Formation of Metal-Based 21- and 22-Membered Macrocycles from

Jun 10, 2019 - Investigaciones. Químicas,. Instituto. de. Investigación. en. Ciencias. Básicas. y. Aplicadas,. Universidad. Autónoma. del. Estado...
0 downloads 0 Views 4MB Size
Article Cite This: Organometallics 2019, 38, 2443−2460

pubs.acs.org/Organometallics

Formation of Metal-Based 21- and 22-Membered Macrocycles from Dinuclear Organotin Tectons and Ditopic Organic Ligands Carrying Carboxylate or Dithiocarbamate Groups Irán Rojas-León,†,‡ Hazem Alnasr,‡ Klaus Jurkschat,*,‡ María G. Vasquez-Ríos,† Gelen Gómez-Jaimes,† Herbert Höpfl,*,† Irán F. Hernández-Ahuactzí,§ and Rosa Santillan∥

Downloaded via IDAHO STATE UNIV on July 18, 2019 at 19:08:19 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



Centro de Investigaciones Químicas, Instituto de Investigación en Ciencias Básicas y Aplicadas, Universidad Autónoma del Estado de Morelos, Av. Universidad 1001, Chamilpa, Cuernavaca 62209, Morelos, México ‡ Fakultät für Chemie und Chemische Biologie, Technische Universität Dortmund, Dortmund 44221, Germany § Centro Universitario de Tonalá, Universidad de Guadalajara, Av. Nuevo Periférico 555, Ejido San José Tatepozco, Tonalá 45425, Jalisco, México ∥ Departamento de Química, Centro de Investigación y de Estudios Avanzados del IPN, Av. Instituto Politécnico Nacional 2508, Ciudad de México, D. F. 07360, México S Supporting Information *

ABSTRACT: Four dinuclear organotin halides of composition X n Ph ( 3 − n ) SnCH 2 Si(Me) 2 −C 1 2 H 8 −Si(Me) 2 CH 2 SnPh(3−n)Xn (X = Cl, I; n = 1, 2) were prepared and combined in 1:1 stoichiometric reactions with potassium 2,5pyridinedicarboxylate, 3,5-pyridinedicarboxylate, and piperazine bis-dithiocarbamate, respectively. The reactions yielded a total of five [1 + 1] aggregates with either 21- or 22membered macrocyclic structures that were fully characterized by elemental analysis, mass spectrometry, IR and NMR (1H, 13 C, 29Si, and 119Sn) spectroscopy, and, in three cases, additionally by single-crystal X-ray diffraction analysis. In solution, the macrocycles exhibit conformational and configurational equilibria being fast on the NMR time scale, which, for one of the macrocycles, were closer examined by variable temperature NMR spectroscopy and DFT calculations.



coordination polymers5 is increasing. When compared to constructs derived from non-organometallic metal complexes, assemblies carrying organic groups attached by covalent binding to the metal atom are appealing due to the opportunity of structural and functional modification. The possibility of modifying the chemical composition of the substituents attached to the metal enables one to vary the electronic and steric properties around the metal atom and to functionalize the interior and periphery of the resulting molecules, which is important for potential applications such as adsorption, molecular recognition, supramolecular catalysis, etc.6 Organotin compounds exhibit relatively inert metal−carbon bonds with aliphatic and aromatic substituents and can be frequently manipulated under non-inert conditions.3−5 Moreover, the coordination chemistry of organotin compounds is quite predictable, in particular when bidentate ligands such as carboxylates or dithiocarbamates are used. Diorganotin dicarboxylates and bis-dithiocarbamates, i.e., R2Sn(OOCR)2 and R2Sn(dtc)2,7 are generally six-coordinate complexes with skewed-trapezoidal bipyramidal coordination environments,

INTRODUCTION The opportunities related to assemblies based on the combination of metal ions and organic ligands are exponentially increasing as seen from the number of publications in this field.1−6 This can be attributed to the large number of metals in the periodic table of the elements with varying properties such as coordination number and geometry, oxidation state, color, magnetic moments, etc., which, in combination with organic ligands, give an almost infinite number of zero-, one-, two-, and three-dimensional (0D, 1D, 2D, and 3D) aggregates.3−5 Self-assembly consists in the spontaneous transformation of the participating reagents to the product, a process in which the reversibility of bond formation and mechanisms of selfcorrection are essential to achieve the formation of frequently only the thermodynamically most stable aggregate.2 Assemblies based on metal−ligand bonds are ideal in this context due to the still relatively strong directional character and the weaker bond strengths when compared to traditional covalent bonds. At the beginning of the 21st century, organotin compounds started to receive attention in the field of metal-coordination driven self-assembly, and since then, the number of reports on macrocyclic,3 cage-type,4 and infinite 1D, 2D, and 3D © 2019 American Chemical Society

Received: February 27, 2019 Published: June 10, 2019 2443

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics

However, only recently, the first macrocyclic and cage-type structures based on the combination of di- and trinuclear organotin building blocks and organic dicarboxylic acid ligands were assembled.19 Herein, we report further details on the synthesis and structural characteristics of a series of four dinuclear organotin compounds (BB1−BB4), which in part were described in a previous communication.19 In addition, BB1−BB4 were employed for the generation of a total of five novel macrocyclic structures, using 2,5-pyridinedicarboxylate, 3,5-pyridinedicarboxylate, and piperazine bis-dithiocarbamate as ligands (MC1−MC5), thus extending the chemistry of the previously reported [1 + 1] macrocyclic analogue derived from BB1 and the potassium salt of terephthalic acid.19

while the triorganotin analogues are essentially four-coordinate with distorted tetrahedral geometry, i.e., R3Sn(OOCR) and R3Sn(dtc).7a,8 The above-mentioned circumstances are appealing for the generation of di- and oligonuclear building blocks (tectons)9 based on tin for molecular self-assembly. The linkage of tin atoms by organic spacer groups generates a kind of secondary building block10 with two or more metal ions suitable for interconnection through inorganic and organic ligands. In comparison to metalla-supramolecular assemblies based on single metal ions, which exercise the role of nodes for the connection with di- or oligotopic ligands in the 1D, 2D, and 3D space,11 oligonuclear specimens enable the establishment of structural characteristics required for the generation of specific assemblies. This is particularly relevant for the assembly of cage-type structures and metal−organic frameworks (MOFs).10,11 The self-assembly of organic building blocks, which are frequently employed for the generation of cages and covalent organic frameworks (COFs),12 requires the formation of reversible covalent bonds. In contrast, metalbased tectons have the advantage that the intermolecular linkage is achieved by means of covalent-coordinate bonds, for which a large number of ligands is available. Although an important number of open-chain and cyclic di-, tri-, tetra-, and even octanuclear organotin compounds, in which the tin atoms are connected by an aliphatic or aromatic spacer group, have been reported in the past,13 their application for the synthesis of macrocycles, cages, and coordination polymers is still little explored. Newcomb’s group explored a series of organotin macrocycles and cages based on building blocks of composition BrPh2Sn−(CH2)n− SnPh2Br (n = 4, 5, 6, 8, 10, and 12) with the main objective to study their recognition properties for halide anions.14 The research group of Jousseaume synthesized a number of related dinuclear tetraorganotin compounds, but carrying aromatic connectors, some of which were then converted into polymeric organotin oxides.15 Jurkschat and co-workers prepared a series of macrocyclic ferrocenocyclophanes starting from 1,1′-Fe(C5H4−SiMe2CH2SnR2Cl)2 (R = Me, Ph)16a,b and, in collaboration with Dakternieks et al., a significant number of bis(tetraorganodistannoxane)-based cages of variable sizes, which were derived mostly from different dinuclear organotin tetrahalides of composition X2RSn−Y−SnRX2 (X = Cl, I, OTf; R = CH2CMe3, CH2CHMe2, CH2SiMe3, Ph; Y = organic spacer group).16c−l In a conjunct research project between the Tuononen and Power research groups, more recently, an interesting tetranuclear cluster containing five ethylene moieties bridging four aryl substituted tin atoms and one tin−tin bond was achieved.17 Finally, Dehnen’s research group achieved the formation of ball-shaped capsules through the interconnection of organotin sulfide and selenide clusters by organic hydrazone-based spacers.18 In the above-listed compounds, the tin atoms are connected only by organic spacer groups and, for the case of Jousseaume’s inorganic−organic hybrid materials and the cage compounds reported by Dakternieks/Jurkschat and Dehnen, additionally by Sn−O, Sn−S, or Sn−Se bonds. Considering the vast literature on metal-coordination polymers, it is evident that the introduction of di- or oligofunctional organic ligands, such as carboxylic acids, dithiocarbamates, phosphates, amines, etc., for the connection between di- and oligonuclear organotin fragments enables the creation of additional and including more complex assemblies.



RESULTS AND DISCUSSION Synthesis and Characterization of Building Blocks BB1−BB4. The {(triphenylstannylmethyl)dimethylsilyl}-substituted biphenyl derivative Ph3SnCH2Si(Me)2−C12H8−Si(Me)2CH2SnPh3 (PRE) (Scheme 1) is an excellent precursor Scheme 1. Syntheses of the Building Blocks BB1−BB319 and BB4a

a

Alternatively, BB1 and BB2 were obtained from the reactions of BB3 and BB4, respectively, with silver chloride, AgCl, in CH2Cl2.

for further functionalization of the tin centers by halogen atoms, which in subsequent reactions can be interchanged by stronger binding donor ligands having oxygen or sulfur atoms. PRE was obtained following a previously established two-step reaction protocol consisting of the synthesis of precursor ClCH2Si(Me)2−C12H8−Si(Me)2CH2Cl from 4,4′-dibromobiphenyl and two molar equiv of chloro(chloromethyl)dimethylsilane in the presence of nBuLi, followed by introduction of the triphenyltin moieties via a Grignard reaction in the presence of triphenyltin chloride.19 Using the reaction sequences developed for the step-by-step halogenation of triphenyltin derivatives,20 the dinuclear 2444

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics organotin building blocks BB1−BB3 19 and the novel compound BB4 were obtained in yields ranging from 95% to 98%. For the preparation of the chlorine-substituted derivatives BB1 and BB2, the Sn−Ph bonds in PRE were cleaved by using HCl dissolved in dichloromethane. On the contrary, the iodine-substituted derivatives BB3 and BB4 were formed by reactions of PRE with elemental iodine (Scheme 1). Alternatively, the chlorine-substituted compounds BB1 and BB2 were obtained from the corresponding iodine-substituted analogues BB3 and BB4 by reaction with silver(I) chloride. These reactions require long reaction times (1 week), but the yields are almost quantitative (Scheme 1). The novel compound BB4 was completely characterized by elemental analysis, IR and NMR spectroscopy, and mass spectrometry (see the Experimental Section and Supporting Information, Figures S1−S7).

Figure 1 shows the molecular structure of PRE, and Figure 2 illustrates in a comparative overview the molecular structures of BB119−BB4. The mono-halogenated analogues BB119 and BB3 are isostructural. Table S1 and Tables 1 and 2 contain crystallographic data and selected interatomic distances, angles, and torsion angles, respectively. In the molecular structures of PRE and BB119−BB4, the Sn−Cph, Sn−Cl, and Sn−I bond distances with values in the ranges of 2.087(12)−2.155(10) Å, 2.3426(6)−2.3778(13) Å, and 2.6786(6)−2.7229(12) Å, respectively, are as expected.13,21 However, despite the sp3-hybrid character, the Sn−CH2 bond distances with values of 2.109(2)−2.137(9) Å are similar to the Sn−Cph bond lengths (Table 1), although the bond energies of the latter are weaker as seen from their susceptibility for reaction with HCl and I2 (vide supra). The bond angles around the tin atoms in compound PRE, with values ranging from 101.4(4)° to 114.2(3)°, deviate less from the ideal tetrahedral angle (109.5°) than the bond angles in the halogenated derivatives BB119−BB4 with values ranging from 99.79(2)° to 129.96(8)°. In compounds BB119−BB4, the bond angles corresponding to the Cph−Sn−CH2 bonds are significantly larger than the Cph−Sn−Cph, C−Sn−X, and X− Sn−X (X = halogen) bond angles (Table 3). The largest difference is observed for the tetra-chlorinated derivative BB2, in which the Cl−Sn−Cl bond angles approach 100° [101.54(3)° and 99.79(2)°], while the Cph−Sn−CH2 bond angles are 129.96(8)° and 124.73(9)°. The Si−CH2−Sn bond angles for BB1−BB4 range from 116.6(2)° to 120.46(11)°. The molecular structures of BB119−BB4 exhibit different conformations (rotamers) of the Si(Me)2CH2SnPh(3−n)Xn

Figure 1. Molecular structure of compound PRE. Hydrogen atoms and some atom labels are omitted for clarity. Thermal ellipsoids drawn at the 30% probability level.

Figure 2. Molecular structures of compounds BB1,19 BB2, BB3, and BB4. Hydrogen atoms and some atom labels are omitted for clarity. Thermal ellipsoids drawn at the 30% probability level. 2445

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics Table 1. Selected Interatomic Distances [Å] for Compounds PRE, BB1,19 BB2, BB3, and BB4 PRE (Sn1)

PRE (Sn2)

Sn−X Sn−CPh

Sn−CH2 Si−Me Si−CH2 Si−Cbiph

2.087(12) 2.144(10) 2.155(10) 2.137(9) 1.836(10) 1.851(10) 1.846(9) 1.891(10)

2.129(10) 2.146(9) 2.154(10) 2.133(9) 1.837(10) 1.855(10) 1.888(9) 1.888(10)

BB119

BB2 (Sn1)

BB2 (Sn2)

BB3

BB4 (Sn1)

BB4 (Sn2)

2.3778(13)

2.3463(6) 2.3495(6) 2.113(2)

2.3426(6) 2.3497(6) 2.113(2)

2.7229(12)

2.6954(5) 2.7220(5) 2.135(6)

2.6786(6) 2.7198(5) 2.128(7)

2.109(2) 1.858(2) 1.860(2) 1.881(2) 1.880(2)

2.109(2) 1.857(2) 1.859(2) 1.876(2) 1.876(2)

2.134(10) 1.844(10) 1.849(11) 1.876(10) 1.901(7)

2.135(5) 1.872(6) 1.856(6) 1.879(6) 1.879(6)

2.124(5) 1.849(6) 1.868(6) 1.877(6) 1.884(6)

2.130(5) 2.133(4) 2.130(4) 1.848(5) 1.852(5) 1.862(4) 1.858(5)

2.131(6) 2.137(5)

Table 2. Selected Interatomic Angles and Torsion Angles [deg] for Compounds PRE, BB1,19 BB2, BB3, and BB4 PRE (Sn1)

PRE (Sn2)

Cph−Sn−X

103.96(13) 105.80(13) 101.61(13)

CH2−Sn−X X−Sn−X Cph−Sn−Cph

Cph−Sn−CH2

Cbiph−Si−CH2 Si−CH2−Sn CH3−Si−CH3 Sn−CH2−Si−Cbiph Si−CH2−Sn−Cph

101.4(4) 108.2(4) 110.9(4) 109.4(4) 112.5(4) 114.2(3) 109.9(4) 118.0(5) 110.1(5) 173.8(5) −107.9(6) 7.0(7) 128.4(6)

107.0(4) 108.0(4) 108.5(4) 110.2(4) 111.1(4) 111.9(3) 110.5(4) 118.3(5) 110.7(5) −172.0(5) −143.2(5) 97.2(6) −24.3(7)

Si−CH2−Sn−X Cbiph−Cbiph−Si−CH2

BB119

−110.0(8) 69.4(10)

40.8(10) −146.1(9)

BB2 (Sn1)

BB2 (Sn2)

104.57(6) 105.17(6) 104.96(7) 107.32(6) 101.54(3)

107.59(6) 104.04(6) 109.47(7) 108.30(6) 99.79(2)

106.10(19)

BB3 105.1(2) 105.6(2) 103.8(3)

BB4 (Sn1)

BB4 (Sn2)

102.9(13) 107.27(13) 104.37(14) 109.54(13) 105.513(16)

107.93(15) 106.69(15) 104.03(13) 111.82(13) 105.67(2)

108.6(3)

121.07(18) 116.42(17)

129.96(8)

124.73(9)

111.7(3) 120.6(4)

125.35(19)

119.79(19)

107.7(2) 116.6(2) 110.0(2) −55.9(3) 87.3(3) −44.0(3)

106.85(10) 119.32(11) 111.45(12) −64.01(15) 50.1(2)

108.02(10) 120.46(11) 111.01(11) 45.92(15) −74.4(2)

109.8(5) 119.0(6) 109.4(6) 58.7(7) −78.9(6) 50.6(7)

108.8(2) 119.0(2) 110.6(3) −41.8(3) 73.7(3)

109.4(2) 117.6(2) 111.7(3) 50.1(3) −52.8(3)

−158.4(2)

−74.2(1) 178.3(1) 133.55(18) −41.9(2)

55.0(1) 162.9(1) 83.85(19) −93.43(19)

163.9(5)

−168.5(2) −55.9(3) −82.1(5) 100.6(5)

−173.4(2) 73.1(3) 60.5(5) −121.7(4)

−76.8(4) 102.7(4)

−105.9(7) 72.4(7)

by a π···π contact (centroid···centroid distance = 3.87 Å). As already indicated by the bond angles around the metal centers, the analysis of the intermolecular interactions in the crystal structures of PRE and BB119−BB4 does not reveal Sn···X bonds, but only weak secondary C−H···π, C−H···X, and van der Waals contacts. Synthesis and Characterization of the Macrocycles MC1−MC5. As reported in our previous communication on the macrocyclic ring formation between BB119 and organic dicarboxylates, dinuclear building blocks such as BB1−BB4 are suitable for tin-coordination driven self-assembly. Terephthalate provided a 22-membered [1 + 1] macrocyclic assembly, whereas dicarboxylates in which the metal-coordination functions are separated by a larger connector give rise to [2 + 2] or even bigger aggregates including 1D coordination polymers.19 In order to explore the supramolecular chemistry of these tectons further, BB119−BB4 were reacted with two nitrogen-containing ligands, viz., dipotassium 2,5- and 3,5pyridinedicarboxylate, and the dipotassium salt of piperazinebis-dithiocarbamate (bis-dtc), respectively. The 2,5-pyridinedicarboxylate was explored based on previous reports showing that 2-pyridinecarboxylate derivatives usually form complexes with N→Sn coordination, giving five-membered chelate rings,

fragments, as illustrated by the Sn−CH2−Si−Cbiph, Si−CH2− Sn−Cph, Si−CH2−Sn−X, and Cbiph−Cbiph−Si−CH2 torsion angles given in Table 3. Within the molecular structures, the Si−CH2 bonds point in the same or opposite directions, giving rise to either a syn- or an anti-conformer. Moreover, the biphenylene connectors deviate significantly from planarity, giving a twist between the mean planes of the C6H4 rings varying from 40.0° to 46.8° (BB1,19 40.3°; BB2, 40.0°; BB3, 41.4°; BB4, 46.8°). In PRE, the twist is significantly smaller (21.6°). The syn-conformation and spatial orientation of the halogen atoms in isostructural BB119 and BB3 illustrates the predisposition of the building blocks for macrocyclic ring formation when combined with an appropriate ditopic organic ligand such as a dicarboxylate or bis-dithiocarbamate (Figure 2). The size of BB119−BB4 as potential tectons can be described by the intramolecular Sn···Sn distances, which are in the range of 11.234(1)−12.4779(6) Å. A further interesting aspect in the molecular structures of BB119−BB4 is that, independent of the molecular conformation, they are all stabilized by intramolecular C−H···π contacts (H···π = 2.56−2.81 Å), which is illustrated representatively for compound BB2 in Figure 3. Compounds BB119−BB3 exhibit two such contacts. In compound BB4, one of them is replaced 2446

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics

project in order to examine if similar structures to the dinuclear tin complexes with organic dicarboxylates can be formed, giving the first macrocyclic ring structures of a dinuclear organotin tecton based on a bis-dtc. Macrocycles MC1 and MC2. Compounds MC1 and MC2 were obtained in good yields starting from the chlorine- or the iodine-substituted organotin compounds BB119 and BB3, respectively (Scheme 2). In combination with the single-crystal X-ray diffraction (SCXRD) analysis of MC1, the microanalytical and spectroscopic data of MC1 and MC2 indicate the formation of [1 + 1] assemblies, in which the pyridine nitrogen atoms are not coordinated to any of the tin atoms, indicating that the aromatic connector is ideal for macrocyclic ring closure (Scheme 2). The 1H, 13C, 29Si, and 119Sn NMR data for MC1 exhibit signals for a completely unsymmetrical molecular structure. Compound MC2 is C2v-symmetric, showing only signals for half of the molecule. Complete assignment of the 1H and 13C NMR signals was achieved by COSY, HSQC, and HMBC NMR experiments (see Figure 4 and Supporting Information, Figures S8−S15 and S18−S22). The 1H, 13C, and 29Si NMR signals for the organotin skeletons in MC1 and MC2 are similar to those observed for the previously reported analogue containing terephthalate.19 The coordination of the carboxylate ligand to the tin atoms in M1 and M2 is evidenced by (i) the 13C NMR chemical shifts for the SnOC(O) carbon atoms (δ = 169.5 ppm for MC1 and δ = 169.1 ppm for MC2) being typical for organotin carboxylates,24 and (ii) by the characteristic νas(COO) and νs(COO) IR vibrations at approximately 1650 and 1320 cm−1, respectively, indicating a monodentate coordination mode.25 The 119Sn NMR spectra measured in CDCl3 gave signals for tetracoordinated tin(IV) atoms,26 δ = −24 and −27 ppm for MC1 and δ = −26 ppm for MC2, thus confirming the absence of N→Sn coordination. Contrary to the [1 + 1] assembly of BB1 with terephthalate, for which the molecular ion was not observed, the mass spectra of MC1 and MC2 show mass clusters for the molecular ions, [M + H]+, at m/z = 1008, which agree reasonably well with the simulated patterns (Figure 4; Figures S16 and S23 in the Supporting Information (SI)). This can be attributed to the presence of the pyridine N atom that is susceptible to protonation by the solvent mixture used for the experiment (CH2Cl2/MeOH). For both compounds, there are also mass clusters centered at m/z 1894.9 (MC1) and 1894.2 (MC2) that are indicative of tetranuclear species. Single crystals of MC1, as its aqua-ethanol complex MC1· H2O·EtOH, suitable for XRD analysis (Table S2) were grown from a solution in ethanol/water (3:1, v/v). Figure 5 shows the molecular structure, and Table 3 contains selected interatomic distances, angles, and torsion angles. The water molecule coordinates the Sn(1) and the ethanol molecule the Sn(2) atom. A similar adduct formation occurred with the macrocycle analogue prepared with terephthalate19 and can be attributed to crystal lattice stabilization effects. In consequence, the tin atoms in the solid-state structures of MC1 have distorted trigonal-bipyramidal coordination geometries with the axial sites occupied by the ligand and solvent oxygen atoms. The organic substituents comprise the equatorial positions. The Sn−Osolv distances are significantly larger [2.480(2) Å for Sn(1)−O(5) and 2.390(3) Å for Sn2−O6] than the Sn−OOCO distances [2.129(2) Å for Sn(1)−O(1) and 2.153(2) Å for Sn(2)−O(3)]. The Sn−OC(O) bond is essentially monodentate [3.307(3) Å for Sn(1)···O(2);

Table 3. Selected Interatomic Distances [Å], Angles [deg], and Torsion Angles [deg] of Compound MC1 Sn−OOCO Sn···OOCO Sn−Osolv Sn−CH2 Sn−Cph OOCO−Sn−OOCO OOCO−Sn−Osolv OOCO···Sn−Osolv OOCO−Sn−Cph OOCO···Sn−Cph Osolv−Sn−Cph OOCO−Sn−CH2 OOCO···Sn−CH2 Osolv−Sn−CH2 Cph−Sn−Cph Cph−Sn−CH2 Sn−CH2−Si−Cbiph Si−CH2−Sn−Cph Si−CH2−Sn−Osolv Cbiph−Cbiph−Si−CH2

Sn1

Sn2

2.129(2) 3.307(3) 2.480(2) 2.122(3) 2.135(3) 2.156(4) 41.98(7) 176.21(8) 140.49(6) 88.43(11) 95.68(10) 130.1(1) 81.7(1) 89.36(11) 82.52(9) 99.94(11) 71.9(1) 83.85(10) 115.81(13) 117.41(13) 124.78(12) 66.2(2) 18.1(3) −178.7(2) 103.0(2) 75.4(3) −100.0(3)

2.153(2) 3.441(3) 2.390(3) 2.112(3) 2.136(3) 2.193(3) 39.47(8) 173.94(9) 146.37(7) 87.47(14) 95.05(12) 126.8(1) 79.9(1) 86.55(13) 88.65(13) 97.01(12) 72.1(1) 85.57(11) 117.86(16) 116.78(14) 124.48(16) −69.1(2) −6.6(3) −175.5(2) −89.3(2) −53.9(3) 126.5(3)

Figure 3. Perspective view of the molecular structure of compound BB2, showing the stabilization of the anti-conformation by intramolecular C−H···π contacts.

which can alter the macrocyclic ring structure significantly.22 In 3,5-pyridinedicarboxylate, the donor atoms of the carboxylate and pyridine functions are disposed at an angle of 120°, which, with mononuclear organotin moieties, has given rise to trinuclear macrocyclic ring structures and 2D coordination polymers.23 The bis-dtc ligand was included in this research 2447

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics Scheme 2. Reaction Sequences for the Synthesis of the Macrocyclic Compounds MC1−MC5a

a

Reaction conditions: (i) H2O/EtOH (1:1, v/v), r.t., 12 h; (ii) EtOH, r.t., 12 h. Note: K2(25pdc) = potassium 2,5-pyridinedicarboxylate, K2(35pdc) = potassium 3,5-pyridinedicarboxylate, K2(bis-dtc) = potassium piperazine bis-dithiocarbamate.

quantum-mechanical DFT calculations using a combination of the B3LYP functional and the def2-SVP basis set. Previous reports have shown that DFT calculations confidently reproduce the geometries of diorganotin complexes obtained from SCXRD analysis.3j,m,19,27 In order to reduce the costs for computing time, geometry optimizations were realized on a modified version of MC2, in which the Sn-phenyl substituents were replaced by methyl substituents (MC2′). The molecular skeleton of compound MC2′ is relatively rigid, exhibiting conformational flexibility essentially only at the C−C junction of the biphenylene spacer and the Si−CH2−Sn moiety. Further, the coordination mode of the carboxylate groups can vary. Realizing optimizations starting from different initial geometries through systematic variation of the conformation and the connectivity of the metal-coordinating carboxylate function, a total of six molecular structures having energy minima were identified (Figure 6). The geometric parameters varying among the six conformers are indicated by arrows. In the most stable conformer, the Si− CH2−Sn methylene groups are mutually anti-oriented, each pointing into the direction of the secondary Sn···O bond of the anisobidentate-coordinated 3,5-pyridinedicarboxylate ligand. The twist between the C6H4-ring mean planes in the −C12H8− spacer is 35.7°. Table 4 summarizes the geometric features of the remaining conformers, which enables the following conclusion: The Si−CH2−Sn methylene groups can be either syn (entries 2, 3, 5, and 6) or anti (entries 1 and 4).

3.441(3) Å for Sn(2)···O(4)]. The distorted character of the coordination polyhedra around Sn(1) and Sn(2) is best illustrated by the O−Sn−O bond angles, with values of 176.21(8)° for O(1)−Sn(1)−O(5) and 173.94(9)° for O(3)− Sn(2)−O(6), and the C−Sn−C bond angles ranging from 115.81(13)° to 124.78(12)°. In accordance with the weak Sn− Osolv bond, the bipyramid is distorted toward a monocapped tetrahedron, as indicated by a comparison of the OOCO−Sn−C and Osolv−Sn−C bond angles, which are significantly larger in the first case [87.47(14)−99.94(11)° versus 82.52(9)− 89.36(11)°]. The molecular structure of MC1 is of cyclophane-type with approximate parallel orientation of the aromatic rings. The twist of the C6H4 rings of the biphenylene spacer is 32.2° (angle between mean planes). Nevertheless, this accommodation enables π−π interactions with centroid···centroid distances of 4.56 and 4.26 Å between the C6H4 moieties and the pyridylene group. The corresponding intra-annular COCO··· centroid distances are 3.69 and 4.11 Å. Within the macrocycle, the Sn···Sn, Si···Si, and Sn···Si distances are 11.0282(6), 10.907(1), and 3.3969(9)/3.4285(9) Å, respectively. The molecular conformation is also influenced by the torsion of the CbiphSiMe2−CH2SnR2 fragments, exhibiting a gauche-conformation along the Si−CH2 bonds (see Table 3). The mutual orientation of the Si−CH2−Sn methylene groups, which is syn in MC1, constitutes a further important element for the definition of the molecular conformation. Since, for MC2, single crystals suitable for XRD analysis could not be grown, the molecular structure was analyzed by 2448

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics

Figure 4. 1H, 13C, 29Si, and fragment [M + H]+.

119

Sn NMR spectra and isotope distribution pattern extracted from the ESI+-TOF mass spectrum of MC1 for the

as the methylene groups (entries 2 and 6) or in the opposite direction (entries 3 and 5). The carboxylate groups are always anisobidentate, but the coordination to the tin atoms can be symmetric (Sn···O bonds have the same orientation; entries 2, 3, and 4) or unsymmetrical (Sn···O bonds have alternate orientation; entries 1, 5, and 6). The biphenylene spacer is always significantly twisted with a torsion angle in the range of 31.1−37.1°. The differences of the relative energies in the geometryoptimized structures are negligible within the error of the method (ΔE = 0.09−1.40 kcal/mol, Figure 6). This indicates that the conformers coexist in solution through fast dynamic equilibria as previously observed for a series of macrocyclic diorganotin dithiocarbamate carboxylate carrying aliphatic chain spacers of varying sizes.3m This is underlined by the observation of a single and symmetric set of signals in the 1H and 13C NMR spectra, the absence of AB systems for the CH2 and SiCH3 hydrogen atoms, and a single 119Sn NMR signal (Figures S18−S22, SI) Macrocycles MC3, MC4, and MC5. Compound MC3 was synthesized by a single-pot two-step reaction, consisting of the in situ generation of the potassium salt of piperazine-bisdithiocarbamate starting from piperazine, carbon disulfide, and potassium hydroxide, followed by addition of 1 equiv of BB1 or BB3. Similarly, MC4 and MC5 were prepared, but using BB2 and BB4 as starting materials, respectively (see the Experimental Section). The compositions of MC3, MC4, and MC5 were established by elemental analysis, mass spectrometry, IR and NMR spectroscopy, and SCXRD analysis for MC4 and MC5.

Figure 5. Molecular structure and coordination geometries about the Sn(1) and Sn(2) atoms in the crystal structure of compound MC1. Hydrogen atoms and some atom labels are omitted for clarity. Disorder is not shown. Thermal ellipsoids are drawn at the 30% probability level.

In the syn-conformers, the nitrogen atom of the 3,5pyridinedicarboxylate ligand can point in the same direction 2449

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics

Figure 6. Geometry-optimized molecular structures for syn- and anti-MC2′ at the B3LYP/def2-SVP level (syn and anti are with respect to the mutual orientation of the Si−CH2−Sn methylene groups).

Table 4. Geometric Features of the Geometry-Optimized Conformers for MC2′ (in the Gas Phase) entry

rel. energy [kcal/mol]

mutual orientation Si−CH2−Sn groups

1 2 3 4 5 6

0.00 0.09 0.53 0.80 0.92 1.40

anti syn syn anti syn syn

orientation Npy with respect to Si−CH2−Sn syn anti anti syn

Contrary to MC1 and MC2, the mass spectra of MC3−MC5 do not exhibit mass clusters for the molecular ion, but there are characteristic mass clusters centered at m/z = 1001, 959, and 1051 that are assigned to the fragments [MC3 − Ph]+, [MC4 − Cl]+, and [MC5 − I]+, respectively (Figures S32, S40, and S47, SI). Besides the signals for the organotin tecton, which have chemical shifts and multiplicities similar to the above-discussed macrocycles containing pyridine dicarboxylates, the 1H NMR spectra of MC3−MC5 contain a single signal at approximately δ = 3.6 ppm for the NCH2 hydrogen atoms. As a result of the delocalization of π-electron density into the CS2 fragment, the signal is low-field shifted when compared to piperazine (δ = 2.78 ppm). In the 13C NMR spectrum, the NCH2 carbon atoms are displaced to approximately δ = 49 ppm giving lowfield shifts of approximately Δδ = 2.0 ppm when compared to piperazine. The signals at δ = 197.1 (MC3), 197.8 (MC4), and 199.5 (MC5) ppm are typical for tin-coordinated dtc groups.28 The 119Sn NMR chemical shifts of δ = −108 ppm for MC3, δ = −251 ppm for MC4, and δ = −333 ppm for MC5 indicate four-coordinate tin atoms.28 Diffusion-ordered spectroscopy (1H-DOSY NMR) enables the determination of the diffusion coefficient of a chemical species in solution.29 Since the diffusion coefficient is related to the molecular volume and weight, the molecule size can be established. The DOSY NMR spectrum of MC3 in C6D6 is illustrated in Figure 7. Using the Stokes−Einstein equation30 and based on the diffusion constant of D = 4.82 × 10−10 m2 s−1, for MC3, a hydrodynamic radius of 7.5 Å and a volume of 1770 Å3 were determined. The molecular weight calculated according to Morris’ procedure (1455 g mol−1) is in good agreement with the molecular weight corresponding to the elemental composition (1078.79 g mol−1).29c The diffusion constant for MC3 is comparable to the value reported by Cohen for a calix[4]arene monomer of similar size (MW = 1160 g mol−1, D = 4.6 × 10−10 m2 s−1, solvent: C6D6 with 3% DMSO).31

coordination mode of OCO groups

symmetry of OOCO−Sn coordination

twist in −C12H8− spacer [deg]

anisobidentate anisobidentate anisobidentate anisobidentate anisobidentate anisobidentate

asymmetric symmetric symmetric symmetric asymmetric asymmetric

35.7 31.1 33.9 37.1 35.6 36.6

Figure 7. 1H-DOSY NMR spectrum of MC3 at a concentration of 1.5 × 10−2 mol/L in C6D6.

Single crystals of MC4, as its tetrahydrofuran solvate MC4· 2THF, and MC5, as its dichloromethane solvate MC5·CH2Cl2, suitable for XRD analysis were grown from THF and dichloromethane, respectively (Table S2). Figure 8 shows the molecular structures, and Table 5 contains selected interatomic distances, angles, and torsion angles. In both compounds, the molecular structures exhibit disorder over two positions for the piperazinylene fragment, indicating the presence of two conformers within the solid state structures (Figure 8). The conformers are different for MC4 and MC5, giving in an approximate 1:1 ratio two conformational isomers with a chair conformation in the crystal structure of MC4 (conformers I and II), while, in MC5, the major isomer (≈2:1 ratio) shows a twisted conformation (conformer III). Aside from the differences in the piperazine-bis-dithiocarbamate ligand, the molecular structures of MC4 and MC5 are 2450

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics

Figure 8. Molecular structures and coordination geometries about the tin atoms in the crystal structures of compounds (a) MC4 and (b) MC5, showing the presence of two different conformational isomers in each case (disorder over two positions of the piperazinylene fragments). Hydrogen atoms are omitted for clarity. Thermal ellipsoids are drawn at the 30% probability level.

quite similar. The tin-bound halogen atoms are trans, and the Si−CH2−Sn methylene groups are mutually disposed anti. Of the Sn−S/Sn···S bonds, the secondary Sn···S bonds are trans to the Sn−X (X = Cl, I) bond in both compounds. As typically observed in triorgano- and diorganotin dithiocarbamates, the short Sn−S bonds have strong covalent character with bond distances in the range of 2.464(3)−2.4816(7) Å. In contrast, the weak Sn···S contacts with distances ranging from 2.6681(7) to 2.717(3) Å are essentially secondary interactions (Table 5). Nevertheless, when compared to diorganotin bisdithiocarbamates, the Sn−S and Sn···S bonds in MC4 and MC5 are relatively short and exhibit a coordination mode that approaches an isobidentate character.3j,m,27e,28d According to the τ-parameter,32 the trigonal-bipyramidal coordination geometries of the tin atoms are distorted more significantly toward a square pyramid than in MC1 (MC1, τ = 0.86 for Sn1 and τ = 0.82 for Sn2; MC4, τ = 0.59; MC5, τ = 0.61 for Sn1 and 0.60 for Sn2). The Sn···Sn, Si···Si, and Sn···Si distances in MC4 and MC5 are similar among each other (Table 6). However, when compared to MC1, the Sn···Sn distance is increased by 0.30 Å for MC4 and 0.35 Å for MC5. As in MC1, the C6H4 rings of the biphenylene spacer are twisted by 5.7° and 36.1° (angle between mean planes) for MC4 and MC5, respectively.

In the crystal lattice of MC4, the formation of doublebridged C−H···Cl hydrogen-bonded one-dimensional polymers in the direction of axis [1 0 0] is observed (Figure 9), which by additional C−H···Cl and S···S contacts are connected into 2D layers (Table 7). On the contrary, in MC5, hydrogenbonded double chains formed through a combination of C− H···I and C−H···S interactions are observed along [0 1 0] (Table 7). The geometric characteristics of the C−H···Hal, C− H···S, and S···S contacts are within the range typically observed for such interactions.28e,33 In order to evaluate the energy differences between the cisand trans-stereoisomers of MC4 and MC5 (concerning the mutual orientation of the halogen atoms), DFT calculations were realized as a representative example for a modified version of MC4, in which the Sn-phenyl groups were replaced by Sn-methyl substituents (MC4′). As for MC2′, geometry optimizations for cis-MC4′ and trans-MC4′ were carried out using the B3LYP/def2-SVP method, starting in each case from different initial geometries in order to establish the most stable conformer. The geometry-optimized molecular structures for cis-MC4′ and trans-MC4′ and their relative energies are illustrated in Figure 10. A comparison between the geometry extracted from the crystal structure of trans-MC4 and the calculated structure shows good overall agreement for the covalent bond distances 2451

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics

organotin carboxylates and dithiocarbamates on the 119Sn NMR time scale. The dynamic behavior of macrocyclic assemblies can be induced by intramolecular and, at higher concentrations, also by intermolecular ligand exchange reactions.34a−c For MC4 and MC5 with five-coordinated metal atoms, Berry-pseudorotation might be an option for the interconversion between the cis- and trans-isomers. However, in this case, this mechanism is excluded due to the small bond angle of the bidentate dithiocarbamate group with the tin atom (≈69° in the solid state structures), which disfavors the generation of the required square pyramid as intermediate. Considering also that, in low-polarity solvents, such as CDCl3, the formation of ionic species is not favored, e.g., by dissociation of the polar Sn−Hal bond, the mechanism outlined in Scheme 3 is proposed for the configuration interchange in nonpolar solvents. In the initial molecular configuration on the left, the Sn−Cl substituents are trans and both tin atoms have distorted trigonal-bipyramidal polyhedra, in which the axial positions are occupied by the chlorine atom and the sulfur atom involved in a secondary bond. The intermediate comprises a species with isobidentate coordination of the dithiocarbamate ligand to the tin atom marked with an asterisk, in which now the organic substituents occupy the axial positions of the trigonal-bipyramidal coordination geometry. This configuration is facilitated by rotation of the −Si−CH2−Sn fragment around the Si−Cbiph, Si−CH2, and Sn−CH2 single bonds and the simultaneous re-accommodation of the chlorine and phenyl substituents. Upon further rotation of the groups involved, the cis-isomer is generated, in which again the chlorine and the sulfur atom of the secondary Sn···S bond are in axial positions of the trigonal bipyramid surrounding the tin atom.

Table 5. Selected Interatomic Distances [Å], Angles [deg], and Torsion Angles [deg] for Compounds MC4 and MC5a Sn−X Sn−S Sn−CH2 Sn−Ph S−Sn−S S−Sn−Cph S−Sn−CH2 X−Sn−Cph X−Sn−CH2 X−Sn−S Cph−Sn−CH2 Sn−CH2−Si−Cbiph Si−CH2−Sn−Cph Si−CH2−Sn−X Cbiph−Cbiph−Si−CH2

MC4

MC5 (Sn1)

MC5 (Sn2)

2.4601(7) 2.4816(7) 2.6681(7) 2.132(2) 2.131(3) 69.55(2) 118.68(7) 95.99(7) 121.42(7) 94.93(7) 95.78(7) 96.86(7) 87.14(3) 156.69(2) 118.94(10) 55.28(17) 172.4(1) 71.9(1) −116.0(2) 66.3(3)

2.8396(9) 2.465(3) 2.717(3) 2.111(9) 2.130(9) 69.14(8) 110.0(3) 88.7(3) 120.0(3) 92.4(3) 98.5(3) 100.8(3) 87.54(6) 156.64(7) 126.8(4) 57.3(7) −171.5(5) 79.2(6) −106.1(8) 77.0(9)

2.8424(10) 2.464(3) 2.689(2) 2.122(10) 2.136(11) 69.44(8) 117.7(3) 93.6(3) 119.0(3) 92.8(3) 99.3(3) 98.4(3) 85.72(7) 155.12(7) 121.4(4) 58.8(7) −177.6(5) 76.1(5) 70.1(9) −113.1(9)

a For the bonds lengths and bond angles containing sulfur, the first value corresponds to the Sn−S and the second to the Sn···S bond.

Table 6. Selected Intramolecular Distances [Å] in the Molecular Structures of Compounds MC1, MC4, and MC5 Sn···Sn Si···Si Sn···Si

MC1

MC4

MC5

11.0282(6) 10.907(1) 3.3969(9) 3.4285(9)

11.3325(6) 10.937(1) 3.4657(8) 3.4657(8)

11.375(1) 10.938(4) 3.474(3) 3.444(3)



CONCLUSIONS This work shows that spacer-bridged biphenylene and SiMe2CH2 moieties-containing dinuclear organotin halides are suitable precursors for the formation of 21- and 22membered macrocycles by reaction with organic dicarboxylate and bis-dithiocarbamate ligands. In solution, the macrocycles are involved in conformational and configurational equilibria. DFT calculations show that the cis- and trans-stereoisomers in the case of the macrocycles carrying four different substituents at the tin atoms have negligible energy differences. The same is true for the conformational isomers resulting from variation of the conformation in the cyclohexylene spacer groups that link the metal-coordinating functions.

and angles (Table S3, SI). The secondary Sn···S bonds in the calculated gas-phase structures are overestimated, but this is a well-known phenomenon.3j The negligible difference of the relative energies indicates that the cis- and trans-stereoisomers coexist in solution, aside from different conformers (see NMR and SCXRD discussion). The observation of a single signal in the room-temperature 119 Sn NMR spectrum of MC4 indicates a dynamic equilibrium being fast on the corresponding NMR time scale. The equilibrium becomes slow at low temperature (T = −40 °C) as two signals with close chemical shifts at δ = −256 and −257 ppm are observed. This is probably not surprising in view that configurational and conformational equilibria are frequent in

Figure 9. 1D chain formation through intermolecular C−H···Cl contacts in the crystal lattice of MC4. For clarity, disorder in the piperazine-bisdithiocarbamate fragment is not shown and part of the hydrogen atoms were omitted. Symmetry operator: 2 − x, 1 − y, 1 − z. 2452

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics Table 7. Geometric Parameters for the Hydrogen Bonding Interactions in the Crystal Structures of MC4 and MC5 compound

intermolecular interaction

D−H [Å]

D···A [Å]

H···A [Å]

∠DHA [deg]

symmetry code

MC4

C8−H8B···Cl1 C3−H3···Cl1 S2···S2 C22−H22B···I1 C18−H18C···I2 C14−H14B···S4

0.98 0.95

3.81 3.51 3.22 3.81 4.03 3.87

2.83 2.77

175 136

3.16 3.17 2.99

126 149 153

2 − x, 1 − y, 1 − z 1 − x, 1 − y, 1 − z 1.5 − x, y, 0.5 − z x, −1 + y, z 0.5 − x, 0.5 + y, 1.5 − z 0.5 − x, −0.5 + y, 1.5 − z

MC5

0.97 0.96 0.97

29

Si, 119Sn) and two-dimensional (COSY, HSQC, HMBC, DOSY) NMR studies were recorded at room temperature on Bruker DRX 400 and 500 spectrometers. Standard internal and external references were used: tetramethylsilane (δ1H = 0, δ13C = 0, and δ29Si = 0) and tetramethyltin (δ119Sn = 0). Mass spectra were recorded on Jeol (FAB+) and Agilent Technologies instruments (ESI+-TOF) with CH2Cl2 or CH2Cl2/MeOH as the mobile phase. Synthesis of 1,4-[I2PhSnCH2(CH3)2Si−C12H8−Si(CH3)2CH2SnPhI2] (BB4). A 5.0 g portion of precursor 1,4-[Ph3SnCH2Si(CH3)2− C12H8−(CH3)2SiCH2SnPh3] (5.0 mmol) and 150 mL of dichloromethane were placed into a 250 mL ball flask equipped with a magnetic stirring bar. The solution was cooled to 0 °C, whereupon elemental iodine (5.2 g, 20 mmol) was added in small portions every 20 min during a period of 10 h. The reaction mixture was allowed to stir overnight. Subsequently, the solvent was evaporated and the byproduct (iodobenzene) was removed under vacuum to give BB4 as a viscous yellow oil in 98% yield (5.9 g, 4.9 mmol). Mp 69−71 °C. IR νmax: 441 (m), 523 (m), 601 (m), 638 (m), 690 (s), 726 (s), 776 (s), 801 (s), 832 (m), 996 (m), 1002 (w), 1067 (w), 1114 (m), 1251 (m), 1330 (w), 1344 (w), 1430 (m), 1479 (w), 1595 (w), 2951 (w), 3065 (w) cm−1. 1H NMR (CDCl3, 400.2 MHz, 298 K) δ: 0.50 (s, 12H, SiCH3), 1.73 (s, 4H, SiCH2) [2J(1H−117/119Sn) = 91/95 Hz], 7.36 (m, 6H, Hm, Hp), 7.54 (m, 12H, H2′, H3′, Ho) ppm. 13C{1H} NMR (CDCl3, 100.6 MHz, 298 K) δ: −0.1 [3J(13C−117/119Sn) = 24 Hz, SiCH3], 13.0 (C2), 127.1 (C3′), 129.2 [3J(13C−117/119Sn) = 76 Hz, Cm], 131.0 [4J(13C−117/119Sn) = 16 Hz, Cp], 134.4 [2J(13C−117/119Sn) = 64 Hz, Co], 134.4 (C2′), 137.3 (Ci), 137.6 (C1′), 142.3 (C4′) ppm. 29 Si{ 1 H} NMR (CDCl 3 , 79.5 MHz, 298 K) δ: −1.90 [2J(29Si−117/119Sn) = 30 Hz] ppm. 119Sn{1H} NMR (CDCl3, 149.2 MHz, 298 K) δ: −207.9 ppm. Anal. Calcd for C30H34I4Si2Sn2 (1195.80 g mol−1): C, 30.1; H, 2.9. Found: C, 30.7; H, 3.0. MS (FAB+) for [C30H34I3Si2Sn2]+ ([M − I]+): m/z (exp.) = 1069 (100%), m/z (calcd.) = 1068.74. Synthesis of MC1. A solution of 1,4-[ClPh2SnCH2Si(CH3)2− C12H8−(CH3)2SiCH2SnPh2Cl] (0.50 g, 0.55 mmol) in methanol (10 mL) was added at room temperature to a previously prepared stirred solution of 2,5-pyridinedicarboxylic acid (0.092 g, 0.55 mmol) and potassium hydroxide (0.062 g, 1.10 mmol) in 15 mL of a solvent mixture of water and ethanol (1:1, v/v). The reaction mixture was stirred overnight. After removal of the solvent, the product was washed with a solvent mixture consisting of water and ethanol (20:3), and dried in vacuo to give a colorless solid in 62% isolated yield (0.34 g, 0.34 mmol). Recrystallization from water/ethanol gave the adduct MC1·H2O·EtOH. Mp 225−228 °C. IR νmax: 448 (m), 527 (m), 601 (w), 642 (w), 696 (s), 728 (s), 776 (s), 804 (s), 814 (s), 832 (m), 998 (w), 1023 (w), 1075 (w), 1115 (m), 1159 (w), 1217 (w), 1251

Figure 10. Geometry-optimized molecular structures for cis- and trans-MC4′ with relative energies at the B3LYP/def2-SVP level (cis and trans are with respect to the mutual orientation of the Sn−Cl substituents).

For organic ligands with significantly shorter or larger organic spacers between the metal-coordination functions, the generation of larger oligo-cyclomeric [2 + 2], [3 + 3], etc., or polymeric assemblies is expected, since ring closure for the [1 + 1] aggregate is not possible anymore. Further, oligodentate organic carboxylate or dithiocarbamate ligands should give rise to cage-type structures or infinite 2D or 3D MOF-type assemblies. Because of the Lewis acidity of the tin atoms, the dinuclear tin building blocks described herein can be explored also for the molecular recognition of mono- and difunctional organic substrates. The potential of this general idea has recently been demonstrated by the use of bidentate boron Lewis acids for the complexation of a variety of diamine ligands.34d



EXPERIMENTAL SECTION

Reagents and Solvents. All solvents were dried by standard methods. 4,4′-Dibromobiphenyl, chloro(chloromethyl)dimethylsilane, triphenyltin chloride, iodine, n-butyllithium, magnesium, silver chloride, 2,5-pyridinedicarboxylic acid, 3,5-pyridinedicarboxylic acid, piperazine, potassium hydroxide, and carbon disulfide were commercially available and used as received without further purification. 1,4-[ClPh2SnCH2Si(CH3)2−C12H8−(CH3)2SiCH2SnPh2Cl] (BB1), 1,4-[Cl2PhSnCH2Si(CH3)2−C12H8−(CH3)2SiCH2SnPhCl2] (BB2), and 1,4-[IPh2SnCH2(CH3)2Si−C12H8−Si(CH3)2CH2SnPh2I] (BB3) were synthesized as reported previously.19 Instrumental Methods. Melting points were determined on a Büchi M-560 melting point apparatus. Elemental analyses were carried out on a LECO-CHNS-932 analyzer using samples dried under high vacuum for 6−8 h. IR spectra were recorded on a PerkinElmer FT-IR spectrophotometer. One-dimensional (1H, 13C,

Scheme 3. Proposed Mechanism for the Dynamic Interconversion between trans-MC4 and cis-MC4

2453

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics

mol−1): C, 53.4; H, 4.9; N, 2.6. Found: C, 53.1; H, 5.2; N, 2.7. MS (ESI+) for [C42H47N2S4Si2Sn2]+ ([M − Ph]+): m/z (exp.) = 1001.3, m/z (calcd.) = 1001.0. Synthesis of MC4. A solution of 1,4-[Cl2PhSnCH2Si(CH3)2− C12H8−(CH3)2SiCH2SnPhCl2] (0.20 g, 0.24 mmol) in methanol (10 mL) was added at room temperature to a previously prepared stirred solution of piperazine (0.019 g, 0.22 mmol), carbon disulfide (0.034 g, 0.44 mmol), and potassium hydroxide (0.027 g, 0.48 mmol) in 15 mL of ethanol. The reaction mixture was stirred overnight. After removal of the solvent, the product was washed with a solvent mixture consisting of water and ethanol (20:3) and dried in vacuo to give a colorless solid in 66% reaction yield (0.16 g, 0.16 mmol). Mp 237− 239 °C. IR νmax: 445 (m), 494 (m), 520 (m), 554 (w), 646 (m), 693 (s), 729 (s), 778 (s), 804 (s), 816 (s), 912 (w), 998 (m), 1020 (m), 1067 (w), 1224 (m), 1250 (m), 1277 (m), 1361 (w), 1382 (w), 1428 (s), 1466 (m), 1597 (w), 2951 (w), 3049 (w) cm−1. 1H NMR (CDCl3, 400.2 MHz, 298 K) δ: 0.42, 0.73 (m, 12H, SiCH3), 0.86, 1.26 (m, 4H, SiCH2), 3.57 (s, 8H, H2″), 7.41 (m, 6H, Hm, Hp), 7.72 (d, 4H, H3′), 7.86 (d, 4H, H2′), 7.90 (m, 4H, Ho) ppm. 13C{1H} NMR (CDCl3, 100.6 MHz, 298 K) δ: −1.4, 1.8 (SiCH3), 11.9, 14.2, 22.7, 31.6 (SiCH 2 ), 48.8 (C2″), 124.9 (C3′), 128.8 [3J(13C−117/119Sn) = 84 Hz, Cm], 130.2 [4J(13C−117/119Sn) = 18 Hz, Cp], 135.1 [2J(13C−117/119Sn) = 66 Hz, Co], 135.7 (C2′), 138.7 (C1′), 139.7 (C4′), 143.6 (Ci), 197.8 (CS2) ppm. 119Sn{1H} NMR (CDCl3, 149.2 MHz, 298 K) δ: −251 ppm (h1/2 = 14 Hz). 119Sn{1H} NMR (CDCl3, 149.2 MHz, 233 K) δ: −256, −257 ppm. Anal. Calcd for C36H42Cl2N2S4Si2Sn2 (995.49 g mol−1): C, 43.4; H, 4.3; N, 2.8. Found: C, 43.8; H, 4.4; N, 2.5. MS (ESI+) for [C36H42ClN2S4Si2Sn2]+ ([M − Cl]+): m/z (exp.) = 959.2, m/z (calcd.) = 959.9. Synthesis of MC5. A solution of 1,4-[I2PhSnCH2Si(CH3)2− C12H8−(CH3)2SiCH2SnPhI2] (0.20 g, 0.17 mmol) in methanol (10 mL) was added at room temperature to a previously prepared stirred solution of piperazine (0.015 g, 0.17 mmol), carbon disulfide (0.037 g, 0.48 mmol), and potassium hydroxide (0.019 g, 0.34 mmol) in 15 mL of ethanol. The reaction mixture was stirred overnight. After removal of the solvent, the product was washed with a solvent mixture consisting of water and ethanol (20:3) and dried in vacuo to give a colorless solid in 60% isolated yield (0.12 g, 0.10 mmol). Mp 234− 237 °C. IR νmax: 552 (w), 614 (m), 646 (w), 691 (s), 723 (s), 774 (s), 800 (s), 913 (w), 990 (m), 1041 (w), 1070 (w), 1112 (m), 1153 (w), 1201 (w), 1251 (m), 1281 (m), 1358 (w), 1382 (w), 1426 (s), 1465 (m), 1473 (m), 1571 (w), 1598 (w), 1725 (w), 2858 (w), 2924 (w), 2956 (w), 3045 (w) cm−1. 1H NMR (CDCl3, 500 MHz, 298 K) δ: 0.61 (s, 12H, SiCH3), 1.58 (s, 4H, SiCH2), 3.61 (s, 8H, H2″), 7.39 (m, 6H, Hm, Hp), 7.73 (d, 4H, H3′), 7.87 (d, 8H, H2′, Ho) ppm. 13 C{1H} NMR (CDCl3, 50 MHz, 298 K) δ: 13.8 (SiCH3), 45.7 (SiCH2), 49.3 (C2″), 125.2 (C3′), 128.9 [3J(13C−117/119Sn) = 86 Hz, Cm], 130.3 [4J(13C−117/119Sn) = 17 Hz, Cp], 135.2 [2J(13C−117/119Sn) = 64 Hz, Co], 136.0 (C2′), 138.8 (C1′), 140.0 (C4′), 142.2 (Ci), 199.5 (CS2) ppm. 119Sn{1H} NMR (CDCl3, 186.5 MHz, 298 K) δ: −333 ppm (h1/2 = 38 Hz). 29Si{1H} NMR (CDCl3, 99.4 MHz, 298 K) δ: −6.6 ppm. Anal. Calcd for C36H42I2N2S4Si2Sn2 (1178.39 g mol−1): C, 36.7; H, 3.6; N, 2.4; S, 10.9, Found: C, 36.1; H, 3.3; N, 2.6; S, 10.9. MS (APCI+) for [C36H42IN2S4Si2Sn2]+ ([M − I]+): m/z (exp.) = 1050.8869, m/z (calcd.) = 1050.8857. Single-Crystal X-ray Diffraction Analysis. Single crystals of compounds BB2−BB4 (CH2Cl2), MC1 (EtOH/H2O), MC4 (THF), and MC5 (CH2Cl2) were grown from solutions in the solvents indicated in parentheses. Intensity data for the crystals were collected at T = 100 or 173 K on Oxford Diffraction Xcalibur Sapphire3 (BB2, BB3, and MC4) and Agilent Technologies SuperNova (BB4, MC1, and MC5) diffractometers equipped with a CCD area detector, using Mo Kα radiation (λ = 0.71073 Å). The measured intensities were reduced to F2 and corrected for absorption using spherical harmonics (SCALE3 ABSPACK in CrysAlisPro).35 Intensities were corrected for Lorentz and polarization effects. The structures were solved with direct methods using SHELXS or SHELXT.36 Refinements were carried out against F2 by using SHELXL37 and the OLEX2 program package.38 All non-hydrogen atoms were refined using anisotropic displacement parameters. C−H hydrogen atoms were positioned with

(m), 1319 (m), 1430 (w), 1481 (w), 1651 (m), 1739 (w), 3016 (w) cm−1. 1H NMR (CDCl3, 400.2 MHz, 298 K) δ: 0.33 (s, 12H, SiCH3), 1.05 (s, 4H, SiCH2) [2J(1H−117/119Sn) 76 Hz], 7.36 (m, 4H, H3′, H6′), 7.49 (m, 12H, Hm, Hm′, Hp, Hp′), 7.55 (m, 8H, H2′, H7′), 7.71 (m, 1H, H6″), 7.74 (m, 4H, Ho, Ho′), 7.94 (m, 1H, H5″), 8.97 (m, 1H, H3″) ppm. 13C{1H} NMR (CDCl3, 100.6 MHz, 298 K) δ: 0.0, 0.1 [3J(13C−117/119Sn) = 28 Hz, SiCH3], 1.0 (SiCH2), 125.0 (C6″), 125.8 (C3′, C6′), 128.2 (C4″), 129.0, 129.2 [3J(13C−117/119Sn = 60 Hz, Cm, Cm′], 130.2, 130.4 [4J(13C−117/119Sn) = 12 Hz, Cp, Cp′], 134.3 (C2′, C7′), 136.5, 136.8 [2J(13C−117/119Sn) = 48 Hz, Co, Co′], 137.8, 137.9 (C1′, C8′), 138.1 (C5″), 139.5, 139.9 (C4′, C5′), 140.2, 140.4 (Ci, Ci′), 150.6 (C2″), 151.5 (C3″), 169.5 (C1″, C7″) ppm. 29 Si{ 1 H} NMR (CDCl 3 , 79.5 MHz, 298 K) δ: −2.10 2 29 [ J( Si−117/119Sn) = 44 Hz] ppm. 119Sn{1H} NMR (CDCl3, 149.2 MHz, 298 K) δ: −24, −27 ppm (h1/2 = 4 Hz). Anal. Calcd for MC1· H2O·EtOH, C51H55NO6Si2Sn2 (1071.58 g mol−1): C, 57.2; H, 5.2; N, 1.3. Found: C, 57.7; H, 4.8; N, 1.3. MS (ESI−TOF) for [C49H48NO4Si2Sn2]+ ([M + H]+): m/z (exp.) = 1008.115935, m/z (calcd.) = 1008.117638. Synthesis of MC2. A solution of 1,4-[ClPh2SnCH2Si(CH3)2− C12H8−(CH3)2SiCH2SnPh2Cl] (0.30 g, 0.33 mmol) in methanol (10 mL) was added at room temperature to a previously prepared stirred solution of 3,5-pyridinedicarboxylic acid (0.055 g, 0.33 mmol) and potassium hydroxide (0.037 g, 0.66 mmol) in 15 mL of a solvent mixture of water and ethanol (1:1, v/v). The reaction mixture was stirred overnight. After removal of the solvent, the product was washed with a solvent mixture consisting of water and ethanol (20:3), and dried in vacuo to give a colorless solid in 70% isolated yield (0.23 g, 0.23 mmol). Mp 245−251 °C. IR νmax: 446 (m), 528 (w), 602 (w), 641 (w), 696 (s), 728 (s), 776 (s), 804 (s), 814 (s), 833 (m), 1002 (w), 1076 (w), 1115 (m), 1217 (w), 1229 (w), 1250 (w), 1283 (m), 1316 (m), 1365 (m), 1430 (w), 1481 (w), 1596 (w), 1652 (w), 1739 (m), 2970 (w) cm−1. 1H NMR (CDCl3, 400.2 MHz, 298 K) δ: 0.39 (s, 12H, SiCH3), 1.03 (s, 4H, SiCH2) [2J(1H−117/119Sn) = 76 Hz], 7.21 (m, 4H, H3′), 7.44 (m, 16H, H2′, Hm, Hp), 7.62 (m, 8H, Ho), 8.00 (s, 1H, H4″), 9.21 (s, 2H, H3″) ppm. 13C{1H} NMR (CDCl3, 100.6 MHz, 298 K) δ: −0.1 [3J(13C−117/119Sn) = 26 Hz, SiCH3], 1.5 (SiCH2), 125.8 (C3′), 126.9 (C2″), 129.1 [3J(13C−117/119Sn) = 60 Hz, Cm], 130.3 [4J(13C−117/119Sn) = 12 Hz, Cp], 134.4 (C2′), 136.5 [2J(13C−117/119Sn) = 48 Hz, Co], 138.0 (C1′), 138.4 (C4″), 140.2 (C4′), 140.6 (Ci), 154.9 (C3″), 169.1 (C1″) ppm. 119Sn{1H} NMR (CDCl3, 149.2 MHz, 298 K) δ: −26 ppm (h1/2 = 8 Hz). 29Si{1H} NMR (CDCl3, 79.5 MHz, 298 K) δ: −2.18 [2J(29Si−117/119Sn) = 44 Hz] ppm. Anal. Calcd for C49H47NO4Si2Sn2 (1007.49 g mol−1): C, 58.4; H, 4.7; N, 1.4. Found: C, 57.2; H, 5.0; N, 1.3. MS (FAB+) for [C49H48NO4Si2Sn2]+ ([M + H]+): m/z (exp.) = 1008.11797, m/z (calcd.) = 1008.11764. Synthesis of MC3. A solution of 1,4-[ClPh2SnCH2Si(CH3)2− C12H8−(CH3)2SiCH2SnPh2Cl] (0.20 g, 0.22 mmol) in methanol (10 mL) was added at room temperature to a previously prepared solution of piperazine (0.019 g, 0.22 mmol), carbon disulfide (0.034 g, 0.44 mmol), and potassium hydroxide (0.025 g, 0.44 mmol) in 15 mL of ethanol. The reaction mixture was stirred overnight. After removal of the solvent, the product was washed with a solvent mixture consisting of water and ethanol (20:3), and dried in vacuo to give a colorless solid in 73% yield (0.17 g, 0.16 mmol). Mp 315−320 °C. IR νmax: 419 (m), 444 (m), 450 (m), 485 (m), 506 (m), 521 (m), 558 (w), 605 (m), 695 (s), 726 (s), 802 (s), 813 (s), 833 (m), 917 (m), 994 (s), 1002 (m), 1045 (m), 1114 (m), 1164 (w), 1200 (m), 1251 (m), 1274 (m), 1349 (w), 1410 (s), 1428 (m) cm−1. 1H NMR (CDCl3, 400.2 MHz, 298 K) δ: 0.29 (s, 12H, SiCH3), 1.11 (s, 4H, SiCH2) [2J(1H−117/119Sn) = 78 Hz], 3.74 (s, 8H, H2″), 7.35 (m, 12H, Hm, Hp), 7.56 (m, 8H, Ho), 7.69 (d, 4H, H3′), 7.78 (m, 4H, H2′) ppm. 13 C{1H} NMR (CDCl3, 100.6 MHz, 298 K) δ: 0.5 (SiCH3), 3.2 (SiCH2), 49.9 (C2″), 125.4 (C3′), 128.7 [3J(13C−117/119Sn) = 56 Hz, Cm], 129.3 [4J(13C−117/119Sn) = 14 Hz, Cp], 135.7 (C2′), 136.5 [2J(13C−117/119Sn) = 44 Hz, Co], 139.0 (C1′), 140.3 (C4′), 144.5 (Ci), 197.1 (CS2) ppm. 119Sn{1H} NMR (CDCl3, 149.2 MHz, 298 K) δ: −108 ppm (h1/2 = 4 Hz). 29Si{1H} NMR (CDCl3, 79.5 MHz, 298 K) δ: −1.66 ppm. Anal. Calcd for C48H52N2S4Si2Sn2 (1078.79 g 2454

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics Notes

idealized geometry and refined using a riding model. Figures were created with Diamond.39 In the dinuclear organotin precursor PRE, one of the Sn-phenyl rings is slightly disordered. For the refinement of this substituent, the ISOR instruction implemented in SHELXL37 was employed. In the macrocyclic structure of MC1, one of the Sn-phenyl rings, the pyridine nitrogen of the ligand, and the hydrogen atoms of the water molecule coordinated to Sn1 are disordered. The disorders were treated using the DFIX, EADP, EXYZ, SUMP, and PART instructions. In macrocycles MC4 and MC5, the piperazine moieties and the solvent molecules are disordered. In this case, the disorders were treated using PART, DFIX, DANG, and Uij restraints (SIMU, ISOR, and/or EADP). Crystallographic data for the crystal structures have been deposited with the Cambridge Crystallographic Data Centre as supplementary publications CCDC 1898130−1898136. Computational Studies. DFT quantum chemical calculations were performed with the quantum mechanical program Gaussian 09 (version B.01),40 using the B3LYP hybrid functional41 in combination with the def2-SVP basis set,42 which is recommended for relativistic effects when a heavier element such as Sn is present.43 Harmonic vibrational frequencies were computed with the same method to characterize stationary points as minima. Geometry optimizations for MC2′ and syn/anti-MC4′ were performed using different initial conformations of the molecular structures, based on the structural characteristics established by the X-ray crystallographic analyses of MC1 and MC4, and considering the Car−Si and Si−CH2 bonds as points of rotation. For the compound having the lowest energy, a relative energy of 0.00 kcal/mol is indicated.



The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support from Consejo Nacional de Ciencia y Tecnologiá (CONACyT) through project Nos. 158098, 229929, and 294810 is gratefully acknowledged. Postgraduate fellowships for M.G.V.-R. and I.R.-L. from CONACyT are acknowledged. The authors are thankful for access to Laboratorio Nacional de Estructura de Macromoléculas (LANEM). The authors thank Dr. Wolf Hiller (TU Dortmund) for recording the NMR spectra, and I. Q. Geiser Cuéllar Rivera and M. C. Maria Eugenia Ochoa Becerra for assistance in the mass spectrometric characterization.



(1) (a) Zhang, M.; Saha, M. L.; Wang, M.; Zhou, Z.; Song, B.; Lu, C.; Yan, X.; Li, X.; Huang, F.; Yin, S.; Stang, P. J. Multicomponent Platinum(II) Cages with Tunable Emission and Amino Acid Sensing. J. Am. Chem. Soc. 2017, 139, 5067−5074. (b) Brown, C. J.; Toste, F. D.; Bergman, R. G.; Raymond, K. N. Supramolecular Catalysis in Metal−Ligand Cluster Hosts. Chem. Rev. 2015, 115, 3012−3035. (c) Schmidt, B. M.; Osuga, T.; Sawada, T.; Hoshino, M.; Fujita, M. Compressed Corannulene in a Molecular Cage. Angew. Chem., Int. Ed. 2016, 55, 1561−1564. (d) Choi, K. M.; Kim, D.; Rungtaweevoranit, B.; Trickett, C. A.; Barmanbek, J. T. D.; Alshammari, A. S.; Yang, P.; Yaghi, O. M. Plasmon-Enhanced Photocatalytic CO2 Conversion within Metal−Organic Frameworks under Visible Light. J. Am. Chem. Soc. 2017, 139, 356−362. (e) Percástegui, E. G.; Mosquera, J.; Nitschke, J. R. Anion Exchange Renders Hydrophobic Capsules and Cargoes Water-Soluble. Angew. Chem., Int. Ed. 2017, 56, 9136−9140. (2) (a) Lehn, J. M. From supramolecular chemistry towards constitutional dynamic chemistry and adaptive chemistry. Chem. Soc. Rev. 2007, 36, 151−160. (b) Steed, J. W.; Atwood, J. L.; Gale, P. A. Supramolecular Chemistry; Steed, J. W., Gale, P. A., Eds.; John Wiley & Sons, Ltd.: Hoboken, NJ, 2012; Vol. 1, pp 3−9. (3) (a) Ma, C.; Han, Y.; Zhang, R.; Wang, D. Self-assembly of diorganotin(IV) moieties and 2-pyrazinecarboxylic acid: syntheses, characterizations and crystal structures of monomeric, polymeric or trinuclear macrocyclic compounds. Dalton Trans. 2004, 1832−1840. (b) Prabusankar, G.; Murugavel, R. Hexameric Organotincarboxylates with Cyclic and Drum Structures. Organometallics 2004, 23, 5644− 5647. (c) Baul, T. S. B.; Singh, K. S.; Lyčka, A.; Holčapek, M.; Linden, A. Synthesis of a cyclic dinuclear organotin carboxylate via simultaneous debenzylation and decarbonylation reactions: X-ray crystal structure of [(PhCH2)2{O2CC6H4{N(H)−N(C6H3-4(O)5-O)}-o}Sn]2. J. Organomet. Chem. 2005, 690, 1581−1587. (d) Ma, C.; Zhang, Q.; Zhang, R.; Wang, D. Self-Assembly of Dialkyltin Moieties and Mercaptobenzoic Acid into Macrocyclic Complexes with Hydrophobic “Pseudo-Cage” or Double-Cavity Structures: Supramolecular Infrastructures Involving Intermolecular C−H···S Weak Hydrogen Bonds and π−π Interactions. Chem.Eur. J. 2006, 12, 420−428. (e) García-Zarracino, R.; Höpfl, H. Self-Assembly of Diorganotin(IV) Oxides (R = Me, nBu, Ph) and 2,5-Pyridinedicarboxylic Acid to Polymeric and Trinuclear Macrocyclic Hybrids with Porous Solid-State Structures: Influence of Substituents and Solvent on the Supramolecular Structure. J. Am. Chem. Soc. 2005, 127, 3120− 3130. (f) Bowen, R. J.; Caddy, J.; Fernandes, M. A.; Layh, M.; Mamo, M. A.; Meijboom, R. Synthesis and characterisation of dialkyltin 2,3bis(diphenylphosphino)maleic acid adducts. J. Organomet. Chem. 2006, 691, 717−725. (g) Chandrasekhar, V.; Thirumoorthi, R. 1,1′Ferrocenedicarboxylate-Bridged Redox-Active Organotin and -tellurium-Containing 16-Membered Macrocycles: Synthesis, Structure, and Electrochemistry. Organometallics 2007, 26, 5415−5422. (h) Du, D.; Jiang, Z.; Liu, C.; Sakho, A. M.; Zhu, D.; Xu, L. Macrocyclic organotin(IV) carboxylates based on benzenedicarboxylic acid derivatives: Syntheses, crystal structures and antitumor activities. J.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.organomet.9b00132. IR, NMR (1H, 13C, 29Si, 119Sn), and mass spectra for compounds BB4, MC1, MC2, MC3, MC4, and MC5 (Figures S1−S47). Tables with crystallographic data for compounds BB4, MC1, MC4, and MC5 (Tables S1 and S2). Tables with comparison of selected bonds lengths and angles for the experimental (SCXRD analysis) and DFT geometry-optimized molecular structures of compounds MC4, syn-MC4′, and anti-MC4′ (Table S3) (PDF) Accession Codes

CCDC 1898130−1898136 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 [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



REFERENCES

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. Phone: +49-2317553787. Fax: +49-231-7555048(K.J.). *E-mail: hhopfl@uaem.mx (H.H.). ORCID

Irán Rojas-León: 0000-0002-1289-5732 Klaus Jurkschat: 0000-0001-9930-858X María G. Vasquez-Ríos: 0000-0003-0890-3194 Gelen Gómez-Jaimes: 0000-0001-9555-3262 Herbert Höpfl: 0000-0002-4027-0131 Irán F. Hernández-Ahuactzí: 0000-0001-9238-4010 2455

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics Organomet. Chem. 2011, 696, 2549−2558. (i) González-Rivas, N.; Cuevas-Yañez, E.; Barba, V.; Beltran, H. I.; Reyes, H. Rectangular bimetallic diorganotin macrocycle obtained through a combination of metallosupramolecular chemistry with imine bond formation. Inorg. Chem. Commun. 2013, 37, 110−113. (j) Celis, N. A.; Villamil-Ramos, R.; Höpfl, H.; Hernández-Ahuactzi, I. F.; Sánchez, M.; ZamudioRivera, L. S.; Barba, V. Dinuclear Monomeric and Macrocyclic Organotin Dithiocarbamates Derived from 1,10-Diaza-18-crown-6 and 4,4′-Trimethylenedipiperidine. Eur. J. Inorg. Chem. 2013, 2013, 2912−2922. (k) Xiao, X.; Yan, L.; Mei, Z.; Zhu, D.; Xu, L. Two novel macrocyclic organotin(IV) carboxylates based on amide carboxylic acids. RSC Adv. 2014, 4, 3096−3101. (l) Torres-Huerta, A.; Rodríguez-Molina, B.; Höpfl, H.; Garcia-Garibay, M. A. Synthesis and Solid-State Characterization of Self-Assembled Macrocyclic Molecular Rotors of Bis(dithiocarbamate) Ligands with Diorganotin(IV). Organometallics 2014, 33, 354−362. (m) Torres-Huerta, A.; Cruz-Huerta, J.; Höpfl, H.; Hernández-Vázquez, L.; Escalante-García, J.; Jiménez-Sánchez, A.; Santillan, R.; Hernández-Ahuactzi, I.; Sánchez, M. Variation of the Molecular Conformation, Shape, and Cavity Size in Dinuclear Metalla-Macrocycles Containing HeteroDitopic Dithiocarbamate−Carboxylate Ligands from a Homologous Series of N-Substituted Amino Acids. Inorg. Chem. 2016, 55, 12451− 12469. (n) Zhang, R.-F.; Wang, F.; Li, Q.-L.; Zhang, S.-L.; Ma, C.-L. Synthesis and structural characterization of organotin complexes derived from 2-(trifluoromethyl)benzeneseleninic acid: tetranuclear macrocycle, 1-D polymeric chain, helical double-chain. J. Coord. Chem. 2016, 69, 704−713. (o) Basu Baul, T. S.; Dutta, D.; Duthie, A.; Guedes da Silva, M. F. C. Influencing the outcome: Diorganotin(IV) ladder to macrocycle conversion through solvent selection. Inorg. Chem. Commun. 2017, 84, 68−71. (4) (a) Zobel, B.; Schürmann, M.; Jurkschat, K.; Dakternieks, D.; Duthie, A. Novel Molecular Organotin Oxides Derived from Alkylidene Bridged Ditin Precursors: Syntheses and Structures. Organometallics 1998, 17, 4096−4104. (b) Chandrasekhar, V.; Thilagar, P.; Bickley, J. F.; Steiner, A. Alternating Hydrophilic and Hydrophobic Pockets in the Channel Structures of Organostannoxane Prismanes: Preferential Confinement of Guest Molecules. J. Am. Chem. Soc. 2005, 127, 11556−11557. (c) Reyes-Martinez, R.; García y Garcia, P.; Lopez-Cardoso, M.; Höpfl, H.; Tlahuext, H. Self-assembly of diphenyltin(IV) and tris-dithiocarbamate ligands to racemic trinuclear cavitands and capsules. Dalton Trans. 2008, 6624−6627. (d) Xie, Y.-P.; Ma, J.-F.; Yang, J.; Su, M.-Z. Monorganooxotin cage, diorganotin ladders, diorganotin double chain and triorganotin single chain formed with phosphonate and arsonate ligands. Dalton Trans. 2010, 39, 1568−1575. (5) (a) Xiong, R.-G.; Zuo, J.-L.; You, X.-Z.; Fun, H.-K.; Raj. A Novel Rhombohedral Grid Based on Tetraorganodistannoxane as Corner Unit. Organometallics 2000, 19, 4183−4186. (b) Wang, R.-H.; Hong, M.-C.; Luo, J.-H.; Cao, R.; Weng, J.-B. Synthesis and Crystal Structures of the First Two Novel Dicarboxylate Organotin Polymers Constructed from Dimeric Tetraorganodistannoxane Units. Eur. J. Inorg. Chem. 2002, 2002, 2082−2085. (c) Ma, C.; Han, Y.; Zhang, R.; Wang, D. Self-Assembled Triorganotin(IV) Moieties with 1,3,5Benzenetricarboxylic Acid: Syntheses and Crystal Structures of Monomeric, Helical, and Network Triorganotin(IV) Complexes. Eur. J. Inorg. Chem. 2005, 2005, 3024−3033. (d) Chandrasekhar, V.; Thirumoorthi, R.; Azhakar, R. New Structural Forms of Organostannoxane Macrocycle Networks. Organometallics 2007, 26, 26−29. (e) Ma, C.; Wang, Q.; Zhang, R. Self-Assembly and Characterization of a Novel 2D Network Polymer Containing a 60-Membered Organotin Macrocycle. Inorg. Chem. 2008, 47, 7060−7061. (f) García-Zarracino, R.; Höpfl, H.; Güizado-Rodríguez, M. Bis(tetraorganodistannoxanes) as Secondary Building Block Units (SBUs) for the Generation of Porous Materials − A ThreeDimensional Honeycomb Architecture Containing Adamantane-type Water Clusters. Cryst. Growth Des. 2009, 9, 1651−1654. (g) Zhang, R.; Ren, Y.; Wang, Q.; Ma, C. Syntheses and Characterization of 2D and 3D Organotin Polymers with Phenylsuccinic Acid and Trimethyltin Chloride Under Different Conditions. J. Inorg. Organo-

met. Polym. Mater. 2010, 20, 399−404. (h) Chandrasekhar, V.; Mohapatra, C.; Butcher, R. J. Synthesis of One- and TwoDimensional Coordination Polymers Containing Organotin Macrocycles. Reactions of (n-Bu3Sn)2O with Pyridine Dicarboxylic Acids. Structure-Directing Role of the Ancillary 4,4′-Bipyridine Ligand. Cryst. Growth Des. 2012, 12, 3285−3295. (i) Mohapatra, C.; Chandrasekhar, V. Two-Dimensional Homometallic- to a Three Dimensional Heterometallic Coordination Polymer: A Metalloligand Approach. Cryst. Growth Des. 2014, 14, 406−409. (j) Li, Q.; Liu, X.; Cheng, S.; Zhang, R.; Shi, Y.; Ma, C. Novel organotin complexes derived from 2,2′-selenodiacetic acid: synthesis and biological evaluation. RSC Adv. 2016, 6, 32484−32492. (k) Rocha, C. S.; De Morais, B. P.; Rodrigues, B. L.; Donnici, C. L.; De Lima, G. M.; Ardisson, J. D.; Takahashi, J. A.; Bitzer, R. S. Spectroscopic and X-ray structural characterization of new polymeric organotin(IV) carboxylates and their in vitro antifungal activities: part II. Appl. Organomet. Chem. 2017, 31, e3645. (l) Basu Baul, T. S.; Kehie, P.; Duthie, A.; Höpfl, H. Dinuclear organotin(IV) coordination polymers derived from Schiff Bases with L-aspartic acid. J. Organomet. Chem. 2017, 838, 24−36. (m) Yan, F.-F.; Zhu, Q.; Li, Q.-L.; Zhang, R.-F.; Ma, C.-L. Triorganotin coordination polymers based on three dicarboxylate ligands containing flexible S−S bonds: synthesis, structures, and in vitro anti-tumor activity. J. Organomet. Chem. 2019, 880, 156−162. (n) Topîrlan, A. V.; Patrascu, A. A.; Sava, A.; Popescu, D.-L.; Silvestru, C.; Haiduc, I.; Andruh, M. Coordination polymers constructed from triorganotin(IV) nodes and fumarato spacers. J. Organomet. Chem. 2019, 882, 58−63. (o) Basu Baul, T. S.; Chaurasiya, A.; Duthie, A.; Vasquez-Rı ́os, M. G.; Höpfl, H. Solvent and intermolecular nitrogen coordination dictated formation of selfassembled organostannane-macrocycles based on monomers and coordination polymers with unsymmetrical 5-[(E))-2-(4-pyridyl)-1diazenyl]-2-hydroxybenzoate ligand: Structural topologies and dimensionality. J. Organomet. Chem. 2018, 872, 87−101. (p) Shankar, R.; Jain, A.; Singh, A. P.; Kociok-Köhn, G.; Molloy, K. C. Facile Synthesis of Novel Two- And Three-Dimensional Coordination Polymers Containing Dialkyltin Phosphonate-Based Tri/TetraNuclear Clusters with Appended Sulfonate Groups. Inorg. Chem. 2009, 48, 3608−3616. (q) Shankar, R.; Singla, N.; Asija, M.; KociokKöhn, G.; Molloy, K. C. Synthesis and Structural Studies of Diorganotin(IV)-Based Coordination Polymers Bearing Silaalkylphosphonate Ligands and Their Transformation into Colloidal Domains. Inorg. Chem. 2014, 53, 6195−6203. (r) Shankar, R.; Jain, A.; Kociok-Köhn, G.; Molloy, K. C. Diorganotin-Based Coordination Polymers Derived from Sulfonate/Phosphonate/Phosphonocarboxylate Ligands. Inorg. Chem. 2011, 50, 1339−1350. (s) Shankar, R.; Mendiratta, S.; Singla, N.; Kociok-Köhn, G.; Lutter, M.; Jurkschat, K. Reactivity of Elemental Tin and Zinc toward Organophosphonic Acid Dialkyl Esters: A New One-Pot Recipe for the Synthesis of Coordination Assemblies Derived from O-Alkylorganophosphonate Ligands. Inorg. Chem. 2017, 56, 721−724. (6) (a) Hamilton, T. D.; MacGillivray, L. R. Enclosed Chiral Environments from Self-Assembled Metal−Organic Polyhedra. Cryst. Growth Des. 2004, 4, 419−430. (b) Fiedler, D.; Leung, D. H.; Bergman, R. G.; Raymond, K. N. Selective Molecular Recognition, C−H Bond Activation, and Catalysis in Nanoscale Reaction Vessels. Acc. Chem. Res. 2005, 38, 349−358. (c) Yu, G.; Zhang, M.; Saha, M. L.; Mao, Z.; Chen, J.; Yao, Y.; Zhou, Z.; Liu, Y.; Gao, C.; Huang, F.; Chen, X.; Stang, P. J. Antitumor Activity of a Unique Polymer That Incorporates a Fluorescent Self-Assembled Metallacycle. J. Am. Chem. Soc. 2017, 139, 15940−15949. (d) Wang, S.; Sawada, T.; Ohara, K.; Yamaguchi, K.; Fujita, M. Capsule−Capsule Conversion by Guest Encapsulation. Angew. Chem., Int. Ed. 2016, 55, 2063−2066. (7) (a) Tiekink, E. R. T. Structural chemistry of organotin carboxylates: a review of the crystallographic literature. Appl. Organomet. Chem. 1991, 5, 1−23. (b) Tiekink, E. R. T. Tin dithiocarbamates: Applications and Structures. Appl. Organomet. Chem. 2008, 22, 533−550. (c) Gupta, A. N.; Kumar, V.; Singh, V.; Rajput, A.; Prasad, L. B.; Drew, M. G. B.; Singh, N. Influence of functionalities on the structure and luminescent properties of 2456

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics organotin(IV) dithiocarbamate complexes. J. Organomet. Chem. 2015, 787, 65−72. (d) Mahato, M.; Mukherji, S.; Van Hecke, K.; Harms, K.; Ghosh, A.; Nayek, H. P. Mononuclear homoleptic organotin(IV) dithiocarbamates: Synthesis, structures and antimicrobial activities. J. Organomet. Chem. 2017, 853, 27−34. (e) De Morais, B. P.; Donnici, C. L.; Rodrigues, B. L.; De Lima, G. M.; Wardell, J. L.; Bitzer, R. S. Transesterification reactions of dimethoxycarbonylethyltin-pyrrolidinodithiocarbamate, [Sn(PDTC)2(CH2CH2COOCH3)2]−Spectroscopic and X-ray structural characterization of new estertin dithiocarbamate complexes−part I. J. Organomet. Chem. 2017, 832, 57−65. (f) De Morais, B. P.; Donnici, C. L.; Rodrigues, B. L.; De Lima, G. M.; Wardell, J. L.; Nascimento, M. A. C.; Bitzer, R. S. Transesterification reactions of dimethoxycarbonylethyltinmorpholinodithiocarbamate, [(MeO 2CCH2CH 2)2Sn(MDTC) 2]: Synthesis, spectroscopy, X-ray structural characterization and DFT calculations of new diestertin dithiocarbamate complexes−part II. Polyhedron 2017, 135, 161−168. (g) Claudio-Catalán, M. A.; Medrano, F.; Rivera-Márquez, K. I.; Rodríguez-Uribe, N. A.; PérezGonzález, R.; Tlahuext, H.; Godoy-Alcántar, C. Polyhedron 2018, 146, 55−64. (h) Pandey, M. D.; Metre, R. K.; Kundu, S.; Mahanti, B.; Kumar, A.; Gopal, K.; Chandrasekhar, V. Luminiscent PyreneDecorated Organotin Compounds: Observation of Monomer and Excimer Emission. Cryst. Growth Des. 2019, 19, 1888−1895. (8) (a) Gielen, M.; El Khloufi, A.; Biesemans, M.; Kayser, F.; Willem, R.; Mahieu, B.; Maes, D.; Lisgarten, J. N.; Wyns, L.; Moreira, A.; Chattopadhay, T. K.; Palmer, R. A. Tri-n-butyltin 2,6Difluorobenzoate, a Unique Macrocyclic Tetramer Containing a 16Membered Sn4C4O8 Ring, and Other Related Compounds. Organometallics 1994, 13, 2849−2854. (b) Basu Baul, T. S.; Dutta, D.; Duthie, A.; Prasad, R.; Kumar Rana, N.; Koch, B.; Tiekink, E. R. T. Triphenyltin(IV) benzoates with diazenyl/imino scaffold exhibiting remarkable apoptosis mediated by reactive oxygen species. J. Inorg. Biochem. 2017, 173, 79−92. (9) (a) Brunet, P.; Simard, M.; Wuest, J. D. Molecular Tectonics. Porous Hydrogen-Bonded Networks with Unprecedented Structural Integrity. J. Am. Chem. Soc. 1997, 119, 2737−2738. (b) Hosseini, M. W. Molecular Tectonics: From Simple Tectons to Complex Molecular Networks. Acc. Chem. Res. 2005, 38, 313−323. (10) (a) Haasnoot, J. G. Mononuclear, oligonuclear and polynuclear metal coordination compounds with 1,2,4-triazole derivatives as ligands. Coord. Chem. Rev. 2000, 200−202, 131−185. (b) Eddaoudi, M.; Moler, D. B.; Li, H.; Chen, B.; Reineke, T. M.; O’Keeffe, M.; Yaghi, O. M. Modular Chemistry: Secondary Building Units as a Basis for the Design of Highly Porous and Robust Metal−Organic Carboxylate Frameworks. Acc. Chem. Res. 2001, 34, 319−330. (c) Abdelbaky, M. S. M.; Amghouz, Z.; Fernández-Zapico, E.; García-Granda, S.; García, J. R. Metal−organic frameworks assembled from lanthanide and 2,5-pyridinedicarboxylate with cubane-like [Ln4(OH)4] building units. J. Solid State Chem. 2015, 229, 197− 207. (d) Kim, J.; Oliver, A. G.; Neumann, G. T.; Hicks, J. C. ZnMOFs Containing Pyridine and Bipyridine Carboxylate Organic Linkers and Open Zn2+ Sites. Eur. J. Inorg. Chem. 2015, 2015, 3011− 3018. (e) Cai, S.-L.; Zheng, S.-R.; Fan, J.; Zeng, R.-H.; Zhang, W.-G. Two new three-dimensional metal−organic frameworks with 4connected diamondoid and unusual (6,16)-connected net topologies based on planar tetranuclear squares as secondary building units. CrystEngComm 2016, 18, 1174−1183. (11) (a) Hartshorn, C. M.; Steel, P. J. Self-assembly and X-ray structure of a ten-component, three-dimensional metallosupramolecular cage. Chem. Commun. 1997, 541−542. (b) Stang, P. J.; Olenyuk, B. Self-Assembly, Symmetry, and Molecular Architecture: Coordination as the Motif in the Rational Design of Supramolecular Metallacyclic Polygons and Polyhedra. Acc. Chem. Res. 1997, 30, 502−518. (c) Batten, S. R.; Robson, R. Interpenetrating Nets: Ordered, Periodic Entanglement. Angew. Chem., Int. Ed. 1998, 37, 1460−1494. (d) Caulder, D. L.; Raymond, K. N. Supermolecules by Design. Acc. Chem. Res. 1999, 32, 975−982. (e) Fujita, M.; Umemoto, K.; Yoshizawa, M.; Fujita, N.; Kusukawa, T.; Biradha, K. Molecular paneling via coordination. Chem. Commun. 2001, 509−518.

(f) Mirtschin, S.; Slabon-Turski, A.; Scopelliti, R.; Velders, A. H.; Severin, K. A Coordination Cage with an Adaptable Cavity Size. J. Am. Chem. Soc. 2010, 132, 14004−14005. (g) Komine, S.; Takahashi, S.; Kojima, T.; Sato, H.; Hiraoka, S. Self-Assembly Processes of Octahedron-Shaped Pd6L4 Cages. J. Am. Chem. Soc. 2019, 141, 3178− 3186. (12) (a) Jelfs, K. E.; Eden, E. G. B.; Culshaw, J. L.; Shakespeare, S.; Pyzer-Knapp, E. O.; Thompson, H. P. G.; Bacsa, J.; Day, G. M.; Adams, D. J.; Cooper, A. I. In silico Design of Supramolecules from Their Precursors: Odd−Even Effects in Cage-Forming Reactions. J. Am. Chem. Soc. 2013, 135, 9307−9310. (b) Klotzbach, S.; Scherpf, T.; Beuerle, F. Dynamic covalent assembly of tribenzotriquinacenes into molecular cubes. Chem. Commun. 2014, 50, 12454−12457. (c) Ono, K.; Johmoto, K.; Yasuda, N.; Uekusa, H.; Fujii, S.; Kiguchi, M.; Iwasawa, N. Self-Assembly of Nanometer-Sized Boroxine Cages from Diboronic Acids. J. Am. Chem. Soc. 2015, 137, 7015−7018. (d) Wang, R.-N.; Zhang, X.-R.; Wang, S.-F.; Fu, G.-S.; Wang, J.-L. Flatbands in 2D boroxine-linked covalent organic frameworks. Phys. Chem. Chem. Phys. 2016, 18, 1258−1264. (e) Waller, P. J.; Lyle, S. J.; Popp, T. M. O.; Diercks, C. S.; Reimer, J. A.; Yaghi, O. M. Chemical Conversion of Linkages in Covalent Organic Frameworks. J. Am. Chem. Soc. 2016, 138, 15519−15522. (f) Beaudoin, D.; Rominger, F.; Mastalerz, M. Chiral Self-Sorting of [2 + 3] Salicylimine Cage Compounds. Angew. Chem., Int. Ed. 2017, 56, 1244−1248. (g) Cao, N.; Wang, Y.; Zheng, X.; Jiao, T.; Li, H. Controllable Self-Assembly of Pills and Cages via Imine Condensation for Silver Cation Detection. Org. Lett. 2018, 20, 7447−7450. (h) Jiao, T.; Wu, G.; Chen, L.; Wang, C.-Y.; Li, H. Precursor Control over the Self-Assembly of Organic Cages via Imine Condensation. J. Org. Chem. 2018, 83, 12404−12410. (i) Pugh, C. J.; Santolini, V.; Greenaway, R. L.; Little, M. A.; Briggs, M. E.; Jelfs, K. E.; Cooper, A. I. Cage Doubling: Solvent-Mediated Re-equilibration of a [3 + 6] Prismatic Organic Cage to a Large [6 + 12] Truncated Tetrahedron. Cryst. Growth Des. 2018, 18, 2759−2764. (j) Turcani, L.; Greenaway, R. L.; Jelfs, K. E. Machine Learning for Organic Cage Property Prediction. Chem. Mater. 2019, 31, 714−727. (13) (a) Kraus, C. A.; Eatough, H. The Reactions of Sodium Triphenylstannide with Polyhalogenated Methanes. J. Am. Chem. Soc. 1933, 55, 5014−5016. (b) Zimmer, H.; Moslé, H.-G. Notiz über die Reaktion zwischen Triphenylzinnchlorid und Di-lithium−bzw. Di− Grignard−Verbindungen. Chem. Ber. 1954, 87, 1255−1257. (c) Davies, A. G.; Tse, M.-W.; Kennedy, J. D.; McFarlane, W.; Pyne, G. S.; Ladd, M. F. C.; Povey, D. C. Formation and X-Ray Crystal Structure of 1,1,6,6-tetraphenyl-1,6-distannacyclodecane. J. Chem. Soc., Chem. Commun. 1978, 791−792. (d) Hawker, D. W.; Wells, P. R. Some reactions of tris(trimethylstannyl)- and tetrakis(trimethylstannyl)methane. J. Organomet. Chem. 1984, 266, 37−44. (e) Austin, M.; Gebreyes, K.; Kuivila, H. G.; Swami, K.; Zubieta, J. A. Complexes of Bis(chlorodimethylstannyl)- and Bis(dichloromethylstannyl)methanes with Aromatic Nitrogen Heterocycles. Organometallics 1987, 6, 834−842. (f) Jurkschat, K.; Kuivila, H. G.; Liu, S.; Zubieta, J. A. 1,5,9-Tristannacyclododecanes as Lewis Acids. Novel Structure of a Chloride Complex. Organometallics 1989, 8, 2755− 2759. (g) Jurkschat, K.; Hesselbarth, F.; Dargatz, M.; Lehmann, J.; Kleinpeter, E.; Tzschach, A.; Meunier-Piret, J. 1,2-Bis(organostannyl)ethanes as powerful bidentate Lewis acids. Crystal structures of (Ph2ClSnCH2)2·(Me2N)2PO and [Ph3PNPPh3][(Ph2ClSnCH2)2·Cl]. J. Organomet. Chem. 1990, 388, 259−271. (h) Chaniotakis, N. A.; Jurkschat, K.; Rühlemann, A. Potentiometric phosphate selective electrode based on a multidendate−tin(IV) carrier. Anal. Chim. Acta 1993, 282, 345−352. (i) Mehring, M.; Schürmann, M.; Reuter, H.; Dakternieks, D.; Jurkschat, K. Trimethylene-Bridged Triand Tetratin Compounds as Building Blocks for Unusual Double and Triple Ladders. Angew. Chem., Int. Ed. Engl. 1997, 36, 1112−1114. (j) Hillwig, R.; Harms, K.; Dehnicke, K. Z. Die Kristallstruktur von Tris(trimethylstannyl)acetonitril. Z. Naturforsch., B: J. Chem. Sci. 1997, 52b, 145−148. (k) Steenwinkel, P.; Jastrzebski, J. T. B. H.; Deelman, B.-J.; Grove, D. M.; Kooijman, H.; Veldman, N.; Smeets, W. J. J.; Spek, A. L.; van Koten, G. Hypercoordinate Aryltrialkylsilanes and -stannanes and Their Use in the Synthesis of Homodinuclear 2457

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics Organometallic Complexes with a 1,4-Phenylene Bridge. Organometallics 1997, 16, 5486−5498. (l) Altmann, R.; Jurkschat, K.; Schürmann, M.; Dakternieks, D.; Duthie, A. Methylene-Bridged Triand Tetratin Compounds as Lewis Acids. Organometallics 1997, 16, 5716−5723. (m) Altmann, R.; Jurkschat, K.; Schürmann, M.; Dakternieks, D.; Duthie, A. o-Bis(haloorganostannyl)benzenes As Powerful Bidentate Lewis Acids toward Halide Ions. Organometallics 1998, 17, 5858−5866. (n) Schulte, M.; Schürmann, M.; Dakternieks, D.; Jurkschat, K. Tetraorganodistannoxanes: Formation of a Novel cis-ladder. Chem. Commun. 1999, 1291−1292. (o) Zobel, B.; Duthie, A.; Dakternieks, D.; Tiekink, E. R. T. α,ω-Bis(trichlorostannyl)alkanes as Bis(monodentate) Lewis Acids toward Halide Ions. Organometallics 2001, 20, 3347−3350. (p) Wang, Z.-H.; Tang, L.F.; Jia, W.-L.; Wang, J.-T.; Wang, H.-G. Synthesis and characterization of bidentate organotin Lewis acid derivatives containing bis(pyrazol1-yl)alkanes: X-ray crystal structure of PhBr2(L)SnCH2SnPhBr2 (L = bis(4-chloropyrazol-1-yl)methane). Polyhedron 2002, 21, 873−878. (q) Schulte, M.; Gabriele, G.; Schürmann, M.; Jurkschat, M. K.; Duthie, A.; Dakternieks, D. Novel 10- and 20-Membered Tin- and Silicon-Containing Rings: Synthesis, Complexation Behavior, and Conversion into a Lewis Acidic Polymer. Organometallics 2003, 22, 328−336. (r) Thodupunoori, S. K.; Alamudun, I. A.; Cervantes-Lee, F.; Gomez, F. D.; Carrasco, Y. P.; Pannell, K. H. Synthesis, structures and preliminary biological screening of bis(diphenyl)chlorotin complexes and adducts: Ph2ClSn−CH2−R−CH2−SnClPh2, R = pC6H4, CH2CH2. J. Organomet. Chem. 2006, 691, 1790−1796. (s) Aksu, Y.; Aksu, S.; Schumann, H. Synthesis and characterization of α,ω-bis[tri-(w-triphenylstannyl)butylstannyl] alkanes as starting materials for organotin dendrimers. Appl. Organomet. Chem. 2007, 21, 521−530. (t) Nayek, H. P.; Niedermeyer, H.; Dehnen, S. Preparation and Conformation of Organo-Bridged Bis[tris(arylchalcogenolato)tin]Compounds−An Experimental and Quantum Chemical Study. Z. Anorg. Allg. Chem. 2008, 634, 2805−2810. (u) Bettens, R. P. A.; Dakternieks, D.; Duthie, A.; Kuan, F. S.; Tiekink, E. R. T. The influence of crystal packing effects upon the molecular structures of Ph 3 Sn(CH 2 ) n SnPh 3 , n = 1 to 8, as determined by X-ray crystallography and DFT molecular orbital calculations. Supramolecular aggregation patterns sustained by C−H···π interactions. CrystEngComm 2009, 11, 1362−1372. (v) Wendji, A. S.; Dietz, C.; Kühn, S.; Lutter, M.; Schollmeyer, D.; Hiller, W.; Jurkschat, K. Silicon- and Tin-Containing Open-Chain and Eight-Membered-Ring Compounds as Bicentric Lewis Acids toward Anions. Chem.Eur. J. 2016, 22, 404−416. (w) Wendji, A. S.; Lutter, M.; Stratmann, L. M.; Jurkschat, K. Syntheses, Structures, and Complexation Studies of Tris(organostannyl)methane Derivatives. ChemistryOpen 2016, 5, 554−565. (14) (a) Newcomb, M.; Azuma, Y.; Courtney, A. R. Macrocycles Containing Tin. Two Syntheses of 1,1,6,6,11,11,16,16-octaphenyl1,6,11,16-tetrastannacycloeicosane and a Synthesis of 1,1,6,6tetraphenyl-1,6-distannacyclodecane. Organometallics 1983, 2, 175− 177. (b) Newcomb, M.; Blanda, M. T.; Azuma, Y.; Delord, T. J. Macrocycles Containing Tin. Synthesis and Structure of 1,10Diphenyl-1,10-distannabicyclo[8.8.8]hexacosane. J. Chem. Soc., Chem. Commun. 1984, 1159−1160. (c) Azuma, Y.; Newcomb, M. Macrocycles Containing Tin. Syntheses of Symmetrical Macrocycles Containing Two or Four Diphenylstanna Units. Organometallics 1984, 3, 9−14. (d) Newcomb, M.; Horner, J. H.; Blanda, M. T. Macrocycles Containing Tin. Through Space Cooperative Binding and High Size Selectivity in the Complexation of Chloride Ion by Lewis Acidic Macrobicyclic Hosts. J. Am. Chem. Soc. 1987, 109, 7878−7879. (e) Newcomb, M.; Madonik, A. M.; Blanda, M. T.; Judice, J. K. Macrocycles Containing Tin. 119Sn NMR Studies of Chloride Binding by Lewis Acidic Tin Compounds. Organometallics 1987, 6, 145−150. (f) Blanda, M. T.; Horner, J. H.; Newcomb, M. Macrocycles Containing Tin. The Preparation of Macrobicyclic Lewis Acidic Hosts Containing Two Tin Atoms and 119Sn NMR Studies of Their Chloride and Bromide Binding Properties in Solution. J. Org. Chem. 1989, 54, 4626−4636. (g) Horner, J. H.; Newcomb, M. Parameters for Force Field Calculations of Compounds Containing Tetrahedral

Tin. Organometallics 1991, 10, 1732−1740. (h) Horner, J. H.; Squatritto, P. J.; McGuire, N.; Riebenspies, J. P.; Newcomb, M. Macrocycles Containing Tin. Crystal Structures and Molecular Mechanics Calculated Structures of Macrobicyclic Lewis Acidic Hosts. Organometallics 1991, 10, 1741−1750. (15) (a) Elhamzaoui, H.; Jousseaume, B.; Riague, H.; Toupance, T.; Dieudonné, P.; Zakri, C.; Maugey, M.; Allouchi, H. Self-Assembled Tin-Based Bridged Hybrid Materials. J. Am. Chem. Soc. 2004, 126, 8130−8131. (b) Elhamzaoui, H.; Jousseaume, B.; Toupance, T.; Zakri, C.; Biesemans, M.; Willem, R.; Allouchi, H. A doubly folded spacer in a self-assembled hybrid material. Chem. Commun. 2006, 1304−1306. (c) Toupance, T.; El Hamzaoui, H.; Jousseaume, B.; Riague, H.; Saadeddin, I.; Campet, G.; Brö t z, J. Bridged Polystannoxane: A New Route toward Nanoporous Tin Dioxide. Chem. Mater. 2006, 18, 6364−6372. (d) Toupance, T.; de Borniol, M.; El Hamzaoui, H.; Jousseaume, B. Alkynylorganotins, versatile precursors of class II hybrid materials. Appl. Organomet. Chem. 2007, 21, 514−520. (16) (a) Gausset, O.; Delpon-Lacaze, G.; Schürmann, M.; Jurkschat, K. 1,1,3,3,14,14,16,16,18,18,29,29-Dodecamethyl-3,14,18,29-tetrasila1,16-distanna[5.5]ferrocenophane. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1998, C54, 1425−1427. (b) Altmann, R.; Gausset, O.; Horn, D.; Jurkschat, K.; Schürmann, M.; Fontani, M.; Zanello, P. Novel Silicon- and Tin-Containing Ferrocenophanes and Related Compounds as Lewis Acids. Organometallics 2000, 19, 430−443. (c) Dakternieks, D.; Jurkschat, K.; Schollmeyer, D.; Wu, H. Synthesis and Structure of the First Alkyl-Bridged Double Ladder: {[R(Cl)Sn(CH2)3Sn(Cl)R]O}4 (R = Me3SiCH2). Organometallics 1994, 13, 4121−4123. (d) Mehring, M.; Schürmann, M.; Paulus, I.; Horn, D.; Jurkschat, K.; Orita, A.; Otera, J.; Dakternieks, D.; Duthie, A. Spacerbridged ladder compounds−syntheses, structures and synthetic applications. J. Organomet. Chem. 1999, 574, 176−192. (e) Dakternieks, D.; Kuan, F. S.; Duthie, A.; Tiekink, E. R. T. The crystal structure of the propylene-bridged double ladder, {[Me3SiCH2Sn(Cl)]2O}{CH2CH2CH2}4{[Me3SiCH2Sn(Cl)]2O} as its toluene solvate (1/1). Main Group Met. Chem. 2000, 23, 731−732. (f) Mehring, M.; Paulus, I.; Zobel, B.; Schürmann, M.; Jurkschat, K.; Duthie, A.; Dakternieks, D. The First Spacer-Bridged Tetraorganodistannoxanes with a Mixed Double Ladder Structure. Eur. J. Inorg. Chem. 2001, 2001, 153−160. (g) Dakternieks, D.; Duthie, A.; Zobel, B.; Jurkschat, K.; Schürmann, M.; Tiekink, E. R. T. Extending Distannoxane Double Ladders Using Rigid Spacers: A Double Ladder with Eight Chiral Tin Atoms − and a Twist. Organometallics 2002, 21, 647−652. (h) Mehring, M.; Gabriele, G.; Hadjikakou, S.; Schürmann, M.; Dakternieks, D.; Jurkschat, K. Solvent-controlled assembling by hydrogen bridges and halogenhalogen interactions of novel organotin oxo clusters. Chem. Commun. 2002, 834−835. (i) Beckmann, J.; Dakternieks, D.; Duthie, A.; Kuan, F. S.; Tiekink, E. R. T. Synthesis and structure of an ether-bridged double ladder compound: potential in host−guest chemistry. J. Organomet. Chem. 2003, 688, 56−61. (j) Dakternieks, D.; Zobel, B.; Jurkschat, K.; Schürmann, M.; Tiekink, E. R. T. A Spacer-Linked Molecular Diorganotin Oxide. Organometallics 2003, 22, 1343−1345. (k) Beckmann, J.; Dakternieks, D.; Duthie, A.; Kuan, F. S.; Tiekink, E. R. T. Hydrolysis of Dinuclear Spacer-Bridged Diorganotin(IV) Triflates. A Novel Cationic Double Ladder with Supramolecular Association. Organometallics 2003, 22, 4399−4404. (l) Beckmann, J.; Dakternieks, D.; Duthie, A.; Kuan, F. S.; Jurkschat, K.; Schürmann, M.; Tiekink, E. R. T. Synthesis and structures of new oligomethylenebridged double ladders. How far can the layers be separated? New J. Chem. 2004, 28, 1268−1276. (17) Vasko, P.; Wang, S.; Tuononen, H. M.; Power, P. P. Addition of ethylene or hydrogen to a main-group metal cluster under mild conditions. Angew. Chem., Int. Ed. 2015, 54, 3802−3805. (18) (a) Fard, Z. H.; Halvagar, M. R.; Dehnen, S. Directed Formation of an Organotin Sulfide Cavitand and Its Transformation into a Rugby-Ball-like Capsule. J. Am. Chem. Soc. 2010, 132, 2848− 2849. (b) Halvagar, M. R.; Fard, Z. H.; Dehnen, S. Controlling the Architecture of Discrete Organotin Sulfide Molecules by Optimiza2458

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics tion of the Intramolecular Spacer Length. Chem.Eur. J. 2011, 17, 4371−4374. (c) Hanau, K.; Rinn, N.; Argentari, M.; Dehnen, S. Organotin Selenide Clusters and Hybrid Capsules. Chem.Eur. J. 2018, 24, 11711−11716. (19) Rojas-León, I.; Alnasr, H.; Jurkschat, K.; Vasquez-Ríos, M. G.; Hernández-Ahuactzi, I. F.; Höpfl, H. Molecular Tectonics with Diand Trinuclear Organotin Compounds. Chem.Eur. J. 2018, 24, 4547−4551. (20) (a) Gielen, M.; Jurkschat, K. Synthesis and Spectroscopic Investigations of bis(organostannyl)methanes. J. Organomet. Chem. 1984, 273, 303−312. (b) Dakternieks, D.; Jurkschat, K.; Tozer, R.; Hook, J.; Tiekink, E. R. T. Novel Coordination Isomerization in Organotin(IV) Compounds. Synthesis, Molecular Structures, and NMR Studies of LSnPhX2 (X = Ph, Cl, Br, I, SPh), LCH2SnPhX2 (X = Ph, Cl, Br, I), and LSiPh3, Where LH Is (2-MeO-3-tBu-5-MeC6H2)2CH2. Organometallics 1997, 16, 3696−3706. (21) (a) Buntine, M. A.; Kosovel, F. J.; Tiekink, E. R. T. Supramolecular Sn···Cl associations in diorganotin dichlorides and their influence on molecular geometry as studied by ab initio molecular orbital calculations. CrystEngComm 2003, 5, 331−336. (b) Gholivand, K.; Gholami, A.; Schenk, K. J.; Esrafili, M. D.; Farshadfar, K. Supramolecular assemblies of organotin(IV)−diphosphoryl adducts: insights from X-rays and DFT. RSC Adv. 2015, 5, 98610−98617. (22) (a) García-Zarracino, R.; Höpfl, H. A 3D Hybrid Network Containing Large Spherical Cavities Formed through a Combination of Metal Coordination and Hydrogen Bonding. Angew. Chem., Int. Ed. 2004, 43, 1507−1511. (b) García-Zarracino, R.; Höpfl, H. Structural characterization of dimethyl- and di-n-butyltin(IV) 2,3-pyridinedicarboxylate in solution and in the solid state. Appl. Organomet. Chem. 2005, 19, 451−457. (c) Gielen, M.; Bouâlam, M.; Tiekink, E. R. T. Structural Chemistry of Organotin Carboxylates. XIX. Crystal Structure of Diphenyltin Bis(Picolinate). Main Group Met. Chem. 1993, 16, 251−256. (d) Jurkschat, K.; Tiekink, E. R. T. Structural Chemistry of Organotin Carboxylates. XXI. Crystal Structure of Ditert-butyltin Bis(Picolinate). Main Group Met. Chem. 1994, 17, 659− 664. (e) Shankar, R.; Kumar, M.; Narula, S. P.; Chadha, R. K. Disproportionation reactions of (methoxy/hydroxy)diorganotin(IV) methanesulfonates with carboxylic acids: Synthesis and structure of new diorganotin(IV) carboxylates. J. Organomet. Chem. 2003, 671, 35−42. (f) Yin, H. D.; Wang, Q. B.; Xing, Q. J. Synthesis and Characterization of Trialkyltin Esters of Pyridinecarboxylic Acids and Crystal Structure of [nBu3SnO2CC5H4N-2]∞. Heteroat. Chem. 2004, 15, 524−529. (g) Corona-Bustamante, A.; Viveros-Paredes, J. M.; Flores-Parra, A.; Peraza-Campos, A. L.; Martínez-Martínez, F. J.; Sumaya-Martínez, M. T.; Ramos-Organillo, A. Antioxidant Activity of Butyl- and Phenylstannoxanes Derived from 2-, 3- and 4Pyridinecarboxylic Acids. Molecules 2010, 15, 5445−5459. (23) (a) Ma, C.; Li, J.; Zhang, R.; Wang, D. Syntheses and characterization of triorganotin complexes: X-ray crystallographic study of triorganotin pyridinedicarboxylates with trinuclear, 1D polymeric chain and 2D network structures. J. Organomet. Chem. 2006, 691, 1713−1721. (b) Chandrasekhar, V.; Mohapatra, C.; Butcher, R. J. Synthesis of One- and Two-Dimensional Coordination Polymers Containing Organotin Macrocycles. Reactions of (nBu3Sn)2O with Pyridine Dicarboxylic Acids. Structure-Directing Role of the Ancillary 4,4′-Bipyridine Ligand. Cryst. Growth Des. 2012, 12, 3285−3295. (c) Kuang, D.-Z.; Yu, J.-X.; Feng, Y.-L.; Zhu, X.-M.; Jiang, W.-J.; Zhang, F.-X. Syntheses, Structures and in Vitro Antitumor Activity of Bis(tricyclohexyltin) Pyridinedicarboxylate with Macrocyclic Supramolecular Structure. Chin. J. Inorg. Chem. 2018, 34, 1035−1042. (24) García-Zarracino, R.; Ramos-Quiñones, J.; Höpfl, H. SelfAssembly of Dialkyltin(IV) Moieties and Aromatic Dicarboxylates to Complexes with a Polymeric or a Discrete Trinuclear Macrocyclic Structure in the Solid State and a Mixture of Fast Interchanging Cyclooligomeric Structures in Solution. Inorg. Chem. 2003, 42, 3835− 3845.

(25) (a) Glowacki, A.; Huber, F.; Preut, H. Preparation and structure of triorganotin, triorganolead and tetraphenylantimony derivatives of acetylenedicarboxylic acid, terephthalic acid and dithioterephthalic acid. Recl. Trav. Chim. Pays-Bas 1988, 107, 278− 285. (b) Hernández-Ahuactzi, I. F.; Cruz-Huerta, J.; Tlahuext, H.; Barba, V.; Guerrero-Alvarez, J.; Höpfl, H. Tuning the Supramolecular Structure through Variation of the Ligand Geometry and Metal Substituents−Diorganotin Macrocycles and Coordination Polymers Derived from cis- and trans-1,2-, 1,3-, and 1,4-Cyclohexanedicarboxylic and cis,cis-1,3,5-Cyclohexanetricarboxylic Acid. Cryst. Growth Des. 2015, 15, 829−847. (26) Wrackmeyer, B. 119Sn-NMR Parameters. Annu. Rep. NMR Spectrosc. 1985, 16, 73−186. (27) (a) Buntine, M. A.; Hall, V. J.; Kosovel, F. J.; Tiekink, E. R. T. Influence of Crystal Packing on Molecular Geometry: A Crystallographic and Theoretical Investigation of Selected Diorganotin Systems. J. Phys. Chem. A 1998, 102, 2472−2482. (b) Naseh, M.; Sedaghat, T.; Tarassoli, A.; Shakerzadeh, E. DFT studies of ONO Schiff bases, their anions and diorganotin(IV) complexes: Tautomerism, NBO and AIM analysis. Comput. Theor. Chem. 2013, 1005, 53− 57. (c) Gholivand, K.; Rajabi, M.; Dorosti, N.; Molaei, F. Synthesis, structural characterization, density functional theory calculations and biological evaluation of a new organotin(IV) complex containing Nisonicotinyl-N′,N″-diaryl phosphorictriamide as N-donor ligand. Appl. Organomet. Chem. 2015, 29, 739−745. (d) Kadu, R.; Roy, H.; Singh, V. K. Diphenyltin(IV) dithiocarbamate macrocyclic scaffolds as potent apoptosis inducers for human cancer HEP 3B and IMR 32 cells: synthesis, spectral characterization, density functional theory study and in vitro cytotoxicity. Appl. Organomet. Chem. 2015, 29, 746−755. (e) Vasquez-Ríos, M. G.; Reyes-Márquez, V.; Höpfl, H.; Torres-Huerta, A.; Guerrero-Á lvarez, J.; Sánchez, M.; HernándezAhuactzi, I. F.; Ochoa-Lara, K.; Jiménez-Sánchez, A.; Santillán, R. 23and 27-Membered Macrocyclic Diorganotin(IV) Bis-dithiocarbamates: Synthesis, Spectroscopic Characterization, DFT Calculations, and Physicochemical Analysis as Anion Receptors. Eur. J. Inorg. Chem. 2016, 2016, 3429−3440. (28) (a) Jung, O.-S.; Sohn, Y. S.; Ibers, J. A. Synthesis and Properties of Ethylenebis(dithiocarbamato)diorganotin(IV) Complexes. Structure of [(t-Bu)2Sn(ebdtc)]2.4THF. Inorg. Chem. 1986, 25, 2273− 2275. (b) Dakternieks, D.; Zhu, H.; Masi, D.; Mealli, C. Stereochemical Nonrigidity and Ligand Dynamics in Hypervalent Tin(IV) Compounds. Heteronuclear NMR and Crystallographic Studies of Triorganoyltin(IV) and Diorganoyltin(IV) Complexes with Dithiolate Ligands. Inorg. Chem. 1992, 31, 3601−3606. (c) Yin, H. D.; Xue, S. C. Synthesis and characterization of organotin complexes with dithiocarbamates and crystal structures of (4-NCC6H4CH2)2Sn(S2CNEt2)2 and (2-ClC6H4CH2)2Sn(Cl)S2CNBz2. Appl. Organomet. Chem. 2006, 20, 283−289. (d) Santacruz-Juárez, E.; Cruz-Huerta, J.; Hernández-Ahuactzi, I. F.; Reyes-Martínez, R.; Tlahuext, H.; Morales-Rojas, H.; Höpfl, H. 24- and 26-Membered Macrocyclic Diorganotin(IV) Bis-Dithiocarbamate Complexes with N,N′-Disubstituted 1,3- and 1,4-Bis(aminomethyl)benzene and 1,1′-Bis(aminomethyl)ferrocene as Spacer Groups. Inorg. Chem. 2008, 47, 9804−9812. (e) Tlahuext, H.; Reyes-Martínez, R.; Vargas-Pineda, G.; López-Cardoso, M.; Höpfl, H. Molecular structures and supramolecular association of chlorodiorganotin(IV) complexes with bisand tris-dithiocarbamate ligand. J. Organomet. Chem. 2011, 696, 693− 701. (29) (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) Avram, L.; Cohen, Y. Diffusion NMR of molecular cages and capsules. Chem. Soc. Rev. 2015, 44, 586−602. (c) Evans, R.; Deng, Z.; Rogerson, A. K.; McLachlan, A. S.; Richards, J. J.; Nilsson, M.; Morris, G. A. Quantitative Interpretation of Diffusion-Ordered NMR Spectra: Can We Rationalize Small Molecule Diffusion Coefficients? Angew. Chem., Int. Ed. 2013, 52, 3199−3202. 2459

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460

Article

Organometallics (30) Einstein, A. On the Motion of Small Particles Suspended in Liquids at Rest Required by the Molecular-Kinetic Theory of Heat. Ann. Phys. 1905, 322, 549−560. (31) Frish, L.; Matthews, S. E.; Böhmer, V.; Cohen, Y. A pulsed gradient spin echo NMR study of guest encapsulation by hydrogenbonded tetraurea calix[4]arene dimers. J. Chem. Soc., Perkin Trans. 2 1999, 669−672. (32) Addison, A. W.; Rao, T. N.; Reedijk, J.; van Rijn, J.; Verschoor, G. C. Synthesis, Structure, and Spectroscopic Properties of Copper(II) Compounds containing Nitrogen−Sulphur Donor Ligands; the Crystal and Molecular Structure of Aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) Perchlorate. J. Chem. Soc., Dalton Trans. 1984, 1349−1356. (33) (a) Tlahuext-Aca, A.; Medrano, F.; Tlahuext, H.; Román-Bravo, P.; Godoy-Alcántar, C. Effect of weak sulfur···C(π) interactions, and hydrogen bonds in supramolecular association of chlorodiphenyltin(IV) dithiocarbamate complexes: Study of their stability in solution. Polyhedron 2012, 33, 223−234. (b) Santacruz-Juárez, E.; CruzHuerta, J.; Torres-Huerta, A.; Hernández-Cruz, M. G.; Barba, V.; Tlahuext, H.; Hö pfl, H. Stabilization of [(nBu2SnCl)(μ-Cl)2(ClSnnBu2)] within the solid-state structure of a Chlorodi-nbutyltin(IV) dithiocarbamate. J. Organomet. Chem. 2014, 770, 121− 129. (34) (a) Wengrovius, J. H.; Garbauskas, M. F. Solution and SolidState Structures of Several Dialkyltin Dicarboxylate Complexes, [R2Sn(O2CR′CO2)]x: Polymorphic Organotin Polymers. Organometallics 1992, 11, 1334−1342. (b) Beckmann, J.; Jurkschat, K. Stannasiloxanes: from rings to polymers. Coord. Chem. Rev. 2001, 215, 267−300. (c) Hernández-Ahuactzi, I. F.; Cruz-Huerta, J.; Barba, V.; Höpfl, H.; Zamudio-Rivera, L. S.; Beltrán, H. I. Sequence of MetalOrganic Oligomer-Polymer Exchange Equilibria in Solution: Supramolecular Isomerism, Self-Assembly Dynamics and Carboxylate Shift of Di-n-butyltin cis-1,4-Cyclohexanedicarboxylate. Eur. J. Inorg. Chem. 2008, 2008, 1200−1204. (d) Niermeier, P.; Blomeyer, S.; Bejaoui, Y. K. J.; Beckmann, J. L.; Neumann, B.; Stammler, H.-G.; Mitzel, N. W. Bidentate Boron Lewis Acids: Selectivity in Host−Guest Complex Formation. Angew. Chem., Int. Ed. 2019, 58, 1965−1969. (35) Agilent Technologies. CrysAlisPro, Versions 1.171.37.34 and 1.171.37.35; Yarnton, Oxfordshire, United Kindom, 2014. (36) (a) Sheldrick, G. M. A short history of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, A64, 112−122. (b) Sheldrick, G. M. SHELXT − Integrated space-group and crystal-structure determination. Acta Crystallogr., Sect. A: Found. Adv. 2015, 71, 3−8. (37) Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (38) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339−341. (39) Brandenburg, K. Diamond, version 4.3.2; Crystal Impact GbR: Bonn, Germany, 2017. (40) 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., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; 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, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision B.01; Gaussian, Inc.: Wallingford, CT, 2009. (41) (a) Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648−5652. (b) Lee,

C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, B37, 785−789. (42) Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297−3305. (43) Metz, B.; Stoll, H.; Dolg, M. Small-core multiconfigurationDirac-Hartree-Fock-adjusted pseudopotentials for post-d main group elements: Application to PbH and PbO. J. Chem. Phys. 2000, 113, 2563−2569.

2460

DOI: 10.1021/acs.organomet.9b00132 Organometallics 2019, 38, 2443−2460