Endo- and Exocyclic Coordination of a 20-Membered N2

Endo- and Exocyclic Coordination of a 20-Membered N2...
8 downloads 3 Views 5MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Endo- and Exocyclic Coordination of a 20-Membered N2O2S2‑Macrocycle and Cascade Complexation of a 40-Membered N4O4S4‑Macrocycle Eunji Lee,† Seul-Gi Lee,† In-Hyeok Park,† Seulgi Kim,† Huiyeong Ju,† Jong Hwa Jung,† Mari Ikeda,‡ Yoichi Habata,*,§ and Shim Sung Lee*,† †

Department of Chemistry and Research Institute of Natural Science, Gyeongsang National University, Jinju 52828, South Korea Education Center, Faculty of Engineering, Chiba Institute of Technology, 2-1-1 Shibazono, Narashino, Chiba 275-0023, Japan § Department of Chemistry, Toho University, 2-2-1 Miyama, Funabashi, Chiba 274-8510, Japan ‡

S Supporting Information *

ABSTRACT: A 20-membered N2O2S2-macrocycle (L1) and a 40membered N4O4S4-macrocycle (L2) were employed as a [1:1] and a [2:2] cyclization product, respectively, for the preparation of diverse types of supramolecular complexes including a cascade complex. Six complexes (1−6) of the smaller macrocycle L1 including discrete to continuous forms, mono- to heteronuclear, and endo- to exo- and endo/ exocoordination were prepared and their coordination modes were discussed systemically. First, the reaction of L1 with CuI in the presence of trifluoroacetic acid afforded an exocyclic 1-D coordination polymer {[(μ4-Cu4I4)(HL1)2](CF3COO)2}n (1). Meanwhile, the reaction of L1 with Cu(ClO4)2·6H2O afforded a typical endocyclic mononuclear complex [CuII(L1)](ClO4)2·H2O (2). In the reactions of L1 with CdX2 (X = Br and I), isostructural sandwich-type complexes [Cd(L1)2Br2] (3) and [Cd(L1)2I2] (4) were isolated. The treatment of L1 with Hg(ClO4)2 also afforded a sandwich-type complex [Hg(L1)2](ClO4)2 (5). One-pot reaction of L1 with a mixture of HgI2 and CdI2 afforded a dumbbell-type heteronuclear complex {[Cd(L1)]2(μ-Hg2I6)}[Hg2I6] (6), in which the Cd(II) ion occupies the macrocyclic cavity. Further, such two endocyclic Cd(II) complex units are bridged by a square-type (μ-Hg2I6)2− cluster remaining another same cluster separately. The comparative NMR data exhibited a higher affinity of Cd(II) over Hg(II) toward L1, in the parallel to the situation occurred in the solid state. Meanwhile, complexations of the extra-large macrocycle L2 is more challenging to afford some interesting dimercury(II) coordination products including a cascade complex. In solution, the dimercury(II) perchlorato complex of L2 as a metalloligand shows a preferential binding of dabco (1,4-diazabicyclo[2,2,2]octane), but its dimercury(II) iodo complex has a much smaller affinity for dabco. In order to explain these results, the solid dimercury(II) complexes with different anions [Hg2(L2)X4] (7: X = I, 8: X = ClO4) were prepared and characterized. Further, the dimercury(II) perchlorato complex 8 reacts with dabco to forms a cascade complex [Hg2(L2)(μ-dabco)(ClO4)2](ClO4)2·2DMF·2ether (9), exhibiting its formation being metal-driven and coordinated anion-regulated. The observed cascade complexation both in solution and solid states is an example of the adaptive guest binding.



INTRODUCTION Conventional crown-type macrocycles, which form metal-incavity type (endocyclic coordination) complexes, show extraordinary stability and high selectivity toward the given guest due to the macrocyclic effect but relatively narrow ranges of structural topologies compared to those with acyclic analogues.1 Meanwhile, thiamacrocycles, including sulfurcontaining mixed donor macrocycles, so often tend to form exocyclic complexes which demonstrate the metal coordination outside the cavity.2 We3 and other groups4 have reported a series of soft and hard/soft metal complexes of sulfurcontaining mixed donor macrocycles with unusual stoichiometries and topologies. This has led us to explore the coordination chemistry of thiamacrocycle system, particularly with reference © XXXX American Chemical Society

to the possible coordination modes of these macrocycles more precisely. In our recent works, the structural characteristics of the thiaphilic metal complexes of the 16−18 membered O2S2and O2S3-macrocycles have been systematically discussed in terms of the possible exocoordination modes.5 In this approach, it is understood that bischelating dithiamacrocycles rigidified with aromatic subunit(s) tend to show convergent or divergent exocoordination modes, which might lead the formation of discrete or continuous coordination species, respectively.6 Some thiaoxaaza-macrocycles also show a variety of coordinaReceived: January 17, 2018

A

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

type cryptands; these semirigid macrobicycles form dinuclear cryptates that also uptake an anion via the cascade complexation process.12 Some other macrocycles exhibiting the adaptive guest binding have been reported.13 In this work, we assumed that the extra-large macrocycle L2 with a flexible cavity might be a candidate for the cascade complex type adaptive guest binding system. From the preliminary studies, we make the following observations for L2:8a (i) The macrocycle is capable of binding two soft metal ions. (ii) The macrocycle is probably capable of binding an organic guest but not in the flexible free form alone. (iii) Upon dimetalation, the macrocycle becomes conformationally rigid. (iv) In the dimetalated complex, the space between the two metal ions might be accessible for an organic guest binding. Thus, the continuing interest on the coordination modes of L1 and the extra-large L2, which is expected to adopt the cascade complexations, have motivated us to extend these studies. Herein, we report the synthesis and structural characterization of nine complexes (1−9) of L1 and L2. Since one of the main purposes in this work is to propose the possible exocoordination modes of L1 and to realize their formations via the diverse reaction conditions, six complexes (1−6) of L1 with copper(I), copper(II), cadmium(II), and mercury(II) were prepared and their coordination modes are discussed. In particular, copper(I) iodide was employed to induce the formation of an exocoordination based network which is linked by CunIn-type clusters. An exciting feature among these is the formation of a dumbbell-type heteronuclear cadmium(II)/ mercury(II) complex in which two endocyclic complexes are bridged by organic or inorganic linker. Three complexes (7−9) of L2 are more challenging because of the anion-regulated formation of a new type cascade complex. Solution studies on the formations of the dumbbell-type heteronuclear complex and the cascade complex are also reported.

tion modes to the thiaphilic soft metals to generate diverse types of the products.3,4 On the other hand, the syntheses for the thiaoxa- and thiaoxaaza macrocycles via the dithiol−dihalide coupling reactions enabled us to isolate not only a 1:1 cyclization product but also a 2:2 cyclization product.7,8 In practice, a 24membered N2S6-macrocycle has been isolated as a minor product and its disilver(I) complex was prepared.7a Recently, we have introduced the isolation of a 20-membered N2O2S2macrocycle (L1) via a 1:1 cyclization and a 40-membered N4O4S4-macrocycle (L2) as a 2:2 cyclization product together with a mono- and a dimercury(II) complexes, respectively (Chart 1).8a More recently, we have reported the coordination behavior of these macrocycles including disilver(I) complexes of L2.8b Chart 1. 20-Membered (L1) and a 40-Membered (L2) Macrocycles

As mentioned, the large or extra-large macrocycles derived from the 2:2 cyclization have merely been regarded as potent dinucleating ligands so far.9 In fact, in biological molecular recognition of enzyme−substrate or protein−RNA interactions, some flexible receptors change their shapes to bind to guests precisely via an adaptive guest binding process.10 In this work, we propose that the dinuclear complexes of the extra-large macrocycle, L2, can be considered as metalloligands which include small organic guest between two metal centers via the adaptive binding process.10−13 Because such large macrocycles are capable of binding a small organic guest but not in the flexible form alone. As the adaptive guest binding process, the endocyclic (di)metalation of the large macrocycles might promote the inclusion of the organic guest.10−13 The dinuclear macrocyclic complexes in which small molecules or ions can be incorporated in the ligand cage as bridges between the metal-ion sites have been termed “cascade” complexes by J.-M. Lehn.11 For instance, they reported bistren



RESULTS AND DISCUSSION The macrocycles L1 and L2 were prepared as reported by us previously.8 We employed four metal species with halide or perchlorate anions and described the preparation and comparative structural properties of six complexes (1-6) for L1 and three complexes (7−9) for L2 with emphasis of the coordination modes characterized by single crystal X-ray analysis (Table S1). The bulk purity of the products was confirmed by PXRD patterns (Figure S1). Coordination Modes for Supramolecular Complexes of L1. We recently reported that L1 forms a series of the endocyclic silver(I) and mercury(II) complexes as well as anion-dependent nickel(II) complexes, adopting a perching-

Chart 2. Possible Coordination Modes of L1 toward Thiaphilic Metal Ions

B

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry type solvato-complex and a sandwich-type complex.8b For L1, four possible coordination modes are proposed as shown in Chart 2. In addition to the endocyclic coordination (Type I), the NS2 donor site using its chelating and bridging abilities can connect thiaphilic metal ions in three ways (Type II-a−c). These types of exocoordination modes seem to induce the diversity of the topological structures for its respective complexes. For example, the symmetrical divergent mode (Type II-a) may induce an infinite network, while the symmetrical convergent mode (Type II-b) or the unsymmetrical convergent mode (Type II-c) favors the formation of the sandwich-type or cyclic oligomer depending on other factors. Given the possibility of the diverse coordination modes, we have tried the preparation of novel complexes of L 1 incorporating some thiaphilic metal ions as well as copper(II). The results serve to both extend and exemplify the crystal engineering based on the coordination modes of sulfurcontaining mixed donor macrocycles. Preparation and Structures of Copper(I) and Copper(II) Complexes of L1 (1 and 2). The complexation behaviors of L1 with copper(I) and copper(II) were compared. Copper(I) iodide and copper(II) perchlorate complexes of L1 which show different coordination modes, topologies, and photophysical property were isolated as depicted in Scheme 1.

Figure 1. Crystal structure of {[(μ4-Cu4I4)(HL1)2](CF3COO)2}n (1): (a) the double-stranded 1-D polymeric chain and (b) basic coordination unit. The noncoordinated anions are omitted.

Scheme 1. Copper(I) and Copper(II) Complexes of L1

that two trifluoroacetate ions in the lattice link two protonated L1 molecules via hydrogen bonds (NH···O 1.777−2.294 Å) to form a square motif arrangement (Figure S2). Considering the growth of the single crystals of 1 in the presence of the trifluoroacetate, the hydrogen bonds stabilize the complex and contribute to the crystal packing. Similar Cu4I4 cluster-linked coordination polymers of 23-membered O4S2-macrocycle14 and 19-membered O2S2-macrocycle15 have been reported by us. Very recently, an adaptive formation of the copper(I) iodide clusters was proposed by employing bis-dithiamacrocycle isomers with different preorganized exocyclic binding sites.16 Under UV irradiation, 1 exhibits a bright-green emission at 520 nm (λex = 360 nm) which is likely due to the presence of a cluster-centered excited state with mixed halide-to-metal charge transfer character (Figure 2).17 The major contribution to the luminescent origin of [Cu4I4] is believed to be the clustercentered excited states (*Cu4), which involve Cu···Cu

Reaction of L1 with CuI in dichloromethane/acetonitrile afforded no product, while in the presence of 3 drops of trifluoroacetic acid a pale-yellow crystalline product 1 was obtained. X-ray analysis reveals that 1 crystallizes in the triclinic space group P-1 with Z = 2 and adopts a double-stranded 1-D coordination polymer with the formula {[(μ4-Cu4I4)(HL1)2](CF3COO)2}n (Figure 1). In case of the reaction of L1 with Cu(ClO4)2·6H2O, a green crystalline product 2 was isolated. The complex 2 crystallizes in the monoclinic space group C2/c with Z = 4, adopting a 1:1 complex with the formula [CuII(L1)](ClO4)2·H2O (Figure 3). In 1, L1 acts as a bidentate ligand linked by cubane-type fourway bridging clusters, (μ4-Cu4I4), via Cu−S bonds (2.285− 2.305 Å) adopting a divergent exocoordination mode (Type IIa in Chart 2). The preference of the generation of the 1-D chain topology in 1 is mainly due to the relatively longer sulfurto-sulfur separations (S1···S2 7.988 Å and S3···S4 7.992 Å) which are favorable to the exobridging mode.6 It is also found

Figure 2. Solid-state photoluminescence spectra of L1 and 1 at room temperature (λex 360 nm). C

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Scheme 2. Bis(macrocycle) Complexes (3−6) of L1

interactions. The Cu···Cu distances in 1 are in the range of 2.624(8)−2.911(8) Å, similar to those reported.18 The Cu(II) center in 2 is six-coordinate, being bound to all donors in L1, which adopts a twisted configuration (Figure 3).

Figure 3. Crystal structure of [CuII(L1)](ClO4)2·H2O (2). Noncoordinated anions and water molecule are omitted.

The coordination geometry in 2 can be described as a distorted octahedron, with the two oxygen atoms and two sulfur atoms from the macrocycle defining the square plane and the axial positions occupied by the two nitrogen atoms [N1−Cu1−N2 176.9(1)°]. Bond lengths of Cu−N [Cu1−N1 2.038(2) and Cu1−N2 2.016(2) Å] are shorter than those of Cu−O [Cu1− O1 2.346(2) and Cu1−O2 2.328(2) Å] and Cu−S [Cu1−S1 2.375(1) and Cu1−S2 2.356(1) Å], due to the Jan-Teller distortion. With respect to the oxidation state of copper, the preferred endocyclic mode (Type 1 in Chart 2) of 2 is due to the hard acid nature of copper(II) showing an oxophilicity, allowing all the N2O2S2-donors to bind tightly toward the metal center in the cavity. Preparation and Structures of Cadmium(II), Mercury(II), and Heteronuclear Cadmium(II)/Mercury(II) Complexes of L1 (3−6). Two heavy metals (Cd2+, Hg2+) were used in the complexations with L1 (Scheme 2). Cadmium(II) halides afforded isostructural anion-coordinated bis(macrocycle) complexes, while mercury(II) perchlorate gave a cationic bis(macrocycle) complex. A mixture of CdI2 and HgI2 afforded a dumbbell-type complex. The reaction of L1 with cadmium(II) halides in dichloromethane/methanol, isostructural 1:2 (metal-to-ligand) complexes [Cd(L1)2X2] (3: X = Br, 4: X = I) adopting a sandwichtype were obtained (Figure 4). In 3 and 4, each Cd2+ center which locates between two macrocycles is six-coordinate, being bound to one sulfur and one secondary nitrogen from L1 with the unsymmetrical convergent exocoordination mode (Type IIc in Chart 2). The coordination sphere is completed by two halide anions. In the macrocycle, one pyridine nitrogen, one sulfur, and two oxygen atoms remain uncoordinated. The Cd2+ coordination can be described as a distorted octahedral geometry, where S2, S2A, and two halide anions form the square plane while N2 and N2A atoms occupy the axial positions [N2−Cd1−N2A 180.00(1)°]. The preferred anioncoordinated sandwich structure of 3 and 4 is mainly associated with the strong anion coordination toward the metal center which induces the unsymmetrical exocoordination mode. When one equivalent of Hg(ClO4)2 was reacted with L1 in dichloromethane/acetonitrile, a colorless product 5 was obtained. X-ray analysis reveals that 5 is a 1:2 (metal-to-ligand) sandwich-type complex of the formula [Hg(L1)2](ClO4)2 (Figure 5). The mercury(II) center is six-coordinate, being bound to one secondary nitrogen and two sulfur donors from

Figure 4. Isostructural bis(macrocycle) cadmium(II) halide complexes [Cd(L1)2X2] (3: X = Br and 4: X = I).

Figure 5. Sandwich-type bis(macrocycle) mercury(II) complex 5, [Hg(L1)2](ClO4)2 (5). Noncoordinated anions are omitted.

D

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 6. Heteronuclear complex {[Cd(L1)]2(μ-Hg2I6)}[Hg2I6] (6) showing a dumbbell-type structure: (a) general view, (b) coordination environment of Cd1 showing a distorted monocapped trigonal prism geometry, and (c) side view. The [Hg2I6]2− part is not shown.

L1 with the symmetric convergent exocoordination mode (Type II-b in Chart 2), forming an approximate octahedral geometry. The ether oxygen and pyridine nitrogen donors remain uncoordinated due to the thiaphilicity of the mercury(II). The bite angles around the Hg atom vary considerably, ranging from [75.99(3)° (S1−Hg1−S2A) to 104.01(3)° (S1− Hg1−S2)] due to the larger intraligand S···S repulsion than that of the interligand. The Hg−S bond lengths [Hg1−S1 2.871(1) Å and Hg1−S2 2.885(1) Å] are slightly longer than reported ones for such bonds (2.6−2.8 Å)6b,19 due to the steric hindrance. The Hg−N bond length [2.243(3) Å] is normal for this bond type.20 Consequently, the metal center is shielded by two L1 ligands and therefore is unable to contact the anion or solvent which have weaker coordinating ability toward the metal center. The preferred symmetrical convergent exomode in 5 is mainly due to the synergic coordination of S2N donorset toward mercury(II) in the flexible segment. Previously we reported an endocyclic 1:1 mercury(II) perchlorate complex [Hg(L1)](ClO4)2, which was recrystallized from ether vapor diffusion into the acetonitrile solution of the solid product isolated from the 1:1 reaction in dichloromethane/methanol.8a While the crystalline complex 5 was directly obtained from the 1:1 reaction in dichloromethane/ acetonitrile. This result is hard to explain, however, mercury(II) is a soft acid which often produces its coordination geometries dictated by the conformational preference of the ligands or solvent coordinating ability than the geometric demands of the metal center. Preparation and Structure of Dumbbell-type Cadmium(II)/Mercury(II) Complex of L1 (6). Having obtained the sandwich-type cadmium(II) and mercury(II) complexes of L1, we proceeded to the preparation of related heterometallic complexes. When a mixture of HgI2 and CdI2 in acetonitrile were used in the one-pot reaction with L1 in dichloromethane, a yellow crystalline product 6 was obtained. X-ray analysis revealed that 6 crystallized in the monoclinic

space group C2/c, with Z = 4, with the formula {[Cd(L1)]2(μHg2I6)}[Hg2I6] (Figure 6). The complex 6 consists of one dumbbell-shaped complex cation of type {[Cd(L1)]2(μHg2I6)}2+ and one [Hg2I6]2− cluster anion (Figure S3). In the complex cation unit, the cadmium(II) center is sevencoordinate, being bound to all six donors from one L1. Two macrocyclic cadmium(II) complexes are bridged by (μHg2I6)2− cluster via the Cd−I bond [Cd1−I1 2.908(1) Å] to complete the coordination environment. The coordination geometry of cadmium(II) atom can be described as a distorted monocapped trigonal prism (Figure 6b) with the five donors of L1 defining the pentagonal plane and the axial positions occupied by the nitrogen atom and iodide ion. It is noteworthy to compare the coordination modes of L1 toward CdI2 in the absence (4) and presence (6) of HgI2. In 4, L1 is a bidentate ligand and the two iodide ions participate to the coordination sphere, remaining NO2S-donors of L1 uncoordinated. In 6, however, the formation of two different types of [Hg2I6]2− clusters enhance the complexation ability of L1 toward cadmium(II) to form an endomode complex units. Because the formation the heteronuclear dumbbell 6 is associated with the higher coordination affinity of cadmium(II) toward the macrocycle and the formation of [Hg2I6]2− clusters, this is a good case of competition and collaboration between cadmium(II) and mercury(II) ions. A large numbers of the dumbbell-type (ML-linker-ML) complexes have been reported. However, those with macrocycles are not so common.21,22 Grant group have reported a diruthenium(II) complex of 12-tetrathia-crown-4 (12S4), {[Ru(12S4)]2(μ-bpy)}, linked with 4,4-bipyridyl (bpy).21 We have reported several dumbbell-type disilver(I) complexes of oxathiamacrocycles linked with different diamine spacers such as dabco (1,4-diazabicyclo[2,2,2]octane, bpy, bpp (1,4-bis(4pyridyl)piperazine).22 The preparations of the dumbbell-type macrocyclic complexes with inorganic linkers are serendipitous because they are associated with the cluster formations which E

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 7. (A) 1H NMR spectra of the aliphatic region for L1 in CD3OD in the presence of (b) cadmium(II), (c) mercury(II), and (d) cadmium(II) and mercury(II) (as iodides), [L1] = 5 mM. (B) Comparison of the corresponding cation-induced chemical shifts.

are hardly predictable.23 We recently have reported the first heteronuclear dumbbell complex [(CdL)2(μ-Hg2Br6)](Hg2Br6) (L = 18-membered N2O2S2 macrocycle).23 Thus, 6 isolated in the present work appears to be the second example of the heteronuclear dumbbell-shaped complex of macrocycles. Comparative NMR Study of Competitive Reactions of Cadmium(II) and Mercury(II) toward L1. For further understanding of the above dumbbell-shaped complex that forms via the competition and collaboration between cadmium(II) and mercury(II), the corresponding complexation studies in solution were performed (Figure 7). The comparative 1H NMR experiments to follow the interaction of cadmium(II) iodide (one equivalent, Figure 7b) and mercury(II) iodide (one equivalent, Figure 7c) with L1 were performed. First, the magnitudes of the chemical shift change for aliphatic protons (H1−H4) of L1 were measured after addition of one equivalent of the metal salt (Figure 7B). Except the H1 peak, the magnitudes of the cadmium(II)-induced Δδ for the corresponding proton peaks (0.0013−0.1604 ppm) are larger than those for mercury(II) (0.001−0.1161 ppm; Figure 7B). This result could be evident that (i) both single cations bind to L1 in the exocyclic mode and (ii) cadmium(II) interacts more strongly with L1 than mercury(II) does. When a mixture of one equivalent of CdI2 and one equivalent of HgI2 was added (Figure 7A-d), all the aliphatic protons including H1 were shifted to downfield indicating the endocoordination of cadmium(II) rather than mercury(II). The NMR data for the aromatic region also support the stronger binding affinity of cadmium(II) over mercury(II) (Figure S4). In this condition, probably the mercury(II) weakly interacts with L1 in the exocyclic mode (Figure S4C). Considering no anion contribution on the complexation, the observed NMR result agrees with the solid state data, because the higher endocyclic affinity of cadmium(II) over mercury(II) toward L is, as mentioned, a primary issue in the formation of the dumbbellshaped complex.

Cascade Complexations of L2. We here report the formation of the dimercury(II) complex of L2 which uptakes a small organic guest. In this two stage (or cascade) reactions, we also found that the binding of the organic guest is anionregulated in solution. In order to reveal this phenomena in solution, our attempts for the isolation and structural characterization of the single crystals two macrocyclic dimercury(II) complexes, 7 (I−-form) and 8 (ClO4−-form), together with the perchlorate-dabco (1,4-diazabicyclo[2,2,2]octane) cascade complex 9 were successful. These results allow us to explain that the observed anion-regulated cascade complexation is regulated by the coordination ability of the anions which compete with the organic guest toward the metal centers, as depicted in Chart 3. To the best of our knowledge, the formation of the cascade complex system whose reactivity for the organic guest is regulated by anions has not been reported. Solution Study for the Cascade Complexation of L2. In comparative NMR experiments for the cascade complexations, two mercury(II) salts (I, ClO 4 ) with different anion coordination abilities were employed. Since anions are the potential factors for the second stage binding, our attention focused on the reaction conditions for employing the anions (Figure 8). The addition of two equivalents of HgI2 (Figure 8Ab) or Hg(ClO4)2 (Figure 8B-b) to L2 caused complexationbased downfield shifts in accordance with fast exchange occurring. However, some line broadening and splitting were observed for the perchlorate system, perhaps due to the exchange caused by slow conformational “flipping” of parts of the molecule.24 For both systems, the magnitudes of the chemical shift changes fall in the order of H1, H2 ≫ H3 > H4, in keeping with the NS2 donor sets acting as stronger binding sites. When 1 equiv of dabco was added to the L2-HgI2 solution, the downfield shift of the dabco signal was small (Δδ = 0.14 ppm, Figure 8A-c). While the dabco signal added to the L2F

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

solution also shows a preferential binding for dabco in the perchlorate system (Figures S7 and S8). Consequently, it is obvious that the inclusion of dabco in the second stage is highly regulated by the anions in solution. Crystallographic Approach for the Anion-Regulated Cascade Complexation (7−9). Since these results raise the question of whether the binding of dabco in solution is in accordance with the anion coordination, we assumed that it might be associated with the structures of the respective complexes when the events occur. To investigate this hypothesis, we carried out the preparation of the corresponding species in the crystalline form to reveal their crystal structures. The reaction of L2 with HgI2 in acetonitrile/chloroform yielded a colorless precipitate and vapor diffusion of diethyl ether into a DMSO solution gave rise to a crystalline product of type [Hg2(L2)I4]·2DMSO (7). Employing a similar procedure, but with Hg(ClO4)2 substituted for HgI2, yielded {[Hg2(L2)(ClO4)2(CH3CN)](ClO4)2·2H2O}n (8) as colorless crystals. The iodo complex (7; Figures 9 and S9) and the perchlorato

Chart 3. Schematic Illustrations of the Cascade Complexation of L2 Showing Cation-Supported and AnionRegulated Inclusion of an Organic Guest

Figure 9. Crystal structure of mercury(II) iodide complex [Hg2(L2)I4]·2DMSO (7), showing the blocking of the central space between two Hg atoms by two iodo ligands (I1 and I1A): (top) front view and (bottom) side view. Noncoordinated solvent molecules are omitted.

complex (8; Figures 10a and S10) crystallize in space groups P21/c and P21/m, respectively. The respective crystal structures of 7 and 8 each confirm that the dimetalation reaction with mercury(II) alters the conformation and rigidity of free L2 (Figure S11).8a In both solid state structures, the coordinated anions to some extent block the respective cavities between the two metals, and this might influence the subsequent binding with an organic guest. The coordination environments of the metal centers in 7 and 8 are different. The iodo complex 7 is quite planar (Figure 9), and the two Hg(II) atoms are crystallographically equivalent. The Hg(II) center is five-coordinates being bound to NS2 donors from L2 and two iodo ions in a distorted square pyramidal geometry (τ = 0.14).25 While, the perchlorato complex 8 has a 2-fold axis and two crystallographically independent Hg(II) atoms (Hg1 and Hg2; Figure 10a). The Hg1 atom is five-coordinate, being bound to a NS2 donor set from the macrocycle, one monodentate (bridging) perchlorate ion, and one acetonitrile molecule to give a distorted square

Figure 8. 1H NMR spectra of A: (a) L2, (b) L2 + HgI2 (2 equiv), (c) L2 + HgI2 (2 equiv) + dabco (1 equiv), (d) dabco; and B: (a) L2, (b) L2 + Hg(ClO4)2 (2 equiv), (c) L2 + Hg(ClO4)2 (2 equiv) + dabco (1 equiv), (d) dabco in DMSO-d6.

Hg(ClO4)2 solution showed a much greater downfield shift (Δδ = 0.45 ppm, Figure 8B-c). These results are in keeping with the dabco being bound inside the cavity of the mercury(II) perchlorate complex more strongly to form a stable cascade complex than occurs for the mercury(II) iodide complex. The NMR spectra for the corresponding solid products show identical results (Figures S5 and S6). Also the addition of pyrazine or pyridine to the HgX2-L2-dabco (X = I or ClO4) G

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 10. Crystal structures of (a) mercury(II) perchlorate complex [Hg2(L2)(ClO4)2(CH3CN)](ClO4)2·2H2O (8), showing the removal of one coordinated acetonitrile molecule and the dislocation of two perchlorato ligands from inside to outside (see the green arrows in the side view) on dabco inclusion and (b) a cascade complex [Hg2(L2)(μ-dabco)(ClO4)2](ClO4)2·2DMF·2ether (9): (top) front view and (bottom) side view. Noncoordinated anions and solvent molecules are omitted.



pyramidal coordination geometry (τ = 0.25).25 The Hg2 atom is six-coordinate, being bound to the remaining NS2 donor set from the macrocycle, two oxygens from a terminal perchlorate ion and a further oxygen (O4B) from the bridging perchlorate ion to form an infinite one-dimensional structure (Figure S10). The two mercury(II) centers in 7 are separated by 9.70 Å apart which is shorter than occurs in 8 (10.65 Å), mainly reflecting to the deviation of the Hg1 atom from the macrocyclic plane in 8. In order to investigate the possible formation of cascade complexes, 7 and 8 were employed for reaction with dabco. However, the iodo complex 7 afforded no meaningful solid state products. This is in accord with the iodo ligands being strongly coordinated to the metal centers thus inhibiting further reaction with the guest (Figure 9), as depicted by the locked door system illustrated in Chart 3. Meanwhile, the perchlorato complex 8 gave a specific product when dabco was used. In this case, the cascade complex 9, [Hg2(L2)(μ-dabco)(ClO4)2](ClO4)2·2DMF·2ether (Figure 10b and Figure S12) was isolated. Upon inclusion of dabco in 9, the coordination environments of the two mercury(II) centers (Hg1 and Hg2) are changed due to the loss of the coordinated acetonitrile (in Hg1) and the Hg2−O4B bond breaking (Figure S12). Thus, the 1-D chain of 8 is considered to be converted to the discrete type of 9 with a higher symmetry. In 9, the Hg−N bond distance [Hg1−N3 2.329(7) Å] is typical and the Hg···Hg separation is 7.21 Å, which is considerably shorter than that observed for 8 (10.65 Å). More interestingly, the binding of dabco inside 8 opens the available cavity between the two mercury(II) atoms to its maximum size via a dislocation of the coordinated perchlorato ligands as well as the flipping of the parts of ligating sites (see the green arrows in Figure 10a). Such cation-mediated and anion-regulated cascade complexation can be expected to provide new applications involving the development of adaptive guest binding system. The related studies for the extra-large macrocycles are in progress.

CONCLUSION We have isolated nine complexes of a 20-membered N2O2S2macrocycle and a 40-membered N4O4S4-macrocycle with unusual topologies which cover from monomer and sandwich to dumbbell, 1-D polymer and cascade types. Although unexplored aspects remain in the understanding of exocoordination-based synthetic approaches, the primary object of the present study is to synthesize a series of complexes where the proposed thiaoxaaza-macrocycle is coordinated in different modes and where the coordination modes can be controlled chemically. As discussed, these results highlight how the coordination modes can propagate to the connectivity patterns and influence their overall structures. Especially, the extra-large macrocycle and its dimercury(II) complexes act as candidates for the formation of the cascade complex exhibiting inclusion of an organic guest. 1H NMR investigations starting from either the iodo or perchlorato precursor as a metalloligand showed that the latter system dominates in terms of cascade complexation with the organic guest. Crystallographic analyses established the precise structures of the both dimercury(II) complexes under pinning the anion-regulated binding of the organic guest. This might be an example of the cascade complexations whose formation is not only cation-mediated but also anion-regulated.



EXPERIMENTAL SECTION

General. All chemicals and solvents employed in the syntheses were of reagent grade and were used without further purification. NMR spectra were recorded on a Bruker 500 spectrometer (500 MHz). The FT-IR spectra were measured with a Nicolet iS10 spectrometer. The ESI-mass spectra were obtained on a Thermo Scientific LCQ Fleet spectrometer (solvent: methanol). MALDI-TOF mass spectrum was obtained in a dithranol matrix on a Voyager Applied Biosystems. Each product obtained in this work was dried in a vacuum before elemental analysis, which was carried out on a Thermo Scientific Flash 2000 Series elemental analyzer. The powder X-ray diffraction (PXRD) experiments were performed in a transmission mode with a Bruker GADDS diffractometer equipped with graphite monochromated CuKα radiation (λ = 1.54073 Å). H

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

C, 32.89; H, 3.23; N, 2.87; S, 9.98. Mass spectrum m/z (ESI): 1232.58 for [Hg(L2)I]+ (calcd 1233.19). Preparation of {[Hg2(L2)(ClO4)3(CH3CN)]·ClO4·2H2O}n (8). In a glass tube, Hg(ClO4)2·xH2O (10.2 mg, 0.025 mmol) in acetonitrile (2 mL) was layered onto a solution of L2 (10.1 mg, 0.011 mmol) in chloroform (2 mL). The X-ray quality single crystals 8 were obtained. Yield (20%); Mp: 170−171 °C (decomp.). IR (KBr pellet): 3529, 2925, 2855, 1637, 193, 1457, 1420, 1381, 1298, 1250, 1191, 1094, 789, 767 cm−1. Anal. Calcd for C52H59Hg2Cl4N5S4O20 (%): C, 35.79; H, 3.41; N, 4.01; S, 7.35. Found: C, 35.80; H, 3.45; N, 4.29; S, 7.38. Mass spectrum m/z (ESI): 1204.58 for [Hg(L2)ClO4]+ (calcd 1205.24). Preparation of [Hg2(L2)(μ-dabco)(ClO4)2](ClO4)2·2DMF·2ether (9). A dabco (1.4 mg, 0.013 mmol) was added to a solution of 8 (20.1 mg, 0.012 mmol) in DMF (1 mL). Vapor diffusion of diethyl ether into a DMF solution afforded a colorless crystalline product 9, suitable for X-ray analysis. Yield (15%); Mp: 173−175 °C. IR (KBr pellet): 3028, 2957, 2878, 2018, 1662, 1463, 1388, 1146, 1109, 1088 (ClO4−), 626 cm−1. Anal. Calcd for C62H82Cl4Hg2N8S4O22 (%): C, 37.94; H, 4.21; N, 5.71; S, 6.54. Found: C, 37.89; H, 4.03; N, 5.58; S, 6.48. Mass spectrum m/z (MALDI-TOF, Figure S13): 1717.57 for [Hg2(L2)(dabco)(ClO4)3]+ (calcd 1717.21) .

CAUTION! Mercury and cadmium complexes are toxic. Perchlorate salts of metal complexes are potentially explosive and should be handled with great care. Preparation of {[(μ4-Cu4I4)(HL1)2](CF3COO)2}n (1). CuI (25.7 mg, 0.135 mmol) in acetonitrile (2 mL) was layered onto a solution of L1 (20.0 mg, 0.044 mmol) in dichloromethane (2 mL), and 3 drops of trifluoroacetic acid were added. 1 was obtained as an X-ray quality pale-yellow crystalline product. Yield (40%); Mp: 209−211 °C (decomp.). IR (KBr pellet): 3064, 3033, 2933, 1679, 1599, 1493, 1455, 1241, 1200, 1137, 1103, 1046, 1015, 782, 753, 721 cm−1. Anal. Calcd for C54H58Cu4I4N4O8S4F6 (%): C, 34.22; H, 3.08; N, 2.96; S, 6.77. Found: C, 33.97; H, 3.03; N, 3.23; S, 7.04. Preparation of [CuII(L1)](ClO4)2·H2O (2). Cu(ClO4)2·6H2O (18.1 mg, 0.049 mmol) in methanol (2 mL) was added to a solution of L1 (20.1 mg, 0.044 mmol) in dichloromethane (2 mL). Slow evaporation of the solution afforded a green crystalline product 2 suitable for X-ray analysis. Yield (35%); Mp: 220−222 °C (decomp.). IR (KBr pellet): 3451, 3197, 3071, 2977, 2936, 1603, 1492, 1454, 1381, 1246, 1144, 1088, 1007, 794, 764, 625 cm−1. Anal. Calcd for C25H28Cu1Cl2N2O10S2 (%): C, 41.99; H, 3.95; N, 3.92; S, 8.97. Found: C, 42.17; H, 4.04; N, 3.93; S, 9.20. Mass spectrum m/z (ESI): 614.32 for [Cu(L1)ClO4]+ (calcd 614.04). Preparation of [Cd(L1)2Br2] (3). CdBr2·4H2O (17.1 mg, 0.050 mmol) in acetonitrile (2 mL) was layered onto a solution of L1 (20.4 mg, 0.045 mmol) in dichloromethane (2 mL), and 3 was obtained as an X-ray quality colorless crystalline product. Yield (34%); Mp: 177− 179 °C. IR (KBr pellet): 3214, 3064, 2926, 2862, 1600, 1578, 1491, 1452, 1291, 1240, 1222, 1103, 1047, 1014, 757 cm−1. Anal. Calcd for C50H56N4S4O4Cd1Br2 (%): C, 51.00; H, 4.79; N, 4.76; S, 10.89. Found: C, 50.60; H, 4.40; N, 4.88; S, 10.51. Mass spectrum m/z (ESI): 644.92 for [Cd(L1)Br]+ (calcd 644.98). Preparation of [Cd(L1)2I2] (4). CdI2 (17.9 mg, 0.049 mmol) in acetonitrile (2 mL) was layered onto a solution of L1 (20.0 mg, 0.044 mmol) in dichloromethane (2 mL), and 4 was obtained as an X-ray quality colorless crystalline product. Yield (38%); Mp: 182−183 °C. IR (KBr pellet): 3199, 2967, 2923, 2856, 2375, 1601, 1577, 1490, 1457, 1451, 1240, 1221, 1103, 1047, 756 cm−1. Anal. Calcd for C50H56N4S4O4Cd1I2 (%): C, 47.23; H, 4.44; N, 4.41; S, 10.09. Found: C, 47.01; H, 4.55; N, 4.89; S, 10.45. Mass spectrum m/z (ESI): 692.83 for [Cd(L1)I]+ (calcd 692.97). Preparation of [Hg(L1)2](ClO4)2 (5). Hg(ClO4)2·xH2O (17.9 mg, 0.045 mmol) in acetonitrile (2 mL) was added to a solution of L1 (20.1 mg, 0.044 mmol) in dichloromethane (2 mL). Slow evaporation of the solution afforded a colorless crystalline product 5 that proved suitable for X-ray analysis. Yield (30%); Mp: 179−180 °C (decomp.). IR (KBr pellet): 3226, 2924, 2868, 1598, 1493, 1454, 1242, 1100, 1019, 762, 623 cm−1. Anal. Calcd for C50H56HgN4S4O12Cl2 (%): C, 46.03; H, 4.33; N, 4.29; S, 9.83. Found: C, 46.55; H, 4.05; N, 4.04; S, 9.53. Mass spectrum m/z (ESI): 752.75 for [Hg(L1)ClO4]+ (calcd 753.08). Preparation of {[Cd(L1)]2(μ-Hg2I6)}[Hg2I6] (6). A mixture of CdI2 (18.0 mg, 0.049 mmol) and HgI2 (22.4 mg, 0.049 mmol) in acetonitrile (2 mL) were layered onto a solution of L1 (20.2 mg, 0.045 mmol) dichloromethane (2 mL). 6 was obtained as an X-ray quality yellow crystalline product. Yield (40%); Mp: 210−211 °C. IR (KBr pellet): 3226, 2921, 1604, 1487, 1454, 1441, 1231, 1221, 1178, 1167, 11 0 4, 1 0 47 , 10 3 2, 7 86 , 7 58 c m − 1 . A na l. C a l c d f o r C50H56Cd2Hg4I12N4S4O4 (%): C, 17.38; H, 1.63; N, 1.62; S, 3.71. Found: C, 17.80; H, 1.61; N, 1.70; S, 3.81. Mass spectrum m/z (ESI): 692.83 for [Cd(L1)I]+ (calcd 692.97) and 780.75 for [Hg(L1)I]+ (calcd 781.03). Preparation of [Hg2(L2)I4]·2DMSO (7). HgI2 (11.2 mg, 0.025 mmol) in acetonitrile (2 mL) was added to a solution of L2 (10.1 mg, 0.011 mmol) in dichloromethane (2 mL). The colorless precipitate obtained was separated and dissolved in DMSO. Vapor diffusion of diethyl ether into the DMSO solution afforded a colorless crystalline product 7, suitable for X-ray analysis. Yield (13%); Mp: 148−150 °C (decomp.). IR (KBr pellet): 3196, 3059, 2918, 2850, 1685, 1662, 1593, 1492, 1447, 1295, 1246, 1111, 1061, 1046, 752 cm−1. Anal. Calcd for C54H68Hg2I4N4S6O6 (%): C, 32.92; H, 3.48; N, 2.84; S, 9.76. Found:



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b00154. Crystal structures, PXRD patterns, comparative NMR spectra, mass spectrum, and X-ray crystallographic files (CIF format) for 1−9 (PDF). Accession Codes

CCDC 1571710−1571712 and 1813888−1813893 contain the supplementary crystallographic data for this paper. CCDC 1813888 (1), 1813889 (2), 1813890 (3), 1813891 (4), 1813892 (5), 1813893 (6), 1571710 (7), 1571711 (8), and 1571712 (9) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_ [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

In-Hyeok Park: 0000-0003-1371-6641 Jong Hwa Jung: 0000-0002-8936-2272 Shim Sung Lee: 0000-0002-4638-5466 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by NRF (2016R1A2A2A05918799 and 2017R1A4A1014595), South Korea, and the Support Program for Strategic Research Foundation at Private Universities (S1201034) and KAKENHI-PROJECT17K05844 from the Ministry of Education, Sports, Science and Technology of Japan are acknowledged. I

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



[PtIV(L1)Br3]Br0.5·2H2O, [AuIII(L1)Cl2][AuCl4] and [Ag4(L1)4](CF3SO3)4. Polyhedron 1991, 10, 19−25. (5) Kim, H. J.; Park, I.-H.; Lee, J.-E.; Park, K.-M.; Lee, S. S. Supramolecular Silver(I), Copper(I), and Mercury(II) Complexes with Thiamacrocycles Exhibiting Different Types of Endo- or Exocoordination Modes: From Monomer and Dimer to OneDimensional and Two-Dimensional Polymers. Cryst. Growth Des. 2014, 14, 6269−6281. (6) (a) Kim, S.; Lee, E.; Park, K.-M.; Lee, S. S. Exocyclic coordination chemistry of an O2S2-macrocycle with copper(I), mercury(II) and palladium(II) ions. CrystEngComm 2013, 15, 8544−8551. (b) Lee, S. Y.; Park, S.; Kim, H. J.; Jung, J. H.; Lee, S. S. Ligand- and AnionDirected Assembly of Exo-Coordinated Mercury(II) Halide Complexes with O2S2-Donor Macrocycles. Inorg. Chem. 2008, 47, 1913− 1915. (7) (a) Habata, Y.; Seo, J.; Otawa, S.; Osaka, F.; Noto, K.; Lee, S. S. Synthesis of Diazahexathia-24-crown-8 Derivatives and Structures of Ag+ Complexes. Dalton Trans. 2006, 2202−2206. (b) Kang, Y.; Park, I.-H.; Ikeda, M.; Habata, Y.; Lee, S. S. A Double Decker Type Complex: Copper(I) Iodide Complexation with Mixed Donor Macrocycles via [1:1] and [2:2] Cyclizations. Dalton Trans. 2016, 45, 4528−4533. (8) (a) Jung, D.; Chamura, R.; Habata, Y.; Lee, S. S. Extra Large Macrocycle: 40-Membered Macrocycle via 2:2 Cyclization and Its Dimercury(II) Complex. Inorg. Chem. 2011, 50, 8392−8396. (b) Lee, S.-G.; Park, K.-M.; Habata, Y.; Lee, S. S. Endo- and Exocyclic Supramolecular Complexes of Mixed-Donor Macrocycles via [1:1] and [2:2] Cyclizations. Inorg. Chem. 2013, 52, 8416−8426. (9) (a) Brooker, S. Complexes of thiophenolate-containing Schiffbase macrocycles and their amine analogues. Coord. Chem. Rev. 2001, 222, 33−56. (b) Cameron, S. A.; Brooker, S. Metal-Free and Dicopper(II) Complexes of Schiff Base [2 + 2] Macrocycles Derived from 2,2′-Iminobisbenzaldehyde: Syntheses, Structures, and Electrochemistry. Inorg. Chem. 2011, 50, 3697−3706. (c) Tandon, S. S.; Thompson, L. K.; Bridson, J. N. Dinuclear Copper (II) and MixedValence Copper(II)-Copper(I) Complexes of 34-Membered Macrocyclic Ligands (H2M1, H2M2) Capable of Forming Endogenous Phenolate and Pyridazino Bridges. X-ray Crystal Structures of the Dinuclear Copper(I1) Complexes [Cu 2 M1][BF 4 ] 2 ·H 2 O and [Cu2M2][BF4]2·CH3OH, Which Exhibit a Remarkable Ligand Twist. Inorg. Chem. 1993, 32, 32−39. (10) (a) Koshland, D. E., Jr Application of a Theory of Enzyme Specificity to Protein Synthesis. Proc. Natl. Acad. Sci. U. S. A. 1958, 44, 98−104. (b) Koshland, D. E., Jr The Key-Lock Theory and the Induced Fit Theory. Angew. Chem., Int. Ed. Engl. 1995, 33, 2375−2378. (c) Boehr, D. D.; Nussinov, R.; Wright, P. E. The Role of Dynamic Conformational Ensembles in Biomolecular Recognition. Nat. Chem. Biol. 2009, 5, 789−796. (d) Williamson, J. R. Induced fit in RNA− Protein Recognition. Nat. Struct. Biol. 2000, 7, 834−837. (11) Lehn, J.-M.; Pine, S. H.; Watanabe, E.; Willard, A. K. Binuclear Cryptates. Synthesis and Binuclear Cation Inclusion Complexes of Bistren Macrobicyclic Ligands. J. Am. Chem. Soc. 1977, 99, 6766−6768. (12) (a) Alibrandi, G.; Amendola, V.; Bergamaschi, G.; Fabbrizzi, L.; Licchelli, M. Bistren Cryptands and Cryptates: Versatile Receptors for Anion Inclusion and Recognition in Water. Org. Biomol. Chem. 2015, 13, 3510−3524. (b) Zhong, D.-C.; Lu, T.-B. Molecular Recognition and Activation by Polyaza Macrocyclic Compounds Based on Host− Guest Interactions. Chem. Commun. 2016, 52, 10322−10337. (c) Esteves, C. V.; Mateus, P.; André, V.; Bandeira, N. A. G.; Calhorda, M. J.; Ferreira, L. P.; Delgado, R. Di- versus Trinuclear Copper(II) Cryptate for the Uptake of Dicarboxylate Anions. Inorg. Chem. 2016, 55, 7051−7060. (d) Kang, S. O.; Llinares, J. M.; Dayc, V. W.; Bowman-James, K. Cryptand-Like Anion Receptors. Chem. Soc. Rev. 2010, 39, 3980−4003. (e) Bond, A. D.; Derossi, S.; Harding, C. J.; McInnes, E. J. L.; McKee, V.; McKenzie, C. J.; Nelson, J.; Wolowska, J. Cascade complexation: a single cyano bridge links a pair of Cu(II) cations. Dalton Trans. 2005, 2403−2409. (f) Ravikumar, I.; Suresh, E.; Ghosh, P. A Perfect Linear Cu−NNN−Cu Unit Inside the Cryptand Cavity and Perchlorate Entrapment within the Channel Formed by the

REFERENCES

(1) (a) Pedersen, C. J. Cyclic Polyethers and their Complexes with Metal Salts. J. Am. Chem. Soc. 1967, 89, 7017−7036. (b) Lindoy, L. F. The Chemistry of Macrocyclic Complexes; Cambridge University Press: Cambridge, U.K., 1989. (c) Lehn, J.-M. Supramolecular Chemistry, Concept and Perspectives; VCH: Weinheim, Germany, 1995. (d) Izatt, R. M.; Pawlak, K.; Bradshaw, J. S. Thermodynamic and Kinetic Data for Macrocycle Interaction with Cations, Anions and Neutral Molecules. Chem. Rev. 1995, 95, 2529−2586. (2) (a) Wolf, R. E., Jr.; Hartman, J. R.; Storey, J. M. E.; Foxman, B. M.; Cooper, S. R. Crown thioether chemistry: structural and conformational studies of tetrathia-12-crown-4, pentathia-15-crown-5, and hexathia-18-crown-6. Implications for ligand design. J. Am. Chem. Soc. 1987, 109, 4328−4335. (b) Blake, A. G.; Schröder, M. Chemistry of Thioether Macrocyclic Complexes. Adv. Inorg. Chem. 1990, 35, 1− 80. (c) Hill, S. E.; Feller, D. Theoretical Conformational Analysis of Thiacrown Macrocycles. J. Phys. Chem. A 2000, 104, 652−660. (3) (a) Park, S.; Lee, S. Y.; Park, K.-M.; Lee, S. S. Supramolecular Networking of Macrocycles Based on Exo-Coordination: From Discrete to Continuous Frameworks. Acc. Chem. Res. 2012, 45, 391−403. (b) Seo, J.; Song, M. R.; Lee, J.-E.; Lee, S. Y.; Yoon, I.; Park, K.-M.; Kim, J.; Jung, J. H.; Park, S. B.; Lee, S. S. Exo-CoordinationBased Supramolecular Silver(I) Complexes of S2O Macrocycles: Effect of Ligand Isomerism on the Structural Diversity. Inorg. Chem. 2006, 45, 952−954. (c) Seo, J.; Song, M. R.; Lee, J.-E.; Lee, S. Y.; Yoon, I.; Park, K.-M.; Kim, J.; Jung, J. H.; Lee, S. Y.; Park, S.; Lee, S. S. Copper(I), Silver(I), and Palladium(II) Complexes of a Thiaoxamacrocycle Displaying Unusual Topologies. Inorg. Chem. 2009, 48, 11335−11341. (d) Lee, H.-H.; Park, I.-H.; Lee, S. S. Cooperative Effect of Anion and Mole Ratio on the Coordination Modes of an NO2S3-Donor Macrocycle. Inorg. Chem. 2014, 53, 4763−4769. (e) Lee, H.-H.; Lee, E.; Ju, H.; Kim, S.; Park, I.-H.; Lee, S. S. Influence of Anion and Mole Ratio on the Coordination Behavior of an NO2S3-Macrocycle: The Formation of a Dumbbell-Shaped Macrocyclic Cadmium(II) Iodide Complex. Inorg. Chem. 2016, 55, 2634−2640. (f) Park, S.; Lindoy, L. F.; Lee, S. S. Assembly of Silver(I) Complexes of Isomeric NS2-Macrocycles Displaying Cyclic Oligomer, Helix, and Zigzag Structures. Cryst. Growth Des. 2012, 12, 1320−1329. (4) (a) Beattie, C.; Farina, P.; Levason, W.; Reid, G. Oxa-thia-, oxaSelena and crown ether macrocyclic complexes of tin(II) tetrafluoroborate and hexafluorophosphate - synthesis, properties and structures. Dalton Trans. 2013, 42, 15183−15190. (b) Sarnicka, A.; Starynowicz, P.; Lisowski, J. Controlling the macrocycle size by the stoichiometry of the applied template ion. Chem. Commun. 2012, 48, 2237−2239. (c) Cao, W.; Wang, H.; Wang, X.; Lee, H. K.; Ng, D. K. P.; Jiang, J. Constructing Sandwich-Type Rare Earth Double-Decker Complexes with N-Confused Porphyrinato and Phthalocyaninato Ligands. Inorg. Chem. 2012, 51, 9265−9272. (d) Berlicka, A.; LatosGra, L. Intramolecular Rotation of Iron(II) Dithiaethyneporphyrin Double-Decker Complex: 1H NMR Studies. Inorg. Chem. 2009, 48, 7922−7930. (e) Kubiak, R.; Janczak, J. Structural and spectroscopic characterisation of the holmium double-decker partially oxidised and bi-radical phthalocyanine complexes with iodine. Inorg. Chim. Acta 2010, 363, 2461−2466. (f) Zhang, X.; Muranaka, A.; Lv, W.; Zhang, Y.; Bian, Y.; Jiang, J.; Kobayashi, N. A Cobalt Metallacrown Anion Host with Guest-Dependent Redox Activity. Chem. - Eur. J. 2008, 14, 4667−4675. (g) Golecki, M.; Kersting, B. Preparation, Properties, and Structures of Pentanuclear [{Ni2L}2(μ-csalen)/M]2+ Complexes (L = Macrocyclic N6S2 Donor Ligand). Z. Anorg. Allg. Chem. 2015, 641, 436−441. (h) Habata, Y.; Noto, K.; Osaka, F. Substituent Effects on the Structures of Silver Complexes with Monoazatrithia-12-Crown-4 Ethers Bearing Substituted Aromatic Rings. Inorg. Chem. 2007, 46, 6529−6534. (i) Yamaguchi, T.; Yamazaki, F.; Ito, T. A cyclic tetramer of [Ag(1,4,7-trithiacyclononane)]+, cyclo-tetra-kis(μ-1,4,7-trithiacyclononane-κ3S1,S4,S7:κS1)tetrasilver(I) tetrakis(trifluoromethanesulfonate) nitromethane disolvate. Acta Crystallogr. 2002, C58, m213−m214. (j) Heinzel, U.; Mattes, R. Binding of the late transition metal ions to the mixed N2S donor macrocyclic ligand 1-thia-4,7-diazacyclononane (= L1). X-ray crystal structures of J

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

and Endo-/Exocyclic Polymer Structures. Cryst. Growth Des. 2012, 12, 4972−4980. (23) Seo, S.; Ju, H.; Kim, S.; Park, I.-H.; Lee, E.; Lee, S. S. Group 12 Metal Complexes of an 18-Membered N2O2S2 Macrocycle Incorporating Two Pyridines: First Examples of an Infinite Mercury(I) Complex and a Dumbbell-Shaped Heteronuclear Complex with a Macrocyclic Ligand. Inorg. Chem. 2016, 55, 11028−11039. (24) Schoepff, L.; Kocher, L.; Durot, S.; Heitz, V. Chemically Induced Breathing of Flexible Porphyrinic Covalent Cages. J. Org. Chem. 2017, 82, 5845−5851. (25) Addison, A. W.; Rao, T. N.; Reedijk, J.; Van Rijin, J.; Verschoor, G. C. Synthesis, Structure, and Spectroscopic Properties of Copper(II) Compounds Containing Nitrogen−Sulphur Donor Ligands; the Crystal and Molecular Structure of Aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) Perchlorate. J. Chem. Soc., Dalton Trans. 1984, 1349−1356.

Cascade Complex. Inorg. Chem. 2006, 45, 10046−10048. (g) Amendola, V.; Bergamaschi, G.; Boiocchi, M.; Fabbrizzi, L.; Poggi, A.; Zema, M. Halide ion inclusion into a dicopper(II) bistren cryptate containing ‘active’ 2,5-dimethylfuran spacers: The origin of the bright yellow colour. Inorg. Chim. Acta 2008, 361, 4038−4046. (13) (a) Anderson, H. L.; Hunter, C. A.; Meah, M. N.; Sanders, J. K. M. Thermodynamics of Induced-Fit Binding Inside Polymacrocyclic Porphyrin Hosts. J. Am. Chem. Soc. 1990, 112, 5780−5789. (b) Corbett, P. T.; Tong, L. H.; Sanders, J. K. M.; Otto, S. Diastereoselective Amplification of an Induced-Fit Receptor from a Dynamic Combinatorial Library. J. Am. Chem. Soc. 2005, 127, 8902− 8903. (c) Poul, L. P.; Mest, Y. L.; Javin, I.; Reinaud, O. Supramolecular Modeling of Mono-copper Enzyme Active Sites with Calix[6]arenebased Funnel Complexes. Acc. Chem. Res. 2015, 48, 2097−2106. (14) Kang, E.-J.; Lee, S. Y.; Lee, H.; Lee, S. S. Sulfur-Containing Mixed-Donor Tribenzo-Macrocycles and Their Endo- and Exocyclic Supramolecular Silver(I) and Copper(I) Complexes. Inorg. Chem. 2010, 49, 7510−7520. (15) Park, I.-H.; Kim, H. J.; Lee, S. S. Anion-dependent coordinative networking of macrocycle with copper(I) halides. CrystEngComm 2012, 14, 4589−4595. (16) Kim, S.; Siewe, A. D.; Lee, E.; Ju, H.; Park, I.-H.; Park, K.-M.; Ikeda, M.; Habata, Y.; Lee, S. S. Ligand-Induced Formation of Copper(I) Iodide Clusters: Exocyclic Coordination Polymers with Bisdithiamacrocycle Isomers. Inorg. Chem. 2016, 55, 2018−2022. (17) (a) Churchill, M. R.; Rotella, F. J. Molecules with an M4X4 core. 9. Crystal structure and molecular geometry of tetrameric (methyldiphenylphosphine)copper(I) iodide, [(PmePh2)CuI]4. Inorg. Chem. 1977, 16, 3267−3273. (b) Babich, O. A.; Kokozay, V. N. Direct synthesis and crystal structure of tetrameric copper(I) iodide with trimethylamine [CuI(Nme3)]4. Polyhedron 1997, 16, 1487−1490. (c) Ashton, P. R.; Burns, A. L.; Claessens, C. G.; Shimizu, G. K. H.; Small, K.; Stoddart, J. F.; White, A. J. P.; Williams, D. J. Thiamacrocyclic chemistry: synthesis of a novel oxadithiacrown and itscopper iodide complex. J. Chem. Soc., Dalton Trans. 1997, 1493− 1496. (d) Graham, P. M.; Pike, R. D. Coordination Polymers of Copper(I) Halides. Inorg. Chem. 2000, 39, 5121−5132. (e) Vitale, M.; Palke, W. E.; Ford, P. C. Origins of the Double Emission of the Tetranuclear Copper(I) Cluster Cu4I4(pyridine)4: An ab Initio Study. J. Phys. Chem. 1992, 96, 8329−8336. (18) (a) Ford, P. C.; Cariati, E.; Bourassa, J. Photoluminescence Properties of Multinuclear Copper(I) Compounds. Chem. Rev. 1999, 99, 3625−3648. (b) Vitale, M.; Ford, P. C. Luminescent mixed ligand copper(I) clusters (CuI)n(L)m (L = pyridine, piperidine): thermodynamic control of molecular and supramolecular species. Coord. Chem. Rev. 2001, 219−221, 3−16. (c) Kim, T. H.; Shin, Y. W.; Jung, J. H.; Kim, J. S.; Kim, J. Crystal-to-Crystal Transformation between Three CuI Coordination Polymers and Structural Evidence for Luminescence Thermochromism. Angew. Chem., Int. Ed. 2008, 47, 685. (d) Peng, R.; Li, M.; Li, D. Copper(I) halides: A versatile family in coordination chemistry and crystal engineering. Coord. Chem. Rev. 2010, 254, 1−18. (19) Tsuchiya, T.; Shimizu, T.; Hirabayashi, K.; Kamigata, N. Formation and Structures of Mercury Complexes of 18-Membered Unsaturated and Saturated Thiacrown Ethers. J. Org. Chem. 2003, 68, 3480−3485. (20) (a) Byriel, K. A.; Gahan, L. R.; Kennard, C. H. L.; Sunderland, C. J. Synthesis, characterization and crystal structures of 1,7-dioxa-4thia-10-azacyclododecane ([12]aneNO2S) and [Hg([12]aneNO2S)(NO3)2]. J. Chem. Soc., Dalton Trans. 1993, 625−629. (b) Hambley, T. W.; Afshar, S.; Marcus, S. T.; Gahan, L. R. Mercury(II) and Lead(II) Complexes of 1-Oxa-7-thia-4,10-diazacyclododecane ([12]aneN2OS. Aust. J. Chem. 1999, 52, 1−6. (21) Janzen, D. E.; Chen, W.; VanDerveer, D. G.; Mehne, L. F.; Grant, G. J. Synthesis and structural studies of ruthenium(II) 12S4 complexes with 4,4’-bipyridine: The crystal structures of [Ru(12S4)(bpy)Cl](Cl)·H2O and [{Ru(12S4)Cl}2-μ-(bpy)](PF6)2·2CH3CN. Inorg. Chem. Commun. 2006, 9, 992−995. (22) Ju, H.; Lee, S. S. Assembly of Soft Metal Complexes of an O4S2Macrocycle Displaying Endocyclic Monomer, Dumbbell-Like Dimer, K

DOI: 10.1021/acs.inorgchem.8b00154 Inorg. Chem. XXXX, XXX, XXX−XXX