guest editors philip gale and thorfinnur gunnlaugsson · anion coordination chemistry has attracted...

19
This article was published as part of the Supramolecular chemistry of anionic species themed issue Guest editors Philip Gale and Thorfinnur Gunnlaugsson Please take a look at the issue 10 2010 table of contents to access other reviews in this themed issue Downloaded by Xiamen University on 29 January 2011 Published on 25 August 2010 on http://pubs.rsc.org | doi:10.1039/B926160P View Online

Upload: others

Post on 15-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This article was published as part of the

Supramolecular chemistry of anionic species themed issue

Guest editors Philip Gale and Thorfinnur Gunnlaugsson

Please take a look at the issue 10 2010 table of contents to access other reviews in this themed issue

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

PView Online

Page 2: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

ISSN 0306-0012

0306-0012(2010)39:10;1-T

www.rsc.org/chemsocrev Volume 39 | Number 10 | October 2010 | Pages 3581–4008

Themed issue: Supramolecular chemistry of anionic species

Chemical Society Reviews

Guest editors: Philip Gale and Thorfi nnur Gunnlaugsson

CRITICAL REVIEWYun-Bao Jiang et al.Anion complexation and sensing using modifi ed urea and thiourea-based receptors

TUTORIAL REVIEWEric V. Anslyn et al.The uses of supramolecular chemistry in synthetic methodology development

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 3: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3729

Anion complexation and sensing using modified urea and thiourea-based

receptorsw

Ai-Fang Li, Jin-He Wang, Fang Wang and Yun-Bao Jiang*

Received 17th March 2010

DOI: 10.1039/b926160p

This critical review highlights recent advances in the structurally modified (thio)urea-based

receptors for anion complexation and sensing. Modifications of the (thio)urea structure are aimed

at a better anion binding in terms of higher binding constant, anion selectivity and feasibility.

Major (thio)urea receptors are reviewed as N-alkyl, N-aryl and N-amido/N-amino (thio)ureas.

Hints for designing (thio)urea-based receptors for anions are discussed (102 references).

Introduction

Anion coordination chemistry has attracted growing attention

in supramolecular chemistry due to the essential roles that

anions play in biology, medicine, catalysis, and environmental

science.1 Much effort has been expended on the design of

receptors for anion complexing and sensing. This has been a

great challenge because anion complexation with the receptor

is quite different from that for metal cations, which operates

through metal coordination, mainly because of the peculiar

characters of anions with varying size, shape and charge, and

pH-dependent species.1 Yet substantial progress has been

made in the development of anion receptors. Several excellent

reviews can be referred to for the recent advances in this

regard.2–9 Basically, the anion receptors can be neutral or

positively charged and the anion–receptor interactions are, in

general, of hydrogen bonding and/or electrostatic interaction,

although deprotonation of the acidic hydrogen bonding donor

occurs in some cases. In neutral anion receptors those bearing

amide, (thio)urea and pyrrole groups that bind anions via

favourable hydrogen bonding have been effective.2,5–9

Since the pioneering work of Wilcox10 and Hamilton11 that

showed urea moieties can act as appropriate binding sites for

anions, in particular oxoanions such as CH3CO2�, a variety of

receptors containing a (thio)urea subunit have been developed

and applied for anion complexation and sensing over the past

years.2,5–9 Therefore, (thio)ureas have been taken as good

hydrogen bond donors for the construction of anion receptors

that may bind anions by two hydrogen bonds of the (thio)-

ureido –NHs.

In this review, we will focus on the N,N0-substitution

modified (thio)urea-based receptors for anion complexation

and sensing. They will be classified according to the structural

motif of one substitution. We aim to provide a general view of

the design of anion receptors containing the (thio)urea

subunit. In the conclusion, we will discuss the future develop-

ments in the structural modifications of (thio)ureas as anion

receptors and their applications in organocatalysis, crystal

engineering, supramolecular gelation and membrane trans-

porters for anions.

Modified (thio)urea-based receptors

(Thio)urea has been the subject of intensive investigations for

its performance in the construction of tailored anion receptors

Department of Chemistry, College of Chemistry and ChemicalEngineering, and the MOE Key Laboratory of Analytical Sciences,Xiamen University, Xiamen 361005, China.E-mail: [email protected]; Fax: +86 592 218 8372;Tel: +86 592 218 8372w Part of a themed issue on the supramolecular chemistry of anionicspecies.

Ai-Fang Li

Ai-Fang Li was born in Fujian,China in 1981. She studiedchemistry at Xiamen Univer-sity and received her PhD in2009 under the supervision ofProf. Yun-Bao Jiang. Shespent 3 months in 2008 as anexchanged student in thegroup of Prof. Andy Hor TziSum at National Universityof Singapore. Her researchinterests are the developmentof molecular sensors for anionicand cationic species and theconstruction of molecularlogic gates and switches.

Jin-He Wang

Jin-He Wang was born inFujian, China in 1984. Hereceived a BS degree in chemis-try from Fuzhou University in2007. Currently he is workingwith Professor Yun-Bao Jiangat Xiamen University for hisPhD degree. In his PhD thesisresearch, he is working on(thio)urea-based anion recogni-tion and anion-responsivesupramolecular organogels, andanion-directed hydrogen-bondingsupramolecular polymers.

CRITICAL REVIEW www.rsc.org/csr | Chemical Society Reviews

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 4: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

3730 Chem. Soc. Rev., 2010, 39, 3729–3745 This journal is c The Royal Society of Chemistry 2010

via double hydrogen bonding interactions. The geometrical

arrangement of the acetate complex of a (thio)urea-based

receptor is illustrated in Scheme 1. It is known that hydrogen

bonding is directional, thereby allowing the design of (thio)-

urea-based receptors for given anions with both high affinity

and selectivity. Taking into account the complicated nature

of anions, such as a wide range of geometries, high Lewis

basicity and high solvation energies in a given solvent, the key

factors that govern the design of (thio)urea-based anion

receptors, for a strong binding affinity and a high selectivity,

are the strength of hydrogen bonding, cis- or trans-conforma-

tion of the two (thio)ureido N–H bonds versus the CQX

(X= S or O) bond, and the topology of the (thio)urea binding

sites in the receptor molecule. Structural modifications on

the (thio)ureas have thus been attempted (i) to increase

the acidity of the (thio)ureido –NH protons by introducing

electron-withdrawing substituent(s), e.g., –NO2, –CF3, via

the mediation of a p-system such as a phenyl ring, (ii) to

include other interactions such as additional hydrogen

bonding, through introducing –NH, –OH, pyrrole or indole

groups, and electrostatic interaction, and/or (iii) to form a

preorganized binding cleft/cavity using a rigid (calixarene,

cholapod or polynorbornane) or flexible structural skeleton

that leads to a binding environment with size/shape com-

plementary for a given anion.

Many N,N0-substitution modified (thio)urea-based receptors

have accordingly been designed so that the binding to the

anion can be probed through a macroscopic optical or electro-

chemical response. 1H NMR, UV-Vis absorption, fluorescence

and CD are among those optical methods. Detailed informa-

tion on the nature of the anion–receptor interactions can be

provided by 1H NMR titrations and in fortunate cases by

X-ray crystal structure analysis.

In general, the acidity of thioureido –NH protons is higher

than that of the ureido –NHs, with pKa of 21.1 and 26.9 for

thiourea and urea in DMSO, respectively.12 This leads to a

stronger hydrogen bonding of an anion with thiourea than

with the urea counterpart, an observation that is quite often

taken to identify the (thio)urea anion binding site in the

(thio)urea-based receptors. However, if the –NH protons are

acidic enough, in particular when an electron-withdrawing

substituent is introduced into the receptor molecule, deproto-

nation may occur in the presence of a highly basic anion,

normally F�. On the occurrence of –NH deprotonation,

dramatic spectral changes with a concomitant solution color

change may be observed. The deprotonation of –NHs by F� in

anion sensing chemistry was first observed by the Gunnlaugsson

group13,14 and the Gale group.15 Later, Fabbrizzi et al. carried

out a detailed study of the deprotonation of (thio)ureas in the

presence anions.16 Two step-wise processes are assumed in the

deprotonation process, equilibria (1) and (2). First, a genuine

receptor–anion hydrogen bonding complex, [LH� � �X]� is

formed. Then, on further addition of the anion, one molecule

of HX is released from the [LH� � �X] complex, leading to a

self-complex [HX2]� and deprotonated receptor L�. Deproto-

nation normally occurs at the (thio)ureido –NH proton close

to the more electron-withdrawing subunit. It appears that

only one of the two (thio)ureido –NHs is deprotonated, if

deprotonation occurs. The deprotonation process can be

probed by 1H NMR titrations and, luckily, by X-ray crystallo-

graphy. As a distinct solution color change may be observed

when deprotonation occurs, a colorimetric method is

easily made available for anion sensing, even by naked eyes.Scheme 1 Complementary hydrogen bonding interactions between

(thio)urea and an acetate anion.

Fang Wang

Fang Wang was born inShandong, China in 1985 andis currently a master student inthe group of Prof. Yun-BaoJiang. She received her bachelordegree in chemistry in 2008from Shandong Normal Univer-sity and joined the group ofProf. Jiang in September2008. Her MS thesis researchis on chiral thiourea-basedreceptors for anion recognitionand anion-mediated chiralitytransferring and amplification.

Yun-Bao Jiang

Yun-Bao Jiang, currently aprofessor at the Departmentof Chemistry, Xiamen Univer-sity, studied chemistry in thesame University for 10 yearsand obtained his PhD in 1990under the supervision of lateProf. Guo-Zhen Chen. He hasbeen a Humboldt fellow withProf. K. A. Zachariasse atthe Max-Planck-Institut furbiophysikalische Chemie inGottingen and a universitypostdoctoral fellow with Prof.Chi-Ming Che at the Univer-sity of Hong Kong. He has

since 2010 been on the editorial board of ‘‘Photochemical &Photobiological Sciences’’. Research in his group focuses onproton and electron transfer photophysics, supramolecularrecognition and sensing, and biomacromolecular interactionsmonitored by FCS.

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 5: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3731

Deprotonation will be pointed out in the examples given in the

latter sections.

LH + X� " [LH� � �X]� (1)

[LH� � �X]� + X� " L� + [HX2]� (2)

A systematic study by Wilcox’s group17 of N-octyl-N0-

arylthioureas 1 on their binding to 4-tributylammonium-1-

butanesulfonate in CHCl3 provided an early example of the

dependence of the anion hydrogen bonding interaction on the

acidity of thioureido –NH protons. 1H NMR investigations

indicated that the introduction of an electron-withdrawing

substituent (X) in the N-phenyl ring led to a higher acidity of

thioureido –NH protons and consequently stronger binding to

anions.

The report of Martınez-Manez et al.18 on (thio)urea deri-

vatives 2a and 2b allowed a comparison of the anion binding

of thiourea and urea receptors. With anions AcO�, BzO�,

H2PO4� and Cl�, the formation of hydrogen bonding

complexes was clearly indicated by absorption spectral and

NMR studies. The binding constants of 2a with AcO�, BzO�,

H2PO4� and Cl� decrease in the order of the anion basicity

and are lower than those of 2b, which is in agreement with

the hydrogen bonding nature of the anion interaction with

the (thio)urea moiety and the lower acidity of the ureido –NH

than that of the thioureido –NH. Deprotonation was observed

with both 2a and 2b in DMSO in the presence of highly basic

F�, leading to a neat color change from yellow to blue.

In this review, we will highlight advances of modified

(thio)urea-based receptors in terms of N-alkyl, N-aryl and

N-amido/N-amino (thio)ureas. This classification is based on

the modification made on the N-alkyl, aryl or amido/amino

moiety. Under this condition it is implied that theN0-substitution

is in general an aryl group such as phenyl. Note that with these

(thio)urea derivatives, the (thio)urea moiety is linked to

the N-substituent via a ‘‘C’’ or ‘‘N’’ atom. In the case of the

‘‘C’’-linkage, N-acyl(thio)ureas are also a big group of the

derivatives. However, as there is a strong six-membered ring

intramolecular hydrogen bond (IHB) between the N-acyl

CQO oxygen and the other (thio)ureido –NH proton, their

hydrogen bonding with a foreign anion is prevented to a high

extent. N-Acyl(thio)ureas are therefore not included.

N-Alkyl(thio)ureas

An alkyl group is not efficient in delivering the electronic effect

for tuning the acidity of the attached (thio)ureido –NH

proton, yet its hydrophobicity allows for positioning the

resulting N-alkyl(thio)urea receptor in a microenvironment,

while its structural flexibility may lead to a conformation of

the two –NH hydrogen binding donors favorable for anion

binding or may afford a good anion binding environment in

terms of size and/or shape matching. The latter is especially

important for a higher binding selectivity for spheric halide

anions and tetrahedral phosphate anions.

Teramae et al.19 have examined anion recognition of

N-alkyl-N0-(p-nitrophenyl)thioureas 3 in cationic vesicle solutions.

Varying the length of alkyl chain has successfully positioned

the anion binding site in the vesicle. The anion-binding site in

receptor with n = 1 is buried deeper within the nonpolar

vesicle core and shows a high selectivity towards Br� anion

(Br� > H2PO4� > Cl� c HCO3

�, CH3CO2�). No distinct

anion selectivity was observed when n = 8. This was the first

report of a position-dependent anion recognition via hydrogen

bonding interaction performed in aqueous vesicle solutions by

thiourea-based receptors.

Coordination of a metal cation has been an important

means of enhancing the hydrogen bond donor tendencies of

the N-alkyl(thio)ureas, thanks to the flexibility of the alkyl

chain. Fabbrizzi and co-workers reported a heteroditopic

p-nitrophenylthiourea 4 that contains a 15-membered NO2S2crown with a flexible –CH2CH2– spacer.20 The thiourea S

atom becomes more electron withdrawing when it is coordinated

with a Ag+ cation (Scheme 2). As a consequence, the acidity

of the thioureido –NH protons is enhanced and unusually

stable hydrogen bonding complexes with Cl� and Br� are

formed. The Cl�/Br� binding constants are increased by at

least 3 orders of magnitude compared to those in the absence

of Ag+. The corresponding N-methyl derivative (3, n = 1)

showed practically no binding with these two anions in the

same solvent, CH3CN.19 The urea counterpart of 4, however,

Scheme 2

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 6: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

3732 Chem. Soc. Rev., 2010, 39, 3729–3745 This journal is c The Royal Society of Chemistry 2010

did not show a similar binding enhancement towards Cl� and

Br� in the presence of Ag+, as AgCl and AgBr precipitates

were instead formed.20 This was further evidence to support

the contribution of the CQS� � �Ag+ coordination in promot-

ing the binding of 4 with Cl� and Br�.

The introduction of an additional anion binding site was also

employed to create (thio)urea-based receptors for better perfor-

mance. Receptors 5a–c containing both thiourea and amide

binding sites were reported by He et al.21 All of them showed a

high selectivity for adipate against dicarboxylate anion, through

the formation of a 1 : 1 hydrogen bonding complex. More

binding sites, together with the relatively flexible structure of

the receptor, allows for an optimal multiple hydrogen bonding

environment of the receptor for adipate anion but not others.

By introducing indole –NH groups, suitably functionalized

indole-based (thio)urea receptors 6a–c were designed by Pfef-

fer et al.22 The interactions of 6a–c with a series of anions (F�,

Br�, Cl�, AcO� and H2PO4�) monitored by 1H NMR titra-

tions indicated that all four –NH protons in 6a–c were

cooperating in their binding to H2PO4�.

The conformation of the two N–H bonds related to the CQX

bond (X = S or O) in the substituted (thio)ureas plays an

important role in anion binding. Qian and Liu23 reported a series

of naphthylthioureas 7 containing a varying number of –CH2OH

groups in the N-alkyl moiety. The fluorescent sensing ability

of 7 for anions was shown to depend on the number of –CH2OH

groups, since the trans–trans or trans–cis conformation of the

thioureido N–H bonds is influenced by it. 7a and 7b, whose N–H

bonds take the trans–trans conformation, form hydrogen bonding

complexes with anions, whereas 7c and 7d in a trans–cis con-

formation are not good at hydrogen bonding with the anions. The

–CH2OH group was suggested to preorganize the N–H bonds so

that the anion–receptor hydrogen bonding interaction is facili-

tated. An early report of sugar thioureas24 suggested that this

observation may relate to the intramolecular hydrogen bonding of

the –CH2OH group with the thiourea moiety.

The flexibility of the alkyl chain has also allowed the

construction of a variety of fluorescent (thio)urea receptors.

In these receptors the (thio)urea moiety, as an anion binding

site, is linked to a fluorophore via an alkyl spacer, for instance

in many cases –CH2–, that allows a through-space electron

transfer in the excited state, the so-called photo-induced

electron transfer (PET) mechanism. Gunnlaugsson’s group

has been among the first to develop such PET fluorescent

(thio)urea-based receptors.25,26 Receptors 8 and 9 are two

examples of this kind, in which the anthracene fluorophore

is linked to the thiourea moiety via a –CH2– spacer. The

fluorescence emission is significantly quenched upon anion

binding to the thiourea unit, since an active PET channel is

operating while the thiourea moiety is within the electron

donor. Mono-thiourea 8 was designed to bind and sense

simple anions F�, CH3CO2� and H2PO4

�.25 The anion

selectivity and the degree of the fluorescence quenching are

in the order of anion basicity, F�>CH3CO2�>H2PO4

�. By

introducing an additional (thio)urea unit at the anthracene

core, the extended design leads to PET fluorescent receptor 9

that shows an enhanced selectivity towards biologically important

dicarboxylates such as malonate, glutarate and pyrophosphate.26

An early example reported by Teramae et al.27 that is struc-

turally similar to 8 but bearing a pyrene fluorophore, shows

the occurrence of electron transfer upon anion binding in

acetone by a red-shifted exciplex fluorescence.

The difference in the anion selectivity of 8 and 9 suggests

another feature of the structural flexibility of the N-alkyl

group in tuning the conformation of the receptor bearing

multiple N-alkyl(thio)urea moieties. Conformational restriction

can thus be a design guideline for the receptors.

With a wide range of geometries, thiourea 10 is such a

receptor constructed by Kondo and Sato28 on the 2,20-bi-

naphthalene skeleton (Scheme 3). Remarkable absorption and

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 7: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3733

fluorescence spectral changes occurred upon the addition of

anions in CH3CN. Characteristic changes in the absorp-

tion spectra probed the conformational restriction of the

2,20-binaphthyl skeleton upon cooperative complexation of

the two thiourea groups with AcO� and F�. A dual-fluorescence

ratiometric assay of F� was also possible.

Bisthiourea receptor 11 by Hong et al.29 bearing an azophenol

chromophore showed a sensitive colorimetric response upon

hydrogen bonding with H2PO4�, F� and CH3CO2

� by red-

shifted absorption in chloroform. In the case of RQNO2, it

also allowed the differentiation of H2PO4� from F� and

CH3CO2� by a further spectral red-shift, despite the similar

basicity of these anions, whereas receptors with a single thiourea

moiety showed similar spectral profiles in the presence of all of

them. This observation is again an indication of the conforma-

tional fitness of the receptors bearing multiple N-alkylthiourea

moieties for given anions.

In the case of several N-alkyl(thio)ureas appended to a rigid

skeleton, the anion binding sites can be preorganized to

provide a favorable cavity for anions that are otherwise

weakly bound by the individual N-alkyl(thio)urea receptor,

such as weakly basic halide anions Cl� and Br� and tetra-

hedral phosphate. Steroidal receptors 12 and 13 with a cholapod

architecture, reported by Davis et al.,30 contain four (thio)urea

–NH groups that are well positioned for anion binding

because of the restricted rotation around the C–N bonds.

These receptors in CHCl3 very effectively bind Cl� and Br�,

that are known to be weakly bound by normal

N-alkyl(thio)ureas such as 3 (n = 1) in the same solvent.

From 12a to 12d, the anion affinity in the order of 107–109 M�1

increases along with the acidity of the (thio)ureido –NH

protons. 13 bearing an additional nitrosulfonamide binding

site shows a much higher anion affinity (109–1011 M�1).

A series of di-, tri- and tetrathiourea functionalized [3]-

and [5]polynorbornane scaffolds were created by Pfeffer

and Lowe31,32 that nicely showed preorganized effects for

anion binding. It was found that the binding affinity for

terephthalate2� was influenced by the size of the preorganized

cleft/cavity, which was imparted by the [n]polynorbornane

scaffold. For example, the [5]polynorbornane scaffold 15a–b

that provides a larger cleft/cavity exhibited increased affinities

for terephthalate2� through a higher degree of host:guest size

complementarity than the [3]polynorbornane-based receptor

14a–b. 14b and 15b, bearing the more electron-withdrawing

p-NO2 substituent, showed a slightly higher affinity due to the

enhanced hydrogen bonding ability of the thiourea –NH

protons. Meanwhile, 14b–c showed high affinity for H2PO4�

and H2P2O72� due to the formation of 1 : 1 stoichiometric

complexes and 2 : 1 self-assembled structures, respectively.33,34

Macrocyclic anion receptors with multidentate (thio)urea

units provide another kind of preorganization of the anion

binding sites into a ‘‘cavity’’ shape. Macrocyclic urea-based

receptors 16a and 16b form 1 : 1 complexes with H2PO4� via

hydrogen bonding interactions.35 16b, of higher struc-

tural flexibility, exhibited a higher selectivity for H2PO4�

Scheme 3

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 8: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

3734 Chem. Soc. Rev., 2010, 39, 3729–3745 This journal is c The Royal Society of Chemistry 2010

(binding constant of 4.0 � 103 M�1) over Cl� (100-fold)

compared to that of 16a (5-fold) in polar organic solvent,

DMSO. The cleft-like bisurea receptors without the R bridge

in 16 showed differing binding stoichiometries for H2PO4�

and Cl�. It was therefore made clear that the size of the

preorganized ‘‘cavity’’ in 16 dictated the preference for the

bound anions.

Metal coordination to the ligand-functionalized (thio)urea

receptors can alternatively preorganize the anion binding sites

for higher selectivity towards given anions. Gale and Loeb

et al.36 reported a urea receptor 17 bearing an isoquinoline

ligand (Scheme 4). Four receptor molecules were assembled

via coordination with Pt2+ so that suitable hydrogen bonding

donor pockets are formed. 17 in the presence of Pt2+ in a

highly competitive solvent, DMSO, exhibited strong binding

to SO42� and H2PO4

� with binding constants over 105 M�1.

Crystal structure analysis indicates that the four urea units in

the Pt(17)42+ complex exist in a cone conformation so that

eight NH groups are oriented towards a single anion. With

spherical halide anions Cl�, Br� and I�, Pt(17)42+ binds

in a 1 : 2 stoichiometry in its 1,2-alternate conformation, with

the first-step binding constants (103–104 M�1) being quite

high too.

Isothiouroniums, derivatives of mainly N-alkylthioureas, as

anion receptors may deserve discussion here too, although not

much attention has been paid to them. Hong and Yeo37,38

reported anion binding of isothiouroniums in 1998; however,

only after the report of Kubo et al.39 on 18 did several later

investigations appeared. Isothiouroniums in general showed

stronger anion binding ability because both hydrogen bonding

and electrostatic interaction with anions existed. For example,

the binding constant of 18 with CH3CO2� in CH3CN of

>106 M�1 is much higher than that of the neutral receptor.

Using the same isothiouronium moiety in 18, Kubo’s group40

developed bis(isothiouronium) receptor 19 based on the 1,10-

binaphthalene fluorescent framework that exhibited higher

affinity for oxoanions with more substantial fluorescence

enhancement than 18.40 They also showed that the ‘‘concave’’-

shaped bis(isothiouronium) fluorescent receptor 20 was able to

bind anions in a highly competitive solvent, 6% H2O–MeCN,

but in a selectivity order of HPO42� > AcO� c H2PO4

�.41

Teramae et al.42 have examined the anion binding in MeOH

of a variety of thiouroniums with varying S-substituents

(21 and 22). Those isothiouroniums with a CH2 spacer

(n = 0) underwent decomposition in protic and nucleophilic

solvents such as ROH upon photoexcitation. It was shown

that the CH2 spacer is not suitable for the isothiouronium-

based PET fluorescent receptors in which the fluorophore is

linked via this spacer to the sulfur atom of the isothiouronium

moiety. Indeed, 21b with an ethylene (CH2CH2) spacer formed

stable 1 : 1 complexes with anions in the selectivity order of

HPO42� (1.1 � 104 M�1) > CH3CO2

� (1.7 � 103 M�1) c

H2PO4�c Cl�. This preference for HPO4

2� may result from

the contribution of electrostatic interaction with anions of the

cationic isothiouronium. Both 18–20 and 21b are PET-type

Scheme 4

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 9: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3735

‘‘turn-on’’ fluorescent receptors that are themselves less fluores-

cent due to an efficient PET with the isothiouronium moiety

being the electron acceptor.

It is a pity that the early research was not followed up

extensively, most of the available reports coming from these

two Japanese groups. Later, isothiouroniums have recently

been employed to bind anions at the CH2Cl2–H2O interface43

and in LB films,44 to construct sensing systems using gold

nanoparticles45,46 and polythiophene47 as signal reporter, and

to template photodimerization.48 A preorganization effect was

also observed in a benzene-based tripodal receptor 23,49 that

showed selective binding of the tetrahedral oxoanion SO42�.

The high anion affinity of isothiouroniums has, in several

cases, allowed anion sensing in highly competitive solvents

including aqueous solutions.

N-Aryl(thio)ureas

N-Aryl(thio)ureas represent major (thio)urea-based anion

receptors hitherto reported in the literature, since the acidity

of the (thio)ureido –NH protons can be easily and efficiently

tuned by way of substitution at the aryl ring. In addition,

substitution at the aryl ring allows other anion binding site to

be attached. The aryl framework can also be employed to

preorganize multiple binding sites. Compared to N-alkyl-

(thio)ureas, N-aryl(thio)ureas in general exhibit stronger anion

binding ability, because of the higher acidity of the (thio)ureido

–NHs in the N-aryl derivatives. Deprotonation of the –NH

proton by basic anions may thereby be observed more

frequently. Simple N-arylthiourea-based receptor 24 bearing

highly electron-withdrawing p-NO2 substituents showed a

high binding selectivity towards acetate (CH3CO2�

c

H2PO4� > Cl�, Br�, I�, SCN�, NO3

�, HSO4�, ClO4

�) in

1% H2O–CH3CN, together with a solution color change from

colorless to yellow.50 In the presence of CH3CO2�, the color

change of 24 in 1% H2O–CH3CN was assigned to deprotona-

tion. Detailed crystal structure analysis, absorption and NMR

titrations performed on the corresponding urea 25 in CH3CN

showed that double hydrogen bonding occurs with CH3CO2�,

whereas deprotonation happens in the presence of F�,51 the

latter being clearly indicated by a dramatic change in the

solution color. In the NMR titration traces recorded in

DMSO-d6 for 25 it was noted that hydrogen bonding inter-

action led to a down-field shift of the 1H NMR signal of

N-phenyl a-CH protons next to the urea moiety, while there

was an up-field shift in the NMR signal of the b-CHs that are a

bit more away from the urea moiety. However, in the case of

deprotonation, both signals underwent an up-field shift.51

Fabbrizzi’s group has also detailed the binding behavior of

(thio)ureas 25–27,16 which nicely showed that the stability of

the 1 : 1 hydrogen bonding complex strictly relates to the

acidity of the –NH protons in the receptor and to the basicity

of the anions. Again, deprotonation of (thio)ureido –NHs

occurs in the presence of an excess amount of highly basic

anions, typically F�, which is characterized by the appearance

of a new and substantially red-shifted absorption band

and a drastic solution color change. More acidic thiourea

27a is much more easily to deprotonate than the less acidic

urea 27b.

Enhancing the hydrogen bonding of N-aryl(thio)ureas was

also proved to be possible by introducing a metal binding site

in the receptor molecules.52,53 Ureidopyridyl receptor 28 that

contains a pyridyl functionality for Ag+ binding and a urea

group for anions, reported by Steed et al.,52 showed much

higher anion binding constants when complexed with Ag+

than those of the free receptor.

Employing lanthanide metal coordination, Gunnlaugsson

et al.54,55 developed a Tb3+/diaryl-urea complex, 29�Tb(CF3SO3

� salt) that exhibited high anion affinity. Absorption

titrations indicated that the anion binding was more compli-

cated than a simple 1 : 1 stoichiometry. The high anion affinity

was attributed to the neighbouring lanthanide ion that acts

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 10: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

3736 Chem. Soc. Rev., 2010, 39, 3729–3745 This journal is c The Royal Society of Chemistry 2010

as a strong Lewis acid to enhance the hydrogen bonding

tendency of the ureido –NH protons. Meanwhile, it was found

that 29�Tb exhibited high selectivity for H2PO4� in CH3CN

because of the existence of multiple interactions of hydrogen

bonding and metal coordination.

The pull-push character of (thio)ureas has recently led to the

design of redox-active N-arylurea receptors 30 and 31 by

Sargent and Sibert et al.56 on the basis of the Wurster’s

reagent, N,N0-tetramethyl-p-phenylenediamine (p-TMPD).

Although they look structurally similar to 24–27, the TMPD

moiety in 30 and 31 allows an electrochemical voltammetric

assay of the anion binding. It is of significance to observe that

urea 31, despite bearing an electron-donating substituent,

shows in CH3CN extremely high anion binding constants up

to 106 M�1. The other feature noted with 30 and 31 was the

oxidation-induced binding enhancement that can be as high as

104-fold for the 31–CH3CO2� complex.

The high sensitivity of fluorescence detection has also

pushed fluorescent N-aryl(thio)ureas to be extensively explored.

The type of fluorophore and the way it is linked to the (thio)-

urea moiety have led to access to fluorescent receptors operating

under a variety of mechanisms. Thioureas 3213 and 3357

bearing a naphthalimide fluorophore represent good examples

of PET fluorescent receptors. 32, in which a N-arylthiourea

moiety is connected by a CH2 spacer to the naphthalimide

fluorophore at its 4-amino position, showed fluorescence

quenching upon addition of F�, CH3CO2� and H2PO4

� due

to a promoted PET. Deprotonation was observed with it at

high F� concentration. 33 differs from 32 in the position at

which the N-arylthiourea moiety is linked to the fluorophore.

An enhanced PET-type fluorescence quenching was also observed

upon anion binding to its thiourea moiety. The anion binding

natures of 32 and 33 are similar in spite of the fact that the

thiourea moiety is located, respectively, at the electron donor

and acceptor side of the fluorophore that has an excited state

of the intramolecular charge transfer (ICT) character.

Thiourea 34 containing an acridinedione fluorophore

represents another example of fluorescent receptors that has both

the PET and ICT channels in the excited state.58 It showed

specific optical signalling for F�, CH3CO2� and H2PO4

� under

different mechanisms. At low concentration of F� and CH3CO2�,

fluorescence quenching occurred due to their selective binding to

the thiourea moiety that triggers the PET from thiourea to the

relatively electron-deficient acridinedione moiety. Further addi-

tion of F� and CH3CO2�, respectively, led to deprotonation and

hydrogen bonding interaction with the amino hydrogen of the

acridinedione moiety, which leads to substantially red-shifted

emission from the ICT states, with the thiourea moiety now

being in the electron acceptor. In the presence of H2PO4�, a

fluorescence enhancement was observed because of its hydrogen

bonding with the acridinedione amino hydrogen that increases the

electron density in the electron donor group. The PET and ICT

channels with 34 can be modulated by addition of different

anions.

Fluorescent N-aryl(thio)urea receptors were also designed that

operate under the mechanism of modulating the excited state

intramolecular proton transfer (ESIPT). 35 is such an example

reported by Peng et al.59 It can distinguish F�, CH3CO2� and

H2PO4� in DMSO by its ratiometric fluorescent response as a

result of inhibition of the ESIPT to differing extents by these

anions. The ESIPT is inhibited either by F�-induced deproto-

nation of the ureido –NH or by strong hydrogen bonding of

CH3CO2� with its urea moiety. In the presence of H2PO4

�, the

ESIPT is not inhibited because the lower basicity and hydrogen

bonding acceptor ability of the anion is such that its interaction

with 35 is not strong enough to destroy the existing intra-

molecular hydrogen bond in 35. These fluorescent receptors

therefore appear to be advantageous as they could provide more

information on the sensed anions, such as basicity, size/shape

and hydrogen bonding ability.

Gunnlaugsson et al.60 developed a group of N-phenylurea

receptors 36a–c in which the urea binding site is coupled with

an amide group at the N-phenyl ring. The position of the

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 11: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3737

amide group has a significant influence on both the anion

affinity and anion binding stoichiometry. Binding of anions at

the urea moiety in 36a and 36b leads to a concomitant

activation of the amide functionality so that the latter becomes

‘‘better’’ at hydrogen bonding with the anion, a consequence

of the ‘‘positive allosteric effect’’. In contrast, the amide moiety

in 36c directly participates the cooperative anion binding event

at the urea site to form a 1 : 1 binding complex, since the urea

and amide groups are in close proximity for hydrogen bonding

to anions.

Sophisticated combination of anion binding sites via substi-

tution at the N-aryl ring can also lead to improved binding

selectivity, in addition to enhanced binding. Thiourea receptors

37 reported by Pfeffer et al.61 showed a binding selectivity for

H2PO4� against CH3CO2

�, despite the lower basicity of the

former anion. 1H NMR titrations clearly indicated that the

naphthalimide –NH at position 4, in conjunction with thioureido

–NHs, cooperatively binds to the tetrahedral H2PO4� in a 1 : 1

stoichiometry. The planar CH3CO2� anion of higher basicity

is less strongly bound by 37, which is opposite to the observion

with simple thiourea receptors without such an additional

binding site.

N-Arylurea-based receptors 38 are another set of examples

in which other binding sites are incorporated.62 38e bearing a

terminal OH group deserves more attention. It shows a higher

affinity and outstanding selectivity for carboxylates containing

a-heteroatoms, presumably due to the formation of an

additional hydrogen bond. An efficient fifth hydrogen bond

between the OH hydrogen of the hydroxamic group in 38e and

the heteroatom in the carboxylate anion was suggested to

contribute to the improved binding affinity and selectivity.62

Macrocyclic receptor 39 reported by the Gale group con-

tains a 2,6-dicarboxamidopyridine group and a N,N0-diphenyl-

urea moiety linked by two phenylamides, thus providing a

convergent set of hydrogen bonding donors for anion com-

plexation.63 1H NMR studies showed that 39 did not bind

carboxylate, dihydrogen phosphate and chloride in the same

mode, with a particularly high affinity and selectivity for

carboxylate over phosphate and chloride. The binding mode

was proposed that one oxygen atom of the carboxylate bound

to the two ureido –NHs while the other oxygen atom bound to

the 2,6-dicarboxamidopyridine –NHs.

The indole NH group has also been employed as an addi-

tional hydrogen bonding donor. A series of (thio)urea-based

receptors 40–45 bearing one or two indole groups were created

by Gale et al.64,65 Crystallographic analysis and/or 1H NMR

titrations confirmed that both the indole and urea groups

participated in the hydrogen-bonding with oxoanions. The

increase in the number of hydrogen bonding donors (2 to 4) in

those urea-based receptors was found to result in a significant

increment in the downfield shift of the ureido –NH proton

NMR signal upon hydrogen bonding to alkylcarbamate

anions.64 This supports the cooperative binding of these

donors with the anion. Diindolylurea 42 shows a remarkably

high affinity for H2PO4� in H2O–DMSO-d6, whereas its

thiourea counterpart 44 exhibits only a moderate affinity and

selectivity. This may result from a conformational inter-

conversion of the thiourea group.65

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 12: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

3738 Chem. Soc. Rev., 2010, 39, 3729–3745 This journal is c The Royal Society of Chemistry 2010

Jurczak et al.66 have recently developed 7,70-diureido-2,20-

diindolylmethane 46 using the same building block. Struc-

turally it looks like a dimer of 42. 46 is able to form anion

hydrogen bonding complexes even in the highly competitive

solvent MeOH. 1H NMR studies indicated that this is due to

the equal participation of the ureido and indolyl NH protons

in the anion binding. The conformation of 46 evidently prefers

the tetrahedral oxoanions as it binds HPO42� with eight

complementary hydrogen bonds. The isopropyl linkage for

the two 42 moieties in 46 may play a role in fixing the

conformation of 46.

Kim et al.67 reported receptors 47a–c with an alternative

combination of urea and indole binding sites, in which the two

urea moieties are attached to the 1,8-positions of a carbazole

fluorophore. They showed differing responses in both their

absorption and fluorescence towards a set of anions that allow

for the differentiation of these anions by a simple 2-dimensional

analytical approach. Although anion binding affinity and

selectivity were not the subject of that report, the preference

of these receptors towards those anions that require more

binding sites to be arranged in a suitable manner, such as

H2PO4� and HP2O7

3�, can be seen from the photograph

that shows the colors of their CH3CN–DMSO (9 : 1, v/v)

solutions.

In receptors 47 the carbazole fluorophore can actually be

considered as a rigid platform that positions the two urea

moieties. A direct consequence of this positioning is the

improvement of anion selectivity. Bisurea 48, that was built

on the rigid naphthalene motif, represents an early example

by Nam et al.68 Two urea moieties stand parallel on the 1,8-

naphthalene structure. An anion binding pocket is thus

formed and 48 displays a highly selective fluorescent response

towards the spherical F� anion in CH3CN–DMSO. The

selectivity for F� was 40 times that for Cl� because the smaller

F� anion fits well into the pocket surrounded by the four

ureido –NHs.

Bis(thio)ureas 49a–f were built on the rigid ortho-phenylene-

diamine framework,69 in which two (thio)urea moieties are

arranged at an angle of 601. Studies carried out in DMSO–

0.5%H2O illustrated that the binding affinity of carboxylate

anions can be modulated by varying the substituents and their

position in the receptor molecules. This observation suggested

that simply increasing the acidity of the (thio)ureido –NH

protons may not necessarily lead to stronger anion binding.

Although crystal structures confirmed that the anion indeed

bound to the four ureido –NHs, the specific conformation of

the receptor molecule represents a more important factor

governing the anion binding. For example, receptor 49b bearing

two electron-withdrawing m-Cl’s on the ortho-phenylene-

diamine framework showed a higher anion affinity, presumably

resulting from a more preorganized conformation by the

formation of CH� � �O IHBs between the aromatic –CHs next

to the m-Cl and the urea carbonyl O atoms. By replacing the

urea group in 49a with a thiourea group, the resultant 49f

showed unexpectedly lower binding constants with carboxylates,

by almost one order of magnitude, than those of 49a. The

larger sulfur atom in 49f that distorts the shape of the binding

site was assumed to be responsible. It appears that a con-

formation difference between thiourea and its urea counter-

part as anion receptors deserves further effort.

Colorimetric receptors 50, despite being built on the same

o-phenylenediamine framework as 49, showed a high selectivity

for the bigger and tetrahedral H2PO4� anion over other oxo-

anions such as OH�, CH3CO2� and PhCO2

�.70 A tweezer-like

binding mode was suggested for oxoanions from ab initio

calculations. Six-membered ring IHB in 50 between the

(thio)ureido –N�H at position 1 and CQ�O at position 9 of

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 13: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3739

the anthraquinone subunit may explain this different anion

selectivity of 50 from 49.

Johnson and co-workers have recently synthesized bisurea

51 based on a new rigid 2,6-bis(2-anilinoethynyl)pyridine

framework with the two urea moieties are arranged to provide

a binding cavity.71 As expected, 51 showed preferred binding

with spherical halide anions. More interestingly, the pyridine

N atom in 51 upon protonation affords an additional binding

site. The protonated 51 is a much better receptor as the

binding constants increase by more than one order of magni-

tude over those of neutral 51 and it shows an altered selectivity

for larger halide anions. The conformation of 51 converges all

available hydrogen bond donors towards the binding pocket

so that the pyridinium –NH could serve as the third point of

contact.

Macrocyclic skeletons, on the other hand, can also lead to

preorganization of the binding sites for improved anion

selectivity. Calix[n]arenes are the most employed macrocycles

since both the size and conformation of them can be varied.

Lhotak et al. attached four N-phenylurea moieties to the rim

of a calix[4]arene in its 1,3-alternate conformation.72 52 binds

exclusively only one halide anion with a distinctive size

selectivity for Cl�. This was revealed to result from a strong

negative allosteric effect, as binding of the first anion induces a

change in the conformation of 52 that disturbs the binding of

the second anion. Later, the group developed biscalix[4]arene

receptor 53 in which two urea moieties are bridged by two

calix[4]arenes on their upper rims.73 A cavity-like binding

environment was created and better halide anion size recogni-

tion was observed as NMR only probed Cl� binding to

53 in CDCl3–DMSO-d6 (4 : 1, v/v), but not for Br� and I�.

A systematic survey by 1H NMR titrations on the anion

binding profile of the ureidocalix[4]arenes with 1–4 urea units

appended to the calix[4]arene skeleton in cone, partial cone and

1,3-alternate conformations indicated that diureidocalix[4]-

arene 54 in 1,3-alternate conformation showed extremely

high binding abilities for bulky and planar acetate and

benzoate anions,74 differing from 52 that bears 4 urea

moieties. This is a clear indication of the preorganization effect

of the calixarene skeleton in the design of (thio)urea-based

receptors, the number of (thio)urea moieties, conformation of

the calixarene and the way of attaching the (thio)urea moieties

to the skeleton being the factors that should be taken into

account.

Metal coordination is another means of preorganizing the

anion binding sites of N-aryl(thio)urea receptors functionalized

with metal binding sites, yet the relatively rigid N-aryl

moiety in the receptor molecule may lead to stricter control

on the anion selectivity. Fabbrizzi et al.75 reported a 1,10-

phenanthroline-functionalized urea-based anion receptor 55.

Cu+ coordinates to the 1,10-phenanthroline ligand in 55 in a

regular tetrahedral mode, creating a rather small and strictly

defined binding pocket for anions. Modeling suggested that

Cl� could receive four hydrogen bonding donors from the two

facing urea subunits in the formed Cu(55)2+ complex

(Scheme 5). Indeed the binding constant of Cl� with

Cu(55)2+ complex is at least 350-fold higher than that with

free 55. The metal-induced enhancement of the anion binding

constant was ascribed to the preorientation led by the Cu+

Scheme 5

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 14: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

3740 Chem. Soc. Rev., 2010, 39, 3729–3745 This journal is c The Royal Society of Chemistry 2010

coordination that affords a suitable binding environment for

spherical anions.

The Fabbrizzi group has developed 56 bearing three phenyl-

urea moieties that are linked to a structurally flexible tri-

(2-aminoethyl)amine (tren) ligand.76 Coordination of 56 to a

Cu2+ ion at the tren subunit preorganizes the three phenylurea

arms, resulting in a pocket suitable for accommodating two

anions in a cascade binding mechanism in three well-defined

stepwise equilibria (Scheme 6). Binding of the first anion (Cl�)

benefits from the strong metal–ligand interaction so that Cl� is

accommodated next to the Cu2+ centre. The Cu2+ coordination

resulting pocket, despite being poor for Cl�, fits well for

efficient hydrogen bonding to the weakly basic H2PO4�,

although steric and electrostatic repulsions from Cl� may

prevent this binding. This is a nice example where metal

coordination preorganizes anion binding sites into a pocket

that accommodates two anions well, although the receptor

itself does not bind efficiently with the individual anion. In that

regard the flexibility of the alkyl components in the ligand unit

may contribute quite a lot.

N-Amido and N-amino(thio)ureas

With N-alkyl and N-aryl(thio)ureas, the (thio)urea moiety is

linked to the rest of the receptor molecule via a ‘‘C’’ atom. In

N-amino and N-amido(thio)ureas, an ‘‘N’’ atom is such a

‘‘linkage’’. The development of these receptors was based on

an idea to create (thio)ureas with high anion affinity yet the

acidity of the (thio)ureido –NH protons remains low so that

they suffer less from deprotonation in the presence of basic

anions. In N-benzamidothioureas 57 and 58 the thiourea

anion-binding site is connected to the benzamide moiety, the

signal reporter, by an N–N bond.77–79 The N–N single bond

was shown to be highly twisted, thus stopping the electronic

communication between the anion binding site and the benzamide

chromophore. The acidity of the thioureido –NH protons is

indeed lower than that of the classic N-phenylthioureas. In

spite of these unfavorable structural characteristics, they

show substantially enhanced anion affinities, compared to

the correspondingN-phenylthiourea counterparts. For example,

the binding constants with AcO� in CH3CN are, at 105–107 M�1

orders of magnitude, 101–102 times those of the corresponding

N-phenylthioureas. Hydrogen bonding of anions with the

thiourea moiety in N-benzamidothioureas was confirmed.

An interesting substituent effects on the anion binding con-

stant was found with 57 and 58. While the substituent located

at the N-benzamido phenyl ring of 58 does not influence the

anion affinity much,78 the effect of the substituent in 57 located

in the N0-phenyl ring is amplified.79 A conformation change

around the N–N bond and the formation of a hydrogen

bonding network in the anion–amidothiourea complex were

shown to be responsible for the enhanced anion affinity and

amplified substituent effect,79 the N-benzamidothioureas

being, therefore, allosteric receptors (Scheme 7). A ground-

state ICT was identified in the anion binding complex by

a substantially red-shifted absorption,78 which generates a

positive charge in the thiourea moiety that in turn reinforces

the anion binding. A substantially increased electron-donating

ability of the N-aminothiourea upon anion binding was thereby

concluded and later verified in a PET-based N-amidothiourea

receptor (69).

Extended examination on N-ferrocamido-N0-(substituted-

phenyl)thioureas 59 containing an electrochemically active

ferrocene moiety led to similar conclusions.80 Although the

electronic communication between the ferrocene moiety and

the thiourea anion-binding site was blocked by the twisted

N–N bond, a ca. �200 mV shift in the oxidation potential was

observed in the presence of anions such as CH3CO2� and F�

in CH3CN.

Because of the substantially enhanced anion affinity in

CH3CN of 105–107 M�1 orders of magnitude for CH3CO2�,

for example, the possibility of anion complexation in a

competitive and practically important solvent, water, was

explored on 60.81 A hydrophobic microenvironment around

the thiourea binding site was shown to exist. The N-pyridine

group in 60 appears to be important because of its electron

deficient character so that both the hydrophobic and electro-

static interactions occur between the two aromatic terminals.

Indeed, 60 in pure water shows efficient binding with anions

such as CH3CO2�, with binding constants in the 103 M�1

order of magnitude that are comparable to or even higher than

those of the classicalN,N0-diphenylthioureas in a noncompetitive

solvent such as CH3CN. This provides a new clue for designing

Scheme 6

Scheme 7

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 15: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3741

N-amidothiourea-based receptors for anion binding in aqueous

solutions by taking into account the receptor molecular con-

formation.

The Gale group reported a series of pyrrole-functionalized

N-amido(thio)ureas 61 and 62.82 The urea receptor 61b

showed hydrogen bonding interactions with anions in DMSO,

whereas no significant interaction was observed for 61a. The

analogous thiourea receptor 61d deprotonated upon the addi-

tion of only one equivalent of relatively weakly basic anions

such as CH3CO2� and H2PO4

�, which can be visualized

by solution color changes. A non-convergent nature of the

hydrogen bonding array in this system was suggested as it fails

to adopt a geometry suitable for stabilizing a hydrogen bonding

complex, thus leading to a neat deprotonation. It is worth-

while to add that both the X-ray crystal structure and NMR

titrations indicate that deprotonation occurs at the thioureido

–NH next to the pyrrole moiety, despite the higher acidity of

the pyrrole –NH proton. DisubstitutedN-amidourea receptors

62 with a more favourable cleft-like binding mode showed

enhanced affinities for anions compared to the mono-substituted

derivative 61b.

Using the same framework, Kim’s group developed

N-(2,7-dichlorofluorescein)lactam-N0-phenylthiourea 63,83 that

underwent an anion-induced ring-opening of the spirolactam.

Therefore, 63 in CH3CN exhibited a distinct fluores-

cence ‘‘turn-on’’ upon CH3CO2� addition as well as a solution

color change from colorless to pink. This represents the first

example of the ring-opening mechanism applied in anion

sensing.

The fact that the substituent on the N-benzamide group of

58 does not exert an appreciable influence on the anion

binding suggests that N-aliphatic counterparts of 58 can be

good receptors too. N-Acetamidothioureas 64 and 6584 are

indeed efficient anion receptors, with even higher anion affinities

(up to 107 M�1 for CH3CO2�) than those of 57 and 58, despite

the fact that the acidity of the thioureido –NHs in 64 and 65 is

lower than that in 57 and 58. The N–N single bond in 64 and

65 was identified to be twisted too, but less than that in 57

and 58. The results of 57, 58 and 64, 65 confirmed that

N-amidothioureas derived from both N-aliphatic and

N-aromatic amides can be a family of efficient hydrogen

bonding anion receptors. An N-acetamidothiourea was shown

to be an efficient organocatalyst too.84

Following the conclusion that a hydrogen bonding network

exists in the anion binding complexes of 57 and 58 (Scheme 7),

an IHB in the N-benzamide moiety was expected to facili-

tate this network and therefore anion binding (Scheme 8).

N-(o-Substituted-benzamido)thioureas bearing an ortho-

substituent capable of forming IHBs were examined.85 The

o-OMe derivative 66b, in which an IHB was confirmed, indeed

exhibited a CH3CO2� binding constant in CH3CN over 107 M�1,

much higher than that of 66c (105 M�1) having no such IHB.

Similar enhancements in anion binding were shown with 66a,

66d, 66e and 66f.85 This was also observed with extended

examples of 67 and 68 with an IHB in the N-benzamide

moiety. Both 67 and 68 in CH3CN showed CH3CO2� binding

constants over 107 M�1 orders of magnitude and a red-shifted

absorption of the anion binding complex compared to those of

N-benzamidothioureas 58. This IHB appears to extend the

conjugation in the N-benzamide moiety.85 It is interesting that

the substituent effect on the anion binding of 67 and 68 is in

general similar to that with 57 and 58.

With 57, 58 and 67, 68, it was concluded that the

N-aminothiourea unit, upon binding to an anion, is a

much better electron donor than a simple thiourea unit.

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 16: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

3742 Chem. Soc. Rev., 2010, 39, 3729–3745 This journal is c The Royal Society of Chemistry 2010

N-Amidothiourea was introduced into neutral PET receptors

69 in which the anion binding site is linked to the pyrene

fluorophore via a relatively longer (CH2)3 spacer.86 Both the

anion quenching and binding constants of 69 are much higher

than those of the corresponding PET receptors bearing a

simple thiourea unit with a shorter CH2 spacer. Because of

the extremely high binding constants, 69c showed a sensitive

spectral response towards CH3CO2� in 8%H2O–CH3CN

(v/v). These observations confirm that N-amidothiourea unit

is much better as a part of the electron donor for constructing

PET anion receptors than thiourea itself.

The preorganization of the N-amido(thio)urea moieties was

also examined. Gunnlaugsson et al.87 developed a cleft-like

bis-amidothiourea receptor 70 using a pyridine skeleton. 70 is

a nice colorimetric receptor that even allows for ‘‘naked-eye’’

assays for several biologically important anions such as AMP

and ADP in 4 : 1 DMSO–H2O solution. High selectivity for

AMP and ADP over ATP was observed and assigned to both

hydrogen bonding interaction and deprotonation.

The macrocyclic calix[4]arene framework was again employed

for preorganizing multiple N-amido(thio)urea moieties.

Calix[4]arene 71 bearing four N-amidourea moieties, reported

by Gunnlaugsson et al.,88 affords a symmetrical and preorganized

hydrogen bonding cavity. It shows highly selective and

efficient binding with pyrophosphate and Br� in 1 : 1 stoichio-

metry. With F� and CH3CO2�, however, 2 : 1 binding

complexes are formed. The preorganization effect does lead

to efficient binding with Cl� but not Br�, differing from the

observation with 52. This might result from the flexible

–OCH2– linkage in 71. Considering the formation of a planar

and therefore rigid hydrogen bonding network in the anion

bonding complexes of N-amido(thio)urea receptors (Scheme 7),

preorganization effects using a macrocyclic skeleton with rigid

linkages can be further explored for better performance in terms

of anion affinity and selectivity.

Clip-like N-amidothiourea-based receptor 72 was able to

distinguish o-phthalate from m-phthalate and p-phthalate

isomers, following a conformational switching mechanism.89

The high selectivity for o-phthalate results from a special

conformation of 72, in that two naphthalene units position

close to fit o-phthalate well. Receptor conformation switching

therefore represents another means that leads to preorganiztion

effect.

Two additional attempts were made to extend the structural

diversity of N-amidothiourea receptors. The first was based on

the comparison of the anion affinity of N-benzamido- and

N-phenylthioureas; the former is much better in hydrogen

bonding with anions such as CH3CO2�. The structural

difference of these two kinds of receptors is an amide group,

–C(O)NH–. N-Benzoylthiourea, because of an IHB between

the N-acyl CQO and the other thioureido –NH, is unable to

bind CH3CO2�. The amide group was therefore inserted into

N-benzoylthiourea to lead toN-benzamido-N0-benzoylthioureas

73 and 74.90 N-Benzoylthioureas 73 and 74 are indeed activated

by this amide group and show anion binding affinity up to one

order of magnitude higher than that of 57 and 58. The anion

binding mode and its consequences for 73 and 74 are similar to

those with 57 and 58 in that an anion binding-induced con-

formation change around the N–N single bond takes place.

Breaking of the IHB in 73 and 74 was therefore revealed to

occur upon anion binding (Scheme 9).

Scheme 8

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 17: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3743

Further extension led to N,N0-bis(substituted-benzamido)-

thioureas 75 that are in the ‘‘A–D–A’’ structural framework.91

75 shows a much higher anion affinity in CH3CN as the

binding constants of F� and CH3CO2� are 107 M�1 or higher

orders of magnitude. The ‘‘–C(O)NH–’’ group was hence

demonstrated to be useful in promoting anion binding of the

thiourea-based receptors.

The development of N-anilinothioureas with a relatively

acidic N-anilino –NH proton was another rewarding attempt

following the N-benzamidothiourea research. Because of the

weaker hydrogen bonding ability of the thiourea CQ�S atom,

the acidity of the N-benzamido –N�H proton of 57 and 58

appears to be more important in maintaining the hydrogen

bonding network in their anion binding complexes (Scheme 7).

A structural design was hence performed such that CQO

in the N-benzamide moiety in 57 and 58 was moved as

an electron-withdrawing group to the other position of the

N-benzamide phenyl ring, leading to N-anilinothioureas that

may function similarly to their precursors 57 and 58 in anion

binding. N-Anilino-N0-phenylthioureas 76 and 77, with the

N-anilino –NH proton highly acidic and of varying acidity,

were developed.92 76 and part of 77 that bears a substituent X

not less electron-withdrawing than m-Cl show dramatically

enhanced anion binding. The CH3CO2� binding constants in

CH3CN are at 106–107 M�1 orders of magnitude. The N–N

bond was found to be twisted and a conformation change was

shown to occur upon anion binding that leads to a hydrogen

bonding network in the anion binding complex, as in the cases

of 57 and 58.

Lin et al. reported a N-anilinourea receptor 78 containing a

2,4-dinitroaniline.93 Anion binding affinities of 78 in dry

DMSO vary in the order H2PO4� > F� > CH3CO2

� BCl� B Br� B I�, despite the lower basicity of H2PO4

� than

that of F� and CH3CO2�. The efficient H2PO4

� binding was

assigned to the 1 : 1 anion complexation by multiple hydrogen

bonds. Two electron-withdrawing –NO2 groups in 78 lead

the 5-CH proton of the same phenyl ring to be acidic to

participate hydrogen bonding with H2PO4�.

Efficient anion binding of N-(naphthylamino)thioureas 79

and ureas 80 bearing an acidic N-naphthanilino –NH were

reported by Gunnlaugsson et al. The recognition and sensing

by 79 for CH3CO2�, H2PO4

� and F� anions can be achieved

in competitive pH-buffered aqueous solutions, through the

hydrogen bonding interaction-modulated ICT of the receptor

molecules.94 This reflects the high anion affinity of 79. 80, as

the urea counterparts of 79, also showed large colorimetric

changes visible to the naked eye upon addition of CH3CO2�,

H2PO4� and F� in DMSO.95 The final absorption spectrum is

fingerprinting for individual anions, since each anion gives a

unique absorption spectrum due to its hydrogen bonding

interaction with and/or deprotonation of the –NH in 80.

However, anion sensing by 80 could not be carried out in

aqueous media.

Conclusions and perspectives

We have shown that N,N0-substitution has made available a

great variety of (thio)urea-based receptors for anion com-

plexation and sensing. While the involved interactions with

anions are mainly of hydrogen bonding nature, electro-

static interactions and deprotonation of the receptors happen

too. With N-alkyl(thio)ureas, the alkyl group is changed to

position the receptor in a microenvironment and to allow the

Scheme 9

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 18: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

3744 Chem. Soc. Rev., 2010, 39, 3729–3745 This journal is c The Royal Society of Chemistry 2010

receptor to take a good conformation or afford a suitable binding

environment. Hydrophobicity and structural flexibility of the

alkyl chain are the reasons for these possibilities. The alkyl

substituent, however, does not allow efficient electronic communi-

cation so that the acidity of the (thio)ureido –NHs can not be well

tuned. With N-aryl(thio)ureas the acidity of these –NHs can be

easily tuned via substitution on the aryl ring. Additional binding

sites can be introduced to, and well arranged on, the aryl ring for

stronger binding and/or higher selectivity. N-Amido and

N-amino(thio)ureas are relatively young receptors in the big

family of (thio)urea-based anion receptors. It is now known that

with N-amido- and N-amino(thio)ureas a conformation change

around the N–N bond occurs upon anion binding, and they are

allosteric receptors. The substituent at the N0-phenyl ring exerts

an amplified effect on the anion binding to the N-amido-

N0-phenylthioureas. In all these three kinds of receptors pre-

organization effects can be achieved by attaching multiple

(thio)urea binding sites to a structural motif, flexible or rigid, to

promote anion binding and to improve anion selectivity.

However, there is space for future modifications to the

(thio)urea core. Combination of two of the described three

types of receptors could be a choice that may lead to better

receptors if the preferred merits of each kind are taken. A

comparison of the conformational differences led by structural

modifications on thiourea and urea cores could be another way

to go. The isothiouroniums deserve further attention, since not

much is yet known of their binding nature, for example the

contribution, tuning and interplay of the hydrogen bonding and

electrostatic interactions with a given anion. The S-substituents

in isothiouroniums available at the moment have mainly been

–CH2Ar (Ar = aryl such as phenyl, 2-naphthyl, 9-anthryl),

while the N,N0-subtituents are mainly alkyl groups. The aryl

substituents on these positions are expected to be examined.

Another issue could be the photostability of the isothio-

uroniums in order to perform credible investigations. In all of

the described (thio)urea-based receptors, the (thio)urea moiety

is linked with the rest of the molecules by either a ‘‘C’’ or an

‘‘N’’ atom. An ‘‘O’’ linkage may thus be considered to examine

N-alkoxy(thio)urea-based receptors.

It shall also be pointed out that, while the structurally modified

(thio)ureas have been extensively explored and employed as

anion receptors, they have recently been increasingly applied to

organocatalysis,84,96,97 crystal engineering,98 supramolecular

gelation,99,100 and membrane transporters for anions.101 These

applications have benefited and will continue to benefit from the

investigations of (thio)urea-based receptors and their interactions

with anions. This will in turn further promote the development of

the general anion-receptor chemistry.1,102

Acknowledgements

We thank the NSFC of China for support through grants

20675069 and 20835005.

References

1 J. L. Sessler, P. A. Gale and W.-S. Cho, Anion ReceptorChemistry, RSC Publishing, Cambridge, UK, 2006.

2 C. Caltagirone and P. A. Gale, Chem. Soc. Rev., 2009, 38,520–563.

3 V. Amendola and L. Fabbrizzi, Chem. Commun., 2009, 513–531.4 J. Perez and L. Riera, Chem. Soc. Rev., 2008, 37, 2658–2667.5 P. A. Gale, S. E. Garcia-Garrido and J. Garric, Chem. Soc. Rev.,

2008, 37, 151–190.6 P. A. Gale, Acc. Chem. Res., 2006, 39, 465–475.7 P. A. Gale and R. Quesada, Coord. Chem. Rev., 2006, 250,

3219–3244.8 T. Gunnlaugsson, M. Glynn, G. M. Tocci, P. E. Kruger and

F. M. Pfeffer, Coord. Chem. Rev., 2006, 250, 3094–3117.9 V. Amendola, M. Bonizzoni, D. Esteban-Gomez, L. Fabbrizzi,

M. Licchelli, F. Sancenn and A. Taglietti, Coord. Chem. Rev.,2006, 250, 1451–1470.

10 P. J. Smith, M. V. Reddington and C. S. Wilcox, TetrahedronLett., 1992, 33, 6085–6088.

11 E. Fan, S. A. Van Arman, S. Kincaid and A. D. Hamilton, J. Am.Chem. Soc., 1993, 115, 369–370.

12 F. G. Bordwell, Acc. Chem. Res., 1988, 21, 456–463.13 T. Gunnlaugsson, P. E. Kruger, T. C. Lee, R. Parkesh,

F. M. Pfeffer and G. M. Hussey, Tetrahedron Lett., 2003, 44,6575–6578.

14 T. Gunnlaugsson, P. E. Kruger, P. Jensen, F. M. Pfeffer andG. M. Hussey, Tetrahedron Lett., 2003, 44, 8909–8913.

15 S. Camiolo, P. A. Gale, M. B. Hursthouse and M. E. Light,Org. Biomol. Chem., 2003, 1, 741–744.

16 V. Amendola, D. Esteban-Gomez, L. Fabbrizzi and M. Licchelli,Acc. Chem. Res., 2006, 39, 343–353.

17 C. S. Wilcox, E. Kim, D. Romano, L. H. Kuo, A. L. Burt andD. P. Curran, Tetrahedron, 1995, 51, 621–634.

18 J. V. Ros-Lis, R. Martınez-Manez, F. Sancenon, J. Soto,K. Rurack and H. Weibhoff, Eur. J. Org. Chem., 2007,2449–2458.

19 T. Hayashita, T. Onodera, R. Kato, S. Nishizawa andN. Teramae, Chem. Commun., 2000, 755–756.

20 V. Amendola, D. Esteban-Gomez, L. Fabbrizzi, M. Licchelli,E. Monzani and F. Sancenon, Inorg. Chem., 2005, 44, 8690–8698.

21 J. Wu, Y. He, Z. Zeng, L. Wei, L. Meng and T. Yang,Tetrahedron, 2004, 60, 4309–4314.

22 F. M. Pfeffer, K. F. Lim and K. J. Sedgwick, Org. Biomol. Chem.,2007, 5, 1795–1799.

23 X. Qian and F. Liu, Tetrahedron Lett., 2003, 44, 795–799.24 J. L. J. Blanco, J. M. Benito, C. O. Mellet and J. M.

G. Fernandez, Org.Lett., 1999, 1, 1217–1220.25 T. Gunnlaugsson, A. P. Davis and M. Glynn, Chem. Commun.,

2001, 2556–2557.26 T. Gunnlaugsson, A. P. Davis, J. E. O’Brien and M. Glynn, Org.

Biomol. Chem., 2005, 3, 48–56.27 S. Nishizawa, H. Kaneda, T. Uchida and N. Teramae, J. Chem.

Soc., Perkin Trans. 2, 1998, 2325–2327.28 S. Kondo and M. Sato, Tetrahedron, 2006, 62, 4844–4850.29 D. H. Lee, H. Y. Lee, K. H. Lee and J. Hong, Chem. Commun.,

2001, 1188–1189.30 A. J. Ayling, M. N. Perez-Payan and A. P. Davis, J. Am. Chem.

Soc., 2001, 123, 12716–12717.31 A. J. Lowe and F. M. Pfeffer, Org. Biomol. Chem., 2009, 7,

4233–4240.32 A. J. Lowe and F. M. Pfeffer, Chem. Commun., 2008, 1871–1873.33 F. M. Pfeffer, T. Gunnlaugsson, P. Jensen and P. E. Kruger, Org.

Lett., 2005, 7, 5357–5360.34 F. M. Pfeffer, P. E. Kruger and T. Gunnlaugsson, Org. Biomol.

Chem., 2007, 5, 1894–1902.35 B. H. M. Snellink-Ruel, M. M. G. Antonisse, J. F. J. Engbersen,

P. Timmerman and D. N. Reinhoudt, Eur. J. Org. Chem., 2000,165–170.

36 C. R. Bondy, P. A. Gale and S. J. Loeb, J. Am. Chem. Soc., 2004,126, 5030–5031.

37 W.-S. Yeo and J.-I. Hong, Tetrahedron Lett., 1998, 39,3769–3772.

38 W.-S. Yeo and J.-I. Hong, Tetrahedron Lett., 1998, 39,8137–8140.

39 Y. Kubo, M. Tsukahara, S. Ishihara and S. Tokita, Chem.Commun., 2000, 653–654.

40 Y. Kubo, S. Ishihara, M. Tsukahara and S. Tokita, J. Chem. Soc.,Perkin Trans. 2, 2002, 1455–1460.

41 Y. Kubo, M. Kato, Y. Misawa and S. Tokita, Tetrahedron Lett.,2004, 45, 3769–3773.

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online

Page 19: Guest editors Philip Gale and Thorfinnur Gunnlaugsson · Anion coordination chemistry has attracted growing attention in supramolecular chemistry due to the essential roles that anions

This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 3729–3745 3745

42 S. Nishizawa, Y.-Y. Cui, Minagawa, K. Morita, Y. Kato,S. Tanikuchi, R. Kato and N. Teramae, J. Chem. Soc., PerkinTrans. 2, 2002, 866–870.

43 R. Kato, Y.-Y. Cui, S. Nishizawa, T. Yokobori and N. Teramae,Tetrahedron Lett., 2004, 45, 4273–4276.

44 Y. Misawa, Y. Kubo, S. Tokita, H. Ohkuma and H. Nakahara,Chem. Lett., 2004, 33, 1118–1119.

45 Y. Kubo, S. Uchida, Y. Kemmochi and T. Okubo, TetrahedronLett., 2005, 46, 4369–4372.

46 T. Minami, K. Kaneko, T. Nagasaki and Y. Kubo, TetrahedronLett., 2008, 49, 432–436.

47 T. Minami and Y. Kubo, Chem.–Asian J., 2010, 5, 605–611.48 Y. Kato, S. Nishizawa and N. Teramae, Org. Lett., 2002, 4,

4407–4410.49 H. R. Seong, D. S. Kim, S. G. Kim, H. J. Choi and K. H. Ahn,

Tetrahedron Lett., 2004, 45, 723–727.50 R. Kato, S. Nishizawa, T. Hayashita and N. Teramae, Tetrahedron

Lett., 2001, 42, 5053–5056.51 M. Boiocchi, L. D. Boca, D. Esteban-Gomez, L. Fabbrizzi,

M. Licchelli and E. Monzani, J. Am. Chem. Soc., 2004, 126,16507–16514.

52 D. R. Turner, B. Smith, E. C. Spencer, A. E. Goeta, I. R. Evans,D. A. Tocher, J. A. K. Howard and J. W. Steed, New J. Chem.,2005, 29, 90–98.

53 N. Qureshi, D. S. Yufit, J. A. K. Howard and J. W. Steed, DaltonTrans., 2009, 5708–5714.

54 C. M. G. dos Santos, P. B. Fernandez, S. E. Plush, J. P. Leonardand T. Gunnlaugsson, Chem. Commun., 2007, 3389–3391.

55 C. M. G. dos Santos and T. Gunnlaugsson, Dalton Trans., 2009,4712–4721.

56 J. P. Clare, A. Statnikov, V. Lynch, A. L. Sargent andJ. W. Sibert, J. Org. Chem., 2009, 74, 6637–6646.

57 E. B. Veale and T. Gunnlaugsson, J. Org. Chem., 2008, 73,8073–8076.

58 V. Thiagarajan, P. Ramamurthy, D. Thirumalai andV. T. Ramakrishnan, Org. Lett., 2005, 7, 657–660.

59 Y. Wu, X. Peng, J. Fan, S. Gao, M. Tian, J. Zhao and S. Sun,J. Org. Chem., 2007, 72, 62–70.

60 C. M. G. dos Santos, T. McCabe, G. W. Watson, P. E. Krugerand T. Gunnlaugsson, J. Org. Chem., 2008, 73, 9235–9244.

61 F. M. Pfeffer, M. Seter, N. Lewcenko and N. W. Barnett,Tetrahedron Lett., 2006, 47, 5241–5245.

62 Ma. F. de la Torre, E. G. Campos, S. Gonzalez, J. R. Moran andMa. C. Caballero, Tetrahedron, 2001, 57, 3945–3950.

63 S. J. Brooks, P. A. Gale and M. E. Light, Chem. Commun., 2006,4344–4346.

64 P. R. Edwards, J. R. Hiscock and P. A. Gale, Tetrahedron Lett.,2009, 50, 4922–4924.

65 C. Caltagirone, J. R. Hiscock, M. B. Hursthouse, M. E. Light andP. A. Gale, Chem.–Eur. J., 2008, 14, 10236–10243.

66 P. Dydio, T. Zielinski and J. Jurczak, Org. Lett., 2010, 12,1076–1078.

67 T. D. Thangadurai, N. J. Singh, I. Hwang, J. W. Lee,R. P. Chandran and K. S. Kim, J. Org. Chem., 2007, 72,5461–5464.

68 E. J. Cho, J. W. Moon, S. W. Ko, J. Y. Lee, S. K. Kim, J. Yoonand K. C. Nam, J. Am. Chem. Soc., 2003, 125, 12376–12377.

69 S. J. Brooks, P. R. Edwards, P. A. Gale and M. E. Light, New J.Chem., 2006, 30, 65–70.

70 D. A. Jose, D. K. Kumar, B. Ganguly and A. Das, TetrahedronLett., 2005, 46, 5343–5346.

71 C. N. Carroll, O. B. Berryman, C. A. Johnson II, L. N. Zakharov,M.M.Haley andD.W. Johnson,Chem. Commun., 2009, 2520–2522.

72 J. Budka, P. Lhotak, V. Michlova and I. Stibor, TetrahedronLett., 2001, 42, 1583–1586.

73 V. Stastny, P. Lhotak, V. Michlova, I. Stibor and J. Sykora,Tetrahedron, 2002, 58, 7207–7211.

74 I. Stibor, J. Budka, V. Michlova, M. Tkadlecova, M. Pojarova,P. Curınova and P. Lhotak, New J. Chem., 2008, 32,1597–1607.

75 V. Amendola, M. Boiocchi, B. Colasson and L. Fabbrizzi, Inorg.Chem., 2006, 45, 6138–6147.

76 M. Allevi, M. Bonizzoni and L. Fabbrizzi, Chem.–Eur. J., 2007,13, 3787–3795.

77 F.-Y. Wu, Z. Li, Z.-C. Wen, N. Zhou, Y.-F. Zhao andY.-B. Jiang, Org. Lett., 2002, 4, 3203–3205.

78 L. Nie, Z. Li, J. Han, X. Zhang, R. Yang, W.-X. Liu, F.-Y. Wu,J.-W. Xie, Y.-F. Zhao and Y.-B. Jiang, J. Org. Chem., 2004, 69,6449–6454.

79 F.-Y. Wu, Z. Li, L. Guo, X. Wang, M.-H. Lin, Y.-F. Zhao andY.-B. Jiang, Org. Biomol. Chem., 2006, 4, 624–630.

80 J. Han, Z. Li, W.-X. Liu, R. Yang and Y.-B. Jiang, Acta Chim.Sin., 2006, 64, 1716–1722.

81 R. Yang, W.-X. Liu, H. Shen, H.-H. Huang and Y.-B. Jiang,J. Phys. Chem. B, 2008, 112, 5105–5110.

82 L. S. Evans, P. A. Gale, M. E. Light and R. Quesada, New J.Chem., 2006, 30, 1019–1025.

83 J.-S. Wu, H. J. Kim, M. H. Lee, J. H. Yoon, J. H. Lee andJ. S. Kim, Tetrahedron Lett., 2007, 48, 3159–3162.

84 W.-X. Liu, R. Yang, A.-F. Li, Z. Li, Y.-F. Gao, X.-X. Luo,Y.-B. Ruan and Y. B. Jiang, Org. Biomol. Chem., 2009, 7,4021–4028.

85 Q.-Q. Jiang, B. Darhkijav, H. Liu, F. Wang, Z. Li andY.-B. Jiang, Chem.–Asian J., 2010, 5, 543–549.

86 W.-X. Liu and Y.-B. Jiang, Org. Biomol. Chem., 2007, 5,1771–1775.

87 R. M. Duke, J. E. O’Brien, T. McCabe and T. Gunnlaugsson,Org. Biomol. Chem., 2008, 6, 4089–4092.

88 E. Quinlan, S. E. Matthews and T. Gunnlaugsson, J. Org. Chem.,2007, 72, 7497–7503.

89 Z.-H. Lin, L.-X. Xie, Y.-G. Zhao, C.-Y. Duan and J.-P. Qu, Org.Biomol. Chem., 2007, 5, 3535–3538.

90 W.-X. Liu and Y.-B. Jiang, J. Org. Chem., 2008, 73,1124–1127.

91 Z. Li, Z. Liu, Q.-X. Liao, Z.-B. Wei, L.-S. Long and Y.-B. Jiang,C. R. Chimie, 2008, 11, 67–72.

92 Z. Li, F.-Y. Wu, L. Guo, A.-F. Li and Y.-B. Jiang, J. Phys. Chem.B, 2008, 112, 7071–7079.

93 J. Shao, H. Lin and H. K. Lin, Spectrochim. Acta, Part A, 2008,70, 682–685.

94 T. Gunnlaugsson, P. E. Kruger, P. Jensen, J. Tierney, H. D. P. Aliand G. M. Hussey, J. Org. Chem., 2005, 70, 10875–10878.

95 H. D. P. Ali, P. E. Kruger and T. Gunnlaugsson, New J. Chem.,2008, 32, 1153–1161.

96 Z. Zhang and P. R. Schreiner, Chem. Soc. Rev., 2009, 38,1187–1198.

97 H. Xu, S. J. Zuend, M. G. Woll, Y. Tao and E. N. Jacobsen,Science, 2010, 327, 986–990.

98 R. Custelcean, Chem. Commun., 2008, 295–307.99 M.-O. M. Piepenbrock, G. O. Lloyd, N. Clarke and J. W. Steed,

Chem. Commun., 2008, 2644–2646.100 M.-O. M. Piepenbrock, G. O. Lloyd, N. Clarke and J. W. Steed,

Chem. Rev., 2010, 110, 1960–2004.101 B. A. McNally, E. J. O’Neil, A. Nguyen and B. D. Smith, J. Am.

Chem. Soc., 2008, 130, 17274–17275.102 P. A. Gale, Chem. Commun., 2011, DOI: 10.1039/c0cc00656d.

Dow

nloa

ded

by X

iam

en U

nive

rsity

on

29 J

anua

ry 2

011

Publ

ishe

d on

25

Aug

ust 2

010

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B92

6160

P

View Online