12.01.2013 Views

Macrocyclic Ligands - Web del Profesor

Macrocyclic Ligands - Web del Profesor

Macrocyclic Ligands - Web del Profesor

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

<strong>Macrocyclic</strong> <strong>Ligands</strong><br />

Kristin Bowman-James<br />

University of Kansas, Lawrence, KS, USA<br />

1 Introduction 1<br />

2 Classification of <strong>Ligands</strong> 1<br />

3 Synthesis 5<br />

4 Thermodynamics and Structural Aspects 9<br />

5 Applications 17<br />

6 Related Articles 18<br />

7 References 18<br />

Glossary<br />

Bis-macrocycles: two macrocycles joined together<br />

Calixarenes: basket-shaped macrocycles with phenyl<br />

backbones<br />

Catenands: two interlocked macrocycles<br />

Compartmental ligands: macrocycles with ‘compartments’<br />

for housing more than one substrate<br />

Crown ethers: polyoxa macrocycles<br />

Cryptands: bicyclic macrocycles with aza bridgeheads<br />

Cyclidenes: lacunar tetraaza macrocycles<br />

Expanded porphyrins: macrocycles based on pyrrolic<br />

frameworks<br />

Lariat ethers: crown ethers with pendant chains<br />

Sepulchrates: bicyclic caged macrocycles<br />

Spherands: macrocycles with phenyl backbones<br />

Abbreviations<br />

Polyaza macrocycles: [n]aneNm: n = Number of ring atoms;<br />

Nm = Number of nitrogen atoms; Polyoxa macrocycles: ncrown-m:<br />

n = Number of ring atoms; m = Number of oxygen<br />

atoms; TRI = Tribenzo[b,f ,j][1,5,9]triazacyclododecine;<br />

TAAB = Tetrabenzo[b,f ,j,n][1,5,9,13]tetraazacyclohexadecine.<br />

1 INTRODUCTION<br />

<strong>Macrocyclic</strong> ligands are defined as cyclic molecules<br />

generally consisting of organic frames into which heteroatoms,<br />

capable of binding to substrates, have been interspersed.<br />

Some reports of ‘synthetic’ macrocycles (as opposed to the<br />

naturally occurring species such as porphyrins, corrins, and<br />

chlorins) appeared as early as 1936, when the first synthesis of<br />

1,4,8,11-tetraazacyclotetradecane was reported. 1 Nonetheless,<br />

the field only began to blossom in the early 1960s with the<br />

pioneering work of Busch 2 and Curtis’ discovery of the nickelmediated<br />

condensation of [Ni(en)3] 2+ with acetone. 3 The<br />

early macrocycles were synthesized with an eye to mimicking<br />

biologically occurring macrocycles such as the porphyrins,<br />

corrins, chlorins, and, more recently, the corphins.<br />

Another area of macrocyclic development began in the late<br />

1960s and initial applications were focused toward mo<strong>del</strong>ing<br />

biological processes such as ion transport. These macrocycles<br />

initially included the oxygen-based crown ethers of Pedersen, 4<br />

and the mixed oxygen–nitrogen bicyclic cryptands of Lehn, 5<br />

both of which exhibit high selectivity toward alkali and<br />

alkaline earth metal ions. Several years later, the concept<br />

of ‘preorganized’ cavities resulted in the synthesis of the<br />

cavitands by Cram. 6<br />

Since its birth, the development of macrocyclic chemistry<br />

has proceeded along two lines:<br />

1. as mo<strong>del</strong>s of the naturally occurring macrocyclic systems,<br />

containing predominantly nitrogen donor atoms; and<br />

2. as receptors designed for recognition and supramolecular<br />

chemistry, with a variety of donor atoms and recognition<br />

capabilities.<br />

<strong>Macrocyclic</strong> chemistry has expanded phenomenally since<br />

the 1960s to provide exciting and novel chemistry. The award<br />

of the 1987 Nobel Prize in Chemistry to Pedersen, Lehn,<br />

and Cram is testimony to the importance of this rapidly<br />

expanding field.<br />

In such a large subject, this article can only focus on<br />

certain aspects, namely those that involve complexation<br />

with inorganic substrates. We only consider the synthetic<br />

macrocycles, with emphasis on transition metal complexation.<br />

Aza, oxa, and, to a lesser extent, thia and phospha<br />

macrocycles are also covered. The naturally occurring<br />

porphyrins, corrins, corphins, chlorins, and phthalocyanins, 7<br />

as well as the cyclodextrins, 8 are not included. Because of the<br />

general complexity of macrocyclic systems and the resulting<br />

complicated systematic names, commonly used abbreviations<br />

or simplified names will be employed. This review will<br />

encompass the synthesis, thermodynamics, structure, and<br />

applications of macrocyclic ligands.<br />

2 CLASSIFICATION OF LIGANDS<br />

Two major areas of complexation have developed over<br />

the years with regard to synthetic macrocycles. Those with<br />

nitrogen, sulfur, phosphorus, and arsenic tend predominantly<br />

to form traditional covalent coordination complexes with<br />

transition metal ions. A notable exception to this tendency,<br />

however, is the rapidly expanding chemistry of the


2 MACROCYCLIC LIGANDS<br />

polyammonium macrocycles, which are capable of forming<br />

a variety of complexes with anionic substrates. Oxygenderived<br />

macrocycles are noted for complexation with alkali<br />

and alkaline earth metal ions, as well as with organic cations<br />

and molecular substrates. In this latter situation, associations<br />

tend to be electrostatic in nature, and in many instances<br />

hydrogen-bonding interactions are vital to complex formation.<br />

<strong>Macrocyclic</strong> ligands will be classified, for the purposes of<br />

this article, as rings with at least nine members and three or<br />

more donor atoms. In a number of cases of unique structural<br />

units, elegant descriptive names have developed, which more<br />

appropriately describe the macrocyclic shape. Macrocycles<br />

will be classified as to donor types and, within the donor<br />

types, specific classifications of macrocycles will be noted<br />

where applicable.<br />

2.1 Polyaza Macrocycles (1)–(10) 9–18<br />

2.1.1 Simple Polyaza Macrocycles<br />

Until recently, the tetraaza macrocycles, such as (1)<br />

(cyclam) and related ligands with extensive varieties of<br />

modifications including differing degrees of saturation and<br />

ring size (2), had been the most studied, primarily because<br />

of the relationship of these molecules to naturally occurring<br />

tetraaza macrocycles, such as the porphyrins and corrins.<br />

Currently, with interest in metal–metal interactions, increased<br />

activity has occurred in the area of larger macrocycles<br />

capable of incorporating more than one metal ion, such as<br />

(3) ([24]aneN8). 18 Interest in the smaller triaza macrocycles,<br />

such as (4) ([9]aneN3) and its variations, has also accelerated<br />

in recent years. 14 Added to the simple polyaza macrocycles has<br />

been the effort to achieve functionalized macrocycles in order<br />

to expand the chemistry of these ligands by combining the<br />

rigid structural aspects of the macrocyclic ring with the more<br />

flexible and kinetically labile properties of pendant chains, as<br />

in (5). 11<br />

NH<br />

NH<br />

(1)<br />

HN<br />

HN<br />

2.1.2 Cyclidenes<br />

N<br />

N<br />

(2)<br />

N<br />

N<br />

Me<br />

Me<br />

NH<br />

NH<br />

NH HN<br />

NH<br />

(3)<br />

HN<br />

HN<br />

HN<br />

Cyclidenes (6) are a subset of the polyaza macrocycles<br />

and are the lacunar ligands first synthesized and extensively<br />

NH HN<br />

H<br />

N<br />

(4)<br />

HO 2C<br />

HO 2C<br />

N<br />

N<br />

(5)<br />

N<br />

N<br />

CO2H<br />

CO2H<br />

studied by Busch. 19 They coordinate a single metal ion and<br />

maintain a ‘persistent void’ which allows access to small<br />

molecules within the vaulted cavity.<br />

N<br />

N<br />

2.1.3 Sepulchrates<br />

N<br />

N<br />

N<br />

N<br />

R<br />

R<br />

NH HN<br />

NH<br />

NH HN<br />

HN<br />

R<br />

(6) (7)<br />

Sepulchrates (7) are polyaza cage macrocycles. They are<br />

noted for their exceptionally strong hold on encapsulated metal<br />

ions. 20<br />

2.1.4 Expanded Porphyrins<br />

Expanded porphyrins are macrocycles based on the pyrrolic<br />

backbone of porphyrins, but are expanded in size to achieve a<br />

larger cavity (8) 21 or binucleating capabilities (9). 22<br />

NH<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

NH NH<br />

HN<br />

HN<br />

(8) (9)<br />

N<br />

N


2.1.5 Bis-Macrocycles<br />

Bis-macrocycles (10) provide another mechanism for<br />

achieving complexation of more than one metal ion. They<br />

are joined by a bridge linking two simple macrocycles. 13,23<br />

Me<br />

N N<br />

N<br />

N<br />

(CH 2) 2<br />

N<br />

Me<br />

(10)<br />

2.2 Polythia, Polyphospha, and Polyarsa Macrocycles<br />

Polythia macrocycles (11), the thioether analogs of the<br />

crown ethers, have been known since the 1930s. 24 These are<br />

the most extensively studied macrocycles in line after the<br />

polyoxa and polyaza macrocycles.<br />

S<br />

S<br />

S<br />

S<br />

N<br />

(11) (12)<br />

The ‘pure’ polyphospha macrocycles (12) (as opposed to<br />

the mixed donor phospha macrocycles) were first reported in<br />

1975. 25 These macrocycles have been found to complex a<br />

variety of transition metals, but have not received the same<br />

attention as the more readily accessible polyaza and polyoxa<br />

macrocycles.<br />

The polyarsa macrocycles (13) comprise one of the least<br />

common type of macrocycles. 26<br />

P<br />

P<br />

P<br />

P<br />

As<br />

As<br />

As<br />

MACROCYCLIC LIGANDS 3<br />

(13)<br />

2.3 Mixed Donor Macrocycles<br />

As<br />

As<br />

As<br />

2.3.1 Simple Mixed Donor Macrocycles<br />

The simple mixed donor macrocycles (14) at one time were<br />

the major source of study of the influence of the incorporation<br />

of ‘soft’ phosphorus and arsenic donors into macrocycles. 27<br />

Mixed oxygen–nitrogen macrocycles have been studied quite<br />

extensively, since they serve as bridges for examining the<br />

coordination tendencies of the aza macrocycles and the oxa<br />

crown ethers. 13<br />

P<br />

P<br />

2.3.2 Cryptands<br />

O<br />

O<br />

HN<br />

HN<br />

HN<br />

N<br />

O<br />

O O<br />

O<br />

(14) (15)<br />

Cryptands (15) are bicyclic macrocycles which can contain<br />

a variety of donor atoms with bridgehead nitrogen atoms. 5<br />

They are highly selective for alkali and alkaline earth<br />

metal ions.<br />

2.3.3 Compartmental <strong>Ligands</strong><br />

Compartmental ligands (16) are macrocyclic ligands (as<br />

well as nonmacrocyclic ligands) which contain ‘compartments’<br />

for housing more than one metal ion. 28 Only the<br />

macrocyclic counterparts will be treated here.<br />

O<br />

O<br />

N


4 MACROCYCLIC LIGANDS<br />

N<br />

N<br />

2.3.4 Catenands<br />

Me<br />

N OH N<br />

N<br />

OH<br />

Me<br />

(16)<br />

O<br />

O O<br />

O<br />

O<br />

O<br />

O O<br />

O<br />

O<br />

(17)<br />

Catenands (17) are interlocked macrocyclic ligands, which<br />

complex a variety of metal ions. 29<br />

2.4 Polyoxa Macrocycles<br />

Polyoxa macrocycles, known more commonly as the crown<br />

ethers, comprise an extensive area of research, with a repertoire<br />

of types and variations. 30 Some of these macrocycles have<br />

been utilized predominantly for purposes other than metal ion<br />

complexation, and these will not be discussed in depth in this<br />

review. Included in this latter category are the polycarbonyls, 31<br />

polylactones, 32 polylactams 33 and carcerands. 34<br />

2.4.1 Polyether Macrocycles<br />

Polyether macrocycles (18) are the simplest of the polyoxa<br />

macrocycles. The commonly used name for these macrocycles<br />

is the crown ethers, due to their crown-like structure in the<br />

solid state. These molecules have been extensively studied<br />

N<br />

O<br />

O<br />

N<br />

N<br />

O<br />

O<br />

O<br />

O<br />

(18)<br />

as complexing agents for the alkali and alkaline earth metal<br />

ions. 30<br />

2.4.2 Lariat Ethers<br />

The lariat ethers comprise a subset of the polyether<br />

macrocycles, and are identified by their pendant chains. 35<br />

They can be categorized as either N-pivot (19) orC-pivot<br />

(20), depending on which type of atom the chain is attached.<br />

As for their polyether parents, much of the focus on these<br />

macrocycles has been on complexation of alkali and alkaline<br />

earth metal ions.<br />

O<br />

O<br />

O<br />

O<br />

(19)<br />

N<br />

O<br />

2.4.3 Spherands and Hemispherands<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

(20)<br />

O<br />

O<br />

MeO<br />

These consist of an arrangement of phenyl groups which<br />

provide a preorganized cavity for complexation, 36 e.g. (21)<br />

and (22).<br />

Me Me<br />

Me<br />

Me<br />

OMe<br />

OMe MeO<br />

OMe MeO<br />

OMe<br />

Me<br />

(21)<br />

Me


Me<br />

2.4.4 Calixarenes<br />

O<br />

Me<br />

OMe<br />

OMe MeO<br />

O<br />

(22)<br />

Calixarenes (23) are the macrocyclic result of condensations<br />

between phenols and formaldehyde 37 and have been<br />

referred to as the most easily accessible molecular basket. 38<br />

t-Bu<br />

t-Bu<br />

OH<br />

OH HO<br />

OH<br />

t-Bu<br />

(23)<br />

H 2N<br />

H2N<br />

O<br />

NH<br />

H<br />

N<br />

HN<br />

Me<br />

t-Bu<br />

NH 2<br />

NH2<br />

NH<br />

NH<br />

NH<br />

NH<br />

3 SYNTHESIS<br />

3.1 Polyaza Macrocycles<br />

MACROCYCLIC LIGANDS 5<br />

3.1.1 Conventional (Nontemplate) Syntheses<br />

Reviews of synthetic procedures can be found for tridentate<br />

and pentadentate macrocyclic ligands with nitrogen donors,<br />

mixed nitrogen donors, and sulfur donor macrocycles, 39 the<br />

techniques of which can be expanded to other ring sizes. The<br />

general procedures will be summarized below.<br />

Cyclic secondary amines, [n]aneNm, are generally prepared<br />

by macrocyclization reactions known as the Richman–Atkins<br />

procedure. 40 These reactions involve ring closure by<br />

condensation of two precursor fragments of the cyclic<br />

molecule. In general, one fragment consists of a salt<br />

of a sulfonamide, while the other contains two terminal<br />

leaving groups, which can vary in identity and include<br />

chloride, bromide, hydroxide, or, more often, a sulfonate ester<br />

(Scheme 1). The reaction is performed in polar aprotic solvents<br />

and may involve high dilution techniques. Simplified routes<br />

to tri-, tetra-, and pentaaza systems have been described. 41 A<br />

handy synthetic technique for the smaller triaza ring has been<br />

described by Alder, where the macrocycle is built by using<br />

a single carbon as template. 42 Treatises on the synthesis of<br />

pyridine-containing macrocycles 43 and imidazole-containing<br />

macrocycles 44 have also been reported.<br />

Functionalized macrocycles (5) with additional ligating<br />

components attached as pendant arms have been an area<br />

of focus in efforts to expand the chemistry of macrocyclic<br />

receptors by incorporating additional recognition sites.<br />

Synthetic techniques for N-functionalized, C-functionalized,<br />

NTs<br />

TsN −<br />

Scheme 1<br />

Ts<br />

N<br />

TsN<br />

OMs MsO<br />

HN<br />

NTs TsN<br />

HN<br />

HN<br />

− NTs


6 MACROCYCLIC LIGANDS<br />

and bis-macrocycles are known. 11 N-Alkylation, the more<br />

common of the functionalization routes, is usually achieved<br />

by alkylation or acylation of the amino nitrogens using<br />

a variety of agents such as chloroacetic acid, ethylene<br />

oxide, acrylonitrile, and formaldehyde. 11 Routes for selective<br />

alkylations are known, including the use of selective protection<br />

and deprotection techniques. 45<br />

Bis-macrocycles (10), another extension of the concept of<br />

functionalized macrocycles, can be made by several different<br />

methods. One of the more commonly used is the condensation<br />

of two precursor chains already joined by the linker. 23<br />

Bis-macrocycles also can be formed from cyclam (1) asa<br />

by-product bis-macrocycle in low yield (24). 46<br />

NH<br />

NH<br />

HN<br />

HN<br />

(24)<br />

NH<br />

NH<br />

3.1.2 Template-Mediated Syntheses<br />

HN<br />

HN<br />

Metal ion template mediation in macrocyclic synthesis has<br />

been a part of the field since its inception, its importance<br />

having been realized early in the development of this area.<br />

Two specific roles for the metal ion in template reactions have<br />

been proposed. These are, in turn, kinetic and thermodynamic<br />

in origin. 47 In the kinetic template effect, the arrangement of<br />

ligands already coordinated to the metal ion provides control<br />

in a subsequent condensation during which the macrocycle<br />

is formed. The thermodynamic effect serves to promote<br />

stabilization of a structure which would not be favored in<br />

the absence of a metal ion. Schiff base condensations tend<br />

to be dependent on this latter type of template effect. Some<br />

of the more routine and general synthetic procedures will be<br />

described here. A more in-depth treatment can be found in a<br />

review by Curtis, with particular emphasis on general methods<br />

as well as modifications of preformed macrocycles. 48<br />

Carbonyl compounds are commonly used precursors for<br />

the polyaza macrocycles, as in the classic synthesis of<br />

the Curtis ligand from the condensation of acetone with<br />

Ni(en)3 (Scheme 2). 3 2,6-Diacetylpyridine has provided the<br />

Ni<br />

H2<br />

N<br />

N<br />

H 2<br />

3<br />

O<br />

N<br />

HN<br />

Ni<br />

Scheme 2<br />

NH<br />

N<br />

+<br />

HN<br />

HN<br />

Ni<br />

N<br />

N<br />

Me<br />

N<br />

Me<br />

O O<br />

+<br />

H2N NH2 N<br />

H<br />

small M n+<br />

large M n+<br />

Scheme 3<br />

Me<br />

N<br />

Me<br />

N<br />

M<br />

N<br />

Me<br />

N<br />

Me<br />

N<br />

M<br />

N<br />

Me<br />

NH HN<br />

precursor for a number of pyridine-derived imine macrocycles<br />

(Scheme 3), where smaller metal ions as templates tend to<br />

implement the formation of 1:1 condensates, while larger<br />

metal ions allow for 2:2 stoichiometries. 49 Schiff base<br />

condensations can be considered as another variation of<br />

the carbonyl condensations (Scheme 4). 50 The expanded<br />

porphyrins (8) and(9) 21,22 and the compartmental ligands<br />

(16) 28 are usually synthesized by Schiff base condensations.<br />

In some instances the macrocyclic analog cannot be obtained<br />

via other methods. This is true for the self-condensation of<br />

o-aminobenzaldehyde, which can yield both tridentate (TRI)<br />

and tetradentate (TAAB) macrocycles (25) and(26). 51<br />

N N<br />

N<br />

(25)<br />

N<br />

N<br />

N<br />

(26)<br />

Template-assisted condensations of amines with formaldehyde<br />

yield a wide variety of macrocyclic products. Sepulchrates<br />

(7) can be synthesized from the template-assisted<br />

condensation of [Co(en)3] 3+ with formaldehyde and ammonia<br />

under basic conditions. 20 Primary aldehydes other than<br />

formaldehyde have also been used. 52 Caged metal ion complexes<br />

in which the metal ion is used as a template are normally<br />

N<br />

N<br />

N H<br />

M<br />

N<br />

N<br />

Me


N<br />

N<br />

CHO<br />

H2N +<br />

CHO<br />

H2N<br />

M n+<br />

extremely inert, so much so that removal of the metal is often<br />

impossible. Nonmetal template syntheses of polyaza cages<br />

have also been reported. 53 A number of interesting variations<br />

utilizing the template-assisted condensation of formaldehyde<br />

and amines have also resulted in structurally new macrocycles<br />

such as (27). 54<br />

Me<br />

N<br />

NH HN<br />

N N<br />

N<br />

(27)<br />

3.2 Polythia, Polyphospha, and Polyarsa Macrocycles<br />

One of the reasons for the relative ‘late-blooming’ of the<br />

thioether macrocycles can be found in synthetic difficulties.<br />

While the polyaza and polyoxa macrocycles can often utilize<br />

template effects in controlling the critical condensations,<br />

polythia condensations are more limited in this area. In<br />

general, these macrocycles are made from condensation of the<br />

appropriate polythiane with a dibromoalkane (Scheme 5). 55<br />

Synthetic procedures and yields have been greatly enhanced by<br />

the addition of high dilution techniques. 56,57 A cage-like sulfur<br />

macrocycle has been reported as an analog of the nitrogencontaining<br />

sepulchrates (28). 58 Mixed nitrogen–sulfur cages<br />

can also be obtained. 58<br />

S<br />

S<br />

N<br />

S<br />

S<br />

N<br />

(28)<br />

N<br />

S<br />

S<br />

Scheme 4<br />

N<br />

N N<br />

M M<br />

N<br />

N N<br />

Ph<br />

N<br />

N<br />

S S<br />

SNa NaS<br />

+<br />

Br Br<br />

P P<br />

Ni<br />

PH HP<br />

Ph Ph<br />

+<br />

Br Br<br />

Ph<br />

MACROCYCLIC LIGANDS 7<br />

Scheme 5<br />

Scheme 6<br />

Ph<br />

S S<br />

S S<br />

P P<br />

Ph<br />

Ni<br />

Ph<br />

P P<br />

Ph<br />

Polyphospha macrocycles can be made via template<br />

condensations of coordinated polyphosphine ligands and<br />

dibromoalkanes (Scheme 6). 59,60<br />

Polyarsa macrocycles can be made by the reaction of<br />

lithiated polyarsanes with a dichloroalkane (Scheme 7). 26,60<br />

3.3 Mixed Donor Macrocycles<br />

Simple mixed donor macrocycles, such as aza–oxa,<br />

aza–thia, oxa–thia, and analogous phospha and arsa analogs<br />

are generally achieved via combinations of the routes used for<br />

synthesis of the ‘pure’ donor analogs. Since the possibilities<br />

are so extensive they will not be treated here, but are found<br />

elsewhere. 16,60 New mixed donor phosphorus techniques<br />

have been devised for phospha–thia and phospha–aza<br />

macrocycles. 61,62


8 MACROCYCLIC LIGANDS<br />

PhAs As(Ph)Li<br />

As(Ph)Li<br />

+<br />

Cl Cl<br />

Scheme 7<br />

PhAs<br />

As<br />

Ph<br />

AsPh<br />

Cryptands are usually synthesized via sequential condensations<br />

between a diamine and an acid chloride, which<br />

yields a diamide, followed by reduction with LiAlH4<br />

to give the macromonocycle. Condensation with another<br />

H2N O O NH2 O O<br />

Cl O O Cl<br />

HN<br />

O O<br />

O O<br />

NH<br />

HN<br />

+<br />

O O<br />

Cl O O Cl<br />

N<br />

O O<br />

O O<br />

O O<br />

O<br />

O O<br />

O<br />

equivalent of acyl chloride will yield the bicyclic precursor,<br />

which can be reduced by B2H6 to give the bicycle<br />

(Scheme 8). 5<br />

Compartmental ligands (16) are derived from diketonates<br />

and triketonates and are usually synthesized from Schiff base<br />

reactions of the ketone with a diamine. 28<br />

Catenands (17) are also made using the metal ion template<br />

effect. A bis-complex is formed from an α,α ′ -disubstituted<br />

o-phenanthroline. Then the initial product is treated with a<br />

diiodoalkane to accomplish the ring closure. 29<br />

3.4 Polyoxa Macrocycles<br />

Polyethers were originally synthesized by template<br />

assistance from an oligo(ethylene glycol), monoglyme, and<br />

potassium t-butoxide (Scheme 9). 4<br />

O O<br />

HN<br />

NH<br />

O O<br />

O O<br />

NH<br />

N<br />

O O<br />

O<br />

N<br />

Scheme 8<br />

O<br />

O O<br />

O<br />

O<br />

N<br />

B2H 6<br />

LAH


HO O O OH<br />

+<br />

Cl O O Cl<br />

Scheme 9<br />

Spherands and hemispherands (21)and(22) are synthesized<br />

by ring closure reactions of aryllithium with Fe(acac)3, often<br />

using high dilution techniques. 36<br />

Calixarenes (23) are obtained from base-catalyzed condensations<br />

of p-substituted phenols with formaldehyde. 37<br />

4 THERMODYNAMICS AND STRUCTURAL<br />

ASPECTS<br />

4.1 Introduction<br />

4.1.1 The <strong>Macrocyclic</strong> Effect<br />

This term refers to the amazing stability of macrocyclic<br />

ligands. It was initially described in studies of tetraaza<br />

macrocycles with copper(II). 63 For polyaza macrocycles this<br />

effect has been attributed to both entropic and enthalpic<br />

considerations and considerable controversy raged for a<br />

number of years as to which was the predominant factor. 64,65<br />

The conflicting reports are now realized to be extremely<br />

dependent on the experimental methods used for the<br />

determination of the thermodynamic parameters. Two main<br />

types of technique have been employed, each of which has its<br />

strengths and weaknesses: the calorimetric titration method<br />

and the use of the temperature variation of the stability<br />

constants. The controversy has been largely settled by more<br />

recent studies. 66,67 Important contributions to the enthalpic<br />

term are now attributed to a number of factors, including<br />

solvation and conformation changes upon bond formation.<br />

Likewise, the entropic considerations include the number of<br />

species present and particularly solvation effects. Detailed<br />

discussions of the historical development can be found. 13,17<br />

Related to the macrocyclic effect are the decreased rates<br />

of dissociation observed for macrocyclic complexes. Busch<br />

and co-workers have coined a term to describe these longterm<br />

stabilities incurred by synthetic macrocycles: multiple<br />

juxtapositional fixedness. The premise is that straight-chain<br />

ligands can undergo dissociative displacements in consecutive<br />

steps starting at one end of the ligand and finishing with the<br />

opposite end. This is not the case for macrocyclic ligands, for<br />

which each dissociated donor is still held in proximity to the<br />

metal ion by the rest of the ligand framework. 68<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

MACROCYCLIC LIGANDS 9<br />

The macrocyclic effect has been observed for polyaza,<br />

polythia, and polyoxa, as well as mixed donor atom,<br />

macrocycles. 69<br />

4.1.2 Selectivity<br />

The selectivity of a macrocycle for either a metal ion or<br />

another substrate is critically dependent on the structure of<br />

the macrocycle and electronic effects, i.e. the types of donor<br />

atoms. Some of the important aspects are described below.<br />

1. The number of binding sites is perhaps one of the<br />

most crucial influences on the binding properties of the<br />

substrate. Electronic effects of the binding of macrocycles<br />

with substrates are charge, polarity, and polarizability<br />

of the binding sites. For metal ion binding, this means<br />

ion pair interactions for negatively charged ligands,<br />

ion–dipole and ion–induced dipole interactions for<br />

neutral ligands, and the hard–soft acid–base criteria.<br />

Nitrogen, phosphorus, and sulfur donors are noted for<br />

their complexation of transition metal ions. Oxygen is<br />

more likely to complex alkali or alkaline earth metal ions.<br />

2. The arrangement of the binding sites should be such as to<br />

maximize the potential ligand–metal ion interactions. In<br />

this regard the selection of spacers between donor atoms<br />

to allow for the formation of five- and six-member chelate<br />

rings has been the most utilized.<br />

3. The preferred conformations of the macrocycles dictate<br />

its propensity to bind a metal ion internally or externally<br />

to the cavity. The propensity of the lone pair to point in<br />

or out of the cavity is also a deciding factor. Hence, it is<br />

not always a foregone conclusion that the metal ion will<br />

be bound within the macrocyclic cavity.<br />

4. The identity of the macrocyclic framework also plays<br />

a major role in structure. For example, saturated<br />

hydrocarbon chains provide considerably more flexibility<br />

than incorporated aromatic units. Likewise the presence<br />

of other functional groups, such as amides or esters, serve<br />

to stiffen the macrocyclic framework. Decreasing the<br />

flexibility of the macrocycle by adding selected ‘shaping<br />

groups’ is the theory behind preorganization, so important<br />

in the cavitands. Another method of creating rigidity is to<br />

increase the dimensionality of the macrocycle, inherent in<br />

cryptand selectivities.<br />

5. The size of the macrocyclic cavity also plays a large role in<br />

governing the flexibility of the ligand, and its propensity<br />

for metal ion binding.<br />

Since the focus of this article is primarily on transition metal<br />

chemistry, the structural aspects related to complexation of<br />

transition metals will be emphasized, and other aspects of<br />

complexation will only be briefly treated.<br />

In addition to the traditional measurement of thermochemical<br />

properties, molecular mechanics calculations are now<br />

available to supplement and correlate with experimental findings.<br />

An extensive review which links the large data base of


10 MACROCYCLIC LIGANDS<br />

thermodynamic and kinetic data with items such as ring size,<br />

number and arrangement of ligand binding sites, and solvent<br />

effects, for all types of donor atoms including coronands,<br />

cryptands, spherands, and nitrogen donors can be found. 69 A<br />

more recent series of molecular mechanics calculations have<br />

added to this base of thermochemical data and point to structural<br />

factors affecting complex stabilities from the viewpoint<br />

of steric strain. 70<br />

4.2 Polyaza Macrocycles<br />

An extensive review of the thermodynamic aspects of<br />

polyaza macrocycles has been reported. 17 Other reviews<br />

include the chemistry of tridentate and pentadentate aza<br />

macrocycles, 16 1,4,7-triazacyclononane and derivatives, 14 and<br />

polyaza macrocycles with pendant chains. 11 In general, the<br />

polyaza macrocycles form extremely stable complexes with<br />

transition metals of the later transition series, but show reduced<br />

affinity for alkali and alkaline earth metal ions compared to<br />

the oxa macrocycles.<br />

4.2.1 Triaza Macrocycles<br />

One of the simplest and smallest of the polyaza macrocycles<br />

according to definition is 1,4,7-triazacyclononane (4). The<br />

geometrical constraints of the triaza macrocycles are such<br />

that they do not allow for the incorporation of the metal ion<br />

within the macrocyclic ring. Hence, these macrocycles are<br />

facially coordinated in either a mono- or bis-ligand complex<br />

with a variety of metal ions. 14 The macrocyclic effect is<br />

observed, and the stability constants of the complexes follow<br />

the Irving–Williams Series. 17 Both microcalorimetric 71 and<br />

stability constant determinations at different temperatures 72<br />

indicate that the effect is most probably enthalpic in origin.<br />

The triaza macrocycles also form extremely stable complexes<br />

with the heavier main group metals (such as Ga III , In III ,<br />

Tl I ,andTl III ) as well as transition metals. The chemistry<br />

of this macrocycle and its derivatives is wide in scope and<br />

is treated extensively in a review which includes the base<br />

compound and its N-functionalized derivatives. 14 Depending<br />

on the appendages employed in N-functionalization, three<br />

more coordination sites are potentially available, rounding out<br />

the coordination to pseudooctahedral.<br />

The formation of dinuclear and higher nuclearity species<br />

is common for the mono-coordinated triazacyclononane, with<br />

bridging acetates, hydroxides, and oxides being very common.<br />

Extensive studies of the chemistry of the variety of bridged<br />

species have been made using the relatively substitutionally<br />

inert chromium(III) ion. Different dimers and trimers have<br />

been isolated and structurally characterized, as in (29) and<br />

(30). 73,74 Higher nuclear clusters such as an octanuclear iron<br />

system are relevant as a mo<strong>del</strong> for the iron storage protein<br />

ferritin (see Iron Proteins for Storage & Transport & their<br />

Synthetic Analogs). 75<br />

N<br />

N<br />

Cr<br />

N O<br />

H<br />

O<br />

O<br />

H<br />

O<br />

(29)<br />

N<br />

Cr<br />

O<br />

N<br />

N<br />

N<br />

N<br />

N<br />

Cr<br />

N<br />

H<br />

O N<br />

O<br />

H Cr<br />

OH<br />

Cr<br />

N<br />

N<br />

(30)<br />

A more ‘preorganized’ ligand system is derived from the<br />

self-condensation of o-aminobenzaldehyde. 51 The tridentate<br />

form of the ligand (25) (TRI) imparts considerable<br />

inertness toward substitution. For example, the salts of<br />

the [Ni(TRI)(H2O)3] 2+ ion can be resolved into optical<br />

isomers. 76 A copper(II) complex of the methyl-substituted<br />

tetradentate macrocycle Me4TAAB, in which bis-coordination<br />

occurs, displays a dynamic Jahn–Teller distortion based on<br />

crystallographic evidence. 77<br />

4.2.2 Tetraaza Macrocycles<br />

Because of the potential relationship to the naturally<br />

occurring porphyrins and porphyrin-analog macrocycles, the<br />

tetraaza macrocycles have been the focus of much attention.<br />

Tetraaza macrocycles often, but not always, form a coplanar<br />

arrangement of the four nitrogen donors. Empirical force field<br />

calculations of free macrocycles from 12- to 16-membered<br />

rings indicate that cyclam (1) exhibits the least strain with the<br />

best planarity. A straightforward assessment of the relationship<br />

of hole size to selectivity is complicated by the conformational<br />

flexibility of the ligands. Results of studies for the tetraaza<br />

macrocycles show that hole size does not appear to be the<br />

predominant factor in metal ion discrimination. Rather, the<br />

selectivity of these macrocycles is governed by the relative<br />

stability of the conformers of the macrocycle which have<br />

different metal ion size preferences. An interesting observation<br />

regarding the relationship of the tetraaza macrocycles with<br />

regard to hole size and metal ion selectivity can be found<br />

for the most studied of the simple tetraaza macrocycles,<br />

cyclam. Cyclam is proposed to have five configurational<br />

isomers, based on the orientation of the amine hydrogens.<br />

From molecular mechanics calculations, where the best M–N<br />

distance is calculated as that giving the minimum energy,<br />

the trans-III analog of [12]aneN4 (31) is found to have an<br />

extremely high strain energy of 19.7 kcal mol −1 with a best-fit<br />

M–N distance of 1.81 ˚A, compared to the trans-I form (32)<br />

(10.8 kcal mol −1 and 2.11 ˚A, respectively). 70,78 In general, the<br />

larger, more flexible, planar coordination is provided by the<br />

trans-I conformer, and often if a metal is too large for the<br />

macrocyclic cavity, it will coordinate lying out of the plane of<br />

the donor atoms. When the metal ion is not incorporated into<br />

the macrocyclic plane, the factors influencing stability are the<br />

same as for the acyclic aza analogs, namely that for larger<br />

metal ions, as the size of the chelate ring increases from five<br />

N<br />

N


to six, the complex stability decreases. A detailed discussion<br />

of the thermodynamics of changing chelate sizes for tetraaza<br />

macrocycles can be found. 17,78<br />

M N<br />

N<br />

H H<br />

H<br />

H<br />

H<br />

H N<br />

M<br />

N N<br />

N<br />

(31) (32)<br />

N<br />

N<br />

H<br />

H<br />

An elegant example of the importance of conformational<br />

changes in tetradentate macrocycles is the blue to yellow<br />

conversion observed for nickel(II) complexes. The yellow<br />

form is the low-spin square planar complex NiL 2+ , while the<br />

blue form is high-spin pseudooctahedral [NiL(H2O)2] 2+ .In<br />

the blue to yellow conversion the Ni–N bonds contract, which<br />

compensates for the breaking of the axial Ni–OH2 bonds. The<br />

reaction is controlled by entropy, and the addition of an inert<br />

salt is such as to favor the dissociation of the water molecules.<br />

At equilibrium in aqueous solution, both [12]aneN4 (cyclam)<br />

(1) and the 15-membered analog, [15]aneN4, have 99% of<br />

the high-spin form present, while the 13- and 14-membered<br />

macrocycles exist in predominantly the low-spin square planar<br />

form (87 and 71%, respectively). 79<br />

For the nonplanar octahedral cis-coordinated macrocycles<br />

[n]aneN4, changes in the ligand field correlate well with the<br />

analogous ligand field strengths for nonmacrocyclic analogs,<br />

specifically as related to the chelate ring size. A general rule<br />

of thumb is that increasing the chelate ring size from five<br />

to six increases the stability of complexes of smaller metals<br />

compared to larger metal ions. The origin of this effect can be<br />

attributed to increases in ring strain energy when metal ions<br />

larger than tetrahedral carbon are part of the ring. 78 For the<br />

planar-coordinated macrocycles, the equatorial ligand field,<br />

as anticipated, is dependent on the ring size. These findings<br />

have been related to the calculation of the optimum hole size<br />

permitting the macrocycle to adopt its most preferable endo<br />

configuration. Thus, it has been found that the macrocyclic<br />

hole size increases by 10–15 pm for each increment in n for<br />

[n]aneN4. 79<br />

In order to introduce a greater rigidity into the<br />

flexible polyaza macrocycles and to implement greater<br />

hole size–metal ion size match correlations, reinforced<br />

macrocycles such as (33) have been created, which contain<br />

fused diaza rings. 80 Crystallographic results for the nickel(II)<br />

complex indicate that the Ni–N bonds are shortened from<br />

the strain-free value of 1.91 ˚A for diamagnetic nickel to<br />

1.86 ˚A. Ligand field strength is also found to increase, and<br />

this has been suggested as being due to the compression<br />

of the bond lengths 80 as well as the presence of tertiary<br />

nitrogen donors. 78 In a more recent comparative study of<br />

nickel macrocycles with two fused 1,3-diazacyclohexane rings<br />

(35) compared to two fused 1,3-diazacyclopentane rings (34),<br />

MACROCYCLIC LIGANDS 11<br />

structural results revealed weaker ligand field strengths for the<br />

1,3-diazacyclohexane compared to 1,3-diazacyclopentane. 54<br />

N<br />

N<br />

NH HN<br />

(33)<br />

N<br />

NH<br />

N<br />

N<br />

(34)<br />

HN<br />

N<br />

N<br />

NH<br />

N<br />

N<br />

(35)<br />

HN<br />

Attempts to achieve macrocycles that are capable of<br />

stabilizing highly oxidized transition metal complexes has<br />

led to the design of ‘noninnocent’ ligands. 81 The structures of<br />

high-valent chromium(V) oxo species with the two tetraamido<br />

N ligands (36)and(37) were determined. Both structures were<br />

found to contain distinctly nonplanar amide groups, and in<br />

(36) all four amides are nonplanar.<br />

O<br />

NH<br />

NH<br />

O<br />

HN<br />

HN<br />

O<br />

Cl Cl<br />

NH<br />

N<br />

O NH HN O<br />

(36) (37)<br />

HN<br />

Polyaza macrocycles with pendant arms have been studied<br />

extensively, in particular with respect to protonation and<br />

complexation as well as to the kinetics of metal complex<br />

formation. These aspects are treated in a review by Kaden. 11<br />

Of particular interest is the fact that metal complex formation<br />

constants of macrocycles with pendant carboxylates can be<br />

10 3 to 10 4 times higher than for the unsubstituted analogs.<br />

4.2.3 Higher Polyaza Macrocycles<br />

Transition metal complexes of the larger polyaza<br />

macrocyclic ligands have been less extensively studied than<br />

for the smaller ring systems. For the pentaaza macrocycles,<br />

[15]aneN5 with ethylene bridges appears to form the most<br />

stable complexes with most metal ions. 17 Structural data<br />

for a variety of pentaaza macrocyclic complexes have<br />

been reviewed. 16 The N–H bonds as well as the different<br />

sized chelate rings must be considered in calculating the


12 MACROCYCLIC LIGANDS<br />

number of possible isomers. For each of the complexes,<br />

three configurations of the in-plane N–H bonds are possible:<br />

(38), (39), and (40). Crystallographic data indicate that<br />

most of the complexes with pentadentate macrocycles have<br />

pseudooctahedral geometries. Pentadentate macrocycles also<br />

tend to stabilize unusual oxidation states. For example, the<br />

nickel(II) complexes of [15]aneN5 (41), [16]aneN5 (42), and<br />

one of the isomers of [17]aneN5 (43), are readily oxidized<br />

to the Ni III analogs. Also, there is little dependence of E1/2<br />

values on the macrocyclic ring size, which has been attributed<br />

to the absence of in-plane ring size effects. 16<br />

NH<br />

H N N H N<br />

N<br />

H<br />

N<br />

M<br />

N<br />

N<br />

N<br />

H<br />

N<br />

M<br />

N<br />

N<br />

N<br />

N<br />

M<br />

N<br />

N<br />

H<br />

H<br />

(38) meso–syn (39) meso–anti (40) racemic<br />

NH HN<br />

NH<br />

HN<br />

NH<br />

NH HN<br />

NH<br />

HN<br />

NH<br />

NH HN<br />

NH<br />

(41) (42) (43)<br />

HN<br />

The hexaaza [18]aneN6 forms complexes with transition<br />

metal ions and with certain alkali and alkaline earth and<br />

lanthanide ions. 82 For the higher aza macrocycles with<br />

seven or more donor atoms, dinuclear complexes become<br />

possible. A systematic investigation of both the structural<br />

and thermodynamic aspects of copper complexes formed with<br />

the larger polyaza macrocycles from heptaaza to dodecaaza<br />

has been published. 18 All of the macrocycles were found to<br />

form hydroxo species as well as polynuclear complexes. A<br />

number of structures have been determined for the higher<br />

polyaza macrocycles, both in complexed and noncomplexed<br />

forms, and structures range from highly boat shaped to nearly<br />

planar. 18,50,83,84<br />

A review of macrocycles possessing subheterocyclic rings<br />

has appeared, which includes pyridine, furan, and thiophene. 85<br />

In a study of formation constants for transition metal ions<br />

with pyridine- and furan-containing macrocycles, (44) and<br />

(45), it was found that the pyridine macrocycles follow the<br />

Irving–Williams series and bind even more effectively than<br />

their saturated analogs (i.e. [18]aneN6 and [18]aneN4O2). The<br />

furan analogs showed little tendency to bind, which has been<br />

attributed to the increased rigidity of the furan ring. 86<br />

NH<br />

NH<br />

N<br />

N<br />

HN<br />

4.2.4 Anion Coordination<br />

NH<br />

HN NH<br />

O<br />

O<br />

(44) (45)<br />

HN<br />

HN<br />

While the initial interest in polyaza macrocycles involved<br />

metal ion coordination, the finding in 1968 by Simmons 87<br />

that diaza bicyclic catapinands can incorporate halide ions<br />

into their cavity opened the door on a vast new area of<br />

chemistry, that of anion complexation. The thermodynamics<br />

of anion binding can be divided into several different<br />

areas: that of simple inorganic anions; more complex<br />

carboxylate and polycarboxylates; corresponding phosphates,<br />

polyphosphates, and nucleotides; and culminating in anionic<br />

metal complexes. 17,88–90 Binding is accomplished via both<br />

electrostatic and hydrogen-bonding interactions between the<br />

protonated macrocyclic amines and the anionic substrates.<br />

The general trend appears to be that the increased flexibility<br />

of larger polyammonium macrocycles tends to facilitate<br />

complexation of more complex anionic substrates.<br />

The results of studies for complexes formed between<br />

polyammonium macrocycles and transition metal complex<br />

anions indicate that cation–anion electrostatic attraction is a<br />

crucial factor in complexation reactions and serves to regulate<br />

the stoichiometry of the complexes formed. Hydrogenbonding,<br />

size, and conformational factors also play major<br />

roles. 89 Anions can be incorporated in or out of the ring.<br />

Two illustrative examples are metal ion complexes with the<br />

octaprotonated macrocycle H8[30]aneN10 (46). In the complex<br />

with Co(CN)6 3− , the anion lies outside the macrocycle. The<br />

PdCl4 2− complex is a true ‘inclusion’ situation, however, in<br />

which the PdCl4 2− is situated along the minor axis of the<br />

macrocyclic cavity, and the Cl atoms are out of the frame,<br />

forming strong hydrogen bonds with the polyammonium<br />

sites. 90<br />

4.2.5 Cyclidenes<br />

Crystallographic results for the cyclidenes (6) show that<br />

a wide variety of structural ranges can result from designed<br />

modifications of the lacunar cavity (or void). 91 The affinity<br />

of the cobalt(II) complexes of the cyclidenes for molecular<br />

oxygen was found to be very dependent on the identity of the<br />

overhead bridge and was found to increase with increasing<br />

bridge length. Further design has also allowed for expanding


NH<br />

NH<br />

NH<br />

NH<br />

NH<br />

NH<br />

(46)<br />

HN<br />

HN<br />

HN<br />

HN<br />

the capability of these macrocycles beyond simple oxygen<br />

binding to oxygenase activity observed for the cytochrome<br />

P-450s. This has been achieved by adding piperazine ‘risers’<br />

to increase the cavity size (9 ˚A high) as well as increasing the<br />

hydrophobicity of the molecules, and by adding anthracene<br />

and durene ‘roofs’. The crystal structure of the anthracenebridged<br />

derivative shows that the macrocycle is indeed capable<br />

of hosting an acetonitrile molecule.<br />

4.2.6 Sepulchrates<br />

Sepulchrates (7) are the most noted of the caged<br />

macrocyclic ligands and are the nitrogen analogs of the<br />

cryptands. The Co–N distances are 1.99 ˚AforCo III and 2.16 ˚A<br />

for Co II from crystallographic data, and do not vary greatly<br />

from other cobalt amines. 20<br />

4.2.7 Expanded Porphyrins<br />

A review of expanded porphyrin ligands can be found. 92<br />

The texaphyrins (8) can be considered as 22-π-electron<br />

benzannulene systems with an 18-π-electron <strong>del</strong>ocalization<br />

path, based on crystal structure data as well as NMR.<br />

The cadmium complex of the macrocycle is found to be<br />

planar with pentadentate coordination of the macrocycle to<br />

cadmium, which becomes seven-coordinate as a result of<br />

axial coordination to two pyridine molecules. The cavity is<br />

nearly circular with a center-to-nitrogen distance of 2.39 ˚A.<br />

Because of the larger size of this macrocycle, metal ion<br />

coordination is generally seen with the larger transition metals<br />

and lanthanides.<br />

A more flexible expanded porphyrin is the ‘accordion’<br />

porphyrin (9). 22 The structural aspects of this macrocycle<br />

illustrate the importance of flexibility in achieving unanticipated<br />

structures. The free-base macrocycle is elliptical with<br />

the inclusion of two water molecules (47), while the dicopper(II)<br />

complex is highly distorted by means of exo and endo<br />

orientations of the imine groups (48).<br />

N 3<br />

N<br />

N N<br />

Cu<br />

MACROCYCLIC LIGANDS 13<br />

Ph<br />

NHNH<br />

O<br />

N<br />

H HN<br />

N H H N<br />

O<br />

HN HN<br />

N<br />

Ph<br />

(47)<br />

(48)<br />

N<br />

N<br />

Cu<br />

N<br />

N3<br />

4.3 Polythia and Polyphospha Macrocycles<br />

4.3.1 Polythia Macrocycles<br />

The coordination chemistry of thioether macrocycles has<br />

expanded greatly only since the mid-1980s, as seen by a number<br />

of reviews. 55,93–95 The macrocyclic effect is also noted for<br />

thioethers, but to a lesser extent than some of the other macrocyclic<br />

ligands. This is due primarily to the reorganizational<br />

energy requirements, since a number of the free-ligand thia<br />

macrocycles have a tendency to adopt ‘exodentate’ conformations<br />

in the uncomplexed form, where the sulfurs are pointed<br />

out of the macrocycle (49). <strong>Macrocyclic</strong> thioethers must then<br />

undergo a reorganization of their exo lone pairs in order to<br />

incorporate metal ions within the cavity. It was found in a study<br />

of the complexation of a number of open-chain thia ligands and<br />

thia macrocycles that the enthalpy changes were essentially<br />

identical for both macrocyclic and nonmacrocyclic ligands.<br />

Hence, the favorable macrocyclic effect is more attributable to<br />

the entropy changes in the sulfur macrocycles. 96 The smaller<br />

trithia analog of the extensively studied nitrogen donor triazacyclononane<br />

does not require such organization and, as such,<br />

has been extensively studied itself. 55 Because of the preference<br />

for exodentate sulfurs, metal ion coordination in many<br />

cases is external to the cavity (50). 97 A comprehensive review<br />

of the structural aspects of thia macrocycles can be found. 55<br />

Considerable effort has been made with regard to<br />

conformation analysis of crown thioethers. It has been found<br />

that ligand strain is most evident in torsion angles, whereby<br />

an examination of the deviations from the optimum values of<br />

N


14 MACROCYCLIC LIGANDS<br />

S<br />

S<br />

S<br />

S Cl<br />

Cl<br />

Hg<br />

S<br />

S<br />

(49) (50)<br />

S<br />

S<br />

Hg Cl<br />

Cl<br />

60 ◦ for gauche or 180 ◦ for anti configurations can lead to an<br />

assessment of the overall strain in the molecules. 93<br />

Examination of the influence of ring size has been<br />

reported for the 12- to 16-membered tetrathia systems with<br />

copper(II). 96,98 The results indicate a marked interrelationship<br />

between ring size and stability. The stability peaks at 14membered<br />

rings, and the rings are large enough to incorporate<br />

the copper only for the 14- to 16-membered systems.<br />

Results from the correlation of stability constants in<br />

conjunction with redox data have led to insights regarding the<br />

coordination chemistry of thia macrocycles. For example, the<br />

electrochemical behavior of a number of copper(II)/(I) redox<br />

couples has been investigated, 99 and redox potentials as well<br />

as protonation and stability constants of Cu I species were<br />

determined for a number of tetradentate and pentadentate thiaderived<br />

macrocycles with thia- and mixed thia–aza rings with<br />

the basic backbones (51)and(52). Results of the examination<br />

of the stability constants in conjunction with the Cu II/I redox<br />

potentials indicate that the stability constants for the Cu I<br />

oxidation state are relatively constant regardless of the mixing<br />

in of nitrogen donor atoms. Hence, the dramatic increase in<br />

the Cu II/I redox potential which is observed in the presence of<br />

the sulfur macrocycles can be attributed to a destabilization of<br />

the Cu II state rather than stabilization of the Cu I state, contrary<br />

to popular belief from the hard–soft acid–base system.<br />

S S<br />

S<br />

S<br />

S<br />

S S<br />

S<br />

(51) (52)<br />

In order to force binding of trithia structural units into<br />

an endodentate conformation, one strategy has been to add<br />

rigid xylyl groups into the ring to limit the flexibility (53). 100<br />

While the conformation of the free ligands is exodentate,<br />

a number of transition metal complexes of this ligand have<br />

been found to exhibit endodentate coordination, including Mo,<br />

Cu, Ag, Pd, and Rh. Results for the bis-macrocyclic silver<br />

complex with a variety of noncoordinating anions, indicate<br />

that the conformational interconversions of the ligand are low<br />

in energy.<br />

S<br />

S<br />

S<br />

(53)<br />

S<br />

S<br />

S<br />

S<br />

(54)<br />

Thiophene units have also been incorporated into the thia<br />

crowns (54). 101<br />

4.3.2 Polyphospha Macrocycles<br />

Phosphorus macrocycles can exist in a variety of<br />

conformations, a number of which are stable. The barrier<br />

for inversion of phosphate is 146.4 kJ mol −1 . 102 Hence there<br />

are five conformations possible for the tetraphosphorus<br />

macrocycle (12). Two are preferred: the one in which the<br />

macrocyclic benzo groups are trans (55) and that in which<br />

they are cis (56). 60,103<br />

P<br />

P<br />

P<br />

P<br />

(55) (56)<br />

4.4 Mixed Donor Macrocycles<br />

4.4.1 Simple Mixed Donors<br />

Much of the work in this area has been reported by Lindoy<br />

and co-workers, who have performed extensive studies on the<br />

role of hole size in complex stability and rates of complex<br />

formation. 13,27,104 Bradshaw, Krakowiak, and Izatt have<br />

published an extensive text on the synthesis of aza crowns. 105<br />

A review of tri- and pentadentate macrocyclic ligands also<br />

includes mixed donor results as well as the influence of<br />

pendant arms. 16 Due to the numerous ramifications of this<br />

area, a few key findings will be cited for the simplest<br />

systems.<br />

A major focus in the study of mixed metal ion systems<br />

has been to examine metal ion discrimination. In particular,<br />

two specific mechanisms can be attributed to metal<br />

ion discrimination: macrocyclic hole size and what Lindoy<br />

has termed as a ‘dislocation’ mechanism. The key to this<br />

S<br />

S<br />

P<br />

P<br />

P<br />

P


mechanism is the assumption that coordination geometry<br />

preferences can be suddenly changed at some point along<br />

a series of ligands where gradual changes in the ligand<br />

framework are made. Because these changes can occur at<br />

different points for different metal ions, discrimination can<br />

be achieved. A particularly appealing aspect of the mixed<br />

donor aza–oxa systems is the lower ligand field which they<br />

provide, which then tends to minimize spin state changes.<br />

These systems are treated in a comprehensive review of<br />

O3N2, O2N3, and other pentadentate macrocycles with N, O,<br />

S heteroatoms. 104<br />

Examples of the use of synthetic mixed donor macrocycles<br />

in heavy metal ion separations are found in the discrimination<br />

of silver from lead. A number of studies indicate that<br />

the inclusion of sulfur in macrocyclic sequestering agents<br />

shifts the discrimination to silver. 104 An example of this<br />

is seen with (57) and (58). For the aza–oxa macrocycle<br />

(57) the log K is 5.9 for both silver and lead ions,<br />

while the thia-incorporated ligand (58) complexes silver<br />

more efficiently (log K = 9.9) compared to lead (log<br />

K = 5.7). 106,107<br />

NH<br />

O O<br />

HN<br />

NH<br />

S S<br />

O<br />

O<br />

(57) (58)<br />

HN<br />

The larger mixed aza–oxa, aza–thia, and aza–and<br />

oxa–phospha macrocycles are noted for their ability to<br />

complex more than one metal ion and to alter the magnetic<br />

properties of bimetallic complexes. 108–110 An example of<br />

tri-metal coordination is the tricopper complex of a 27member<br />

ring system (59). 108 A classic series of dicopper<br />

complexes which illustrates the influence of donor atoms<br />

on magnetism are the dicopper structures (60)–(62). 110 The<br />

magnetic properties were found to be extremely dependent<br />

on the mode of azide coordination, which is thought to be<br />

influenced by the orientation of the orbitals on the metal ions.<br />

In complex (60), the two copper ions are ferromagnetically<br />

coupled with a triplet ground state; in (61), the metal ions<br />

are antiferromagnetically coupled; and in (62), the two copper<br />

ions are not coupled.<br />

4.4.2 Cryptands<br />

Cryptands (15) are noted for their highly selective<br />

complexation of alkaline earth metal ions, and for their<br />

ring size–metal ion match ability. 5 The thermodynamic<br />

properties of these macrocycles have been extensively<br />

investigated, and results indicate that the high stability of the<br />

N<br />

H<br />

N<br />

H<br />

O<br />

Cu<br />

HOH O<br />

HN Cu Cu NH<br />

HN NH<br />

O<br />

(59)<br />

H<br />

N<br />

N3<br />

O<br />

N3<br />

H<br />

N<br />

HN Cu Cu<br />

N3<br />

N<br />

H<br />

O<br />

(61)<br />

N<br />

H<br />

N 3<br />

NH<br />

MACROCYCLIC LIGANDS 15<br />

HN<br />

N 3<br />

O<br />

Cu<br />

O<br />

O<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

O<br />

(60)<br />

S<br />

N3 N N N<br />

S<br />

O<br />

Cu<br />

O<br />

S<br />

S<br />

N 3<br />

NH<br />

HN Cu Cu NH<br />

N<br />

N N N3<br />

(62)<br />

bicyclic macrocycles compared to their monocyclic analogs is<br />

enthalpic in origin. 88<br />

4.4.3 Compartmental <strong>Ligands</strong><br />

Compartmental ligands (16) provide extensive opportunities<br />

for multiple metal ion complexation. An example of a<br />

mixed donor ligand incorporating different metal ions is the<br />

macrocyclic trinucleating ligand (63), which is capable of<br />

complexing two ‘soft’ donor metal centers in addition to a<br />

‘hard’ alkali or alkaline earth metal. 111<br />

4.5 Oxa Macrocycles<br />

4.5.1 Crown Ethers<br />

In the crown ethers (18) the interactions between the ligand<br />

and metal ion are considered to be more electrostatic in nature,<br />

rather than the covalent binding observed for the transition<br />

metal complexes of the aza, thia, and phospha macrocycles.<br />

The thermodynamic properties of these macrocycles have<br />

been extensively studied, with numerous reviews covering<br />

complexation, selectivity, and structural aspects, some with<br />

extensive tables of thermodynamic data. 69,70,112–119 Considerable<br />

efforts have been made to correlate the interrelationship<br />

between cavity size of the macrocycles and stability of alkali<br />

and alkaline earth metal complexes. From X-ray and CPK<br />

mo<strong>del</strong>s, cavity radii are determined as 0.86–0.92 ˚A for 15crown-5<br />

(64), 1.34–1.43 ˚A for 18-crown-6 (65), and about<br />

1.7 ˚A for 21-crown-7 (66). 69 For complex formation between<br />

the alkali metal ions and 18-crown-6, the maximum stability


16 MACROCYCLIC LIGANDS<br />

O<br />

S<br />

O<br />

N N<br />

Cu<br />

O O<br />

O<br />

N<br />

Ba<br />

Cu<br />

(63)<br />

occurs for the potassium ion, which has a radius of 1.38 ˚A, thus<br />

correlating well with the cavity radius. However, 18-crown-6<br />

forms extremely stable complexes with all of the alkali and<br />

alkaline earth metal ions. Hence, Gokel argues that the data<br />

indicate that the hole size concept is inapplicable, since the<br />

binding constants for sodium, potassium, ammonium, and<br />

calcium ions are the largest for the 18-crown-6 compared to<br />

almost all of the other simple crown ethers. 119 Hancock has<br />

proposed that chelate ring size is the critical factor, and that the<br />

high stabilities observed for the crown ethers with large metal<br />

ions is a result of the presence of five-membered chelate rings.<br />

Thus the high affinity of these macrocycles for the potassium<br />

ion is explained by the fact that potassium is the right size for<br />

the five-membered chelate rings of the crown ethers. 78<br />

O<br />

O<br />

O<br />

(64)<br />

O<br />

O<br />

O<br />

O<br />

O<br />

(65)<br />

O<br />

O<br />

O<br />

N<br />

O<br />

O<br />

S<br />

O<br />

O<br />

O<br />

O<br />

O<br />

(66)<br />

A number of reviews of the structural aspects of crown<br />

ethers can be found. 115–117 These structures vary considerably<br />

in complexity. An example of the flexibility of the crown ethers<br />

can be seen in the variation in the structures as a result of ring<br />

size of three different benzo crowns. When the cavity of the<br />

O<br />

O<br />

O<br />

crown matches the radius of the metal ion, the metal ion can be<br />

readily incorporated in the cavity, such as in the structure of the<br />

rubidium thiocyanate complex with the dibenzo-18-crown-6<br />

(67). In cases where the cavity of the crown is too large to<br />

surround the metal ion snugly, a folded structure can result,<br />

as with the dibenzo-30-crown-10 (68) and the potassium ion.<br />

For very large metal ions incapable of fitting into smaller<br />

macrocyclic cavities, sandwich-type structures can occur, as<br />

in the benzo-15-crown-5 (69) with the potassium ion. 115<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

(67)<br />

O<br />

O<br />

O O O<br />

O O O<br />

O<br />

(68)<br />

O<br />

O O<br />

(69)<br />

Molecular mechanics studies indicate that the lowest energy<br />

conformer of the uncomplexed ligand is not necessarily<br />

that required for complexation, i.e. oxygen donors may be<br />

exodentate as in the thia macrocycles. This means that in order<br />

for complex formation to occur, the ligand must undergo both<br />

reorganization as well as desolvation. A general rule of thumb<br />

with respect to size, however, is that the larger macrocycles<br />

are more flexible and subject to adaptability, while the smaller<br />

macrocycles are more rigid and, in that sense, ‘preorganized’.<br />

Cram has provided an excellent treatise on preorganization. 118<br />

His principle of preorganization is that ‘the more highly<br />

hosts and guests are organized for binding and low solvation<br />

prior to their complexation, the more stable will be their<br />

complexes.’ 118 �G values for a variety of macrocyclic oxygen<br />

donors indicate that the ‘prearranged’ ligands in general bind<br />

O<br />

O<br />

O


their guests more strongly and are, in sequence, the spherands<br />

> cryptaspherands ≈ cryptands > hemispherands > crown<br />

ethers. 116<br />

A useful correlation of enthalpy–entropy considerations for<br />

complexation has been shown by Inoue, Liu, and Hakushi. 113<br />

The treatment reflects enthalpy–entropy relationships for<br />

given types of ligands. The general concept is that as the<br />

enthalpic contributions become strong, a higher level of<br />

organization is obtained, which will result in unfavorable<br />

entropy changes. For a given type of system with similar<br />

entropic versus enthalpic considerations, the T�S and �H<br />

values determined for a series of ligands should thus exhibit<br />

a linear relationship. This is found for the macrocyclic crown<br />

ethers, the cryptands, lariat ethers, and bis-crown ethers, as<br />

well as the acyclic polyethers known as podands. The slopes<br />

are all positive with high correlation coefficients. Gokel has<br />

suggested that these slopes can be used to assess the ligand<br />

flexibility: glymes and podands (0.86) > crown ethers (0.76)<br />

> cryptands (0.51). 119<br />

4.5.2 Lariat Ethers<br />

The lariat ethers (19) and (20) known to date consist<br />

of macrocycles with many different types of podand<br />

groups, and much of their complexation chemistry involves<br />

electrostatic binding of guests. Reviews of both structural and<br />

thermodynamic aspects of the lariat ethers can be found. 120–122<br />

The trends are noted to be relatively similar for both the<br />

carbon-pivot and nitrogen-pivot types of lariat ethers. Binding<br />

strengths and selectivities are dependent on ring size and in<br />

general increase as ligand size increases. Strong selectivities<br />

are noted for the potassium ion, as in the crown ethers.<br />

4.5.3 Spherands and Hemispherands<br />

The spherands (21) were specifically designed using the<br />

concept of ‘preorganization’ wherein the oxygen donors are<br />

arranged in an enforced spherical cavity. Totally prearranged<br />

(spherand) and partially arranged (hemispherand, (22))<br />

complexes are possible. 118 Due to the structural restraints<br />

imposed by the rigidly joined phenyl rings, the spherands are<br />

considered to be highly ‘preorganized’ binding sites. In these<br />

macrocycles the lone pair of electrons will always be pointed<br />

toward the center of the macrocyclic cavity.<br />

4.5.4 Calixarenes<br />

The calixarenes (23) are also highly preorganized<br />

molecules which are capable of forming different<br />

conformational isomers. The conformational flexibility is<br />

determined by the size of the ring, with the preferred conformation<br />

becoming more planar as the ring size increases. 37<br />

5 APPLICATIONS<br />

MACROCYCLIC LIGANDS 17<br />

As macrocyclic chemistry has developed, the variety and<br />

scope of the applications of these molecules have continued<br />

to multiply. This concluding section is an attempt to provide<br />

an overview of only three of the applications of synthetic<br />

macrocycles. A particularly insightful treatment can be<br />

found in the Nobel Lecture of Jean-Marie Lehn, 123 which<br />

describes the concept of supramolecular chemistry from<br />

simple recognition, to cation and anion receptors, multiple<br />

recognition, catalysis, transport, and molecular devices.<br />

5.1 Ion Transport<br />

Ion transport, especially cation transport, was one of the<br />

early focal points in macrocyclic chemistry, revolving primarily<br />

around the crown ethers and cryptands. Later efforts have<br />

been to provide switches to control the rates of cation transport.<br />

Two examples of the types of switches that have been<br />

developed include photo switches using cryptands, 124 and<br />

electrochemical switches using anthraquinone-derived lariat<br />

ethers. 125<br />

Related to transport capabilities is the use of synthetic<br />

macrocycles in analytical chemistry. Because of their selective<br />

complexation of a variety of cations, the crown ethers and<br />

related macrocycles have been wi<strong>del</strong>y used for separations<br />

and analyses. 126<br />

While transport efforts have largely involved metal<br />

cations, more recent developments have led to the use of<br />

macrocycles for transport of more complex molecules such as<br />

nucleosides. 127<br />

5.2 Catalysis<br />

Catalysis can be broken down into a number of areas,<br />

depending on the substrate and the catalytic reaction. One of<br />

the prime areas of the initial effort in catalysis has been small<br />

molecule activation, such as oxygen with a number of transition<br />

metal ion macrocycles 128,129 and carbon dioxide, the latter<br />

particularly with cobalt(I) and nickel(I) macrocycles. 130,131<br />

Once the polyammonium macrocycles were found to be able<br />

to recognize substrates other than metal ions, other catalysis<br />

applications evolved. For example, phosphoryl transfer catalysis<br />

with simple polyammonium macrocycles has become<br />

quite accessible. 132<br />

5.3 Magnetic Resonance Imaging<br />

<strong>Macrocyclic</strong> complexes have gained recognition in<br />

magnetic resonance imaging. 133,134 In order to be effective<br />

imaging agents, complexes must provide a significant<br />

enhancement in the proton relaxation rates of water,<br />

as well as be nontoxic, and thermodynamically stable.<br />

Hence, macrocyclic ligands with pendant carboxylates, such


18 MACROCYCLIC LIGANDS<br />

as (5), have been examined primarily because of their<br />

thermodynamic stability.<br />

6 RELATED ARTICLES<br />

Ammonia & N-donor <strong>Ligands</strong>; Mixed Donor <strong>Ligands</strong><br />

P-donor <strong>Ligands</strong>; S-donor <strong>Ligands</strong>; Water & O-donor <strong>Ligands</strong>.<br />

7 REFERENCES<br />

1. J. Van Alphen, Recl. Trav. Chim. Pays-Bas, 1936, 55, 835.<br />

2. M. C. Thompson and D. H. Busch, Chem. Eng. News, 1962,<br />

57.<br />

3. N. F. Curtis, J. Chem. Soc., 1960, 4409.<br />

4. C. J. Pedersen, J. Am. Chem. Soc., 1967, 89, 7017.<br />

5. B. Dietrich, J.-M. Lehn, and J.-P. Sauvage, Tetrahedron Lett.,<br />

1969, 2889.<br />

6. J. Almy, D. C. Garwood, and D. J. Cram, J. Am. Chem. Soc.,<br />

1973, 95, 2961.<br />

7. T. Mashiko and D. Dolphin, in ‘Comprehensive Coordination<br />

Chemistry’, eds. G. Wilkinson, R. D. Gillard, and<br />

8.<br />

J. A. McCleverty, Pergamon, Oxford, 1982, Vol. 2, p. 813.<br />

V. T. Souza and M. L. Bender, Acc. Chem. Res., 1987, 20,<br />

146.<br />

9. G. A. Melson ed., ‘Coordination Chemistry of <strong>Macrocyclic</strong><br />

Compounds’, Plenum, New York, 1979.<br />

10. D. H. Busch, Acc. Chem. Res., 1978, 11, 393.<br />

11. T. A. Kaden, Top. Curr. Chem., 1984, 121, 157.<br />

12. E. Kimura, Top. Curr. Chem., 1985, 128, 113.<br />

13. L. F. Lindoy, ‘The Chemistry of <strong>Macrocyclic</strong> Ligand<br />

14.<br />

Complexes’, Cambridge University Press, Cambridge, MA,<br />

1989.<br />

P. Chaudhuri and K. Wieghardt, Prog. Inorg. Chem., 1987,<br />

35, 329.<br />

15. D. Parker, Chem. Soc. Rev., 1990, 19, 271.<br />

16. R. Bhula, P. Osvath, and D. C. Weatherburn, Coord. Chem.<br />

Rev., 1988, 91, 89.<br />

17. A. Bianchi, M. Micheloni, and P. Paoletti, Coord. Chem. Rev.,<br />

1991, 110, 17.<br />

18. A. Bianchi, M. Micheloni, and P. Paoletti, Pure Appl. Chem.,<br />

1988, 60, 525.<br />

19. D. H. Busch and C. Cairns, in ‘Progress in <strong>Macrocyclic</strong><br />

Chemistry’, eds. R. M. Izatt and J. J. Christensen, Wiley, New<br />

York, 1987, Vol. 3, Chap. 1, p. 1.<br />

20. A. M. Sargeson, Pure Appl. Chem., 1984, 56, 1603.<br />

21. J. L. Sessler, T. Murai, and V. Lynch, Inorg. Chem., 1989, 28,<br />

1333.<br />

22. F. V. Acholla, F. Takusagawa, and K. B. Mertes, J. Am. Chem.<br />

Soc., 1985, 107, 6902.<br />

23. I. Murase, K. Hamada, and S. Kida, Inorg. Chim. Acta, 1981,<br />

54, L171.<br />

24. N. B. Tucker and E. E. Reid, J. Am. Chem. Soc., 1933, 55,<br />

775.<br />

25. L. Horner, H. Kunz, and P. Walach, Phosphorus Relat. Group<br />

BElem., 1975, 6, 63.<br />

26. J. Ennen and T. Kauffmann, Angew. Chem., Int. Ed. Engl.,<br />

1981, 28, 118.<br />

27. L. Lindoy, in ‘Cation Binding by Macrocycles’, eds. Y. Inoue<br />

and G. W. Gokel, Dekker, New York, 1990, Chap. 16,<br />

p. 599.<br />

28. D. E. Fenton, U. Casellato, P. A. Vigato, and M. Vidali, Inorg.<br />

Chim. Acta, 1982, 62, 57.<br />

29. C. O. Dietrich-Buchecker, J.-M. Kern, and J.-P. Sauvage, J.<br />

Am. Chem. Soc., 1984, 106, 3043.<br />

30. Y. Inoue and G. W. Gokel eds, ‘Cation Binding by<br />

31.<br />

Macrocycles’, Dekker, New York, 1990.<br />

I. Tabushi, Y. Kobuki, and T. Nishiya, Tetrahedron Lett.,<br />

1979, 20, 3515.<br />

32. A. Shanzer, J. Libman, and F. Frolow, J. Am. Chem. Soc.,<br />

1981, 103, 7339.<br />

33. E. Schwartz and A. Shanzer, J. Chem. Soc., Chem. Commun.,<br />

1981, 634.<br />

34. D. J. Cram, S. Karbach, Y. H. Kim, L. Baczynskyj, and<br />

35.<br />

W. Kalleymeyn, J. Am. Chem. Soc., 1985, 107, 2575.<br />

G. W. Gokel, D. M. Dishong, and C. J. Diamond, J. Chem.<br />

Soc., Chem. Commun., 1980, 1053.<br />

36. D. J. Cram, Angew. Chem., Int. Ed. Engl., 1986, 25, 1039.<br />

37. C. D. Gutsche, in ‘Calixarenes’, ed. G. F. Stoddart, Royal<br />

Society Chemistry, Cambridge, MA, 1989.<br />

38. C. D. Gutsche, I. Alam, M. Iqbal, T. Mangiafico, K. C. Nam,<br />

J. Rogers, and K. A. See, J. Inclusion Phenom., 1989 7, 61.<br />

39. R. Bhula, P. Osvath, and D. C. Weatherburn, Coord. Chem.<br />

Rev., 1988, 91, 89.<br />

40. J. E. Richman and T. J. Atkins, J. Am. Chem. Soc., 1974, 96,<br />

2268.<br />

41. F. Chavez and A. D. Sherry, J. Org. Chem., 1989, 54, 2990.<br />

42. R. W. Alder, R. W. Mowlam, D. J. Vachon, and G. R.<br />

Weisman, J. Chem. Soc., Chem. Commun., 1992, 507.<br />

43. M. W. Hosseini, J. Comarmond, and J.-M. Lehn, Helv. Chim.<br />

Acta, 1989, 72, 1066.<br />

44. Y.-D. Choi and J. P. Street, J. Org. Chem., 1992, 57, 1258.<br />

45. L. Qian, Z. Sun, M. P. Mertes, and K. B. Mertes, J. Org.<br />

Chem., 1991, 56, 4904.<br />

46. E. K. Barefield, D. Chueng, D. vanDerveer, and F. Wagner, J.<br />

Chem. Soc., Chem. Commun., 1981, 302.<br />

47. M. C. Thompson and D. H. Busch, J. Am. Chem. Soc., 1964,<br />

86, 3651.


48. N. F. Curtis, in ‘Comprehensive Coordination Chemistry’,<br />

eds. G. Wilkinson, R. D. Gillard, and J. A. McCleverty,<br />

49.<br />

Pergamon, Oxford, 1982, Vol. 2, Chap. 21.1, p. 899.<br />

S. M. Nelson, Pure Appl. Chem., 1980, 52, 461.<br />

50. K. Krakowiak, J. S. Bradshaw, W. Jiang, N. K. Dalley, G. Wu,<br />

and R. M. Izatt, J. Org. Chem., 1991, 56, 2675.<br />

51. G. A. Melson and D. H. Busch, J. Am. Chem. Soc., 1964, 86,<br />

4834.<br />

52. A. Hohn, R. J. Geue, and A. M. Sargeson, J. Chem. Soc.,<br />

Chem. Commun., 1990, 1473.<br />

53. A. Bencini, A. Bianchi, A. Borselli, M. Ciampolini,<br />

P. Dapporto, E. Garcia-España, M. Micheloni, P. Paoli,<br />

54.<br />

J. A. Ramirez, and B. Valtancoli, J. Chem. Soc., Perkin Trans.<br />

2, 1989, 1131.<br />

M. P. Suh, S.-G. Kang, V. L. Goedken, and S.-J. Park, Inorg.<br />

Chem., 1991, 30, 365.<br />

55. A. J. Blake and M. Schröder, Adv. Inorg. Chem., 1990, 35,1.<br />

56. D. P. Riley and J. D. Oliver, Inorg. Chem., 1983, 22, 3361.<br />

57. R. E. Wolf, J. R. Hartman, J. M. E. Storey, B. M. Foxman,<br />

and<br />

4328.<br />

S. R. Cooper, J. Am. Chem. Soc., 1987, 109,<br />

58. P. Osvath, A. M. Sargeson, B. W. Skelton, and A. H. White,<br />

J. Chem. Soc., Chem. Commun., 1991, 1036.<br />

59. T. A. DelDonno and W. Rosen, J. Am. Chem. Soc., 1977, 99,<br />

8051.<br />

60. C. A. McAuliffe, in ‘Comprehensive Coordination Chemistry’,<br />

eds. G. Wilkinson, R. D. Gillard, and J. A. McCleverty,<br />

Pergamon, Oxford, 1982, Vol. 2, Chap. 14, p. 989.<br />

61. T. L. Jones, A. C. Willis, and S. B. Wild, Inorg. Chem., 1992,<br />

31, 1411.<br />

62. F. Gonce, A.-M. Caminade, F. Boutonnet, and J.-P. Majoral,<br />

J. Org. Chem., 1992, 57, 970.<br />

63. D. K. Cabbiness and D. W. Margerum, J. Am. Chem. Soc.,<br />

1969, 91, 6540.<br />

64. M. Kodama and E. J. Kimura, J. Chem. Soc., Dalton Trans.,<br />

1976, 2341.<br />

65. F. P. Hinz and D. W. Margerum, Inorg. Chem., 1974, 13,<br />

2941.<br />

66. A. Bianchi, L. Bologni, P. Dapporto, M. Micheloni, and<br />

P. Paoletti, Inorg. Chem., 1984, 23, 1201.<br />

67. M. Micheloni, P. Paoletti, and A. Sabatini, J. Chem. Soc.,<br />

Dalton Trans., 1985, 1169.<br />

68. D. H. Busch, K. Farmery, V. Goedken, V. Katovic, A. C.<br />

Melnyk, C. R. Sperati, and N. Tokel, Adv. Chem. Ser., 1971,<br />

100, 44.<br />

69. R. M. Izatt, J. S. Bradshaw, S. A. Nielsen, J. D. Lamb,<br />

J. J.<br />

271.<br />

Christensen, and D. Sen, Chem. Rev., 1985, 85,<br />

70. R. D. Hancock and A. E. Martell, Chem. Rev., 1989, 89,<br />

1875.<br />

71. M. Nonoyama and K. Nonoyama, Inorg. Chim. Acta, 1979,<br />

35, 231.<br />

MACROCYCLIC LIGANDS 19<br />

72. L. Fabrizzi and L. J. Zompa, Inorg. Nucl. Chem. Lett., 1977,<br />

13, 28.<br />

73. K. Wieghardt, W. Schmidt, R. van Eldik, B. Nuber, and<br />

J. Weiss, Inorg. Chem., 1980, 19, 2922.<br />

74. K. Wieghardt, W. Schmidt, H. Endres, and C. R. Wolfe,<br />

75.<br />

Chem. Ber., 1979, 112, 2837.<br />

K. Wieghardt, K. Pohl, I. Jibril, and G. Huttner, Angew.<br />

Chem., Int. Ed. Engl., 1984, 23, 77.<br />

76. L. T. Taylor and D. H. Busch, J. Am. Chem. Soc., 1967, 89,<br />

5372.<br />

77. R. I. Sheldon, A. J. Jircitano, M. A. Beno, J. M. Williams, and<br />

K. B. Mertes, J. Am. Chem. Soc., 1983, 105, 3028.<br />

78. R. D. Hancock, Acc. Chem. Res., 1990, 23, 257.<br />

79. L. Fabrizzi, M. Micheloni, and P. Paoletti, Inorg. Chem., 1982,<br />

19, 535.<br />

80. K. P. Wainwright and A. J. Ramasubbu, J. Chem. Soc., Chem.<br />

Commun., 1982, 277.<br />

81. T. J. Collins, C. Slebodnick, and E. S. Uffelman, Inorg.<br />

82.<br />

Chem., 1990, 29, 3433.<br />

M. Kodama and E. Kimura, J. Chem. Soc., Dalton Trans.,<br />

1978, 1081.<br />

83. L. Qian, Z. Sun, J. Gao, B. Movassagh, L. Morales, and<br />

K. B. Mertes, J. Coord. Chem., 1991, 23, 155.<br />

84. A. Bencini, A. Bianchi, E. Garcia-España, E. C. Scott,<br />

L. Morales, B. Wang, M. P. Mertes, and K. B. Mertes, Bioorg.<br />

Chem., 1992, 20,8.<br />

85. G. R. Newkome, J. D. Sauer, J. M. Roper, and D. C. Hager,<br />

Chem. Rev., 1977, 77, 513.<br />

86. G. L. Rothermel, L. Miao, A. L. Hill, and S. C. Jackels, Inorg.<br />

Chem., 1992, 31, 4854.<br />

87. E. Simmons and C. H. Park, J. Am. Chem. Soc., 1968, 90,<br />

2428.<br />

88. K. B. Mertes and J.-M. Lehn, in ‘Comprehensive<br />

Coordination Chemistry’, eds. G. Wilkinson, R. D. Gillard,<br />

and J. A. McCleverty, Pergamon, Oxford, 1982, Vol. 2,<br />

Chap. 21.3, p. 915.<br />

89. A. Bencini, A. Bianchi, P. Dapporto, E. Garcia-España,<br />

90.<br />

M. Micheloni, J. A. Ramirez, P. Paoletti, and P. Paoli, Inorg.<br />

Chem., 1992, 31, 1902.<br />

A. Bencini, A. Bianchi, M. Micheloni, P. Paoletti, P. Dapporto,<br />

P. Paoli, and E. Garcia-España, J. Inclusion Phenom., 1992,<br />

12, 291.<br />

91. D. H. Busch, Acc. Chem. Res., 1978, 11, 392.<br />

92. J. L. Sessler and A. K. Burrell, Top. Curr. Chem., 1991, 161,<br />

177.<br />

93. S. R. Cooper and S. C. Rawle, Struct. Bonding, 1990, 72,1.<br />

94. S. R. Cooper, Acc. Chem. Res., 1988, 21, 141.<br />

95. M. Schröder, Pure Appl. Chem., 1988, 60, 517.<br />

96. S. Lucia, W. L. Sokol, L. A. Ochrymowycz, and D. B.<br />

Rorabacher, Inorg. Chem., 1981, 20, 3189.


20 MACROCYCLIC LIGANDS<br />

97. N. W. Alcock, H. Heron, and P. Moore, J. Chem. Soc., Chem.<br />

Commun., 1976, 886.<br />

98. V. B. Pett, L. L. Diaddario, E. R. Dockal, P. W. Corfield,<br />

C. Ceccarelli, and M. D. Glick, Inorg. Chem., 1983, 22,<br />

661.<br />

99. M. M. Bernardo, M. J. Heeg, R. R. Schroeder, L. A.<br />

100.<br />

Ochrymowycz, and D. B. Rorabacher, Inorg. Chem., 1992,<br />

31, 191.<br />

B. de Groot, H. A. Jenkins, and S. J. Loeb, Inorg. Chem.,<br />

1992, 31, 203.<br />

101. C. M. Lucas, L. Shuang, M. J. Newlands, J.-P. Charland, and<br />

E. J. Gabe, Can. J. Chem., 1989, 66, 639.<br />

102. K. Mislow and R. Baechler, J. Am. Chem. Soc., 1970, 92,<br />

3090.<br />

103. E. P. Kyba, R. E. Davis, C. W. Hudson, A. M. John, S. B.<br />

Brown, M. J. McPhaul, L. K. Liu, and A. C. Glover, J. Am.<br />

Chem. Soc., 1981, 103, 3868.<br />

104. L. F. Lindoy, in ‘Progress in <strong>Macrocyclic</strong> Chemistry’, eds.<br />

R. M. Izatt and J. J. Christensen, Wiley, New York, 1987,<br />

Vol. 3, Chap. 2, p. 53.<br />

105. J. S. Bradshaw, K. Krakowiak, and R. M. Izatt, ‘Aza Crown<br />

Macrocycles’, Wiley, New York, 1993.<br />

106. F. Arnaud-Neu, M. J. Schwing-Weill, R. Louis, and R. Weiss,<br />

Inorg. Chem., 1979, 18, 2956.<br />

107. F. Arnaud-Neu, B. Spiess, and M. J. Schwing-Weill, Helv.<br />

Chim. Acta, 1977, 60, 2633.<br />

108. J. Comarmond, B. Dietrich, J.-M. Lehn, and D. Parker, J.<br />

Chem. Soc., Chem. Commun., 1985, 75.<br />

109. A. Bencini, A. Bianchi, E. Garcia-España, M. Micheloni, and<br />

P. Paoletti, Inorg. Chem., 1987, 26, 1243.<br />

110. J. Comarmond, P. Plumeré, J.-M. Lehn, Y. Agnus, R. Louis,<br />

R. Weiss, O. Kahn, and I. Morgenstern-Badaru, J. Am. Chem.<br />

Soc., 1982, 104, 6330.<br />

111. F. C. J. M. van Veggel, M. Bos, S. Harkema, H. van de<br />

112.<br />

Bovenkamp, W. Verboom, J. Reedijk, and D. N. Reinhoudt,<br />

J. Org. Chem., 1991, 56, 225.<br />

A. I. Popov and J.-M. Lehn, in ‘Coordination Chemistry of<br />

<strong>Macrocyclic</strong> Compounds’, ed. G. A. Melson, Plenum, New<br />

York, 1979.<br />

113. Y. Inoue and T. Hakushi, in ‘Cation Binding by Macrocycles’,<br />

eds. Y. Inoue and G. W. Gokel, Dekker, New York, 1990,<br />

Chap.1,p.1.<br />

114. R. L. Brueining, R. M. Izatt, and J. S. Bradshaw, in ‘Cation<br />

Binding by Macrocycles’, eds. Y. Inoue and G. W. Gokel,<br />

Dekker, New York, 1990, Chap. 2, p. 111.<br />

115. M. R. Truter, Struct. Bonding, 1973, 16, 71.<br />

116. I. Goldberg, in ‘The Chemistry of Functional Groups’, ed.<br />

S. Patai, Wiley, New York, 1980, p. 175.<br />

117. R. Hilgenfeld and W. Saenger, Top. Curr. Chem., 1982, 101,<br />

1.<br />

118. D. J. Cram, Science, 1988, 242, 760.<br />

119. G. W. Gokel and J. E. Trafton, in ‘Cation Binding by<br />

120.<br />

Macrocycles’, eds. Y. Inoue and G. W. Gokel, Dekker, New<br />

York, 1990, Chap. 6, p. 253.<br />

F. R. Fronczek and R. D. Gandour, in ‘Cation Binding by<br />

Macrocycles’, eds. Y. Inoue and G. W. Gokel, Dekker, New<br />

York, 1990, Chap. 7, p. 311.<br />

121. J. Tsukube, J. Coord. Chem., 1987, 16, 101.<br />

122. K. E. Krakowiak, J. S. Bradshaw, and D.-J. Zamecka-<br />

123.<br />

Krakowiak, Chem. Rev., 1989, 89, 929.<br />

J.-M. Lehn, Angew. Chem., Int. Ed. Engl., 1988, 27, 89.<br />

124. S. Shinkai and O. Manabe, ‘Host Guest Complex<br />

p. 67.<br />

Chemistry III’, Springer-Verlag, Berlin, 1984,<br />

125. L. E. Echegoyen, H. K. Yoo, V. J. Gatto, G. W. Gokel,<br />

and<br />

2440.<br />

L. Echegoyen, J. Am. Chem. Soc., 1989, 111,<br />

126. K. Kimura and T. Shono, in ‘Cation Binding by Macrocycles’,<br />

eds. Y. Inoue and G. W. Gokel, Dekker, New York, 1990,<br />

Chap. 10, p. 429.<br />

127. H. Furuta, K. Furuta, and J. L. Sessler, J. Am. Chem. Soc.,<br />

1991, 113, 4706.<br />

128. C. J. Burrows, in ‘Inclusion Phenomena and Molecular<br />

129.<br />

Recognition’, ed. J. L. Atwood, Plenum, New York, 1990,<br />

p. 199.<br />

L. D. Margerum, K. I. Liao, and J. S. Valentine, in ‘Metal<br />

Clusters in Proteins’, ed. L. Que Jr, American Chemical<br />

Society, Washington, DC, 1988, Chap. 6, p. 105.<br />

130. M. H. Schmidt, G. M. Miskelly, and N. S. Lewis, J. Am.<br />

Chem. Soc., 1990, 112, 3420.<br />

131. E. Fujita, C. Creutz, N. Sutin, and D. J. Szalda, J. Am. Chem.<br />

Soc., 1991, 113, 343.<br />

132. M. P. Mertes and K. B. Mertes, Acc. Chem. Res., 1990, 23,<br />

413.<br />

133. D. Parker, Chem. Soc. Rev., 1990, 19, 271.<br />

134. J. F. Carvalho, S.-H. Kim, and C. A. Chang, Inorg. Chem.,<br />

1992, 31, 4065.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!