The Chalcogen Bond in Crystalline Solids: A World Parallel to

May 13, 2019 - ... small molecules to proteinaceous targets or the organization of molecular ... Being based on the groups of the periodic table, thes...
0 downloads 0 Views 9MB Size
Article Cite This: Acc. Chem. Res. XXXX, XXX, XXX−XXX

pubs.acs.org/accounts

The Chalcogen Bond in Crystalline Solids: A World Parallel to Halogen Bond Patrick Scilabra, Giancarlo Terraneo, and Giuseppe Resnati*

Acc. Chem. Res. Downloaded from pubs.acs.org by UNIV AUTONOMA DE COAHUILA on 05/13/19. For personal use only.

Department of Chemistry, Materials, and Chemical Engineering ’’Giulio Natta’’, Politecnico di Milano, via Mancinelli 7, I-20131 Milano, Italy CONSPECTUS: The distribution of the electron density around covalently bonded atoms is anisotropic, and this determines the presence, on atoms surface, of areas of higher and lower electron density where the electrostatic potential is frequently negative and positive, respectively. The ability of positive areas on atoms to form attractive interactions with electron rich sites became recently the subject of a flurry of papers. The halogen bond (HaB), the attractive interaction formed by halogens with nucleophiles, emerged as a quite common and dependable tool for controlling phenomena as diverse as the binding of small molecules to proteinaceous targets or the organization of molecular functional materials. The mindset developed in relation to the halogen bond prompted the interest in the tendency of elements of groups 13−16 of the periodic table to form analogous attractive interactions with nucleophiles. This Account addresses the chalcogen bond (ChB), the attractive interaction formed by group 16 elements with nucleophiles, by adopting a crystallographic point of view. Structures of organic derivatives are considered where chalcogen atoms form close contacts with nucleophiles in the geometry typical for chalcogen bonds. It is shown how sulfur, selenium, and tellurium can all form chalcogen bonds, the tendency to give rise to close contacts with nucleophiles increasing with the polarizability of the element. Also oxygen, when conveniently substituted, can form ChBs in crystalline solids. Chalcogen bonds can be strong enough to allow for the interaction to function as an effective and robust tool in crystal engineering. It is presented how chalcogen containing heteroaromatics, sulfides, disulfides, and selenium and tellurium analogues as well as some other molecular moieties can afford dependable chalcogen bond based supramolecular synthons. Particular attention is given to chalcogen containing azoles and their derivatives due to the relevance of these moieties in biosystems and molecular materials. It is shown how the interaction pattern around electrophilic chalcogen atoms frequently recalls the pattern around analogous halogen, pnictogen, and tetrel derivatives. For instance, directionalities of chalcogen bonds around sulfur and selenium in some thiazolium and selenazolium derivatives are similar to directionalities of halogen bonds around bromine and iodine in bromonium and iodonium compounds. This gives experimental evidence that similarities in the anisotropic distribution of the electron density in covalently bonded atoms translates in similarities in their recognition and self-assembly behavior. For instance, the analogies in interaction patterns of carbonitrile substituted elements of groups 17, 16, 15, and 14 will be presented. While the extensive experimental and theoretical data available in the literature prove that HaB and ChB form twin supramolecular synthons in the solid, more experimental information has to become available before such a statement can be safely extended to interactions wherein elements of groups 14 and 15 are the electrophiles. It will nevertheless be possible to develop some general heuristic principles for crystal engineering. Being based on the groups of the periodic table, these principles offer the advantage of being systematic.



potential is typically positive or negative at these regions,1 which are involved in the formation of interactions with regions of opposite polarities in surrounding molecules. The position of positive and negative regions is related to the position of the covalent bonds given by corresponding atoms and, consequently, to their atomic orbitals. The periodic change of the orbitals in isolated atoms thus translates into a periodic change in the interactions formed by these atoms in molecules containing

INTRODUCTION The electron density distribution in isolated atoms is anisotropic, and the shape and size of atomic orbitals, the regions of space where electrons are localized, change in a periodic way among different elements. These features are encoded in the periodic table, and the position occupied by elements in the table gives information on the number, type, and geometry of covalent bond(s) typically formed by elements. The electron density distribution is anisotropic also in bonded atoms, which frequently present, at their surface, regions of lower or higher electron density. The electrostatic © XXXX American Chemical Society

Received: January 18, 2019

A

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 1. Classification of attractive interactions formed by elements of groups 1, 2, and 13−18 of the periodic table.

wherein elements of groups 13, 14, 15, and 18 attractively interact with nucleophiles. This Account discusses the ChB, the attractive interaction wherein elements of group 16 are the electrophiles,6 by considering the close contacts between chalcogens and lone pair possessing atoms in crystalline solids. Systems presented here are not a comprehensive collection of the respective short contacts. They have been chosen thanks to their simple and minimal structures or their applicative and functional importance. An early crystallographic contribution to the understanding of interactions involving chalcogen atoms was reported in 1977. It described that electrophiles attack divalent sulfur in R−S−R′ moieties roughly perpendicular to the R−S−R′ plane and nucleophiles do so roughly along the extension of the S−R or S−R′ covalent bonds.12 However, major attention began to be devoted to interactions driven by electrophilic chalcogens only after 2007, when a model of the electronic basis of the ChB was proposed.13,14 Information described in this Account gives meaningful insights to the attractive interactions preferentially formed by chalcogens and complements information made available by computational studies,15 and the use of the ChB in catalysis16 and in recognition17 or transport18 of anions.

Figure 2. Schematic representation of ChB formation.

them and a periodic classification of interactions is enabled (Figure 1). 2 For instance, halogens, chalcogens, and pnictogens, the elements of groups 17, 16, and 15 of the periodic table, typically form one, two, and three covalent bonds. Regions of positive electrostatic potential, commonly named σ-holes,1 are frequently present opposite to these bonds and can attractively interact with electron rich sites (e.g., lone pair possessing atoms, anions, π-bond electrons). Some years ago, we proposed to designate noncovalent interactions between electrophilic and nucleophilic sites by referring to the name of the group of the periodic table to which the electrophilic atom belongs.2 This proposal generalizes a criterion that underlies the IUPAC definition of hydrogen bond (HB)3 and is explicitly adopted by the IUPAC definitions of halogen bond (HaB)4,5 and chalcogen bond (ChB)6 (Figure 2). Following this convention, the terms triel bond (TrB),7 tetrel bond (TtB),8 pnictogen bond (PnB),9,10 and noble gas bond (NgB)11 are more and more commonly used for interactions



GENERAL FEATURES OF ChBs A quite general feature of attractive interactions between nucleophiles and elements of groups 14−18 of the periodic table is that the more polarizable and less electronegative the element is within its group, the greater the tendency of that element is to form close contacts with nucleophiles. Fluorine can hardly work as an electrophile and rarely functions as a B

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 6. Halogen bonded adducts of 18-crown-6 with a λ3bromane(III) derivative (left) and a λ3-iodane(III) analogue (right). Anions are omitted for clarity. Color codes: brown, bromine; light green, fluorine; violet, iodine; other colors as Figures 3.

Figure 3. ChBs formed by a bis-furoxan (left) and a methylsulfonyloxaziridine (right). Here and in all other figures, the ball and stick representations were obtained with Mercury 3.10.3; the normalized contacts (Nc)22 of the ChBs (black dotted lines) and other interactions are given as well as the ChB angles. Hydrogen atoms are omitted for clarity. Color codes: gray, carbon; red, oxygen; blue, nitrogen, green, chlorine.

Figure 7. Examples showing the correlation between electron withdrawing ability of chalcogen substituents and ChB length.

Figure 8. One-dimensional networks formed by ChBs in chloromethyl- (left), trichloromethyl- (middle), and dichlorofluoromethyl-thiocyanate (right).

Figure 4. Two-dimensional networks formed by ChBs in Se(CN)2 (top) and Te(CN)2 (bottom). Color codes: ocher, selenium; dark ocher, tellurium; other colors as Figure 3.

HaB donor;19 the same holds for oxygen, which is a poor ChB donor. Very specific bonding patterns can nevertheless elicit oxygen electrophilicity and close contacts with nucleophiles can be found in certain solid systems20,21 (Figure 3). In O(CN)2, S(CN)2, and Se(CN)2, two σ-holes are present on chalcogen atoms opposite to the cyano groups, and the respective electrostatic potentials are 31.0, 42.7, and 46.9 kcal·mol−1.13 The relative lengths of ChBs present in pure chalcogen dicyanides23,24 (Figure 4) and in their cocrystals, for example, with 18-crown-625,26 (Figure 5), are consistent with the electrostatic potential at respective σ-holes and with expectation from chalcogen polarizability.

Figure 5. Chalcogen bonded adducts of 18-crown-6 with S(CN)2 (left) and with Se(CN)2 (right). Color codes: light ocher, sulfur; other colors as Figures 3 and 4 C

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 9. Infinite chains formed by HaB (brown dotted lines) in cyanogen bromide (top); PnB (violet dotted lines) in cyanodimethyl arsine (middle); TtB (gray dotted lines) in cyanotrimethyl tin (bottom). Color codes: light violet, arsenic; teal, tin; other colors as in Figures 3 and 6.

Figure 11. Infinite chains formed via ChB by SF4 with pyridinium fluoride (top); 4-(dimethylamino)pyridinium hydrogen difluoride (bottom). Cations are omitted for simplicity.

Figure 12. Chalcogen bonded cocrystals formed by SF4 with pyridine (top, left); cyclopentanone (top, right); ethylene glycol dimethyl ether (bottom).

Figure 10. Chalcogen bonded 1D network formed by p-nitrobenzylthiocyanate (top) and 2D net formed by p-nitrobenzyl-tellurocyanate (bottom).

Bromonium and iodonium salts are hypervalent halogen derivatives wherein halogen atoms forms two covalent bonds. Two σ-holes are present opposite to these bonds, and two charge assisted HaBs are typically formed between these holes and anions.27 In the presence of 18-crown-6, the crown oxygen atoms substitute for the anion in getting close to the holes28,29 (Figure 6) and adducts quite similar to those given by chalcogen dicyanides are formed. The covalent bond pattern around chalcogens and halogens in chalcogen dicyanides and halonium salts is similar and translates into a similar pattern of noncovalent interactions. Another general feature of the attractive interactions between nucleophiles and positive σ-holes, ChBs included, is that when the electron withdrawing ability of a residue bound to an atom is increased, the attractive interaction between the positive σ-hole opposite to the residue and a nucleophile

Figure 13. Chalcogen bonded cocrystals formed: by dicyano-difluoroselenium with propionitrile (top, left) and with tetrahydrofuran (top, right); by diiodo-bis(4-methoxyphenyl)-tellurium with dimethyl sulfoxide (bottom).

becomes more likely or shorter, as the electrostatic potential at the σ-hole formed by the residue becomes more positive.30 The intramolecular Se···N and Te···N contacts in 1,8disubstituted naphthalene derivatives (Figure 7) exemplify this trend. In crystalline chloromethyl-thiocyanate34 the assembly of infinite chains is driven by ChBs formed on the elongation of the S−CN bond (Figure 8). Analogous halogen, D

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research pnictogen, and tetrel bonded chains are present in the crystals of cyanogen bromide, cyanodimethyl arsine, and cyanotrimethyl tin (Figure 9) making apparent the similarities among ChBs and HaBs, PnBs, and TtBs.35,36 The group electronegativity of −CCl3 and −CCl2F is greater than that of −CN, and in trichloromethyl- and dichlorofluoromethyl-thiocyanate, the effect of the S···N ChB is on the elongation of the S−CCl3 and S−CCl2F covalent bonds.37,38 The group electronegativity of −CCl2F is greater than of −CCl3, and the ChB opposite to the former residue is shorter than that opposite to the latter. As exemplified by several structures presented in this Account, it is quite general that the presence of fluorine atoms and cyano groups near to a chalcogen atom promotes its ChB donor ability and enables the interaction to become a determining factor for the crystal packing adopted in the solid.39 Consistent with the different electron withdrawing ability of carbon atoms having different hybridization, the tendency to form a ChB opposite to a carbon residue appended to the chalcogen increases in the order C(sp3) < C(sp2) < C(sp). If no strong electron withdrawing substituent is present on an alkyl or aryl substituent of a chalcogen, no ChB is usually formed opposite to the substituent. For instance, in crystalline p-nitrobenzyl-thiocyanate, sulfur functions as a monodentate ChB donor,40 and the interaction is opposite to the S−CN covalent bond (Figure 10). In crystalline p-nitrobenzyltellurocyanate, tellurium, a better ChB donor than sulfur, functions as a bidentate ChB donor, and the interaction opposite to the cyano group is shorter than that opposite to the benzyl residue.41 Chalcogens can afford hypervalent compounds, and also these derivatives can give ChBs. For instance, SF4, a particularly useful reagent in fluoroorganic chemistry, adopts a seesaw conformation and forms ChBs on the extensions of the two equatorial S−F covalent bonds, for example, in cocrystals with anions (Figure 11)42 and lone pair possessing nitrogen43 or oxygen44 atoms (Figure 12). Selenium45 and tellurium46 hypervalent derivatives can afford similar cocrystals with donors of electron density (Figure 13).

Figure 15. ChBs involving a tetrafluorodithietane molecule (left) and side view of tetrafluorodithietane electrostatic potential computed on the 0.001 au molecular surface (right). Sulfur is in the foreground and CF2 groups are at the left and right.

Figure 16. ChBs in trans-difluoro-bis(trifluoromethyl)-1,3-diselenetane (top, left), tetrakis(trifluoromethyl)-1,3-diselenetane (top, middle), tetrakis(acetyl)-1,3-diselenetane (top, right), tetrafluoro1,3-ditelluretane (bottom, left), and tetrakis(trifluoromethyl)-1,3ditelluretane (bottom, right).



SULFIDES, DISULFIDES, AND SELENIUM AND TELLURIUM ANALOGUES Thio-, seleno-, and telluroethers can form ChBs in the solid as exemplified by the 1D net present in 6-aminopenicillanic acid

Figure 14. Chalcogen bonded infinite chain in (2S,5R,6R)-6aminopenicillanic acid.

Figure 17. ChB driven binding of iodide anion in macrocycles containing the 1-methyl-4-methylselanyl-1,2,3-triazolium moiety.

47

(Figure 14). ChBs formed by thioethers are the structuredetermining interactions in 2,2,4,4-tetrafluoro-1,3-dithietane (Figure 15).39 The molecule functions as a tetradentate ChB donor consistent with the presence of four σ-holes on its electrostatic potential surface. These holes are the cumulative effects of two overlapping holes, one at sulfur and one at carbon, on the elongation of a C−S and a S−C bond,

respectively. This overlapping explains the deviation from linearity of the C−S···F ChBs.48 Similar deviations are a distinctive feature of ChBs compared to HaBs and TtBs. Similar ChBs are observed in crystals of conveniently substituted 1,3-diselenetane49−51 and 1,3-ditelluretane52,53 E

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 18. ChBs in L-cystine dimethyl ester dihydrochloride (top) and L-selenocystine dihydrochloride (bottom). Figure 21. Chalcogen bonded infinite chain formed by phenylselenylthiocyanate (top), tetramer formed by 2-(((5-bromo-2methoxyphenyl)tellanyl)sulfanyl)-4,5-dihydro-1H-imidazol-3-ium tetrabromo-(4-methoxyphenyl)-tellurium (bottom, left), and dimer formed by 2-((phenylselanyl)sulfanyl)-4,5-dihydro-1H-imidazol-3ium tetraiodo-phenyl-tellurium (bottom, right).

(Figure 16), steric hindrance around the chalcogen resulting in longer interactions. Macrocycles containing a 1-methyl-4-methylselanyl-1,2,3triazolium moiety have been used for the ChB driven binding of anions in solution and in the solid. Thanks to the strong electron withdrawing ability of the triazolium ring, the selenium electrophilicity is boosted to the point that the charge-assisted C−Se···I− ChB is the only notable short contact17 involving iodide anion (Figure 17), despite the possibility of HBs with the electron deficient triazolium Nmethyl protons as well. The C−Se···I− angle approaches linearity, highlighting the influence of the σ-hole on selenium in governing the geometry of the Se···I− contact. Organodisulfides, -diselenides, and -ditellurides and mixed organodichalcogenides can all show ChBs in respective crystals. These interactions occur in crystalline cystine and selenocystine, two naturally occurring amino acids (Figure 18).54,55 The biological relevance of these compounds and of some others discussed above (Figure 14) suggests that the ChB donor ability of the moieties discussed in this paragraph may have a nonminor biological importance. Dicyanodiselenide functions as a tetradentate ChB donor consistent with the computed presence of two σ-holes on any selenium atom (Figure 19).56 In bis(o-anilinium)diselenide salts, the positively charged anilinium residues on selenium atoms boost the electrophilicity of the diselenide moiety, and up to four close contacts with anions can be formed (Figure 20).57 As expected, in mixed organodichalcogenides, ChBs preferentially involve the more polarizable chalcogen,58 even when the pendant at the less polarizable chalcogen is positively charged (Figure 21).59,60

Figure 19. ChBs formed by a dicyanodiselenide molecule (left) and electrostatic potential of this compound computed on the 0.001 au molecular surface (right). One selenium is in front (top, left). There are two sites of most positive potential (black hemispheres) on each selenium, though only three are visible in this picture (VS,max(C−Se), 40.9 kcal/mol; VS,max(Se−Se), 37.4 kcal/mol).



CHALCOGEN-CONTAINING HETEROAROMATICS An analysis of the Cambridge Structural Database (CSD)61 shows that chalcogen atoms in thiophene and selenophene

Figure 20. Chalcogen bonded 2D network in bis(o-anilinium)diselenide dibromide. F

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 24. Convergent ChBs bind an oxygen in two oligothienoacenes.

Figure 22. ChBs formed by 2-dicyanovinyl-thiophene (A), 5tricyanovinyl-2,2′-bithiophene (B), 5-tricyanovinyl-2,2′:5′,2″-terthiophene (C), 5,5‴-bis(dicyanovinyl)-2,2′:5′,2″:5″,2‴-quaterthiophene (D), 5,5⁗-bis(dicyanovinyl)-2,2′:5′,2″:5″,2‴:5‴,2⁗-quinquethiophene (E).

Figure 25. ChBs in crystalline frentizole (top) and riluzole malonate (bottom).

Figure 23. ChBs formed by 5,5‴-bis(dicyanovinyl)-2,2′:5′,2″:5″,2‴quaterselenophene.

Figure 26. ChBs in crystalline sulfamethizole (top) and acetazolamide (bottom).

derivatives can function as ChB donors.62 Di- or tricyanovinyl substituted oligothiophenes have been studied for their useful electrical and photophysical properties. Figures 22 and 23 show the crystalline structures of parent compounds containing from one63 to five64,65 thiophene66,67 or selenophene68 rings. These structures exemplify how thiophene and selenophene rings are general and reliable ChB donors. Oligothiophenes with convenient covalent bridges between adjacent thiophenes may adopt the s-cis conformation, and

thanks to this preorganization, sulfur atoms can function as converging ChB donors toward neutral and lone pair possessing acceptors. This is the case, for instance, in oligothienoacenes, useful materials for optoelectronic applications (Figure 24).69 In the remaining part of this section, we focus mainly on 1,3benzochalcogenazoles70 due to the relevance of these ring systems as bioactive compounds and molecular materials. G

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 27. Chalcogen bonded chains in 2-(1H-pyrrol-1-yl)-1,3benzothiazole (top), 2-phenyl-1,3-benzoselenazole (middle), and 2isobutyl-1,3-benzotellurazole (bottom).

For instance, frentizole is used clinically in rheumatoid arthritis and systemic lupus erythematosus, and riluzole is used in amyotrophic lateral sclerosis. Sulfur atoms work as monodentate71 and bidentate72 ChB donors in the former and latter compound (Figure 25). In frentizole the intramolecular ChB locks the conformation of the residue appended on C2, and an analogous behavior is quite common whenever an oxygen atom, conveniently positioned on a pendant at C2 or C5 of the chalcogenazole, can form a fivemembered ring via an intramolecular ChB. Sulfamethizole73 is an antibiotic drug, and acetazolamide74 is used to treat glaucoma, and they both show the intramolecular ChB discussed above (Figure 26). Alternatively, the ChB acceptor can be the azole nitrogen, and Figure 27 presents three chains assembled thanks to such intermolecular ChBs.75−77 These examples prove that this chalcogen···N supramolecular synthon is particularly reliable as it remains unchanged while the chalcogen atom and the substituent at C2 change. The robustness of this synthon is further proven by its occurrence in other azoles, for example, in 1,2,5-thiadiazole derivatives and their selena- and telluraanalogues (Figure 28). Sulfur, a moderate ChB donor, acts in benzo[c][1,2,5]thiadiazole as a monodentate site;78 selenium acts in benzo[c][1,2,5]selenadiazole as a monodentate or bidentate donor as a function of the presence of electron withdrawing groups;79,80 tellurium, an efficient ChB donor, acts in benzo[c][1,2,5]telluradiazole as a bidentate site also when nonactivated.81 The supramolecular synthon shown by benzotelluradiazole is robust enough to secure the formation of capsules via self-assembly of benzotelluradiazole units conveniently preorganized on resorcin[4]arene cavitands (Figure 29).82 Isotellurazole N-oxides have been employed for the selfassembly of macrocyles via Te···O−N ChBs.83 Chalcogenazolium derivatives are, in general, much better donors than neutral chalcogenazoles. The C−S···OH2 ChBs in the two derivatives of Figure 30 adopt very similar arrangements, but the interaction is much longer in the benzothia-

Figure 28. Chalcogen bonded chains in benzo[c][1,2,5]thiadiazole (A), benzo[c][1,2,5]selenadiazole (B), 6,7,8,9-tetrafluorobenzo[c][1,2,5]-selenadiazole (C), and benzo[c][1,2,5]telluradiazole (D).

zole84 than in the benzothiazolium85 derivative, the latter compound forming also a second ChB, again shorter than the ChB in the neutral analogue. Cyanine dyes containing benzochalcogenazolium moieties find useful applications spanning photographic emulsions and contrast agents. Consistent with the polarizability of respective chalcogens, no ChB has been observed in benzoxazolium dyes, while benzothiazolium (Figure 31)86,87 and benzoselenazolium (Figure 32)88 analogues frequently give rise to one charge assisted ChB with the anion on the extension of the chalcogen−C(N) bond, just where modeling localizes a remarkably positive σ-hole. Chalcogen atoms in chalcogenazolium derivatives and halogen atoms in halonium salts are both divalent and both possess two σ-holes.36 Topologies of adducts formed by the two classes of derivatives can be quite different, spanning discrete adducts and infinite chains,89 and typically parallel each others (Figure 33), once again confirming similarities between ChB and HaB. H

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research

Figure 32. ChB in 3-ethyl-5-methoxy-2-[(E)-(1-ethyl-6-methyl2(1H)-quinolylidene)methyl]-benzoselenazolium p-toluenesulfonate (left) and computed electrostatic potential on the 0.001 au molecular surface of 3-ethyl-5-methoxy-2-[(E)-(1-ethyl-6-methoxyl-2(1H)-quinolylidene)-methyl]benzoselenazolium cation. The 5-methoxy pendant on the benzoselenazole ring is at the top left of the picture. Gray dots indicate the atom positions. Black hemispheres show the locations of the maximum positive potentials (Vs,max, 83 kcal/mol for both hemispheres). Figure 29. Chalcogen bonded capsule assembled via N−Te···N ChBs. Atoms in the background are in darker colors and included benzene rings in lighter color.

Figure 33. Square tetramers formed by ChBs in 3-methylbenzoselenazolium iodide (left) and by HaBs in diphenyliodonium iodide (right).

Figure 30. Hydrates formed by ChBs in 2-amino-6-nitro-benzothiazole (left) and infinite chain of 2-amino-6-sulfamoyl-benzothiazolium (right, the anion is omitted for the sake of simplicity).

W. Hofmann, who described the adducts between SO2 and amines.90 Analogous complexes of SO2 with phenols were reported in 1882,91 with anilines in 1891,92 and with other donors of electron density soon after.93 Crystallographic confirmation is now available that in adducts between SO2 and electrophiles, the latter species approach sulfur on one or both molecular faces in directions orthogonal to the molecular plane.94,95 Modeling showed that sulfur regions involved in these interactions have a positive electrostatic potential (32.9 kcal·mol−1). These regions are named π-holes96 to stress that they are perpendicular to the molecular framework, while other holes described in this paper are named σ-holes as they are along the extension of a σ covalent bond at the chalcogen. The attractive nature of ChBs is mainly due to electrostatic effects, but polarization, charge-transfer, and dispersion contributions all play a role in determining interpenetration of van der Waals volumes, and the relative relevance of this role varies from one case to the other.97 This Account proves that chalcogen atoms frequently behave as electrophiles and ChBs may be strong enough to control the conformation and packing of molecules in the solid. It may be expected that in the coming years the ChB will further demonstrate its impact in all areas where recognition and self-assembly play a role, once again performing in analogy with HaBs.

Figure 31. Dimers formed by ChBs in 1-methyl-2-((3-methyl-1,3benzothiazol-2(3H)-ylidene)methyl)quinolinium iodide (left) and 3ethyl-2-(5-(3-ethyl-3H-benzothiazol-2-ylidene)-penta-1,3-dienyl)benzothiazolium iodide (right).



PERSPECTIVES Crystallographic analyses of close contacts wherein elements of groups 14,8 15,10 and 1730 are interacting with nucleophiles have afforded valuable information on key features of TtB, PnB, and HaB. Here we have presented an analogous analysis of close contacts involving group 16 elements as electrophiles, and we have drawn information on some general features of the ChB. Interest in recognition phenomena driven by electrophilic chalcogens boomed15−18,82 after the seminal paper of J. S. Murray et al. in 2007,13 but complexes that we now recognize as formed under ChB control have long been known. To the best of our knowledge, the earliest complexes wherein a chalcogen atom is the electrophile were reported in 1843 by A.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. I

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research ORCID

(11) Hou, C.; Wang, X.; Botana, J.; Miao, M. Noble Gas Bond and the Behaviour of XeO3 under Pressure. Phys. Chem. Chem. Phys. 2017, 19, 27463−27467. (12) Rosenfield, R. E.; Parthasarathy, R.; Dunitz, J. D. Directional Preferences of Nonbonded Atomic Contacts with Divalent Sulfur. 1. Electrophiles and Nucleophiles. J. Am. Chem. Soc. 1977, 99, 4860− 4862. (13) Murray, J. S.; Lane, P.; Clark, T.; Politzer, P. σ-Hole Bonding: Molecules Containing Group VI Atoms. J. Mol. Model. 2007, 13, 1033−1038. (14) Soon after the ChB rationalization, the same authors, proposed similar models for the PnBs and TtBs effecting an analogous increase of interest in these interactions. Murray, J. S.; Lane, P.; Politzer, P. A Predicted New Type of Directional Noncovalent Interaction. Int. J. Quantum Chem. 2007, 107, 2286−2292. Murray, J. S.; Lane, P.; Politzer, P. Expansion of the σ-hole concept. J. Mol. Model. 2009, 15, 723−729. (15) Pascoe, D. J.; Ling, K. B.; Cockroft, S. L. The Origin of Chalcogen-Bonding Interactions. J. Am. Chem. Soc. 2017, 139, 15160−15167. (16) 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. (17) 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. (18) 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. (19) Metrangolo, P.; Murray, J. S.; Pilati, T.; Politzer, P.; Resnati, G.; Terraneo, G. Fluorine-Centered Halogen Bonding: A Factor in Recognition Phenomena and Reactivity. Cryst. Growth Des. 2011, 11, 4238−4246. (20) Sitzmann, M. E.; Bichay, M.; Fronabarger, J. W.; Williams, M. D.; Sanborn, W. B.; Gilardi, R. Hydroxynitrobenzodifuroxan and Its Salts. J. Heterocycl. Chem. 2005, 42, 1117−1125. (21) Forni, A.; Moretti, I.; Torre, G.; Bruckner, S.; Malpezzi, L. XRay Structure and Stereochemical Properties of (S,S)-(−)-2Methylsulphonyl-3-Phenyloxaziridine and of (S,S)-(−)-2-Methylsulphonyl-3-(2-Chloro-5-Nitrophenyl)Oxaziridine. J. Chem. Soc., Perkin Trans. 2 1987, 2, 699−704. (22) A normalized contact (Nc) is defined as the ratio of the experimental separation of the interacting atoms to the sum of their respective van der Waals radii (or Pauling ionic radii for charged species). Bondi, A. van der Waals Volumes and Radii. J. Phys. Chem. 1964, 68, 441−451. Shannon, R. D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976, 32, 751−767. Nc is a useful indicator because it allows distances between different interacting partners to be compared in a more reliable way than when absolute values of separations are used. A value of Nc < 1 is usually assumed to designate an attractive interaction. (23) Klapotke, T. M.; Krumm, B.; Gálvez-Ruiz, J. C.; Noth, H.; Schwab, I. Experimental and Theoretical Studies of Homoleptic Tellurium Cyanides Te(CN)x: Crystal Structure of Te(CN)2. Eur. J. Inorg. Chem. 2004, 2004, 4764−4769. (24) Klapotke, T. M.; Krumm, B.; Scherr, M. Homoleptic Selenium Cyanides: Attempted Preparation of Se(CN)4 and Redetermination of the Crystal Structure of Se(CN)2. Inorg. Chem. 2008, 47, 7025− 7028. (25) Ilyukhin, A.; Petrosyants, S. CSD Communication, 2011; CSD Refcode GOJZUC. (26) Fritz, S.; Ehm, C.; Lentz, D. Structure and Chemistry of SeFx(CN)4‑x Compounds. Inorg. Chem. 2015, 54, 5220−5231.

Patrick Scilabra: 0000-0003-1972-620X Giancarlo Terraneo: 0000-0002-1225-2577 Giuseppe Resnati: 0000-0002-0797-9296 Notes

The authors declare no competing financial interest. Biographies Patrick Scilabra (b. 1991) is a doctoral candidate at Politecnico di Milano, Department of Chemistry, Materials, and Chemical Engineering. In 2016, he joined the Laboratory of Nanostructured Fluorinated Materials (NFMLab). His research interests are in interactions involving group 14−16 elements as electrophilic sites. Giancarlo Terraneo (b. 1975) received his Ph.D. in chemistry from Università degli Studi di Milano in 2006 and was appointed as associate professor at Politecnico di Milano in 2014. His research interests are in crystal engineering, supramolecular chemistry, and crystallography. Giuseppe Resnati (b. 1955) is full professor at Politecnico di Milano. His research interests are in fluorine chemistry and supramolecular chemistry. He has served as coordinator of IUPAC projects that defined the halogen and chalcogen bonds. He is coordinating the UNESCO-UNITWIN network GREENOMIcS.



REFERENCES

(1) The electrostatic potential frequently but not always follows the electronic density, since it also reflects the effect of the nuclei: Politzer, P.; Murray, J. S. σ-Hole Interactions: Perspectives and Misconceptions. Crystals 2017, 7, 212−226. (2) Cavallo, G.; Metrangolo, P.; Pilati, T.; Resnati, G.; Terraneo, G. Naming Interactions from the Electrophilic Site. Cryst. Growth Des. 2014, 14, 2697−2702. (3) 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. (4) Desiraju, G. R.; Ho, P. S.; Kloo, L.; Legon, A. C.; Marquardt, R.; Metrangolo, P.; Politzer, P.; Resnati, G.; Rissanen, K. Definition of the Halogen Bond (IUPAC Recommendations 2013). Pure Appl. Chem. 2013, 85, 1711−1713. (5) The acronym XB is typically used for the halogen bond, while three-letter coding is adopted for triel, chalcogen, pnictogen, and tetrel bonds. The HaB acronym is employed for the halogen bond for the sake of consistency. (6) Aakeroy, C. B.; Bryce, D. L.; Desiraju, G. R.; Frontera, A.; Legon, A. C.; Nicotra, F.; Rissanen, K.; Scheiner, S.; Terraneo, G.; Metrangolo, P.; Resnati, G. Definition of the chalcogen bond (IUPAC Recommendation 2019). Pure Appl. Chem. 2019, Submitted for publication. (7) Escudero-Adán, E. C.; Bauzá, A.; Lecomte, C.; Frontera, A.; Ballester, P. Boron Triel Bonding: A Weak Electrostatic Interaction Lacking Electron-Density Descriptors. Phys. Chem. Chem. Phys. 2018, 20, 24192−24200. (8) Scilabra, P.; Kumar, V.; Ursini, M.; Resnati, G. Close Contacts Involving Germanium and Tin in Crystal Structures: Experimental Evidence of Tetrel Bonds. J. Mol. Model. 2018, 24, 37−55. (9) Scheiner, S. The Pnicogen Bond: Its Relation to Hydrogen, Halogen, and Other Noncovalent Bonds. Acc. Chem. Res. 2013, 46, 280−288. (10) Scilabra, P.; Terraneo, G.; Resnati, G. Fluorinated Elements of Group 15 as Pnictogen Bond Donor Sites. J. Fluorine Chem. 2017, 203, 62−74. J

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research (27) Catalano, L.; Cavallo, G.; Metrangolo, P.; Resnati, G.; Terraneo, G. Halogen Bonding in Hypervalent Iodine Compounds. Top. Curr. Chem. 2016, 373, 289−310. (28) Ochiai, M.; Nishi, Y.; Goto, S.; Shiro, M.; Frohn, H. J. Synthesis, Structure, and Reaction of 1-Alkynyl(aryl)-λ3-bromanes. J. Am. Chem. Soc. 2003, 125, 15304−15305. (29) Ochiai, M.; Suefuji, T.; Miyamoto, K.; Tada, N.; Goto, S.; Shiro, M.; Sakamoto, S.; Yamaguchi, K. Secondary Hypervalent I(III)···O Interactions: Synthesis and Structure of Hypervalent Complexes of Diphenyl-λ3-iodanes with 18-Crown-6. J. Am. Chem. Soc. 2003, 125, 769−773. (30) Metrangolo, P.; Resnati, G. Halogen Bonding: A Paradigm in Supramolecular Chemistry. Chem. - Eur. J. 2001, 7, 2511−2519. (31) Panda, A.; Mugesh, G.; Singh, H. B.; Butcher, R. J. Synthesis, Structure, and Reactivity of Organochalcogen (Se, Te) Compounds Derived from 1-(N,N-Dimethylamino)naphthalene and N,N-Dimethylbenzylamine. Organometallics 1999, 18, 1986−1993. (32) Rakesh, P.; Singh, H. B.; Butcher, R. J. Synthesis of Selenenium Ions: Isolation of Highly Conjugated, pH-Sensitive 4,4′-Bis(methylimino)-1,1′-binaphthylene-5-diselenenium(II) Triflate. Organometallics 2013, 32, 7275−7282. (33) Menon, S. C.; Singh, H. B.; Jasinski, J. M.; Jasinski, J. P.; Butcher, R. J. Intramolecularly Coordinated Low-Valent Organotellurium Complexes Derived from 1-(Dimethylamino)naphthalene. Organometallics 1996, 15, 1707−1712. (34) Martínez, Y. B.; Rodríguez Pirani, L. S.; Erben, M. F.; Boese, R.; Reuter, C. G.; Vishnevskiy, Y. V.; Mitzel, N. W.; Della Védova, C. O. Gas and Crystal Structures of CCl2FSCN. J. Mol. Struct. 2017, 1132, 175−180. (35) Murray, J. S.; Resnati, G.; Politzer, P. Close Contacts and Noncovalent Interactions in Crystals. Faraday Discuss. 2017, 203, 113−130. (36) Cavallo, G.; Murray, S.; Politzer, P.; Pilati, T.; Ursini, M.; Resnati, G. Halogen bonding in hypervalent iodine and bromine derivatives: halonium salts. IUCrJ 2017, 4, 411−419. (37) Berrueta Martínez, Y.; Rodríguez Pirani, L. S.; Erben, M. F.; Boese, R.; Reuter, C. G.; Vishnevskiy, Y. V.; Mitzel, N. W.; Della Védova, C. O. Structures of Trichloromethyl Thiocyanate, CCl3SCN, in Gaseous and Crystalline State. ChemPhysChem 2016, 17, 1463− 1467. (38) Berrueta Martínez, Y.; Rodríguez Pirani, L. S.; Erben, M. F.; Reuter, C. G.; Vishnevskiy, Y. V.; Stammler, H. G.; Mitzel, N. W.; Della Védova, C. O. The Structure of Chloromethyl Thiocyanate, CH2ClSCN, in Gas and Crystalline Phases. Phys. Chem. Chem. Phys. 2015, 17, 15805−15812. (39) Nayak, S. K.; Kumar, V.; Murray, J. S.; Politzer, P.; Terraneo, G.; Pilati, T.; Metrangolo, P.; Resnati, G. Fluorination Promotes Chalcogen Bonding in Crystalline Solids. CrystEngComm 2017, 19, 4955−4959. (40) Maartmann-Moe, K.; Sanderud, K. A.; Songstad, J. The Crystal Structure of 4-Nitrobenzyl Tellurocyanate, 4-Nitrobenzyl Selenocyanate, 4-Nitrobenzyl Thiocyanate and Benzyl Selenocyanate. Acta Chem. Scand. 1984, 38a, 187−200. (41) Maartmann-Moe, K.; Sanderud, K. A.; Songstad, J. 4Nitrobenzyl Tellurocyanate. Preparation and Crystal Structure. Acta Chem. Scand. 1981, 35a, 151−153. (42) Goettel, J. T.; Kostiuk, N.; Gerken, M. Interactions between SF4 and Fluoride: A Crystallographic Study of Solvolysis Products of SF4·Nitrogen-Base Adducts by HF. Inorg. Chem. 2016, 55, 7126− 7134. (43) Chaudhary, P.; Goettel, J. T.; Mercier, H. P. A.; SowlatiHashjin, S.; Hazendonk, P.; Gerken, M. Lewis Acid Behavior of SF4: Synthesis, Characterization, and Computational Study of Adducts of SF4 with Pyridine and Pyridine Derivatives. Chem. - Eur. J. 2015, 21, 6247−6256. (44) Goettel, J. T.; Gerken, M. Synthesis and Characterization of Adducts between SF4 and Oxygen Bases: Examples of O···S(IV) Chalcogen Bonding. Inorg. Chem. 2016, 55, 12441−12450.

(45) Fritz, S.; Ehm, C.; Lentz, D. Structure and Chemistry of SeFx(CN)4‑x Compounds. Inorg. Chem. 2015, 54, 5220−5231. (46) Farran, J.; Alvarez-Larena, A.; Piniella, J. F.; Capparelli, M. V.; Friese, K. Twinning by merohedry in bis(4-methoxyphenyl)tellurium(IV) diiodide dimethyl sulfoxide hemisolvate. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2008, 64, o257−o260. (47) Saouane, S.; Buth, G.; Fabbiani, F. P. A. Crystal Structure and Packing Energy Calculations of (+)-6-Aminopenicillanic Acid. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2013, 69, 1238−1242. (48) Politzer, P.; Murray, J.; Clark, T.; Resnati, G. The σ-Hole Revisited. Phys. Chem. Chem. Phys. 2017, 19, 32166−32178. (49) Boese, R.; Haas, A.; Spehr, M. Preparation of (Perfluoroalkyl)Halogeno-1,3-Diselenetanes from the Corresponding Selenocarbonyl Fluorides and Reactions with Boron-Trichloride or Arsenic Pentafluoride. Chem. Ber. 1991, 124, 51−61. (50) Rabe, G.; Keller, K.; Roesky, H. W.; Lagow, R. J.; Pauer, F.; Stalke, D. Structure of 2,2,4,4-tetrakis(trifluoromethyl)-1,3-diselenane. Z. Naturforsch., B: J. Chem. Sci. 1991, 46, 157−160. (51) Adrien, R. J.; Gable, R. W.; Hoskins, B. F.; Dakternieks, D. Crystal and molecular structures of the acetylacetonate derivatives [Se(CH3C(O)CC(O)CH3)]2 and [PhSe]2[CH3C(O)CC(O)CH3]. J. Organomet. Chem. 1989, 359, 33−39. (52) Boese, R.; Haas, A.; Limberg, C. Synthesis and Characterization of 2,2,4,4-Tetrafluoro-1,3-Ditelluretane and −1-Selena-3-Telluretane via the Intermediate Difluorotelluroketone. J. Chem. Soc., Chem. Commun. 1991, 19, 1378−1379. (53) Baum, M.; Beck, J.; Haas, A.; Herrendorf, W.; Monse, C. Perfluoroalkyltellurocarbonyl fluorides, their cyclic dimers and perfluoroalkanetellurenyl iodides: preparation and reactivity. J. Chem. Soc., Dalton Trans. 2000, 11−15. (54) Vijayalakshmi, B. K.; Srinivasan, R. Crystal and molecular structure of L-cystine dimethyl ester dihydrochloride monohydrate. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1975, 31, 993−998. (55) Gorbitz, C. H.; Levchenko, V.; Semjonovs, J.; Sharif, M. Y. Crystal structure of seleno-L-cystine dihydrochloride. Acta Crystallogr., Sect. E: Cryst. Commun. 2015, 71, 726−729. (56) Burchell, C. J.; Kilian, P.; Slawin, A. M. Z.; Woollins, J. D.; Tersago, K.; Van Alsenoy, C.; Blockhuys, F. E2(CN)2(E = S, Se) and Related Compounds. Inorg. Chem. 2006, 45, 710−716. (57) Scilabra, P.; Murray, J. S.; Terraneo, G.; Resnati, G. Chalcogen Bonds in Crystals of Bis(o-Anilinium)Diselenide Salts. Cryst. Growth Des. 2019, 19, 1149−1154. (58) Barnes, N. A.; Godfrey, S. M.; Halton, R. T. A.; Mushtaq, I.; Parsons, S.; Pritchard, R. G.; Sadler, M. A. Comparison of the SolidState Structures of a Series of Phenylseleno-Halogen and Pseudohalogen Compounds, PhSeX (X = Cl, CN, SCN). Polyhedron 2007, 26, 1053−1060. (59) Casagrande, G. A.; Raminelli, C.; Lang, E. S.; Lemos, S. D. S. A Novel Organotellurium Halide with Tellurium Presenting Mixed Oxidation States: Synthesis and Structural Characterization. Inorg. Chim. Acta 2011, 365, 492−495. (60) Manzoni de Oliveira, G.; Antônio Casagrande, G.; Schulz Lang, E. S.; Marques Muzzi, R. M.; Lemos, S. de S. New examples of mixed organochalcogene compounds: Synthesis, structural features and spectroscopic data of [PhSe(etu)][PhTeX4] (Ph = phenyl; etu = ethylenethiourea; X = Br, I). J. Organomet. Chem. 2009, 694, 2463− 2466. (61) Groom, C. R.; Bruno, I. J.; Lightfoot, M. P.; Ward, S. C. Cambridge Structural Database. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 2016, B72, 171−179. (62) Murray, J. S.; Lane, P.; Politzer, P. Simultaneous σ-Hole and Hydrogen Bonding by Sulfur- And Selenium-Containing Heterocycles. Int. J. Quantum Chem. 2008, 108, 2770−2781. (63) Mukherjee, A. K.; Mukherjee (née Mondal), M.; De, A.; Bhattacharyya, S. P. 2-Thienylmethylenemalononitrile, C8H4N2S. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1984, 40, 991−992. (64) Fitzner, R.; Mena-Osteritz, E.; Mishra, A.; Schulz, G.; Reinold, E.; Weil, M.; Körner, C.; Ziehlke, H.; Elschner, C.; Leo, K.; Riede, M.; K

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX

Article

Accounts of Chemical Research Pfeiffer, M.; Uhrich, C.; Bäuerle, P. Correlation of π-Conjugated Oligomer Structure with Film Morphology and Organic Solar Cell Performance. J. Am. Chem. Soc. 2012, 134, 11064−11067. (65) Fitzner, R.; Elschner, C.; Weil, M.; Uhrich, C.; Korner, C.; Riede, M.; Leo, K.; Pfeiffer, M.; Reinold, E.; Mena-Osteritz, E.; Bauerle, P. Interrelation between Crystal Packing and Small-Molecule Organic Solar Cell Performance. Adv. Mater. 2012, 24, 675−680. (66) Bader, M. M.; Custelcean, R.; Ward, M. D. TricyanovinylSubstituted Oligothiophenes. Chem. Mater. 2003, 15, 616−618. (67) Pham, P.-T. T.; Bader, M. M. Inter- and Intramolecular Interactions in Some Bromo- and Tricyanovinyl-Substituted Thiophenes and Ethylenedioxythiophenes. Cryst. Growth Des. 2014, 14, 916−922. (68) Haid, S.; Mishra, A.; Weil, M.; Uhrich, C.; Pfeiffer, M.; Bäuerle, P. Synthesis and Structure-Property Correlations of DicyanovinylSubstituted Oligoselenophenes and Their Application in Organic Solar Cells. Adv. Funct. Mater. 2012, 22, 4322−4333. (69) Suzuki, Y.; Okamoto, T.; Wakamiya, A.; Yamaguchi, S. Electronic Modulation of Fused Oligothiophenes by Chemical Oxidation. Org. Lett. 2008, 10, 3393−3396. (70) Biot, N.; Bonifazi, D. Programming Recognition Arrays through Double Chalcogen-Bonding Interactions. Chem. - Eur. J. 2018, 24, 5439−5443. (71) Rooth, W.; Srikrishnan, T. Crystal Structure and Conformation of Frentizole, [1-(6-Methoxy-2-Benzothiazolyl)-3-Phenylurea], an Antiviral Agent and an Immunosuppressive Drug. J. Chem. Crystallogr. 1999, 29, 1187−1192. (72) Mondal, P. K.; Rao, V.; Mittapalli, S.; Chopra, D. Exploring Solid State Diversity and Solution Characteristics in a FluorineContaining Drug Riluzole. Cryst. Growth Des. 2017, 17, 1938−1946. (73) Thomas, S. P.; Veccham, S. P. K. P.; Farrugia, L. J.; Guru Row, T. N. “Conformational Simulation” of Sulfamethizole by Molecular Complexation and Insights from Charge Density Analysis: Role of Intramolecular S···O Chalcogen Bonding. Cryst. Growth Des. 2015, 15, 2110−2118. (74) Thomas, S. P.; Jayatilaka, D.; Guru Row, T. N. S···O Chalcogen Bonding in Sulfa Drugs: Insights from Multipole Charge Density and X-Ray Wavefunction of Acetazolamide. Phys. Chem. Chem. Phys. 2015, 17, 25411−25420. (75) Kremer, A.; Fermi, A.; Biot, N.; Wouters, J.; Bonifazi, D. Supramolecular Wiring of Benzo-1,3-chalcogenazoles through Programmed Chalcogen Bonding Interactions. Chem. - Eur. J. 2016, 22, 5665−5675. (76) Hedidi, M.; Bentabed-Ababsa, G.; Derdour, A.; Roisnel, T.; Dorcet, V.; Chevallier, F.; Picot, L.; Thiery, V.; Mongin, V. Synthesis of C,N’-linked bis-heterocycles using a deprotometalation-iodinationN-arylation sequence and evaluation of their antiproliferative activity in melanoma cells. Bioorg. Med. Chem. 2014, 22, 3498−3507. (77) Cozzolino, A. F.; Vargas-Baca, I. Parametrization of a Force Field for Te-N Secondary Bonding Interactions and Its Application in the Design of Supramolecular Structures Based on Heterocyclic Building Blocks. Cryst. Growth Des. 2011, 11, 668−677. (78) Suzuki, T.; Tsuji, T.; Okubo, T.; Okada, A.; Obana, Y.; Fukushima, T.; Miyashi, T.; Yamashita, Y. Preparation, Structure, and Amphoteric Redox Properties of p-Phenylenediamine-Type Dyes Fused with a Chalcogenadiazole Unit. J. Org. Chem. 2001, 66, 8954− 8960. (79) Gomes, A. C.; Biswas, G.; Banerjee, A.; Duax, W. L. Structure of a Planar Organic Compound: 2,1,3-Benzoselenadiazole (Piaselenole). Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1989, 45, 73− 75. (80) Mikhailovskaya, T. F.; Makarov, A. G.; Selikhova, N. Yu.; Makarov, A. Yu.; Pritchina, E. A.; Bagryanskaya, I. Yu.; Vorontsova, E. V.; Ivanov, I. D.; Tikhova, V. D.; Gritsan, N. P.; Slizhov, Y. G.; Zibarev, A. V. Carbocyclic functionalization of quinoxalines, their chalcogen congeners 2,1,3-benzothia/selenadiazoles, and related 1,2diaminobenzenes based on nucleophilic substitution of fluorine. J. Fluorine Chem. 2016, 183, 44−58.

(81) Cozzolino, A. F.; Britten, J. F.; Vargas-Baca, I. The Effect of Steric Hindrance on the Association of Telluradiazoles through Te-N Secondary Bonding Interactions. Cryst. Growth Des. 2006, 6, 181− 186. (82) Riwar, L.-J.; Trapp, N.; Root, K.; Zenobi, R.; Diederich, F. Supramolecular Capsules: Strong versus Weak Chalcogen Bonding. Angew. Chem., Int. Ed. 2018, 57, 17259−17264. (83) Ho, P. C.; Szydlowski, P.; Sinclair, J.; Elder, P. J. W.; Kübel, 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. (84) Lynch, D. E. 2-Amino-6-Nitrobenzo-1,3-Thiazole Hydrate. Acta Crystallogr., Sect. E: Struct. Rep. Online 2002, 58, o1139−o1141. ̇ (85) Alkaya, Z. A.; Ilkimen, H.; Yenikaya, C.; Tunca, E.; Bülbül, M.; Tunç, T.; Sarı, M. Synthesis and Characterization of Cu(II) Complexes of 2-Amino-6-Sulfamoylbenzothiazole and Their Inhibition Studies on Carbonic Anhydrase Isoenzymes. Polyhedron 2018, 151, 199−205. (86) Guan, L.; Anyang, G.; Song, L. Y.; Yan, M.; Gao, D.; Zhang, X.; Li, B.; Wang, L. Nonplanar Monocyanines: Meso-Substituted Thiazole Orange with High Photostability and Their Synthetic Strategy as well as a Cell Association Study. J. Org. Chem. 2016, 81, 6303−6313. (87) Scilabra, P.; Konidaris, K. F.; Terraneo, G.; Resnati, G. Thiazoliums and selenazoliums as Chalcogen Bond donors in crystals. Acta Crystallogr., Sect. A: Found. Adv. 2018, A74, e108. (88) Konidaris, K. F.; Pilati, T.; Terraneo, G.; Politzer, P.; Murray, J. S.; Scilabra, P.; Resnati, G. Cyanine Dyes: Synergistic Action of Hydrogen, Halogen and Chalcogen Bonds Allows Discrete I42‑ Anions in Crystals. New J. Chem. 2018, 42, 10463−10466. (89) Nesmeyanov, A. N.; Khotsyanova, T. L.; Saatsazov, V. V.; Tolstaya, T. P.; Isaeva, L. S. Dokl. Akad. Nauk SSSR 1974, 218, 140− 148. (90) Hofmann, A. W. Chemische Untersuchung der organischen Basen im Steinkohlen-Theeröl. Ann. Chem. Pharm. 1843, 47, 37−87. (91) Meyer, L.; Kolbe, H. Ueber eine Verbindung des Schwefligsäureanhydrids mit Phenol. J. Prakt. Chem. 1882, 25, 462−464. (92) Michaelis, A. Ueber die Thionylamine. Ber. Dtsch. Chem. Ges. 1891, 24, 745−757. (93) André, G. Action des acides sulfureux et sulfhydrique sur la pyridine. Compt. Rend. 1900, 130, 1714−1716. (94) Kumar, A.; McGrady, G. S.; Passmore, J.; Grein, F.; Decken, A. Reversible SO2 uptake by Tetraalkylammonium Halides: Energetics and Structural Aspects of Adduct Formation between SO2 and Halide Ions. Z. Anorg. Allg. Chem. 2012, 638, 744−753. (95) Reuter, K.; Rudel, S. S.; Buchner, M. R.; Kraus, F.; von Hänisch, C. Crown Ether Complexes of Alkali-Metal Chlorides from SO2. Chem. - Eur. J. 2017, 23, 9607−9617. (96) Murray, J. S.; Lane, P.; Clark, T.; Riley, K. E.; Politzer, P. σHoles, π-Holes and Electrostatically-Driven Interactions. J. Mol. Model. 2012, 18, 541−548. (97) Cozzolino, A. F.; Vargas-Baca, I.; Mansour, S.; Mahmoudkhani, A. H. The Nature of the Supramolecular Association of 1,2,5Chalcogenadiazoles. J. Am. Chem. Soc. 2005, 127, 3184−3190.

L

DOI: 10.1021/acs.accounts.9b00037 Acc. Chem. Res. XXXX, XXX, XXX−XXX