Next Article in Journal / Special Issue
Sensor for Silver(I) Ion Based on Schiff-base-p-tertbutylcalix[4]arene
Previous Article in Journal / Special Issue
Dibenzocyclamnickel(II) as Ionophore in PVC-Matrix for Ni2+-Selective Sensor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Calixarene-Based Molecules for Cation Recognition

by
Rainer Ludwig
* and
Nguyen Thi Kim Dzung
Institute of Chemistry/Inorg. and Analyt. Chem., Freie Universitaet Berlin, Fabeckstr. 34-36, 14195 Berlin, Germany
*
Author to whom correspondence should be addressed.
Sensors 2002, 2(10), 397-416; https://doi.org/10.3390/s21000397
Submission received: 24 September 2002 / Accepted: 1 October 2002 / Published: 11 October 2002
(This article belongs to the Special Issue Ion Sensors)

Abstract

:
This review discusses molecular design principles of calixarene-type macrocycles for ion recognition and gives examples on the relationship between structure and selectivity, however without attempting to cover all the different approaches.

Introduction

Macrocyclic compounds such as crown ethers and cryptands are well-known for their relationship be-tween molecular design and selectivity in complexation of ions. When aromatic subunits are bridged by spacers forming a ring, these macrocycles are called cyclophanes. Phenolic units bridged by methylene spacers in meta-position are called calixarenes, although in recent years macrocycles with related subunits such as resorcin or pyrrol and other spacers such as sulfur are also considered to belong to the same class. Their synthesis, properties and applications are subject of several books, e.g. [1,2,3,4,5,6] and reviews such as [7,8,9,10,11] and many more as cited in [in 3, p.4-5].
Figure 1. tert-Butylcalix[6]arene, comprising 6 CH2-bridged phenyl rings. The conformation, where substituents at the phenyl rings show into the same direction, is called ‘cone’.
Figure 1. tert-Butylcalix[6]arene, comprising 6 CH2-bridged phenyl rings. The conformation, where substituents at the phenyl rings show into the same direction, is called ‘cone’.
Sensors 02 00397 g001
The macrocyclic ring in calixarenes acts as a molecular backbone to which ligating functional groups are attached. There are many examples however where the ring itself engages in binding such as with its π-basic phenolic cavities to class B cations, to class A cations with the phenolic oxygen atoms or to organic molecules by CH-π or π-π-stacking interactions.
The molecular design allows the rational control of binding properties such as complex stability and selectivity. The main design features are:
a) 
The size of the cavity is large enough to accommodate the metal cation. During complexation, the hydration shell is removed from the cation and substituted by the donor atoms of the ligand.
b) 
There is a sufficient number of donor atoms in the ligand in order to match the coordination number.
c) 
The donor atoms are held by the calixarene backbone with limited flexibility in positions suitable to match the shape of the coordination sphere.
d) 
The ‘classical’ mechanisms of complexation apply to calixarenes as well, ion exchange or ion pairing, with the difference that now several iononizable, chelation or solvating donor groups are combined within one molecule.
e) 
Often, synergy is observed if these are different functional groups, e.g. neutral and ionizable.
f) 
Using the macrocyclic backbone, an excess of certain ligating groups can be avoided, thus improving the selectivity. E.g. for a divalent cation two ionizable groups are introduced into calix[4]arene which could hold a maximum of four of them. The remaining two neutral binding groups fill up the coordina-tion sphere.
g) 
There is a possibility to attach chelating moieties onto calixarenes, thus combining the chelate and the macrocyclic effect. However, the chelating groups increase the flexibility of the ligand and may reduce the overall selectivity. Our own approach is therefore the use of short functional groups.
h) 
Branched alkyl groups are attached to the phenyl rings for high hydrophobicity and to avoid crystal-lized membrane phases.
Only a few examples for ion recognition were selected for the following chapters out of the well-over 3000 references on calixarenes so far.

Alkali Ion Recognition

Lithium. As it is the case for crown ethers, calixarenes with oxygen donor atoms turned out to be suitable for selectively binding alkali ions. The ligands are more hydrophobic compared with crown ethers and the membranes therefore are more stable. Although the ring formed by phenolic oxygens in calix[4]arenes (2 Å diameter [12]) is well suited for the diameter of lithium ion (1.36 Å), due to its strong hydration, it is not situated inside the cavity of 1 but at its rim instead [13]. The stronger binding of Li+ to the phenolate anion compared with the other alkali ions can be visual-ized by chromo- and fluoroionophores comprised of a calix[4]arene back-bone and ligating as well as signalling groups [14,15] such as in ligands 2a and 3.
Sensors 02 00397 i001
Sensors 02 00397 i002
Nitrophenol or azophenol moieties on calix[4]arenes equipped with additional ester groups [16,17,18,19] transform the Li/Na-selectivity in organic solutions into a batho-chromic shift from 350 to 425 nm with the help of an auxiliary base to support deprotonation. Ligand 4a as example is characterized by a sensitivity of 10-5 M Li+ and selectiv-ity of log KLi,Na = -1.5. This value improves to log KLi,Na = -1.9 (THF/H2O) with the related ligand 4b [17]. PVC-optodes, which apply the 520 nm absorbance, retain this selectivity [20]. The calixarene with a nitrophenylazophenol group is rather versatile: It not only detects lithium ions in the presence of an weak base, but in turn detects weak bases such as volatile amines when Li+ is already present in the membrane [20,21,22].
Sensors 02 00397 i003
Despite the ligand structure, the solute-solvent interactions are an important factor contributing to the alkali ion selectivity [23]. For example, the selectivity of ligand 2b for Li+ in solution changes towards Na+ in optode membranes due to solvent effects (log KNa,Li = -2.8) [24].
Sodium. There are many different, successful approaches to Na+-binding by calixarenes and only a few can be men-tioned here. In calix[4]arenes bearing carboxymethoxy-groups such as in ligand 5 the carbonyl and ether oxygens form a cavity with a size suitable for Na+. Earlier reviews [25,26,27] deal with these structures and a good example is 5a with log KNa,M = -3.1 (K+), -4.6 (Li+), -5.1 (Rb+), -5.8 (Cs+) (SSM) [28,29]. Covalently attaching calixarenes to poly-siloxane in the membrane further improves the long-term stability while retaining the selectivity (log KNa,M = -2.6 (K), -3.5 (Li), -3.8 (Mg), -3.4 (Ca), FIM) [30,31]. Copolymerizing Na+-selective, ester and methacrylamide groups containing calixarenes also increases the lifetime without a negative impact on selectivity compared with the monomer (log KNa,M = -2.4 (K), -3.4 (Ca), -4.3 (Mg)) [32].
Sensors 02 00397 i004
In recent years we observe the development of optical sensors and the current goal is to improve their properties so that they can compete with potentiometric sensors. A straightforward way is the mixing of a neutral ligand and an ionizable dye in a membrane which responds to cations with the release of protons. Sodium ions can be detected selectively by ligand 5d as change of the 660 nm absorption of ETH 5294 (log KNa,K= -3.06; Li, NH4 and group II: no response) [33,34] under pH-controlled condi-tions, or by 5a-c and various other ligands.
Calixarenes with azophenol [24] or indoaniline [35] as chromogenic groups or acylmethoxy-naphtalene [36] as fluorogenic group as part of the macrocyclic ring transform the binding of Na+ to the calixarene into an optical signal, such as a 42 nm bathochromic shift in case of ligand 6 [35].
The fluorescence emission of pyrene is quenched by an adjacent nitrophenol in 7, but the emission intensifies 8 times when Na+ is complexed, which widens the distance between the carbonyl oxygens and separates both groups (log Kass for MSCN = 4.3 (Na), 2.9 (K), 2.2 (Cs), 1.9 (NH4), 1.2 (Li), Et2O:MeCN 97:3) [37].
Various other mechanisms involving fluorescence were applied to calix[4]arenes for sodium sens-ing, e.g.:
-
Intramolecular energy transfer from a pyrene donor to an anthroyloxy acceptor pair (SfNa,K = 59, MeOH:THF) [38];
-
the intersystem crossing from the lowest excited singlet to the triplet state in acyl-methoxynaphtalene groups, which is prevented by Na+ binding to the acyl group because the energy level of the nπ* excited state increases [36];
-
the decrease of fluorescence emission in 5e, when Na+ is bound [39,40]; or
-
the increase of monomer:excimer emission ratio (30 times) in ligand 8 due to a larger distance upon complexation (Na/K selectivity = 154) [41].
Crowned calixarenes are characterized by a high degree of molecular preorganization and there-fore achieve even higher selectivities, especially among alkali ions. The crown moiety and the calixarene restrict each others molecular flexibility, leading to a better discrimination by ion size. Examples are 9a (log KNa,M = -5.0 (K+), -2.8 (Li+), -4.8 (Rb+), -4.4 (NH4 +), -4.5 (Mg2+), -4.4 (Ca2+), -5.4 (H+), FIM) [42,43] and a similar ligand [44] with a narrow cavity for maximum Na/K-selectivity. Mixing it with fluorophore ETH 5294 renders 9a to become an optode (detection limit 10-7 M Na+, log KNa,M <-3.06 (K), <-2.4 (Ca), <-3.9 (Mg), response time 10s) which can be miniaturized for intracellular measure-ments [45].
Sensors 02 00397 i005a Sensors 02 00397 i005b
Another mechanism for fluorescence detection applied in 9b, where the phenyl groups contribute to the selectivity, is reduced molecu-lar motion upon Na+ binding, increasing the emission intensity [46]. On the other hand, the monomer/excimer ratio increases in case of pyrene-derivative 9c (log Kass = 5.48(Na+), 4.84(Li+), < detection limit(K+), MeCN, MClO4) [47]. This crowned calixarene skeleton can be converted into a chromoionophore as demonstrated for the ligand 9d (Na/K-selectivity 103.1 in extraction) [48].
Sensors 02 00397 i006
The fluorescence emission wavelength of a conjugated poly(phenylene bithiophene) with calix[4]arene-crown-4 moieties as selective binding sites switches by 50 nm above 1 mM Na+ [49].
Potassium. The same idea of a crowned calix[4]arene can be applied for K+ recognition. For example, ligand 10 is a t-butylcalix[4]arene bearing a crown-5 group and has a K/Na selectivity of 103.1 in chemically modified field effect transistors [50].
Sensors 02 00397 i007
In combination with a nitrophenolazophenol moiety , the extraction of K+ by such a crowned calixarene can be observed photometrically (selectivity as log(Kex,K/Kex,Na) = 3) [51].
However, these are cone (or distorted cone) conformers, where the four phenyl rings look into the same direction, so the ligating groups are on one side of the molecule and the hydrophobic groups form the other side. Chang-ing the conformation to partial cone (one phenyl ring inverted) or alternating (phenyl rings look up and down, alternating), the ligand becomes even better pre-organized for binding K+. Alkyl groups larger than methyl are required as substituents at the phenolic oxygens, otherwise the high selectivity (log KK,Na = -3.7, SSM) [52] decreases with time due to conformational isomerism. In addition, removal of sterically bulky t-butyl groups improves the selectivity and the highly selective ligands 11a, b [53] were characterized by various meth-ods (e.g. 11a, log(Kex,K/Kex,Na) = 5.5, CHCl3, picrate extr.; log KK,M = -4.4 (Na), -4 (Mg), -3.9 (Ca), FIM) [31,53,54].
Sensors 02 00397 i008
The same alternate conformer but with azacrown-5 moieties, ligand 12 and its monocrown relatives show K+-selectivity in transport as well as in ISEs, which translates into an optical signal with the help of the nitrophenol chromophoric group [55,56,57].
Sensors 02 00397 i009
Rubidium and Caesium. There are two main approaches for Rb+and Cs+ recognition by calixarenes. The first one applies calix[6]- or calix[8]-arenes bearing mono- or bidentate groups. Examples are ligand 13 (log KCs,M = -3.9 (Na), -2.7 (K), -1.9 (Rb), -2.8 (NH4), SSM) and related structures [58,59,60,61], with improved selectivity at the expense of hydrophobicity upon re-moval of the p-t-butyl groups [62].
Sensors 02 00397 i010
The second approach utilizes crowned calixarenes such as 14 in 1,3-alternate conformation [63,64,65,66,67,68,69] which provide high Cs/K-selectivity (e.g. log KCs,M = -4.5 (Na), -2.2 (K)). The selectivity is still reasonably high after covalent linking to polymeric membranes in CHEMFETs (log KCs,M = -3.2 (Na), -2.3 (K), -4 (Ca), -2 (NH4), FIM) [31]. The absence of t-butyl groups is important for a good Cs+-selectivity because it enables cation-π-interaction. This type of interaction is also responsible for a strong decrease of 1π π* fluorescence emission of 14 upon complexation with Cs+ [70].
Sensors 02 00397 i011
Attaching benzo or naphto groups to the crown moiety increases the Cs/Na-selectivity [71]. The Cs/K-selectivity on the other hand im-proves to over 4000 after removal of the two phenolic oxygens outside the crown cavity [72].
Photo-induced electron transfer (PET) is a common mechanism in nature which can also be ap-plied for sensitive detection such as in ligand 15a (Φ=0.03(Cs+ added), Φ=0.003(no M+), log(βCs/βK) = 1.3, detection limit 10-5 M, CH2Cl2:MeOH) and the related biscrown-calix[4]arene 15b [73,74,75]. The crown moiety is rigidified by the calixarene and contributes to the Cs+-selectivity, while PET from benzocrown oxygens or aza-crown nitrogens to the excited singlet state of anthracene or cyanoanthracene groups quenches the fluorescence emission in the uncomplexed state. Complexation triggers an in-crease in fluorescence emission. This can be detected with more accuracy and sensitivity compared to a fluorescence ‘switch-off’ mechanism.
Sensors 02 00397 i012
Instead of calix[4]arene, the larger calix[5]-arene can be used as molecular platform. Then, the smaller crown-5 ring is suitable for Cs+-binding. With azophe-nol as chromophoric groups it re-sponds selectively to Cs+ with a new absorption at 475 nm due to deprotonation (log KCs,M = -2.2 (K), -0.8 (Rb), Li and Na no effect, detection limit 10-5 M, methoxyethanol/H2O 9:1) [76].
Doubly crowned calix[4]arene such as 16 avoid the need for protective groups during syn-thesis and can bind two metal ions in one molecule. The selectivity of 14 and similar ligands having aromatic or aliphatic groups on the crown moiety towards Cs+ was shown in extraction and transport experiments [77,78,79,80,81,82,83,84,85,86,87] and in ISEs (log KCs,M = -4.88 (Na), -2.2 (K), -1.94 (NH4), -0.9 (Rb), <-5 (group II), FIM, lin. range 0.1-10-6 M) [88].
An optode composed of 17 and chromophore ETH5294 is sta-ble and selective (log KCs,M = -2.1(Na), -2.3(K), -2.5(Ca), -3.2(Li), -2.8(Mg), SSM, lin. range 10-2 - 10-6 M), but has a long response time of 5 min [89].
Sensors 02 00397 i013

Alkaline Earth Ion Recognition

Among ester, acid or amide groups anchored to calixarenes, the latter turned out to form most stable complexes with alkaline earth ions due to the high carbonyl group polarity, e.g. [62,90]. A good selectiv- ity for Ca2+ over Mg2+ is achieved with tertiary amides such as 18a in electrodes (log KCa,Mg = -1.9, FIM) [91] and with 18b in complexation (log(KCa/KMg) > 7.8, MeOH) [92,93]. Ca and Sr cannot be discriminated by these ligands due to the enthalpy-entropy compensation. As far as complexation is concerned, the carboxylic acid groups in 18c provide also a good Ca/Mg-selectiv-ity (log K = 22.44 (Ca), 20.92 (Sr), 17.96 (Ba), 11.02 (Mg), 9.94 (Na), 9.05 (K), MeOH) [94].
Sensors 02 00397 i014
With phosphine oxide groups appended to the lower rim in 18d, good Ca2+/Na+-selectivity and membrane durability (>7 weeks) is achieved in ISEs (log KCa,M = -2.2 (Na), -2.7 (K), -2 (NH4), -1.6 (Li), -2.6 (Mg), SSM) [95]. In an optode comprising a mixture of 18d and the acidochromic dye ETH 5294 even higher selectivity coefficients were reported (log KCa,M = -4.77 (Na), -3.84 (K), -3.65 (NH4), -2.3 (Li), -1.4 (Mg), -1.1 (Ba)) [33,34].
Several papers were published about chromoionophoric calix-arenes for Ca2+. The ligand 19 [96] for example contains indoaniline groups the polar quinone carbonyl oxygens of which coordinate diva-lent cations well. The large bathochromic shift ( 110 nm) and the higher association constants log(Kass,Ca/Kass,Mg) = 4.9, log(Kass,Ca/Kass,K) = 1.4, K+: 28 nm shift, EtOH) reflect the Ca-selectivity, which is a result of the size matching between host and guest. Even a logical switch using Ca2+ and K+ has been designed based on ditopic recogni-tion [97].
Sensors 02 00397 i015
Another chromoionophore, ligand 20, utilizes azophenol groups and works well at neutral pH. Upon binding of Ca2+ an 168 nm bathochromic shift is induced as deter-mined by extraction into CHCl3 (detection limit 10-4 M, logKCa,M = -2.3(Sr), -2.9(Mg), -2.8(Na)) [98]. Sr2+ can be detected by ISEs containing underivatized t-butylcalix[8]arene (logKSr,M = -1.2 (Ca), -1.3(Ba), -1.4 (Mg, Li), -1.5 (K), -0.3 (Na), -1.3 (Pb), FIM, linear range 2.10-5 - 0.1 M) [99].
Sensors 02 00397 i016
Expanding the calix[n]arene cavity from n= 4 over 5 to 6 changes the Sr/Na-selectivity of the amide derivatives from 0.09 over 2.8 to 760 in extraction (DSr/DNa), which further improves with alkylether instead of t-butyl groups in 4-position of the phenyl rings [100].
Amide derivatives of calixarenes have also been used for Sr2+-separation in synergistic mixtures with hydrophobic anions [101,102], which should work in ISE membranes as well.
Calix[4]diquinone-based ionophores were proposed for amperometric sensors for Ba2+ due to the selectivity over alkali and ammonium complexes as shown by the largest anodic shift in voltammetry [103,104].

Ammonium, Guanidium Ion and Amino Acid Recognition

Ammonium ions can be included in the cavity of calixarenes and discriminated by size and steric rea-sons. A phenolic OH-group can transfer a proton [105] to primary, secondary or tertiary amines and complexation occurs due to electrostatic, dispersive, H-bonding, CH3-π and for aromatic amines also π-π interaction. N+-π-interactions were proven to take part as well, e.g. [106,107,108,109]. Alternatively, deprotonated calixarenes can bind quaternary amines [110].
The phenolic OH-groups in the water-soluble calix[6]arene with p-sulphonato instead of t-butyl groups were employed for fluorimetric sensing [111]: The ligand is selective for acetylcholine over primary and secondary amines and was mixed with an auxiliary fluorescent cation which competes for binding. Binding of acetylcholine releases the fluorophore, the emission of which is quenched only in the complexed state by the phenolate anions (detection limit 10-6 M acetylcholine, H2O/MeOH 1:1).
Hydrophobic ether derivatives such as ligand 21 discriminate between structural isomers such as butylammonium ions in ISEs (log Kn,x = -1.20 (i), -1.31 (s), -1.65 (t), x-BuNH3 +, FIM) and in their complex stabilities [112].
Sensors 02 00397 i017
Ester derivatives such as 22, the synthesis of which can be accomplished smoothly, bind amines in the order n- > s-, i- > t-butylamine and have a preference for shorter amines [113,114,115,116]. The ligand discriminates according to guest hydropho-bicity and shows selectivity for Phe- and Typ-esters over Gly, Ala and 4-aminobu-tyric acid. The analogous decylester deriva-tive in ISEs responds to primary amines due to tripodal NH3 +…O=C hydrogen bonding. Selectivity is observed for amines without bulky substituents in α-position, such as to dopamine (log K(dopamin/adrenalin) = -1.55, SSM) [117].
Sensors 02 00397 i018
The more hydrophobic and larger p-adamantylcalix[8]arene is characterized by higher sensitivity (0.1 μg/ml L-Phe compared with 2.8 μg/ml (22)) and complex stability [118]. Steric effects are dis-cussed to cause the selectivity for L-Leu over L-Phe ester [119,120].
Ligand 23a is a stronger H-bond acceptor as compared with 22 and binds guanidinium as shown in transport [121] and at interfaces [122]. The suitable cavity size and binding geometry contribute to the selectivity over alkali ions. Having excessive binding sites removed while retaining C3-symmetry, 23b has a higher selectivity in CHEMFETs (log Kgu,M = -1.8 (K), -1.85 (Na), -1.8 (NH4), -2.8(Ca), FIM) [123] and in transport across liquid membranes [124]. Another approach towards favourite gu+-binding (log Kass = 3.4) utilizes a ‘cap’ for rigidifying calix[6]arenes [125] or homotrioxacalix[3]arene [126], although in the first case it slows down the dissociation rate. The calix[6]arene and the homotrioxacalix[3]arene were also used as molecular backbones to anchor fluorophoric groups such as in 24, where the binding of gu+ changes the monomer-excimer intensity ratio [127,128].
Sensors 02 00397 i019
Chirality can be introduced into calix[4]arene by 3 or more different substituents such as in 25 [129], where the excimer emission intensifies upon amino acid binding.
Combined hydrogen bond donors and acceptors, such as α-aminophosphonate groups held by the calix[4]arene molecular platform allow the binding of hydrophilic amino acid zwitter ions with a selectiv-ity order His > Phe, Tyr, Typ, which can slightly be influenced by the substituent in para-position [130,131].
Sensors 02 00397 i020
Ammonium and alkylammonium ions can be sensed by calixarenes where diquinone is part of the macrocycle together with ligating sites such as ester or amide for hydrogen bonding [132,133].
Crowned Calixarenes for Ammonium Recognition. Crowned calix[4]arenes retain the capability of hydrogen bonding to guests while discriminating better according to steric reasons such as bulkiness due to the rigidified cavity. Examples are the preference for n-butyl over tert-butylamines in transport and extraction [134,135] and the selectivity for primary over tertiary ones [136]. The crown also contributes to the complex stability [137] as it forms a more closed cavity shielding the ion from its hydration shell. Modifying the crown moiety with hydrophobic groups and attaching it to the ‘upper rim’ rather than to the phenolic oxygens also improves the selectivity for primary ammonium over alkali ions [138,139].
With inherently chiral, crowned calix[4]arenes the chiral rec-ognition of ammonium ions may be possible [140]. An interesting example for the design of a chiral host is seen in the structures 26a and b. The symmetrical ligand discriminates among butylamines due to steric and electrostatic reasons of the binaphtyl hinge (log Kass = 3.0(t-), 2.8 (s-), 2.5 (i-), 2 (n-butylamine), EtOH) 141]. Compared with ligand 22 before, this order is nearly reversed. R-isomers of amines and amino acids are preferably bound by the unsymmetrical ligand 26b [142,143]. In both cases sensing is easily accomplished by a bathochromic shift (red to blue colour change) and a new ab-sorption, mainly due to proton transfer.
Sensors 02 00397 i021
Crowned calixarene 27 interacts with amines primarily by means of H-bonding, while bulkiness and hydrophobicity also contribute to the observed selectivity n- >> s-> t-) [144]. The proton transfer is sensed as a 93 nm bathochromic shift.
Sensors 02 00397 i022
Host-guest proton transfer also cause the colour change from yellow to orange (monoamines) or to red (di-, triamines) in bisazobiphenyl-bridged calix[n]arenes (n = 4, 8) [145,146,147]. However, the mechanism of proton transfer cannot be applied to aromatic amines due to their lower basicity.

Silver Ion recognition

Although good Ag+-sensitive electrodes are well-established, also calixarenes were investigated for this purpose. One reason is the specific interaction between the π-clouds and ‘soft’ cations [148,149,150] which leads to selective complexation without the need for other donor groups.
Allyl ethers of calix[4]arene utilize the same type of interaction, this time between the allyl and phenyl π-clouds and Ag+ [151]. The selectivity over ‘hard’ cations is excellent (log KAg,M = -3.3 (H), -3.6 (Pb), -1.8 (Hg), -0.6 (Tl), <-4 (group I+II, NH4), FIM, lin. range 10-4 - 10-1 M Ag+). The selectivity was also confirmed by extraction using various allyl and benzyl ether derivatives [152].
The second approach to Ag+-selectivity consists of the introduction of ‘soft’ donor atoms such as sulfur and nitrogen. As an example, thioether 28 was used in ISEs (log KAg,M = -2.5 (Hg), -4.7 (K, Pb), -5 (Na, Ni), < -5 (group II, NH4, Cu, Cd, Zn), FIM, response time 10 s, linear range 10-4.5 M Ag+) [153,154] as well as in CHEMFETs with a polysiloxane matrix (log KAg,M = -4.3 (Ca), -4.4 (Cu), -4.7 (K), -2.5 (H), -2.4 (Hg), -4 (Cd), FIM, lin. range above 10-4.2 M) [155]. The membrane conditioning can further increase the selectivity [156].
Sensors 02 00397 i023
Thioester groups in calix[4]arenes such as 29 lower the selectivity over sodium ions to some extent due to their carbonyl oxygens (log KAg,Na = -1.2 in ISE, lin. range 10-4 to 10-1 M) [157,158].
Thiocarbamoyl groups [159,160] or dithiocarbamoyl groups when attached to the calixarene skel-eton such as in ligand 30 [161,162] provide selectivity for Ag+, Pd2+, Hg2+ and other ‘soft’ metal ions over ‘hard’ ones.
Alternatively, nitrogen as donor atom for Ag+-complexation was successfully applied to avoid interference by alkali and ammonium ions. The pyridinophane 31 resembles a calixarene, selected log KAg,M values are -2.7 (K), -3.1(Pb), -3.4 (Cd) (SSM) [163].
Sensors 02 00397 i024
The Ag+-complex of 28 acts as reversible electron-transfer agent and by this way stabilizes the phase boundary potential in solid-state sen-sors for polyanions [164]. Thioether-substituted calixarenes self-assem-ble on surfaces of Au or Ag, allowing mono- or multilayered structures with recognition sites such as the molecular cavity.

Other Heavy Metal Ion Recognition

Although many calixarene derivatives were characterized by solvent extraction and membrane transport with respect to their heavy metal selectivity [1,7,165,166], limited data are available for sensors. Examples are the determination of Hg2+, Pb2+ and Cu2+ by anodic stripping voltammetry [167] and with CHEMFETs [153] utilizing thioamide groups on a calixarene skeleton. For Pb2+, thioamides such as ligand 32a was investigated in CHEMFETs (log KPb,M =-4.2 (Cd), -3.4 (Cu), -5.2 (K), -4.3 (Ca), detec-tion limit 10-6M, FIM) [153], in CHEMFETs with polysiloxane membranes which avoid the need for a plasticizer (log KPb,M = -3.3 to -3.7 (K), -3.0 to -3.3 (Cu), -3.8 (Cd), -4.2 (Ca), FIM) and linear range (above 10-5M) [31,155,168], and in ISEs (log KPb,M < -3, pH-range 3-6) [169].
Similar ligands were applied in anodic stripping voltammetry for Pb2+-determination [167,170]. The thiol 32b, also based on the calix[4]arene frame, was successfully used for analyzing Pb2+ by the same method using screen-printed carbon electrodes. The detection limit of 5 ng per ml after preconcentration is sufficient for water quality control [171].
Sensors 02 00397 i025
Thioamides based on the resorc[4]arene skeleton instead work in CHEMFETs for Pb2+, although with reduced selectivities compared with some of the above-mentioned calixarenes (log KPb,M = -1.9 (K), -2.1 (Cu), -2.4 (Cd), -4.3 (Ca), FIM) [172].
The selectivity of thiophosphate derivatized calix[6]arenes (log KPb,M = -2.65 (Cd), -2.4 (Cu), -1.4 (Ca), sensitivity (10-4 M Pb, FIM) depend also on the ligand’s substitution pattern [173]. A wider linear range (10-5M Pb2+) was demonstrated with ethyleneoxydiphenyl phosphine oxide groups, the interfer-ence by Ca2+ can be avoided by expanding the calixarene ring size [174].
Calixarene 33 contains two thioamide groups and was used for the preparation of Cd-selective CHEMFETs (log KCd,M = -3.2 (Ca), -0.6 (Cu), >0 (K, Pb), FIM) [153]. Adding two more amide groups improves the selectivity over K+ at the expense of Cu2+ (log KCd,K = -2.6, detection limit 10-5.5 M Cd). With polysiloxane membranes, the interference by Pb2+ could be avoided (log KCd,M = -4.1 (Ca), -2.6 (K), -2.0 (Pb), >0 (Cu), FIM) [155]. To achieve selectivity over Pb2+ and Cu2+ in PVC or polysiloxane- based membranes, the ligand 34 with an alternating conformation was developed (log KCd,M = -0.7 (Pb), -4.2 (Ca), -3.0 (K), -1.6 (Cu), lin. range 10-5 M, Nernstian slope, polysiloxane, FIM) [168].
The dansyl-groups render 35 a fluorophore which responds to Cd2+ with increased emission inten-sity when the dansyl groups are inside the hydrophobic cavity (ΔI/I0 = 0.42, Cd, 0.23(Al3+), 0.2(Cr3+), 0.17(Zn2+, Cu2+), <0 (for alkali, Co, Ni ions), H2O, 489nm, 3x10-4M metal, pH 7) [175].
1,2-Diquinone containing calixarenes were proposed for the voltammetric detection of transition metal ions due to their interaction with them as exemplified by Ag+ [176].
Dithioamide groups on the calix[4]arene backbone, represented by ligands 36a and b were applied for Cu2+-selective CHEMFETs, although an interference by K+ occurs (log KCu,M = -1.9 (Ca), -2.1 (Cd), -1.7 (Pb), >0 (K), linear range to 10-4 M Cu2+, FIM, 36a) [153].
From N-(X)sulfonyl carboxamide derivatized calix[4]-arene, which is selective for Hg2+ in extraction [177] the fluorophore 37 was derived, the 520 nm emission of which is quenched upon complexation due to electron transfer [178].
Sensors 02 00397 i026
Sensors 02 00397 i027
Utilizing the cation-π-interaction of the phenyl rings, Tl+ could be selectively sensed by the propyl ether of de-t-butylated calix[4]arene in an ISE with log KTl,M = -1.35 (Ag), -2.7 (Hg), <-4.0 (K, NH4, H, Pb), <-5 (Ca, Na, Mg), lin. range 0.1 - 3.10-6 M, FIM) [179].
Calix[4]arenes can be bridged with a dioxotetraaza group hav-ing an anthracene group appended [180]. The complexation is moni-tored by means of the fluorescence emission, depending on whether quenching (Ni2+) or emission-enhancing cations (Zn2+) are present. f-Elements.For f-element ions, a large amount of data has been accu-mulated in the area of extraction and transport, and rather selective ligands were developed [1]. These molecular designs are based on calixarenes bearing (i) bidentate groups such as carbamoylphosphine oxide, (ii) monodentate groups such as phosphine oxide, esters of phosphoric, phosphonic or phosphinic acids, or (iii) different monodentate groups such as acid and amide within the same molecule.
Of special interest are the luminescent properties of lanthanide ions encapsulated in calixarenes, because they are shielded from the hydration shell by the macrocyclic cavity. If the energy levels match each other, energy transfer from excited ligand energy state to a lanthanide ion energy level or between two different lanthanide ions occurs with high efficiency. FIAs could be based on this approach. Data on luminescence properties of Tb3+ and Eu3+ [181,182,183,184,185,186,187,188,189,190,191,192,193] and also other lanthanides [194,195,196,197,198,199]with quantum yields of up to 30% show that the design of the functional group acting as molecular ‘antenna’ to absorb the energy is very important.
The pentahydroxycalix[6]arene acts as a pseudoplanar pentadentate ligand, for UO2 2+ and with the indoaniline group introduced responds selectively with a bathochromic shift (59 nm) in the presence of a base [200].

References

  1. Asfari, Z.; Böhmer, V.; Harrowfield, J.; Vicens, J. (Eds.) Calixarenes 2001; Kluwer Acad. Publ.: Dordrecht, 2001.
  2. Lumetta, G. J.; Rogers, R. D. (Eds.) Calixarene Molecules for Separations; American Chemical Society: Wash-ington DC, 1999.
  3. Gutsche, C. D. Calixarenes Revisited; The Royal Society of Chemistry: Cambridge, UK, 1998; Vol. 6. [Google Scholar]
  4. Gokel, G. W. (Ed.) Molecular Recognition: Receptors for Cationic Guests, 1 ed.; Pergamon Press: New York, Oxford, 1996; Vol. 1.
  5. Vicens, J.; Asfari, Z.; Harrowfield, J. M. (Eds.) Calixarenes 50th Anniversary: Commemorative Issue; Kluwer Academic Publ.: Dordrecht, Holland, 1994.
  6. Vicens, J.; Böhmer, V. (Eds.) Calixarenes. A Versatile Class of Macrocyclic Compounds; Kluwer Academic Publ.: Dortrecht/Boston/London, 1991.
  7. Ludwig, R. Calixarenes in analytical and separation chemistry. Fresenius’ J. Analyt. Chem. 2000, 367, 103–128. [Google Scholar] [CrossRef]
  8. Diamond, D.; Nolan, K. Calixarenes: Designer Ligands for Chemical Sensors. Analyt. Chem. 2001, 22 A–29 A. [Google Scholar]
  9. Arnaud-Neu, F.; Schwing-Weill, M.-J. Calixarenes, new selective molecular receptors. Synthetic Metals 1997, 90, 157–164. [Google Scholar] [CrossRef]
  10. Roundhill, D. M. Progr. Inorg. Chem.; Karlin, K. D., Ed.; Wiley: New York, 1995; Vol. 43, pp. 533–591. [Google Scholar]
  11. Ludwig, R. Review on calixarene-type macrocycles and metal extraction data. JAERI-Review 95-022. JAERI: Japan, 1995. [Google Scholar]
  12. Gutsche, C. D. Calixarenes; The Royal Society of Chemistry: Cambridge, UK, 1989; Vol. 1. [Google Scholar]
  13. Abidi, R.; Baker, M. V.; Harrowfield, J. M.; Ho, D. S.-C.; Richmond, W. R.; Skelton, B. W.; White, A. H.; Varnek, A.; Wipff, G. Complexation of the p-t-butyl-calix[4]arene anion with alkali metal cations in polar, non-aqueous solvents: Experimental and theoretical studies. Inorg. Chim. Acta 1996, 246, 275–286. [Google Scholar] [CrossRef]
  14. Nakamoto, Y.; Nakayama, T.; Yamagishi, T.; Ishida, S. Abstr. Workshop on Calixarenes and Related Compounds, Mainz (Ger.), 28-30 Aug. 1991; 1.
  15. Shimizu, H.; Iwamoto, K.; Fujimoto, K.; Shinkai, S. Chromogenic calix[4]arene. Chem. Letters 1991, 2147–2150, J. Chem. Soc., Perkin Trans I 1992, 1885-1887. [Google Scholar] [CrossRef]
  16. McCarrick, M.; Wu, B.; Harris, S. J.; Diamond, D.; Barrett, G.; McKervey, M. A. Novel chromogenic ligands for lithium and sodium based on calix[4]arene tetraesters. J. Chem. Soc., Chem. Commun. 1992, 1287–1289. [Google Scholar] [CrossRef]
  17. McCarrick, M.; Harris, S. J.; Diamond, D.; Barret, G.; McKervey, M. A. Assessment of three azophenol calix[4]arenes as chromogenic ligands for optical detection od alkali metal ions. Analyst 1993, 118, 1127–1130. [Google Scholar] [CrossRef]
  18. McCarrick, M.; Wu, B.; Harris, S. J.; Diamond, D.; Barrett, G.; McKervey, M. A. Chromogenic ligands for lithium based on calix[4]arene tetraesters bearing nitrophenol residues. J. Chem. Soc., Perkin Trans. II 1993, 1963–1968. [Google Scholar] [CrossRef]
  19. Harris, S. J.; Diamond, D.; McKervey, M. A. IR Pat. Appl. S922577 (1992), 1997.
  20. Grady, R.; Butler, T.; MacCraith, B. D.; Diamond, D.; McKervey, M. A. Optical sensor for gaseous ammonia with tunable sensitivity. Analyst 1997, 122, 803–806. [Google Scholar] [CrossRef]
  21. McCarrick, M.; Harris, S. J.; Diamond, D. Assessment of a chromogenic calix[4]arene for the rapid colorimetric detection of trimethylamine. J. Mater. Chem. 1994, 4, 217–221. [Google Scholar] [CrossRef]
  22. Loughran, M.; Diamond, D. Monitoring of volatile bases in fish sample headspace using an acidochromic dye. Food Chemistry 2000, 69, 97–103. [Google Scholar] [CrossRef]
  23. de Namor, A. F. D.; Hutcherson, R. G.; Sueros Velarde, F. J.; Zapata-Ormachea, M. L.; Salazar, L. E. P.; Al Jammaz, I.; Al Rawi, N. An overview on the solution thermodynamics of lower rim functionalised calixarene derivatives and metal cations. New derivatives containing amino and thioalkyl functional groups. Pure & Appl. Chem. 1998, 79, 769–778. [Google Scholar]
  24. Toth, K.; Lan, B. T. T.; Jeney, J.; Horvath, M.; Bitter, I.; Grün, A.; Agai, B.; Töke, L. Chromogenic calix[4]arene as ionophore for potentiometric and optical sensors. Talanta 1994, 41, 1041–1049. [Google Scholar] [CrossRef]
  25. Diamond, D. Calixarene-based sensing agents. J. Inclusion Phen. and Molec. Recognition in Chem. 1994, 19, 149–166. [Google Scholar] [CrossRef]
  26. O’Connor, K. M.; Cherry, M.; Svehla, G.; Harris, S. J.; McKervey, M. A. Symmetrical, unsymmetrical and bridged calix[4]arene derivatives as neutral carrier ionophores in PVC membrane sodium selective electrodes. Talanta 1994, 41, 1207–1217. [Google Scholar] [CrossRef]
  27. Bühlmann, P.; Pretsch, E.; Bakker, E. Carrier-based ion-selective electrodes and bulk optodes. 2. Ionophores for potentiometric and optical sensors. Chem. Rev. 1998, 98, 1593–1687. [Google Scholar] [CrossRef] [PubMed]
  28. Sakaki, T.; Harada, T.; Deng, G.; Kawabata, H.; Kawahara, Y.; Shinkai, S. Molecular design of calixarene-based ion-selective electrodes. J. Inclusion Phen. and Molec. Recognition in Chem. 1992, 14, 285–302. [Google Scholar] [CrossRef]
  29. Sakaki, T.; Harada, T.; Kawahara, Y.; Shinkai, S. On the selection of the optimal plasticizer for calix[n]arene-based ion-selective electrodes: Possible correlation between the ion selectivity and the “softness” of the plasticizer. J. Inclusion Phen. and Molec. Recognition in Chem. 1994, 17, 377–392. [Google Scholar] [CrossRef]
  30. Reinhoudt, D. N. Sens. Actuators B 1995, 24-25, 197. [CrossRef]
  31. Lugtenberg, R. J. W.; Egberink, R. J. M.; ven den Berg, A.; Engbersen, J. F. J.; Reinhoudt, D. N. The effects of covalent binding of the electroactive components in durable CHEMFET membranes - impedance spectroscopy and ion sensitivity studies. J. Electroanalyt. Chem. 1998, 452, 69–86. [Google Scholar] [CrossRef]
  32. Parzuchowski, P.; Malinowska, E.; Rokicki, G.; Brzozka, Z.; Böhmer, V.; Arnaud-Neu, F.; Souley, B. Calix[4] arene derived tetraester receptors modified at their wide rim by polymerizable groups. New J. Chem. 1999, 23, 757–763. [Google Scholar] [CrossRef]
  33. O’Neil, S.; Kane, P.; Diamond, D. Comparison of the performance of calix[4]arene phosphine oxide and ester derivatives in ion-selective optode membranes. Analyt. Commun. 1998, 35, 127–131. [Google Scholar] [CrossRef]
  34. O’Neill, S.; Conway, S.; Twellmeyer, J.; Egan, O.; Nolan, K.; Diamond, D. Ion-selective optode membranes using 9-(4-diethylamino-2-octadecanoatestyryl)-acridine acidichromic dye. Analyt. Chimica Acta 1999, 398, 1–11. [Google Scholar] [CrossRef]
  35. Kubo, Y.; Hamaguchi, S.-I.; Kotani, K.; Yoshida, K. New chromoionophores based on indoaniline dyes containing calix[4]arene. Tetr. Lett. 1991, 32, 7419–7420. [Google Scholar] [CrossRef]
  36. Leray, I.; O’Reilly, F.; Jiwan, J. L. H.; Soumillion, J. P.; Valeur, B. A new calix[4]arene-based fluorescent sensor for sodium ion. J. Chem. Soc., Chem. Commun. 1999, 795–796. [Google Scholar] [CrossRef]
  37. Aoki, I.; Sakaki, T.; Shinkai, S. A new metal sensory system based on intramolecular fluorescence quenching on the ionophoric calix[4]arene ring. J. Chem. Soc., Chem. Commun. 1992, 730–732. [Google Scholar] [CrossRef]
  38. Jin, T. A new Na+-sensor based on intramolecular fluorescence energy transfer derived from calix[4]arene. J. Chem. Soc., Chem. Commun. 1999, 2491–2492. [Google Scholar] [CrossRef]
  39. Perez-Jimenez, C.; Harris, S. J.; Diamond, D. A novel calix[4]arene tetraester with fluorescent response to complexation with alkali metal cations. J. Chem. Soc., Chem. Commun. 1993, 480–483. [Google Scholar] [CrossRef]
  40. Perez-Jiminez, C.; Harris, S. J.; Diamond, D. New fluoroionophores for alkali-metal cations based on tetrameric calixarenes. J. Mater. Chem. 1994, 4, 145–151. [Google Scholar] [CrossRef]
  41. Jin, T.; Ichikawa, K.; Koyama, T. A fluorescent calix[4]arene as an intramolecular excimer-forming Na+ sensor in nonaqueous solution. J. Chem. Soc., Chem. Commun. 1992, 499–501. [Google Scholar] [CrossRef]
  42. Yamamoto, H.; Shinkai, S. Molecular design of calix[4]arene-based sodium-selective electrodes which show re-markably high 10^5.0 - 10^5.3 sodium/potassium selectivity. Chem. Letters 1994, 1115–1118. [Google Scholar] [CrossRef]
  43. Yamamoto, H. JP Appl. H6-23700. JRDC, 1994. [Google Scholar]
  44. Yamamoto, H.; Ueda, K.; Suenaga, H.; Sakaki, T.; Shinkai, S. Exploitation of Na+-selective electrodes for protein solutions from crown-bridged calix[4]quinones. Chem. Letters 1996, 39–40. [Google Scholar] [CrossRef]
  45. Shortreed, M.; Bakker, E.; Kopelman, R. Miniature sodium-selective ion-exchange optode with fluorescent pH chromoionophores and tunable dynamic range. Analyt. Chem. 1996, 68, 2656–2662. [Google Scholar] [CrossRef]
  46. Matsumoto, H.; Shinkai, S. Alkali metal sensing with a fluorescent 5,11,17,23-tetraphenylcalix[4]arene and the mechanism of the fluorescence change. Chem. Letters 1994, 2431–2434. [Google Scholar] [CrossRef]
  47. Matsumoto, H.; Shinkai, S. Metal-induced conformational change in pyrene-appended calix[4]crown-4 which is useful for metal sensing and guest tweezing. Tetr. Lett. 1996, 37, 77–80. [Google Scholar] [CrossRef]
  48. Yamamoto, H.; Ueda, K.; Sandanayaka, K.; Shinkai, S. Molecular design of chromogenic calix[4]crowns which show very high Na+ selectivity. Chem. Letters 1995, 497–498. [Google Scholar] [CrossRef]
  49. Crawford, K. B.; Goldfinger, M. B.; Swager, T. M. Na+ Specific Emission Changes in an Ionophoric Conjugated Polymer. J. Amer. Chem. Soc. 1998, 120, 5187–5192. [Google Scholar] [CrossRef]
  50. Sudholter, E. J. R.; vander Wal, P. D.; Skovronska-Ptasinska, M.; vander Berg, A.; Berveld, P.; Reinhoudt, D. N. Transuction of host-guest complexation into electronic signals: favoured complexation of potassium ions by syn-thetic macrocyclic polyethers using membrane-modified, ion-sensitive field-effect transistors. Recl.Trav.Chim.Pays- Bas 1990, 109, 222–225. [Google Scholar] [CrossRef]
  51. King, A. M.; Moore, C. P.; Sandanayake, K. S.; Sutherland, I. O. A highly selective chromoionophore for potassium based upon a bridged calix[4]arene. J. Chem. Soc., Chem. Commun. 1992, 582–584. [Google Scholar] [CrossRef]
  52. Brzozka, Z.; Lammerink, B.; Reinhoudt, D. N.; Ghidini, E.; Ungaro, R. Transduction of selective recognition by preorganized ionophores, K+ selectivity of the different 1,3-diethoxycalix[4]arene crown ether conformers. J. Chem. Soc., Perkin Trans. II 1993, 1037–1040. [Google Scholar] [CrossRef]
  53. Casnati, A.; Pochini, A.; Ungaro, R.; Bocchi, C.; Ugozzoli, F.; Egberink, R. J. M.; Struijk, H.; Lugtenberg, R.; de Jong, F.; Reinhoudt, D. N. 1,3-Alternate calix[4]arenecrown-5 conformers: New synthetic ionophores with better K+/Na+ selectivity than valinomycin. Chem. Europ. J. 1996, 2, 436–445. [Google Scholar] [CrossRef]
  54. Arnaud-Neu, F.; Deutsch, N.; Fanni, S.; Schwing-Weill, M. J.; Casnati, A.; Ungaro, R. Thermodynamic studies on alkali metal complexes of calix[4]-crowns-5: Influence of the conformation. Gazz. Chim. Ital. 1997, 127, 693–697. [Google Scholar]
  55. Kim, J. S.; Yu, I. Y.; Suh, A. H.; Ra, D. Y.; Kim, J. W. Novel calix[4]arene azacrown ether. Synth. Commun. 1998, 28, 2937–2944. [Google Scholar] [CrossRef]
  56. Kim, J. S. Abstr. 5th Int. Conf. Calixarene Chem.; Univ. Western Australia: Perth 19-23.9, 1999; p. L-24. [Google Scholar]
  57. Kim, J. S.; Shon, O. J.; Ko, J. W.; Cho, M. H.; Yu, I. Y.; Vicens, J. Synthesis and metal ion complexation studies of proton-ionizable calix[4]azacrown ethers in the 1,3-alternate conformation. J. Org. Chem. 2000, 65, 2386–2392. [Google Scholar] [CrossRef] [PubMed]
  58. Diamond, D. Anal. Chem. Symp. Ser. 1986, 25, 155.
  59. Telting-Diaz, M.; Regan, F.; Diamond, D.; Smyth, M. R. J. Pharm. Biomed. Anal. 1990, 115, 1207.
  60. Cadogan, A.; Diamond, D.; Smyth, M. R.; Svehla, G.; McKervey, M. A.; Seward, E. M.; Harris, S. J. Caesium-selective poly(vinyl chloride) membrane electrodes based on calix(6)arene esters. Analyst 1990, 115, 1207–1210. [Google Scholar] [CrossRef]
  61. Cunningham, K.; Svehla, G.; Harris, S. J.; McKervey, M. A. Analyst 1993, 111, 341.
  62. Fanni, S.; Arnaud-Neu, F.; McKervey, M. A.; Schwing-Weill, M. J.; Ziat, K. Dramatic effects of p-dealkylation on the binding abilities of p-tert-butylcalix[6]arenes: New Cs+ and Sr2+ selective receptors. Tetr. Lett. 1996, 37, 7975–7978. [Google Scholar] [CrossRef]
  63. Ungaro, R.; Casnati, A.; Ugozzoli, F.; Pochini, A.; Dozol, J.-F.; Hill, C.; Rouquette, H. 1,3-Dialkoxycalix[4]aren-kronen-6 in 1,3-alternate Konformation: Caesium-selektive Liganden, die Kation-Aren-Wechselwirkungen nutzen. Angew. Chemie 1994, 106, 1551–1553. [Google Scholar] [CrossRef]
  64. Dozol, J.-F.; Rouquette, H.; Ungaro, R.; Casnati, A. CA 122, 239730n (1995); 1994. [Google Scholar]
  65. Lugtenberg, R. J. W.; Brzozka, Z.; Casnati, A.; Ungaro, R.; Enbersen, J. F. J.; Reinhoudt, D. N. Cesium-selective chemically modified field effect transistors with calix[4]arene-crown-6 derivatives. Analyt. Chimica Acta 1995, 310, 263–267. [Google Scholar] [CrossRef]
  66. Bocchi, C.; Careri, M.; Casnati, A.; Mori, G. Selectivity of calix[4]arene-crown-6 for cesium ion in ISE: Effect of the conformation. Analyt. Chem. 1995, 67, 4234–4238. [Google Scholar] [CrossRef]
  67. Casnati, A.; Pochini, A.; Ungaro, R.; Ugozzoli, F.; Arnaud, F.; Fanni, S.; Schwing, M.-J.; Egberink, R. J. M.; de Jong, F.; Reinhoudt, D. N. Synthesis, Complexation, and membrane transport studies of 1,3-alternate calix[4]arene-crown-6 conformers: A new class of cesium selective ionophores. J. Amer. Chem. Soc. 1995, 117, 2767–2777. [Google Scholar] [CrossRef]
  68. Schwing-Weill, M.-J.; Arnaud-Neu, F. Calixarenes for radioactive waste management. Gazz. Chim. Ital. 1997, 127, 687–692. [Google Scholar]
  69. Haverlock, T. J.; Bonnesen, P. V.; Sachleben, R. A.; Moyer, B. A. Analysis of equilibria in the extraction of cesium nitrate by calix[4]arene-bis(t-octylbenzo-crown-6) in 1,2-dichloroethane. J. Inclusion Phen. and Molec. Recogni-tion in Chem. 2000, 36, 21–37. [Google Scholar] [CrossRef]
  70. Prodi, L.; Bolletta, F.; Montalti, M.; Casnati, A.; Sansone, F.; Ungaro, R. Photophysics of 1,3-alternate calix[4]arene-crowns and their metal ion complexes: evidence for cation-π interactions in solution. New J. Chem. 2000, 24, 155–158. [Google Scholar] [CrossRef]
  71. Kim, J. S.; Pang, J. H.; Yu, I. Y.; Lee, W. K.; Suh, I. H.; Kim, J. K.; Cho, M. H.; Kim, E. T.; Ra, D. Y. Calix[4]arene dibenzocrown ethers as cesium selective extractants. J. Chem. Soc., Perkin Trans. 2 1999, 837–846. [Google Scholar] [CrossRef]
  72. Sachleben, R. A.; Urvoas, A.; Bryan, J. C.; Haverlock, T. J.; Hay, B. P.; Moyer, B. A. Dideoxygenated calix[4]arene crown-6 ethers prefer the 1,3-alternate conformation and exhibit enhanced selectivity for cesium over potassium and rubidium. J. Chem. Soc., Chem. Commun. 1999, 1751–1752. [Google Scholar] [CrossRef]
  73. Ji, H.-F.; Brown, G. M.; Dabestani, R. Calix[4]arene-based Cs+ selective optical sensor. J. Chem. Soc., Chem. Commun. 1999, 609–610. [Google Scholar] [CrossRef]
  74. Ji, H.-F.; Brown, G. M.; Dabestani, R. Spacer length effect on the PET fluorescent probe for alkali metal ions. Photochem. Photobiol. 1999, 69, 513–516. [Google Scholar] [CrossRef]
  75. Ji, H.-F.; Dabestani, R.; Brown, G. M. Optical sensing of cesium using 1,3-alternate calix[4]mono and di(anthrylmethyl)aza-crown-6. Photochem. Photobiol. 1999, 70, 882–886. [Google Scholar] [CrossRef]
  76. Gordon, J. L. M.; Böhmer, V.; Vogt, W. A calixarene-based chromoionophore for the larger alkali metals. Tetr. Lett. 1995, 36, 2445–2448. [Google Scholar] [CrossRef]
  77. Asfari, Z.; Wenger, S.; Vicens, J. Calixcrowns and related molecules. J. Inclusion Phen. and Molec. Recognition in Chem. 1994, 19, 137–148. [Google Scholar] [CrossRef]
  78. Hill, C.; Dozol, J. F.; Lamare, V.; Rouquette, H.; Eymard, S.; Tournois, B.; Vicens, J.; Asfari, Z.; Bressot, C.; Ungaro, R.; Casnati, A. Nuclear waste treatment by means of supported liquid membranes containing calixcrown compounds. J. Inclusion Phen. and Molec. Recognition in Chem. 1994, 19, 399–408. [Google Scholar] [CrossRef]
  79. Asfari, Z.; Bressot, C.; Vicens, J.; Hill, C.; Dozol, J.-F.; Rouquette, H.; Eymard, S.; Lamare, V.; Tournois, B. Doubly crowned calix[4]arenes in the 1.3-alternate conformation as cesium-selective carriers in supported liquid membranes. Analyt. Chem. 1995, 67, 3133–3139. [Google Scholar] [CrossRef]
  80. Hill, C.; Dozol, J.-F.; Rouquette, H.; Tournois, B.; Vicens, J.; Asfari, Z.; Bressot, C.; Ungaro, R.; Casnati, A. J. Inclusion Phen. and Molec. Recognition in Chem. 1995, 18, 1.
  81. Hill, C.; Dozol, J.-F.; Rouquette, H.; Eymard, S.; Tournois, B. J. Membr. Sci. 1996, 114, 83–80.
  82. Haverlock, T. J.; Bonnesen, P. V.; Sachleben, R. A.; Moyer, B. A. Applicability of a calixarene-crown compound for the removal of cesium from alkaline tank waste. Radiochimica Acta 1997, 76, 103–108. [Google Scholar] [CrossRef]
  83. Dozol, J.-F.; Asfari, Z.; Hill, C.; Vicens, J. Commissariat A L’Energie Atomique, P., FP Appl. No. 92 14245. 1997. [Google Scholar]
  84. Lamare, V.; Bressot, C.; Dozol, J.-F.; Vicens, J.; Asfari, Z.; Ungaro, R.; Casnati, A. Sep. Sci. Technology 1997, 32, 175–191.
  85. Kim, J. S.; Suh, I. H. K.; Jong, Kuk; Cho, M. H. Selective sensing of caesium ions by novel calix[4]arene bis(dibenzocrown) ethers in an aqueous environment. J. Chem. Soc., Perkin Trans. 1 1998, 2307–2312. [Google Scholar] [CrossRef]
  86. Moyer, B. A.; Bonnesen, P. V.; Delmau, L. H.; Heverlock, T. J.; Sachleben, R. A.; Leonard, R. A.; Conner, C.; Lumetta, G. L. Proc. ISEC’99; Barcelona 11-16. July; Cox, M., et al., Eds.; Society of Chem. Ind. SCI: London, UK, 2001. [Google Scholar]
  87. Delmau, L. H.; Bonnesen, P. V.; Van Berkel, G. J.; Moyer, B. A. Improved performance of the alkaline-side CSEX process for cerium extraction from alkaline high-level waste obtained by characterization of the effect of surfactant impurities. ORNL/TM-1999/209. ORNL: USA, 1999. [Google Scholar]
  88. Kim, J. S.; Ohki, A.; Ueki, R.; Ishizuka, T.; Shimotashiro, T.; Maeda, S. Cesium-ion selective electrodes based on calix[4]arene dibenzocrown ethers. Talanta 1999, 48, 705–710. [Google Scholar] [CrossRef]
  89. Zeng, H.; Dureault, B. Cesium-selective optode membrane based on the lipophilic calix[4]biscrown in the 1,3-alternate conformation. Talanta 1998, 46, 1485–1491. [Google Scholar] [CrossRef]
  90. Chang, S.-K.; Kwon, S.-K.; Cho, I. Calixarene-based amide ionophores for group IIA metal cations. Chem. Letters 1987, 947–948. [Google Scholar] [CrossRef]
  91. Kimura, K.; Matsuo, M.; Shono, T. Lipophilic calix[4]areneester and amide derivatives as neutral carriers for sodium ion-selective electrodes. Chem. Letters 1988, 615–616. [Google Scholar] [CrossRef]
  92. Arnaud-Neu, F.; Schwing-Weill, M.-J.; Ziat, K.; Cremin, S.; Harris, S. J.; McKervey, M. A. Selective alkali and alkaline earth cation complexation by calixarene amides. New J. Chem. 1991, 15, 33–37. [Google Scholar]
  93. Arnaud-Neu, F.; Barrett, G.; Fanni, S.; Marrs, D.; McGregor, W.; McKervey, M. A.; Schwing-Weill, M.-J.; Vetrogon, V.; Wechsler, S. Extraction and solution thermodynamics of complexation of alkali and alkaline earth cations by calix[4]arene amides. J. Chem. Soc., Perkin Trans. II 1995, 453–461. [Google Scholar] [CrossRef]
  94. Arnaud-Neu, F.; Barrett, G.; Harris, S. J.; Owens, M.; McKervey, M. A.; Schwing-Weill, M.-L.; Schwinte, P. Cation complexation by chemically modified calixarenes. 5. Protonation constants for calixarene carboxylates and stability constants of their alkali and alkaline-earth complexes. Inorg. Chem. 1993, 32, 2644–2650. [Google Scholar] [CrossRef]
  95. McKittrick, T.; Diamond, D.; Marrs, D. J.; O’Hagan, P.; McKervey, M. A. Calcium-selective electrode based on a calix[4]arene tetraphosphine oxide. Talanta 1996, 43, 1145–1148. [Google Scholar] [CrossRef]
  96. Kubo, Y.; Hamaguchi, S.; Niimi, A.; Yoshida, K.; Tokita, S. Synthesis of a 1,3-Bis(indoaniline)-derived Calix[4]arene as an Optical Sensor for Calcium Ion. J. Chem. Soc., Chem. Commun. 1993, 305, 305–307. [Google Scholar] [CrossRef]
  97. Kubo, Y.; Obara, S.; Tokita, S. Effective signal control (off-on-off action) by metal ionic inputs on a new chromoionophore-based calix[4]crown. J. Chem. Soc., Chem. Commun. 1999, 2399–2400. [Google Scholar] [CrossRef]
  98. Kim, N. Y.; Chang, S.-K. Calix[4]arenes Bearing Two Distal Azophenol Moieties: Highly Selective Chromogenic Ionophores for the Recognition of Ca2+ Ion. J. Org. Chem. 1998, 63, 2362–2364. [Google Scholar] [CrossRef]
  99. Gupta, V. K.; Jain, A. K.; Khurana, U.; Singh, L. P. PVC-based neutral carrier and organic exchanger membranes as sensors for the determination of Ba2+ and Sr2+. Sensors and Actuators B: Chemical 1999, B 55, 201–211. [Google Scholar] [CrossRef]
  100. Dozol, J. F.; GArcia-Carrera, A.; Lamare, V.; Rouquette, H. Proceedings of ISEC 2002; C. v. Rensburg Publ. Ltd.: Melville, 2002; pp. 1168–1173.
  101. Vanura, P.; Stibor, I. Extraction of Alkaline and Alkaline-Earth Metal Cations at Synergistic Action of Bis(undecahydro-7,8-dicarbaundecaborato(2-))cobaltate(1-) and Substituted Calix(n)arenes in Nitrobenzene. Col-lect. Czech. Chem. Commun. 1998, 63, 2009–2014. [Google Scholar] [CrossRef]
  102. Arena, G.; Contino, A.; Magri, A.; Sciotto, D.; Lamb, J. D. Abstr. 4th International Conference on Calixarenes; Parma, Italy 31.8.-4.9., 1997; p. 131. [Google Scholar]
  103. Beer, P. D.; Gale, P. A.; Chen, Z.; Drew, M. G. B.; Heath, J. A.; Ogden, M. I.; Powell, H. R. New ionophoric calix[4]diquinones: coordination chemistry, electrochemistry, and X-ray crystal structure. Inorg. Chem. 1997, 36, 5880–5893. [Google Scholar] [CrossRef] [PubMed]
  104. Beer, P. D.; Gale, P. A.; Chen, G. Z. Electrochemical molecular recognition: pathways between complexation and signalling. J. Chem. Soc., Dalton Trans. 1999, 1897–1909. [Google Scholar] [CrossRef]
  105. Gutsche, C. D.; Iqbal, M.; Alam, I. Calixarenes. 20. The interaction of calixarenes and amines. J. Amer. Chem. Soc. 1987, 109, 4314–4320. [Google Scholar] [CrossRef]
  106. Shinkai, S.; Mori, S.; Koreichi, H.; Tsubaki, T.; Manabe, O. Hexasulfonated calix[6]arene derivatives: A new class of catalysts, surfactants, and host molecules. J. Amer. Chem. Soc. 1986, 108, 2409–2416. [Google Scholar] [CrossRef] [PubMed]
  107. Shinkai, S. Calixarenes - The third generation of supramolecules. Tetrahedron 1993, 49, 8933–8968. [Google Scholar] [CrossRef]
  108. Arnecke, R.; Böhmer, V.; Cacciapaglia, R.; Dalla Cort, A.; Mandolini, L. Cation-π interactions between neutral calix[5]arene hosts and cationic organic guests. Tetrahedron 1997, 53, 4901–4908. [Google Scholar] [CrossRef]
  109. Murayama, K.; Aoki, K. Molecular recognition involving multiple cation-π interactions: The inclusion of the ace-tylcholine trimethylammonium moiety in resorcin[4]arene. J. Chem. Soc., Chem. Commun. 1997, 119–120. [Google Scholar] [CrossRef]
  110. Harrowfield, J. M.; Richmond, W. R.; Sobolev, A. N. Inclusion of quaternary ammonium compounds by calixarenes. J.Inclusion Phen. and Molec. Recognition in Chem. 1994, 19, 257–276. [Google Scholar] [CrossRef]
  111. Koh, K. N.; Araki, K.; Ikeda, A.; Otsuka, H.; Shinkai, S. Reinvestigation of calixarene-based artificial-signaling acetylcholine receptors useful in neutral aqueous (water/methanol) solution. J. Amer. Chem. Soc. 1996, 118, 755–758. [Google Scholar] [CrossRef]
  112. Giannetto, M.; Mori, G.; Notti, A.; Pappalardo, S.; Parisi, M. F. Abstr. 4th International Conference on Calixarenes; Parma, Italy 31.8.-4.9., 1997; p. 98. [Google Scholar]Analyt. Chem. 1998, 70, 4631–4635.
  113. Chang, S.-K.; Jang, M.; Han, S. Y.; Lee, J. H.; Kang, M. H.; No, K. T. Molecular recognition of butylamines by calixarene-based ester ligands. Chem. Letters 1992, 1937–1940. [Google Scholar] [CrossRef]
  114. Chan, W. H.; Shiu, K. K.; Gu, X. H. Analyst 1993, 118, 863.
  115. Han, S.-Y.; Kang, M.-H.; Jung, Y. E.; Chang, S.-K. Molecular recognition of alkylamines: conformational and binding properties of calix[6]arene-based ester ligands. J. Chem. Soc., Perkin Trans. II 1994, 835–839. [Google Scholar] [CrossRef]
  116. Ahn, S.; Chang, S.-K.; Kim, T.; Lee, J. W. An NMR study on complexation of p-tert-butylcalix[6]arene ester derivatives with ethylammonium picrate. Chem. Letters 1995, 297–298. [Google Scholar] [CrossRef]
  117. Odashima, K.; Yagi, K.; Tohda, K.; Umezawa, Y. Potentiometric discrimination of organic amines by a liquid membrane electrode based on a lipophilic hexaester of calix[6]arene. Analyt. Chem. 1993, 65, 1074–1083. [Google Scholar] [CrossRef]
  118. Zolotov, Y. A.; Pletnev, I. V.; Torocheshnikova, I. I.; Shvedene, N. V.; Nemilova, M. Y.; Kovalev, V. V.; Shokova, E. A.; Smirnova, S. V. Extraction and determination of amino compounds with calix[8]arenes. Solvent Extraction Research and Development, Japan 1994, 1, 123–131. [Google Scholar]
  119. Shvedene, N. V.; Nemilova, M. Y.; Zatonskaya, V. L.; Pletnev, I. V.; Baulin, V. E.; Lyubitov, I. E.; Shvyadas, V. K. J. Analyt. Chem. 1995, 50, 402.
  120. Shvedene, N. V.; Nemilova, M. Y.; Kovalev, V. V.; Shokova, E. A.; A.K., R.; Pletnev, I. V. Sens. Actuators B 1995, 372.
  121. Casnati, A.; Minari, P.; Pochini, A.; Ungaro, R.; Nijenhuis, W. F.; deJong, F.; Reinhoudt, D. N. Selective complexation and membrane transport of guanidinium salts by calix[6]arene amides. Isr. J. Chem. 1992, 32, 79–87. [Google Scholar] [CrossRef]
  122. Dei, L.; Casnati, A.; Lo Nostro, P.; Pochini, A.; Ungaro, R.; Baglioni, P. Complexation properties of p-tert-butylcalix[6]arene hexamide in monolayers at the air-water interface. Langmuir 1996, 12, 1589–1593. [Google Scholar] [CrossRef]
  123. Kremer, F. J. B.; Chiosis, G.; Engbersen, J. F. J.; Reinhoudt, D. N. Improved guanidinium ion-selectivity by novel calix[4]arene and calix[8]arene receptor molecules on CHEMFETs. J. Chem. Soc., Perkin Trans. II 1994, 677–681. [Google Scholar] [CrossRef]
  124. Casnati, A. Calixarenes: From chemical curiosity to a rich source for molecular receptors. Gazz. Chim. Ital. 1997, 127, 637–649. [Google Scholar]
  125. Araki, K.; Akao, K.; Otsuka, H.; Nakashima, K.; Inokuchi, F.; Shinkai, S. Immobilization of the ring inversion motion in calix[6]arene by a cap with C3-symmetry. Chem. Letters 1994, 1251–1254. [Google Scholar] [CrossRef]
  126. Takeshita, M.; Inokuchi, F.; Shinkai, S. C3-symmetrically capped homotrioxacalix[3]arene. A preorganized host molecule for inclusion of primary ammonium ions. Tetr. Lett. 1995, 36, 3341–3344. [Google Scholar] [CrossRef]
  127. Takeshita, M.; Shinkai, S. Novel fluorometric sensing of ammonium ions by pyrene functionalized homotrioxacalix[3]arenes. Chem. Letters 1994, 125–128. [Google Scholar] [CrossRef]
  128. Takeshita, M.; Shinkai, S. A selective fluorometric sensing system for guanidinium ion in the presence of primary ammonium ions. Chem. Letters 1994, 1349–1352. [Google Scholar] [CrossRef]
  129. Jin, T.; Monde, K. Synthesis and optical resolution of a fluorescent chiral calix[4]arene with two pyrene moieties forming an intramolecular excimer. J. Chem. Soc., Chem. Commun. 1998, 1357–1358. [Google Scholar] [CrossRef]
  130. Antipin, I. S.; Stoikov, I. I.; Garifzyanov, A. R.; Konovalov, A. I. Dokl. Akad. Nauk Russ. 1996, 347, 626–628.
  131. Antipin, I. S.; Stoikov, I. I.; Pinkhassik, E. M.; Fitseva, N. A.; Stibor, I.; Konovalov, A. I. Calix[4]arene based alpha-aminophosphonates: Novel carriers for zwitterionic amino acids transport. Tetr. Lett. 1997, 38, 5865–5868. [Google Scholar] [CrossRef]
  132. Beer, P. D.; Chen, Z.; Gale, P. A.; Heath, J. A.; Knubley, R. J.; Ogden, M. I. Cation recognition by new diester- and diamide-calix[4]arenediquinones and a diamide-benzo-15-crown-5-calix[4]arene. J. Inclusion Phen. and Molec. Recognition in Chem. 1994, 19, 343–359. [Google Scholar] [CrossRef]
  133. Beer, P. D.; Chen, Z.; Gale, P. A. Diester-calix[4]arenediquinone complexation and electrochemical recognition of group 1 and 2, ammonium and alkyl ammonium guest cations. Tetrahedron 1994, 50, 931–940. [Google Scholar] [CrossRef]
  134. Song, B. M.; Chang, S.-K. Bull. Korean Chem. Soc. 1993, 14, 540.
  135. Jung, Y. E.; Song, B. M.; Chang, S.-K. Molecular recognition of alkyl- and arylalkyl-amines in dichloromethane and chloroform by calix[4]-crown ethers. J. Chem. Soc., Perkin Trans. II 1995, 2031–2034. [Google Scholar] [CrossRef]
  136. Arduini, A.; McGregor, W. M.; Paganuzzi, D.; Pochini, A.; Secchi, A.; Ugozzoli, F.; Ungaro, R. Rigid cone calix[4]arenes as π-donor systems: Complexation of organic molecules and ammonium ions in organic media. J. Chem. Soc., Perkin Trans. II 1996, 839–846. [Google Scholar] [CrossRef]
  137. Casnati, A.; Jacopozzi, P.; Pochini, A.; Ugozzoli, F.; Cacciapaglia, R.; Mandolini, L.; Ungaro, R. Bridged calix[6]arenes in the cone conformation: New receptors for quaternary ammonium cations. Tetrahedron 1995, 51, 591–598. [Google Scholar] [CrossRef]
  138. Paek, K.; Ihm, H. Synthesis and binding property of tunable upper-rim calix[4]crowns. Chem. Letters 1996, 311–312. [Google Scholar] [CrossRef]
  139. Ihm, H.; Kim, H.; Paek, K. Molecular engineering. Part 2. Influence of side-chain substituents in lariat-type upper-rim calix[4]crowns on their binding properties and the reversal of these. J. Chem. Soc., Perkin Trans. 1 1997, 1997–2003. [Google Scholar] [CrossRef]
  140. Pappalardo, S.; Parisi, M. F. Inherently chiral calix[4]crown ethers. Tetr. Lett. 1996, 37, 1493–1496. [Google Scholar] [CrossRef]
  141. Kubo, Y.; Maruyama, S.; Ohhara, N.; Nakamura, M.; Tokita, S. Molecular recognition of butylamines by a binaphtyl-derived chromogenic calix[4]crown. J. Chem. Soc., Chem. Commun. 1995, 1727–1728. [Google Scholar] [CrossRef]
  142. Kubo, Y.; Maeda, S.; Tokita, S.; Kubo, M. Colorimetric chiral recognition by a molecular sensor. Nature 1996, 382, 522–524. [Google Scholar] [CrossRef]
  143. Kubo, Y. Binaphthyl-appended chromogenic receptors: Synthesis and application to their colorimetric recognition of amines. Syn. Lett. 1999, 161–174. [Google Scholar] [CrossRef]
  144. Zheng, Q.-Y.; Chen, C.-F.; Huang, Z.-T. Synthesis of new chromogenic calix[4]crowns and molecular recognition of alkylamines. Tetrahedron 1997, 53, 10345–10356. [Google Scholar] [CrossRef]
  145. Chawla, H. M.; Srinivas, K. Molecular diagnostics: Synthesis of new chromogenic calix[8]arenes as potential reagents for detection of amines. J. Chem. Soc., Chem. Commun. 1994, 2593–2594. [Google Scholar] [CrossRef]
  146. Chawla, H. M.; Srinivas, K. Synthesis of new chromogenic calix[8]arenes. Tetr. Lett. 1994, 35, 2925–2928. [Google Scholar] [CrossRef]
  147. Chawla, H. M.; Srinivas, K. Synthesis of new chromogenic calix[4]arenes bridged at the upper rim through bisazophenyl linkages. J. Org. Chem. 1996, 61, 8464–8467. [Google Scholar] [CrossRef]
  148. Ikeda, A.; Shinkai, S. On the origin of high ionophoricity of 1,3-alternate calix[4]arenes: π-Donor participation in complexation of cations and evidence for metal-tunneling through the calix[4]arene cavity. J. Amer. Chem. Soc. 1994, 116, 3102–3110. [Google Scholar] [CrossRef]
  149. Ikeda, A.; Tsuzuki, H.; Shinkai, S. NMR spectroscopic and x-ray crystallographic studies of calix[4]arene Ag+ complexes. Influence of bound Ag+ on C2v - C2v interconversion in cone-calix[4]arenes. J. Chem. Soc., Perkin Trans. II 1994, 2073–2080. [Google Scholar] [CrossRef]
  150. Inokuchi, F.; Miyahara, Y.; Inazu, T.; Shinkai, S. Selektive Erkennung von Alkalimetall-Kationen durch π-basische, molekulare Hohlräume und einfacher massenspektrometrischer Nachweis von Kation-Aren-Komplexen. Angew. Chemie 1995, 107, 1459–1461. [Google Scholar] [CrossRef]
  151. Kimura, K.; Tatsumi, K.; Yajima, S.; Miyake, S.; Sakamoto, H.; Yokoyama, M. Calix[4]arene derivative bearing pi-coordinate substituents as neutral carrier for silver ion sensors. Chem. Letters 1998, 833–834. [Google Scholar] [CrossRef]
  152. Couton, D.; Mocerino, M.; Rapley, C.; Kitamura, C.; Yoneda, A.; Ouchi, M. Silver and thallium ion complexation with allyloxycalix[4]arenes. Aust. J. Chem. 1999, 52, 227–229. [Google Scholar] [CrossRef]
  153. Cobben, P. H. L. M.; Egberink, R. J. M.; Bomer, J. G.; Bergvelt, P.; Verboom, W.; Reinhoudt, D. N. Transduction of selective recognition of heavy metal ions by chemically modified field effect transistors (CHEMFETs). J. Amer. Chem. Soc. 1992, 114, 10573–10582. [Google Scholar] [CrossRef]
  154. Malinowska, E.; Brzozka, Z.; Kasiura, K.; Egberink, R. J. M.; Reinhoudt, D. N. Silver electrodes based on thioether functionalized calix[4]arene as ionophores. Analyt. Chim. Acta 1994, 298, 245–251. [Google Scholar] [CrossRef]
  155. Lugtenberg, R. J. W.; Antonisse, M. M. G.; Egberink, R. J. M.; Engbersen, J. F. J.; Reinhoudt, D. N. Polysiloxane based CHEMFETs for the detection of heavy metal ions. J. Chem. Soc., Perkin Trans. 2 1996, 1937. [Google Scholar] [CrossRef]
  156. Bakker, W. I. M. Determination of unbiased selectivity coefficients of neutral carrier-based cation-selective elec-trodes. Analyt. Chem. 1997, 69, 1061–1069. [Google Scholar] [CrossRef]
  157. O’Connor, K. M.; Svehla, G.; Harris, S. J.; McKervey, M. A. Calixarene-based potentiometric ion-selective elec-trodes for silver. Talanta 1992, 39, 1549–1554. [Google Scholar] [CrossRef]
  158. O’Connor, K. M.; Svehla, G.; Harris, S. J.; McKervey, M. A. Evaluation of calixarene-based ionophores for silver ions. Analyt. Proc. 1993, 30, 137–139. [Google Scholar]
  159. Arnaud-Neu, F.; Barrett, G.; Corry, D.; Cremin, S.; Ferguson, G.; Gallagher, J. F.; Harris, S. J.; McKervey, M. A.; Schwing-Weill, M.-J. Cation complexation by chemically modified calixarenes. Part 10. Thioamide derivatives of p-tert-butylcalix[4]-, [5]- and [6]-arenes with selectivity for copper, silver, cadmium and lead. X-ray molecular structures of calix[4]arene thioamide-lead(II) and calix[4]arene amide-copper(II) complexes. J. Chem. Soc., Perkin Trans. 2 1997, 575–579. [Google Scholar]
  160. Harris, S. J. Europ. Patent Appl. EP 432.989, CA 182813r (115) IE Appl. 3.982, 1991.
  161. Yordanov, A. T.; Roundhill, D. M.; Mague, J. T. Extraction selectivities of lower rim substituted calix[4]arene hosts induced by variations in the upper rim substituents. Inorg. Chim. Acta 1996, 250, 295–302. [Google Scholar] [CrossRef]
  162. Yordanov, A. T.; Falana, O. M.; Koch, H. F.; Roundhill, D. M. (Methylthio)methyl and (N,N-dimethylcarbamoyl)methyl upper rim substituted calix[4]arenes as potential extractants for Ag(I), Hg(II), Ni(II), Pd(II), Pt(II), and Au(III). Inorg. Chem. 1997, 36, 6468–6471. [Google Scholar] [CrossRef]
  163. Haase, W.; Ahlers, B.; Reinbold, J.; Cammann, K.; Brodesser, G.; Vögtle, F. Sens. Actuators, B 1994, 18-19, 380.
  164. Lutze, O.; Meruva, R. K.; Frielich, A.; Ramamurthy, N.; Brown, R. B.; Hower, R.; Meyerhoff, M. E. Stabilized potentiometric solid-state polyion sensors using silver-calixarene complexes as additives within ion-exchanger-based polymeric films. Fresenius J. Analyt. Chem. 1999, 364, 41–47. [Google Scholar] [CrossRef]
  165. Roundhill, M. Extraction of Metals from Soils and Waters, 1. ed.; Kluwer Academic/Plenum Publ.: New York, 2001; p. 375. [Google Scholar]
  166. McKervey, M.A.; Schwing-Weill, M.J.; Arnaud-Neu, F. Cation Binding by Calixarenes, in [4]. 537–603.
  167. Arrigan, D. W. M.; Svehla, G.; Harris, S. J.; McKervey, M. A. Use of calixarenes as modifiers of carbon paste electrodes for voltammetric analysis. Electroanalysis 1994, 6, 97–106. [Google Scholar] [CrossRef]
  168. Lugtenberg, R. J. W.; Egberink, R. J. M.; Engbersen, J. F. J.; Reinhoudt, D. N. Pb2+ and Cd2+ selective chemically modified field effect transistors based on thioamide functionalized 1,3-alternate calix[4]arenes. J. Chem. Soc., Perkin Trans. 2 1997, 1353–1357. [Google Scholar] [CrossRef]
  169. Malinowska, E.; Brzozka, Z.; Kasiura, K.; Egberink, R. J. M.; Reinhoudt, D. N. Lead selective electrodes based on thioamide functionalized calix[4]arene as ionophores. Analyt. Chim. Acta 1994, 298, 253–258. [Google Scholar] [CrossRef]
  170. Arrigan, D. W. M.; Svehla, G.; Harris, S. J.; McKervey, M. A. Stripping voltammetry with a polymeric calixarene modified carbon paste electrode. Analyt. Proc. 1992, 29, 27. [Google Scholar]
  171. Honeychurch, K. C.; Hart, J. P.; Cowell, D. C.; Arrigan, D. W. M. Voltammetric studies at calixarene modified screen-printed carbon electrodes and its trace determination in water by stripping voltammetry. Sens. Act. B 2001, 77, 642–652. [Google Scholar] [CrossRef]
  172. Boerrigter, H.; Lugtenberg, R. J. W.; Egberink, R. J. M.; Verboom, W.; Spek, A. L. Resorcinarene cavitands as building blocks for cation receptors. Gazz. Chim. Ital. 1997, 127, 709–716. [Google Scholar]
  173. Wroblewski, W.; Brzozka, Z.; Janssen, R. G.; Verboom, W.; Reinhoudt, D. N. Lead versus cadmium selectivity of ion selective electrodes based on thiophosphorylated calix[6]arene ionophores. New J. Chem. 1996, 20, 419–426. [Google Scholar]
  174. Cadogan, F.; Diamond, D. Abstr. 5th Int. Conf. Calixarene Chem.; Univ. Western Australia: Perth 19-23.9., 1999; p. L-20. [Google Scholar]
  175. Narita, M.; Higuchi, Y.; Hamada, F.; Kumagai, H. Metal sensor of water soluble dansyl-modified thiacalix[4]arenes. Tetr. Lett. 1998, 39, 8687–8690. [Google Scholar] [CrossRef]
  176. Vataj, R.; Ridaoui, H.; Louati, A.; Gabelica, V.; Steyer, S.; Matt, D. First synthesis of a ‘1,2-diquinone-calix[4]arene’. Interaction of its reduced form with Ag+. J. Electroanal. Chem. 2002, 519, 123–129. [Google Scholar] [CrossRef]
  177. Talantova, G. G.; Hwang, H.-S.; Talantov, V. S.; Bartsch, R. A. Calix[4]arene with hard donor groups as efficient soft cation extractants. Remarkable extraction selectivity of calix[4]arene N-(X)sulfonylcarboxamides for Hg(II). J. Chem. Soc., Chem. Commun. 1998, 1329–1330. [Google Scholar] [CrossRef]
  178. Talantova, G. G.; Elkarim, N. S. A.; Talantov, V. S.; Bartsch, R. A. A calixarene-based fluorogenic reagent for selective mercury(II) recognition. Analyt. Chem. 1999, 71, 3106–3109. [Google Scholar] [CrossRef] [PubMed]
  179. Kimura, K.; Tatsumi, K.; Yokoyama, M.; Ouchi, M.; Mocerino, M. Remarkable thallium(I) selectivity for ion sensors based on -coordination of calix[4]arene neutral carriers. Analyt. Commun. 1999, 36, 229–230. [Google Scholar] [CrossRef]
  180. Unob, F.; Asfari, Z.; Vicens, J. An Anthracene-Based Fluorescent Sensor for Transition Metal Ions Derived From Calix[4]arene. Tetr. Lett. 1998, 39, 2951–2954. [Google Scholar] [CrossRef]
  181. Sato, N.; Yoshida, I.; Shinkai, S. Energy-transfer luminescence of lanthanide ions complexed with water-soluble calix[n]arenes. Chem. Letters 1993, 1261–1264. [Google Scholar] [CrossRef]
  182. Yoshida, I.; Koyabu, K.; Nishida, M.; Sagara, F.; Ishii, D.; Shinkai, S. Highly selective spectrofluorometric detec-tion of terbium(III) with tetrasodium calix[4]arene-p-sulfonate. Analyt. Sci. 1994, 10, 353–354. [Google Scholar] [CrossRef]
  183. Matsumoto, H.; Shinkai, S. Energy-transfer luminescence of lanthanide ions encapsulated in calix[4]arenes. Corre-lation between energy level of sensitizers and the quantum yield. Chem. Letters 1994, 901–904. [Google Scholar] [CrossRef]
  184. Casnati, A.; Fischer, C.; Guardigli, M.; Isernia, A.; Manet, I.; Sabbatini, N.; Ungaro, R. Synthesis of calix[4]arene receptors incorporating (2,2’-bipyridin-6-yl)methyl and (9-methyl-1,10-phenanthrolin-2-yl)methyl chromophores and luminescence of their Eu3+ and Tb3+ complexes. J. Chem. Soc., Perkin Trans. 2 1996, 395–399. [Google Scholar] [CrossRef]
  185. Sabbatini, N.; Casnati, A.; Fischer, C.; Girardini, R.; Guardigli, M.; Manet, I.; Sarti, G.; Ungaro, R. Luminescence of Eu and Tb complexes of new macrobicyclic ligands derived from p-t-butylcalix[4]arene. Inorg. Chim. Acta 1996, 252, 19–24. [Google Scholar] [CrossRef]
  186. Charbonniere, L. J.; Balsiger, C.; Schenk, K. J.; Bünzli, J.-C. G. Complexes of p-t-butylcalix[5]arene with lantha-nides: synthesis, structure and photophysical properties. J. Chem. Soc., Dalton Trans. 1998, 505–510. [Google Scholar] [CrossRef]
  187. Georgiev, E. M.; Clymire, J.; McPherson, G. L.; Roundhill, D. M. Luminescent europium(III) and terbium(III) ions encapsulated in a 2-aminoethoxy or carbamoyloxy substituted calixarene host. Inorg. Chim. Acta 1994, 227, 293–296. [Google Scholar] [CrossRef]
  188. Steemers, F. J.; Meuris, H. G.; Verboom, W.; Reinhoudt, D. N.; van der Tol, E. B.; Verhoeven, J. W. Water-soluble neutral calix[4]arene-lanthanide complexes: Synthesis and luminescence properties. J. Org. Chem. 1997, 62, 4229–4235. [Google Scholar] [CrossRef] [PubMed]
  189. Ulrich, G.; Ziessel, R.; Manet, I.; Guardigli, M.; Sabbatini, N.; Fraternali, F.; Wipff, G. Calix[4]arene podands and barrelands incorporating 2,2’-bipyridine moieties and their lanthanide complexes: luminescence properties. Chem. Eur. J. 1997, 3, 1815–1822. [Google Scholar] [CrossRef]
  190. Sato, N.; Shinkai, S. Energy-transfer luminescence of lanthanide ions encapsulated in sensitizer-modified calix[4]arenes. J. Chem. Soc., Perkin Trans. II 1993, 621–624. [Google Scholar] [CrossRef]
  191. Rudkevich, D. M.; Verboom, W.; van der Tool, E.; van Staveren, C. J.; Kaspersen, F. M.; Verhoeven, J. W.; Reinhoudt, D. N. Calix[4]arene-triacids as receptors for lanthanides; synthesis and luminescence of neutral Eu3+ and Tb3+ complexes. J. Chem. Soc., Perkin Trans. II 1995, 131–134. [Google Scholar] [CrossRef]
  192. Sabbatini, N.; Guardigli, M.; Manet, I.; Ziessel, R.; Ulrich, G.; Ungaro, R.; Casnati, A.; Fischer, C. Synthesis and luminescence of Eu3+ and Tb3+ complexes with novel calix[4]arene ligands carrying 2,2’-bipyridine subunits. New J. Chem. 1995, 19, 137–140. [Google Scholar]
  193. Steemers, F. J.; Verboom, W.; Reinhoudt, D. N.; van der Tol, E. B.; Verhoeven, J. W. New sensitizer-modified calix[4]arenes enabling near-UV excitation of complexed luminescent lanthanide ions. J. Amer. Chem. Soc. 1995, 117, 9408–9414. [Google Scholar] [CrossRef]
  194. Oude Wolbers, M. P.; van Veggel, F. C. J. M.; Peters, F. G. A.; van Beelen, E. S. E.; Hofstraat, J. W.; Geurts, F. A. J.; Reinhoudt, D. N. Sensitized Near-Infrared Emission from Nd3+ and Er3+ Complexes of Fluorescein-Bearing Calix[4]arene Cages. Chem. Eur. J. 1998, 4, 772–780. [Google Scholar] [CrossRef]
  195. Shevchuk, S.; Bacherikov, V.; Turianskaya, A.; Korovin, Y.; Alexeeva, E.; Rusakova, N.; Gren, A. Proc. Int. Conf. Calixarenes; Perth, 1999. [Google Scholar]
  196. Shevchuk, S.; Alexeeva, E.; Rusakova, N. V.; Korovin, Y. V.; Bacherikov, V. A.; Gren, A. Some new properties of calixarenes: the luminescence of four lanthanide ions in calixarene complexes. Mendeleev Commun. 1998, 112–113. [Google Scholar] [CrossRef]
  197. Ludwig, R.; Matsumoto, H.; Takeshita, M.; Ueda, K.; Shinkai, S. Study on monolayers of metal complexes of calixarenes and their luminescence properties. Supramol. Chem. 1995, 4, 319–327. [Google Scholar] [CrossRef]
  198. Froidevaux, P.; Bünzli, J.-C. G. Energy-transfer processes in lanthanide dinuclear complexes with p-tert-butyl-calix[8]arene: An example of dipole-dipolar mechanism. J. Phys. Chem. 1994, 98, 532–536. [Google Scholar] [CrossRef]
  199. Bünzli, J.-C.; Ihringer, F. Photophysical properties of lanthanide dinuclear complexes with p-nitro-calix[8]arene. Inorg. Chim. Acta 1996, 246, 195–205. [Google Scholar] [CrossRef]
  200. Kubo, Y.; Maeda, S.; Nakamura, M.; Tokita, S. A uranyl ion sensitive chromoionophore based on calix[6]arene. J. Chem. Soc., Chem. Commun. 1994, 1725–1726. [Google Scholar] [CrossRef]
  • Sample availability: Some of the compounds are available from the author.

Share and Cite

MDPI and ACS Style

Ludwig, R.; Dzung, N.T.K. Calixarene-Based Molecules for Cation Recognition. Sensors 2002, 2, 397-416. https://doi.org/10.3390/s21000397

AMA Style

Ludwig R, Dzung NTK. Calixarene-Based Molecules for Cation Recognition. Sensors. 2002; 2(10):397-416. https://doi.org/10.3390/s21000397

Chicago/Turabian Style

Ludwig, Rainer, and Nguyen Thi Kim Dzung. 2002. "Calixarene-Based Molecules for Cation Recognition" Sensors 2, no. 10: 397-416. https://doi.org/10.3390/s21000397

Article Metrics

Back to TopTop