Chemical Science EDGE ARTICLE O pe n A cc es s A rti cl e. P ub lis he d on 1 1 N ov em be r 2 01 4. D ow nl oa de d on 1 9/ 11 /2 01 4 16 :2 3: 20 . Th is ar tic le is li ce ns ed u nd er a C re at iv e Co m m on s A ttr ib ut io n 3. 0 U np or te d Li ce nc e. View Article Online View JournalLarge, heteromeaInstitut des Sciences et Inge´nierie Chim Lausanne (EPFL), 1015 Lausanne, Switzer +41 21-693-9305 bGZG, Abteilung Kristallographie, Goldschmidtstr. 1, 37077 Go¨ttingen, Germa cSwiss-Norwegian Beamline, ESRF, F-38043 dLaboratory of Crystallography, Ecole Poly 1015 Lausanne, Switzerland eLaboratory of X-ray Crystallography, Un Switzerland fGlobal Phasing Ltd., Sheraton House, Castl † Electronic supplementary information ( For ESI and crystallographic data in CI 10.1039/c4sc03046j Cite this: DOI: 10.1039/c4sc03046j Received 3rd October 2014 Accepted 3rd November 2014 DOI: 10.1039/c4sc03046j www.rsc.org/chemicalscience This journal is © The Royal Society oftallic coordination cages based on ditopic metallo-ligands with 3-pyridyl donor groups† Matthew D. Wise,a Julian J. Holstein,bf Philip Pattison,cd Celine Besnard,e Euro Solari,a Rosario Scopelliti,a Gerard Bricognef and Kay Severin*a Ditopic N-donor ligands with terminal 4-pyridyl groups are omnipresent in coordination-based self- assembly. The utilization of ligands with 3-pyridyl donor groups is significantly less common, because the intrinsic conformational flexibility of these ligands tends to favor the formation of small aggregates. Here, we show that large Pd6L12 12+ cages can be obtained by reaction of Pd(II) salts with metallo-ligands L bearing terminal 3-pyridyl groups. The easy-to-access metallo-ligands contain an Fe(II) clathrochelate core. These sterically demanding clathrochelate complexes prevent the formation of smaller aggregates, which is observed for less bulky analogous building blocks. The cages were shown to bind BF4  and BPh4  anions in aqueous solvent mixtures, whilst the lateral size of the clathrochelate significantly affects their guest encapsulation behavior.Introduction Coordination-based self-assembly is underpinned by a set of well-established design principles. These geometrical tenets have enabled chemists to rationally target molecular architec- tures possessing an enormous range of structural and func- tional characteristics.1,2 The directional bonding approach is perhaps the most intuitive design strategy.1g,3 This approach entails the combination of donor (ligand) and acceptor (metal complex) units, the geometries of which determine the struc- ture of the assembly formed. A fundamental requirement of this strategy is that these building blocks are shape-persistent; possessing conformational rigidity and well-dened coordinate vectors. The self-assembly behavior of building blocks without these characteristics is less predictable and, consequently, their strategic incorporation into supramolecular structures inher- ently difficult.iques, Ecole Polytechnique Fe´de´rale de land. E-mail: kay.severin@ep.ch; Fax: Georg-August-Universita¨t Go¨ttingen, ny Grenoble, Switzerland technique Fe´de´rale de Lausanne (EPFL), iversity of Geneva, CH-1211-Geneva 4, e Park Cambridge CB3 0AX, England ESI) available. CCDC 1024735–1024740. F or other electronic format see DOI: Chemistry 2014Ditopic N-donor ligands with terminal 4-pyridyl groups are amongst the most extensively exploited building blocks in the preparation of coordination-based supramolecular assemblies. The simplest member of this family of tectons is 4,40-bipyridine, which has been incorporated into countless discrete1 and polymeric self-assembled structures.3 The insertion of a linker group between the 4-pyridyl moieties has led to ever more elaborate assemblies.4,5 In contrast to the ubiquity of 4-pyridyl terminated N-donor ligands, closely related 3-pyridyl termi- nated ligands are far less common.1–3 Scheme 1 alludes to why this is the case. The coordinate vectors of a generic ditopic N-donor ligand L1 with terminal 4-pyridyl groups are xed, regardless of the linking group R between the heterocycles, provided that R is rigid. Free rotation about the R–pyridine bond does not affect the orientation of the non-bonding sp2Scheme 1 For ditopic ligands L2 with 3-pyridyl donor groups, the relative orientation of the coordinate vectors is more flexible than for ligands L1 with 4-pyridyl donor groups. Chem. Sci. Scheme 3 Synthesis of the bipyridyl ligands 1 and 2. Chemical Science Edge Article O pe n A cc es s A rti cl e. P ub lis he d on 1 1 N ov em be r 2 01 4. D ow nl oa de d on 1 9/ 11 /2 01 4 16 :2 3: 20 . Th is ar tic le is li ce ns ed u nd er a C re at iv e Co m m on s A ttr ib ut io n 3. 0 U np or te d Li ce nc e. View Article Onlineorbitals of the N atoms (Scheme 1a). However, in the case of a generic 3-pyridyl terminated ligand L2, rotation about this bond changes the coordinate vectors of the pyridine N atoms (Scheme 1b). As a consequence of the conformational exibility of 3-pyridyl terminated N-donor ligands, entropically favored small6 aggre- gates are obtained upon coordination to metal ions. Several research groups have shown that the combination of Pd2+ with ‘banana-shaped’ ligands of type L2 (a < 180) typically results in the formation of dinuclear aggregates of the formula [Pd2(L2)4] 4+ (Scheme 2).7,8 Fujita and co-workers have examined the reaction of Pd2+ with linear ligands of type L2 (a¼ 180) and observed the formation of [Pd3(L2)6] 6+ and [Pd4(L2)8] 8+ complexes.9 These results are in contrast to reactions of Pd2+ with ligands of type L1, which were found to give polynuclear coordination cages of type [Pd12(L1)24] 24+ and [Pd24(L1)48] 48+.4 Inspection of the assemblies obtained with L2-type ligands reveals that they all feature Pd(L2)2Pd macrocycles as part of their structure.10 We hypothesized that increasing the steric bulk of L2-type ligands would overcome their entropic propensity to form small aggregates. Bulky lateral R group substituents should disfavor Pd(L2)2Pd macrocycles and thus result in more expanded assemblies. Below we show that this strategy can indeed be used to access unprecedented, large coordination cages based on ditopic 3-pyridyl ligands. In particular, we demonstrate that clathrochelate-based metal- loligands enable the synthesis of octahedral [Pd6(L2)12] 12+ cages with a diameter of more than 20 A˚. Despite the lateral size of the metalloligands, the encapsulation of reasonably large guest molecules such as tetraphenylborate (BPh4 ) is possible.Results and discussion Bipyridyl ligands 1 and 2 were isolated in 96% and 79% yield respectively, following a one-pot synthesis similar to that previously described for 4-pyridyl functionalized clathrochelates (Scheme 3).11 All the starting materials used inScheme 2 Reactions of Pd2+ with ligands of type L2 give di-, tri- and tetranuclear aggregates, whereas large coordination cages are obtained upon reaction with ligands of type L1. Chem. Sci.this procedure—1,2-cyclohexanedione dixoime (nioxime) or dimethylglyoxime, anhydrous iron(II) chloride and pyr- idin-3-ylboronic acid—are commercially available. Single crystal X-ray diffraction unambiguously conrmed the formation of 1 and 2 (Fig. 1). The crystal structures reveal the conformational exibility of the coordination vectors of these 3-pyridyl appended ligands. In 1, the two planes of the pyridine rings are close to orthogonal (85.8). Conversely, in 2, they are almost coplanar (7.8), yet the nonbonding sp2 orbitals of the pyridine N atoms are essentially parallel and divergent. The bond lengths and angles found for the trigonal prismatic Fe core of 1 and 2 are within the range observed for other Fe-based clathrochelate complexes.12 The inherent conformational exibility of 1 and 2makes the prediction of the structure of an assembly comprising these building blocks difficult, even when combined with a metal acceptor of well-dened geometry, such as a naked Pd2+ ion (square planar geometry). As outlined above, we envisaged that the considerable lateral steric bulk of 1 and 2 would play a structure-directing role and prevent the formation of macrocy- clic Pd(L2)2Pd motifs. Consequently, we anticipated that larger cage structures could be formed by combination of 1 or 2 and a naked Pd2+ ion, rather than small aggregates. NMR-scale experiments were subsequently undertaken to investigate this hypothesis. A solution of [Pd(CH3CN)4](BF4)2 inFig. 1 Molecular structures of clathrochelate-based bipyridyl ligands 1 (left) and 2 (right) as determined by X-ray crystallography. Color coding: C: gray, B: green, Fe: orange, N: blue, O: red. Hydrogen atoms and solvent molecules are omitted for clarity. This journal is © The Royal Society of Chemistry 2014 Fig. 2 Molecular structure of cages 5 (top) and 6 (bottom) determined by X-ray crystallography. Color coding: C: gray, B: green, Fe: orange, N: blue, O: red, Pd: cyan. Anions and hydrogen atoms are omitted for clarity. Edge Article Chemical Science O pe n A cc es s A rti cl e. P ub lis he d on 1 1 N ov em be r 2 01 4. D ow nl oa de d on 1 9/ 11 /2 01 4 16 :2 3: 20 . Th is ar tic le is li ce ns ed u nd er a C re at iv e Co m m on s A ttr ib ut io n 3. 0 U np or te d Li ce nc e. View Article Onlineacetonitrile-d3 was added to two equivalents of solid 1 or 2, both of which are poorly soluble in acetonitrile-d3 at room temper- ature, giving a pale yellow suspension. These samples were then heated at 70 C and 1H NMR spectra were recorded periodically. Over time, the reaction mixtures progressively became less turbid and turned red. Aer 3 hours, neither solution contained any precipitate and a single set of sharp peaks was observed in the 1H NMR spectrum in both cases, suggesting the formation of a single highly symmetric, discrete species (ESI, Fig. S19 and S20†). The syntheses were subsequently performed on a preparative scale, without deuterated solvent (Scheme 4). The ESI-MS spectra of the products show major peaks which can be attributed to complexes of the formula (Pd6(L2)12) (BF4)n 12n+ (n¼ 4–6, L2¼ 1 or 2, ESI Fig. S1 and S2†). This result conrmed that higher order assemblies (3, 4) had indeed been formed from 1 and 2 respectively, rather than small aggregates. The complexes 5 and 6 were prepared in an analogous fashion by heating a suspension of 1 or 2 and Pd(NO3)2(H2O)n in a 5 : 1 mixture of acetonitrile : water for 2 hours at 70 C. The 1H NMR spectra once more showed a single set of signals for each product and ESI-MS conrmed the molecular formulas of the cations to be [(Pd6(1)12)] 12+ and [(Pd6(2)12)] 12+ (ESI, Fig. S3 and S4†). Isolation of all complexes was achieved by precipitation with diethyl ether (isolated yields: 82–94%). Single crystals of 3 suitable for X-ray diffraction were grown by diffusion of diethylether into a solution of the assembly in acetonitrile, whilst those of 5 and 6 were grown by layering an acetonitrile solution onto toluene. Synchrotron radiation was required to obtain diffraction data of sufficient quality, which were collected at the Swiss Norwegian Beamline at ESRF Gre- noble. Assemblies of this size pose a number of difficulties due in part to their extensive inherent disorder and high solvent content within the crystal. Synchrotron radiation mitigates some of these difficulties, but we have also employed a series of carefully and rigorously adapted macromolecular renement techniques13 in order to build a molecular model (see ESI† for details). Fig. 2 shows the molecular structures of 5 and 6 in the crystal (the structure of 3 is similar to that of 5 and thus not shown). In all structures, the six Pd2+ ions of the assembly form an octa- hedron of close to regular geometry, with Pd–Pd distances of 2.19–2.32 nm across the body of the framework, and 1.52–1.63Scheme 4 Synthesis of the cages 3–6. This journal is © The Royal Society of Chemistry 2014nm within each triangular face. The internal angles of the facial triangles vary from 57.3 to 61.4. The structural regularity of the Pd framework is not, however, reected in the conformation of the bridging ligands themselves. In 5, the planes of the pyridine rings of the six ligands along the edges of two opposing faces of the octahedral framework deviate signicantly from parallel (plane dihedral angles 90.0, 89.2 and 66.8). In contrast, the planes of the pyridine rings of the remaining six ligands are much closer to parallel (20.6, 9.3 and 13.7). The situation is yet more disordered in the cases of 3 and 6, where a range of dihedral angles between pyridine rings within each clathrochelate is observed (10.8–89.9). This illustrates the intrinsic conformational exibility of ligands of the type L2. Their variable coordination vectors determine the most ener- getically favorable conformation of the assembly as a whole, and enable this arrangement to be attained. However, the different conformations adopted by 1 and 2 in the self-assembled architectures 3–6 are not manifested in the solution phase NMR spectra at room temperature. Only one set of peaks is visible, implying that the pyridyl rings of the ligandsChem. Sci. Chemical Science Edge Article O pe n A cc es s A rti cl e. P ub lis he d on 1 1 N ov em be r 2 01 4. D ow nl oa de d on 1 9/ 11 /2 01 4 16 :2 3: 20 . Th is ar tic le is li ce ns ed u nd er a C re at iv e Co m m on s A ttr ib ut io n 3. 0 U np or te d Li ce nc e. View Article Online1 and 2, as well as the tris(dioxime) clathrochelate core, undergo rapid conformational changes (on the NMR timescale). The 1H NMR spectra of 3 and 4 were also recorded at 233 K (CD3CN). Signicant line broadening was observed in all cases (ESI, Fig. S23 and S24†). The extent of line broadening was far greater in the case of 4 than 3. This result is unsurprising, given the bulky cyclohexyl substituents of ligand 2 impose a higher steric barrier to rotation than the relatively small methyl substituents of ligand 1. The octahedral cages possess internal void space, which is largely occupied by solvent molecules and anions (both ordered and disordered) in the crystal. The X-ray structures of 3, 5 and 6 show NO3  or BF4  anions in interior binding pockets adjacent to the Pd2+ ions at the vertices of the octahedral assemblies. The presence of close anion–Pd contacts is in line with what has been observed for many cages of type [Pd2(L2)4] 4+.7 Simple geometric approximations reveal the volume of the octahedra described by the six Pd2+ ions in 3, 5 and 6 to be 5.2 nm3. However, much of the interior volume of the assembly is occupied by clathrochelate oxime substituents.14 The inuence of the lateral size of the clathrochelate complexes over the void volume within each assembly was quantied by VOIDOO calculations (see ESI† for details).15 In the solid state, assembly 3 possesses a cavity volume of 1.4 nm3 or 1.8 nm3 (two inde- pendent cages found in the asymmetric unit of the crystal structure), cage 5 a cavity volume of 1.3 nm3 and cage 6 a much smaller cavity volume of 1.1 nm3. From these values it is clear that it is not only the oxime substituents that signicantly affect the internal space within each assembly, but also the confor- mation of the clathrochelate metalloligands. However, given the dynamic and uxional nature of these cages evidenced by 1H NMR, this conformational inuence was not expected to contribute to the solution phase void volume, and hence guest binding behavior, of assemblies 3–6. Any differences in solution phase behavior can be attributed to the oxime substituents alone. To investigate whether oxime steric bulk would inuence the kinetics and thermodynamics of guest binding, 19F NMR spectra of 3 and 4 were recorded at 298 K in acetonitrile-d3. A single peak was observed at 150.55 ppm for the BF4 anion in 3, whereas peaks at 146.37 ppm and 151.98 ppm were observed in the case of 4 (referenced to hexauorobenzene at 164.90 ppm). These data suggest that the rate of exchange between internal and external BF4  is fast on the NMR time- scale at 298 K in the case of 3 (one signal), but slow in the case of 4 (two signals). 19F NMR spectra of 3 and 4 were subsequently recorded over a range of temperatures (see ESI, Fig. S25 and S26†). These experiments revealed a coalescence temperature of approximately 255 K for the BF4  anion signals of 3, and 345 K for those of 4. The difference of 90 K corresponds to a difference of 16 kJ mol1 for the activation free energy16 of the anion exchange process and neatly illustrates one key characteristic of clathrochelate-based building blocks in general—the ease with which we can modulate their steric bulk. In addition to the anion binding sites situated adjacent to the Pd2+ metal centers, cages 3–6 possess a central cavity. The internal environment of this cavity is hydrophobic due to the methyl or cyclohexyl substituents of the clathrochelate ligands,Chem. Sci.yet the overall charge of the cages is 12+. Consequently, we expected coulombic interactions and the hydrophobic effect to dominate the guest binding behavior of assemblies 3–6. Therefore, the lipophilic tetraphenylborate (BPh4 ) anion appeared to be a potentially suitable guest molecule. The BPh4  anion was of particular interest not only given its perceived size and electrostatic match, but also the limited number of host complexes that have been found to bind BPh4  as a guest. To the best of our knowledge, only certain cyclodextrins,17 cav- itands,18 liposomes19 and the external face of a metal- losupramolecular cube20 have been observed to associate with BPh4  in a host—guest manner. Complexation studies with cage 4 were hampered by the formation of precipitates upon addition of NaBPh4, and ulti- mately abandoned. However, addition of one equivalent of NaBPh4 to 3 (0.62 mM) in CD3CN did not lead to precipitation, and signicant changes in the 1H NMR signals of both species were apparent.21 Two sets of peaks were immediately visible for the protons of the phenyl groups of the anionic guest, and a new set of signals was also observed for aromatic protons and methyl groups of the clathrochelate ligands. These observations were consistent with the formation of a host—guest complex between the cage and the BPh4  anion, the rate of exchange of bound and unbound BPh4  being slow on the NMR timescale. To conrm the encapsulation of the BPh4  anion within the interior cavity of 3, DOSY and NOESY NMR experiments were performed. DOSY NMR revealed a common diffusion coefficient (4.94 106 cm2 s1) for the assembly and one of the two sets of BPh4  protons (ESI, Fig. S40–S43†), whilst NOESY cross peaks were observed between the CH3 groups of the clathrochelate oxime ligand and the same set of BPh4  peaks (ESI, Fig. S46– S49†). The apparent association constant for the complexation of BPh4  by 3 was calculated through integration of the NMR signals corresponding to bound and unbound guest across a range of concentrations, using a 1 : 1 binding model (see ESI† for details). The resulting value is Ka ¼ 2.4(0.3)  103 M1. We subsequently recorded the 1H NMR spectra of 3 in the presence of one equivalent of NaBPh4 in a 2 : 1 mixture of CD3CN and D2O, and a 1 : 1 mixture of CD3CN and CDCl3 in order to elucidate the effect of solvent polarity upon BPh4  binding. In 2 : 1 CD3CN : D2O, the binding constant was calculated to be of the order of 105 M1, whereas in 1 : 1 CD3CN : CDCl3, the intensities of the peaks corresponding to bound BPh4  were too low to integrate reliably. The large increase in binding affinity observed upon changing the solvent from neat CD3CN to the more polar 2 : 1 CD3CN : D2O mixture, as well as the dramatic drop observed in the much less polar 1 : 1 CD3CN : CDCl3 mixture, enforces the hypothesis that the hydrophobic effect plays a signicant role in guest binding. These apparent asso- ciation constants take into account not only the encapsulation of BPh4  by 3, but also the thermodynamically unfavorable concomitant ejection of internal BF4 . Consequently, the affinity of 3 for BPh4  is, in isolation, presumably greater than the composite association constant recorded as discussed above.This journal is © The Royal Society of Chemistry 2014 Edge Article Chemical Science O pe n A cc es s A rti cl e. P ub lis he d on 1 1 N ov em be r 2 01 4. D ow nl oa de d on 1 9/ 11 /2 01 4 16 :2 3: 20 . Th is ar tic le is li ce ns ed u nd er a C re at iv e Co m m on s A ttr ib ut io n 3. 0 U np or te d Li ce nc e. View Article OnlineEncapsulation of BPh4  by 3 was unambiguously conrmed by single crystal X-ray crystallography. A ten- fold excess of NaBPh4 was added to a solution of 3 in 2 : 1 CD3CN : D2O, and from the resulting suspension single crystals of a mixed salt (7) containing both BPh4  and BF4  anions were isolated. Despite the relatively poor quality of the diffraction data, due to the size and complexity of the struc- ture, a sufficient solution was obtained, which unquestion- ably established the connectivity within the [Pd6112] assembly, as well as the presence of encapsulated BPh4 . Unsurprisingly, it was impossible to locate all BF4  and BPh4  anions due to the extent to which they are disordered, hence precise formulation of this salt cannot be undertaken using the X-ray diffraction data. Rather, the general formula [BPh4@Pd6112][BF4]m[BPh4]n, where m + n ¼ 11, must be used to describe 7. Two independent cages were found in the asymmetric unit, each of which contained a single encapsu- lated BPh4  (see ESI, Fig. S31†). One of the two independent cages additionally contained three ordered BF4  (Fig. 3), whilst a single ordered BF4  was located in the second. The positions of the different anions in the former illustrate the differences in size of the two interior cavities of 3—the smaller BF4  anions occupy binding pockets adjacent to the Pd2+ vertices of one face of the octahedral framework, whilst the non-coordinating BPh4  resides within the larger central cavity of the assembly. Two of the phenyl groups of BPh4  sit snugly between the methyl substituents of the clathrochelate oxime ligands of two neighboring faces of the [Pd6112] 12+ assembly. This conformation alludes to the contribution of hydrophobic interactions to BPh4  encapsulation, and is consistent with the upeld change in chemical shi observed for the protons of the CH3 groups of the oxime substituents as they are exposed to the diamagnetic ring current of the phenyl rings (see ESI, Fig. S37†).Fig. 3 Part of the X-ray crystal structure of 7, highlighting the encapsulated BF4  (yellow-green) and BPh4  (magenta) anions. External anions, hydrogen atoms and solvent molecules are omitted for clarity. This journal is © The Royal Society of Chemistry 2014Conclusions Herein, we have reported the syntheses and the structures of two clathrochelate-basedmetalloligands with terminal 3-pyridyl groups (1 and 2). Reactions of these ligands with Pd2+ afford large octahedral coordination cages (3–6), which were compre- hensively characterized (including crystallographic analysis for three of the four cages). The formation of octahedral Pd cages from ditopic ligands containing 3-pyridyl groups is unprece- dented. Typically, conformationally exible ligands with terminal 3-pyridyl groups give rise to small aggregates (Scheme 2). In our case, the lateral size of the metalloligands prevents the formation of macrocyclic Pd(L2)2Pd structures. This enthalpic effect (steric hindrance) is sufficient to overcome the entropi- cally driven propensity of a coordination-based supramolecular system to form small aggregates. Consequently, larger octahe- dral structures are obtained, the assembly of which is favored to such an extent that no side products are observed. Despite the fact that we have used steric effects to enlarge the assembly, we were still able to obtain cages with rather large cavities. Notably, cage 3 was found to bind the ‘non-coordinating’22 anion BPh4  with a solvent-dependent association constant of the order of 105 M1. Our work also highlights the advantageous characteristics of clathrochelate complexes as building blocks for supramolecular chemistry.11,23 These complexes are straightforward to synthe- size and their lateral size and functionality24 can be modulated by the boronic acid capping groups and the oxime substituents. The ability to modify the structure of the metalloligands without substantial synthetic efforts represents an important advantage for implementing or optimizing a certain function of the nal assembly. As a proof-of-principle study, we have shown that the kinetics of BF4  exchange are strongly affected by the nature of the oxime-derived side chain (methyl vs. cyclohexyl). We believe that judicious design of clathrochelate building blocks will enable many more structurally and functionally novel self-assembled architectures to be obtained, and we are continuing to pursue the applications of these ligands in coordination-based self-assembly. Acknowledgements The research leading to these results has received funding from the People Programme (Marie Curie Actions) of the European Union's Seventh Framework Programme FP7/2007–2013/under REA grant agreement no 264645 (MDW and JJH). We thank Mathieu Marmier, EPFL, for help with the graphical material. Notes and references 1 For selected recent review articles about molecularly dened assemblies see: (a) S. Mukherjee and P. S. Mukherjee, Chem. Commun., 2014, 50, 2239; (b) M. D. Ward and P. R. Raithby, Chem. Soc. Rev., 2013, 42, 1619; (c) J. G. Hardy, Chem. Soc. Rev., 2013, 42, 7881; (d) N. J. Young and B. P. Hay, Chem. Commun., 2013, 49, 1354; (e) T. K. Ronson, S. Zarra, S. P. Black and J. R. Nitschke, Chem. Commun., 2013, 49,Chem. Sci. Chemical Science Edge Article O pe n A cc es s A rti cl e. P ub lis he d on 1 1 N ov em be r 2 01 4. D ow nl oa de d on 1 9/ 11 /2 01 4 16 :2 3: 20 . Th is ar tic le is li ce ns ed u nd er a C re at iv e Co m m on s A ttr ib ut io n 3. 0 U np or te d Li ce nc e. View Article Online2476; (f) H. Amouri, C. Desmarets and J. Moussa, Chem. Rev., 2012, 112, 2015; (g) R. Chakrabarty, P. S. Mukherjee and P. J. Stang, Chem. Rev., 2011, 111, 6810; (h) M. J. Wiester, P. A. Ulmann and C. A. Mirkin, Angew. Chem., Int. Ed., 2011, 50, 114. 2 For selected recent review articles about polymeric assemblies see: (a) W. Lu, Z. Wei, Z.-Y. Gu, T.-F. Liu, J. Park, J. Park, J. Tina, M. Zhang, Q. Zhang, T. Gentle III, M. Bosch and H.-C. Zhou, Chem. Soc. Rev., 2014, 43, 5561– 5593; (b) V. Guillerm, D. Kim, J. F. Eubank, R. Luebke, X. Liu, K. Adil, M. S. Lah and M. Eddaoudi, Chem. Soc. Rev., 2014, 43, 6141; (c) H. Furukawa, K. E. Cordova, M. O'Keeffe and O. M. Yaghi, Science, 2013, 341, 974; (d) T. R. Cook, Y.-R. Zheng and P. J. Stang, Chem. Rev., 2013, 113, 734; (e) W. Xuan, C. Zhu, Y. Liu and Y. Cui, Chem. Soc. Rev., 2012, 41, 1677; (f) N. N. Adarsh and P. Dastidar, Chem. Soc. Rev., 2012, 41, 3039; (g) F. A. Almeida, J. Klinowski, S. M. F. Vilela, J. P. C. Tome´, J. A. S. Cavaleiro and J. Rocha, Chem. Soc. Rev., 2012, 41, 1088; (h) N. Stock and S. Biswas, Chem. Rev., 2012, 112, 933. 3 B. H. Northrop, D. Chercka and P. J. Stang, Tetrahedron, 2008, 64, 11495. 4 Review: K. Harris, D. Fujita and M. Fujita, Chem. Commun., 2013, 49, 6703. 5 Selected examples: (a) S.-L. Huang, Y.-J. Lin, T. S. A. Hor and G.-X. Jin, J. Am. Chem. Soc., 2013, 135, 8125; (b) D. Fujita, K. Suzuki, S. Sato, M. Yagi-Utsumi, Y. Yamaguchi, N. Mizuno, T. Kumasaka, M. Takata, M. Noda, S. Uchiyama, K. Kato and M. Fujita, Nat. Commun., 2012, 3, 1093; (c) J. Bunsen, J. Iwasa, P. Bonakdarzadeh, E. Numata, K. Rissanen, S. Sato and M. Fujita, Angew. Chem., Int. Ed., 2012, 51, 3161; (d) Q.-F. Sun, T. Murase, S. Sato and M. Fujita, Angew. Chem., Int. Ed., 2011, 50, 10318; (e) Q.-F. Sun, J. Iwasa, D. Ogawa, Y. Ishido, S. Sato, T. Ozeki, Y. Sei, K. Yamaguchi and M. Fujita, Science, 2010, 328, 1144. 6 The notion ‘small’ refers here to the formation of aggregates containing a small number of building blocks. They can be large in terms of size if large ligands are employed. 7 Review: M. Han, D. M. Engelhard and G. H. Clever, Chem. Soc. Rev., 2014, 43, 1848. 8 Selected examples: (a) J. E. M. Lewis, A. B. S. Elliott, C. J. McAdam, K. C. Gordonab and J. D. Crowley, Chem. Sci., 2014, 5, 1833; (b) C. Desmarets, G. Gontard, A. L. Cooksy, M. N. Rager and H. Amouri, Inorg. Chem., 2014, 53, 4287; (c) M. Han, R. Michel, B. He, Y.-S. Chen, D. Stalke, M. John and G. H. Clever, Angew. Chem., Int. Ed., 2013, 52, 1319; (d) G. H. Clever, W. Kawamura, S. Tashiro, M. Shiro and M. Shionoya, Angew. Chem., Int. Ed., 2012, 51, 2606; (e) J. E. M. Lewis, E. L. Gavey, S. A. Cameron and J. D. Crowley, Chem. Sci., 2012, 3, 778; (f) M. Han, J. Hey, W. Kawamura, D. Stalke, M. Shionoya and G. H. Clever, Inorg. Chem., 2012, 51, 9574; (g) N. Kishi, Z. Li, K. Yoza, M. Akita and M. Yoshizawa, J. Am. Chem. Soc., 2011, 133, 11438; (h) P. Liao, B. W. Langloss, A. M. Johnson, E. R. Knudsen, F. S. Tham, R. R. Julian and R. J. Hooley, Chem. Commun., 2010, 46, 4932; (i) G. H. Clever, S. Tashiro and M. Shionoya, Angew. Chem., Int. Ed., 2009, 48, 7010.Chem. Sci.9 D. K. Chand, K. Biradha, M. Kawano, S. Sakamoto, K. Yamaguchi and M. Fujita, Chem.–Asian J., 2006, 1, 82. 10 The reaction of linear ligands of type L2 with cis-blocked (L- L)Pd2+ complexes gives also (L-L)Pd(L2)2Pd(L-L) macrocycles. See: M.-E. Moon, H.-J. Lee, B.-S. Ko and K.-W. Chi, Bull. Korean Chem. Soc., 2007, 28, 311. 11 M. D. Wise, A. Ruggi, M. Pascu, R. Scopelliti and K. Severin, Chem. Sci., 2013, 4, 1658. 12 Recent examples include: (a) O. A. Varzatskii, S. V. Shul'ga, A. S. Belov, V. V. Novikov, A. V. Dolganov, A. V. Vologzhanina and Y. Z. Voloshin, Dalton Trans., 2014, DOI: 10.1039/c4dt01557f; (b) O. A. Varzatskii, I. N. Denisenko, A. S. Belov, A. V. Vologzhanina, Y. N. Bubnov, S. V. Volkov and Y. Z. Voloshin, Inorg. Chem. Commun., 2014, 44, 134; (c) V. V. Novikov, I. V. Ananyev, A. A. Pavlov, M. V. Fedin, K. A. Lyssenko and Y. Z. Voloshin, J. Phys. Chem. Lett., 2014, 5, 496; (d) O. A. Varzatskii, I. N. Denisenko, S. V. Volkov, A. V. Dolganov, A. V. Vologzhanina, Y. N. Bubnov and Y. Z. Voloshin, Inorg. Chem. Commun., 2013, 33, 147; (e) O. A. Varzatskii, I. N. Denisenko, S. V. Volkov, A. S. Belov, A. V. Dolganov, A. V. Vologzhanina, V. V. Novikov, Y. N. Bubnov and Y. Z. Voloshin, Eur. J. Inorg. Chem., 2013, 3178. 13 (a) A. Thorn, B. Dittrich and G. M. Sheldrick, Acta Cryst. A, 2012, 68, 448; (b) O. S. Smart, T. O. Womack, C. Flensburg, P. Keller, W. Paciorek, A. Sharff, C. Vonrhein and G. Bricogne, Acta Cryst. D, 2012, 68, 368. 14 For examples of octahedral Pd cages see: (a) C. Gu¨tz, R. Hovorka, C. Klein, Q.-Q. Jiang, C. Bannwarth, M. Engeser, C. Schmuck, W. Assenmacher, W. Mader, F. Topic´, K. Rissanen, S. Grimme and A. Lu¨tzen, Angew. Chem., Int. Ed., 2014, 53, 1693; (b) K. Li, L.-Y. Zhang, C. Yan, S.-C. Wei, M. Pan, L. Zhang and C.-Y. Su, J. Am. Chem. Soc., 2014, 136, 4456; (c) K. Suzuki, M. Tominaga, M. Kawano and M. Fujita, Chem. Commun., 2009, 1638. 15 G. J. Kleywegt and T. A. Jones, Acta Crystallogr., Sect. D: Biol. Crystallogr., 1994, 50, 178–185. 16 J. Sandstro¨m, Dynamic Nmr Spectroscopy, Academic Press, 1982. 17 (a) M. D. Johnson and V. C. Reinsborough, Aust. J. Chem., 1992, 45, 1961; (b) T. Nhujak and D. M. Goodall, Electrophoresis, 2001, 22, 117. 18 S. S. Zhu, H. Staats, K. Brandhorst, J. Grunenberg, F. Gruppi, E. Dalcanale, A. Lu¨tzen, K. Rissanen and C. A. Schalley, Angew. Chem., Int. Ed., 2008, 47, 788. 19 H. Osanai, T. Ikehara, S. Miyauchi, K. Shimono, J. Tamogami, N. Nara and N. Kamo, J. Biophys. Chem., 2013, 4, 11. 20 W. J. Ramsay and J. R. Nitschke, J. Am. Chem. Soc., 2014, 136, 7038. 21 Binding studies were not performed with cage 5 and 6 because they display a lower solubility compared to 3 and 4. 22 Receptors for weakly-coordinating anions are described in: (a) S. Lee, C.-H. Chen and A. H. Flood, Nat. Chem., 2013, 5, 704; (b) Y. R. Hristova, M. M. J. Smulders, J. K. Clegg, B. Breiner and J. R. Nitschke, Chem. Sci., 2011, 2, 638.This journal is © The Royal Society of Chemistry 2014 Edge Article Chemical Science O pe n A cc es s A rti cl e. P ub lis he d on 1 1 N ov em be r 2 01 4. D ow nl oa de d on 1 9/ 11 /2 01 4 16 :2 3: 20 . Th is ar tic le is li ce ns ed u nd er a C re at iv e Co m m on s A ttr ib ut io n 3. 0 U np or te d Li ce nc e. View Article Online23 (a) Y.-Y. Zhang, Y.-J. Lin and G.-X. Jin, Chem. Commun., 2014, 50, 2327; (b) M. Pascu, M. Marmier, C. Schouwey, R. Scopelliti, J. J. Holstein, G. Bricogne and K. Severin, Chem.–Eur. J., 2014, 20, 5592.This journal is © The Royal Society of Chemistry 201424 For clathrochelate complexes with different functional groups see: E. G. Lebed, A. S. Belov, A. V. Dolganov, A. V. Vologzhanina, A. Szebesczyk, E. Gumienna-Kontecka, H. Kozlowski, Y. N. Bubnov, I. Y. Dubey and Y. Z. Voloshin, Inorg. Chem. Commun., 2013, 30, 53.Chem. Sci.