Anion coordinated capsules and pseudocapsules of tripodal amide, urea and thiourea scaffolds

Sandeep Kumar Dey , Arghya Basu , Romen Chutia and Gopal Das *
Department of Chemistry, Indian Institute of Technology Guwahati, 781039 Assam, India. E-mail: gdas@iitg.ernet.in

Received 5th January 2016 , Accepted 2nd March 2016

First published on 3rd March 2016


Abstract

This review aims to deliver a detailed and comparative account of the reported examples of anion (halide and oxyanions) coordinated capsules and pseudocapsules of tripodal receptors that employ hydrogen bonds and/or electrostatic hydrogen bonds offered by specific binding sites from amide, urea and thiourea functionalities. The review discusses both the structural aspects of anion binding and solution-state anion binding affinities of N-bridged and aryl-bridged tripodal receptors. Discussions relating to selective anion recognition and separation, carbondioxide uptake, and transmembrane anion transport, as demonstrated by some of these tripodal receptors have also been included in this review.


image file: c6ra00268d-p1.tif

Sandeep Kumar Dey

Dr Sandeep Kumar Dey received his MSc degree in 2007 from North Eastern Hill University (NEHU) Shillong, India and PhD in Supramolecular Chemistry from Indian Institute of Technology Guwahati in 2013 under the supervision of Prof. Gopal Das. He then joined the group of Prof. Shih-Sheng Sun as a postdoctoral research fellow in Institute of Chemistry, Academia Sinica, Taiwan. Presently, he is an Alexander von Humboldt (AvH) research fellow working in the group of Prof. Dr Christoph Janiak at Heinrich-Heine-Universität Düsseldorf, Germany. His research interests are mainly in the field of supramolecular chemistry and functional porous materials for various applications.

image file: c6ra00268d-p2.tif

Arghya Basu

Dr Arghya Basu received his MSc degree in 2008 from Indian Institute of Engineering Science and Technology, Shibpur, India. He obtained his PhD in Supramolecular Chemistry from Indian Institute of Technology Guwahati in 2014 under the supervision of Prof. Gopal Das. He then joined the group of Prof. Hiroyuki Furuta as a postdoctoral research fellow in Kyushu University, Japan (2014–2015). Presently, he is a DST Young Scientist, working in the group of Prof. Rahul Banerjee at CSIR-NCL Pune, India. His research interests are mainly in the field of supramolecular chemistry and π-conjugated systems for material applications.

image file: c6ra00268d-p3.tif

Romen Chutia

Dr Romen Chutia received his BSc degree in 2008 from Sibsagar College affiliated to Dibrugarh University and MSc degree in 2010 from Gauhati University, India. He obtained his PhD in Supramolecular Chemistry from Indian Institute of Technology Guwahati in 2016 under the supervision of Prof. Gopal Das. Presently, he is a faculty member in the Department of Chemistry, Sibsagar College, and his research interests are mainly focused on supramolecular chemistry.

image file: c6ra00268d-p4.tif

Gopal Das

Prof. Gopal Das received his MSc degree in 1993 from Calcutta University, India. He obtained his PhD from Indian Institute of Technology Kanpur in 1999 under the supervision of Prof. Parimal K. Bharadwaj. He then joined the group of Prof. Stefan Matile as a postdoctoral research associate in University of Geneva, Switzerland. In 2004, he joined the Indian Institute of Technology Guwahati as an Assistant Professor and from 2013 he is Full Professor. He has published about 150 research papers till date and his current research interests are in the field of supramolecular chemistry and development of sensors and biosensors.


Introduction

The significance of anion coordination chemistry has been increasingly appreciated in recent years because of the important roles that anions play in many biological, environmental, and chemical processes.1–4 Over the past two decades, supramolecular chemistry has emerged to achieve recognition and sensing of anionic guests mostly in organic solvent media,5–10 and also in competitive aqueous environments.11–15 Indeed, cationic macrocyclic hosts had been employed in some early elegant works to achieve anion binding in water by electrostatic interactions.15 One of the most extended approaches towards the design of anion receptor is based on the use of –NH functions that interact through hydrogen bonds with the anionic guests. Along this line, neutral anion receptors containing multiple hydrogen bond donor groups, such as amide, urea/thiourea and pyrrole has been extensively used for the effective and selective binding of anions,5–10 although deprotonation of the most acidic hydrogen bond donor occurs in some cases.16–21 In 2005, Bowman-James et al. has categorized the binding of anions based on their coordination numbers which is helpful in defining the notions of complementarity for a given anion and can aid to the design of optimal anion binding host structures.22 At this juncture, it would not be erroneous to say that the development and systematic studies of numerous acyclic and macrocyclic anion receptors have at least partially solved the problem of selective anion binding and, perhaps more relevantly, the problem of discrimination between different anions. More recently, there has been a growing interest in the development of highly selective anion transporters23–25 with the potential to become future therapeutics and anion induced formation of supramolecular self-assemblies and materials.26–29 The concept of anion coordination chemistry has been proficiently employed to generate a range of exciting supramolecular structures such as catenanes, rotaxanes, foldamers, helices, and capsules.30–35

Since the beginning of anion coordination chemistry, macrocyclic and acyclic podand type hosts functionalized with polarized –NH and –CH donors have widely been employed for anion binding and sensing. However, it has been realized that most of the inorganic anions have very high hydration energies that must be compensated for by the host molecule for effective anion recognition.36 Thus, it is the design of sophisticated three dimensional receptors that is essential to fully encapsulate an anion by creating a highly specific anion binding cavity as observed in various anion binding proteins.37,38 In comparison to macrocyclic hosts, podands are more readily synthesized, can display rapid complexation/decomplexation kinetics, may undergo significant conformational changes upon anion binding and thus, may form the basis of switchable sensing devices.39 Along this line, numerous hydrogen bonding tripodal scaffolds have been developed to form stable host–guest complexes with anions of different geometries such as, spherical halides, planar carbonate/nitrate and tetrahedral sulfate/phosphate.40–43 Tripodal receptors constitute a special class of acyclic ionophores with C3v symmetry, whose side arms upon functionalization with appropriate anion binding elements can selectively recognize one or more anionic guest species via topological complementarity. The geometry and orientation of the receptor side arms in N-bridged and aryl-bridged tripodal scaffolds favour the formation of capsules or pseudocapsules upon anion encapsulation by multiple hydrogen bonding. Based on its selectivity, a tripodal receptor capsule can create a distinct microenvironment to isolate an anionic guest from the bulk of solvent media by encapsulation, and thereby, leads to the phenomenon of selective anion encapsulation when possible formation of different capsules are present in the same solution. Numerous reviews describing the coordination and supramolecular chemistry of anions have been published in recent years and over the last decade. However, reviews that discussed the anion induced capsular assembly formation are only few30–32,41,42 and require more attention to be paid. Anion induced capsules and pseudocapsules of tripodal scaffolds have shown a range of exciting properties such as, encapsulation of hydrated anion clusters, fixation of aerial carbon dioxide as carbonate, liquid–liquid extraction of anions from water, selective separation of anions by crystallization and transmembrane anion transportation. This review aims to deliver a comprehensive compilation of the examples reported till date related with the anion induced capsules and pseudocapsules of tripodal receptors that employ hydrogen bonds and/or electrostatic hydrogen bonds offered by specific binding sites from amide, urea, thiourea and hybrid urea–amide functionalities (Schemes 1–5). The review has been divided into four main sections based on the hydrogen bond donating functionality incorporated in the tripodal scaffold for anion coordination. The primary objective of the review is to provide a detailed and comparative account of the solid- and solution-states binding of anions (halides and oxyanions) within the cavity of hydrogen bonding tripodal scaffolds. Some of these tripodal receptors have been found to display selective anion recognition, aerial carbondioxide capture, and selective anion separation, which have also been discussed here. This review covers some of the significant research works from our group and from the groups of Ghosh et al., Custelcean et al., Gale et al., Wu et al., Steed et al., Sun et al., Gunnlaugsson et al. and Hossain et al.


image file: c6ra00268d-s1.tif
Scheme 1 Structures of tripodal tris-amide receptors 1–12.

image file: c6ra00268d-s2.tif
Scheme 2 Structures of tripodal tris-urea receptors 13–36.

image file: c6ra00268d-s3.tif
Scheme 3 Structures of hexaurea receptors 37, 40–42, and pyridinium-based tris-urea receptors 38–39.

image file: c6ra00268d-s4.tif
Scheme 4 Structures of tripodal tris-thiourea receptors 44–57.

image file: c6ra00268d-s5.tif
Scheme 5 Structures of tripodal urea–amide hybrid receptors 58–65 and squaramide based tripodal receptors 66–68.

Design principles of tripodal receptors

Tripodal tris-amide receptors (Scheme 1) offer three –NH hydrogen bonds for the recognition of anionic guest. Tris(2-aminoethyl)amine (Tren)-based amide receptors (1–5), due to its small cavity size mostly recognize a spherical F/Cl within its molecular cavity. In comparison to Tren-based receptors (1–5), aryl-bridged tris-amide receptors (7–10) provide a larger cavity size and thereby, can encapsulate even larger anions such as AcO, NO3 and SiF62− in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1/2[thin space (1/6-em)]:[thin space (1/6-em)]2 or 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest stoichiometry. The cavity size of N-bridged tripodal amide receptors can be increased by certain structural modifications (6 and 11) to encapsulate larger anions such as, NO3, SiF62− and even discrete halide–water clusters.

On the other hand, tripodal tris-urea/thiourea receptors (Schemes 2 and 4) offer six –NH hydrogen bonds and hence, can recognize anions of different geometry including spherical (F/Cl/Br), trigonal planar (AcO/NO3), trigonal pyramidal (SO32−) and tetrahedral (SO42−/PO43−) anions in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1/2[thin space (1/6-em)]:[thin space (1/6-em)]2 or 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest stoichiometry. In general, the acidity of thiourea –NH proton is higher than that of urea (pKa of 21.1 and 26.9 in DMSO, respectively),44 and thus thiourea receptors are expected to establish stronger hydrogen bonds and form more stable complexes with anions than their urea analogues. However, if the –NH protons are acidic enough due to the incorporation of a highly electron-withdrawing substituent into the receptor framework; deprotonation may occur in the presence of a highly basic anion such as, F and HCO3. The use of thiourea groups offers additional advantages, including generally enhanced solubility in lower polarity solvents and a lower propensity to undergo self-association.

In 2005, theoretical investigation by Hay et al. showed that the incorporation of electron withdrawing substituents on the peripheral aryl function of tripodal receptors significantly enhance the stability of the anion complexes.45 Notably, a nitro group at the meta-position of the peripheral aryl function increases the acidity of the ortho-CH protons to participate in anion binding. The lipophilicity of the receptors can be significantly enhanced by incorporation of pyridyl or fluorinated aryl functions to the tripodal scaffold, to achieve anion recognition in aqueous or semi-aqueous medium. Attachment of peripheral pyridyl functions offers additional advantage of metal coordination, and hence, anion complexation can be achieved with metal salts in aqueous or semi-aqueous media. The lipophilicity of the receptors can also be increased by functionalization of each receptor side-arm with cationic moieties such as pyridinium functions. Further, due to the charged nature of the pyridinium ring, the –CH protons subsequent to the pyridinium nitrogen can act as potent hydrogen bond donor to the encapsulated anion.

Besides, a few tripodal hexaurea receptors have also been synthesized based on the concepts of complementarity and hydrophobic effect. The ortho-phenylene bridged tripodal hexaurea receptors (40–42) due to its tetrahedral cone conformation and 12 –NH hydrogen bond donors are ideal for tetrahedral oxyanion encapsulation.

Tripodal amide receptors

In the early nineties, Beer et al. has first reported the anion recognition properties of tripodal tris-amide receptors 1a–b (Scheme 1) having redox-active ferrocene and cobaltocene as the electrochemical signaling unit.46–48 Cyclic voltammetry and 1H NMR experiments revealed that the combinations of positively charged metallocene units together with amide groups are the essential components for the electrochemical anion recognition effects. Nonetheless, 1a could detect H2PO4 by large cathodic shifts of up to 240 mV in the presence of a tenfold excess of Cl and HSO4 ions in acetonitrile. During the same period, Reinhoudt et al. reported the anion binding properties of a series of neutral tripodal tris-amide receptors 2a–d (Scheme 1) by 1H NMR and conductometric experiments.49 The association constants for the anion complexation of 2a–d revealed a preferential binding to dihydrogenphosphate (H2PO4 > Cl > HSO4), and the highest binding affinity was observed for 2a (Table 1). They rationalized that the simple structure of the tripodal scaffold offers the possibility for synthetic manipulation to achieve higher association constants and anion selectivity. Since then, a range of N-bridged and aryl-bridged tripodal amide receptors have been developed to accomplish selective anion recognition, although selectivity has been achieved only in few cases. For example, a systematic study with a set of three positional isomers of nitro-phenyl functionalized tris-amide receptor 3a–c (Scheme 1) showed that the ortho-isomer (3a) is capable of selective fluoride recognition, as revealed from 1H NMR titration experiments (DMSO-d6) with different halides.50 The ortho-isomer experienced a large downfield shift of amide –NH resonance (Δδ = 1.05 ppm) upon titration with F, and yielded an association constant (log[thin space (1/6-em)]K) of 5.63 for a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 stoichiometry. However, the other two isomers (3b and 3c) were found to recognize both fluoride and chloride anions, with greater selectivity for fluoride (Table 1). In recent years, studies employing tripodal receptors generally highlight the structural aspects of anion binding along with detailed solution-state anion binding studies mostly using 1H NMR spectroscopy. Tris(2-aminoethyl)amine (Tren) based tris-amide receptors demonstrated the halide induced formation of monomeric capsules (semi-capsules) in its neutral form, whereas, aryl-bridged tris-amide receptors have been established to form either monomeric or dimeric capsules depending on the nature of the anion and/or terminal aryl function.
Table 1 Binding constants (log[thin space (1/6-em)]K) of amide receptors with different anions
Receptors F Cl Br HSO4 H2PO4
2a 3.24 2.23 3.78
3b 3.76 3.32
3c 4.06 2.29
5a 3.35 3.26 1.56
5b 3.94 2.87 2.66
9 1.59 1.58 1.56 1.96 0.87
10 4.86 3.83 2.97


Monomeric capsules (semi-capsules) of tripodal amide receptors

In our work with tripodal amide receptors, we have shown that the dinitrophenyl-functionalized tris-amide receptor 4 (Scheme 1) serves as an efficient fluoride sensor with characteristic solvent dependent absorptions in the optical spectrum.51 Single crystals of fluoride complex [TBA(4·F)] were obtained from a library of polar aprotic solvents viz. acetonitrile (MeCN), tetrahydrofuran (THF), dimethylformamide (DMF), dimethylsulfoxide (DMSO) and acetone (Me2CO) as solvates. Solvatomorphs of [TBA(4·F)] showed the encapsulation of a fluoride anion within the receptor cavity by six strong hydrogen bonds involving the amide –NH protons and three aryl –CH protons, irrespective of the solvent of crystallization (Fig. 1a). In solution state, the π-acidic cavity gives rise to intense colorimetric changes (colourless to red/blue) upon fluoride recognition, due to strong anion–π charge transfer interactions. 1H NMR analyses of the isolated fluoride complexes showed a significant downfield shift of the amide –NH and ortho-CH resonances (Δδ ∼3.50 and ∼0.70 ppm, respectively), suggesting a strong solution-state binding of fluoride. The high fluoride selectivity of 4 (log[thin space (1/6-em)]K > 7.0) has been employed in the transformation of charged anion complexes [(H4)+A] (A = Cl, Br, ClO4, HSO4) into [TBA(4·F)], as evidenced in 1H NMR titration experiments (DMSO-d6).52 It is to be noted that, attempted crystallization of a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 KF complex in solvents such as, DMSO and DMF yielded the respective solvates of 4.53 However, it is possible to crystallize the KF complex from an acetonitrile media with the composition [4·KF(H2O)2], where a hydrated KF contact ion-pair is coordinated to the receptor molecule(s) outside the tripodal cavity.53 In the 1H NMR spectrum of KF complex (CD3CN/DMSO-d6), the amide –NH resonance could not be observed and the ortho-CH resonance showed a comparatively minor downfield shift of ∼0.30 ppm in comparison to [TBA(4·F)]. Thus, a binding discrepancy of fluoride in quaternary ammonium (TBAF) and alkali (KF) salts has been demonstrated both in solid and solution-states by tris-amide receptor 4.
image file: c6ra00268d-f1.tif
Fig. 1 X-ray structures of monomeric capsules, (a) [4·(F)], and (b) [5a·(F·H2O)]. TBA cations are omitted for clarity. Colour code: C = green, N = blue, O = red, F = chartreuse yellow.

Another interesting example of fluoride binding within a tripodal tris-amide scaffold has been established with a pentafluorophenyl-based receptor, 5a reported by Ghosh et al.54,55 Solvent sealed monomeric F capsule [TBA(5a·F)S] (S = H2O/CHCl3), can either be obtained from an aqueous or chloroform solution. Whereas, solvent sealed monomeric Cl capsule [TBA(5a·Cl)CHCl3], can only be obtained from a chloroform solution and an aqueous solution give non-solvated complex [TBA(5a·Cl)]. In the solvent sealed monomeric capsules, a solvent molecule (H2O/CHCl3) is coordinated to the encapsulated halide anion (Fig. 1b). Notably, 5a together with 18-crown-6-ether has efficiently been employed in the liquid–liquid extraction (H2O/CHCl3) of KF and KCl from aqueous solution by a dual host approach.

Sun et al. has demonstrated the reversible encapsulation of nitrate anion within the preorganized cavity of a series of N-bridged tripodal amide receptors 6a–6k (Scheme 1), via solvent polarity controlled formation of dimeric receptor capsules.56,57 The structural scaffold responsible for the capsule formation was prepared by in situ nitration of the N-terminal aromatic ring of the tripodal receptors bearing methyl (6a) and methoxy (6b) substituents at para-position. The studies showed that the nitro group at the ortho-position of the N-terminal aromatic ring is mandatory for hydrogen bonded capsule formation and the mode of hydrogen bonding depends upon the electronic nature of the N-terminal aromatic ring. In the crystal structure of a nitrate-encapsulated complex [(H6a)+NO3], the amide –NH protons form strong hydrogen bonds with the nitrate anion (Fig. 2) and, the etheric oxygen atoms are hydrogen bonded to the proton attached to the bridgehead N-atom. They have also demonstrated that the incorporation of dansyl groups at the periphery of the tripodal amide scaffold resulted in a nitrate selective fluorescent probe (6g) in an aqueous DMSO solution.58 They rationalized that the ionization of sulfonic acid in the dansyl moiety assists the transformation of the fluorescent receptor into its zwitterion form by protonation of the bridgehead nitrogen and thereby, attain a cone shape conformation through intramolecular hydrogen bonding.


image file: c6ra00268d-f2.tif
Fig. 2 X-ray structure of complex [(H6a)+(NO3)]. Colour code: C = green, N = blue, O = red.

Ghosh et al. and coworkers have further explored the anion recognition properties of two positional isomers of an aryl-bridged tris-amide receptor having ortho-nitrophenyl (7) and para-nitrophenyl (8) peripheral functions (Scheme 1).59,60 The aryl-platform provides a larger preorganized cavity in comparison to the Tren-based receptors and thereby, enables the encapsulation of hydrated anions. The ortho-isomer 7 showed encapsulation of a monohydrated chloride/acetate within the tripodal cavity in complexes [TBA(7·A)H2O] (A = Cl/AcO), while monotopic recognition of anion has been observed in the nitrate complex [TBA(7·NO3)] (Fig. 3).59


image file: c6ra00268d-f3.tif
Fig. 3 X-ray structures of monomeric capsules, (a) [7·(AcO·H2O)], and (b) [7·(NO3)]. Colour code: C = green, N = blue, O = red. TBA cations are not shown.

The same group has also reported the detailed solution-state isothermal calorimetry (ITC) studies of a pentafluorophenyl-based tris-amide receptor 10 with halides and oxyanions.61 Fluoride, chloride, bromide, acetate and benzoate showed an exothermic binding profile with 1[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest stoichiometry. The binding of bromide and acetate is equally facilitated by entropy and enthalpy factors, whereas chloride and benzoate binding is strongly enthalpy driven. Structures of halide complexes [TBA(10·X)] (X = F/Cl), has further confirmed the 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding via encapsulation of an anion within the tripodal cavity by 3 –NH hydrogen bonds.

A cationic tripodal receptor 12, based on amide-pyridinium as recognition sites and nitrobenzene as signaling unit showed high selectivity and strong binding affinity for AcO (log[thin space (1/6-em)]K > 5.0) over other anions.62 UV-Vis and 1H NMR experiments indicated that the selectivity can be attributed to synergistic effects arising from hydrogen bonding, electrostatic interactions and conformational change.

Dimeric capsules of tripodal amide receptors

To gain further insight into the anion binding properties of aryl-bridged tris-amide scaffold, the para-isomer 8 was crystallized in presence of fluoride, chloride, acetate, and nitrate, to obtain the respective complexes from dioxane solutions.60 Fluoride complex [TBA(8·F)(H2O)3], showed the formation of a dimeric capsular assembly via encapsulation of a fluoride–water cluster [F2(H2O)6]2− (Fig. 4a). Similarly, chloride and acetate complexes showed encapsulation of [A2(H2O)4]2− (A = Cl/AcO) anion–water cluster within the dimeric capsular assembly of 8. Each capsular assembly is stabilized by several receptor–anion, receptor–water and anion–water interactions. Nitrate complex [TBA(8·NO3)], also showed the formation of a dimeric capsule (Fig. 4b). The capsular size of acetate and nitrate complexes (10.24 Å and 10.02 Å, respectively) is significantly larger than the fluoride and chloride complexes (9.45 Å and 9.56 Å, respectively).
image file: c6ra00268d-f4.tif
Fig. 4 X-ray structures of dimeric capsules, (a) [(8)2·(F)2(H2O)6], (b) [(8)2·(NO3)2]. Colour code: C = green/gray, N = blue, O = red, F = chartreuse yellow. TBA cations are not shown.

Although receptor 7 did not showcase the formation of dimeric capsules upon complexation with anions such as, chloride, acetate and nitrate, however, hydrated fluoride ion [F2(H2O)6]2− induced formation of dimeric capsule has been observed in complex [TBA(7·F)(H2O)3].59 Thus, simple variation in the position of the electron withdrawing nitro-substituent in the aryl terminals of the tripodal amide receptor showed substantial alterations in the anion/hydrated anion induced capsular assembly formation. The 1H NMR titration experiments (DMSO-d6) with AcO showed 1[thin space (1/6-em)]:[thin space (1/6-em)]3 host–guest binding for receptor 7 and 8, whereas the single-crystal structure showed 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complex.

Another aryl-bridged tris-amide receptor with a pyridyl terminal 9 (Scheme 1) has recently been reported by the same group, which showed encapsulation of hydrated halides [X2(H2O)4]2− (X = F/Cl) within its dimeric capsular assembly when crystallized from 1[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) aqueous acetone solution.63 On the other hand, attempted complexation with fluoride in 1[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) dioxane–acetone solvent system resulted in the formation of [SiF6(H2O)2]2− encapsulated dimeric assembly. The binding constants calculated from the 1H NMR titration experiments with halides and oxyanions revealed that the receptor shows a comparable binding constant with halides (log[thin space (1/6-em)]K ∼ 1.56), whereas NO3 and HSO4 show slightly higher values (log[thin space (1/6-em)]K ∼ 2.0) (Table 1).

Recently, we have shown the encapsulation of a cyclic halide–water tetramer [X2(H2O)2]2− (X = Br/Cl) inside the dimeric capsular assembly of a new N-bridged tripodal amide receptor 11 (Scheme 1) with para-nitrophenyl terminals.64 The encapsulated halide–water tetramer in complexes [(H11)·X·H2O] is hydrogen bonded to both the receptor cations of the dimeric capsule of size ∼20.60 Å (Fig. 5a). It is important to mention that the etheric oxygen atoms are hydrogen bonded to the proton attached to the bridgehead-N of receptor cation, which eventually organizes the flexible receptor arms in one direction and drives the formation of capsular assembly. However, the receptor cation prefers to adopt a non-capsular polymeric aggregation in the presence of higher homologous iodide, which confirms the anion selective conformational adaptability of the molecule.


image file: c6ra00268d-f5.tif
Fig. 5 X-ray structures of dimeric capsules, (a) [(H11)2+·(Br·H2O)2], and (b) [(H11)2+·(SiF6)2−]. Colour code: C = green, N = blue, O = red, Br = brown, F = chartreuse yellow.

Further studies with receptor 11 have revealed the solvent dependent encapsulation of an octahedral hexafluorosilicate anion within its dimeric capsular assembly in the solid-state.65 Two hexafluorosilicate complexes [(H11)2(SiF6)] and [(H11)2(SiF6)·2DMF·4H2O] were obtained upon reaction of 11 with HF in DMSO and DMF respectively, presumably as a result of glass corrosion. In the non-solvated complex crystallized from DMSO, two protonated receptor molecules create a dimeric capsular assembly of 19.33 Å to encapsulate a hexafluorosilicate anion by multiple hydrogen bonds (Fig. 5b). Whereas, in the solvated complex crystallized from DMF, a hexafluorosilicate anion is involved in side-cleft binding with four receptor cations, outside the receptor cavity. Thus, by a mere change of crystallizing solvent, we have been able to isolate two structurally different SiF62− complexes of receptor 11. It is important to mention that the capsular size of [(H11)2(SiF6)] complex is less by ca. 1.30 Å than the [(H11)·X·H2O] (X = Cl/Br) complexes.

Tripodal urea receptors

After the seminal papers by Wilcox66 and Hamilton67 on urea–anion interactions, a variety of anion receptors have been reported in which one or more urea/thiourea fragments are incorporated in an acyclic or macrocyclic framework. In 1995, Morán et al. has demonstrated that the Tren-based tris-phenylurea receptor 13 (Scheme 2) and its thiourea analogue 44 (Scheme 4) can efficiently bind tetrahedral oxyanions due to topological complementarity and directionality of hydrogen bonding.68 Based on the 1H NMR experiments, he suggested a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complex formation upon binding of a dihydrogenphosphate anion within the tripodal cavity via six hydrogen bonds. Based on similar ideas, Wu et al. has reported the recognition of tetrahedral oxyanions (H2PO4 and HSO4) by complexation-prompted fluorescence enhancement of the Tren-based tris-naphthylurea receptor 14 (Scheme 2) in DMF solution.69 He has also reported the thiourea analogue 45 (Scheme 4) of the naphthylurea receptor, which showed substantial absorption spectral changes upon complexation with H2PO4 and HSO4 in DMF solutions.70 The association constants obtained by fluorescence/UV-Vis titrations showed a higher selectivity for the H2PO4 anion with a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complex stoichiometry for both the receptors.69,70 However, no structural data were available to support the host–guest complementarity and the shape recognition for these tetrahedral oxyanions. Nonetheless, following these solution-state anion binding studies and theoretical calculations by Hay et al.,45 a number of tripodal urea and thiourea receptors mostly with electron withdrawing aryl terminals have been developed for the recognition and separation of anionic guests, particularly sulfate and phosphate. Most of these studies highlight the structural aspects of anion binding with the tripodal urea/thiourea receptors, which were then related to the observed selectivity in solution. Oxyanions such as, CO32−, SO42− and HPO42− induced formation of dimeric capsular assembly and halide induced formation of monomeric capsules have commonly been observed in most of the structural reports on anion complexes of tripodal urea/thiourea receptors. However, oxyanion binding within the unimolecular receptor cavity has also been perceived in certain cases.

Dimeric capsules of tripodal urea receptors

Tren-based tris-urea receptor 15 with peripheral 3-cyanophenyl group (Scheme 2) has been recognized as a versatile anion receptor accounting for the selective separation of SO42− over other anions via selective crystallization of Ag+ coordinated framework,71 encapsulation of staggered conformer of divalent oxalate anion,72 and uptake of aerial CO2 as CO32− complex that efficiently extracts sulfate, thiosulfate and chromate from water by anion exchange metathesis.73

Complexation of 15 with 0.5 equivalents of Ag2SO4 in water/acetone binary solvent mixture yielded a coordination polymer of capsular aggregate having the composition [Ag2SO4(15)2](Me2CO)1.5(H2O)3.7, reported by Custelcean et al.71 Two receptor molecules that encapsulate a SO42− anion are held together by CN–Ag+ and C[double bond, length as m-dash]O–Ag+ coordinative bonds to form a rigidified capsule of 9.81 Å size. Attempted complexation of 15 with other soluble silver salts having anions of different geometry and basicity such as, AgBF4, AgNO3, AgMeSO3 and AgMeCO2 failed to produce any coordination complexes and yielded crystals of 15 in each case.

The same cyanophenyl-based receptor 15 has also been utilized by Ghosh et al. to capture the staggered conformer of divalent oxalate anion within its dimeric capsular assembly of 9.81 Å size (Fig. 6a).72 They have further showed that the planar conformer of the divalent oxalate anion can be captured with the fluorophenyl-based receptor 19. Encapsulation of planar form of the anion resulted in an increase of capsular size by ∼1.0 Å due to the more C–C single bond character of planar C2O42− than its staggered conformer.


image file: c6ra00268d-f6.tif
Fig. 6 X-ray structures of oxyanion encapsulated dimeric assemblies, (a) [(15)2·(C2O4)2−], (b) [(15)2·(CO3)2−], (c) [(15)2·(SO4)2−], and (d) [(15)2·(CrO4)2−]. Colour code: C = green/gray, N = blue, O = red, S = yellow, Cr = purple. TBA cations are not shown.

Thus, recognition of different conformers of C2O42− inside the dimeric capsule of two different tripodal receptors has been accomplished by simple tuning of the peripheral aromatic substituents. 1H NMR titration data showed that 15 has slightly higher binding affinity towards C2O42− (log[thin space (1/6-em)]K = 4.82) compared to 19 (log[thin space (1/6-em)]K = 4.29).

However, the most interesting property of 15 is its ability to capture atmospheric CO2 as CO32− within its dimeric capsular assembly. The aerial CO2 uptake is facilitated in a DMSO solution of the receptor containing equivalent amount of (TBA)OH, which eventually form diamonded crystals of CO32− encapsulated complex [(TBA)2CO3(15)2] (Fig. 6b) in about 85% yield.73 Because of its excellent solubility in chloroform and dichloromethane, [(TBA)2CO3(15)2] has been efficiently employed for the extraction of SO42−, S2O32− and CrO42− from water by anion-exchange metathesis.73

Quantification by weighing the bulk extract shows that the CO32− capsule can separately extract ca. 90% of the above three anions from aqueous solutions. In each case, the bulk extracted solid was crystallized from DMSO to obtain 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest capsular complexes [(TBA)2SO4(15)2] (Fig. 6c), [(TBA)2S2O3(15)2] and [(TBA)2CrO4(15)2] (Fig. 6d), of identical capsular size (9.61–9.71 Å). The extraction of SO42− ions from water has also been demonstrated under alkaline conditions (pH 12.5) and in the presence of excess nitrate ions.

Tris(3-pyridylurea) receptor 17 upon complexation with Mg2+ salts of divalent oxyanions in aqueous methanol solutions, self-assemble into crystalline capsules with composition [Mg(H2O)6·(17)2A] (A = SO42−, SeO42−, CO32−, SO32−), as reported by Custelcean et al.74 Competitive crystallization experiments in the presence of these anions established the anion selectivity order, SO42− > SeO42− > CO32− > SO32− which is different from both the Hofmeister bias and anion basicity scale, whereas the trend was in agreement with the lattice energy calculations on the crystal structures. The lower selectivity of the [Mg(H2O)6]2+ bridged capsule for planar CO32− and pyramidal SO32− was attributed to the mismatch of the anion binding cavity with these anions. Oxyanion encapsulation within the urea-lined dimeric cavity in a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest stoichiometry is characteristic to the solid-state, as only 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complexation could be detected in solution. The selective separation of SO42− from an aqueous solution containing excess of NO3 has also been achieved with this capsule. However, the capsule has limited utility for sulfate separation from alkaline nuclear wastes, as they do not form under basic conditions (pH > 10), due to Mg(OH)2 precipitation. Structurally similar [M(H2O)6]2+ (M = Mn/Zn) bridged dimeric SO42− capsules have also been reported by Wu and Janiak et al.75

It has been realized that, the oxyanion binding cavity in [Mg(H2O)6]2+ bridged capsule is large enough to accommodate either SO42− (Fig. 7a) or SeO42−, which resulted in poor selectivity for sulfate against selenate. Thus, to improve the selectivity for SO42− against SeO42−, the anion binding cavity has been fine-tuned by substituting cationic [Mg(H2O)6]2+ subunits with [Li(H2O)]+, which yielded a coordination capsule [Li2SO4(17)2(H2O)2] (Fig. 7b) with substantially reduced cavity size (capsule size drops by ∼0.50 Å).76 Such allosteric regulation of the cavity size by the judicious choice of external bridging cation has allowed for the effective separation of SO42− over SeO42− by competitive crystallization, with an observed selectivity exceeding that of sulfate-binding protein. Thus, Custelcean et al. has been able to establish both shape recognition of SO42− against SO32− and CO32−, and size recognition against SeO42− using tripodal urea receptor.


image file: c6ra00268d-f7.tif
Fig. 7 X-ray structures of sulfate encapsulated metal complexes, (a) [Mg(SO4)(H2O)6·(17)2], (b) [Li2(SO4)(H2O)2(17)2]. Colour code: C = green, N = blue, O = red, S = yellow, Mg = pink, Li = purple.

Custelcean et al. has also demonstrated sulfate separation from a highly competitive aqueous alkaline solution (∼6 M Na+, pH = 14) by selective crystallization of sodium-based coordination capsule [Na2SO4(17)2(H2O)4] and thereby, provided for the first time a viable approach to sulfate separation from nuclear wastes.77 The alkali metal coordination capsules are held together by coordination and hydrogen bonding water bridges, with a SO42− anion encapsulated inside urea-lined dimeric cavity.

Very recently, Wu et al. has reported the tris(4-pyridylurea) receptor 18 and its self-assembly with various metal chloride salts.78 The receptor has shown the formation of 2[thin space (1/6-em)]:[thin space (1/6-em)]2 dimeric capsules with chloride salts of alkali (Na/K) and alkaline earth metals (Mg/Ca), where two Cl encapsulated semi-capsule formed a staggered dimeric capsule by hydrogen bonding interactions between the peripheral pyridyl nitrogen and water molecules of the cationic [M2(H2O)6]2+/[M(H2O)6]2+ subunit. However, formation of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest dimeric capsules were observed with carbonate salts of alkali metals (Na/K). The sizes of the (Cl)2 encapsulated coordination capsules of 18 (11.18–11.33 Å) are significantly larger than that of oxyanion (CO32−/SO42−) encapsulated coordination capsules of 17 and 18 (9.20–9.96 Å).

A pentafluorophenyl-substituted tris-urea receptor 21 has been extensively studied by Ghosh et al. and established as a versatile receptor for anions of different geometry.79–83 Combined 1H NMR, isothermal titration calorimetry (ITC) and single crystal X-ray diffraction studies revealed that monovalent anions such as, F, OH and H2PO4 formed 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complexes, and divalent oxyanions such as, CO32−, SO42− and HPO42− formed 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest capsular complexes. It is interesting to note that, a DMSO solution of 21 upon treatment with (TBA)CN formed crystals of hydroxide-encapsulated complex [TBA(21·OH)].79 Two hydroxide-encapsulated receptor units are held together by intermolecular π–π interactions to form a pseudo-dimeric capsule of 14.95 Å. It is to be noted that, no crystals were formed from a DMSO solution of 21 and (TBA)CN in anaerobic condition, suggesting that atmospheric moisture must have a role in the formation of OH complex. Pseudo-dimeric capsule formation has also been observed in fluoride complex [TBA(21·F)] via halogen bonding interactions between the pentafluorophenyl rings.80 In dihydrogenphosphate complex [TBA(21·H2PO4)], a (H2PO4)2 dimer is encapsulated within a dimeric assembly of 13.79 Å, demonstrating the formation of 2[thin space (1/6-em)]:[thin space (1/6-em)]2 host–guest capsular complex in the solid-state (Fig. 8a).81 However, hydrogenphosphate complex [(TBA)2HPO4(21)2], obtained from a DMSO solution of the receptor containing TBA salts of H2PO4 and OH ions, showed encapsulation of a HPO42− anion within a dimeric assembly of 9.92 Å (Fig. 8b).79 Thus, the receptor has been shown to encapsulate both monovalent and divalent phosphate anion(s) within dimeric capsular assemblies of different sizes. Arsenate-encapsulated complex [(TBA)2HAsO4(21)2] having capsular size 10.22 Å, was obtained from a receptor solution containing 2 equiv. of TBAI and excess sodium hydrogen arsenate in a 5% H2O–DMSO solvent system.82 Thus, the authors were able to furnish the first structural evidence of arsenate encapsulation by a tripodal urea receptor.


image file: c6ra00268d-f8.tif
Fig. 8 X-ray structures of phosphate encapsulated dimeric assemblies, (a) [(21)2·(H2PO4)2], and (b) [(21)2·(HPO4)2−]. Colour code: C = green/gray, N = blue, O = red, P = orange, F = chartreuse yellow. TBA cations are not shown.

The most exciting property of 21 is its ability to capture atmospheric CO2 as CO32− encapsulated complex [(TBA)2CO3(21)2] in almost quantitative yield (∼98%), from a DMSO solution of the receptor containing equivalent amount of (TBA)OH.83 The disassembly of the CO32− capsule and recovery of the free receptor (∼85%) has been accomplished by simply treating the complex with aqueous methanol. As expected, no crystals were formed from a DMSO solution of 21 and (TBA)OH under anaerobic condition in a glove box.

The carbonate complex [(TBA)2CO3(21)2] has been efficiently employed in the liquid–liquid (chloroform–water) extraction of SO42− from water via an anion exchange process. Almost quantitative and clean extraction of SO42− from water (99% from extracted pure mass and >95% shown by gravimetrically) has been achieved by this process.79 Selective SO42− extraction from water, using CO32− capsule has also been demonstrated in the presence of other competitive oxyanions such as, H2PO4 and NO3. The composition of the extracted mass was confirmed by 1H NMR and single crystal X-ray diffraction, which showed sulfate-encapsulated dimeric assembly [(TBA)2SO4(21)2] of 9.18 Å. It is to be noted that, the size of the capsular assembly did not change upon exchange of tetrahedral SO42− for trigonal-planar CO32−, which perhaps plays a key role in the facile extraction of SO42− due to topological complementarity with the urea-functionalized anion-binding cavity. However, an impure SO42− extraction has been observed when an organic layer containing an equivalent mixture of 21 and (TBA)Cl (phase transfer agent) was used as an extractant, further validating the significance of persistent dimeric capsule in SO42− extraction and separation.

Detailed 1H NMR titration experiments of 21 with halides and oxyanions in DMSO-d6 solutions yielded the anion selectivity order as: H2PO4 > SO42− > CH3COO > F > Cl (Table 2). Similarly, isothermal titration calorimetric (ITC) studies in dry DMSO solutions showed the highest binding affinity towards H2PO4 and the anion binding affinity follows the order: H2PO4 > SO42− ∼ F > CH3COO > Cl, which is nearly in agreement with the 1H NMR titration results.80,81 The anion binding affinity of the receptor has also been studied in aqueous DMSO (DMSO-d6/D2O, 9[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v) by performing 1H NMR titrations with 1[thin space (1/6-em)]:[thin space (1/6-em)]1 D2O/DMSO-d6 solutions of various anions as their sodium salts.82 A 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding stoichiometry has been observed with all investigated oxyanions e.g. HAsO42−, HPO42−, SO42−, CO32−, and showed the highest affinity for HAsO42− (log[thin space (1/6-em)]K = 4.42) followed by H2PO42− (log[thin space (1/6-em)]K = 3.62) > SO42− (log[thin space (1/6-em)]K = 3.48) > CO32− (log[thin space (1/6-em)]K = 2.68).

Table 2 Binding constants (log[thin space (1/6-em)]K) of urea receptors with different anions
Receptors F Cl H2PO4 SO42− AcO
a K determined by 1H NMR titration experiments.b K determined by UV-Vis/fluorescence titration experiments.
13a 2.94 4.04 >4.0
14b 4.60 3.43
16a 4.51 3.09 4.20 4.70
20a 2.75 2.65 >4.0
21a 4.06 3.42 5.52 4.73 4.45
22a 2.60 2.38 >4.0
25a 3.87 4.38 4.78 3.27
26a 4.41 4.97 3.21
27b 5.31 4.28 4.62 6.70 4.95
33a 3.38 4.39 5.74 3.41


Gale et al. has reported the transmembrane anion transport properties of a series of fluorinated tripodal tris-urea receptors 20–23 (Scheme 2) and their thiourea analogues 46–49 (Scheme 4), along with the detailed solid- and solution-state anion binding studies.84 Based on the experimental results obtained, he has suggested that increasing the degree of fluorination increases the lipophilicity of the tripodal receptors and this has shown to be the major contributing factor in the superior transport activity of the fluorinated compounds, with a maximum transport rate achieved with compound 23. Vesicle anion transport assays using ion-selective electrodes showed that, this class of fluorinated urea/thiourea compounds are capable of transporting chloride through a lipid bilayer via a variety of mechanisms, including Cl/H+ cotransport and Cl/NO3, Cl/HCO3 antiport processes. The 1H NMR studies revealed that the fluorinated receptors can bind anions in DMSO-d6 solutions according to the trend SO42− > H2PO4 > Cl > HCO3 ≫ NO3 (Table 2), with various deprotonation events occurring upon titration with H2PO4 and HCO3. Receptor 23 showed the formation of a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest capsular complex [(TEA)2CO3(23)2] of size 9.16 Å, when crystallized in presence of (TEA)HCO3.

Recently, a set of three positional isomers of nitro-phenyl functionalized tris-urea receptor (24–26) has been extensively studied by our group, towards recognition of various oxyanions.85–87 The para-isomer 26 has previously been shown to form a large capsular complex with SO42− via encapsulation of a rugby ball shaped (SO42−·3H2O·SO42−) adduct within a dimeric receptor capsule assembled by electrostatic nitro–nitro interactions (Fig. 9a).88 In our studies, we have shown that the para-isomer can encapsulate a carbonate or a divalent terephthalate anion in a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest stoichiometry.85 The terephthalate complex [(TBA)2C8H4O4(26)2] was obtained from a DMSO solution of 26 containing excess of 1[thin space (1/6-em)]:[thin space (1/6-em)]2 mixture of terephthalic acid and (TBA)OH, while carbonate complex [(TEA)2CO3(26)2] was obtained from a DMSO solution of 26 containing excess of (TEA)HCO3. In both the complexes, a total of 14 –NH hydrogen bonds stabilize an anion within the respective capsular assemblies (Fig. 9b). Interestingly, the para-isomer in the presence of excess TBA(H2PO4) self-assembled into a tetrahedral molecular cage by encapsulation of a tetrameric tetrahedral mixed phosphate cluster (H2PO4·HPO42−)2 via 24 –NH hydrogen bonds (Fig. 10a).87 In the tetrahedral cage complex [(26)4(H2PO4·HPO42−)2], each of the four receptor units coordinate to a phosphate anion (H2PO4/HPO42−) axially by six –NH hydrogen bonds donated from the three urea functions (Fig. 10b). Further, the fluoride complex [TEA(26·F)] of the para-isomer showed the formation of a self-assembled (2 + 2) pseudodimeric cage with 1[thin space (1/6-em)]:[thin space (1/6-em)]1 host[thin space (1/6-em)]:[thin space (1/6-em)]guest stoichiometry.


image file: c6ra00268d-f9.tif
Fig. 9 X-ray structures of oxyanion encapsulated dimeric assemblies, (a) [(26)2·[(SO42−)2·(H2O)3], (b) [(26)2·(C8H4O4)2−]. C = green/gray, N = blue, O = red, S = yellow. TBA cations are not shown.

image file: c6ra00268d-f10.tif
Fig. 10 X-ray structures of tetrahedral cage complex [(26)4(H2PO4·HPO42−)2], (a) tetrameric (H2PO4·HPO42−)2 encapsulated tetrahedral cage of 26, and (b) hydrogen bonding interactions between (H2PO4·HPO42−)2 cluster and a receptor unit. C = green/gray/yellow/pink, N = blue, O = red, P = orange. TBA cations are not shown.

Unlike 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complexation of SO42− by the para-isomer 26, the meta-isomer 25 can encapsulate SO42− in a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest stoichiometry to form a dimeric capsule [(TBA)2SO4(25)2].86 However, the meta-isomer is incapable to self-assemble into a dimeric capsule in presence of terephthalate dianion, and adopts a flat trigonal planar like geometry where each receptor side arm is in interaction with a carboxylate group of the dianion in complex [(TBA)3(C8H4O4)1.5(25)]. The most important property of the meta-isomer is its ability to capture atmospheric CO2 as CO32− to form capsular complex [(TBA)2CO3(25)2], triggered by the presence of (TBA)OH/(TBA)F in a DMSO solution.86 It is remarkable to note that above 90% yield of CO32− encapsulated complex can be obtained from the basic (OH) solution mixture after 3 days of exposure to an unmodified atmosphere. Structural comparison revealed an increase of 1.0 Å in the capsular size of carbonate complex of the meta- isomer (9.0 Å) as compared to its para-analogue (8.0 Å), although both the capsular assemblies has the same symmetry and encapsulate a CO32− with an array of 14 –NH hydrogen bonds, which is indeed a consequence of positional isomeric effect. The effect of positional isomerism has also been observed for sulfate binding by the isomeric receptors, where the para-isomer in presence of SO42− formed a large dimeric capsule of 15.77 Å engulfing a (SO42−·3H2O·SO42−) adduct, whereas the meta-isomer formed a SO42− encapsulated dimeric assembly of 9.60 Å. Above all, the terephthalate complexes of the isomeric receptors revealed the greatest evidence of positional isomerism, where the para-isomer formed a dimeric capsular assembly of 12.25 Å and the meta-isomer formed a terephthalate coordinated polymeric assembly.

The meta-isomer has also been shown to encapsulate HPO42− within a dimeric assembly of 9.86 Å, assembled by π–π interactions of the nitrophenyl aromatics.85 The complex [(TBA)2HPO4(25)2] was obtained from a DMSO solution of the receptor containing equal amounts of excess (TBA)H2PO4 and (TBA)OH, since crystallization of 25 in the presence of excess (TBA)H2PO4 was not fruitful. However, successful crystallization of the ortho-isomer 24, in the presence of TBA salts of different oxyanions was not fruitful presumably due to the steric effect provided by the nitro groups at the ortho-position, which hinders the facile inclusion and coordination of an anion due to electrostatic factor, as confirmed by 2D NOESY NMR analysis of the free receptor.

Recently, we have reported a tripodal heteroditopic tris-urea receptor 29 which is capable of both anion and cation recognition.89 The receptor in the presence of (TEA)HCO3 self-assemble into a molecular capsule having small cavity size of 3.97 Å via intermolecular CH–π interactions and carbonate coordination (Fig. 11a).


image file: c6ra00268d-f11.tif
Fig. 11 X-ray structures of complexes (a) [29·(TEA)2CO3] and (b) [29·K2HPO4]. Colour code: C = green/gray, N = blue, O = red, P = orange, K = purple.

Interestingly, the receptor when treated with K2HPO4 in a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 DMSO-H2O solution, afforded crystals of K+ coordinated centrosymmetric capsule assembled by weak π–π interactions of the phenyl rings, and resulted in three C2-symmetric urea-lined clefts for the recognition of HPO42− anion (Fig. 11b). The binding constants calculated from 1H NMR titration experiments with different oxyanions showed the anion selectivity order as HCO3 (log[thin space (1/6-em)]K = 2.88) > H2PO4 (log[thin space (1/6-em)]K = 2.77) > HSO4 (log[thin space (1/6-em)]K = 2.07) for 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complexation.

The tripodal urea receptors 30–33 on a cyanuric acid-platform has been shown to display different modes of sulfate coordination by Ghosh et al.90–92 SO42− encapsulation within dimeric assembly of 30 and 31 were observed in the presence of MgSO4 and (TBA)2SO4 respectively. Interestingly, 30 in the presence of ZnSO4·7H2O formed a 3D coordination polymer [(30)3Zn3(SO4)4][Zn(H2O)6][(H2O)18], the secondary building unit (SBU) of which constitutes four Zn2+ coordinated receptor units in a bowl shaped conformation encapsulating a sulfate–water [(SO4)4(H2O)12]8− cluster in its cavity (Fig. 12a).90 Binding constant calculation yielded log[thin space (1/6-em)]K value of 5.47 for 30 with ZnSO4 in DMSO-d6. In contrast, receptors 32 and 33 showed a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding with SO42−, which will be discussed in the next section.91,92


image file: c6ra00268d-f12.tif
Fig. 12 X-ray structures of (a) the secondary building unit (SBU) of [(30)3Zn3(SO4)4][Zn(H2O)6][(H2O)18], (b) tetrahedral cage complex [(37)4(PO43−)4]. Colour code: C = green/gray/yellow/pink, N = blue, O = red, P = orange.

Davis et al. has recently reported three new families of tripodal urea receptors, 34, 35 and 36 (Scheme 2) and their thiourea analogues (55, 56 and 57, Scheme 4) for transmembrane anion transport studies using vesicle anion transport assays.93,94 For biologically relevant anion transport the major target is chloride, so the binding properties of these receptors were assessed for chloride (Bu4N+Cl) by 1H NMR titrations in DMSO-d6. It was found that, the use of more electron-withdrawing aryl groups increases the chloride affinities of all three series, while the thioureas are nearly twice as powerful as the ureas. The Cl binding affinity of the urea receptors are typically in the range of log[thin space (1/6-em)]K 1.40–2.60. All of these results will be discussed in detail under the “thiourea receptor” section.

Wu et al. has reported an interesting example of self-assembly where a hexaurea receptor 37 (Scheme 3) in presence of PO43− self-assemble into a tetrahedral anion cage [(37)4(PO4)4]12− held together by 48 hydrogen bonds (Fig. 12b).95 In the complex, the phosphate ions occupy the vertices of a tetrahedron and the ligands lie on the faces with an estimated internal volume of 181 Å and PO43−–PO43− distance of 15 Å. The complex is stable up to 80 °C and can persist in different aprotic solvents such as, acetone, acetonitrile and DMSO.

Monomeric capsules (semi-capsules) of tripodal urea receptors

The para-isomer of the cyanophenyl-based tris-urea receptor 16 has been extensively studied by Hossain et al. for halide and oxyanion binding using 1H NMR spectroscopy, X-ray structure analyses of the protonated receptor complexes and density functional theory (DFT) calculations.96 The 1H NMR titration studies revealed that the receptor forms 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complexes with all studied anions, showing a binding trend in the order of F > H2PO4 > Cl ∼ HSO4 > Br (Table 2). DFT calculations showed formation of monomeric halide capsules, revealing the highest binding energy for the fluoride complex. Likewise, the para-isomer of the fluorophenyl-based tris-urea receptor 20 has revealed the formation of monomeric chloride capsule [TBA(20·Cl)].84

In the oxyanion-encapsulated metal (M = Mg/Mn/Zn) complexes of tris(3-pyridylurea) receptor 17, it has been observed that the pyridyl groups of the tripodal receptor do not coordinate with the metal ion, but interact with the hydrated metal cation [M(H2O)6]2+ through hydrogen bonding.74,75 These results implied that the oxyanion capsules may also be generated via second-sphere coordination, i.e. by employing metal complexes of hydrogen bond donating ligands having free counter anion(s). Based on this concept, Wu, Janiak and Yang et al. has employed [Fe(DABP)3]SO4 (DABP = 5,5′-diamino-2,2′-bipyridine) and [Fe(BP)3]SO4 (BP = 2,2′-bipyridine) complexes to self-assemble with 17.97 However, crystal structures of both the isolated complexes [Fe(DABP)3][17·SO4]·10H2O and [Fe(BP)3][17·SO4]·9H2O, showed formation of monomeric sulfate capsules (Fig. 13a).


image file: c6ra00268d-f13.tif
Fig. 13 X-ray structures of monomeric capsules, (a) [17·(SO4)2−], and (b) [27·(TBA)(SO4)]. C = green/gray, N = blue, O = red, S = yellow, Fe = orange.

The encapsulation of sulfate by a single receptor molecule rather than two is rationalized by the better crystal packing of the cationic units [Fe(DABP)3]2+ or [Fe(BP)3]2+ without sufficient secondary hydrogen bond donors and anion-encapsulated receptor units.

Integration of the well-defined tripodal urea scaffold and the redox-active ferrocene group has resulted in an electrochemical sensor 27 for SO42− and H2PO4 anions, as demonstrated by Wu and Yang et al.98 Large cathodic shifts of both the reduction and oxidation peaks were observed after the addition of 1.0 equivalents of SO42−Epc = −190 and ΔEpa = −150 mV) to a 1.0 mM chloroform solution containing 0.10 M (TBA)PF6 salt. Sulfate complex [(TBA)2SO4(27)H2O], showed the formation of a cation (TBA) sealed semi-capsule, where the TBA cations provide –CH hydrogen bonds to the encapsulated SO42− anion (Fig. 13b). Whereas, fluoride complex [TBA(27·F)] showed the formation of semi-capsule. Detailed 1H NMR experiments showed that the binding affinity of different anions follows the order SO42− > F > AcO ∼ H2PO4 > HSO4 ∼ Cl > Br, in accordance with the UV/Vis titration profiles (Table 2). The receptor has been found to selectively bind a SO42− anion (log[thin space (1/6-em)]K = 6.70) even in the presence of highly competitive fluoride ions (log[thin space (1/6-em)]K = 5.31), as confirmed by competitive 1H NMR experiments. Furthermore, a quinoline-based tripodal tris-urea receptor 28 in conjunction with the ferrocene-based receptor 27 has been employed to control intermolecular SO42− transfer between the two receptors driven by an electrochemical stimulus, which is detected by the change of the fluorescence intensity before and after electrochemical oxidation of the ferrocene groups.99

Ghosh et al. has shown the solid- and solution-state evidences of temperature dependent formation of a cation (TBA) sealed semi-capsule [(TBA)2SO4(33)], employing a pentafluorophenyl-based urea receptor 33 on a cyanuric acid platform (Fig. 14).92 The solution-state recognition of (TBA)2SO4 was made evident from the diffusion coefficient values observed for the receptor and (TBA)2SO4 from the 1H DOSY NMR analyses at variable temperatures. The authors suggested that the higher binding affinity of 33 with SO42− (log[thin space (1/6-em)]K = 5.74) compared to H2PO4 (log[thin space (1/6-em)]K = 4.39) could be due to the existence of ion–pair interaction without the loss of electrostatic energy arising from charge separation.


image file: c6ra00268d-f14.tif
Fig. 14 X-ray structure of cation sealed sulfate capsule, [(TBA)2SO4(33)]. C = green/gray, N = blue, O = red, S = yellow, F = chartreuse yellow.

Steed et al. has reported two sets of cationic tripodal tris-urea receptors on an aryl platform, each arm of which is functionalized with both urea and pyridinium functions.100,101 The first set of receptors 38a and 38b (Scheme 3) containing p-tolyl and n-octyl substituents, displayed significant binding affinity towards halides (Cl and Br) and oxyanions (NO3, AcO and H2PO4) with the formation of pyridinium C–H⋯A interactions despite the presence of six urea –NH donors, as evidenced in 1H NMR titration experiments.100 Density functional theory (DFT) calculations suggested that, in the lowest energy conformation of [(38a)Cl]2+, the pyridinium–urea receptor formed a monomeric capsule that encapsulate a chloride ion in its cavity by three C–H⋯Cl and three N–H⋯Cl hydrogen bonds, and is sealed at the periphery by C–H⋯π interactions between the p-tolyl rings (Fig. 15a). Halide binding with a –NH and a –CH proton from each receptor side arm rather than the urea –NH protons has also been supported by the magnitude of chemical shift changes in the 1H NMR experiments in DMSO-d6. Whereas, in the lowest energy conformation of [38b·Cl]2+ complex, a less symmetrical “2-up, 1-down” conformation is favoured where a chloride ion is bound to urea functions of the two receptor side arms that are oriented upwards, and this has further been supported by 1H and variable temperature (VT) NMR experiments in CD3CN.


image file: c6ra00268d-f15.tif
Fig. 15 Chloride-coordinated monomeric capsules of pyridinium-based tris-urea receptors 38a and 39a, as obtained from DFT optimization studies.

The second set of receptors 39a and 39b (Scheme 3) has been designed with two well-separated binding compartments on the same aryl platform, the bottom one comprising a 3-aminopyridinium pocket and the upper one involves three remote urea groups.101 DFT optimization and 1H NMR experiments (CD3CN) suggested that the 3-aminopyridinium pocket can bind strongly to a chloride ion via charge assisted N–H⋯Cl and C–H⋯Cl hydrogen bonding, and the upper urea groups zip-up the anion bound monomeric capsule by N–H⋯O[double bond, length as m-dash]C tape motif (Fig. 15b). The sharpening of the downfield shifted pyridinium –CH resonance upon addition of one equivalent of chloride suggested that the anion has a role in locking-in the 3-up conformation, working synergically with the intramolecular urea hydrogen bonding. The authors rationalized that in less polar solvents, the tendency of the urea groups to self-associate favoured a more preorganized receptor molecule for anion binding, whereas in more polar media, less preorganization may result due to poor self-association between the urea groups.

In their pioneering work on tripodal anion receptors, Wu and Li et al. has demonstrated the highly efficient extraction of sulfate from an aqueous to an organic phase (chloroform) by employing a nitrophenyl-substituted tripodal hexaurea receptor 40 (Scheme 3) to obtain complex [(TBA)2(40·SO4)DMSO].102 The receptor is capable of encapsulating a sulfate anion within a complementary tetrahedral cage that is protected by aromatic rings, and represents a successful strategy for overcoming the “Hofmeister bias” by taking advantage of a combination of complementarity, the chelate effect, and the hydrophobic effect. This is the first example of a receptor that satisfies the optimal coordination of 12 hydrogen bonds for a sulfate anion by encapsulation within its complementary molecular cavity with each urea group chelating an edge of the tetrahedral sulfate (Fig. 16a). The association constant (log[thin space (1/6-em)]K) estimated from the 1H NMR titration data (DMSO-d6) was found to be larger than 4.0, even in the presence of 25% D2O. The selective extraction of sulfate from a nitrate rich aqueous solution was achieved by thoroughly mixing with a equimolar solution of 40 and (TBA)Cl in CDCl3 and immediate analysis of the organic phase to ascertain the quantitative formation of the sulfate-encapsulated complex.


image file: c6ra00268d-f16.tif
Fig. 16 X-ray structures of monomeric capsules, (a) [40·(SO4)2−], and (b) [41·(CO3)2−]. C = green/gray, N = blue, O = red, S = yellow, F = chartreuse yellow. Countercations are not shown.

A pentafluorophenyl-substituted tripodal hexaurea receptor 41 has recently been reported by Hossain et al. for the absorption of atmospheric CO2 as CO32−, induced by the presence of excess TBAF in DMSO solution to obtain complex [(TBA)2(41·CO3)] (Fig. 16b).103 The 1H NMR titration data with (TEA)HCO3 yielded a binding constant (log[thin space (1/6-em)]K) of 3.25 for a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 (host/guest) binding model. Besides, the receptor showed strong affinity for HSO4 with an association constant (log[thin space (1/6-em)]K) of 3.35, presumably due to receptor-anion complementarity as established by Wu and Li et al.

Wu et al. has recently reported a set of three ferrocenyl-functionalized tripodal hexaurea receptors with ortho-, meta- and para-phenylene bridges for the bis-urea arms, designed for anion binding and electrochemical signalling purposes.104 In sulfate complex [(TBA)2(42·SO4)2H2O], the ortho-phenylene bridged hexaurea receptor 42 adopts a folded cone conformation for optimal coordination of sulfate by 12 –NH hydrogen bonds. Interestingly, the meta-phenylene bridged receptor 43 can encapsulate two SO42− anions in its “inner” and “outer” tripodal clefts, as evidenced by distinct stepwise 1[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]2 host/guest binding in the 1H NMR titration experiments (DMSO-d6), and was also supported by molecular modelling studies (Fig. 17). Such a two-step binding process is facilitated by the specific trigonal bipyramidal conformation of the meta-phenylene bridged receptor, which features two clefts rather than one tetrahedral cage as observed in ortho-phenylene bridged analogue. The association constants for the first and second step sulfate coordination were evaluated to be log[thin space (1/6-em)]K >4.0 and >2.0, respectively. Molecular modelling studies revealed that the para-phenylene bridged receptor forms a cavity that is too large for the second sulfate coordination, which may rationalize its weaker binding than the meta-analogue.


image file: c6ra00268d-f17.tif
Fig. 17 1[thin space (1/6-em)]:[thin space (1/6-em)]2 host–guest capsule of hexaurea receptor 43 with sulfate anion.

Tripodal thiourea receptors

Several Tren-based tris-thiourea receptors (Scheme 4) mostly analogous to the above discussed urea receptors have been synthesized, and their anion binding affinity and stoichiometry have been assessed both in solution and solid-states. It is interesting to note that, a few of these thiourea receptors mainly 46, 53 and 54 behaved very differently towards their anion binding affinity in solution-state and also solid-state host–guest complex stoichiometry, as compared to their urea analogues. Thus, the increased acidity of the thiourea –NH protons has a significant role to play in influencing the anion binding affinity and complexation, which is different from their urea analogues.

Dimeric capsules of tripodal thiourea receptors

As stated in the previous section, Gale et al. and coworkers have extensively studied the solid- and solution-state anion binding properties of a series of fluorinated tripodal tris-urea receptors 20–23 (Scheme 2) and their thiourea analogues 46–49 (Scheme 4).84 Among the fluorinated thiourea receptors, the anion complexes of the pentafluorophenyl-based receptor 46 have been more elaborately studied in the solid-state than the others. The authors were able to crystallize the anion complexes of 46 with Cl, NO3, SO42− and H2PO4 as their quaternary ammonium salts.84 The receptor has shown the formation of 1[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest complexes with spherical chloride, trigonal-planar nitrate and tetrahedral sulfate anions. Attempted crystallization of a bicarbonate complex has failed due to the mono-deprotonation of one of the thiourea –NH groups, presumably because HCO3 functioned as a Brønsted base. However, such deprotonation event has not been observed in the pentafluorophenyl-based urea analogue 21 even in the presence of (TBA)OH, in which case a CO32− complex was obtained by aerial CO2 fixation.83 Complexation with H2PO4, revealed the formation of a 2[thin space (1/6-em)]:[thin space (1/6-em)]2 host–guest stoichiometric capsule, where a (H2PO4)2 dimer is encapsulated within a dimeric assembly of 13.44 Å, similar to the urea analogue 21.

Sulfate and hydrogenphosphate complexes of the fluorophenyl-substituted thiourea receptor 47 showed the formation of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest capsular complexes,84 identical to the urea receptors 21 and 25. Hydrogenphosphate complexes [(TBA)2HPO4(47)2] and [(TBA)2HPO4(49)2] (Fig. 18a) with 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest stoichiometry were obtained in the presence of excess (TBA)H2PO4 indicating the deprotonation of H2PO4 anion in the solution-state while crystallization, as evidenced from 1H NMR titration experiments (DMSO-d6). In the 1H NMR titrations of 48 and 49 with (TBA)H2PO4, a combination of fast and slow exchange processes has been observed in wet DMSO-d6 (0.5/10% D2O), similar to urea compound 23. This behaviour was explained due to the deprotonation of bound H2PO4, and subsequent formation of a HPO42− complex.


image file: c6ra00268d-f18.tif
Fig. 18 X-ray structures of oxyanion encapsulated dimeric assemblies, (a) [(49)2·(HPO4)2−], and (b) [(46)2·(NO3)2]. C = green/gray, N = blue, O = red, P = orange, F = chartreuse yellow. Countercations are not shown.

A nitrate complex [(TBA)(46·NO3)] of thiourea receptor has also been obtained in the solid state, although no solution-state nitrate binding has been observed with 46–49.84 The complex revealed the formation of a (2 + 2) pseudo-dimeric capsule, where two nitrate encapsulated units are held together via anion–π interactions (Fig. 18b). Formation of (2 + 2) pseudo-dimeric capsules has also been observed in the F complexes of urea receptor 21 and 26.

The encapsulation of divalent tetrahedral oxyanion via deprotonation of its monovalent state has also been observed in the solid-state by employing the para-tolyl and para-methoxyphenyl-substituted thiourea receptors 50 and 51.105 Solution-state evidences of the deprotonation events have largely been provided by the tolyl-substituted receptor 50 with HCO3, HSO4 and H2PO4 anions in DMSO-d6. Both these receptors formed 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest SO42− capsules of identical size (∼9.90 Å), when crystallized in the presence of excess (TBA)HSO4 in DMSO. Similarly, crystallization of 51 in the presence of excess (TBA)H2PO4 in DMSO yielded a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest capsular complex with HPO42−. The same complex has also been obtained from a DMSO solution of the receptor mixed with K2HPO4 and TBAI. Capsule size of HPO42− complexes of 47, 49 and 51 are almost same, 10.02–10.12 Å (distance between the bridgehead-N atoms of dimeric assembly).

Sulfate and thiosulfate complexes of a chlorophenyl-substituted tris-thiourea receptor 52 also showed the formation of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest capsular complexes,106 identical to the urea receptor 15. Thiosulfate complex [(TEA)2S2O3(52)2] (Fig. 19b) was obtained from an DMSO–H2O (∼2[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v) solution of 52 containing an excess (10 equiv.) of equimolar quantities of (TEA)Cl and Na2S2O3, and sulfate complex [(TBA)2SO4(52)2] was obtained from a DMSO solution containing 52 and excess (TBA)2SO4 or (TBA)HSO4. The capsular size of the thiosulfate complex (10.12 Å) is slightly larger (∼0.8 Å) than the sulfate complex (9.33 Å), which may be attributed to the larger size and lower charge density of the S2O32− as compared to SO42−. Considering a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding model, the association constants (log[thin space (1/6-em)]K) were determined to be 4.54 and 3.35 for SO42− and S2O32− respectively, using 1H NMR titration data.


image file: c6ra00268d-f19.tif
Fig. 19 X-ray structures of oxyanion encapsulated dimeric assemblies, (a) [(53)2·(PO4)3−], and (b) [(52)2·(S2O3)2−]; C = green/gray, N = blue, O = red, S = yellow, P = orange. Countercations are not shown.

We have further established that the para-nitrophenyl-substituted tris-thiourea receptor 53 functions as a competent hydrogen bonding scaffold that can selectively encapsulate a trivalent phosphate anion (PO43−) within the persistent dimeric capsules assembled by aromatic π-stacking interactions between the receptor side-arms.107,108 An acetonitrile (MeCN) solution of 53 in the presence of excess (TBA)H2PO4 crystallized into 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest capsular complex [(TBA)3PO4(53)2·2MeCN], where a PO43− anion is encapsulated by 12 –NH hydrogen bonds (Fig. 19a).107 The deprotonation of bound H2PO4 was also evident from the origin of a new set of signals in the 1H NMR titration experiments (DMSO-d6). Similar solution-state behaviour has not been observed with the urea analogue 26. Further, an acetonitrile solution of 53 in the presence of phosphoric acid (H3PO4) yielded a non-capsular complex [(H53)2HPO4·3H2O], with two receptor cations and an externally bound HPO42− anion.108 Phosphate complex [(TBA)3PO4(53)2·2MeCN], can reproducibly be crystallized quantitatively in much higher yield simply by adding an excess of (TBA)F/(TBA)AcO into an acetonitrile solution of complex [(H53)2HPO4·3H2O]. Similarly, the TEA salt of PO43− encapsulated complex [(TEA)3PO4(53)2], can reproducibly be crystallized from an acetonitrile solution of complex [(H53)2HPO4·3H2O] charged with excess (TEA)AcO/(TEA)HCO3.108 The solution-state deprotonation of receptor cations can be attributed to the high basicity of crystallization medium induced by the presence of excess fluoride/acetate/bicarbonate ions whereas, deprotonation of HPO42− anion can be explained due to the formation of multiple hydrogen bonding interactions with the in situ generated neutral receptor that lowers the pKa of the bound HPO42− to the extent that it is deprotonated by excess fluoride/acetate/bicarbonate in solution. Competitive crystallization experiments performed in the presence of an excess of anions such as HCO3, HSO4, AcO, NO3, F and Cl further establish the phenomenon of selective PO43− encapsulation as confirmed by 1H NMR, 31P NMR, and powder X-ray diffraction patterns of the isolated crystals. The size of the PO43− capsules were measured to be 9.60 Å and 9.53 Å having TBA and TEA countercations, respectively. However, a much smaller capsular size of 7.92 Å has been observed in the carbonate complex [(TEA)2CO3(53)2], comparable to the urea analogue [(TEA)2CO3(26)2].

The combined solution-phase, solid-phase and phase-interface anion binding properties of a para-cyanophenyl-substituted thiourea 54 has been thoroughly explored by Ghosh et al.109 Solution-state anion binding studies using 1H NMR spectroscopy (DMSO-d6) and ITC measurements in dry acetonitrile showed relatively higher association constants with halides over oxyanions, and follows the binding trend in the order F > Cl > HSO4 > H2PO4 (Table 3), which is different from its urea analogue 16. The receptor has further been exploited in the liquid–liquid extraction of sulfate and fluoride from aqueous media. Using an anion-exchange based liquid–liquid extraction strategy the receptor has shown ca. 70% extraction of fluoride and ca. 40% extraction of sulfate from aqueous solutions by employing (TBA)I as the phase transfer agent. Identical to the other thiourea receptors (47, 50, 51, and 52), 54 formed a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host-guest capsular complex with SO42− anion with a dimension of 9.51 Å.

Table 3 Binding constants (log[thin space (1/6-em)]K) of thiourea receptors with different anions
Receptors F Cl H2PO4 SO42−
a K determined by 1H NMR titration experiments.b K determined by UV-Vis/fluorescence titration experiments.
44a 2.28 2.40 >4.0
45b 5.03 3.93
46a 2.10 2.11 >4.0
47a 2.25 2.35 >4.0
53a 4.47 3.54 5.87 3.75
54a 3.57 3.06 2.20 2.83


It has been mentioned earlier that three new families of tripodal urea (34–36) and thiourea (55–57) receptors have been reported by Davis et al. showing the chloride binding affinities are strongly dependent on the nature of the binding groups (thiourea > urea) and the aryl substituent (electron-withdrawing groups being more effective).93,94 In DMSO-d6, benzene-based 57F2 (log[thin space (1/6-em)]K = 2.65) proved stronger than the desmethyl cyclohexane system 55F2 (log[thin space (1/6-em)]K = 2.0) but slightly weaker than the hexamethyl cyclohexane analogues 56F2 (log[thin space (1/6-em)]K = 2.82). However, in chloroform-d the chloride binding affinity of 57F2 (log[thin space (1/6-em)]K = 8.83) was found to be higher than that of 56F2 (log[thin space (1/6-em)]K = 7.47) and 55F2 (log[thin space (1/6-em)]K = 6.38) as well. Ab initio calculations suggested that the tris-(axial/NH-in) conformation of the cyclohexane-based tris-urea/thiourea systems can form six hydrogen bonds to Cl with lengths in the range 2.6–2.8 Å. Crystal structure analysis of the chloride complex of 57F2 revealed the formation of a dimeric receptor capsule surrounding a cyclic (Cl–H2O)2 dianionic cluster (Fig. 20). Receptor 56F2 has been found to be the most effective anionophore, while 57F2 possesses roughly 70% the activity of 56F2, as measured by approximate half-lives t1/2 (s) for chloride–nitrate exchange across the large unilamellar vesicles (LUVs, 200 nm), and specific initial rates [I] (s−1) obtained from the fluorescence decay data. The hexamethyl cyclohexane system 56 promotes anion transport by improving binding without substantially affecting structural flexibility.


image file: c6ra00268d-f20.tif
Fig. 20 X-ray structure of dimeric capsule [(57F2)2(Cl·H2O)2], C = gray, N = blue, O = red, F = chartreuse yellow. Countercations are not shown.

Monomeric capsules (semi-capsules) of thiourea receptors

Thiourea receptors 46 and 47 showed the formation of monomeric capsules with chloride anion [(TBA)(L·Cl)] (L = 46/47), as observed in urea analogue 20.84 Formation of monomeric halide capsules have also been observed in tolyl-based receptor 50 in complexes [TBA(50·X)DMSO] (X = Cl/Br).105 In all these complexes, an anion is coordinated by six –NH hydrogen bonds inside the receptor cavity.

Interestingly, the structural aspects of fluoride binding by a nitrophenyl-based tris-thiourea 53 revealed the formation of solvent (DMSO/MeCN) sealed monomeric capsules.107,108 Both the complexes [TBA(53·F)S] (S = DMSO/MeCN) showed F encapsulation within the receptor cavity and hydrogen bonding interactions between a lattice solvent and peripheral nitro-aromatic functions, demonstrating the ditopic binding of a negatively charged species (F) and a neutral molecule DMSO/MeCN (Fig. 21a). Whereas, fluoride complex [TBA(54·F)MeCN] of cyanophenyl-based receptor 54 showed the monotopic recognition of fluoride.109


image file: c6ra00268d-f21.tif
Fig. 21 X-ray structures of (a) solvent sealed monomeric capsule [53·(F)DMSO], and (b) cation sealed monomeric capsule [53·(TBA)(SO4)]. C = green/gray, N = blue, O = red, S = yellow, F = chartreuse yellow. All TBA cations are not shown.

Receptors 46 and 53 showed the formation of cation (TBA) sealed monomeric capsules with SO42− anion [(TBA)2(L·SO4)] (L = 46/53).84,108 In both the complexes, a sulfate anion is bound within the receptor cavity by six –NH hydrogen bonds, and one of the TBA cations is held in close contact with both the receptor and anion by –CH hydrogen bonds (Fig. 21b).

Tripodal urea–amide hybrid receptors

Recently, a set of three tripodal urea-amide hybrid receptors 58–60 (Scheme 5) based upon the N-methyl-1,3,5-benzenetricarboxamide platform has been reported by Gunnlaugsson et al.110 This ligand platform having N-methyl tertiary amides gives rise to highly preorganized structures and this preorganization has resulted in the formation of self-assembled dimeric capsules with SO42− and H2PO4 anions. Structure analyses revealed the formation of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest stoichiometric capsule with SO42− anion, and 2[thin space (1/6-em)]:[thin space (1/6-em)]2 capsules with H2PO4 anion, where the hydrogen bonded H2PO4 dimer is bridged by a water molecule within the dimeric capsular assembly (Fig. 22). Sulfate capsules of the receptors were obtained from MeCN–DMF solutions, and showed identical mode of anion encapsulation involving all of the 12 urea –NH protons. 1H NMR titrations of 58–60 with SO42− showed the existence of two binding stoichiometry in solution, an initial formation of the desired 2[thin space (1/6-em)]:[thin space (1/6-em)]1 host–guest complex followed by the formation of a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 stoichiometry.
image file: c6ra00268d-f22.tif
Fig. 22 X-ray structures of dimeric capsules of urea–amide hybrid receptors, (a) [(58)2·(SO4)2−] and (b) [(58)2·(H2PO4)2·H2O]. C = green/gray, N = blue, O = red, S = yellow, P = orange, F = chartreuse yellow. TBA cations are not shown.

The same group has previously reported another set of tripodal urea–amide hybrid receptors 61–65 (Scheme 5) based on the 1,3,5-benzenetricarboxamide platform111.1H NMR titration studies showed that these receptors can bind anions such as, Cl, AcO, SO42− and H2PO4 in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 stoichiometry, but failed to form any self-assembled anion capsules in the solid state. Molecular modelling studies indicated that the secondary amides prevented the formation of the desired preorganized cavity, instead giving rise to the formation of more open or bowl-shaped structures. All receptors showed high binding affinities for AcO, and H2PO4.

Finally, an anion binding tripodal scaffold that needs to be addressed is the squaramide-based receptors 66–68 (Scheme 5).112 Structure analysis of complex [(TBA)2(66·SO4)] revealed an unusual formation of 2[thin space (1/6-em)]:[thin space (1/6-em)]2 host–guest stoichiometric capsule with SO42− anion, unlike those have been discussed for tris-urea/thiourea receptors. Each SO42− anion is coordinated to both the receptors engaging two squaramide arms of one molecule and another arm from the second molecule of the dimeric capsule. DOSY NMR experiments indicated that dimeric structure does not exist as a stable complex in solution, and suggested a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding stoichiometry, supported by 1H-NMR titration studies. Detailed 1H NMR titration experiments (DMSO-d6) yielded the anion selectivity order (log[thin space (1/6-em)]K) SO42− (4.75) > H2PO4 (4.15) > HSO4 (3.65) > CH3COO (2.82) > Cl (2.58) for receptor 66.

Systematic developments and relative anion binding affinities of tripodal receptors

Following their systematic developments over the years, tripodal receptors can be broadly categorized into first and second generation receptors. A vast majority of the tripodal receptors are based on the tris(2-aminoethyl)amine (Tren) scaffold and are regarded as the first generation tripodal receptors (FGTRs) because of their first use in anion receptor chemistry. The growth of anion coordination chemistry has witnessed the development of several other tripodal receptors based on 1,3,5-trialkyl (methyl/ethyl) benzene, cyclohexane and cyanuric acid platform. Because of their structural similarity with the Tren-based system (the bridgehead tertiary-N is replaced by a six member ring structure), these can also be regarded as the FGTRs. Solution and solid state studies of the FGTRs with different anions portrayed a clear picture of the concepts of preorganization, complementarity and hydrophobic effect essential for effective anion recognition by these families of flexible receptors. Based on these acquired information, several groups have designed and synthesized some second generation tripodal receptors (SGTRs) either by incorporation of additional functional groups and/or aromatic rings into the tripodal side arms or increasing the number of hydrogen bond donor sites on each side arm of the receptor molecule. The development of SGTRs have brought into light a range of exciting structures such as, tetrahedral phosphate cage formed by 37 and sulfate encapsulation within the complementary tetrahedral cavity of 40 and 42. However, solution state anion binding affinity of the SGTRs did not vary significantly in comparison to most FGTRs. This could be the reason why FGTRs (mostly with fluorinated aryl functions) have mostly been employed for transmembrane anion transportation applications, with advantageous simple and easy synthetic accessibility. A year wise development of different tripodal scaffolds for anion recognition has been portrayed in Scheme 6, and Table 4 shows a comparative account of the solution state anion binding affinities (log[thin space (1/6-em)]K) of the tripodal receptors for the commonly studied halides (F and Cl) and oxyanions (AcO, SO42−/HSO4 and H2PO4).
image file: c6ra00268d-s6.tif
Scheme 6 Selected scaffolds of first generation and second generation tripodal receptors (FGTRs and SGTRs), showing their systematic developments over the years (1992–2015).
Table 4 A comparative account of the tripodal receptors based on their increasing binding constant values for an anion
Anions Tripodal receptors and their relative binding constants (log[thin space (1/6-em)]K)
log[thin space (1/6-em)]K = 1–3 log[thin space (1/6-em)]K = 3–4 log[thin space (1/6-em)]K = 4–5 log[thin space (1/6-em)]K = 5–6 log[thin space (1/6-em)]K > 6
F 9 3b, 5a, 5b, 25, 54 3c, 16, 21, 53 3a, 10, 27 4
Cl 3c, 5b, 9, 13, 20, 22, 44, 46, 47, 61, 62, 64 2a, 3b, 5a, 10, 16, 21, 33, 53, 54 27 55–57F2
AcO 61, 62, 66 25, 26, 33, 63, 64, 65 21, 27 12
H2PO42− 29, 44, 46, 47, 54 2a, 61, 63, 64, 65 13, 14, 16, 25, 26, 27, 33, 62, 66 45, 53
SO42−/HSO4 2a, 9, 29, 54, 61, 63, 64, 65 14, 41, 45, 53, 62 13, 16, 20, 21, 22, 25, 26, 40, 44, 46, 47, 52, 66 30, 33 27


Conclusions

We have been involved in the research targeting the anion induced supramolecular self-assemblies of hydrogen bonding tripodal scaffolds since almost a decade, and obtained some interesting results employing a few N-bridged tripodal receptors having either amide or urea/thiourea functions for anion coordination. The purpose of this review is to discuss and generalize the anion coordination induced formation of dimeric and monomeric capsular assemblies in hydrogen bonding tripodal scaffolds. Tren-based tris-amide receptors (1–5) have always been shown to form monomeric capsules with halides with high selectivity towards fluoride and thus, selective fluoride sensing and efficient fluoride extraction from water has been achieved with receptors 4 and 5a, respectively. Whereas, tris-amide receptors (7, 8 and 9) on a mesitylene platform has been established as an ideal scaffold for encapsulation of halide-water clusters via formation of dimeric capsular assembly. Tren-based tris-urea receptors (13–28) have frequently been shown to form dimeric capsules with oxyanions showing high selectivity towards sulfate and thereby, selective sulfate separation by competitive crystallization and liquid–liquid extraction has been achieved with receptors 15, 17 and 21 respectively. Tren-based hexaurea receptors with an ortho-phenylene bridge (40–42) provides an ideal cavity for sulfate encapsulation and as such, receptors 40 and 42 have been employed in the selective separation of sulfate from aqueous solution containing high concentration of nitrate by liquid–liquid extraction. Another interesting property that has been observed with few tripodal urea receptors (tris/hexa-urea) is the capture of atmospheric carbondioxide as carbonate inside the molecular cavity of the tripodal scaffold, as demonstrated by tris-urea receptors 15, 21, 25 and hexaurea 41 in basic DMSO solution mixtures (Table 5).
Table 5 Tripodal amide and urea/thiourea receptors for selective anion recognition, anion separation and transmembrane anion transport
Receptors Applications Method(s) involved/remarks
3a Selective fluoride recognition 1H-NMR spectroscopy
4 Selective fluoride recognition 1H-NMR and UV-vis spectroscopy
5a Extraction of KF and KCl from water Extraction by 18-crown-6-ether as a cation host
6g Selective nitrate sensing Fluorescence spectroscopy
15 Carbondioxide uptake as carbonate complex DMSO solution containing (TBA)OH
15 Extraction of sulfate, thiosulfate and chromate Extraction by carbonate complex (in organic phase)
17 Selective sulfate separation Competitive crystallization
21 Carbondioxide uptake as carbonate complex DMSO solution containing (TBA)OH
21 Extraction of sulfate from aqueous to organic phase Extraction by carbonate complex (in organic phase)
23 Transmembrane chloride transport Vesicle anion transport assay
25 Carbondioxide uptake as carbonate complex DMSO solution containing (TBA)OH/F
27 Electrochemical sensing of sulfate and phosphate Cyclic voltammetry
40 Extraction of sulfate from water (TBA)Cl as phase transfer agent
41 Carbondioxide uptake as carbonate complex DMSO solution containing (TBA)F
53 Selective phosphate separation Competitive crystallization
54 Extraction of fluoride and sulfate from water (TBA)I as phase transfer agent
56–57F2 Transmembrane chloride transport Vesicle anion transport assay


The structural aspects of anion binding by Tren-based tris-thiourea receptors (44–54) are analogues to the tris-urea receptors as observed in fluorinated urea and thiourea receptors. However, receptor 53 (thiourea analogue of tris-urea receptor 26) showed unusual selectivity for trivalent phosphate encapsulation even in the presence of other competitive anions, which has not been observed with any other tripodal urea/thiourea receptors. Most importantly, some significant results have also been achieved in the area of transmembrane anion transport using fluorinated tris-urea/thiourea receptors as anion carriers. Overall, a significant amount of research has been pursued towards the progress of anion induced capsular self-assemblies with hydrogen bonding tripodal scaffolds. Based on the information collected from X-ray structures of numerous anion complexes, chemists have always tried to optimize the receptor-anion complementarity by suitable functioning of the tripodal scaffold with highly ordered convergent binding groups. Although a few examples of anion separation has been discussed in this review, but the utility of molecular capsules for the binding and removal of anionic species is undoubtedly in its infancy. Further research in toxic anion separation using the basic concept of anion receptor chemistry is much waited, which could solve some of the practical aspects of drinking water purification. The expansion of biological applications for synthetic anion transporters is also in its infancy, and there will certainly be more opportunities in the future for using the tripodal receptors to address biochemical and biomedical problems. In consideration of the interesting results achieved, tripodal receptors will continue to play an active role in anion coordination chemistry and will continue to be one of the first options to be considered in the design of receptors for halides and oxyanions.

Acknowledgements

This work was supported by CSIR and SERB through grant 01/2727/13/EMR-II and SR/S1/OC-62/2011, New Delhi, India. We also thank Central Instrument Facility (CIF) at Indian Institute of Technology (IIT) Guwahati for providing instrument facilities, and DST-FIST for sponsoring a single crystal X-ray diffractometer at Department of Chemistry, IIT Guwahati.

Notes and references

  1. J. L. Sessler, P. A. Gale and W.-S. Cho, Anion Receptor Chemistry, RSC Publishing, Cambridge, 2006 Search PubMed.
  2. C. J. Avers, Molecular Cell Biology, Addison–Wesley, Reading, 1986 Search PubMed.
  3. J. L. Atwood and A. Szumna, J. Am. Chem. Soc., 2002, 124, 10646 CrossRef CAS PubMed.
  4. P. A. Gale, Chem. Commun., 2011, 47, 82 RSC.
  5. P. A. Gale, N. Busschaert, C. J. E. Haynes, L. E. Karagiannidis and I. L. Kirby, Chem. Soc. Rev., 2014, 43, 205 RSC.
  6. M. Wenzel, J. R. Hiscock and P. A. Gale, Chem. Soc. Rev., 2012, 41, 480 RSC.
  7. P. A. Gale, Chem. Soc. Rev., 2010, 39, 3746 RSC.
  8. C. Caltagirone and P. A. Gale, Chem. Soc. Rev., 2009, 38, 520 RSC.
  9. P. A. Gale, S. E. García-Garrido and J. Garric, Chem. Soc. Rev., 2008, 37, 151 RSC.
  10. T. Gunnlaugsson, M. Glynn, G. M. Tocci, P. E. Kruger and F. M. Pfeffer, Coord. Chem. Rev., 2006, 250, 3094 CrossRef CAS.
  11. S. Kubik, Chem. Soc. Rev., 2010, 39, 3648 RSC.
  12. T. H. Rehm and C. Schmuck, Chem. Soc. Rev., 2010, 39, 3597 RSC.
  13. C. Bazzicalupi, A. Bencini and V. Lippolis, Chem. Soc. Rev., 2010, 39, 3709 RSC.
  14. M. Cametti and K. Rissanen, Chem. Commun., 2009, 2809 RSC.
  15. E. García-Espana, P. Diaz, J. M. Llinares and A. Bianchi, Coord. Chem. Rev., 2006, 250, 2952 CrossRef.
  16. M. Boiocchi, L. D. Boca, D. Esteban-Gómez, L. Fabbrizzi, M. Licchelli and E. Monzani, J. Am. Chem. Soc., 2004, 126, 16507 CrossRef CAS PubMed.
  17. D. Esteban-Gómez, L. Fabbrizzi and M. Licchelli, J. Org. Chem., 2005, 70, 5717 CrossRef PubMed.
  18. L. S. Evans, P. A. Gale, M. E. Light and R. Quesada, Chem. Commun., 2006, 965 RSC.
  19. P. Bose and P. Ghosh, Chem. Commun., 2010, 46, 2962 RSC.
  20. A. Basu and G. Das, Inorg. Chem., 2012, 51, 882 CrossRef CAS PubMed.
  21. A. Basu, S. K. Dey and G. Das, RSC Adv., 2013, 3, 6596 RSC.
  22. K. Bowman-James, Acc. Chem. Res., 2005, 38, 671 CrossRef CAS PubMed.
  23. J. T. Davis, O. Okunola and R. Quesada, Chem. Soc. Rev., 2010, 39, 3843 RSC.
  24. P. A. Gale, Acc. Chem. Res., 2011, 44, 216 CrossRef CAS PubMed.
  25. P. A. Gale, R. Pérez-Tomás and R. Quesada, Acc. Chem. Res., 2013, 46, 2801 CrossRef CAS PubMed.
  26. H. Maeda, T. Shirai and S. Uemura, Chem. Commun., 2013, 49, 5310 RSC.
  27. H. Maeda, K. Naritani, Y. Honsho and S. Seki, J. Am. Chem. Soc., 2011, 133, 8896 CrossRef CAS PubMed.
  28. B. Dong, T. Sakurai, Y. Honsho, S. Seki and H. Maeda, J. Am. Chem. Soc., 2013, 135, 1284 CrossRef CAS PubMed.
  29. B. Dong, T. Sakurai, Y. Bando, S. Seki, K. Takaishi, M. Uchiyama, A. Muranaka and H. Maeda, J. Am. Chem. Soc., 2013, 135, 14797 CrossRef CAS PubMed.
  30. K. M. Mullen and P. D. Beer, Chem. Soc. Rev., 2009, 38, 1701 RSC.
  31. P. Ballester, Chem. Soc. Rev., 2010, 39, 3810 RSC.
  32. R. Vilar, Struct. Bonding, 2008, 129, 175 CrossRef CAS.
  33. H. Juwarker and K.-S. Jeong, Chem. Soc. Rev., 2010, 39, 3664 RSC.
  34. S. J. Coles, J. G. Frey, P. A. Gale, M. B. Hursthouse, M. E. Light, K. Navakhun and G. L. Thomas, Chem. Commun., 2003, 568 RSC.
  35. T. Basu and R. Mondal, CrystEngComm, 2010, 12, 366 RSC.
  36. J. W. Steed and J. L. Atwood, Supramolecular Chemistry: An Introduction, Wiley, Chichester, U.K, 2000 Search PubMed.
  37. Z. Wang, H. Luecke, N. Yao and F. A. Quiocho, Nat. Struct. Biol., 1997, 4, 519 CrossRef CAS PubMed.
  38. J. W. Pflugrath and F. A. Quiocho, Nature, 1985, 314, 257 CrossRef CAS PubMed.
  39. J. W. Steed, Chem. Commun., 2006, 2637 RSC.
  40. B. Kuswandi, Nuriman, W. Verboom and D. N. Reinhoudt, Sensors, 2006, 6, 978 CrossRef CAS.
  41. M. Arunachalam and P. Ghosh, Chem. Commun., 2011, 47, 8477 RSC.
  42. R. Dutta and P. Ghosh, Chem. Commun., 2014, 50, 10538 RSC.
  43. A. E. Hargrove, S. Nieto, T. Zhang, J. L. Sessler and E. V. Anslyn, Chem. Rev., 2011, 111, 6603 CrossRef CAS PubMed.
  44. F. G. Bordwell, Acc. Chem. Res., 1988, 21, 456 CrossRef CAS.
  45. V. S. Bryantsev and B. P. Hay, Org. Lett., 2005, 7, 5031 CrossRef CAS PubMed.
  46. P. D. Beer, D. Hesek, J. Hodacova and S. E. Stokes, J. Chem. Soc., Chem. Commun., 1992, 270 RSC.
  47. P. D. Beer, C. Hazlewood, D. Hesek, J. Hodacova and S. E. Stokes, J. Chem. Soc., Dalton Trans., 1992, 1327 Search PubMed.
  48. P. D. Beer, Z. Chen, A. J. Goulden, A. Graydon, S. E. Stokes and T. Wearb, J. Chem. Soc., Chem. Commun., 1993, 1834 RSC.
  49. S. Valiyaveettil, J. F. J. Engbersen, W. Verboom and D. N. Reinhoudt, Angew. Chem., Int. Ed., 1993, 32, 900 CrossRef.
  50. I. Ravikumar, P. S. Lakshminarayanan and P. Ghosh, Inorg. Chim. Acta, 2010, 363, 2886 CrossRef CAS.
  51. S. K. Dey and G. Das, Chem. Commun., 2011, 47, 4983 RSC.
  52. S. K. Dey and G. Das, Cryst. Growth Des., 2011, 11, 4463 CAS.
  53. S. K. Dey, B. K. Datta and G. Das, CrystEngComm, 2012, 14, 5305 RSC.
  54. I. Ravikumar, S. Saha and P. Ghosh, Chem. Commun., 2011, 47, 4721 RSC.
  55. S. Saha, B. Akhuli, I. Ravikumar, P. S. Lakshminarayanan and P. Ghosh, CrystEngComm, 2014, 16, 4796 RSC.
  56. A. S. Singh and S.-S. Sun, Chem. Commun., 2011, 47, 8563 RSC.
  57. A. S. Singh and S.-S. Sun, RSC Adv., 2012, 2, 9502 RSC.
  58. A. S. Singh and S.-S. Sun, J. Org. Chem., 2012, 77, 1880 CrossRef CAS PubMed.
  59. M. Arunachalam and P. Ghosh, Chem. Commun., 2009, 5389 RSC.
  60. M. Arunachalam and P. Ghosh, Inorg. Chem., 2010, 49, 943 CrossRef CAS PubMed.
  61. R. Dutta and P. Ghosh, Eur. J. Inorg. Chem., 2012, 3456 CrossRef CAS.
  62. S. Bao, W.-T. Gong, W.-D. Chen, J.-W. Ye, Y. Lin and G.-L. Ning, J. Inclusion Phenom. Macrocyclic Chem., 2011, 70, 115 CrossRef CAS.
  63. S. Chakraborty, R. Dutta, M. Arunachalam and P. Ghosh, Dalton Trans., 2014, 43, 2061 RSC.
  64. A. Basu and G. Das, Chem. Commun., 2013, 49, 3997 RSC.
  65. A. Basu, R. Chutia and G. Das, CrystEngComm, 2014, 16, 4886 RSC.
  66. P. J. Smith, M. V. Reddington and C. S. Wilcox, Tetrahedron Lett., 1992, 33, 6085 CrossRef CAS.
  67. E. Fan, S. A. V. Arman, S. Kincaid and A. D. Hamilton, J. Am. Chem. Soc., 1993, 115, 369 CrossRef CAS.
  68. C. Raposo, M. Almaraz, M. Martin, V. Weinrich, M. L. Mussons, V. Alcazar, M. C. Caballero and J. R. Morán, Chem. Lett., 1995, 37, 2795 Search PubMed.
  69. H. Xie, S. Yi, X. Yang and S. Wu, New J. Chem., 1999, 23, 1105 RSC.
  70. H. Xie, S. Yi and S. Wu, J. Chem. Soc., Perkin Trans. 2, 1999, 2751 RSC.
  71. R. Custelcean, B. A. Moyer and B. P. Hay, Chem. Commun., 2005, 5971 RSC.
  72. P. Bose, R. Dutta and P. Ghosh, Org. Biomol. Chem., 2013, 11, 4581 CAS.
  73. R. Dutta, S. Chakraborty, P. Bose and P. Ghosh, Eur. J. Inorg. Chem., 2014, 25, 4134 CrossRef.
  74. R. Custelcean, P. Remy, P. V. Bonnesen, D. Jiang and B. A. Moyer, Angew. Chem., Int. Ed., 2008, 47, 1866 CrossRef CAS PubMed.
  75. B. Wu, J. Liang, J. Yang, C. Jia, X.-J. Yang, H. Zhang, N. Tang and C. Janiak, Chem. Commun., 2008, 1762 RSC.
  76. R. Custelcean and P. Remy, Cryst. Growth Des., 2009, 9, 1985 CAS.
  77. A. Rajbanshi, B. A. Moyer and R. Custelcean, Cryst. Growth Des., 2011, 11, 2702 CAS.
  78. R. Zhang, Y. Zhao, J. Wang, L. Ji, X.-J. Yang and B. Wu, Cryst. Growth Des., 2014, 14, 544 CAS.
  79. B. Akhuli, I. Ravikumar and P. Ghosh, Chem. Sci., 2012, 3, 1522 RSC.
  80. I. Ravikumar, P. S. Lakshminarayanan, M. Arunachalam, E. Suresh and P. Ghosh, Dalton Trans., 2009, 4160 RSC.
  81. P. S. Lakshminarayanan, I. Ravikumar, E. Suresh and P. Ghosh, Chem. Commun., 2007, 5214 RSC.
  82. R. Dutta, P. Bose and P. Ghosh, Dalton Trans., 2013, 42, 11371 RSC.
  83. I. Ravikumar and P. Ghosh, Chem. Commun., 2010, 46, 1082 RSC.
  84. N. Busschaert, M. Wenzel, M. E. Light, P. Iglesias-Hernandez, R. Perez-Tomas and P. A. Gale, J. Am. Chem. Soc., 2011, 133, 14136 CrossRef CAS PubMed.
  85. R. Chutia, S. K. Dey and G. Das, Cryst. Growth Des., 2013, 13, 883 CAS.
  86. S. K. Dey, R. Chutia and G. Das, Inorg. Chem., 2012, 51, 1727 CrossRef CAS PubMed.
  87. R. Chutia, S. K. Dey and G. Das, Cryst. Growth Des., 2015, 15, 4993 CAS.
  88. D. A. Jose, D. K. Kumar, B. Ganguly and A. Das, Inorg. Chem., 2007, 46, 5817 CrossRef CAS PubMed.
  89. A. Basu and G. Das, J. Org. Chem., 2014, 79, 2647 CrossRef CAS PubMed.
  90. R. Dutta, B. Akhuli and P. Ghosh, Dalton Trans., 2015, 44, 15075 RSC.
  91. R. Dutta and P. Ghosh, Eur. J. Inorg. Chem., 2013, 2673 CrossRef CAS.
  92. I. Ravikumar and P. Ghosh, Chem. Commun., 2010, 46, 6741 RSC.
  93. J. A. Cooper, S. T. G. Street and A. P. Davis, Angew. Chem., Int. Ed., 2014, 53, 5609 CrossRef CAS PubMed.
  94. H. Valkenier, C. M. Dias, K. L. P. Goff, O. Jurcek, R. Puttreddy, K. Rissanen and A. P. Davis, Chem. Commun., 2015, 51, 14235 RSC.
  95. B. Wu, F. Cui, Y. Lei, S. Li, N. de, S. Amadeu, C. Janiak, Y.-J. Lin, L.-H. Weng, Y.-Y. Wang and X.-J. Yang, Angew. Chem., Int. Ed., 2013, 52, 5096 CrossRef CAS PubMed.
  96. A. Pramanik, D. R. Powell, B. M. Wong and M. A. Hossain, Inorg. Chem., 2012, 51, 4274 CrossRef CAS PubMed.
  97. F. Zhuge, B. Wu, J. Liang, J. Yang, Y. Liu, C. Jia, C. Janiak, N. Tang and X.-J. Yang, Inorg. Chem., 2009, 48, 10249 CrossRef CAS PubMed.
  98. M. Li, B. Wu, C. Jia, X. Huang, Q. Zhao, S. Shao and X.-J. Yang, Chem.–Eur. J., 2011, 17, 2272 CrossRef CAS PubMed.
  99. M. Li, Y. Hao, B. Wu, C. Jia, X. Huang and X.-J. Yang, Org. Biomol. Chem., 2011, 9, 5637 CAS.
  100. D. R. Turner, M. J. Paterson and J. W. Steed, J. Org. Chem., 2006, 71, 1598 CrossRef CAS PubMed.
  101. D. R. Turner, M. J. Paterson and J. W. Steed, Chem. Commun., 2008, 1395 RSC.
  102. C. Jia, B. Wu, S. Li, X. Huang, Q. Zhao, Q.-S. Li and X.-J. Yang, Angew. Chem., Int. Ed., 2011, 50, 486 CrossRef CAS PubMed.
  103. A. Pramanik, M. E. Khansari, D. R. Powell, F. R. Fronczek and M. A. Hossain, Org. Lett., 2014, 16, 366 CrossRef CAS PubMed.
  104. X. Huang, B. Wu, C. Jia, B. P. Hay, M. Li and X.-J. Yang, Chem.–Eur. J., 2013, 19, 9034 CrossRef CAS PubMed.
  105. A. Gogoi and G. Das, Supramol. Chem., 2013, 25, 819 CrossRef CAS.
  106. A. Basu and G. Das, Dalton Trans., 2012, 41, 10792 RSC.
  107. S. K. Dey and G. Das, Dalton Trans., 2011, 40, 12048 RSC.
  108. S. K. Dey and G. Das, Dalton Trans., 2012, 41, 8960 RSC.
  109. P. Bose, R. Dutta, S. Santra, B. Chowdhury and P. Ghosh, Eur. J. Inorg. Chem., 2012, 35, 5791 CrossRef.
  110. K. Pandurangan, J. A. Kitchen, S. Blasco, E. M. Boyle, B. Fitzpatrick, M. Feeney, P. E. Kruger and T. Gunnlaugsson, Angew. Chem., Int. Ed., 2015, 54, 4566 CrossRef CAS PubMed.
  111. C. Maria Gomes dos Santos, E. M. Boyle, S. D. Solis, P. E. Kruger and T. Gunnlaugsson, Chem. Commun., 2011, 47, 12176 RSC.
  112. C. Jin, M. Zhang, L. Wu, Y. Guan, Y. Pan, J. Jiang, C. Lin and L. Wang, Chem. Commun., 2013, 49, 2025 RSC.

Footnotes

Present address: Institut für Anorganische Chemie und Strukturchemie, Heinrich-Heine-Universität Düsseldorf, 40204 Düsseldorf, Germany.
Present address: Physical/Materials Chemistry Division, CSIR-National Chemical Laboratory, Dr. Homi Bhabha Road, Pune 411008, India.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.