<br />

H 2O (3 ” 100 mL), dried (Na 2SO 4), concentrated to dryness, and then placed under high vacuum<br />

to remove a small amount of residual oil. The residue was dissolved in TFA (30 mL),<br />

and this soln was stirred for 5 min and then diluted with CH 2Cl 2 (100 mL) before being<br />

for references see p 1223


1210 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

awashed with H 2O (3 ” 100 mL). Concentration of the organic fraction gave a residue,<br />

which was chromatographed on thick-layer plates (silica gel, 3.5% MeOH in CH 2Cl 2). Three<br />

major bands were eluted. The least mobile fraction was crystallized (CH 2Cl 2/heptane) to<br />

give the bisacrylate 474; 146 mg (37%); mp 259–2618C.<br />

17.8.4.1.5 Variation 5:<br />

Reactions with Carbenes and Nitrenes<br />

Electrophilic additions of carbenes and nitrenes to porphyrins [e.g., (2,3,7,8,12,13,17,18octaethylporphyrinato)zinc(II)<br />

(475)] tend to occur at the inner nitrogen atoms with formation<br />

of N-substituted porphyrin derivatives 476. Metal chelation at the central nitrogens<br />

induces intramolecular migration of the N-substituent to the macrocyclic periphery,<br />

affording chlorins and meso-substituted porphyrins (e.g., 477, Scheme 113). Ring expanded<br />

meso-homoporphyrins are believed to be intermediates in the formation of the mesosubstituted<br />

porphyrins. [489–495] Treatment of 2,3,7,8,12,13,17,18-octaethylporphyrin (21)<br />

with O-mesitylsulfonylhydroxylamine (OMSH) affords the stable N-aminoporphyrin 478<br />

(Scheme 113). [496] Examples of the formation of cyclopropanechlorins by additions of carbenes<br />

to a porphyrin peripheral b—b¢ double bond are also known. [306,497]<br />

Scheme 113 Carbene/Nitrene Reactions with Porphyrins [489–496]<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

Zn<br />

475<br />

N<br />

N<br />

Et<br />

Et Et<br />

Et<br />

NH<br />

N<br />

21<br />

N<br />

HN<br />

Et<br />

Et Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

1. N2CH2CO2Et 2. HCl<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

45%<br />

OMSH, CHCl3<br />

41%<br />

Et<br />

Et<br />

Ni(acac)2, benzene<br />

39%<br />

Et<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

476<br />

N<br />

CO 2Et<br />

HN<br />

Et<br />

Et Et<br />

Et<br />

Et<br />

N N<br />

NH2 N<br />

478<br />

HN<br />

Et<br />

N<br />

N<br />

Et<br />

Et<br />

Ni<br />

CO2Et Et<br />

N<br />

N<br />

Et Et<br />

477<br />

Et<br />

Et Et<br />

Et<br />

Et<br />

Et<br />

Et


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1211<br />

aN-Ethoxycarbonylmethyl-2,3,7,8,12,13,17,18-octaethylporphyrin (476); Typical Procedure:<br />

[492]<br />

A refluxing soln of (2,3,7,8,12,13,17,18-octaethylporphyrinato)zinc(II) (475; 340 mg,<br />

0.57 mmol) in bromobenzene (10 mL) was treated dropwise with ethyl diazoacetate<br />

(0.7 mL, 6.1 mmol) over 12 min and then refluxed for a further 30 min. The solvent was<br />

removed under vacuum and the residue was dissolved in CH 2Cl 2 (40 mL) and treated<br />

with concd aq HCl (4 mL). The soln was neutralized with aq (NH 4) 2CO 3, washed with H 2O<br />

(” 2), and concentrated to dryness. Chromatography [alumina (200 g), toluene/cyclohexane<br />

1:1] gave some 2,3,7,8,12,13,17,18-octaethylporphyrin (21), followed by 476, which<br />

was crystallized (CHCl 3/MeOH); yield: 160 mg (45%).<br />

(5-Ethoxycarbonylmethyl-2,3,7,8,12,13,17,18-octaethylporphyrinato)nickel(II) (477);<br />

Typical Procedure: [492]<br />

N-Ethoxycarbonylmethyl-2,3,7,8,12,13,17,18-octaethylporphyrin (476; 40mg,<br />

0.064 mmol) in benzene (12 mL) (CAUTION: carcinogen) was treated with nickel(II) acetylacetonate<br />

(120 mg) and refluxed for 12 h. The solvent was removed under vacuum and the<br />

residue was chromatographed (alumina or silica gel, benzene), to give<br />

(2,3,7,8,12,13,17,18-octaethylporphyrinato)nickel(II) (443); yield: 6 mg, followed by the<br />

5-substituted porphyrin. Crystallization (CH 2Cl 2/MeOH) gave 477; yield: 17 mg (39%); mp<br />

207–2088C.<br />

N-Amino-2,3,7,8,12,13,17,18-octaethylporphyrin (478); Typical Procedure: [496]<br />

An excess of O-mesitylsulfonylhydroxylamine (OMSH; 500 mg) and 2,3,7,8,12,13,17,18-octaethylporphyrin<br />

(21; 500 mg, 0.93 mmol) were stirred for 16 h at 208C in CHCl 3. Concentration<br />

and chromatography of the residue [alumina (100 g), CH 2Cl 2] gave recovered<br />

2,3,7,8,12,13,17,18-octaethylporphyrin (21); yield: 216 mg, and a green fraction. Crystallization<br />

of the green fraction (MeOH) gave 478; yield: 120 mg (41% based on recovered 21).<br />

17.8.4.2 Method 2:<br />

Reactions with Nucleophiles<br />

The porphyrin macrocycle also undergoes reactions with nucleophiles. As would be expected,<br />

electron-withdrawing peripheral substituents and electronegative chelated metal<br />

ions facilitate nucleophilic attack on the porphyrin ring. For example, 2-nitro-5,10,15,20tetraarylporphyrins<br />

and their metal complexes display useful chemical reactivity at the 2and<br />

3-positions, undergoing ipso-substitution of the nitro group [243,412,422,424] and nucleophilic<br />

Michael-type additions [242,243,411,425–429] with a variety of nucleophiles, as previously<br />

mentioned. Several functionalized macrocycles, such as chlorins, 2-substituted porphyrins<br />

and 3-substituted 2-nitroporphyrins and chlorins have been obtained by using this<br />

methodology. [498]<br />

The most studied nucleophilic addition reactions occur on the p-cation radicals of<br />

metal porphyrinates, which can be obtained by mild chemical or electrochemical oxidation<br />

(Scheme 114). The magnesium(II) and zinc(II) complexes of porphyrins are usually<br />

employed since these display the lowest oxidation potentials, and iodine (usually in the<br />

presence of silver perchlorate or silver hexafluorophosphate), tris(4-bromophenyl)ammoniumyl<br />

hexachloroantimonate, or N-chlorobenzotriazole are the oxidants of choice. Although<br />

relatively stable in methanol, the p-cation radicals of metal porphyrinates react<br />

with a variety of soft nucleophiles to produce the corresponding meso- and b-substituted<br />

metal porphyrinates. For example, meso-nitroporphyrins (e.g., 480) are obtained by treatment<br />

of metal porphyrinates [e.g., dipyridine magnesium(II) etioporphyrin I (479)] with<br />

nitrogen dioxide, [408–411] iodine/silver nitrite, [397,412–414] or with thallium(III) nitrate or cerium(IV)<br />

ammonium nitrate. [499] Pyridinium porphyrin salts (e.g., 481) are obtained by reaction<br />

of p-cation radicals from, for example, (2,3,7,8,12,13,17,18-octaethylporphyrinato)-<br />

for references see p 1223<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1212 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

azinc(II) (475), with various pyridines. [500–504] Other nucleophiles (cyanide, thiocyanate,<br />

chloride, acetate, imidazole, triphenylphosphine) also react with the p-cation radicals of<br />

porphyrins to give, for example, 5-chloro-2,3,7,8,12,13,17,18-octaethylporphyrin<br />

(482); [505] dimeric and higher oligomeric porphyrin systems have been obtained by the<br />

use of bidentate nucleophiles, such as bipyridine. [506–508]<br />

Scheme 114 Reactions of Porphyrin p-Cation Radicals with Nucleophiles [397,412–414,500–505]<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

N<br />

N<br />

py<br />

Mg<br />

py<br />

479<br />

Zn<br />

475<br />

N<br />

N<br />

N<br />

N<br />

Et<br />

Et<br />

Et Et<br />

Et<br />

Et<br />

Et<br />

1. I2, MeCN, CH 2Cl 2<br />

2. NaNO 2, MeCN<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

98%<br />

1. Tl(NO3) 3, THF, MeOH<br />

2. py, SO2 (g)<br />

60%<br />

1. ammoniumyl salt<br />

THF, CH2Cl2 2. Et4NCl, CH2Cl2 31%<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

NH<br />

N<br />

NO 2<br />

480<br />

Et<br />

N<br />

HN<br />

NH<br />

N<br />

Et<br />

N +<br />

481<br />

N<br />

HN<br />

Et<br />

Cl −<br />

Et<br />

Et Et<br />

Et<br />

NH<br />

N<br />

Cl<br />

482<br />

N<br />

HN<br />

Et<br />

Et Et<br />

Chemical and electrochemical oxidative couplings of zinc(II) 5,15-diarylporphyrinates<br />

lead to the formation of meso–meso and meso–b linked porphyrin arrays, through nucleophilic<br />

attack of the p-cation radicals by neutral porphyrins; [414,509] (5,10,15-triphenylporphyrinato)zinc(II)<br />

(483) produces dimer 484 in the presence of silver hexafluorophosphate<br />

(Scheme 115). [510] Treatment of zinc(II) 5,15-diarylporphyrinates with silver hexafluorophosphate<br />

and iodine in the presence of pyridine leads to regioselective meso-iodination.<br />

[511]<br />

Et<br />

Et<br />

Et<br />

Et


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1213<br />

aScheme 115 Radical Dimerization of Porphyrins [510]<br />

Ph<br />

N<br />

N<br />

Ph<br />

Zn<br />

Ph<br />

483<br />

N<br />

N<br />

AgPF6<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Ph<br />

N<br />

N<br />

Ph<br />

Zn<br />

Ph<br />

N<br />

N<br />

484<br />

Ph<br />

N N<br />

Zn<br />

N N<br />

The more electrophilic p-dications of porphyrins are formed under stronger oxidation<br />

conditions and react with, for example, methanol, preferentially at the meso positions,<br />

leading to isoporphyrin derivatives, such as 486 from (5,10,15,20-tetraphenylporphyrinato)zinc(II)<br />

(485) (Scheme 116). [512,513]<br />

Scheme 116 Reaction of p-Dications of Porphyrins with Methanol [512,513]<br />

Ph<br />

N<br />

N<br />

Ph<br />

Zn<br />

Ph<br />

485<br />

N<br />

N<br />

Ph<br />

2e − oxidation<br />

MeOH<br />

Ph<br />

Ph OMe<br />

Reactions of organometallic reagents with iron(III) and cobalt(III) porphyrins occur at the<br />

metal center with formation of s-bonded species; [514,515] migration of the s-bonded groups<br />

(alkyl, vinyl, aryl) from the metal ion to the nitrogen atom is induced by oxidation or by<br />

acid treatment. [516–522] Nucleophilic attack upon unsubstituted meso positions gives phlorins<br />

[523] and porphodimethenes, [524] which can be oxidized to the corresponding meso-substituted<br />

porphyrins; for example, (2,3,7,8,12,13,17,18-octaethylporphyrinato)nickel(II)<br />

(443) is mono- (487), di- (488, 489), tri- (490), and tetra-meso-alkylated (491) by successive<br />

additions of alkyllithium at low temperature, followed by hydrolysis and oxidation with<br />

2,3-dichloro-5,6-dicyanobenzo-1,4-quinone (Scheme 117). 5,5¢-Linked bisporphyrins are<br />

isolated when the hydrolysis step is omitted. [525]<br />

N<br />

N<br />

+<br />

Zn<br />

Ph<br />

486<br />

N<br />

N<br />

Ph<br />

Ph<br />

Ph<br />

for references see p 1223


1214 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

aScheme 117 Addition of Butyllithium to (2,3,7,8,12,13,17,18-<br />

Octaethylporphyrinato)nickel(II) [525]<br />

Et<br />

Et<br />

Et Et<br />

Et<br />

BuLi<br />

N<br />

N<br />

Ni<br />

443<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Bu<br />

Et<br />

Et<br />

Et<br />

Et Bu Et<br />

Et<br />

N<br />

N<br />

Ni<br />

488<br />

N<br />

N<br />

490<br />

100%<br />

Et<br />

Et Bu Et<br />

Et<br />

Et<br />

N<br />

N<br />

Ni<br />

Bu<br />

1. BuLi, THF<br />

2. DDQ<br />

N<br />

N<br />

Et<br />

Et Bu Et<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Bu<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

N<br />

N<br />

N<br />

N<br />

Ni<br />

487<br />

Ni<br />

489<br />

N<br />

N<br />

N<br />

N<br />

Et<br />

Et Bu Et<br />

Et<br />

Bu<br />

Et Et<br />

Et<br />

N N<br />

BuLi<br />

Bu<br />

Ni<br />

Bu<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et Bu Et<br />

meso-Tetrasubstituted porphyrins react with alkyllithiums to produce porphodimethenes,<br />

[524] phlorins, [526] and chlorins. [526] Nucleophilic addition of hydroxide ion to<br />

gold(III) 5,10,15,20-tetraphenylporphyrinate (492) occurs at a meso position with formation<br />

of hydroxyphlorin 493 (Scheme 118), but the copper(II), palladium(II), cadmium(II),<br />

and manganese(III) complexes of 5,10,15,20-tetraphenylporphyrin (22) are unreactive under<br />

the same conditions. [527] A regioselective 15-meso-alkylation of (5-formyl-<br />

2,3,7,8,12,13,17,18-octaethylporphyrinato)zinc(II) (494) to produce 495 via a readily oxidizable<br />

phlorin intermediate is accomplished by addition of methyl (or ethyl) magnesium<br />

bromide (Scheme 118). [528] Under similar conditions the metal-free and nickel(II) derivatives<br />

react preferentially at the meso-formyl group, as might be expected, to yield the corresponding<br />

1-hydroxyalkyl derivatives.<br />

+<br />

Et<br />

N<br />

Bu<br />

491<br />

N<br />

BuLi<br />

Et<br />

Et


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1215<br />

aScheme 118 Reactions of Metal Porphyrinates with Nucleophiles [528]<br />

N Cl N<br />

Ph Au Ph<br />

Et<br />

Et<br />

N<br />

Ph<br />

Ph<br />

492<br />

N<br />

Et CHO Et<br />

N<br />

N<br />

Zn<br />

494<br />

N<br />

N<br />

Et Et<br />

Et<br />

Et<br />

1. NaBH4, MeOH<br />

2. NaOH<br />

MeMgBr<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Ph<br />

N<br />

N<br />

Ph<br />

493<br />

OH<br />

N<br />

Ph Au Ph<br />

Et CHO Et<br />

5-Nitro-2,7,12,17-tetraethyl-3,8,13,18-tetramethylporphyrin (480); Typical Procedure: [505]<br />

Dipyridine (2,7,12,17-tetraethyl-3,8,13,18-tetramethylporphyrinato)magnesium(II) (479;<br />

50 mg, 0.076 mmol) in CH 2Cl 2 (50 mL) was treated with a soln of I 2 (21 mg, 0.08 mmol) in<br />

MeCN (10 mL) and CH 2Cl 2 (40 mL). The mixture was then treated with NaNO 2 (100 mg,<br />

0.65 mmol) in MeCN (10 mL), and the resulting red porphyrin soln was washed immediately<br />

with H 2O. The organic phase was dried (Na 2SO 4) and concentrated to dryness before<br />

addition of TFA (1 mL) and neutralization by washing several times with H 2O. The product<br />

was chromatographed [short neutral alumina column (Brockmann Grade III), CH 2Cl 2]. The<br />

red-brown eluates were collected and concentrated, and the residue was crystallized<br />

(CH 2Cl 2/MeOH) to give 5-nitroetioporphyrin I (480); yield: 39 mg (98%); mp >3008C.<br />

5-N-Pyridinium-2,3,7,8,12,13,17,18-octaethylporphyrin Chloride (481); Typical Procedure:<br />

[505]<br />

A soln of (2,3,7,8,12,13,17,18-octaethylporphyrinato) zinc(II) (475; 300 mg, 0.50 mmol) in<br />

THF (30 mL) was flushed with N 2 for 10 min prior to addition of Tl(NO 3) 3 (234 mg,<br />

0.60 mmol) in MeOH (25 mL). After stirring for 30 s, an excess of pyridine (10 mL) was added.<br />

The mixture was stirred for 30 min and then SO 2(g) was bubbled into the soln. HCl<br />

(20 mL) was added to the residue obtained by evaporation of the solvents, and after stirring<br />

for 5 min, the mixture was poured into CH 2Cl 2 (300 mL), washed with brine<br />

(300 mL), dried (CaCl 2), and concentrated. The residue was chromatographed [alumina<br />

(Brockmann Grade V), CHCl 3]. A red forerun contained 2,3,7,8,12,13,17,18-octaethylporphyrin<br />

(21; 17 mg). Further elution (CHCl 3/MeOH 20:1) gave red eluates which were washed<br />

with brine, dried (CaCl 2), and concentrated. Crystallization (CH 2Cl 2/hexanes) gave 481;<br />

yield: 157 mg (60% based on consumed 2,3,7,8,12,13,17,18-octaethylporphyrin); mp<br />

>300 8C.<br />

5-Chloro-2,3,7,8,12,13,17,18-octaethylporphyrin (482); Typical Procedure: [505]<br />

A soln of (2,3,7,8,12,13,17,18-octaethylporphyrinato)zinc(II) (475; 300 mg, 0.50 mmol) in<br />

CH 2Cl 2 (100 mL) and THF (30 mL) was treated with tris(4-bromophenyl)ammoniumyl hexachloroantimonate<br />

(902 mg, 1.10 mmol) in THF (30 mL) and CH 2Cl 2 (100 mL). This soln<br />

was treated slowly with Et 4NCl (375 mg, 2.26 mmol) in CH 2Cl 2 (50 mL) and the mixture<br />

Et<br />

Et<br />

N<br />

N<br />

Zn<br />

495<br />

N<br />

N<br />

Et<br />

N<br />

Et<br />

Et<br />

for references see p 1223


1216 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

awas stirred for 2 h. A soln of HCl (5 mL) in THF (10 mL) was added and, after stirring for<br />

5 min, the mixture was poured into H 2O. The organic phase was washed with H 2O<br />

(300 mL), dried (Na 2SO 4), and concentrated to dryness. Chromatography of the residue<br />

on TLC plates (silica gel, petroleum ether/toluene 3:2) gave three bands. The middle<br />

band contained 482; yield: 99 mg (31%); mp 270–2728C.<br />

(5-Butyl-2,3,7,8,12,13,17,18-octaethylporphyrinato)nickel(II) (487); Typical Procedure: [524]<br />

(2,3,7,8,12,13,17,18-Octaethylporphyrinato)nickel(II) (443; 100 mg, 0.17 mmol) in THF<br />

(60 mL) at –70 8C was treated, immediately after cooling, dropwise with 2 M BuLi in cyclohexane<br />

(0.6 mmol). The cooling bath was removed and H 2O (1 mL) in THF (5 mL) was added<br />

dropwise. The mixture was stirred for 10 min, 0.06 M DDQ in CH 2Cl 2 (10 mL) was added,<br />

and the mixture was stirred for another 20 min. The mixture was filtered [neutral alumina<br />

(Brockmann Grade I), CH 2Cl 2] and then chromatographed [neutral alumina (Brockmann<br />

Grade III), hexanes/CH 2Cl 2 4:1] to give a quantitative yield of 487. Preparation of<br />

higher alkylated porphyrins (e.g., 488–491) required lower temperatures (–80 to –1008C)<br />

and less solvent.<br />

17.8.4.3 Method 3:<br />

Oxidation Reactions<br />

Oxidations of porphyrins and metal porphyrinates using osmium tetroxide to give vicinal<br />

diols 340 and 2-oxochlorins 341 have been mentioned previously (Section 17.8.2.1.3).<br />

Porphyrins can also be oxidized at the meso positions to afford, for example, oxophlorins<br />

(e.g., 497) and oxochlorins (e.g., 498). 5-Oxophlorins (such as 497) are produced<br />

from hydrolysis of metal meso-trifluoroacetoxyporphyrinates 496, obtained from zinc(II)<br />

and magnesium(II) b-substituted porphyrinates (e.g., 475) by reaction with thallium(III)<br />

trifluoroacetate. [529,530] Benzoyl peroxide also accomplishes meso-oxidation, although less<br />

selectively. [531,532] More extensive meso-oxidation leads to the formation of di-, tri-, and tetraoxoporphyrins<br />

(the so-called xanthoporphyrinogens, 499) (Scheme 119).<br />

Scheme 119 Syntheses of meso-Oxoporphyrin Systems [529,530]<br />

Et<br />

Et<br />

Et Et<br />

N<br />

N<br />

Zn<br />

N<br />

N<br />

Et Et<br />

475<br />

Et<br />

Et<br />

Tl(OCOCF 3) 3<br />

THF, CH2Cl2<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

Et<br />

SO2, H3O +<br />

Et Et<br />

Et<br />

N<br />

N<br />

Zn<br />

496<br />

Et<br />

Et<br />

N<br />

N<br />

Et<br />

Et<br />

O<br />

Et O<br />

NH<br />

N<br />

HN<br />

HN<br />

497 79%<br />

CF 3<br />

Et Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

O


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1217<br />

Et<br />

O<br />

Et<br />

Et Et<br />

Et<br />

NH<br />

N<br />

HN<br />

HN<br />

a499<br />

498<br />

Et<br />

Et<br />

Et<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Et<br />

O<br />

Et<br />

Et O Et<br />

2,3,7,8,12,13,17,18-Octaethyl-5-oxophlorin (497); Typical Procedure: [530]<br />

(2,3,7,8,12,13,17,18-Octaethylporphyrinato)zinc(II) (475; 415 mg, 0.69 mmol) in THF<br />

(30 mL) and CH 2Cl 2 (100 mL) was treated with a dry soln of thallium(III) trifluoroacetate<br />

(480 mg, 0.88 mmol) in THF (20 mL) and then stirred for 1 min to give 496. H 2O (0.25 mL)<br />

in THF (10 mL) was added and, after stirring for 10 min, the soln was treated briefly with<br />

SO 2(g). Concd HCl (2 mL) was added and the soln was stirred for 5 min before being poured<br />

into H 2O (250 mL) and extracted with CH 2Cl 2 (250 mL). The organic phase was washed<br />

with H 2O (2 ” 500 mL), dried (Na 2SO 4), and concentrated to dryness. The residue was chromatographed<br />

[neutral alumina (200 g), CH 2Cl 2]. A small forerun containing pink material<br />

was discarded and the blue eluates were collected and concentrated to dryness. Crystallization<br />

[CH 2Cl 2/MeOH] gave 497; yield: 302 mg (79%); mp 254–2568C.<br />

17.8.4.4 Method 4:<br />

Cycloaddition Reactions<br />

17.8.4.4.1 Variation 1:<br />

Intermolecular Reactions<br />

Numerous cycloaddition reactions of porphyrins are described in Section 17.8.2.3.2. Pyrroloporphyrin<br />

500, obtained from cleavage and decarboxylation of the benzyl ester 442<br />

(Scheme 104), reacts with dimethyl acetylenedicarboxylate in refluxing 1,2,4-trichlorobenzene<br />

to produce benzoporphyrin 501 as the main product (Scheme 120), along with<br />

a bis-adduct. [533] A N-tert-butoxycarbonyl-protected derivative of 500 also undergoes cycloaddition<br />

reactions with dimethyl acetylenedicarboxylate. [534]<br />

Porphyrins (e.g., 22) react with pyrrolo-fused 3-sulfolene 502 [291] and related systems,<br />

[535] to afford pyrrolochlorins 503, pyrroloporphyrins 504, isoindoloporphyrins<br />

505, and dipyrrolobacteriochlorins (Scheme 120). [536] A porphyrin bearing a b-fused 3-sulfolene<br />

has also been reported; this reacts with dienophiles to give the corresponding<br />

Diels–Alder adducts. [537]<br />

Scheme 120 Intermolecular Cycloadditions [533,536]<br />

Ph<br />

N<br />

N<br />

Ph<br />

Ni<br />

Ph<br />

500<br />

N<br />

N<br />

NH<br />

Ph<br />

MeO2C<br />

heat<br />

Et<br />

CO2Me<br />

NH<br />

NH<br />

O<br />

HN<br />

HN<br />

Ph<br />

Et<br />

Et<br />

Et<br />

O<br />

N<br />

N<br />

Ph<br />

Ni<br />

Ph<br />

N<br />

N<br />

501<br />

CO2Me<br />

Ph<br />

CO2Me<br />

for references see p 1223


1218 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

Ph<br />

NH<br />

N<br />

a503<br />

22<br />

Ph<br />

Ph<br />

Ph<br />

N<br />

HN<br />

NH<br />

N<br />

Ph<br />

Ph<br />

Ph<br />

N<br />

HN<br />

504<br />

17.8.4.4.2 Variation 2:<br />

Intramolecular Reactions<br />

+<br />

O O<br />

S<br />

N<br />

H<br />

502<br />

Ph<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

CO 2R 1<br />

NH<br />

, heat<br />

CO 2R 1<br />

+<br />

Ph<br />

Ph<br />

NH<br />

N<br />

NH<br />

N<br />

Ph<br />

Ph<br />

Ph<br />

Ph<br />

N<br />

HN<br />

N<br />

HN<br />

505<br />

Ph<br />

Ph<br />

NH<br />

NH<br />

CO 2R 1<br />

CO 2R 1<br />

Acid-catalyzed intramolecular cyclizations of 5-[2-(ethoxycarbonyl)vinyl]porphyrins 506<br />

lead to purpurins (e.g., 507) in the absence of a chelated metal ion (Scheme 121). When<br />

nickel(II) or copper(II) 5-(2-formylvinyl)porphyrinates (e.g., 453) are used, benzochlorins<br />

such as 508 result. [212,441,442] The metal-free benzochlorins 509 are obtained either by demetalation<br />

or by acid-catalyzed cyclization of meso-(2-hydroxymethyl)vinylporphyrins<br />

(Scheme 121). [538] meso-Substituted benzochlorins, [539–541] benzoisobacteriochlorins, [442]<br />

and benzobacteriochlorins [442] are similarly produced. Hydroxybenzochlorins and oxobenzochlorins<br />

(e.g. 511, obtained from 510 by reaction with borontrifluoride–diethyl<br />

ether complex) [542] are produced when the electrophilic cyclization is directed toward an<br />

unsubstituted b-position. [441,442] Methods for the synthesis of meso-(2-formylvinyl)porphyrins<br />

were mentioned earlier; the 5-[2-(ethoxycarbonyl)vinyl]porphyrins (e.g., 506) are usually<br />

prepared by treatment of nickel(II) 5-formyloctaethylporphyrinate 444 (Scheme 105)<br />

with (carbethoxymethylene)triphenylphosphorane, followed by demetalation. [213] The<br />

thermodynamically favored trans-purpurins are obtained from regioselective cyclizations,<br />

in refluxing acetic acid, of 5-(2-formylvinyl) or meso-acrylic ester porphyrins. [212–<br />

214,218] Cyclizations, using triethylamine or potassium hydroxide/methanol, have been<br />

used for the preparation of 5,15-disubstituted purpurins. [216,217] Macrocycles containing<br />

both purpurin and benzochlorin species have also been obtained from porphyrin precursors<br />

bearing both meso-(2-formylvinyl) and meso-acrylic ester substituents; [543] meso-[3-(dimethylamino)prop-1-enyl]porphyrin<br />

512 undergoes thermal intramolecular cyclization<br />

to give the ethylenechlorin 513 as a Z/E mixture (Scheme 121). [544] Porphyrin 512 is obtained<br />

by electrophilic substitution of nickel(II) octaethylporphyrinate 443 with 3dimethylacetamide/phosphoryl<br />

chloride followed by reduction of the resulting iminium<br />

salt with sodium borohydride and final quaternization with iodomethane.<br />

b-Vinylporphyrins undergo acid-catalyzed cyclization onto either a vicinal meso position<br />

[545] (514fi515) or to an adjacent 2-phenyl position (516fi517) (Scheme 121). [546]<br />

Naphthochlorin 519, can be obtained [547] from 518 by reduction with lithium aluminum


17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1219<br />

ahydride/aluminum trichloride followed by treatment with base. Porphyrin 518 is in turn<br />

obtained by acid-catalyzed cyclization of the 2-formyl group of 446 onto one of the ortho<br />

positions of the adjacent phenyl group. [434–436,548,549] A related cyclization of a b-methoxycarbonyl<br />

group in the presence of acetic acid, copper(II) acetate and sodium acetate has<br />

also been described. [550] A naphthochlorin 520 is produced from condensation of tert-butyl<br />

isocyanoacetate with (2-nitro-5,10,15,20-tetraphenylporphyrinato)nickel(II) 441; it<br />

subsequently undergoes radical dimerization in the presence of air, or with benzoyl peroxide<br />

to afford 521 (Scheme 121). [394,551]<br />

Scheme 121 Intramolecular Cyclizations<br />

Et<br />

Et<br />

Et<br />

Et<br />

M = Cu, Ni<br />

Et Et<br />

NH<br />

N<br />

N<br />

HN<br />

Et Et<br />

N<br />

N<br />

506<br />

Et Et<br />

Et Et<br />

453<br />

Ar 1<br />

Cu<br />

Ar 2<br />

510<br />

N<br />

N<br />

N<br />

N<br />

M<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

Et<br />

CHO<br />

CO 2Et<br />

CHO<br />

BF3 OEt2<br />

AcOH, heat<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

68%<br />

18% H2SO4/TFA<br />

H +<br />

Et<br />

Et<br />

N<br />

N<br />

Ar 1<br />

Cu<br />

Ar 2<br />

511<br />

Et Et<br />

NH<br />

N<br />

N<br />

HN<br />

Et Et<br />

Et<br />

Et<br />

N<br />

N<br />

Et<br />

Et<br />

Et<br />

507<br />

N<br />

N<br />

Et<br />

Et<br />

M<br />

NH<br />

N<br />

Et<br />

N<br />

N<br />

Et<br />

Et<br />

Et Et<br />

508<br />

O<br />

Et<br />

N<br />

HN<br />

509 70%<br />

CO2Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

Et<br />

for references see p 1223


1220 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a512<br />

Et<br />

Et<br />

Ph<br />

Ph<br />

Et<br />

N<br />

N<br />

Ni<br />

N<br />

N<br />

Et<br />

Et Et<br />

NH<br />

N<br />

N<br />

N<br />

514<br />

N<br />

N<br />

N<br />

HN<br />

Ph<br />

Ni<br />

Ph<br />

516<br />

Ph<br />

Cu<br />

Ph<br />

446<br />

N<br />

N<br />

N<br />

N<br />

Et<br />

Et<br />

Ph<br />

CHO<br />

Ph<br />

I<br />

NMe3 −<br />

+<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

heat<br />

TsOH, benzene-1,2-dicarbonitrile<br />

160 oC 1% H2SO4, TFA<br />

TsOH, p-chloroaniline<br />

1. LiAlH4, AlCl3<br />

2. base<br />

Et<br />

Et<br />

Ph<br />

Ph<br />

Et<br />

N<br />

N<br />

Ni<br />

N<br />

N<br />

Et Et<br />

513<br />

Ph<br />

N<br />

N<br />

Ph<br />

Ni<br />

Ph<br />

N<br />

N<br />

N<br />

N<br />

517<br />

Ph<br />

Cu<br />

Ph<br />

518<br />

N<br />

N<br />

NH<br />

N<br />

N<br />

N<br />

Ph<br />

Cu<br />

Ph<br />

519<br />

515<br />

N<br />

N<br />

Et<br />

N<br />

HN<br />

Et<br />

O


a441<br />

17.8.4 Isomeric, Contracted, and Expanded Porphyrin Systems 1221<br />

Ph<br />

N<br />

N<br />

Ph<br />

Ni<br />

Ph<br />

N<br />

N<br />

NO 2<br />

Ph<br />

CNCH 2CO 2t-Bu<br />

CH 2Cl 2, O 2<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag<br />

Ph<br />

Ph<br />

N<br />

N<br />

Bu<br />

Ph<br />

tO2C N<br />

N<br />

Ni<br />

Ph<br />

Ph<br />

Ni<br />

Ph<br />

N<br />

N<br />

N<br />

N<br />

520<br />

521<br />

Ph<br />

N N<br />

CO 2Bu t<br />

N N<br />

Ni Ph<br />

CO2But Ph<br />

5-[2-(Ethoxycarbonyl)vinyl]-2,3,7,8,12,13,17,18-octaethylporphyrin (506); Typical Procedure:<br />

[213]<br />

A soln of (5-formyl-2,3,7,8,12,13,17,18-octaethylporphyrinato)nickel(II) (444; 506 mg,<br />

0.82 mmol) and an excess of (carbethoxymethylene)triphenylphosphorane (1.024 g) in xylene<br />

(50 mL) was refluxed for 18 h. The soln was cooled, the solvent removed in vacuo, and<br />

the residue was chromatographed (silica gel, CH 2Cl 2). A minor fraction of<br />

(2,3,7,8,12,13,17,18-octaethylporphyrinato)nickel(II) was collected first, followed by a<br />

major red band. Collection, removal of the solvent, and crystallization (CH 2Cl 2/MeOH)<br />

gave Cu•506; yield: 455 mg (81%). Compound Cu•506 (621 mg) was then dissolved in<br />

concd H 2SO 4 (10 mL) and kept at rt for 2 h. CH 2Cl 2 (100 mL) was added, followed by sat.<br />

aq NaHCO 3. After neutralization the organic phase was collected, washed with H 2O, dried<br />

(Na 2SO 4), and concentrated to dryness. The residue was crystallized (CH 2Cl 2/MeOH) to give<br />

506; yield: 552 mg (100%).<br />

2,3,7,8,12,13,17,18-Octaethylpurpurin 3¢-Ethyl Ester (507); Typical Procedure: [213]<br />

5-[2-(Ethoxycarbonyl)vinyl]-2,3,7,8,12,13,17,18-octaethylporphyrin (506; 100 mg,<br />

0.158 mmol) in AcOH (20 mL) was refluxed in a N 2 atmosphere for 24 h. After cooling the<br />

solvent was removed under reduced pressure and replaced with a little CH 2Cl 2. Chromatography<br />

(silica gel, CH 2Cl 2) gave a major green band which was concentrated to dryness.<br />

The product was crystallized (CH 2Cl 2/MeOH) to give 507; yield: 68 mg (68%); mp 135–<br />

1388C.<br />

for references see p 1223


1222 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a2,2,3,4,5,6,7,8-Octaethylbenzochlorin (509); Typical Procedure: [442]<br />

[5-(2-Formylvinyl)-2,3,7,8,12,13,17,18-octaethylporphyrinato]copper(II) (453, M = Cu;<br />

50 mg, 0.077 mmol) was stirred in 18% H 2SO 4/TFA (10 mL) for 15 min. The mixture was<br />

then neutralized with 20% aq NaHCO 3 (100 mL), extracted with CH 2Cl 2 and the combined<br />

organic layers were washed with H 2O (3 ” 200 mL). The soln was dried (Na 2SO 4), the solvent<br />

was removed and the resulting residue was chromatographed (silica gel, petroleum<br />

ether/CH 2Cl 2 7:3). The product was collected and crystallized (CH 2Cl 2/MeOH) to give 509;<br />

yield: 32 mg (70%); mp 241–2438C.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


aReferences<br />

References 1223<br />

[1] The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic: Boston, MA,<br />

(2000); Vols. 1–10.<br />

[2] Porphyrins and Metalloporphyrins, Smith, K. M., Ed.; Elsevier: Amsterdam, (1975).<br />

[3] Falk, J. E., Porphyrins and Metalloporphyrins, Elsevier: Amsterdam, (1964).<br />

[4] The Porphyrins, Dolphin, D., Ed.; Wiley: New York, (1975); 6 Vols.<br />

[5] Smith, K. M., In Rodd s Chemistry of the Carbon Compounds, Suppl. to 2nd Ed; Sainsbury, M., Ed.;<br />

Elsevier: Amsterdam, (1977); Vol. IVB, pp 237–327.<br />

[6] Smith, K. M., In Rodd s Chemistry of the Carbon Compounds, 2nd. Suppl. to 2nd Ed; Sainsbury, M.,<br />

Ed.; Elsevier: Amsterdam, (1997); Vol. IVB, pp 277–357.<br />

[7] Jackson, A. H.; Smith, K. M., In The Total Synthesis of Natural Products, ApSimon, J. W., Ed.; Wiley:<br />

New York, (1973); Vol. 1; pp 143–278.<br />

[8] Jackson, A. H.; Smith, K. M., In The Total Synthesis of Natural Products (1973–1980), ApSimon, J. W.,<br />

Ed.; Wiley: New York, (1984); Vol. 6; pp 237–280.<br />

[9] IUPAC Nomenclature for Tetrapyrroles, Pure Appl. Chem., (1987) 59, 779.<br />

[10] Senge, M. O., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 10.<br />

[11] Scheidt, W. R., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 3, pp 49–112.<br />

[12] Medforth, C. J., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 5, pp 1–80.<br />

[13] Reimers, J. R.; Lu, T. X.; Crossley, M. J.; Hush, N. S., J. Am. Chem. Soc., (1995) 117, 2855.<br />

[14] Boronat, M.; Ortí, E.; Viruela, P. M.; Tomµs, F., J. Mol. Struct. (Theochem.), (1997) 390, 149.<br />

[15] Medforth, C. J., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 5, pp 61–65.<br />

[16] Almlof, J.; Fischer, T. H.; Gassman, P. G.; Ghosh, A.; Haser, M., J. Phys. Chem., (1993) 97, 10964.<br />

[17] Woodward, R. B.; Skaric, V., J. Am. Chem. Soc., (1961) 83, 4676.<br />

[18] Smith, K. M., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 1, pp 1–44.<br />

[19] Lindsey, J. S., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 1, pp 45–118.<br />

[20] Smith, K. M., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 1, pp 119–148.<br />

[21] Clezy, P. S., Aust J. Chem., (1991) 44, 1163.<br />

[22] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1934); Vol. I.<br />

[23] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i.<br />

[24] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1940); Vol. II, part ii.<br />

[25] Arsenault, G. P.; Bullock, E.; MacDonald, S. F., J. Am. Chem. Soc., (1960) 82, 4384.<br />

[26] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i, p 2.<br />

[27] Fischer, H.; Böckh, S., Justus Liebigs Ann. Chem., (1935), 190.<br />

[28] Fischer, H.; Bäumler, R., Justus Liebigs Ann. Chem., (1929) 468, 73.<br />

[29] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i,<br />

pp 73, 106.<br />

[30] Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1972), 1471.<br />

[31] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i, p 3.<br />

[32] Fischer, H.; Sturm, E.; Friedrich, H., Justus Liebigs Ann. Chem., (1928) 461, 261.<br />

[33] Fischer, H.; Weichmann, H. K., Justus Liebigs Ann. Chem., (1931) 492, 61.<br />

[34] Fischer, H.; Riedl, H. J., Z. Physiol. Chem., (1932) 207, 200.<br />

[35] Cavaleiro, J. A. S.; Rocha Gonsalves, A. M. dA.; Kenner, G. W.; Smith, K. M., J. Chem. Soc., Perkin<br />

Trans. 1, (1973) 2471.<br />

[36] Clezy, P. S.; Liepa, A. J., Aust. J. Chem., (1970) 23, 2443.<br />

[37] Tarlton, E. J.; MacDonald, S. F.; Baltazzi, E., J. Am. Chem. Soc., (1960) 82, 4389.<br />

[38] Jackson, A. H.; Pandey, R. K.; Rao, K. R. N.; Roberts, E., Tetrahedron Lett., (1985) 26, 793.<br />

[39] Freeman, B. A.; Smith, K. M., Synth. Commun., (1999) 29, 1843.<br />

[40] Hawker, C. J.; Philippides, A.; Battersby, A. R., J. Chem. Soc., Perkin Trans. 1, (1991), 1833.<br />

[41] Hayes, A.; Kenner, G. W.; Williams, N. R., J. Chem. Soc., (1958), 3779.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1224 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a[42] Jackson, A. H.; Kenner, G. W.; Warburton, D., J. Chem. Soc., (1965), 1328.<br />

[43] Pandey, R. K.; Jackson, A. H.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1991), 1211.<br />

[44] Woodward, R. B.; Ayer, W. A.; Beaton, J. M.; Bickelhaupt, F.; Bonnett, R.; Buchschacher, P.; Closs,<br />

G. L.; Dutler, H.; Hannah, J.; Hauck, F. P.; Ito, S.; Langemann, A.; Le Goff, E.; Leimgruber, W.;<br />

Lwowski, W.; Sauer, J.; Valenta, Z.; Volz, H., Tetrahedron, (1990) 46, 7599.<br />

[45] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1934); Vol. I, p 333.<br />

[46] Mironov, A. F.; Ovsepyan, T. R.; Evstigneeva, R. P.; Preobrazenskii, N. A., Zh. Obshch. Khim., (1965)<br />

35, 324.<br />

[47] Swanson, K. L.; Snow, K. M.; Jeyakumar, D.; Smith, K. M., Tetrahedron, (1991) 47, 685.<br />

[48] Cavaleiro, J. A. S.; Rocha Gonsalves, A. M. dA.; Kenner, G. W.; Smith, K. M., J. Chem. Soc., Perkin<br />

Trans. 1, (1974), 1771.<br />

[49] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1934); Vol. I, p 332.<br />

[50] Fischer, H.; Nenitzescu, C., Justus Liebigs Ann. Chem., (1925) 443, 124.<br />

[51] Lee, D. A.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1997), 1215.<br />

[52] Lee, C.-H.; Lindsey, J. S., Tetrahedron, (1994) 50, 11427.<br />

[53] Ballantine, J. A.; Jackson, A. H.; Kenner, G. W.; McGillivray, G., Tetrahedron, (1966) 22 (Suppl. 7),<br />

241.<br />

[54] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1934); Vol. I, p 362.<br />

[55] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. I, p 361.<br />

[56] Clezy, P. S.; Smythe, G. A., Aust. J. Chem., (1969) 22, 239.<br />

[57] Osgerby, J. M.; MacDonald, S. F., Can. J. Chem., (1962) 40, 1585.<br />

[58] Clezy, P. S.; Liepa, A. J.; Nichol, A. W.; Smythe, G. A., Aust. J. Chem., (1970) 23, 589.<br />

[59] Grigg, R.; Johnson, A. W.; Wasley, J. W. F., J. Chem. Soc., (1963), 359.<br />

[60] Grigg, R.; Johnson, A. W., J. Chem. Soc., (1964), 3315.<br />

[61] Sessler, J. L.; Cyr, M. J.; Burrell, A. K., Tetrahedron, (1992) 48, 9661.<br />

[62] Sessler, J. L.; Hoehner, M. C., Synlett, (1994), 211.<br />

[63] Gonsalves, A. M. dA. R.; Kenner, G. W.; Smith, K. M., Tetrahedron Lett., (1972), 2203.<br />

[64] Bauer, V. J.; Clive, D. L. J.; Dolphin, D.; Paine, J. B., III.; Harris, F. L.; King, M. M.; Loder, J.; Wang, S.-<br />

W. C.; Woodward, R. B., J. Am. Chem. Soc., (1983) 105, 6429.<br />

[65] Sessler, J. L.; Lynch, V.; Johnson, M. R., J. Org. Chem., (1987) 52, 4394.<br />

[66] Sessler, J. L.; Cyr, M. J.; Lynch, V.; McGhee, E.; Ibers, J. A., J. Am. Chem. Soc., (1990) 112, 2810.<br />

[67] Jackson, A. H.; Kenner, G. W.; Sach, G. S., J. Chem. Soc. C, (1967), 2045.<br />

[68] Jackson, A. H.; Kenner, G. W.; Wass, J., J. Chem. Soc., Perkin Trans. 1, (1972), 1475.<br />

[69] Jackson, A. H.; Kenner, G. W.; McGillivray, G.; Smith, K. M., J. Chem. Soc. C, (1968), 294.<br />

[70] Fischer, H.; Kurzinger, A., Z. Physiol. Chem., (1931), 196, 213.<br />

[71] Johnson, A. W.; Kay, I. T., J. Chem. Soc., (1961), 2418.<br />

[72] Dicker, I. D.; Grigg, R.; Johnson, A. W.; Pinnock, H.; Richardson, K.; van den Broek, P., J. Chem. Soc.<br />

C, (1971), 536.<br />

[73] Badger, G. M.; Harris, R. L. N.; Jones, R. A., Aust. J. Chem., (1964) 17, 1013.<br />

[74] Clezy, P. S.; Liepa, A. J., Aust. J. Chem., (1971) 24, 1027.<br />

[75] Conlon, J. M.; Elix, J. A.; Feutrill, G. I.; Johnson, A. W.; Roomi, M. W.; Whelan, J., J. Chem. Soc.,<br />

Perkin Trans. 1, (1974), 713.<br />

[76] Jackson, A. H.; Kenner, G. W.; Smith, K. M., J. Chem. Soc. C, (1971), 502.<br />

[77] Cox, M. T.; Jackson, A. H.; Kenner, G. W., J. Chem. Soc. C, (1971), 1974.<br />

[78] Almeida, J. A. P. B.; Kenner, G. W.; Smith, K. M.; Sutton, M. J., J. Chem. Soc., Chem. Commun., (1975),<br />

111.<br />

[79] Almeida, J. A. P. B.; Kenner, G. W.; Rimmer, J.; Smith, K. M., Tetrahedron, (1976) 32, 1793.<br />

[80] Smith, K. M.; Craig, G. W., J. Org. Chem., (1983) 48, 4302.<br />

[81] Kenner, G. W.; Rimmer, J.; Smith, K. M.; Unsworth, J. F., J. Chem. Soc., Perkin Trans. 1, (1977), 332.<br />

[82] Engel, J.; Gossauer, A., J. Chem. Soc., Chem. Commun., (1975), 570.<br />

[83] Evstigneeva, R. P.; Mironov, A. F.; Fleiderman, L. I., Dokl. Akad. Nauk SSSR, (1973) 210, 1090.<br />

[84] Harris, R. L. N.; Johnson, A. W.; Kay, I. T., J. Chem. Soc. C, (1966), 22.<br />

[85] Rothemund, P., J. Am. Chem. Soc., (1935) 57, 2010.<br />

[86] Rothemund, P.; Menotti, A. R., J. Am. Chem. Soc., (1941) 63, 267.<br />

[87] Adler, A. D.; Longo, F. R.; Finarelli, J. D.; Goldmacher, J.; Assour, J.; Korsakoff, L., J. Org. Chem.,<br />

(1967) 32, 476.<br />

[88] Lindsey, J. S.; Schreiman, I. C.; Hsu, H. C.; Kearney, P. C.; Marguerettaz, A. M., J. Org. Chem., (1987)<br />

52, 827.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


References 1225<br />

a[89] Badger, G. M.; Jones, R. A.; Laslett, R. L., Aust. J. Chem., (1964) 17, 1028.<br />

[90] Barnett, G. H.; Hudson, M. F.; Smith, K. M., Tetrahedron Lett., (1973), 2887.<br />

[91] Barnett, G. H.; Hudson, M. F.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1975), 1401.<br />

[92] Evans, B.; Smith, K. M.; Fuhrhop, J.-H., Tetrahedron Lett., (1977), 443.<br />

[93] Medforth, C. J.; Berber, M. D.; Smith, K. M.; Shelnutt, J. A., Tetrahedron Lett., (1990) 31, 3719.<br />

[94] Medforth, C. J.; Smith, K. M., Tetrahedron Lett., (1990) 31, 5583.<br />

[95] Kadish, K. M.; Caemelbecke, E. V.; D Souza, F.; Lin, M.; Nurco, D. J.; Medforth, C. J.; Forsyth, T. P.;<br />

Krattinger, B.; Smith, K. M.; Fukuzumi, S.; Nakanishi, I.; Shelnutt, J. A., Inorg. Chem., (1999) 38,<br />

2188.<br />

[96] Senge, M. O., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 1, pp 239–348.<br />

[97] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i,<br />

p 175.<br />

[98] Sessler, J. L.; Mozaffari, A.; Johnson, M. R., Org. Synth., (1991) 70, 68.<br />

[99] Ono, M.; Lattmann, R.; Imomota, K.; Lehmann, C.; Früh, T.; Eschenmoser, A., Croat. Chem. Acta,<br />

(1985) 58, 627.<br />

[100] Schreiber, J.; Maag, H.; Hashimoto, N.; Eschenmoser, A., Angew. Chem., (1971) 83, 355; Angew.<br />

Chem. Int. Ed. Engl., (1971) 10, 330.<br />

[101] Whitlock, H. W.; Hanauer, R., J. Org. Chem., (1968) 33, 2169.<br />

[102] Inhoffen, H. H.; Fuhrhop, J.-H.; Voigt, H.; Brockmann, H., Jr., Justus Liebigs Ann. Chem., (1966) 695,<br />

183.<br />

[103] Barton, D. H. R.; Zard, S. Z., J. Chem. Soc., Chem. Commun., (1985), 1098.<br />

[104] Barton, D. H. R.; Kervagoret, J.; Zard, S. Z., Tetrahedron, (1990) 46, 7587.<br />

[105] Ono, N.; Kawamura, H.; Bougauchi, M.; Maruyama, K., Tetrahedron, (1990) 46, 7483.<br />

[106] Aoyagi, K.; Haga, T.; Toi, H.; Aoyama, Y.; Mizutani, T.; Ogoshi, H., Bull. Chem. Soc. Jpn., (1997) 70,<br />

937.<br />

[107] Mauzerall, D., J. Am. Chem. Soc., (1960) 82, 2601.<br />

[108] Scott, A. I., Angew. Chem., (1993) 105, 1281; Angew. Chem. Int. Ed. Engl., (1993) 32, 1223.<br />

[109] Battersby, A. R., Science, (1994) 264, 1551.<br />

[110] Scott, A. I., Tetrahedron, (1994) 50, 13315.<br />

[111] Blanche, F.; Cameron, B.; Crouzet, J.; Debussche, L.; Thibaut, D.; Vuilhorgne, M.; Leeper, F. J.;<br />

Battersby, A. R., Angew. Chem., (1995) 107, 421; Angew. Chem. Int. Ed. Engl., (1995) 34, 383.<br />

[112] Scott, A. I., Heterocycles, (1998) 47, 1051.<br />

[113] Bag, N.; Chern, S.-S.; Peng, S.-M.; Chang, C. K., Tetrahedron Lett., (1995) 36, 6409. For comparison,<br />

see also: Chang, C. K.; Bag, N., J. Org. Chem., (1995) 60, 7030.<br />

[114] Nguyen, L. T.; Smith, K. M., Tetrahedron Lett., (1996) 37, 7177.<br />

[115] Nguyen, L. T.; Senge, M. O.; Smith, K. M., Tetrahedron Lett., (1994) 35, 7581.<br />

[116] Nguyen, L. T.; Senge, M. O.; Smith, K. M., J. Org. Chem., (1996) 61, 998.<br />

[117] Fischer, H.; Klarer, J., Ann. Chem., (1926) 448, 178.<br />

[118] Fischer, H.; Friedrich, H.; Lamatsch, W.; Morgenroth, K., Ann. Chem., (1928) 466, 147.<br />

[119] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i,<br />

p 169.<br />

[120] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i,<br />

p 193.<br />

[121] Fischer, H.; Baumler, R., Justus Liebigs Ann. Chem., (1929) 468, 58.<br />

[122] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i,<br />

p 188.<br />

[123] Bonnett, R.; McManus, K. A., J. Chem. Soc., Chem. Commun., (1994), 1129.<br />

[124] Fischer, H.; Zeile, K., Ann. Chem., (1929) 468, 98.<br />

[125] Fischer, H.; Orth, H., Die Chemie des Pyrrols, Akademische Verlag: Leipzig, (1937); Vol. II, part i,<br />

p 165.<br />

[126] Hudson, M. F.; Smith, K. M., J. Chem. Soc., Chem. Commun., (1973), 515.<br />

[127] Hudson, M. F.; Smith, K. M., Tetrahedron, (1975) 31, 3077.<br />

[128] Cavaleiro, J. A. S.; Kenner, G. W.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1974), 1188.<br />

[129] Abraham, R. J.; Barnett, G. H.; Bretschneider, E. S.; Smith, K. M., Tetrahedron, (1973) 29, 553.<br />

[130] Little, R. G.; Anton, J. A.; Loach, P. A.; Ibers, J. A., J. Heterocycl. Chem., (1975) 12, 343.<br />

[131] Mullen, K.; Hammel, D.; Kautz, C., Chem. Ber., (1990) 123, 1353.<br />

[132] Wallace, D. M.; Leung, S. H.; Senge, M. O.; Smith, K. M., J. Org. Chem., (1993) 58, 7245.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1226 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a[133] Lee, C.-H.; Li, F.; Iwamoto, K.; Dadok, J.; Bothner-By, A. A.; Lindsey, J. S., Tetrahedron, (1995) 51,<br />

11645.<br />

[134] Woodward, R. B., Angew. Chem., (1960) 72, 651.<br />

[135] Woodward, R. B., Pure Appl. Chem., (1961) 2, 883.<br />

[136] Boudif, A.; Momenteau, M., J. Chem. Soc., Chem. Commun., (1994), 2096.<br />

[137] Boudif, A.; Momenteau, M., J. Chem. Soc., Perkin Trans. 1, (1996), 1235.<br />

[138] Hu, Z.; Lash, T. D., Synlett, (1994), 909.<br />

[139] Lash, T. D.; Novak, B. H., Angew. Chem., (1995) 107, 723; Angew. Chem. Int. Ed. Engl., (1995) 34,<br />

683.<br />

[140] Lash, T. D.; Novak, B. H., Tetrahedron Lett., (1995) 36, 4381.<br />

[141] Lash, T. D., J. Porphyrins Phthalocyanines, (1997) 1, 29.<br />

[142] Lash, T. D., Chem.–Eur. J., (1996) 2, 1197.<br />

[143] Sessler, J. L.; Genge, J. W.; Urbach, A.; Sanson, P., Synlett, (1996), 187.<br />

[144] Clezy, P. S.; Liepa, A. J.; Smythe, G. A., Aust. J. Chem., (1970) 23, 603.<br />

[145] Bonnett, R.; Dimsdale, M. J.; Stephenson, G. F., J. Chem. Soc. C, (1969), 564.<br />

[146] Chong, R.; Clezy, P. S.; Liepa, A. J.; Nichol, A. W., Aust. J. Chem., (1969) 22, 229.<br />

[147] Jackson, A. H.; Kenner, G. W.; Smith, K. M., J. Chem. Soc. C, (1968), 302.<br />

[148] Broadhurst, M. J.; Grigg, R.; Johnson, A. W., J. Chem. Soc. C, (1971), 3681.<br />

[149] Tardieux, C.; Bolze, F.; Gros, C. P.; Guilard, R., Synthesis, (1988), 267.<br />

[150] Corwin, A. H.; Coolidge, E. C., J. Am. Chem. Soc., (1952) 74, 5196.<br />

[151] Smith, K. M.; Eivazi, F., J. Org. Chem., (1979) 44, 2591.<br />

[152] Jackson, A. H.; Kenner, G. W.; Wass, J., J. Chem. Soc., Perkin Trans. 1, (1974), 480.<br />

[153] Carr, R. P.; Jackson, A. H.; Kenner, G. W.; Sach, G. S., J. Chem. Soc. C, (1971), 487.<br />

[154] Cox, M. T.; Jackson, A. H.; Kenner, G. W.; McCombie, S. W.; Smith, K. M., J. Chem. Soc., Perkin<br />

Trans. 1, (1974), 516.<br />

[155] Ellis, J.; Jackson, A. H.; Jain, A. C.; Kenner, G. W., J. Chem. Soc., (1964), 1935.<br />

[156] Clezy, P. S.; Diakiw, V.; Webb, N. W., J. Chem. Soc., Chem. Commun., (1972), 413.<br />

[157] Clezy, P. S.; Diakiw, V.; Webb, N. W., Aust. J. Chem., (1973) 26, 2697.<br />

[158] Clezy, P. S.; Diakiw, V., J. Chem. Soc., Chem. Commun., (1973), 453.<br />

[159] Clezy, P. S.; Liepa, A. J.;. Webb, N. W., Aust. J. Chem., (1972) 25, 1991.<br />

[160] Flaugh, M. E.; Rapoport, H., J. Am. Chem. Soc., (1968) 90, 6877.<br />

[161] Clezy, P. S.; Fookes, C. J. R., Aust. J. Chem., (1974) 27, 371.<br />

[162] Grigg, R.; Johnson, A. W.; Kenyon, R. W.; Math, V. B.; Richardson, K., J. Chem. Soc. C, (1969), 176.<br />

[163] Smith, K. M.; Minnetian, O. M., J. Chem. Soc., Perkin Trans. 1, (1986), 277.<br />

[164] Jackson, A. H.; Pandey, R. K.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1987), 299.<br />

[165] Jeyakumar, D.; Snow, K. M.; Smith, K. M., J. Am. Chem. Soc., (1988) 110, 8562.<br />

[166] Lin, J. J.; Holmes, R. T.; Smith, K. M., J. Porphyrins Phthalocyanines, (1998) 2, 363.<br />

[167] Boschi, T.; Paolesse, R.; Tagliatesta, P., Inorg. Chim. Acta, (1990) 168, 83.<br />

[168] Gerzevske, K. R.; Lin, J. J.; Smith, K. M., Heterocycles, (1994) 37, 207.<br />

[169] Lin, J. J.; Gerzevske, K. R.; Liddell, P. A.; Senge, M. O.; Olmstead, M. M.; Khoury, R. G.; Weeth, B. E.;<br />

Tsao, S. A.; Smith, K. M., J. Org. Chem., (1997) 62, 4266.<br />

[170] Kulish, M. A.; Mironov, A. F.; Rozynov, B. V.; Evstigneeva, R. P., Zh. Obshch. Khim., (1971) 41, 2743.<br />

[171] Smith, K. M.; Kehres, L. A., J. Chem. Soc., Perkin Trans. 1, (1983), 2329.<br />

[172] Smith, K. M.; Minnetian, O. M., J. Org. Chem., (1985) 50, 2073.<br />

[173] Liddell, P. A.; Gerzevske, K. R.; Lin, J. J.; Olmstead, M. M.; Smith, K. M., J. Org. Chem., (1993) 58,<br />

6681.<br />

[174] Dolphin, D.; Johnson, A. W.; Leng, J.; van den Broek, P., J. Chem. Soc. C, (1966), 880.<br />

[175] Licoccia, S.; Paolesse, R., Struct. Bonding (Berlin), (1995) 84, 71.<br />

[176] Johnson, A. W., In Porphyrins and Metalloporphyrins, Smith, K. M., Ed.; Elsevier: Amsterdam,<br />

(1975); Chapter 18 (pp 729–754).<br />

[177] Mironov, A. F.; Kulish, M. A.; Kobak, V. V.; Rozynov, B. V.; Evstigneeva, R. P., Zh. Obshch. Khim.,<br />

(1974) 44, 1407.<br />

[178] Smith, K. M.; Miura, M.; Morris, I. K., J. Org. Chem., (1986) 51, 4660.<br />

[179] Mironov, A. F.; Rumyantseva, V. P.; Fleiderman, L. I.; Evstigneeva, R. P., Zh. Obshch. Khim., (1975)<br />

45, 1150.<br />

[180] Battersby, A. R.; Hodgson, G. L.; Ihara, M.; McDonald, E.; Saunders, J., J. Chem. Soc., Perkin Trans. 1,<br />

(1973), 531.<br />

[181] Maruyama, K.; Nagata, T.; Ono, N.; Osuka, A., Bull. Chem. Soc. Jpn., (1989) 62, 3167.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


References 1227<br />

a[182] Senge, M. O.; Liddell, P. A.; Smith, K. M., Acta Crystallogr., Sect. C, (1992) 48, 581.<br />

[183] Burns, D. H.; Caldwell, T. M.; Burden, M. W., Tetrahedron Lett., (1993) 34, 2883.<br />

[184] Closs, G. L.; Closs, L. E., J. Am. Chem. Soc., (1963) 85, 818.<br />

[185] Buchler, J. W.; Schneehage, H. H., Tetrahedron Lett., (1972), 3803.<br />

[186] Whitlock, H. W.; Oester, M. Y., J. Am. Chem. Soc., (1973) 95, 5738.<br />

[187] Schmidt, W.; Montforts, F.-P., Synlett, (1997), 903.<br />

[188] Abel, Y.; Montforts, F.-P., Tetrahedron Lett., (1997) 38, 1745.<br />

[189] Strachan, J.-P.; O Shea, D. F.; Balasubramanian, T.; Lindsey, J. S., J. Org. Chem., (2000) 65, 3160.<br />

[190] Balasubramanian, T.; Strachan, J.-P.; Boyle, P. D.; Lindsey, J. S., J. Org. Chem., (2000) 65, 7919.<br />

[191] Battersby, A. R.; Dutton, C. J.; Fookes, C. J. R., J. Chem. Soc., Perkin Trans. 1, (1988), 1569.<br />

[192] Montforts, F.-P.; Schwartz, U. M., Liebigs Ann. Chem., (1991), 709.<br />

[193] Inhoffen, H. H.; Buchler, J. W.; Thomas, R., Tetrahedron Lett., (1969), 1141.<br />

[194] Smith, K. M.; Lai, J.-J., J. Am. Chem. Soc., (1984) 106, 5746.<br />

[195] Burns, D. H.; Lai, J.-J.; Smith, K. M., J. Chem. Soc., Perkin Trans.1, (1988), 3119.<br />

[196] Whitlock, H. W., Jr.; Hanauer, R.; Oester, M. Y.; Bower, B. K., J. Am. Chem. Soc., (1969) 91, 7485.<br />

[197] Ulman, A.; Gallucci, J.; Fisher, D.; Ibers, J. A., J. Am. Chem. Soc., (1980) 102, 6852.<br />

[198] Vasudevan, J.; Stibrany, R. T.; Bumby, J.; Knapp, S.; Potenza, J. A.; Emge, T. J.; Schugar, H. J., J. Am.<br />

Chem. Soc., (1996) 118, 11676.<br />

[199] Stolzenberg, A. M.; Simerly, S. W.; Steffey, B. D.; Haymond, G. H., J. Am. Chem. Soc., (1997) 119,<br />

11843.<br />

[200] Whitten, D. G.; Yau, J. C.; Carroll, F. A., J. Am. Chem. Soc., (1971) 93, 2291.<br />

[201] Fuhrhop, J.-H.; Lumbantobing, T., Tetrahedron Lett., (1970), 2815.<br />

[202] Harel, Y.; Manassen, J., J. Am. Chem. Soc., (1978) 100, 6228.<br />

[203] Wolf, H.; Scheer, H., Justus Liebigs Ann. Chem., (1973), 1710.<br />

[204] Smith, K. M.; Simpson, D. J., J. Chem. Soc., Chem. Commun., (1987), 613.<br />

[205] Iakovides, P.; Simpson, D. J.; Smith, K. M., Photochem. Photobiol., (1991) 54, 335.<br />

[206] Seely, G. R.; Talmadge, K., Photochem. Photobiol., (1964) 3, 195.<br />

[207] Seely, G. R., J. Am. Chem. Soc., (1966) 88, 3417.<br />

[208] Smith, K. M.; Simpson, D. J.; Snow, K. M., J. Am. Chem. Soc., (1986) 108, 6834.<br />

[209] Simpson, D. J.; Smith, K. M., J. Am. Chem. Soc., (1988) 110, 2854.<br />

[210] Smith, K. M.; Goff, D. A., J. Am. Chem. Soc., (1985) 107, 4954.<br />

[211] Smith, K. M.; Simpson, D. J., J. Am. Chem. Soc., (1987) 109, 6326.<br />

[212] Arnold, D. P.; Gaete-Holmes, R.; Johnson, A. W.; Smith, A. R. P.; Williams, G. A., J. Chem. Soc., Perkin<br />

Trans. 1, (1978), 1660.<br />

[213] Morgan, A. R.; Tertel, N. C., J. Org. Chem., (1986) 51, 1347.<br />

[214] Morgan, A. R.; Rampersaud, A.; Garbo, G. M.; Keck, R. W.; Selman, S. H., J. Med. Chem., (1989) 32,<br />

904.<br />

[215] Gunter, M. J.; Robinson, B. C., Tetrahedron Lett., (1990) 31, 285.<br />

[216] Gunter, M. J.; Robinson, B. C., Aust. J. Chem., (1990) 43, 1839.<br />

[217] Forsyth, T. P.; Nurco, D. J.; Pandey, R. K.; Smith, K. M., Tetrahedron Lett., (1995) 36, 9093.<br />

[218] Morgan, A. R.; Skalkos, D.; Garbo, G. M.; Keck, R. W.; Selman, S. H., J. Med. Chem., (1991) 34, 2126.<br />

[219] Montforts, F.-P.; Zimmermann, G., Angew. Chem., (1986) 98, 451; Angew. Chem. Int. Ed. Engl.,<br />

(1986) 25, 458.<br />

[220] Montforts, F.-P.; Meier, A.; Haake, G.; Hoper, F., Tetrahedron Lett., (1991) 32, 3481.<br />

[221] Haake, G.; Meier, A.; Montforts, F.-P.; Scheurich, G.; Zimmermann, G., Liebigs Ann. Chem., (1992),<br />

325.<br />

[222] Kusch, D.; Meier, A.; Montforts, F.-P., Liebigs Ann. Chem., (1995), 1027.<br />

[223] Montforts, F.-P.; Meier, A.; Scheurich, G.; Haake, G.; Bats, J. W., Angew. Chem., (1992) 104, 1650;<br />

Angew. Chem. Int. Ed. Engl., (1992) 31, 1592.<br />

[224] Bats, J. W.; Haake, G.; Meier, A.; Montforts, F.-P.; Scheurich, G., Liebigs Ann. Chem., (1995), 1617.<br />

[225] Montforts, F.-P.; Mai, G.; Romanowski, F.; Bats, J. W., Tetrahedron Lett., (1992) 33, 765.<br />

[226] Romanowski, F.; Mai, G.; Kusch, D.; Montforts, F.-P.; Bats, J. W., Helv. Chim. Acta, (1996) 79, 1572.<br />

[227] Collier, G. L.; Jackson, A. H.; Kenner, G. W., J. Chem. Soc., Chem. Commun., (1966), 299.<br />

[228] Collier, G. L.; Jackson, A. H.; Kenner, G. W., J. Chem. Soc. C, (1967), 66.<br />

[229] Buchler, J. W.; Puppe, L., Justus Liebigs Ann. Chem., (1970) 740, 142.<br />

[230] Dwyer, P. N.; Buchler, J. W.; Scheidt, W. R., J. Am. Chem. Soc., (1974) 96, 2789.<br />

[231] Shulg a, A. M.; Sinyakov, G. N.; Filatov, I. V.; Gurinovich, G. P.; Dzilinski, K., J. Mol. Struct., (1995)<br />

348, 65.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1228 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a[232] Bonnett, R.; Nizhnik, A. N.; Berenbaum, M. C., J. Chem. Soc., Chem. Commun., (1989), 1822.<br />

[233] Adams, K. R.; Berenbaum, M. C.; Bonnett, R.; Nizhnik, A. N.; Salgado, A.; VallØs, M. A., J. Chem.<br />

Soc., Perkin Trans. 1, (1992), 1465.<br />

[234] Chang, C. K.; Sotiriou, C., J. Org. Chem., (1985) 50, 4989.<br />

[235] Chang, C. K.; Sotiriou, C., J. Heterocycl. Chem., (1985) 22, 1739.<br />

[236] Nakamura, K.; Osamura, Y., J. Phys. Org. Chem., (1990) 3, 737.<br />

[237] Kozyrev, A. N.; Pandey, R. K.; Medforth, C. J.; Zheng, G.; Dougherty, T. J.; Smith, K. M., Tetrahedron<br />

Lett., (1996) 37, 747.<br />

[238] Pandey, R. K.; Isaac, M.; MacDonald, I.; Medforth, C. J.; Senge, M. O.; Dougherty, T. J.; Smith, K. M.,<br />

J. Org. Chem., (1997) 62, 1463.<br />

[239] Stolzenberg, A. M.; Glazer, P. A.; Foxman, B. M., Inorg. Chem., (1986) 25, 983.<br />

[240] Chang, C. K.; Wu, W., J. Org. Chem., (1986) 51, 2134.<br />

[241] Callot, H. J., Bull. Soc. Chim. Fr., (1974), 1492.<br />

[242] Crossley, M. J.; King, G. K.; Newsom, I. A.; Sheehan, C. S., J. Chem. Soc., Perkin Trans. 1, (1996),<br />

2675.<br />

[243] Beavington, R.; Rees, P. A.; Burn, P. L., J. Chem. Soc., Perkin Trans. 1, (1998), 2847.<br />

[244] Chang, C. K.; Sotiriou, C.; Wu, W., J. Chem. Soc., Chem. Commun., (1986), 1213.<br />

[245] Bruckner, C.; Dolphin, D., Tetrahedron Lett., (1995) 36, 9425.<br />

[246] Pandey, R. K.; Shiau, F.-Y.; Isaac, M.; Ramaprasad, S.; Dougherty, T. J.; Smith, K. M., Tetrahedron<br />

Lett., (1992) 33, 7815.<br />

[247] Pandey, R. K.; Shiau, F.-Y.; Sumlin, A. B.; Dougherty, T. J.; Smith, K. M., Bioorg. Med. Chem. Lett.,<br />

(1994), 4, 1263.<br />

[248] Pandey, R. K.; Constantine, S.; Tsuchida, T.; Zheng, G.; Medforth, C. J.; Aoudia, M.; Kozyrev, A. N.;<br />

Rodgers, M. A. J.; Kato, H.; Smith, K. M.; Dougherty, T. J., J. Med. Chem., (1997) 40, 2770.<br />

[249] Battersby, A. R.; Reiter, L. A., J. Chem. Soc., Perkin Trans. 1, (1984), 2743.<br />

[250] Harrison, P. J.; Sheng, Z. -C.; Fookes, C. J. R.; Battersby, A. R., J. Chem. Soc., Perkin Trans. 1, (1987),<br />

1667.<br />

[251] Battersby, A. R.; Block, M. H.; Fookes, C. J. R.; Harrison, P. J.; Henderson, G. B.; Leeper, F. J., J.<br />

Chem. Soc., Perkin Trans. 1, (1992), 2175.<br />

[252] Battersby, A. R.; Block, M. H.; Leeper, F. J.; Zimmerman, S. C., J. Chem. Soc., Perkin Trans. 1, (1992),<br />

2189.<br />

[253] Aucken, C. J.; Leeper, F. J.; Battersby, A. R., J. Chem. Soc., Perkin Trans. 1, (1997), 2099.<br />

[254] Mackman, R. L.; Micklefield, J.; Block, M. H.; Leeper, F. J.; Battersby, A. R., J. Chem. Soc., Perkin<br />

Trans. 1, (1997), 2111.<br />

[255] Micklefield, J.; Beckmann, M.; Mackman, R. L.; Block, M. H.; Leeper, F. J.; Battersby, A. R., J. Chem.<br />

Soc., Perkin Trans. 1, (1997), 2123.<br />

[256] Gibson, C. L.; Doyle, M. J.; Raithby, P. R.; Battersby, A. R., J. Chem. Soc., Perkin Trans. 1, (1994), 1893.<br />

[257] Helberger, J. H., Justus Liebigs Ann. Chem., (1937) 529, 205.<br />

[258] Helberger, J. H.; von Rebay, A., Justus Liebigs Ann. Chem., (1937) 531, 279.<br />

[259] Barrett, P. A.; Linstead, R. P.; Tuey, G. A. P., J. Chem. Soc., (1939), 1809.<br />

[260] Dent, C. E., J. Chem. Soc., (1938), 1.<br />

[261] Helberger, J. H.; von Rebay, A.; HevØr, D. B., Justus Liebigs Ann. Chem., (1938) 533, 197.<br />

[262] Barrett, P. A.; Linstead, R. P.; Rundall, F. G.; Tuey, G. A. P., J. Chem. Soc., (1940), 1079.<br />

[263] Helberger, J. H.; HevØr, D. B., Justus Liebigs Ann. Chem., (1938) 536, 173.<br />

[264] Linstead, R. P.; Weiss, F. T., J. Chem. Soc., (1950), 2975.<br />

[265] Barrett, P. A.; Linstead, R. P.; Leavill, J.; Rowe, G., J. Chem. Soc., (1940), 1076.<br />

[266] Edwards, L.; Gouterman, M.; Rose, C. B., J. Am. Chem. Soc., (1976) 98, 7638.<br />

[267] Koehorst, R. B. M.; Kleibeuker, J. F.; Schaafsma, T. J.; de Bie, D. A.; Geurtsen, B.; Henrie, R. N.; van<br />

der Plas, H. C., J. Chem. Soc., Perkin Trans. 2, (1981), 1005.<br />

[268] Hanack, M.; Zipplies, T., J. Am. Chem. Soc., (1985) 107, 6127.<br />

[269] Kopranenkov, V. N.; Makarova, E. A. Luk yanets, E. A., Zh. Obshch. Khim., (1981) 51, 2727.<br />

[270] Kopranenkov, V. N.; Makarova, Dashkevich, S. N.; Luk yanets, E. A., Zh. Obshch. Khim., (1988), 630.<br />

[271] Carter, T. P.; Brauchle, C.; Lee, V. Y.; Manavi, M.; Moerner, W. E., J. Phys. Chem., (1987) 91, 3998.<br />

[272] Vogler, A.; Kunkely, H., Angew. Chem., (1978) 90, 808; Angew. Chem. Int. Ed. Engl., (1978) 17, 760.<br />

[273] Vogler, A.; Kunkely, H.; Rethwisch, B., Inorg. Chim. Acta, (1980) 46, 101.<br />

[274] Martinsen, J.; Pace, L. J.; Phillips, T. E.; Hoffman,B. M.; Ibers, J. A., J. Am. Chem. Soc., (1982) 104, 83.<br />

[275] Bender, C. O.; Bonnett, R.; Smith, R. G., Chem. Commun., (1969), 345.<br />

[276] Bender, C. O.; Bonnett, R.; Smith, R. G., J. Chem. Soc. C, (1970), 1251.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


References 1229<br />

a[277] Bender, C. O.; Bonnett, R.; Smith, R. G., J. Chem. Soc., Perkin Trans. 1, (1972), 771.<br />

[278] VallØs, M. A.; Gómez, A. M., Proc. SPIE, (1995) 2325, 24.<br />

[279] VallØs, M. A.; Biolo, R.; Bonnett, R.; Caæete, M.; Gómez, A. M.; Jori, G.; Juarranz, A.; McManus, K.<br />

A.;. Okolo, K. T.; Soncin, M.; Villanueva, A., Proc. SPIE, (1996) 2625, 11.<br />

[280] Remy, D. E., Tetrahedron Lett., (1983) 24, 1451.<br />

[281] Cheng, R.-J.; Chen, Y.-R.; Chuang, C.-E., Heterocycles, (1992) 34, 1.<br />

[282] Cheng, R.-J.; Chen, Y.-R.; Chuang, C.-E., Polyhedron, (1993) 12, 1353.<br />

[283] Ichimura, K.; Sakuragi, M.; Morii, H.; Yasuike, M. Fukui, M.; Ohno, O., Inorg. Chim. Acta, (1990)<br />

176, 31.<br />

[284] Kopranenkov, V. N.; Dashkevich, S. N.; Luk yanets, E. A., Zh. Obshch. Khim., (1981) 51, 2513.<br />

[285] Ichimura, K.; Sakuragi, M.; Morii, H.; Yasuike, M.; Fukui, M.; Ohno, O., Inorg. Chim. Acta, (1991)<br />

182, 83.<br />

[286] Yasuike, M.; Yamaoka, T.; Ohno, O.; Ichimura, K.; Morii, H.; Sakuragi, M., Inorg. Chim. Acta, (1991)<br />

185, 39.<br />

[287] Vinogradov, S. A.; Wilson, D. F., J. Chem. Soc., Perkin Trans.2, (1995), 103.<br />

[288] Ito, S.; Murashima, T.; Uno, H.; Ono, N., Chem. Commun. (Cambridge), (1998), 1661.<br />

[289] Ito, S.; Ochi, N.; Murashima, T.; Uno, H.; Ono, N., Heterocycles, (2000) 52, 399.<br />

[290] Finikova, O.; Cheprakov, A.; Beletskaya, I.; Vinogradov, S., Chem. Commun. (Cambridge), (2001),<br />

261.<br />

[291] Vicente, M. G. H.; Tome, A. C.; Walter, A.; Cavaleiro, J. A. S., Tetrahedron Lett., (1997) 38, 3639.<br />

[292] Baker, E. W., J. Am. Chem. Soc., (1966) 88, 2311.<br />

[293] Baker, E. W.; Yen, T. F.; Dikie, J. P.; Rhodes, R. E.; Clarke, L. F., J. Am. Chem. Soc., (1967) 89, 3631.<br />

[294] Kaur, S.; Chicarelli, M. I.; Maxwell, J. R., J. Am. Chem. Soc., (1986) 108, 1347.<br />

[295] Kaur, S.; Gill, J. P.; Evershed, R. P.; Eglinton, G.; Maxwell, J. R., J. Chromatogr., (1989) 473, 135.<br />

[296] Quirke, J. M. E.; Dale, T.; Britton, E. D.; Yost, R. A.; Trichet, J.; Belayouni, H., Org. Geochem., (1990)<br />

15, 169.<br />

[297] Callot, H. J.; Ocampo, R., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.;<br />

Academic: Boston, MA, (2000); Vol. 1, pp 349–398.<br />

[298] Clezy, P. S.; Fookes, C. J. R.; Mirza, A. H., Aust. J. Chem., (1977) 30, 1337.<br />

[299] Clezy, P. S.; Leung, C. W. F., Aust. J. Chem., (1993) 46, 1705.<br />

[300] Clezy, P. S.; Mirza, A. H., Aust. J. Chem., (1982) 35, 197.<br />

[301] May, D. M., Jr.; Lash, T. D., J. Org. Chem., (1992) 57, 4820.<br />

[302] Lash, T. D., Energy & Fuels, (1993) 7, 166.<br />

[303] Vicente, M. G. H.; Tan, K.; Smith, K. M., unpublished results.<br />

[304] Bonnett, R.; McManus, K. A., J. Chem. Soc., Perkin Trans. 1, (1996), 2461.<br />

[305] Grigg, R.; Johnson, A. W.; Sweeney, A., Chem. Commun., (1968), 697.<br />

[306] Callot, H. J.; Johnson, A. W.; Sweeney, A., J. Chem. Soc., Perkin Trans. 1, (1973), 1424.<br />

[307] DiNello, R. K.; Dolphin, D., J. Org. Chem., (1980) 45, 5196.<br />

[308] Cavaleiro, J. A. S.; Jackson, A. H.; Neves, M. G. P. M. S.; Rao, K. R. N., J. Chem. Soc., Chem. Commun.,<br />

(1985), 776.<br />

[309] Pangka, V. S.; Morgan, A. R.; Dolphin, D., J. Org. Chem., (1986) 51, 1094.<br />

[310] Shiau, F.-Y.; Pandey, R. K.; Dougherty, T. J.; Smith, K. M., Proc. SPIE, (1991) 330, 1426.<br />

[311] Morgan, A. R.; Kohli, D. H., Tetrahedron Lett., (1995) 36, 7603.<br />

[312] Faustino, M. A. F.; Neves, M. G. P. M. S.; Vicente, M. G. H.; Silva, A. M. S.; Cavaleiro, J. A. S., Tetrahedron<br />

Lett., (1996) 37, 3569.<br />

[313] Morgan, A. R.; Pangka, V. S.; Dolphin, D., J. Chem. Soc., Chem. Commun., (1984), 1047.<br />

[314] Yon-Hin, P.; Wijesekera, T. P.; Dolphin, D., Tetrahedron Lett., (1989) 30, 6135.<br />

[315] Pandey, R. K.; Shiau, F.-Y.; Ramachandran, K.; Dougherty, T. J.; Smith, K. M., J. Chem. Soc., Perkin<br />

Trans. 1, (1992), 1377.<br />

[316] Meunier, I.; Pandey, R. K.; Walker, M. M.; Senge, M. O.; Dougherty, T. J.; Smith, K. M., Bioorg. Med.<br />

Chem. Lett., (1992) 2, 1575.<br />

[317] Pandey, R. K.; Jagerovic, N.; Ryan, J. M.; Dougherty, T. J.; Smith, K. M., Bioorg. Med. Chem. Lett.,<br />

(1993) 3, 2615.<br />

[318] Meunier, I.; Pandey, R. K.; Senge, M. O.; Dougherty, T. J.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1,<br />

(1994), 961.<br />

[319] Pandey, R. K.; Jagerovic, N.; Ryan, J. M.; Dougherty, T. J.; Smith, K. M., Tetrahedron, (1996) 52,<br />

5349.<br />

[320] Zheng, G.; Kozyrev, A. N.; Dougherty, T. J.; Smith, K. M.; Pandey, R. K., Chem. Lett., (1996), 1119.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1230 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a[321] Yon-Hin, P.; Wijesekera, T. P.; Dolphin, D., Tetrahedron Lett., (1991) 32, 2875.<br />

[322] Ma, L.; Dolphin, D., Can. J. Chem., (1997) 75, 262.<br />

[323] The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic: Boston, MA,<br />

(2000); Vol. 2, pp 1–300.<br />

[324] Sessler, J. L.; Weghorn, S. J., Expanded, Contracted and Isomeric Porphyrins; Tetrahedron Org. Chem.<br />

Series; Pergamon: Oxford, (1997); Vol. 15.<br />

[325] Vogel, E.; Köcher, M.; Schmickler, H.; Lex, J., Angew. Chem., (1986) 98, 262; Angew. Chem. Int. Ed.<br />

Engl., (1986) 25, 257.<br />

[326] Vogel, E.; Balci, M.; Pramod, K.; Koch, P.; Lex, J.; Ermer, O., Angew. Chem., (1987) 99, 909; Angew.<br />

Chem. Int. Ed. Engl., (1987) 26, 928.<br />

[327] Vogel, E.; Koch, P.; Hou, X.-L.; Lex, J.; Lausmann, M.; Kisters, M.; Aukauloo, M. A.; Richard, P.;<br />

Guilard, R., Angew. Chem., (1993) 105, 1670; Angew. Chem. Int. Ed. Engl., (1993) 32, 1600.<br />

[328] Barbe, J.-M.; Richard, P.; Aukauloo, M. A.; Lecomte, C.; Petit, P.; Guilard, R., J. Chem. Soc., Chem.<br />

Commun., (1994), 2757.<br />

[329] Lausmann, M.; Zimmer, I.; Lex, J.; Lueken, H.; Weinghardt, K.; Vogel, E., Angew. Chem., (1994)<br />

106, 776; Angew. Chem. Int. Ed. Engl., (1994) 33, 736.<br />

[330] Richert, C.; Wessels, J. M.; Müller, M.; Kisters, M.; Benninghaus, T.; Goetz, A. E., J. Med. Chem.,<br />

(1994) 37, 2797.<br />

[331] Sessler, J. L.; Brucker, E. A.; Weghorn, S. J.; Kisters, M.; Schäfer, M.; Lex, J.; Vogel, E., Angew. Chem.,<br />

(1994) 106, 2402; Angew. Chem. Int. Ed. Engl., (1994) 33, 2308.<br />

[332] Aukauloo, M. A.; Guilard, R., New J. Chem., (1994) 18, 1205.<br />

[333] Falk, H.; Chen, Q.-Q., Monatsh. Chem., (1996) 127, 69.<br />

[334] Vogel, E.; Will, S.; Tilling, A. S.; Neumann, L.; Lex, J.; Bill, E.; Trautwein, A. X.; Wieghardt, K., Angew.<br />

Chem., (1994) 106, 771; Angew. Chem. Int. Ed. Engl., (1994) 33, 731.<br />

[335] Paolesse, R., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 2, pp 201–232.<br />

[336] Paolesse, R.; Licocia, S.; Bandoli, G.; Dolmella, A.; Boschi, T., Inorg. Chem., (1994) 33, 1171.<br />

[337] Liccocia, S.; Tassoni, E.; Paolesse, R.; Boschi, T., Inorg. Chim. Acta, (1995) 235, 15.<br />

[338] Paolesse, R.; Tassoni, E.; Liccocia, S.; Paci, M.; Boschi, T., Inorg. Chim. Acta, (1996) 241, 55.<br />

[339] Rose, E.; Kossanyi, A.; Quelquejeu, M.; Soleilhavoup, M.; Duwavran, F.; Bernard, N.; Lecas, A., J.<br />

Am. Chem. Soc., (1996) 118, 1567.<br />

[340] Loim, N. M.; Grishko, E. V.; Pyshnograeva, N. I.; Vorontsov, E. V.; Sokolov, V. I., Izv. Akad. Nauk.<br />

Ser. Khim., (1994) 5, 925.<br />

[341] Paolesse, R.; Jaquinod, L.; Nurco, D. J.; Mini, S.; Sagone, F.; Boschi, T.; Smith, K. M., Chem. Com-<br />

mun. (Cambridge), (1999), 1307.<br />

[342] Gross, Z.; Galili, N.; Saltsman, I., Angew. Chem., (1999) 111, 1530; Angew. Chem. Int. Ed., (1999) 38,<br />

1427.<br />

[343] Conlon, M.; Johnson, A. W.; Overend, W. R.; Rajapaksa, D., J. Chem. Soc., Perkin Trans. 1, (1973),<br />

2281.<br />

[344] Vogel, E.; Bröring, M.; Fink, J.; Rosen, D.; Schmickler, H.; Lex, J.; Chan, K. W. K.; Wu, Y.-D.; Plattner,<br />

D. A.; Nendel, M.; Houk, K. N., Angew. Chem., (1995) 107, 2705; Angew. Chem. Int. Ed. Engl.,<br />

(1995) 34, 2511.<br />

[345] Johnson, A. W.; Kay, I. T., J. Chem. Soc., (1965), 1620.<br />

[346] Paolesse, R.; Pandey, R. K.; Forsyth, T. P.; Jaquinod, L.; Nurco, D. J.; Senge, M. O.; Licocia, S.; Boschi,<br />

T.; Smith, K. M., J. Am. Chem. Soc., (1996) 118, 3869.<br />

[347] Paolesse, R.; Licocia, S.; Fanciullo, M.; Morgante, E.; Boschi, T., Inorg. Chim. Acta, (1993) 203, 107.<br />

[348] Paolesse, R.; Jaquinod, L.; Senge, M. O.; Smith, K. M., J. Org. Chem., (1997) 62, 6193.<br />

[349] Pandey, R. K.; Zhou, H.; Gerzevske, K. R.; Smith, K. M., J. Chem. Soc., Chem. Commun., (1992), 183.<br />

[350] Pandey, R. K.; Gerzevske, K. R.; Zhou, H.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1994), 971.<br />

[351] Engel, J.; Gossauer, A., J. Chem. Soc., Perkin Trans. 1, (1978), 871.<br />

[352] Broadhurst, M. J.; Grigg, R.; Johnson, A. W., J. Chem. Soc., Perkin Trans. 1, (1972), 1124.<br />

[353] Woodward, R. B., In Aromaticity Special Publication No. 21; Chemical Society: London, (1966).<br />

[354] Broadhurst, M. J.; Grigg, R.; Johnson, A. W., J. Chem. Soc., Perkin Trans. 1, (1972), 2111.<br />

[355] Shiau, F.-Y.; Liddell, P. A.; Vicente, M. G. H.; Ramana, N. V.; Ramachandran, K.; Lee, S.-J.; Pandey,<br />

R. K.; Dougherty, T. J.; Smith, K. M., Proc. SPIE, (1989) 6, 71.<br />

[356] Sessler, J. L.; Lisowski, J.; Boudreaux, K. A.; Lynch, V.; Barry, J.; Kodadek, T. J., J. Org. Chem., (1995)<br />

60, 5975.<br />

[357] Shevchuk, S. V.; Davis, J. M.; Sessler, J. L., Tetrahedron Lett., (2001) 42, 2447.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


References 1231<br />

a[358] Paolesse, R.; Licoccia, S.; Spagnoli, M.; Boschi, T.; Khoury, R. G.; Smith, K. M., J. Org. Chem., (1997)<br />

62, 5133.<br />

[359] Chmielewski, P. J.; Latos-Grazynski, L.; Rachlewicz, K., Chem.–Eur. J., (1995) 1, 68.<br />

[360] Rexhausen, H.; Gossauer, A., J. Chem. Soc., Chem. Commun., (1983), 275.<br />

[361] Gossauer, A., Bull. Soc. Chim. Belg., (1983) 92, 793.<br />

[362] Gossauer, A., Chimia, (1983) 37, 341.<br />

[363] Gossauer, A., Chimia, (1984) 38, 45.<br />

[364] Burrell, A. K.; Hemmi, G.; Lynch, V.; Sessler, J. L., J. Am. Chem. Soc., (1991) 113, 4690.<br />

[365] Kral, V.; Brucker, E. A.; Hemmi, G.; Sessler, J. L.; Kralova, J.; Bose, H., Jr., Bioorg. Med. Chem., (1995)<br />

3, 573.<br />

[366] Franck, B., Angew. Chem., (1982) 94, 327; Angew. Chem. Int. Ed. Engl., (1982) 21, 343.<br />

[367] Bringmann, G.; Franck, B., Liebigs Ann. Chem., (1982), 1272.<br />

[368] Vicente, M. G. H., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 1, pp 149–199.<br />

[369] Jaquinod, L., In The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic:<br />

Boston, MA, (2000); Vol. 1, pp 201–237.<br />

[370] Buchler, J. W., In Porphyrins and Metalloporphyrins, Smith, K. M., Ed.; Elsevier: Amsterdam, (1975),<br />

pp 157–231.<br />

[371] Naruta, Y.; Tani, F.; Maruyama, K., Tetrahedron Lett., (1992) 33, 1069.<br />

[372] Andrews, L. E.; Bonnett, R.; Kozyrev, A. N.; Appelman, E. H., J. Chem. Soc., Perkin Trans. 1, (1988),<br />

1735.<br />

[373] Tsuchiya, S.; Seno, M., Chem. Lett., (1989), 263.<br />

[374] Wijesekera, T.; Matsumono, A.; Dolphin, D.; Lexa, D., Angew. Chem., (1990) 102, 1073; Angew.<br />

Chem. Int. Ed. Engl., (1990) 29, 1028.<br />

[375] Gonsalves, A. M. dA. R.; Johnstone, R. A. W.; Pereira, M. M.; Shaw, J.; Sobral, J. F. N., Tetrahedron<br />

Lett., (1991) 32, 1355.<br />

[376] Chorghade, M. S.; Dolphin, D.; DuprØ, D.; Hill, D. R.; Lee, E. C.; Wijesekera, T. P., Synthesis, (1996),<br />

1320.<br />

[377] Wijesekera, T.; DuprØ, D.; Cader, M. R.; Dolphin, D., Bull. Soc. Chim. Fr., (1996) 133, 765.<br />

[378] Vicente, M. G. H.; Smith, K. M., Tetrahedron, (1991) 47, 6887.<br />

[379] Hoffmann, P.; Robert, A.; Meunier, B., Bull. Soc. Chim. Fr., (1992) 129, 85.<br />

[380] Bonnett, R.; Gale, I. A. D.; Stephenson, G. F., J. Chem. Soc. C, (1966), 1600.<br />

[381] Ali, H.; van Lier, J. E., Tetrahedron Lett., (1991) 32, 5015.<br />

[382] Hoffmann, P.; Labat, G.; Robert, A.; Meunier, B., Tetrahedron Lett., (1990) 31, 1991.<br />

[383] Traylor, T. G.; Tsuchiya, S., Inorg. Chem., (1987) 26, 1338.<br />

[384] DiMagno, S. G.; Lin, V. S.-Y.; Therien, M. J., J. Org. Chem., (1993) 58, 5983.<br />

[385] Bonnett, R.; Campion-Smith, I. H.; Kozyrev, A. N.; Mironov, A. F., J. Chem. Res., Synop., (1990), 138;<br />

J. Chem. Res., Miniprint, (1990), 1015.<br />

[386] Tagliatesta, P.; Li, J.; Autret, M.; Van Caemelbecke, E.; Villard, A.; DSouza, F.; Kadish, K. M., Inorg.<br />

Chem., (1996) 35, 5570.<br />

[387] Shea, K. M.; Jaquinod, L.; Khoury, R. G.; Smith, K. M., Chem. Commun. (Cambridge), (1998), 759.<br />

[388] Bhyrappa, P.; Krishan, V., Inorg. Chem., (1991) 30, 239.<br />

[389] Hariprasad, G.; Dahal, S.; Maiya, B. G., J. Chem. Soc., Dalton Trans., (1996), 3429.<br />

[390] Jaquinod, L.; Khoury, R. G.; Shea, K. M.; Smith, K. M., Tetrahedron, (1999) 55, 13151.<br />

[391] Crossley, M. J.; Harding, M. M.; Sternhell, S., J. Am. Chem. Soc., (1986) 108, 3608.<br />

[392] Boyle, R. W.; Johnson, C. K.; Dolphin, D., J. Chem. Soc., Chem. Commun., (1995), 527.<br />

[393] Vicente, M. G. H.; Smith, K. M., Synlett, (1990), 579.<br />

[394] Vicente, M. G. H.; Jaquinod, L.; Smith, K. M., Chem. Commun. (Cambridge), (1999), 1771.<br />

[395] Hyslop, A. G.; Kellett, M. A.; Iovine, P. M.; Therien, M. J., J. Am. Chem. Soc., (1998) 120, 12676.<br />

[396] Callot, H. J., Tetrahedron Lett., (1973), 4987.<br />

[397] Arnold, D. P.; Bott, R. C.; Eldridge, H.; Elms, F. M.; Smith, G.; Zojaji, M., Aust. J. Chem., (1997) 50,<br />

495.<br />

[398] Crossley, M. J.; Burn, P. L.; Chew, S. S.; Cuttance, F. B.; Newsom, I. A., J. Chem. Soc., Chem. Commun.,<br />

(1991), 1564.<br />

[399] Wickramasinghe, A.; Jaquinod, L.; Nurco, D. J.; Smith, K. M., Tetrahedron, (2001) 57, 4261.<br />

[400] Dahal, S.; Krishnan, V., Photochem. Photobiol. A: Chem., (1995) 89, 105.<br />

[401] Bartoli, J.-F.; Battioni, P.; De Foor, W. R.; Mansuy, D., J. Chem. Soc., Chem. Commun., (1994), 23.<br />

[402] Bonnett, R.; Stephenson, G. F., J. Org. Chem., (1965) 30, 2791.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1232 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a[403] Johnson, A. W.; Oldfield, D., Tetrahedron Lett., (1964), 1549.<br />

[404] Johnson, A. W.; Oldfield, D., J. Chem. Soc., (1965), 4303.<br />

[405] Drach, J. E.; Longo, F. R., J. Org. Chem., (1974) 39, 3282.<br />

[406] Watanabe, E.; Nishimura, S.; Ogoshi, H.; Yoshida, Z., Tetrahedron, (1975) 31, 1385.<br />

[407] Cavaleiro, J. A. S.; Neves, M. G. P. M. S.; Hewlins, M. J. E.; Jackson, A. H., J. Chem. Soc., Perkin Trans.<br />

1, (1986), 575.<br />

[408] Fanning, J. C.; Mandel, F. S.; Gray, T. L.; Datta-Gupta, N., Tetrahedron, (1979) 35, 1251.<br />

[409] Gong, L.-C.; Dolphin, D., Can J. Chem., (1985) 63, 401.<br />

[410] Catalano, M. M.; Crossley, M. J.; Harding, M. M.; King, L. G., J. Chem. Soc., Chem. Commun., (1984),<br />

1535.<br />

[411] Shea, K. M.; Jaquinod, L.; Smith, K. M., J. Org. Chem., (1998) 63, 7013.<br />

[412] Baldwin, J. E.; Crossley, M. J.; DeBernardis, J. F., Tetrahedron, (1982) 38, 685.<br />

[413] Barnett, G. H.; Smith, K. M., J. Chem. Soc., Chem. Commun., (1974), 772.<br />

[414] Osuka, A.; Shimidzu, H., Angew. Chem., (1997) 109, 93; Angew. Chem. Int. Ed. Engl., (1997) 36, 135.<br />

[415] Ozette, K.; Battioni, P.; Leduc, P.; Bartoli, J.-F.; Mansuy, D., Inorg. Chim. Acta, (1998) 272, 4.<br />

[416] Crossley, M. J.; Govenlock, L. J.; Prashar, J. K., J. Chem. Soc., Chem. Commun., (1995), 2379.<br />

[417] Baldwin, J. E.; DeBernardis, J. F., J. Org. Chem., (1977) 42, 3986.<br />

[418] Hombrecher, H. K.; Gherdan, V. M.; Ohm, S.; Cavaleiro, J. A. S.; Neves, M. G. P. M. S.; Condesso, M.<br />

F., Tetrahedron, (1993) 49, 8569.<br />

[419] Hombrecher, H. K.; Gherdan, V. M.; Cavaleiro, J. A. S.; Neves, M. G. P. M. S., J. Prakt. Chem., (1994)<br />

336,1.<br />

[420] Crossley, M. J.; Hambley, T. W.; Mackay, L. G.; Try, A. C.; Walton, R., J. Chem. Soc., Chem. Commun.,<br />

(1995), 1077.<br />

[421] Johnson, C. K.; Dolphin, D., Tetrahedron Lett., (1998) 39, 4619.<br />

[422] Crossley, M. J.; King, L. G.; Pyke, S. M., Tetrahedron, (1987) 43, 4569.<br />

[423] Crossley, M. J.; Burn, P. L.; Langford, S. J.; Pyke, S. M.; Stark, A. G., J. Chem. Soc., Chem. Commun.,<br />

(1991), 1567.<br />

[424] Crossley, M. J.; Gosper, J. J.; Wilson, M. G., J. Chem. Soc., Chem. Commun., (1985), 1798.<br />

[425] Crossley, M. J.; King, L. G., J. Org. Chem., (1993) 58, 4370.<br />

[426] Crossley, M. J.; Harding, M. M.; Tansey, C. W., J. Org. Chem., (1994) 59, 4433.<br />

[427] Crossley, M. J.; King, L. G., J. Chem. Soc., Perkin Trans. 1, (1996), 1251.<br />

[428] Crossley, M. J.; King, L. G.; Simpson, J. L., J. Chem. Soc., Perkin Trans. 1, (1997), 3087.<br />

[429] Jaquinod, L.; Gros, C. P.; Khoury, R. G.; Smith, K. M., Chem. Commun. (Cambridge), (1996), 2581.<br />

[430] Jaquinod, L.; Gros, C. P.; Olmstead, M. M.; Antolovich, M.; Smith, K. M., Chem. Commun. (Cam-<br />

bridge), (1996), 1475.<br />

[431] Gros, C. P.; Jaquinod, L.; Khoury, R. G.; Olmstead, M. M.; Smith, K. M., J. Porphyrins Phthalocya-<br />

nines, (1997) 1, 201.<br />

[432] Smith, K. M.; Bisset, G. M. F.; Case, J. J.; Tabba, H. D., Tetrahedron Lett., (1980) 21, 3747.<br />

[433] Smith, K. M.; Bisset, G. M. F.; Tabba, H. D., J. Chem. Soc., Perkin Trans. 1, (1982), 581.<br />

[434] Momenteau, M.; Loock, B.; Bisagni, E.; Rougee, M., Can. J. Chem., (1979) 57, 1804.<br />

[435] Buchler, J. W.; Dreher, C.; Herget, G., Liebigs Ann. Chem., (1988), 43.<br />

[436] Henrich, K.; Owston, P. G.; Peters, R.; Tasker, P. A.; Dell, A., Inorg. Chim. Acta, (1980) 45, L161.<br />

[437] Brockmann, H., Jr.; Bliesener, K.-M.; Inhoffen, H. H., Justus Liebigs Ann. Chem., (1968) 718, 148.<br />

[438] Smith, K. M.; Langry, K. C., J. Chem. Soc., Perkin Trans. 1, (1983), 439.<br />

[439] Montforts, F.-P.; Scheurich, G.; Meier, A.; Haake, G.; Hoper, F., Tetrahedron Lett., (1991) 32, 3477.<br />

[440] Nichol, A. W., J. Chem. Soc. C, (1970), 903.<br />

[441] Vicente, M. G. H.; Rezzano, I. N.; Smith, K. M., Tetrahedron Lett., (1990) 31, 1365.<br />

[442] Vicente, M. G. H.; Smith, K. M., J. Org. Chem., (1991) 56, 4407.<br />

[443] Johnson, A. W.; Oldfield, D., J. Chem. Soc. C, (1966), 794.<br />

[444] Clezy, P. S.; Lim, C. L.; Shannon, J. S., Aust. J. Chem., (1974) 27, 1103.<br />

[445] Abraham, R. J.; Eivazi, F.; Nayyir-Mazhir, R.; Pearson, H.; Smith, K. M., Org. Magn. Reson., (1978)<br />

11, 52.<br />

[446] Callot, H. J., Tetrahedron, (1973) 29, 899.<br />

[447] Callot, H. J., Bull. Soc. Chim. Fr., (1973) 12, 3413.<br />

[448] Callot, H. J.; Catro, B.; Selve, C., Tetrahedron Lett., (1978), 2877.<br />

[449] Dolphin, D.; Sivasothy, R., Can. J. Chem., (1981) 59, 779.<br />

[450] Ponomarev, G. V.; Maravin, G. B., Chem. Heterocycl. Comp. (Engl. Transl.), (1982) 18, 50.<br />

[451] Arnold, D. P.; Nitschinsk, L. J., Tetrahedron, (1992) 48, 8781.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


References 1233<br />

a[452] Gosper, J. J.; Ali, M., J. Chem. Soc., Chem. Commun., (1994), 1707.<br />

[453] Kahl, S. B.; Schaeck, J. J.; Koo, M.-S., J. Org. Chem., (1997) 62, 1875.<br />

[454] Burrell, A. K.; Officer, D. L., Synlett, (1998), 1297.<br />

[455] Arnold, D. P.; Nitschinsk, L. J., Tetrahedron Lett., (1993) 34, 693.<br />

[456] Taylor, P. N.; Huuskonen, J.; Rumbles, G.; Aplin, R. T.; Williams, E.; Anderson, H. L., Chem. Commun.<br />

(Cambridge), (1998), 909.<br />

[457] Arnold, D. P.; Johnson, A. W.; Mahendran, M., J. Chem. Soc., Perkin Trans. 1, (1978), 366.<br />

[458] Runge, S.; Senge, M. O., Tetrahedron, (1999) 55, 10375.<br />

[459] Hoper, F.; Montforts, F.-P., Liebigs Ann. Chem., (1995), 1033.<br />

[460] Smith, K. M.; Bisset, G. M. F.; Bushell, M. J., Bioorg. Chem., (1980) 9,1.<br />

[461] Smith, N. W.; Smith, K. M., Energy & Fuels, (1990) 4, 675.<br />

[462] Arnold, D. P.; Johnson, A. W.; Winter, M., J. Chem. Soc., Perkin Trans. 1, (1977), 1643.<br />

[463] Ponomarev, G. V.; Borovkov, V. V.; Sugiura, K.-I.; Sakata, Y.; Shulga, A. M., Tetrahedron Lett., (1993)<br />

34, 2153.<br />

[464] Shulga, A. M.; Ponomarev, G. V., Chem. Heterocycl. Comp. (Engl. Transl.), (1988) 24, 276.<br />

[465] Higuchi, H.; Shimizu, K.; Ojima, J.; Sugiura, K.-I.; Sakata, Y., Tetrahedron Lett., (1995) 36, 5359.<br />

[466] Yashunsky, D. V.; Ponomarev, G. V.; Arnold, D. P., Tetrahedron Lett., (1995) 36, 8485.<br />

[467] Zhilina, Z. I.; Ishkov, Y. V.; Voloshanovskii, I. S.; Andronati, A. S. A., Dokl. Akad. Nauk USSR, (1988)<br />

303, 326.<br />

[468] Senge, M. O.; Gerzevske, K. R.; Vicente, M. G. H.; Forsyth, T. P.; Smith, K. M., Angew. Chem., (1993)<br />

105, 745; Angew. Chem. Int. Ed. Engl., (1993) 32, 750.<br />

[469] Jaquinod, L.; Nurco, D. J.; Medforth, C. J.; Pandey, R. K.; Forsyth, T. P.; Olmstead, M. M.; Smith, K.<br />

M., Angew. Chem., (1996) 108, 1085; Angew. Chem. Int. Ed. Engl., (1996) 35, 1013.<br />

[470] Smith, K. M.; Bisset, G. M. F., J. Org. Chem., (1979) 44, 2077.<br />

[471] Smith, K. M.; Bisset, G. M. F.; Bushell, M. J., Int. J. Biochem., (1980) 12, 695.<br />

[472] Smith, K. M.; Bisset, G. M. F., J. Chem. Soc., Perkin Trans. 1, (1981), 2625.<br />

[473] Bonfantini, E. E.; Officer, D. L., Tetrahedron Lett., (1993) 34, 8531.<br />

[474] Omote, M.; Ando, A.; Koyama, M.; Takagi, T.; Kumadaki, I., Heterocycles, (1994) 39, 381.<br />

[475] Smith, K. M.; Fujinari, E. M.; Langry, K. C.; Parish, D. W.; Tabba, H. D., J. Am. Chem. Soc., (1983)<br />

105, 6638.<br />

[476] Jeandon, C.; Bauder, C.; Callot, H. J., Energy & Fuels, (1990) 4, 665.<br />

[477] Shiau, F.-Y.; Whyte, B. J.; Castelfranco, P. A.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1991),<br />

1781.<br />

[478] Kusch, D.; Montforts, F.-P., Tetrahedron: Asymmetry, (1995) 6, 867.<br />

[479] Kusch, D.; Tollner, E.; Lincke, A.; Montforts, F.-P., Angew. Chem., (1995) 107, 874; Angew. Chem.<br />

Int. Ed. Engl., (1995) 34, 784.<br />

[480] Kusch, D.; Montforts, F.-P., Liebigs Ann., (1996), 83.<br />

[481] Jiang, X.; Smith, K. M., J. Chem. Soc., Chem. Commun., (1993), 1054.<br />

[482] Jiang, X.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1996), 1601.<br />

[483] Smith, K. M.; Langry, K. C., J. Org. Chem., (1983) 48, 500.<br />

[484] Smith, K. M.; Langry, K. C.; Minnetian, O. M., J. Org. Chem., (1984) 49, 4602.<br />

[485] Minnetian, O. M.; Morris, I. K.; Snow, K. M.; Smith, K. M., J. Org. Chem., (1989) 54, 5567.<br />

[486] Morris, I. K.; Snow, K. M.; Smith, N. W.; Smith, K. M., J. Org. Chem., (1990) 55, 1231.<br />

[487] Jiang, X.; Pandey, R. K.; Smith, K. M., Tetrahedron Lett., (1995) 36, 365.<br />

[488] Jiang, X.; Pandey, R. K.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1996), 1607.<br />

[489] Callot, H. J., J. Chem. Soc., Chem. Commun., (1975), 163.<br />

[490] McLaughlin, G. M., J. Chem. Soc., Perkin Trans. 2, (1974), 136.<br />

[491] Callot, H. J.; Tschamber, T., J. Am. Chem. Soc., (1975) 97, 6175.<br />

[492] Smith, K. M.; Pandey, R. K.; Snow, K. M., J. Chem. Soc., Chem. Commun., (1986), 1498.<br />

[493] Grigg, R., J. Chem. Soc. C, (1971), 3664.<br />

[494] Ichimura, K., Bull. Chem. Soc. Jpn., (1978) 51, 1444.<br />

[495] Callot, H. J.; Chevrier, B.; Weiss, R., J. Am. Chem. Soc., (1978) 100, 4733.<br />

[496] Callot, H. J., Tetrahedron, (1979) 35, 1455.<br />

[497] Callot, H. J., Tetrahedron Lett., (1972), 1011.<br />

[498] Gong, L.-C.; Dolphin, D., Can. J. Chem., (1985) 63, 406.<br />

[499] Evans, B.; Smith, K. M.; Cavaleiro, J. A. S., J. Chem. Soc., Perkin Trans. 1, (1978), 768.<br />

[500] Barnett, G. H.; Evans, B.; Smith, K. M.; Besecke, S.; Fuhrhop, J.-H., Tetrahedron Lett., (1976), 4009.<br />

[501] Fuhrhop, J.-H.; Wanja, U.; Bunzel, M., Liebigs Ann. Chem., (1984), 426.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


1234 Science of Synthesis 17.8 Porphyrins and Related Compounds<br />

a[502] Fuhrhop, J.-H.; Kruger, W.; Meding, U., Liebigs Ann. Chem., (1981), 1367.<br />

[503] Shine, H. J.; Padilla, A. G.; Wu, S.-M., J. Org. Chem., (1979) 44, 4069.<br />

[504] El Kahef, L.; Giraudeau, A., Can. J. Chem., (1991) 69, 1161.<br />

[505] Smith, K. M.; Barnett, G. H.; Evans, B.; Martynenko, Z., J. Am. Chem. Soc., (1979) 101, 5953.<br />

[506] Ruhlman, L.; Giraudeau, A., Chem. Commun., (1996), 2007.<br />

[507] Giraudeau, A.; Ruhlman, L.; El Kahef, L.; Gross, M., J. Am. Chem. Soc., (1996) 118, 2969.<br />

[508] Ruhlman, L.; Lobstein, S.; Gross, M.; Giraudeau, A., J. Org. Chem., (1999) 64, 1352.<br />

[509] Ogawa, T.; Nishimoto, Y.; Yoshida, N.; Ono, N.; Osuka, A., Angew. Chem., (1999) 111, 140; Angew.<br />

Chem. Int. Ed., (1999) 38, 176.<br />

[510] Yoshida, N.; Shimidzu, H.; Osuka, A., Chem. Lett., (1998), 55.<br />

[511] Nakano, A.; Shimidzu, H.; Osuka, A., Tetrahedron Lett., (1998) 39, 9489.<br />

[512] Dolphin, D.; Felton, R. H.; Borg, D. C.; Fajer, J., J. Am. Chem. Soc., (1970) 92, 743.<br />

[513] Lee, W. A.; Bruice, T. C., Inorg. Chem., (1986) 25, 131.<br />

[514] Ogoshi, H.; Sugimoto, H.; Yoshida, Z.-I.; Kobayashi, H.; Sakai, H.; Maeda, Y., J. Organomet. Chem.,<br />

(1982) 234, 185.<br />

[515] Cocolios, P.; Lagrange, G.; Guilard, R., J. Organomet. Chem., (1983) 253, 65.<br />

[516] Callot, H. J.; Schaeffer, E., Tetrahedron Lett., (1980) 21, 1335.<br />

[517] Dolphin, D.; Halko, D. J.; Johnson, E., Inorg. Chem., (1981) 20, 4348.<br />

[518] Lange, M.; Mansuy, D., Tetrahedron Lett., (1981) 22, 2561.<br />

[519] Ortiz de Montellano, P. R.; Kunze, K. L.; Augusto, O., J. Am. Chem. Soc., (1982) 104, 3545.<br />

[520] Callot, H. J.; Metz, F., J. Chem. Soc., Chem. Commun., (1982), 947.<br />

[521] Setsune, J.-I.; Dolphin, D., Can. J. Chem., (1987) 65, 459.<br />

[522] Kadish, K. M.; Caemelbecke, E. V.; DSouza, F.; Medforth, C. J.; Smith, K. M.; Tabard, A.; Guilard, R.,<br />

Organometallics, (1993) 12, 2411.<br />

[523] Setsune, J.-I.; Yazawa, T.; Ogoshi, H.; Yoshida, Z.-I., J. Chem. Soc., Perkin Trans. 1, (1980), 1641.<br />

[524] Kalisch, W. W.; Senge, M. O., Angew. Chem., (1998) 110, 1156; Angew. Chem. Int. Ed., (1998) 37,<br />

1107.<br />

[525] Senge, M. O.; Feng, X., Tetrahedron Lett., (1999) 40, 4165.<br />

[526] Krattinger, B.; Callot, H. J., Tetrahedron Lett., (1996) 37, 7699.<br />

[527] Segawa, H.; Azumi, R.; Shimidzu, T., J. Am. Chem. Soc., (1992) 114, 7564.<br />

[528] Jiang, X.; Nurco, D. J.; Smith, K. M., Chem. Commun. (Cambridge), (1996), 1759.<br />

[529] McCombie, S. W.; Smith, K. M., Tetrahedron Lett., (1972), 2463.<br />

[530] Barnett, G. H.; Hudson, M. F.; McCombie, S. W.; Smith, K. M., J. Chem. Soc., Perkin Trans. 1, (1973),<br />

691.<br />

[531] Bonnett, R.; Cornell, P.; McDonagh, A. F., J. Chem. Soc., Perkin Trans. 1, (1976), 794.<br />

[532] Jackson, A. H.; Rao, K. R. N.; Wilkins, M., J. Chem. Soc., Perkin Trans. 1, (1987), 307.<br />

[533] Vicente, M. G. H.; Jaquinod, L.; Khoury, R. G.; Madrona, A. Y.; Smith, K. M., Tetrahedron Lett.,<br />

(1999) 40, 8763.<br />

[534] Knapp, S.; Vasudevan, J.; Emge, T. J.; Arison, B. H.; Potenza, J. A.; Schugar, H. J., Angew. Chem.,<br />

(1998) 110, 2537; Angew. Chem. Int. Ed., (1998) 37, 2368.<br />

[535] Tome, A. C.; Lacerda, P. S. S.; Neves, M. G. P. M. S.; Cavaleiro, J. A. S., Chem. Commun. (Cambridge),<br />

(1997), 1199.<br />

[536] Vicente, M. G. H.; Cancilla, M. T.; Lebrilla, C. B.; Smith, K. M., Chem. Commun. (Cambridge), (1998),<br />

2355.<br />

[537] Gunter, M. J.; Tang, H.; Warrener, R. N., Chem. Commun. (Cambridge), (1999), 803.<br />

[538] Morgan, A. R.; Skalkos, D.; Maguire, G.; Rampersaud, A.; Garbo, G. M.; Keck, R.; Selman, S. H.,<br />

Photochem. Photobiol., (1992) 55, 133.<br />

[539] Gunter, M. J.; Robinson, B. C.; Gulbis, J. M.; Tiekink, E. R. T., Tetrahedron, (1991) 47, 7853.<br />

[540] Osuka, A.; Ikawa, Y.; Maruyama, K., Bull. Chem. Soc. Jpn., (1992) 65, 3322.<br />

[541] Lee, D. A.; Brisson, J. M.; Smith, K. M., Heterocycles, (1995) 40, 131.<br />

[542] Boyle, R. W.; Dolphin, D., J. Chem. Soc., Chem. Commun., (1994), 2463.<br />

[543] Morgan, A. R.; Gupta, S., Tetrahedron Lett., (1994) 35, 4291.<br />

[544] Yashunsky, D. V.; Ponomarev, G. V.; Moskovkin, A. S.; Arnold, D. P., Aust. J. Chem., (1997) 50, 487.<br />

[545] Bauder, C.; Ocampo, R.; Callot, H. J., Synlett, (1990), 335.<br />

[546] Faustino, M. A. F.; Neves, M. G. P. M. S.; Vicente, M. G. H.; Silva, A. M. S.; Cavaleiro, J. A. S., Tetrahedron<br />

Lett., (1995) 36, 5977.<br />

[547] Ishkov, Y. V.; Zhilina, Z. I., Zh. Org. Khim., (1995) 31, 136.<br />

[548] Callot, H. J.; Schaeffer, E.; Cromer, R.; Metz, F., Tetrahedron, (1990) 46, 5253.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag


References 1235<br />

a[549] Barloy, L.; Dolphin, D.; DuprØ, D.; Wijesekera, T. P., J. Org. Chem., (1994) 59, 7976.<br />

[550] Atkinson, E. J.; Brophy, J. J.; Clezy, P. S., Aust. J. Chem., (1990) 43, 383.<br />

[551] Krattinger, B.; Nurco, D. J.; Smith, K. M., Chem. Commun. (Cambridge), (1998), 757.<br />

K. M. Smith and M. G. H. Vicente, Section 17.8, Science of Synthesis, 2003 Georg Thieme Verlag

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!