Halides Held by Bifurcated Chalcogen–Hydrogen Bonds. Effect of μ(S

15 hours ago - Synopsis. Unreported bifurcated chalcogen−hydrogen bonding, μ(S,N−H)Cl, provides solid-state and solution dimerization of Cl(carbe...
0 downloads 2 Views 5MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Halides Held by Bifurcated Chalcogen−Hydrogen Bonds. Effect of μ(S,N−H)Cl Contacts on Dimerization of Cl(carbene)PdII Species Alexander S. Mikherdov, Alexander S. Novikov, Mikhail A. Kinzhalov, Vadim P. Boyarskiy,* Galina L. Starova, Alexander Yu. Ivanov, and Vadim Yu. Kukushkin* Saint Petersburg State University, 7/9 Universitetskaya Nab., Saint Petersburg 199034, Russian Federation S Supporting Information *

ABSTRACT: The reaction of cis-[PdCl2(CNCy)2] (1) with thiazol-2amines (2−10) leads to the C,N-chelated diaminocarbene-like complexes [PdCl{C(N(H)4,5-R2-thiazol-2-yl)NHCy}(CNCy)] (11−14; 82−91%) in the case of 4,5-R2-thiazol-2-amines (R, R = H, H (2), Me, Me (3), −(CH2)4− (4)) and benzothiazol-2-amine (5) or gives the diaminocarbene species cis[PdCl2{C(N(H)Cy)N(H)4-R-thiazol-2-yl}(CNCy)] (15−19; 73−93%) for the reaction with 4-aryl-substituted thiazol-2-amines (R = Ph (6), 4-MeC6H4 (7), 4-FC6H4 (8), 4-ClC6H4 (9), 3,4-F2C6H3 (10)). Inspection of the singlecrystal X-ray diffraction data for 15−17 and 19 suggests that the structures of all these species exhibit previously unrecognized bifurcated chalcogen− hydrogen bonding μ(S,N−H)Cl and also PdII···PdII metallophilic interactions. These noncovalent interactions collectively connect two symmetrically located molecules of 15−17 and 19, resulting in their solid-state dimerization. The existence of the μ(S,N−H)Cl system and its strength (6−9 kcal/mol) were additionally verified/estimated by a Hirshfeld surface analysis and DFT calculations combined with a topological analysis of the electron density distribution within the formalism of Bader’s theory (AIM method) and NBO analysis. The observed noncovalent interactions are jointly responsible for the dimerization of 15−19 not only in the solid phase but also in CHCl3 solutions, as predicted theoretically by DFT calculations and confirmed experimentally by FTIR, HRESI-MS, 1H NMR, and diffusion coefficient NMR measurements. Available CCDC data were processed under the new moiety angle, and the observed μ(S,E−H)Cl systems were classified accordingly to E (E = N, O, C) type atoms.



INTRODUCTION The field of noncovalent interactions has grown explosively in the past decade, warranting a recent special issue of Chemical Reviews devoted exclusively to this topic.1 The hydrogen,2 halogen,3 chalcogen,4 pnictogen,5 and tetrel5a bonding, tetrellike σ-hole interactions involving Pb2+,6 stacking,7 cation− and anion−π,8 and metallophilic interactions9 play key roles in many chemical, physical, and biochemical processes due to their ability to control the structures and properties of associates and supramolecular systems from water dimerization10 to protein folding.11 In particular, chalcogen bonding (CB) is based on the interaction between a localized positive region, the so-called σhole, on a chalcogen atom and an electron donor species serving as a CB acceptor. The heavier chalcogens Te and Se12 are more prone to act as CB donors due to their higher polarizability,4e,13 but most recent reports deal with CBs with more abundant S centers. Most common types of CBs with sulfur centers involve S···S, S···O, and S···N interactions,4a,14 whereas S···Hal (Hal = halide) contacts are substantially less common. While intermolecular Te···Hal and Se···Hal CBs are rather strong and these contacts are preserved even in solution,12d for S © XXXX American Chemical Society

centers weak but noticeable S···Hal contacts are feasible only for bifurcated CBs: viz. μ(Ch,Ch)Hal (Ch = S, Figure 1, A). It is noteworthy that bifurcated types of noncovalent contacts play an active role in some organocatalytic transformations (for a recent review see ref 15; halogen bonding (XB) catalysts,16 CB

Figure 1. Bifurcated bonding patterns (Ch = S, Se, Te; E = C, N, O). Received: January 23, 2018

A

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Scheme 1. Two Directions of the Reactions of cis-[PdCl2(CNR1)2] with Thiazol-2-amines

catalysts17) and in anion recognition and transfer (for a recent review see ref 18; XB receptors,19 CB receptors20). Whereas the classic bifurcated hydrogen bond (Figure 1, C) is the subject of a large number of studies,18,21 noncovalent bifurcated chalcogen−hydrogen bonding μ(Ch,E−H)Hal (Figure 1, B) has never been investigated as a separate phenomenon. Although more than 600 examples of structures featuring the μ(Ch,E−H)Hal moieties (B) can be found in the CCDC database, the presence of both types of contacts, i.e. both Ch···Hal (CB) and E−H···Hal (HB) within a single structure, was only briefly mentioned in 20 reports (Table S1 in the Supporting Information). None of these studies, except our previous work on CB,22 discuss the nature of the Ch···Hal contacts or recognize them as CBs. Thus, the μ(Ch,E−H)Hal functionality has never been considered as a specific moiety with heterobifurcated chalcogen−hydrogen μ(Ch,E−H)Hal (B) bonding. In this work, we found that the complexes cis-[PdCl2{C(N(H)Cy)N(H)4-R-thiazol-2-yl}(CNCy)] (15−19; Scheme 1) exhibit dimeric structures held together by the previously overlooked bifurcated chalcogen−hydrogen bonding μ(S,N−H)Cl (B) between two monomeric parts. The existence of these interactions, observed by X-ray diffraction, was additionally confirmed by DFT calculations combined with a topological analysis of the electron density distribution within the formalism of Bader’s theory (AIM method). The observed noncovalent interactions are jointly responsible for the dimerization of 15−19 not only in the solid phase but also in CHCl3 (CDCl3) solutions, as predicted theoretically by DFT calculations and confirmed experimentally by combined FTIR, HRESI-MS, 1H NMR, and diffusion NMR methods. We have additionally processed the available CCDC data using the new moiety angle and verified and classified the μ(S,E−H)X moieties on the basis of the nature of the E atom (E = N, O, C). These results are discussed accordingly in the sections that follow.

in a box in the scheme.24a The formed species were not isolated because of their higher reactivity in comparison to that of thiazol-2-amines, and they are subject to further reaction with the other isocyanide complex, giving a mixture of two regioisomeric binuclear diaminocarbenes. The equilibrium (bottom left, Scheme 1) between these species mainly depends on the energy difference between two different types of CB: viz., S···Cl and S···N. Being interested in an extension of these reactions to other isocyanide species, we employed another isocyanide complex, viz. cis-[PdCl2(CNCy)2] (1), and found that seemingly small modification of the nature of the substituent R at CNR (Xyl was replaced with Cy) led to a substantial change in the reactivity and, what is even more important, to totally different types of noncovalent interactions in the generated coupling products (route B). Reaction between cis-[PdCl2(CNCy)2] and Thiazol-2amines. We found that the reaction between 1 and 4,5-R2thiazol-2-amines (R1, R2 = H, H (2), Me, Me (3), −(CH2)4− (4)) or benzothiazol-2-amine (5) gives mononuclear C,Nchelated complexes 11−14 (82−91%; route C, Scheme 2). In 11−14, the newly formed ligands are of aminocarbene-like Scheme 2. Studied Coupling



RESULTS AND DISCUSSION In the framework of our ongoing projects focused on metalinvolving reactions of isocyanides (for our reviews see ref 23) on the one hand and on noncovalent interactions on the other hand (for recent works see refs 22 and 24), we recently reported on the reaction between cis-[PdCl2(CNXyl)2] (Xyl = 2,6-Me2C6H4) and various thiazol-2-amines (route A, Scheme 1) that leads to C,N-chelated aminocarbene complexes shown B

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

is possibly due to involvement of the N−H protons in hydrogen bonding;28 this was additionally confirmed by theoretical calculations (see below). Addition of 2−10 to a CNCy ligand of 1 is accompanied by disappearance of the isocyanide quaternary carbon signal at δ 116.1 and appearance of a new signal from the carbene NCN fragment at around δ 180 for 11−19. These signals are in the range specific for Pd− Ccarbene (δC 160−224) in acyclic diaminocarbenes.29 The crystal data, data collection parameters, and structure refinement data for 11, 15−17, and 19 are given in Table S1 in the Supporting Information. The plots of the structures are given in Figure 2 (complex 11) and Figure 3 (complexes 15−

type, as formally they are not carbenes because of their negative charge.25 Concurrently, the treatment of 1 with any one of the 4-aryl-substituted thiazol-2-amines (R1 = Ph (6), 4-MeC6H4 (7), 4-FC6H4 (8), 4-ClC6H4 (9), 3,4-F2C6H3 (10), R2 = H; route D, Scheme 2) leads to open-chain aminocarbene species 15−19 featuring neutral carbene ligands (73−93%). In all cases, these reactions are palladium-mediated insofar as uncomplexed CyNC does not react with 2−10. Formation of two types of the products upon coupling with unsubstituted and substituted thiazol-2-amines can be rationalized by the different reactivity patterns of these heterocyclic systems. Thiazol-2-amines (2−4) and benzothiazol-2-amine (5) react with cis-[PdCl2(CNCy)2] (1) as ambident nucleophiles, initially replacing the chloride ligand by the heterocyclic nitrogen (route C, Scheme 2) and then attacking the isocyanide ligand by the amino group. The second molecule of 2−5 acts as a base, and it deprotonates the formed aminocarbene complex, giving 11−14 and corresponding (2−5)·HCl. Introduction of bulky substituents in the fourth position of the thiazole ring (6−10) prevents the intramolecular nucleophilic addition of the amino group and promotes the decoordination of 6−10 from palladium, and therefore these species react with CNCy in 1 only by the amino group (route D, Scheme 2). Yet more evidence of the alternative mechanisms is the quite different rates of these two reactions. While the reaction with 2 occurs in C2H4Cl2 at room temperature for 2 h, the addition of 6 proceeds upon reflux in C2H4Cl2 for 24 h. The higher rate of the former reaction is probably due to the intramolecular character of the addition.26 The difference in the reactivity between the Xyl and Cy palladium(II) complexes could be rationalized by the different acidities of the intermediate cationic C,N-chelated aminocarbene complexes (Schemes 1 and 2). In the case of the Xyl derivative, the formed aminocarbene complex bearing the more acidic N−H protons is subject to easy two-step deprotonation by a thiazol-2-amine followed by coupling with the other XylCN complex, giving the binuclear species (Scheme 1). The less acidic Cy-substituted diaminocarbene species is subject to only one-step deprotonation, giving complexes 11−14, which are not involved in further reactions (Scheme 2). Characterization of 11−19. The pure complexes were obtained as colorless or yellow solids, and they were characterized by microanalyses (C, H, N), high-resolution ESI-MS, FTIR, TG/DTA, and 1D (1H, 13C{1H}, 19F{1H}) and 2D (1H,1H-COSY, 1H,1H-NOESY, 1H,13C-HSQC, and 1H,13CHMBC) NMR spectroscopy and also by X-ray diffraction for five species (11, 15−17, and 19) (Supporting Information). The HRMS-ESI+ spectra of all new complexes display fragmentation patterns corresponding to [M − Cl]+ and/or [M + H]+ with the characteristic isotopic distribution. In addition, the HRMS-ESI+ spectra of 15−19 display fragments from [2M − Cl]+. The 1H NMR spectra exhibit resonances from the two characteristic C−H protons of the cyclohexyl groups as two overlapped signals at δ ca. 4.0−4.2 for 11−14 or two separate signals in the range from δ 4.0 to 4.9 for 15−19. The positions of the NH protons for both types of complexes are also different. In the case of 11−14, only one signal is located in the range δ 5.2−5.7, whereas for 15−19 two NH protons were observed as a singlet at δ ca. 11.4−11.6 and as a doublet at δ ca. 11.7−11.8. The strong downfield shift of these signals is unusual for palladium aminocarbene complexes,27 and this shift

Figure 2. View of 11 with the atomic numbering scheme. Thermal ellipsoids are drawn at the 50% probability level.

17 and 19). The geometry of these complexes is in agreement with the structures proposed from the solution NMR experiments (Supporting Information). In all cases, the Pd centers adopt slightly distorted square-planar geometries with the isocyanide ligand in the cis-position to the diaminocarbene fragment. The bond lengths of the coordinated CN groups are in the range 1.132−1.157 Å, which is the typical interval for CN triple-bond lengths in related isocyanide palladium complexes: e.g., cis-[PdCl2(CNR)2] (R = Cy, 1.128−1.142 Å;30 R = tBu, 1.108−1.149 Å;31 R = Xyl, 1.145−1.156 Å).32 The Pd−Ccarbene distances (11, 2.004(3) Å; 15, 1.988(3) Å; 16, 1.978(2) Å; 17, 1.999(6) Å; 19, 1.992(6) Å) are comparable to those reported for related palladium aminocarbene complexes: e.g., cis[PdCl2(CNCy){C(ProOtBu)NHCy}] (1.983(2) Å),27e cis[PdCl2{C(NHNHS(O)2Ph)N(H)Cy}(CNCy)] (1.9781(16) Å), 27b and cis-[PdCl 2 {C(Ind)N(H)Cy}(CNCy)] (1.976(3) Å).27c The C−N bond lengths in the NCN fragment are nearly equal (11, 1.332(3) and 1.334(4) Å; 15, 1.362(5) and 1.308(5) Å; 16, 1.349(4) and 1.306(4) Å; 17, 1.354(10) and 1.288(9) Å; 19, 1.353(6) and 1.290(5) Å) and are intermediate between the typical double (1.260(9) Å)33 and single (1.452(8) Å)33 bonds, thus reflecting their diaminocarbene nature.27a,b,e,h,34 C

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. Views of dimers of 15 (a), 16 (b), 17 (c), and 19 (d). Dotted lines indicate the bifurcated μ(S,N−H)Cl and Pd···Pd contacts and intramolecular N−H···N HBs. Thermal ellipsoids are shown with 50% probability. The cyclohexyl substituents are omitted for simplicity.

Table 1. Distances (Å) and Angles (deg) for Two Types of Noncovalent Contacts in the Structures of 15−17 and 19 intermolecular bifurcated μ(S,N−H)Cl

intramolecular HB

d(S···Cl)

∠(C−S···Cl)

d(H···Cl)

15 16 17 19

3.4509(11) 3.4985(9) 3.495(2) 3.5428(12)

166.37(11)° 172.35(13)° 171.7(3)° 167.44(13)°

2.27(5) 2.30(5) 2.33(5) 2.37(5)

Bondi’s vdW36 Rowland’s vdW37

3.55 3.57

180 180

2.95 2.86

compound

d(N···Cl) 3.108(3) 3.119(3) 3.133(6) 3.149(3) Comparison 3.30 3.40

∠(N−H···Cl)

d(H···N)

d(N···N)

∠(N−H···N)

164(1)° 157(1)° 153(1)° 149(1)°

1.98(5) 1.91(5) 1.98(5) 1.96(5)

2.692(3) 2.649(3) 2.679(7) 2.662(5)

139(1)° 141(1)° 136(1)° 137(1)°

2.75 2.74

3.10 3.28

180 180

180 180

(Pd + Pd 3.26 Å36), but less than the sum of Allinger vdW radii (Pd + Pd 4.66 Å39), and the latter suggests the possibility of existence of metallophilic Pd···Pd interactions. The availability of these contacts was confirmed theoretically by a DFT study (see below), which also gave an estimate of their energy. In the structures of 15−17 and 19, one of the two amino groups from the diaminocarbene fragment is involved in the intramolecular HB N2−H2···N3, thus forming a six-membered cycle. The distances between the N3 and H2 atoms and between the N3 and N1 atoms (Table 1) are substantially less than the sum of their Rowland37 or Bondi36 vdW radii. The corresponding angles are in the range 136−141°, and these contacts could be interpreted as HB in according with the IUPAC criterion.40 Moreover, in the structures of 15 and 16 geometric parameters for two weak C−H···Cl contacts, viz. between CyNC and chloride ligands, disagree with the HB IUPAC criterion40 for HB, as both H18−Cl1 (3.00(5) Å for 15, 2.81(5) Å for 16) and C18−Cl1 (3.598(4) Å for 15, 3.547(3) Å for 16) distances are greater than the sum of their Rowland37 and Bondi vdW36 radii of these atoms. However, we recognized these contacts as HBs by DFT calculations (see below). Theoretical Approach. Inspection of the crystallographic data indicates the presence of several types of noncovalent interactions in the dimers of 15−17 and 19: viz., the bifurcated CB−HB μ(S,N−H)Cl, intermolecular metallophilic interactions Pd···Pd, and intramolecular HB N−H···N. To confirm or disprove the hypothesis of the existence of these weak

All other bond lengths are usual, and their values agree with those reported for related palladium(II) carbene and isocyanide complexes.27a,b,e,h,34,35 Noncovalent Interactions in the Crystal Structures of 15−17 and 19. Experimental Approach. The XRD data for 15−17 and 19 indicate several types of noncovalent interactions: namely, intermolecular interactions between two symmetrically located neighboring molecules, i.e. bifurcated μ(S,N−H)Cl bonding, metallophilic PdII···PdII interactions, and intramolecular N−H···N hydrogen bonding (HB) (Figure 3). The geometric parameters for these noncovalent contacts are given in Table 1. The intermolecular interactions collectively connect two symmetrically located molecules in 15−17 and 19, providing their solid-state dimerization. In all cases, the S1 and H1 atoms in each molecule are linked with PdII-bound chloride (Cl2) in the neighboring molecule. The distances (Table 1) between the Cl2 and H1 and between the Cl2 and N1 atoms are less than the sum of their Rowland37 or even Bondi36 vdW radii, and the corresponding angles N1−H1−Cl2 are within the 110−180° range that is specific for a classic HB. The S1−Cl2 distances are also less than the sum of their Rowland37 or Bondi36 vdW radii of these atoms, and the angles C3−S1−C2 are in the range 150−180°; thus, this contact was explicitly recognized as CB4d,e,38 between the sulfur and the chlorine centers. The distances between the palladium atoms in the dimers (15, 3.9608(5) Å; 16, 4.1533(6) Å; 17, 4.1728(13) Å; 19: 4.1996(7) Å) are greater than the sum of the Bondi vdW radii D

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

(RDG) analysis43 was carried out, and RDG isosurfaces were plotted (Figures 5 and 6). The Poincare−Hopf relationship is satisfied; thus, all critical points have been found. In order to exclude the crystal-packing effects from consideration, we optimized the structure of the isolated supramolecular dimeric associate of 15 in the gas phase using the experimental X-ray geometry as a starting point. If the crystal-packing effects are significant, the structure should change appreciably on going from the solid state to the gas phase; otherwise the geometry is preserved in the isolated form. The geometry optimization leads to a very slight asymmetrical distortion of the dimer of 15 and to significant shortening of the H···Cl contacts (by 0.10−0.12 and 0.31−0.33 Å); the S···Cl contacts contracted to a much lesser degree (by 0.05 Å), whereas the Pd···Pd separation remained unchanged. The AIM analysis demonstrates the presence of bond critical points (BCPs) (3,−1) for all noncovalent interactions discussed above and also for one additional pair of intermolecular hydrogen bonds H···Cl, whose parameters disagree with the IUPAC criterion40 (C···Cl and H···Cl distances are greater than the sum of the Rowland37 or Bondi36 radii of these atoms) (Figure 6). The low magnitude of the electron density (0.005−0.034 hartree), positive values of the Laplacian (0.011−0.108 hartree), and close to zero energy density (−0.002−0.002 hartree) in these BCPs are typical for noncovalent interactions. We have defined energies for these contacts according to the procedures proposed by Espinosa et al.44 and Vener et al.45 (Table 2), and one can state that metallophilic interactions Pd···Pd and CB S···Cl in the dimer of 15 are very weak, whereas HBs N−H···N and H···Cl may be classified as moderate and weak strength contacts following the classification of Jeffrey (“strong” H bonds, 40−15 kcal/mol; “moderate” H-bonds, 15−4 kcal/mol; “weak” H-bonds, 1 is satisfied, then the nature of appropriate interaction is purely noncovalent; if −G(r)/V(r) < 1, some covalent component is present.47 On the basis of this criterion one can state that in 15 the covalent contribution in the metallophilic interaction Pd···Pd, CB S···Cl, and HB H···Cl is absent, but it is available (although almost negligible) in HBs H···Clbif and N−H···N. The negligible nonzero values of the Wiberg bond indices for Pd···Pd, S···Cl, H···Clbif, H···Cl, and N−H···N contacts in the optimized structures of the dimer of 15 (Table 2) computed by using the natural bond orbital (NBO) partitioning scheme48 additionally confirm the electrostatic nature of these noncovalent interactions. Since the chlorine and sulfur atoms have electron pairs and σ-holes, they both can act as electron donors and/or acceptors.49 To define the direction of charge transfer (CT) in this system, we applied NBO analysis. In the case of μ(S,N−H)Cl, the NBO analysis revealed the intermolecular charge transfers from the Cl ligands to the {S−C} and {N− H} moieties in the optimized gas-phase equilibrium structure for the dimer of 15. A second-order perturbation theory analysis indicates that these intermolecular charge transfers are mainly associated with the transitions LP(Cl) → σ*(S−C) and LP(Cl) → σ*(N−H); the total E(2) values are 2.6 and 36.2 kcal/mol, respectively. The calculated NBO atomic charges on the appropriate Cl, S, and H atoms are −0.47, 0.49, and 0.47, correspondingly.

interactions in the solid state, to quantify their energies, and to determine the influence of crystal-packing effects on the geometric features of the dimers, we carried out a Hirshfeld surface analysis of the crystal structure and DFT calculations and performed topological analysis of the electron density distribution within the formalism of Bader’s theory (AIM method)41 for the dimeric structure of 15 taken as a model. This approach has already been successfully used by us in studies of noncovalent interactions and properties of coordination bonds in various transition-metal complexes.22,24a,d,f−j,42 The molecular Hirshfeld surface represents an area where molecules come into contact, and its analysis gives the possibility of an additional insight into the nature of intermolecular interactions in the crystal state. For visualization, we have used a mapping of the normalized contact distance (dnorm); a negative value enables identification of molecular regions of substantial importance for the detection of short contacts. Figure 4 depicts the Hirshfeld surface of one complex

Figure 4. Hirshfeld surface of one PdII complex molecule in the crystal structure of 15 with a colored scale, which corresponds to values ranging from −0.5 Å (red) to 1.8 Å (blue).

molecule in the crystal structure of 15. In this Hirshfeld surface, the regions of bifurcated CB−HB, μ(S,N−H)Cl, are visualized by small (CB) and large (HB) red circled areas. The following intermolecular contacts give the largest contributions to the Hirshfeld surface: H−H, 52.1%; H−Cl, 20.2%; H−C, 12.8%; H−S, 4.0%; H−N, 2.3%; C−C, 2.0%; C−N, 1.7%; S−Cl, 1.7%. Individual contributions of other intermolecular contacts (viz., Pd−Pd, Pd−N, Pd−H, Pd−C, N−Cl, C−Cl, and S−S) are less than 1% and their additive contribution is 3.1%. Thus, the Hirshfeld surface analysis of 15 confirms that the solid dimeric associates are formed by various types of intermolecular contacts and, in particular, by the bifurcated μ(S,N−H)Cl linkage. However, this analysis does not answer the question of the energies of these contacts and, therefore, the DFT calculations needed to be further performed. To study the nature and strengths of these noncovalent interactions, we carried out DFT calculations and topological analysis of the electron density distribution within the formalism of Bader’s theory (AIM method).41 The results are summarized in Table 2; the contour line diagrams of the Laplacian distribution ∇2ρ(r), bond paths and selected zeroflux surfaces for 15 are shown in Figures 5 and 6. To visualize studied noncovalent interactions, reduced density gradient E

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Table 2. Values of the Density of All Electrons (ρ(r)), Laplacian of Electron Density (∇2ρ(r)), Energy Density (Hb), Potential Energy Density (V(r)), and Lagrangian Kinetic Energy (G(r), hartree) at the Bond Critical Points (3,−1), Corresponding to Noncovalent Interactions in 15 (M06/6-31G* Levels of Theory, MWB28 Pseudopotentials on the Pd Atoms), Appropriate Bond Lengths (d, Å), and Wiberg Bond Indices (WI), as well as Energies for These Contacts (Eint, kcal/mol), Defined by Three Approaches: Quasi-Solid-State (Experimental X-ray Geometrya), Optimized Geometry in the Gas Phase, and Optimized Geometry in CHCl3 Solution phase

ρ(r)

∇2ρ(r)

solid state

0.008 0.008 0.009 0.009 0.008 0.008

0.027 0.027 0.030 0.031 0.028 0.027

0.001 0.001 0.002 0.002 0.001 0.001

0.022 0.022 0.029 0.028 0.026 0.026

0.075 0.075 0.078 0.077 0.072 0.072

0.000 0.000 −0.002 −0.001 −0.001 −0.001

0.006 0.006 0.010 0.011 0.008 0.008

0.021 0.021 0.035 0.037 0.026 0.026

0.001 0.001 0.002 0.002 0.001 0.001

0.028 0.028 0.034 0.034 0.032 0.033

0.107 0.107 0.107 0.108 0.103 0.106

0.001 0.001 −0.001 −0.001 0.000 −0.001

0.006 0.006 0.005

0.012 0.014 0.011

0.000 0.000 0.000

gas phase CHCl3 solution

solid state gas phase CHCl3 solution

solid state gas phase CHCl3 solution

solid state gas phase CHCl3 solution

solid state gas phase CHCl3 solution

Hb

V(r) S···Cl −0.004 −0.004 −0.005 −0.005 −0.004 −0.004 H···Clbif −0.018 −0.018 −0.023 −0.022 −0.020 −0.020 H···Cl −0.003 −0.003 −0.006 −0.006 −0.004 −0.004 N−H···N −0.025 −0.025 −0.028 −0.028 −0.027 −0.027 Pd···Pd −0.003 −0.003 −0.002

G(r)

Eintb

Eintc

0.005 0.005 0.006 0.006 0.005 0.005

1.3 1.3 1.6 1.6 1.3 1.3

1.4 1.4 1.6 1.6 1.4 1.4

3.45 3.45 3.40 3.40 3.45 3.46

(0.01) (0.01) (0.01) (0.01)

0.018 0.018 0.021 0.021 0.019 0.019

5.7 5.7 7.2 6.9 6.3 6.3

4.9 4.9 5.7 5.7 5.1 5.1

2.27 2.27 2.15 2.17 2.20 2.20

(0.09) (0.08) (0.07) (0.07)

0.004 0.004 0.007 0.008 0.005 0.005

0.9 0.9 1.9 1.9 1.3 1.3

1.1 1.1 1.9 2.2 1.4 1.4

3.00 3.00 2.69 2.67 2.84 2.84

(0.01) (0.01) (0.01) (0.01)

0.026 0.026 0.027 0.028 0.026 0.027

7.8 7.8 8.8 8.8 8.5 8.5

7.0 7.0 7.3 7.5 7.0 7.3

1.98 1.98 1.91 1.90 1.93 1.91

(0.05) (0.05) (0.05) (0.05)

0.003 0.003 0.002

0.9 0.9 0.6

0.8 0.8 0.5

3.96 3.96 (0.03) 4.08 (0.03)

d(WI)

a

We also carried out additional DFT calculations and performed topological analysis of the electron density distribution within the formalism of Bader’s theory (AIM method) for the dimeric structure of 15 (experimental X-ray geometry) using the popular dispersion-corrected hybrid functional ωB97XD.46 The results are presented in Table S5 in the Supporting Information; the obtained values of the density of all electrons, Laplacian of electron density, energy density, potential energy density, and Lagrangian kinetic energy at the bond critical points (3, −1), corresponding to noncovalent interactions S···Cl, H···Clbif, H···Cl, N−H···N, and Pd···Pd, are almost the same as those found when the M06 functional is used. Thus, the estimated energies for the studied noncovalent interactions are almost independent of the DFT functional used. bEint = −V(r)/2.44 cEint = 0.429G(r).45

structure of trans-[PtCl2(NCSEt)2].22 All examples retrieved from the CCDC data and obtained from the literature could be systematized according to two parameters: viz., HB donor (C− H, N−H, O−H; Figure 7A) and source of Cl such as chloride anion, metal chloride complexes, and organic R−Cl species (Figure 7B). The dominant part of the published examples relates to the structures featuring both S···Cl and N−H···Cl contacts with metal-bound chlorides or chloride anion; these μ(S,N−H)(Cl−[M]) and μ(S,N−H)(Cl−) groups were identified by us in 6 species. Only a few reports deal with μ(S,C−H)Cl bifurcates (5 structures), and only 1 example of the bifurcate featuring μ(S,O−H)Cl− moiety has been reported. Dimerization of 15−19 in Solution: Experimental and Theoretical Considerations. The dimeric structure of 15−17 and 19 in the solid state allowed the assumption that the dimerization also occurs in solution. The relevant dimeric

The calculated vertical total energy (Ev,gas) for the dimer cleavage in gas phase is 57.4 kcal/mol (Figure S3 in the Supporting Information), which is comparable with, for instance, the energy of typical covalent C−N bonding in nitro compounds (50.5−72.5 kcal/mol).50 Retrieved and Processed CCDC Data for Bifurcated Chalcogen−Hydrogen μ(S,E−H)Cl Bonding (E = C, N, O). Our processing of available CCDC data explicitly indicates that the μ(S,E−H)Cl (E = C, N, O) pattern is the most abundant among the chalcogen−hydrogen bonding μ(Ch,E−H)Hal (Ch = S, Se, Te). However, even the μ(S,E−H)Cl functionality has not been classified as a specific moiety and the presence of both short contacts, namely separately S···Cl and separately E−H··· Cl in one structure, were mentioned only in 11 reports (Table 3).22,51 As indicated in the Introduction, the nature of CB in these species was recognized only by us in the solid state F

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Cl contact remained almost unchanged. The results of AIM analysis of the equilibrium geometry of the dimer in a CHCl3 solution are given in Table 2. To estimate the stability of the dimer in solution, we have employed two different approaches. First, we calculated the vertical total energy for the dimer cleavage on two monomeric parts with frozen coordinates in CHCl3 solution (Ev,solv = ∑Em,solv − Ed,solv; Figure S3) and found that it is also very high (56.5 kcal/mol). The contribution of the intermolecular bifurcated CB−HB μ(S,N−H)Cl to the dimer stabilization in a CHCl3 solution is 23−27%. Second, we performed full geometry optimization of the monomer of 15 and estimated the energy profit of its dimerization in CHCl3 solution in terms of Gibbs free energy (ΔGsolv = Gd,solv − 2Gm,solv; Figure S4), and this process is thermodynamically favorable by −9.3 kcal/mol. It is notable that the quantitative data might be overestimated because of the known uncertainty in the description of solvent effects53 and entropy terms.54 Nevertheless, the obtained theoretical results indicate that the dimerization is highly probable and these data encouraged the conductance of further experimental studies to prove the dimerization in solution. Experimental Verification. Comparison of the FTIR spectra in the solid state (KBr) with those in a CHCl3 solution does not indicate any difference in positions for ν(N−H) bands (ca. 3180 cm−1). The HRESI+ mass spectra of 15−19 exhibit the [2M − Cl]+ peaks, which were not previously reported for other relevant diaminocarbene complexes. In addition, strong downfield shifts of the NH protons in the 1H NMR spectra (δ 11.4−11.8 for 15−19 in CDCl3) are unusual for conventional complexes with acyclic diaminocarbene ligands (δ ranges from δ ca. 5.5 for electron-donor substituents to δ ca. 10.5 for strongly electron withdrawing substituents) and more resemble diaminocarbene species where the N−H protons are involved in intra- or intermolecular HB (δ 9.0−15.527g,h,52,55). These data collaterally confirm the presence of the dimeric associates in solution. On the other hand, complex 15 (which was chosen again as a model, now because of its good solubility in CDCl3) exhibits negligible concentration dependence of the 1H NMR signal of the NH proton involved in intermolecular HB: viz., from δ 11.53 to 11.44 in the concentration range from 100 to 1.5 mM (Figure S5). The presence of such concentration dependence reveals fast exchange between the monomeric and dimeric forms of 15. However, such weak dependence between the chemical shift and concentration probably suggests that the equilibrium between these two forms is shifted to one species in the studied concentration range.56 In order to study in more detail the dimerization of 15 in solution, we conducted NMR measurements of diffusion coefficients (D). The D values exhibit fair sensitivity to the effective size of supramolecular associates in solution.57 Under the conditions of fast exchange between the monomeric and dimeric forms of 15, the observed diffusion coefficient (Dobs) is an average of the species excising in a solution weighted with their relative amount. A series of D measurements were performed for 15 in CDCl3, at concentrations ranging from 159 to 26 mM. To eliminate the effect of variable viscosity at different concentrations on the Dobs value, tetramethylsilane (TMS) was used as an internal standard, which not subject to association, and the values of DTMS/Dobs were applied for further calculations. The first parameter we used to estimate the degree of dimerization of 15 in solution is an aggregation number N, which is defined as (D 0 /D obs ) 3 , where D 0

Figure 5. Contour line diagrams of the Laplacian distribution ∇2ρ(r), bond paths and selected zero-flux surfaces (left) and RDG isosurfaces (right) referring to the noncovalent interactions in the solid-state structure of 15 (intermolecular metallophilic interactions Pd···Pd, top; intermolecular heterobifurcated chalcogen/hydrogen bonding, middle; intramolecular hydrogen bonding N−H···N, bottom). Bond critical points (3,−1) are shown in blue, nuclear critical points (3,−3) in in pale brown, and ring critical points (3,+1) in orange. Lengths are given in Å and RDG isosurface values in hartrees.

structure pattern has previously been observed by Espinet et al.52 for the (diaminocarbene)AuI chloride (Figure 8), where the dimer is stabilized by the two intermolecular N−H···Cl HBs and by one aurophilic interaction. The presence of these contacts in the solid state was proved by X-ray diffraction, whereas for solution it was confirmed by 1H NMR spectroscopy (strong downfield shifts of the NH protons: viz., δ 12.4) and luminescence spectroscopy (characteristic emission shift specific for the aurophilic interactions). To confirm or disprove the presence of the dimeric form of 15−19 in solution, we have employed combined experimental and theoretical approaches. Theoretical Results. The high value of Ev,gas for the dimer of 15, taken as a model, implies the high stability of such an associate in the gas phase, and to study theoretically the influence of solvation effects on geometrical features and stability of the dimer of 15, we carried out its full geometry optimization in CHCl3 solution. In this case, the Pd···Pd contact was elongated (by 0.12 Å), H···Clbif and H···Cl contacts were shortened (by 0.07 and 0.16 Å, respectively), and the S··· G

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 6. (top) View of the dimeric structure of 15. Dotted red lines indicate the additional pair of intermolecular hydrogen bonds H···Cl verified by the AIM analysis. (bottom left) Contour line diagram of the Laplacian distribution ∇2ρ(r), bond paths and selected zero-flux surfaces and (bottom right) RDG isosurface referring to these noncovalent interactions in the solid-state structure of 15. Bond critical points (3,−1) are shown in blue and nuclear critical points (3,−3) in pale brown. Lengths are given in Å and RDG isosurface values in hartrees.

Table 3. Verification of the Chalcogen−Hydrogen Bonding μ(S,E−H)Cl via Our CCDC Data Processing and Inspection of Corresponding Geometrical Parameters compound/structure in CCDC [PtCl2(NCSEt)2]22/OJIMIF [ZnCl2{NCNC(MeS)2}2]51a/MUZYIR [CoCl2{NCNC(MeS)2}2]51b/PAFZOO [InCl3(4-bithiazole)(DMSO)]51c/PAKYEH [Zn(μ3-4-pyridylthio)(acetate)Cl]n51d/LALYEE Bondi’s vdW36 Rowland’s vdW37 [2-aminobenzo[d]thiazol-3-ium][SnCl6]51e/WAZJOY [5-bromo-2-aminobenzo[d]thiazol-3-ium][PbCl4Ph2]51f/ YEKQIP [5-amino-2-sulfonamidethiadiaz-4-ium]Cl51g/OXUVOT [6{tHPMT2}Cl3I4(I3)3]51h/FEBZUJ [2{tHPMT2}Cl(I3)3]51h/FECBAS [2{TU}(Cl)(I)(I3)]51i/GITPUW Bondi’s vdW36 Rowland’s vdW37 (H-thiamine)2[PtCl6]Cl2•2H2O51j/ WECBUB Bondi’s vdW36 Rowland’s vdW37

d(S···Cl)

∠(X−S···Cl)

μ(S,C−H)Cl 3.2987(7) 168.12(9) 3.3812(6) 177.83(6) 3.3743(9) 167.93(7) 3.5039(10) 158.46(9) 3.5136(11) 170.58(9) Comparison 3.55 180 3.57 180 μ(S,N−H)Cl 3.385(3) 177.2(2) 3.412(4) 170.9(5) 3.4599(17) 168.31(14) 3.482(2) 166.3(1) 3.314(2) 163.21(12) 3.4229(13) 162.38(9) Comparison 3.55 180 3.57 180 μ(S,O−H)Cl 3.487(3) 169.2(3) Comparison 3.55 180 3.57 180

corresponds to the formal absence of association upon infinite dilution, obtained by extrapolating the DTMS/Dobs value to zero

d(E···Cl)

d(H···Cl)

∠(E−H···Cl)

3.507(4) 3.4487(19) 3.450(2) 3.517(3) 3.503(3)

2.9887(6) 2.7276(4) 2.8845(6) 2.9029(7) 2.8240(8)

114.7(2) 130.77(10) 124.93(12) 116.68(15) 127.84(16)

3.53 3.45

2.95 2.86

180 180

3.283(7) 3.503(10)

2.486(2) 2.663(3)

154.4(4) 147.5(6)

Sn−Cl moiety Pb−Cl moiety

3.148(4) 3.141(6) 3.200(6) 3.155(5)

2.3420(12) 2.2873(14) 2.3695(4) 2.31261(9)

156.3(3) 163.5(4) 162.3(4) 166.5(4)

uncomplexed uncomplexed uncomplexed uncomplexed

3.3 3.4

2.95 2.86

180 180

3.137(8)

2.190(3)

150.1(4)

3.27 3.34

2.95 2.86

180 180

source of Cl Pt−Cl moiety Zn−Cl moiety Co−Cl moiety In−Cl moiety Zn−Cl moiety

Cl− Cl− Cl− Cl−

Uncomplexed Cl−

concentration of 15 (Figure 9) and is equivalent to the diffusion coefficient of the monomer.57a,b,58 At the highest H

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

and benzothiazol-2-amine 5) or open-chain aminocarbene species 15−19 (for the 4-aryl-substituted thiazol-2-amines 6− 10) (Scheme 1). Inspection of single-crystal X-ray diffraction data (Figure 3) indicates that 15−17 and 19 exhibit dimeric structures stabilized by previously unreported bifurcated chalcogen−hydrogen bonding μ(S,N−H)Cl accompanied by Pd···Pd and hydrogen bond interactions between the two monomeric parts. The presence of a μ(S,N−H)Cl moiety in the XRD structure of 15 was confirmed by Hirshfeld surface analysis (Figure 4) and DFT calculations (M06/6-31G* level of theory, MWB28 pseudopotentials on the Pd atoms) combined with a topological analysis of electron density distribution within the formalism of Bader’s theory (AIM method). The natural bond orbital analysis revealed that both LP(Cl)→σ*(S−C) and LP(Cl)→σ*(N−H) charge transfers are important in the bifurcated μ(S,N−H)Cl contacts. In a CHCl3 solution, this μ(S,N−H)Cl linkage, again along with metallophilic interactions and hydrogen bonding, is responsible for the dimerization of 15−19, as predicted theoretically by DFT calculations and confirmed experimentally upon analysis of the obtained FTIR, HRESI-MS, 1H NMR, and diffusion coefficient NMR data (for complex 15 taken as a model). Our processing of available CCDC data along with a Chemical Abstracts search indicates that μ(S,N−H)Cl, and more generally the noncovalent bifurcated heteroatom−hydrogen bonding μ(Ch,E−H)X (Ch = chalcogen; X = halide), have never been studied as separate phenomena. Our work is thus first to identify this bonding and to recognize its role in noncovalent associations.

Figure 7. Bifurcated chalcogen−hydrogen bonding μ(S,E−H)Cl.

Figure 8. Dimeric diaminocarbene complex of gold(I) stabilized by intermolecular N−H···Cl HBs and aurophilic interaction.52



EXPERIMENTAL SECTION

Materials and Instrumentation. Solvents, PdCl2, thiazol-2amines 2−4 and benzothiazol-2-amine 5, ketones S1−S5 (see the Supporting Information), and CNCy were obtained from commercial sources and used as received. The complex cis-[PdCl2(CNCy)2] (1) and thiazol-2-amines 6−10 were synthesized by the known procedures.29c,59 The C, H, and N elemental analyses were carried out on a Euro EA 3028 HT CHNSO analyzer. The mass spectra were obtained on a Bruker micrOTOF spectrometer equipped with electrospray ionization (ESI) source; CH2Cl2 was used as a solvent. The instrument was operated at positive ion mode using an m/z range of 50−3000. The capillary voltage of the ion source was set at −4500 V (ESI+) and the capillary exit at ±(70−150) V. The nebulizer gas pressure was 0.4 bar and drying gas flow 4.0 L/min. The most intense peak in the isotopic pattern is reported. Infrared spectra were recorded on a Shimadzu IRAffinity-1S FT-IR spectrometer (4000−400 cm−1) in KBr pellets for solid samples and in IR cells (NaCl; l = 0.13 mm) for CHCl3 solutions. TG/DTA measurements were performed with a NETZSCH TG 209 F1 Libra thermoanalyzer. The initial weights of the samples were in the 0.6−1.6 mg range. The experiments were run in an open alumina crucible under a stream of argon at a heating rate of 10 K min−1. The final temperature of the experiments was 450 °C. 1D (1H, 400 MHz; 13C{1H}, 101 MHz; 19F{1H}, 376 MHz; diffusion coefficient measurements) NMR spectra were acquired on a Bruker Avance 400 MHz spectrometer, while 2D (1H, 13C-HSQC, 1H,13CHMBC) NMR correlation experiments were recorded on a Bruker Avance III 500 MHz spectrometer. All NMR spectra were acquired in CDCl3 or DMSO-d6 at 298K. Synthetic Work. Synthesis of 11−14. A solution of 1 (20 mg, 5 mmol) in CH2Cl2 (3 mL) was added to any of the thiazol-2-amines 2−5 (10 mmol) placed in a 10 mL flask, and the mixture was stirred in air at room temperature for 2−4 h. The reaction mixtures gradually turned from pale yellow to intense lemon yellow, and solid (2−5)·HCl precipitated. The formed solution was filtered and evaporated at 20− 25 °C to dryness, and then the solid residue was washed with Et2O (three 1 mL portions) and dried in air at room temperature. Pure

Figure 9. Dependence of DTMS/Dobs vs concentration of 15 (CDCl3, 298 K).

acceptable concentration of 159 mM the aggregation number (N) of 15 is 1.33, revealing that it probably exists mostly in the monomeric form at 298 K (Table S6). Additionally, knowing N, C0, and the ratio of hydrodynamic volumes of monomer and dimerwhich could be taken from computation (for details see the Supporting Information)we estimated the dimerization constant of 15 in CDCl3. The value obtained at 298 K is 1.5 ± 0.5 M−1, and it indicates that ca. 25% of 15 exists as a dimer at the maximum solution concentration. Thus, our combined theoretical and experimental results demonstrate that the bifurcated chalcogen−hydrogen bonding μ(S,N−H)Cl is prone to dimerization not only in the solid state but also in solution.



CONCLUSIONS In this work, we found that the PdII-mediated reaction of CyNC with various thiazol-2-amines (2−4, 6−10) and benzothiazol-2-amine (5) proceeds differently in comparison to the respective reaction of XylNC,24a giving aminocarbenelike complexes 11−14 (in the cases of thiazol-2-amines 2−4 I

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry complexes 11−14 were isolated by dissolution of the raw products in a CH2Cl2 (2 mL)/Et2O (1.5 mL) mixture and then the resulting solutions were subjected to crystallization on slow evaporation. Synthesis of 15−19. A solution of 1 (20 mg, 5 mmol) in C2H4Cl2 (5 mL) was added to any of the thiazol-2-amines 6−10 (5 mmol) placed in a 10 mL flask, and the mixture was stirred in air upon reflux for 1−2 days. The reaction mixtures gradually turned from pale yellow to intense orange, and the solid thiazol-2-amines were dissolved. In the cases of 9 and 10, the final products 18 and 19 were precipitated as colorless solids from the appropriate solutions. After cooling, the solid residues were filtered off and 18 and 19 were then washed with diethyl ether (three 1 mL portions) and dried in air at room temperature. The reaction mixtures of 15−17 were evaporated at 20−25 °C to dryness, and then the solid residues were washed with Et2O (three 1 mL portions) and dried in air at room temperature. Pure complexes 15− 17 were isolated by dissolution of the formed residues in a CH2Cl2 (2 mL)/Et2O (1.5 mL) mixture, and then the resulting solutions were subjected to crystallization on slow evaporation. The pure complexes were obtained as yellow (11−14) or colorless (15−19) solids and were characterized by microanalyses (C, H, N), high-resolution ESI-MS, FTIR, TG/DTA, and 1D (1H, 13C{1H}, 19 1 F{ H}) and 2D (1H,1H-COSY, 1H,1H-NOESY, 1H,13C-HSQC, and 1 13 H, C-HMBC) NMR spectroscopy. Characterization. Characterization data and 1H, 13C{H}, and 19 1 F{ H} NMR spectra are included in the Supporting Information. X-ray Structure Determination. Single crystals of 11 and 15−17 were grown by the slow evaporation of solvent mixtures (Et2O/ CH2Cl2 for 11, 15, and 17 and Et2O/CHCl3 for 16) at room temperature. Single crystals of 19 were grown from CHCl3 at room temperature during the reaction. The X-ray experiments were conducted on a SuperNova, Dual, Cu at zero, Atlas diffractometer (CuKα, λ = 1.54184) for 15 and on a Xcalibur, Eos diffractometer (Mo Kα, λ = 0.71073) for 11, 16, 17, and 19. The crystals were kept at 100(2) K during data collection. Using Olex2,60 the structures were solved with ShelXS using direct methods (11, 15, and 16) and with Superflip using the Charge Flipping structure solution program (19), and all structures were refined with the ShelXL refinement package using least-squares minimization.61 The unit cell of 17 contains disordered molecules of solvent (CH2Cl2), which were treated as a diffuse contribution to the overall scattering without specific atom positions by SQUEEZE/PLATON.62 The carbon-bound H atoms were placed in calculated positions and were included in the refinement in the “riding” model approximation, with Uiso(H) set to 1.5[Ueq(C)] and C−H 0.96 Å for the CH3 groups, with Uiso(H) set to 1.2[Ueq(C)] and C−H 0.97 Å for the CH2 groups and C−H 0.93 Å for the CH groups, and Uiso(H) set to 1.2[Ueq(N)] and N−H 0.86 Å for the NH groups. Empirical absorption correction was applied with the CrysAlisPro63 program complex using spherical harmonics implemented in the SCALE3 ABSPACK scaling algorithm. Computational Details. The Hirshfeld molecular surface was generated by the CrystalExplorer 3.1 program64 on the basis of the results of the X-ray study. The normalized contact distances, dnorm,65 based on the Bondi van der Waals radii,36 were mapped into the Hirshfeld surface. In the color scale, negative values of dnorm are visualized by a red color indicating contacts shorter than the sum of van der Waals radii. The white color denotes intermolecular distances close to van der Waals contacts with dnorm equal to zero. In turn, contacts longer than the sum of van der Waals radii with positive dnorm values are colored in blue. The full geometry optimization in the gas phase and in CHCl3 solution and single-point calculations based on the experimental X-ray geometry (quasi-solid-state approach) of 15 have been carried out at the DFT level of theory using the M06 functional66 (this functional was specifically developed to describe weak dispersion forces and noncovalent interactions) with the help of the Gaussian-09 program package.67 The calculations were performed using the quasi-relativistic Stuttgart pseudopotential that described 28 core electrons and the appropriate contracted basis sets68 for the palladium atoms and the 631G* basis sets for other atoms. No symmetry restrictions have been applied during the geometry optimization, and experimental X-ray data

were used as starting point. The solvent effects were taken into account using the SMD continuum solvation model by Truhlar et al.69 The Hessian matrix was calculated analytically for the optimized structure in order to prove the location of correct minima (no imaginary frequencies) and to estimate the thermodynamic parameters, the latter being calculated at 25 °C. The topological analysis of the electron density distribution with the help of the atoms in molecules (AIM) method developed by Bader41 has been performed by using the Multiwfn program (version 3.3.8).70 The Cartesian atomic coordinates of the calculated equilibrium structures of 15 in the gas phase and CHCl3 solution are given in Table S4 in the Supporting Information. Processing of CCDC Data. Processing of the Cambridge Structure Database (v 5.38) was performed using the ConQuest module (v 1.19). Only structures with bifurcated chalcogen−hydrogen μ(Ch,E−H)Hal bonding with halogen atoms were included in the search (Ch = S, Se, Te; Hal = F, Cl, Br, I; Table S1). The analysis for μ(Ch,E−H)Hal was based on five parameters: viz., the distances Ch···X (d1), H···X (d2), and E···X (d3) and angles Y(E)−Ch−X (a1) and E− H−X (a2) (Figure 10). For S-containing systems the Ch···X and H···X

Figure 10. Parameters taken for processing of relevant CCDC data. distances were restricted by the sum of Rowland vdW radii and for Se and Te systems by the sum of Bondi vdW radii. The Y(E)−Ch−X and E−H−X angles was restricted to be greater than 150 and 110°, respectively. Only structures with determined 3D and with no error were included in the search query. In addition to this, disordered structures and powder structures were also excluded from the search. To ensure that we only had high-quality structures, the R factorwhich represents the agreement between the obtained crystallographic model and the experimental diffraction datawas kept at ≤0.1.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b00190. Experimental details for the synthesis and characterization of the complexes, crystallographic and computational information, and NMR spectra (PDF) Accession Codes

CCDC 1499892, 1535254, 1543620, and 1557909−1557910 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_request@ccdc. cam.ac.uk, 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 for V.P.B.: [email protected]. *E-mail for V.Y.K.: [email protected]. ORCID

Alexander S. Mikherdov: 0000-0002-6471-5158 J

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

(c) Servati Gargari, M.; Stilinović, V.; Bauzá, A.; Frontera, A.; McArdle, P.; Van Derveer, D.; Ng, S. W.; Mahmoudi, G. Design of Lead(II) Metal−Organic Frameworks Based on Covalent and Tetrel Bonding. Chem. - Eur. J. 2015, 21, 17951−17958. (7) Mahadevi, A. S.; Sastry, G. N. Cooperativity in Noncovalent Interactions. Chem. Rev. 2016, 116, 2775−2825. (8) (a) Mahadevi, A. S.; Sastry, G. N. Cation−π Interaction: Its Role and Relevance in Chemistry, Biology, and Material Science. Chem. Rev. 2013, 113, 2100−2138. (b) Dougherty, D. A. Y. The Cation − π Interaction. Acc. Chem. Res. 2013, 46, 885−893. (c) Wang, H.; Wang, W.; Jin, W. J. σ-Hole Bond vs π-Hole Bond: A Comparison Based on Halogen Bond. Chem. Rev. 2016, 116, 5072−5104. (d) Schottel, B. L.; Chifotides, H. T.; Dunbar, K. R. Anion-π interactions. Chem. Soc. Rev. 2008, 37, 68−83. (9) (a) Yam, V. W.-W.; Au, V. K.-M.; Leung, S. Y.-L. Light-Emitting Self-Assembled Materials Based on d8 and d10 Transition Metal Complexes. Chem. Rev. 2015, 115, 7589−7728. (b) Sculfort, S.; Braunstein, P. Intramolecular d10-d10 interactions in heterometallic clusters of the transition metals. Chem. Soc. Rev. 2011, 40, 2741−2760. (c) Eryazici, I.; Moorefield, C. N.; Newkome, G. R. Square-Planar Pd(II), Pt(II), and Au(III) Terpyridine Complexes: Their Syntheses, Physical Properties, Supramolecular Constructs, and Biomedical Activities. Chem. Rev. 2008, 108, 1834−1895. (10) (a) Cisneros, G. A.; Wikfeldt, K. T.; Ojamäe, L.; Lu, J.; Xu, Y.; Torabifard, H.; Bartók, A. P.; Csányi, G.; Molinero, V.; Paesani, F. Modeling Molecular Interactions in Water: From Pairwise to ManyBody Potential Energy Functions. Chem. Rev. 2016, 116, 7501−7528. (b) Ch’ng, L. C.; Samanta, A. K.; Czakó, G.; Bowman, J. M.; Reisler, H. Experimental and Theoretical Investigations of Energy Transfer and Hydrogen-Bond Breaking in the Water Dimer. J. Am. Chem. Soc. 2012, 134, 15430−15435. (11) (a) Wagner, J. R.; Lee, C. T.; Durrant, J. D.; Malmstrom, R. D.; Feher, V. A.; Amaro, R. E. Emerging Computational Methods for the Rational Discovery of Allosteric Drugs. Chem. Rev. 2016, 116, 6370− 6390. (b) Venditti, V.; Egner, T. K.; Clore, G. M. Hybrid Approaches to Structural Characterization of Conformational Ensembles of Complex Macromolecular Systems Combining NMR Residual Dipolar Couplings and Solution X-ray Scattering. Chem. Rev. 2016, 116, 6305− 6322. (12) (a) Chivers, T.; Laitinen, R. S. Tellurium: a maverick among the chalcogens. Chem. Soc. Rev. 2015, 44, 1725−1739. (b) Cozzolino, A. F.; Elder, P. J. W.; Vargas-Baca, I. A survey of tellurium-centered secondary-bonding supramolecular synthons. Coord. Chem. Rev. 2011, 255, 1426−1438. (c) Ho, P. C.; Szydlowski, P.; Sinclair, J.; Elder, P. J. W.; Kubel, J.; Gendy, C.; Lee, L. M.; Jenkins, H.; Britten, J. F.; Morim, D. R.; Vargas-Baca, I. Supramolecular macrocycles reversibly assembled by Te···O chalcogen bonding. Nat. Commun. 2016, 7, 11299. (d) Garrett, G. E.; Gibson, G. L.; Straus, R. N.; Seferos, D. S.; Taylor, M. S. Chalcogen Bonding in Solution: Interactions of Benzotelluradiazoles with Anionic and Uncharged Lewis Bases. J. Am. Chem. Soc. 2015, 137, 4126−4133. (13) (a) Wang, W.; Ji, B.; Zhang, Y. Chalcogen Bond: A Sister Noncovalent Bond to Halogen Bond. J. Phys. Chem. A 2009, 113, 8132−8135. (b) Bauza, A.; Quinonero, D.; Deya, P. M.; Frontera, A. Halogen bonding versus chalcogen and pnicogen bonding: a combined Cambridge structural database and theoretical study. CrystEngComm 2013, 15, 3137−3144. (14) (a) Geboes, Y.; De Vleeschouwer, F.; De Proft, F.; Herrebout, W. Exploiting the σ-hole Concept: an Infrared and Raman based characterization of the S···O Chalcogen Bond between 2,2,4,4Tetrafluoro-1,3-dithiethane and Dimethyl Ether. Chem. - Eur. J. 2017, 23, 17384−17392. (b) Peacock, H.; Luo, J.; Yamashita, T.; Luccarelli, J.; Thompson, S.; Hamilton, A. D. Non-covalent S···O interactions control conformation in a scaffold that disrupts islet amyloid polypeptide fibrillation. Chem. Sci. 2016, 7, 6435−6439. (c) Xiao, Q.; Sakurai, T.; Fukino, T.; Akaike, K.; Honsho, Y.; Saeki, A.; Seki, S.; Kato, K.; Takata, M.; Aida, T. Propeller-Shaped Fused Oligothiophenes: A Remarkable Effect of the Topology of Sulfur Atoms on Columnar Stacking. J. Am. Chem. Soc. 2013, 135, 18268−

Alexander S. Novikov: 0000-0001-9913-5324 Mikhail A. Kinzhalov: 0000-0001-5055-1212 Vadim Yu. Kukushkin: 0000-0002-2253-085X Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Support from the Russian Science Foundation (grant 14-4300017P) is gratefully acknowledged. Physicochemical studies were performed at the Center for Magnetic Resonance, the Center for Chemical Analysis and Materials Research, the Center for Thermogravimetric and Calorimetric Research, and the Center for X-ray Diffraction Studies (all belonging to Saint Petersburg State University). A.S.N. is thankful to the joint Saint Petersburg State University and Santander Bank program for the opportunity to conduct a computational part of this work during two months’ sabbatical on leave at Instituto Superior Tećnico, Universidade de Lisboa (Lisbon, Portugal).



DEDICATION Dedicated to Irina Petrovna Beletskaya (Full Member of RAS), “Mother and Father of Russian organometallic catalysis”, on the occasion of her jubilee.



REFERENCES

(1) Hobza, P.; Rezac, J. Introduction: Noncovalent Interactions. Chem. Rev. 2016, 116, 4911−4912. (2) (a) Kollman, P. A.; Allen, L. C. Theory of the hydrogen bond. Chem. Rev. 1972, 72, 283−303. (b) Grabowski, S. J. What Is the Covalency of Hydrogen Bonding? Chem. Rev. 2011, 111, 2597−2625. (3) (a) Gilday, L. C.; Robinson, S. W.; Barendt, T. A.; Langton, M. J.; Mullaney, B. R.; Beer, P. D. Halogen Bonding in Supramolecular Chemistry. Chem. Rev. 2015, 115, 7118−7195. (b) Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. The Halogen Bond. Chem. Rev. 2016, 116, 2478−2601. (c) Kolar, M. H.; Hobza, P. Computer Modeling of Halogen Bonds and Other sigma-Hole Interactions. Chem. Rev. 2016, 116, 5155− 5187. (4) (a) Gleiter, R.; Haberhauer, G.; Werz, D. B.; Rominger, F.; Bleiholder, C. From Noncovalent Chalcogen−Chalcogen Interactions to Supramolecular Aggregates: Experiments and Calculations. Chem. Rev. 2018, DOI: 10.1021/acs.chemrev.7b00449. (b) Pascoe, D. J.; Ling, K. B.; Cockroft, S. L. The Origin of Chalcogen-Bonding Interactions. J. Am. Chem. Soc. 2017, 139, 15160−15167. (c) Mahmudov, K. T.; Kopylovich, M. N.; Guedes da Silva, M. F. C.; Pombeiro, A. J. L. Chalcogen bonding in synthesis, catalysis and design of materials. Dalton Trans. 2017, 46, 10121−10138. (d) Zhao, H.; Gabbaï, F. P. A bidentate Lewis acid with a telluronium ion as an anion-binding site. Nat. Chem. 2010, 2, 984−990. (e) Bleiholder, C.; Werz, D. B.; Koppel, H.; Gleiter, R. Theoretical investigations on chalcogen-chalcogen interactions: What makes these nonbonded interactions bonding? J. Am. Chem. Soc. 2006, 128, 2666−2674. (5) (a) Bauza, A.; Mooibroek, T. J.; Frontera, A. The Bright Future of Unconventional sigma/-Hole Interactions. ChemPhysChem 2015, 16, 2496−2517. (b) Scheiner, S. The Pnicogen Bond: Its Relation to Hydrogen, Halogen, and Other Noncovalent Bonds. Acc. Chem. Res. 2013, 46, 280−288. (c) Zahn, S.; Frank, R.; Hey-Hawkins, E.; Kirchner, B. Pnicogen Bonds: A New Molecular Linker? Chem. - Eur. J. 2011, 17, 6034−6038. (6) (a) Mahmoudi, G.; Bauzá, A.; Amini, M.; Molins, E.; Mague, J. T.; Frontera, A. On the importance of tetrel bonding interactions in lead(II) complexes with (iso)nicotinohydrazide based ligands and several anions. Dalton Trans. 2016, 45, 10708−10716. (b) Mahmoudi, G.; Bauzá, A.; Frontera, A. Concurrent agostic and tetrel bonding interactions in lead(II) complexes with an isonicotinohydrazide based ligand and several anions. Dalton Trans. 2016, 45, 4965−4969. K

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Bonding in Metal-Bound Alkylthiocyanates. Cryst. Growth Des. 2016, 16, 2979−2987. (23) (a) Boyarskiy, V. P.; Bokach, N. A.; Luzyanin, K. V.; Kukushkin, V. Y. Metal-Mediated and Metal-Catalyzed Reactions of Isocyanides. Chem. Rev. 2015, 115, 2698−2779. (b) Boyarskiy, V. P.; Luzyanin, K. V.; Kukushkin, V. Y. Acyclic diaminocarbenes (ADCs) as a promising alternative to N-heterocyclic carbenes (NHCs) in transition metal catalyzed organic transformations. Coord. Chem. Rev. 2012, 256, 2029−2056. (24) (a) Mikherdov, A. S.; Kinzhalov, M. A.; Novikov, A. S.; Boyarskiy, V. P.; Boyarskaya, I. A.; Dar’in, D. V.; Starova, G. L.; Kukushkin, V. Y. Difference in Energy between Two Distinct Types of Chalcogen Bonds Drives Regioisomerization of Binuclear (Diaminocarbene)PdII Complexes. J. Am. Chem. Soc. 2016, 138, 14129−14137. (b) Bikbaeva, Z. M.; Ivanov, D. M.; Novikov, A. S.; Ananyev, I. V.; Bokach, N. A.; Kukushkin, V. Y. Electrophilic− Nucleophilic Dualism of Nickel(II) toward Ni···I Noncovalent Interactions: Semicoordination of Iodine Centers via Electron Belt and Halogen Bonding via σ-Hole. Inorg. Chem. 2017, 56, 13562− 13578. (c) Adonin, S. A.; Gorokh, I. D.; Novikov, A. S.; Abramov, P. A.; Sokolov, M. N.; Fedin, V. P. Halogen Contacts-Induced Unusual Coloring in Bi-III Bromide Complex: Anion-to-Cation Charge Transfer via Br···Br Interactions. Chem. - Eur. J. 2017, 23, 15612− 15616. (d) Ivanov, D. M.; Novikov, A. S.; Ananyev, I. V.; Kirina, Y. V.; Kukushkin, V. Y. Halogen bonding between metal centers and halocarbons. Chem. Commun. 2016, 52, 5565−5568. (e) Novikov, A. S.; Ivanov, D. M.; Avdontceva, M. S.; Kukushkin, V. Y. Diiodomethane as a halogen bond donor toward metal-bound halides. CrystEngComm 2017, 19, 2517−2525. (f) Ivanov, D. M.; Kinzhalov, M. A.; Novikov, A. S.; Ananyev, I. V.; Romanova, A. A.; Boyarskiy, V. P.; Haukka, M.; Kukushkin, V. Y. H2C(X)-X···X- (X = CI, Br) Halogen Bonding of Dihalomethanes. Cryst. Growth Des. 2017, 17, 1353−1362. (g) Ding, X.; Tuikka, M. J.; Hirva, P.; Kukushkin, V. Y.; Novikov, A. S.; Haukka, M. Fine-tuning halogen bonding properties of diiodine through halogen-halogen charge transfer - extended [Ru(2,2’-bipyridine)(CO)2 × 2]···I2 systems (X = Cl, Br, I). CrystEngComm 2016, 18, 1987− 1995. (h) Ivanov, D. M.; Novikov, A. S.; Starova, G. L.; Haukka, M.; Kukushkin, V. Y. A family of heterotetrameric clusters of chloride species and halomethanes held by two halogen and two hydrogen bonds. CrystEngComm 2016, 18, 5278−5286. (i) Mikhaylov, V. N.; Sorokoumov, V. N.; Korvinson, K. A.; Novikov, A. S.; Balova, I. A. Synthesis and Simple Immobilization of Palladium(II) Acyclic Diaminocarbene Complexes on Polystyrene Support as Efficient Catalysts for Sonogashira and Suzuki−Miyaura Cross-Coupling. Organometallics 2016, 35, 1684−1697. (j) Serebryanskaya, T. V.; Novikov, A. S.; Gushchin, P. V.; Haukka, M.; Asfin, R. E.; Tolstoy, P. M.; Kukushkin, V. Y. Identification and H(D)-bond energies of CH(D)center dot center dot center dot Cl interactions in chloridehaloalkane clusters: a combined X-ray crystallographic, spectroscopic, and theoretical study. Phys. Chem. Chem. Phys. 2016, 18, 14104− 14112. (25) Moss, G. P.; Smith, P. A. S.; Tavernier, D. Glossary of Class Names of Organic-Compounds and Reactive Intermediates Based on Structure. Pure Appl. Chem. 1995, 67, 1307−1375. (26) Kinzhalov, M. A.; Starova, G. L.; Boyarskiy, V. P. Interaction of benzene-1,2-diamines with isocyanide complexes of palladium(II): Insight into the mechanism. Inorg. Chim. Acta 2017, 455, 607−612. (27) (a) Timofeeva, S. A.; Kinzhalov, M. A.; Valishina, E. A.; Luzyanin, K. V.; Boyarskiy, V. P.; Buslaeva, T. M.; Haukka, M.; Kukushkin, V. Y. Application of palladium complexes bearing acyclic amino(hydrazido)carbene ligands as catalysts for copper-free Sonogashira cross-coupling. J. Catal. 2015, 329, 449−456. (b) Kinzhalov, M. A.; Luzyanin, K. V.; Boyarskiy, V. P.; Haukka, M.; Kukushkin, V. Y. ADC-Based Palladium Catalysts for Aqueous Suzuki Miyaura CrossCoupling Exhibit Greater Activity than the Most Advantageous Catalytic Systems. Organometallics 2013, 32, 5212−5223. (c) Kinzhalov, M. A.; Boyarskiy, V. P.; Luzyanin, K. V.; Dolgushin, F. M.; Kukushkin, V. Y. Metal-mediated coupling of a coordinated isocyanide and indazoles. Dalton Trans. 2013, 42, 10394−10397. (d) Anisimova,

18271. (d) Wang, X.-Y.; Jiang, W.; Chen, T.; Yan, H.-J.; Wang, Z.-H.; Wan, L.-J.; Wang, D. Molecular evidence for the intermolecular S ··· S interaction in the surface molecular packing motifs of a fused thiophene derivative. Chem. Commun. 2013, 49, 1829−1831. (15) Breugst, M.; von der Heiden, D.; Schmauck, J. Novel Noncovalent Interactions in Catalysis: A Focus on Halogen, Chalcogen, and Anion-pi Bonding. Synthesis 2017, 49, 3224−3236. (16) (a) Jungbauer, S. H.; Huber, S. M. Cationic Multidentate Halogen-Bond Donors in Halide Abstraction Organocatalysis: Catalyst Optimization by Preorganization. J. Am. Chem. Soc. 2015, 137, 12110− 12120. (b) Jungbauer, S. H.; Walter, S. M.; Schindler, S.; Rout, L.; Kniep, F.; Huber, S. M. Activation of a carbonyl compound by halogen bonding. Chem. Commun. 2014, 50, 6281−6284. (c) He, W.; Ge, Y.C.; Tan, C.-H. Halogen-Bonding-Induced Hydrogen Transfer to C N Bond with Hantzsch Ester. Org. Lett. 2014, 16, 3244−3247. (d) Kniep, F.; Rout, L.; Walter, S. M.; Bensch, H. K. V.; Jungbauer, S. H.; Herdtweck, E.; Huber, S. M. 5-Iodo-1,2,3-triazolium-based multidentate halogen-bond donors as activating reagents. Chem. Commun. 2012, 48, 9299−9301. (17) (a) Benz, S.; López-Andarias, J.; Mareda, J.; Sakai, N.; Matile, S. Catalysis with Chalcogen Bonds. Angew. Chem., Int. Ed. 2017, 56, 812−815. (b) Benz, S.; Mareda, J.; Besnard, C.; Sakai, N.; Matile, S. Catalysis with chalcogen bonds: neutral benzodiselenazole scaffolds with high-precision selenium donors of variable strength. Chem. Sci. 2017, 8, 8164−8169. (c) Wonner, P.; Vogel, L.; Düser, M.; Gomes, L.; Kniep, F.; Mallick, B.; Werz, D. B.; Huber, S. M. Carbon-Halogen Bond Activation by Selenium-Based Chalcogen Bonding. Angew. Chem., Int. Ed. 2017, 56, 12009−12012. (18) Molina, P.; Zapata, F.; Caballero, A. Anion Recognition Strategies Based on Combined Noncovalent Interactions. Chem. Rev. 2017, 117, 9907−9972. (19) (a) Barendt, T. A.; Docker, A.; Marques, I.; Félix, V.; Beer, P. D. Selective Nitrate Recognition by a Halogen-Bonding Four-Station [3]Rotaxane Molecular Shuttle. Angew. Chem., Int. Ed. 2016, 55, 11069−11076. (b) Zapata, F.; Caballero, A.; White, N. G.; Claridge, T. D. W.; Costa, P. J.; Félix, V. t.; Beer, P. D. Fluorescent Charge-Assisted Halogen-Bonding Macrocyclic Halo-Imidazolium Receptors for Anion Recognition and Sensing in Aqueous Media. J. Am. Chem. Soc. 2012, 134, 11533−11541. (c) Robinson, S. W.; Mustoe, C. L.; White, N. G.; Brown, A.; Thompson, A. L.; Kennepohl, P.; Beer, P. D. Evidence for Halogen Bond Covalency in Acyclic and Interlocked Halogen-Bonding Receptor Anion Recognition. J. Am. Chem. Soc. 2015, 137, 499−507. (d) Walter, S. M.; Kniep, F.; Rout, L.; Schmidtchen, F. P.; Herdtweck, E.; Huber, S. M. Isothermal Calorimetric Titrations on ChargeAssisted Halogen Bonds: Role of Entropy, Counterions, Solvent, and Temperature. J. Am. Chem. Soc. 2012, 134 (20), 8507−8512. (20) (a) Benz, S.; Macchione, M.; Verolet, Q.; Mareda, J.; Sakai, N.; Matile, S. Anion Transport with Chalcogen Bonds. J. Am. Chem. Soc. 2016, 138, 9093−9096. (b) Scheiner, S. Highly Selective Halide Receptors Based on Chalcogen, Pnicogen, and Tetrel Bonds. Chem. Eur. J. 2016, 22, 18850−18858. (c) Garrett, G. E.; Carrera, E. I.; Seferos, D. S.; Taylor, M. S. Anion recognition by a bidentate chalcogen bond donor. Chem. Commun. 2016, 52, 9881−9884. (d) Lim, J. Y. C.; Marques, I.; Thompson, A. L.; Christensen, K. E.; Félix, V.; Beer, P. D. Chalcogen Bonding Macrocycles and [2]Rotaxanes for Anion Recognition. J. Am. Chem. Soc. 2017, 139, 3122− 3133. (21) (a) Feldblum, E. S.; Arkin, I. T. Strength of a bifurcated H bond. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 4085−4090. (b) Schneider, H. J. Binding Mechanisms in Supramolecular Complexes. Angew. Chem., Int. Ed. 2009, 48, 3924−3977. (c) Steiner, T. The Hydrogen Bond in the Solid State. Angew. Chem., Int. Ed. 2002, 41, 48−76. (d) Rozas, I.; Alkorta, I.; Elguero, J. Bifurcated Hydrogen Bonds: Three-Centered Interactions. J. Phys. Chem. A 1998, 102, 9925−9932. (e) Aakeröy, C. B.; Rajbanshi, A.; Desper, J. Hydrogen-bond driven assembly of a molecular capsule facilitated by supramolecular chelation. Chem. Commun. 2011, 47, 11411−11413. (22) Yandanova, E. S.; Ivanov, D. M.; Kuznetsov, M. L.; Starikov, A. G.; Starova, G. L.; Kukushkin, V. Y. Recognition of S···CI Chalcogen L

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

expected from van der Waals radii. J. Phys. Chem. 1996, 100, 7384− 7391. (38) (a) Cozzolino, A. F.; Vargas-Baca, I.; Mansour, S.; Mahmoudkhani, A. H. The nature of the supramolecular association of 1,2,5-chalcogenadiazoles. J. Am. Chem. Soc. 2005, 127, 3184−3190. (b) Burling, F. T.; Goldstein, B. M. Computational Studies of Nonbonded Sulfur Oxygen and Selenium Oxygen Interactions in the Thiazole and Selenazole Nucleosides. J. Am. Chem. Soc. 1992, 114, 2313−2320. (c) Rosenfield, R. E.; Parthasarathy, R.; Dunitz, J. D. Directional Preferences of Nonbonded Atomic Contacts with Divalent Sulfur 0.1. Electrophiles and Nucleophiles. J. Am. Chem. Soc. 1977, 99, 4860−4862. (39) Allinger, N. L.; Zhou, X. F.; Bergsma, J. Molecular Mechanics Parameters. J. Mol. Struct.: THEOCHEM 1994, 312, 69−83. (40) Arunan, E.; Desiraju, G. R.; Klein, R. A.; Sadlej, J.; Scheiner, S.; Alkorta, I.; Clary, D. C.; Crabtree, R. H.; Dannenberg, J. J.; Hobza, P.; Kjaergaard, H. G.; Legon, A. C.; Mennucci, B.; Nesbitt, D. J. Definition of the hydrogen bond (IUPAC Recommendations 2011). Pure Appl. Chem. 2011, 83, 1637−1641. (41) Bader, R. F. W. A Quantum-Theory of Molecular-Structure and Its Applications. Chem. Rev. 1991, 91, 893−928. (42) Novikov, A. S.; Kuznetsov, M. L.; Pombeiro, A. J. L. Theory of the Formation and Decomposition of N-Heterocyclic Aminooxycarbenes through Metal-Assisted 2 + 3 -Dipolar Cycloaddition/ Retro-Cycloaddition. Chem. - Eur. J. 2013, 19, 2874−888. (43) Johnson, E. R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A. J.; Yang, W. Revealing Noncovalent Interactions. J. Am. Chem. Soc. 2010, 132, 6498−6506. (44) Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen bond strengths revealed by topological analyses of experimentally observed electron densities. Chem. Phys. Lett. 1998, 285, 170−173. (45) Vener, M. V.; Egorova, A. N.; Churakov, A. V.; Tsirelson, V. G. Intermolecular hydrogen bond energies in crystals evaluated using electron density properties: DFT computations with periodic boundary conditions. J. Comput. Chem. 2012, 33, 2303−2309. (46) Chai, J. D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615−6620. (47) Espinosa, E.; Alkorta, I.; Elguero, J.; Molins, E. From weak to strong interactions: A comprehensive analysis of the topological and energetic properties of the electron density distribution involving X-H center dot center dot center dot F-Y systems. J. Chem. Phys. 2002, 117, 5529−5542. (48) Glendening, E. D.; Landis, C. R.; Weinhold, F. Natural bond orbital methods. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 1− 42. (49) (a) Sánchez-Sanz, G.; Trujillo, C. Improvement of Anion Transport Systems by Modulation of Chalcogen Interactions: The influence of solvent. J. Phys. Chem. A 2018, 122, 1369−1377. (b) Sanchez-Sanz, G.; Trujillo, C.; Alkorta, I.; Elguero, J. Weak interactions between hypohalous acids and dimethylchalcogens. Phys. Chem. Chem. Phys. 2012, 14, 9880−889. (50) Blanksby, S. J.; Ellison, G. B. Bond dissociation energies of organic molecules. Acc. Chem. Res. 2003, 36, 255−263. (51) (a) Diop, M. B.; Diop, L.; Oliver, A. G. IUCr, Crystal structure of di-chlorido-bis-(dimethyl N-cyano-dithio-imino-carbonate)zinc. Acta Crystallogr., Sect. E: Crystallogr. Commun. 2016, 72, 417−419. (b) Diop, M. B.; Diop, L.; Oliver, A. G. Crystal structure of dichloridobis(dimethyl N-cyanodithioiminocarbonate)cobalt(II). Acta Crystallogr., Sect. E: Crystallogr. Commun. 2016, 72, 66−68. (c) Abedi, A.; Safari, N.; Amani, V.; Khavasi, H. R. Indium(III) complexes containing bithiazole derivatives, chloride, methanol, and dimethyl sulfoxide: X-ray studies, spectroscopic characterization, and thermal analyses. J. Coord. Chem. 2012, 65, 325−338. (d) Zhang, Y. H.; Wang, J.; Zheng, L. L.; Hu, S.; Lin, Z. J.; Tong, M. L. Structural diversity and reactivity of d(10) metal-(4-pyridylthio)acetate system. Chin. Sci. Bull. 2009, 54, 4277−4284. (e) Jin, S.; Wang, D. Syntheses and molecular structures of four inorganic−organic hybrid compounds from Ncontaining aromatic bases and perchlorometallates of Zn, Cu, Sn, and

T. B.; Guedes da Silva, M. F. C.; Kukushkin, V. Y.; Pombeiro, A. J. L.; Luzyanin, K. V. Metal-mediated coupling of amino acid esters with isocyanides leading to new chiral acyclic aminocarbene complexes. Dalton Trans. 2014, 43, 15861−15871. (e) Luzyanin, K. V.; Pombeiro, A. J. L.; Haukka, M.; Kukushkin, V. Y. Coupling between 3Iminoisoindolin-1-ones and Complexed Isonitriles as a MetalMediated Route to a Novel Type of Palladium and Platinum Iminocarbene Species. Organometallics 2008, 27, 5379−5389. (f) Chay, R. S.; Luzyanin, K. V.; Kukushkin, V. Y.; Guedes da Silva, M. F. C.; Pombeiro, A. J. L. Novel Palladium-Aminocarbene Species Derived from Metal-Mediated Coupling of Isonitriles and 1,3Diiminoisoindoline: Synthesis and Catalytic Application in SuzukiMiyaura Cross-Coupling. Organometallics 2012, 31, 2379−2387. (g) Valishina, E. A.; Guedes da Silva, M. F. C.; Kinzhalov, M. A.; Timofeeva, S. A.; Buslaeva, T. M.; Haukka, M.; Pombeiro, A. J. L.; Boyarskiy, V. P.; Kukushkin, V. Y.; Luzyanin, K. V. Palladium-ADC complexes as efficient catalysts in copper-free and room temperature Sonogashira coupling. J. Mol. Catal. A: Chem. 2014, 395, 162−171. (h) Kinzhalov, M. A.; Luzyanin, K. V.; Boyarskiy, V. P.; Haukka, M.; Kukushkin, V. Y. Coupling of C-amino aza-substituted heterocycles with an isocyanide ligand in palladium(II) complex. Russ. Chem. Bull. 2013, 62, 758−766. (28) (a) Abraham, R. J.; Mobli, M. An NMR, IR and theoretical investigation of 1H Chemical Shifts and hydrogen bonding in phenols. Magn. Reson. Chem. 2007, 45, 865−877. (b) Haushalter, K. A.; Lau, J.; Roberts, J. D. An NMR Investigation of the Effect of Hydrogen Bonding on the Rates of Rotation about the C−N Bonds in Urea and Thiourea. J. Am. Chem. Soc. 1996, 118, 8891−8896. (29) (a) Martinez-Martinez, A. J.; Chicote, M. T.; Bautista, D.; Vicente, J. Synthesis of Palladium(II), -(III), and -(IV) Complexes with Acyclic Diaminocarbene Ligands. Organometallics 2012, 31, 3711−3719. (b) Wanniarachchi, Y. A.; Slaughter, L. M. One-step assembly of a chiral palladium bis(acyclic diaminocarbene) complex and its unexpected oxidation to a bis(amidine) complex. Chem. Commun. 2007, 0, 3294−3296. (c) Tskhovrebov, A. G.; Luzyanin, K. V.; Dogushin, F. M.; Guedes da Silva, M. F. C.; Pombeiro, A. J. L.; Kukushkin, V. Y. Novel Reactivity Mode of Metal Diaminocarbenes: Palladium(II)-Mediated Coupling between Acyclic Diaminocarbenes and Isonitriles Leading to Dinuclear Species. Organometallics 2011, 30, 3362−3370. (30) Kitano, Y.; Hori, T. Structure of cis-Dichlorobis(Cyclohexyl Isocyanide) Palladium(II). Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1981, 37, 1919−1921. (31) Perreault, D.; Drouin, M.; Michel, A.; Harvey, P. D. Relationships between Metal-Metal Force-Constants and MetalMetal Separations for Ag2 and Pd2 Systems - Crystal and Molecular-Structures of Ag2(Dmb)2 × 2 (X = Cl, Br, I Dmb = 1,8Diisocyano-P-Menthane) And Cis-Pd(Cnc(CH3)3)2Cl2 Complexes. Inorg. Chem. 1993, 32, 1903−1912. (32) Drouin, M.; Perreault, D.; Harvey, P. D.; Michel, A. Quasi-OneDimensional Structure of Cis-Dichlorobis(2,6-Dimethylphenyl Isocyanide)-Palladium(Ii), [Pd(2,6-(CH3)2C6H3NC)2Cl2]. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1991, 47, 752−754. (33) Allen, F. H.; Kennard, O.; Watson, D. G.; Brammer, L.; Orpen, A. G.; Taylor, R. Tables of Bond Lengths Determined by X-Ray and Neutron-Diffraction 0.1. Bond Lengths in Organic-Compounds. J. Chem. Soc., Perkin Trans. 2 1987, 2, S1−S19. (34) Luzyanin, K. V.; Guedes da Silva, M. F. C.; Kukushkin, V. Y.; Pombeiro, A. J. L. First example of an imine addition to coordinated isonitrile. Inorg. Chim. Acta 2009, 362, 833−838. (35) Moncada, A. I.; Manne, S.; Tanski, J. M.; Slaughter, L. M. Modular chelated palladium diaminocarbene complexes: Synthesis, characterization, and optimization of catalytic Suzuki-Miyaura crosscoupling activity by ligand modification. Organometallics 2006, 25, 491−505. (36) Bondi, A. Van Der Waals Volumes + Radii. J. Phys. Chem. 1964, 68, 441−451. (37) Rowland, R. S.; Taylor, R. Intermolecular nonbonded contact distances in organic crystal structures: Comparison with distances M

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Fe. J. Coord. Chem. 2012, 65, 3188−3204. (f) Garcia-Hernandez, Z.; Flores-Parra, A.; Grevy, J. M.; Ramos-Organillo, A.; Contreras, R. 2Aminobenzothiazole phosphorus amides: Molecular and supramolecular structures, hydrogen bonds and sulfur donor-acceptor interactions. Polyhedron 2006, 25, 1662−1672. (g) Díaz, J. R. A.; Camí, G. E.; González, M. L.; Pedregosa, J. C.; Vega, D. R. Two Novel Potential Anticonvulsant Salts of 5-Amino-2-Sulfonamide-1,3,4Thiadiazole. J. Chem. Crystallogr. 2011, 41, 1114−1119. (h) Owczarzak, A. M.; Kubicki, M.; Kourkoumelis, N.; Hadjikakou, S. K. Stabilization of poly-iodides: structural influences of the cationic disulfides of 2mercapto-3,4,5,6-tetrahydro-pyrimidine and 2-mercatpo-pyrimidine. RSC Adv. 2012, 2, 2856−2867. (i) Biesiada, M.; Kourkoumelis, N.; Kubicki, M.; Owczarzak, A. M.; Balas, V.; Hadjikakou, S. K. Fundamental chemistry of iodine. The reaction of di-iodine towards thiourea and its methyl-derivative: formation of aminothiazoles and aminothiadiazoles through dicationic disulfides. Dalton Trans. 2014, 43, 4790−4806. (j) Aoki, K.; Tokuno, T.; Takagi, K.; Hirose, Y.; Suh, I.-H.; Adeyemo, A. O.; Williams, G. N. The [PtIV Cl6]2−/thiamine system. Crystal structures of (H-thiamine)[PtCl6]·H2O and (Hthiamine)2[PtCl6]Cl2·2H2O: hostguest-like anion coordination. Inorg. Chim. Acta 1993, 210, 17−25. (52) Bartolomé, C.; Carrasco-Rando, M.; Coco, S.; Cordovilla, C.; Martín-Alvarez, J. M.; Espinet, P. Luminescent Gold(I) Carbenes from 2-Pyridylisocyanide Complexes: Structural Consequences of Intramolecular versus Intermolecular Hydrogen-Bonding Interactions. Inorg. Chem. 2008, 47, 1616−1624. (53) (a) Tomasi, J. Selected features of the polarizable continuum model for the representation of solvation. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2011, 1, 855−867. (b) Mennucci, B. Polarizable continuum model. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 386−404. (c) Jinich, A.; Rappoport, D.; Dunn, I.; Sanchez-Lengeling, B.; Olivares-Amaya, R.; Noor, E.; Bar Even, A.; Aspuru-Guzik, A. Quantum Chemical Approach to Estimating the Thermodynamics of Metabolic Reactions. Sci. Rep. 2015, 4, 7022. (54) (a) Suarez, D.; Diaz, N. Direct methods for computing singlemolecule entropies from molecular simulations. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2015, 5, 1−26. (b) Zhou, H. X.; Gilson, M. K. Theory of Free Energy and Entropy in Noncovalent Binding. Chem. Rev. 2009, 109, 4092−4107. (c) Biedermann, F.; Schneider, H. J. Experimental Binding Energies in Supramolecular Complexes. Chem. Rev. 2016, 116, 5216−5300. (55) Kinzhalov, M. A.; Timofeeva, S. A.; Luzyanin, K. V.; Boyarskiy, V. P.; Yakimanskiy, A. A.; Haukka, M.; Kukushkin, V. Y. Palladium(II)Mediated Addition of Benzenediamines to Isocyanides: Generation of Three Types of Diaminocarbene Ligands Depending on the Isomeric Structure of the Nucleophile. Organometallics 2016, 35, 218−228. (56) (a) Pastor, A.; Martinez-Viviente, E. NMR spectroscopy in coordination supramolecular chemistry: A unique and powerful methodology. Coord. Chem. Rev. 2008, 252, 2314−2345. (b) Maugeri, L.; Asencio-Hernandez, J.; Lebl, T.; Cordes, D. B.; Slawin, A. M. Z.; Delsuc, M. A.; Philp, D. Neutral iodotriazoles as scaffolds for stable halogen-bonded assemblies in solution. Chem. Sci. 2016, 7, 6422− 6428. (57) (a) Macchioni, A.; Ciancaleoni, G.; Zuccaccia, C.; Zuccaccia, D. Determining accurate molecular sizes in solution through NMR diffusion spectroscopy. Chem. Soc. Rev. 2008, 37, 479−489. (b) Bellachioma, G.; Ciancaleoni, G.; Zuccaccia, C.; Zuccaccia, D.; Macchioni, A. NMR investigation of non-covalent aggregation of coordination compounds ranging from dimers and ion pairs up to nano-aggregates. Coord. Chem. Rev. 2008, 252, 2224−2238. (c) Cohen, Y.; Avram, L.; Frish, L. Diffusion NMR spectroscopy in supramolecular and combinatorial chemistry: An old parameter - New insights. Angew. Chem., Int. Ed. 2005, 44, 520−554. (58) Sivchik, V. V.; Grachova, E. V.; Melnikov, A. S.; Smirnov, S. N.; Ivanov, A. Y.; Hirva, P.; Tunik, S. P.; Koshevoy, I. O. Solid-State and Solution Metallophilic Aggregation of a Cationic Pt(NCN)L (+) Cyclometalated Complex. Inorg. Chem. 2016, 55, 3351−3363. (59) (a) Chen, J.; Liu, D.; Butt, N.; Li, C.; Fan, D.; Liu, Y.; Zhang, W. Palladium-Catalyzed Asymmetric Hydrogenation of α-Acyloxy-1-

arylethanones. Angew. Chem., Int. Ed. 2013, 52 (44), 11632−11636. (b) Zheng, S. L.; Zhong, Q.; Jiang, Q.; Mottamal, M.; Zhang, Q.; Zhu, N. J.; Burow, M. E.; Worthylake, R. A.; Wang, G. D. Discovery of a Series of Thiazole Derivatives as Novel Inhibitors of Metastatic Cancer Cell Migration and Invasion. ACS Med. Chem. Lett. 2013, 4, 191−196. (60) Dolomanov, O.; Bourhis, L.; Gildea, R.; Howard, J.; Puschmann, H. OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339−341. (61) Sheldrick, G. M. A short history of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (62) Spek, A. L. PLATON SQUEEZE: a tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71 (1), 9−18. (63) CrysAlisPro, 1.171.36.20; Agilent Technologies, 2012. (64) (a) Wolff, S. K.; Grimwood, D. J.; McKinnon, J. J.; Turner, M. J.; Jayatilaka, D.; Spackman, M. A. CrystalExplorer, Version 3.1; University of Western Australia, 2012. (b) Spackman, M. A.; Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 2009, 11, 19−32. (65) McKinnon, J. J.; Jayatilaka, D.; Spackman, M. A. Towards quantitative analysis of intermolecular interactions with Hirshfeld surfaces. Chem. Commun. 2007, 0, 3814−3816. (66) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215−241. (67) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, N. K.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, J.; Foresman, B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision C.01; Gaussian, Inc., Wallingford, CT, 2010. (68) Andrae, D.; Haussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-Adjusted Abinitio Pseudopotentials for the 2nd and 3rd Row Transition-Elements. Theor. Chim. Acta 1990, 77, 123−141. (69) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378−6396. (70) Lu, T.; Chen, F. W. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580−592.

N

DOI: 10.1021/acs.inorgchem.8b00190 Inorg. Chem. XXXX, XXX, XXX−XXX