Next Article in Journal
Field-Induced Single Molecule Magnetic Behavior of Mononuclear Cobalt(II) Schiff Base Complex Derived from 5-Bromo Vanillin
Next Article in Special Issue
The Applications of Metallacycles and Metallacages
Previous Article in Journal
Copper (II)-Catalyzed Oxidation of Ascorbic Acid: Ionic Strength Effect and Analytical Use in Aqueous Solution
Previous Article in Special Issue
Chloride Binding Properties of a Macrocyclic Receptor Equipped with an Acetylide Gold(I) Complex: Synthesis, Characterization, Reactivity, and Cytotoxicity Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controlling Chiral Self-Sorting in Truxene-Based Self-Assembled Cages

1
Laboratoire MOLTECH-Anjou, UMR CNRS 6200, Université d’Angers, SFR MATRIX, 2 Bd Lavoisier, 49045 Angers, France
2
LCP-A2MC, FR 2843 Institut Jean Barriol de Chimie et Physique Moléculaires et Biomoléculaires, FR 3624 Réseau National de Spectrométrie de Masse FT-ICR à très haut champ, Université de Lorraine, ICPM, 1 Boulevard Arago, CEDEX 03, 57078 Metz, France
*
Authors to whom correspondence should be addressed.
Inorganics 2022, 10(7), 103; https://doi.org/10.3390/inorganics10070103
Submission received: 4 July 2022 / Revised: 14 July 2022 / Accepted: 15 July 2022 / Published: 19 July 2022
(This article belongs to the Special Issue Metallamacrocycles and Metallacages: Foundations and Applications)

Abstract

:
Coordination driven self-assembly of achiral components, i.e., hexa-alkylated truxene ligands (L) with bis-metallic complexes (M2), afforded three chiral face-rotating stereoisomer polyhedra (M6L2). By tuning the length of the alkyl chains as well as the distance between both ligands facing each other in the self-assemblies (M6L2), one can control the diastereomeric distribution between the expected homo- and hetero-chiral structures.

Graphical Abstract

1. Introduction

The labile nature of the coordination bond has been extensively exploited for the self-assembly, under thermodynamic control, of simple building blocks into sophisticated 2D or 3D architectures [1,2,3]. The latter can be used to target applications such as molecular recognition, drug delivery, molecular separation or even catalysis in confined spaces [4,5,6,7,8,9,10,11,12,13,14]. In line with the work developed in the field of molecular machines, researchers are particularly interested in designing stimuli-responsive self-assembled structures, whose properties or shape can be tuned through light irradiation, application of a current or addition of a chemical [15,16,17,18,19,20,21,22,23]. Nevertheless, most of these self-assembled structures are achiral. Developing new synthetic strategies to convert symmetric ligands into chiral self-assemblies is highly desirable since the latter can be in principle useful for enantioselective sensing or asymmetric transformations [24,25,26,27,28,29]. While the most common synthetic strategy relies on the use of one chiral building block along the self-assembly process [27,30,31], examples in which achiral components are self-sorted [32,33] to form either homochiral [34,35,36,37,38,39,40,41] or heterochiral [34,35,42,43] self-assemblies are also depicted.
In this context, C3 symmetrical cages have emerged as an interesting class of supramolecular compounds for studying chirality dynamics [44]. Among others, this phenomenon was recently observed with the family of face-rotating polyhedral [45,46,47,48,49,50,51]. They are obtained by the combination of chiral or achiral linkers with face-rotating moieties such as triazatruxene [45,46] or truxene [40,48,49,50,51,52] for example. Hexa-alkylated truxene derivatives exhibit a C3h-symmetry and show both clockwise (C, green) and anti-clockwise (A, blue) faces as defined by the rotation sense of the three sp3 bridges along the C3 axis (Figure 1a). Functionalization of the latter with three pyridine units allows for producing prochiral ligands able to self-assemble with bis-metallic complexes into M6L2 cages (Figure 1b). Upon self-assembling, the ligands lose their mirror symmetry, which results in several possible stereoisomers, i.e., an enantiomers couple (CC/AA) and the meso form (AC). We showed recently that the hexabutyl truxene ligand LBu reacts with the hydroxynaphtoquinonato diruthenium complex Ru to afford, thanks to a chiral self-sorting process, only the CC/AA enantiomers couple [53]. We hypothesized this selectivity was due to the through space interactions occurring between the butyl chains located within the cavity, which self-organize in an alternated way to minimize the constraints. Herein, we investigate further the role of those interactions while studying other combinations of new hexa-alkyl truxene-based ligands and bis-metallic complexes of various lengths.

2. Results and Discussion

Ligands LEt and LBu (Scheme 1) were synthetized from previously described hexaalkyl truxene derivatives 1a and 1b respectively [54,55,56]. After a selective tri-bromination of the truxene moiety using Br2 that affords compounds 2a and 2b (>80% yields), pallado-catalysed Suzuki-Miyaura cross-coupling reactions with 4-pyridinylboronic were carried out in a mixture of toluene and ethanol in basic conditions. The target ligands LEt and LBu were obtained in good yields considering that three sites are functionalized during the reaction (60% and 46%, respectively).
Single crystals of ligand LEt were obtained by slow evaporation of its solution in chloroform. The corresponding crystallographic structure is depicted in Figure 2 and compared to the previously described ligand LBu. Both ligands show nearly similar structural characteristics with a planar truxene core which places the three peripherical nitrogen atoms in the same plane. To minimize H-H interactions, the pyridine shows an average rotation angle of 38.6(1)° (LEt) and 40.8(1)° (LBu) with the central truxene core [57]. In accordance with the C3 symmetry of both ligands, an angle of 120° is measured between each pyridine axis. Finally, the n-alkyl chains are arranged on both opposite sides, almost perpendicularly to the truxene plane.
The self-assembly reactions were proceeded between two equivalents of ligands and three equivalents of bis-metallic complexes, with the objective of forming M6L2 cages which associate two truxene entities face to face (Scheme 1). The resulting cavity is thus intended to be filled by the n-alkyl groups which are intercalated between the two aromatic platforms. Our previous work showed the importance of through space interactions between n-alkyl chains to drive the relative spatial organization of the ligands within the self-assembly. With the aim of controlling the ratio between AA/CC and AC species, we studied new ligand-complex combinations, i.e., LEt with Ru and LBu with Rh. Based on the X-ray structures of LEt and LBu, we extracted the distance “a” between the sp3 carbon of the Ar-Ar bridge in the truxene moiety and the sp3 carbon of the terminal methyl group of the alkyl chain (Figure 2). On the other hand, “b” is defined as the intermetallic distance within the Ru and Rh bimetallic complexes as extracted from literature data. From these values, we calculated “c” (= b − 2a) which reflects the minimal distance between terminal methyl groups of opposite dangling alkyl chains as function of the dinuclear bridge (Table 1). While this value is negative (c = −1.6 Å) for BuRu cage, for which only the CC and AA enantiomers are observed, it increases to 2.9 Å for BuRh, meaning the alkyl chains are too far away to interact in this case. The situation for EtRu appears intermediate.
All reactions were carried out in methanol-d4 at C = 10−3 M and well-resolved 1H NMR spectra, suggesting the formation of one discrete species, were observed in the three cases after 4 h at 50 °C (Figures S7, S10 and S14). The corresponding cages BuRu, EtRu and BuRh were isolated by precipitation with diethyl ether in ca. 80% yield. ESI-FTICR-HRMS spectrometry experiments were carried out in methanol at C = 10−4 M on each sample (Figure 3). These measurements confirm the expected M6L2 stoichiometry in the three cases, with characteristic multi-charged isotopic patterns localized at m/z = 789.65155, 1024.30257 and 1415.38762 for compound BuRu (main contributions, Figure S17), at m/z = 721.97871, 939.71148 and 1302.59947 for EtRu (main contributions, Figure S18) and at m/z = 735.20408, 882.73054, 1089.06712 and 1398.57199 for BuRh (main contributions, Figure S19).
The 1H NMR, 1H COSY NMR and 1H DOSY NMR spectra of the three cages are shown in Figure 4 (high-field region) and Figures S7–S16. The diffusion measurements revealed in each case one single diffusion value (D). BuRu and EtRu self-assemblies exhibit a similar diffusion coefficient close to 3.0 × 10−10 m2·s−1 while this value decreases for BuRh cage (D = 2.3 × 10−10 m2·s−1), as expected for a larger edifice. Hydrodynamic radii of ca. 13 Å and 17 Å were calculated from the Stokes-Einstein equation [58], in agreement with the formation of M6L2 species. The case of BuRu was deeply investigated in a previous work, establishing the exclusive formation of the D3 symmetric CC and AA enantiomers [53].
This was assigned to through-space van der Waals interaction between butyl chains facing each other inside the cavity (c = −1.6 Å). While in this case only one set of signals for the truxene backbone (Figure S7, protons α, β, a, b and c) and for methyl groups from the alkyl chains is observed, these signals are splitted in the cases of EtRu (Figures S10, S11 and Figure 4d,d’) and BuRh (Figure S14 and Figure 4e), suggesting the presence of the mixture of diastereoisomers. The relative integrals of protons 4 (BuRu and BuRh) and protons 2 (EtRu) were used to determine the proportion of enantiomers AA/CC vs. the meso form AC. On this basis, a 50/50 statistical mixture of the diastereoisomers is determined for BuRh (Figure 4e), which confirms the absence of interaction between both opposite faces in the structure (c = 2.9 Å). A similar ratio is observed for EtRu at short reaction time (Figure 4d) but the latter evolves to 66/33 after several hours (Figure 4d’). This indicates that the reaction is thermodynamically driven and confirms the crucial role of the inter-ligand interactions inside the cavity. Regarding (i) the evolution of the ratio over time, (ii) the selectivity observed for BuRu and (iii) the statistical mixture observed for BuRh, the dominant species in EtRu should be the couple of enantiomers AA/CC.
Single crystals were obtained for self-assemblies BuRu and BuRh from slow diffusion of methyl tert-butyl ether in their methanolic solutions (Figure 5). The BuRu cage crystallized in the non-centrosymmetric P3121 space group with a Flack parameter of 0.44 indicating enantio-enriched crystal (enriched in AA) and the BuRh cage in the centrosymmetric C 2/c space group. Interestingly, if BuRh solution speciation revealed a statistical mixture of all stereoisomers, only AA/CC enantiomeric pair was observed in the crystalline phase. Despite numerous attempts, no single crystals of EtRu suitable for XRD measurements could be obtained. We have therefore modeled the corresponding compound by MM+ calculations using the CC-BuRu structure as a starting-point. Unlike topologically similar metalla-structures in which a large tilt of the bis-ruthenium complexes allows the facing ligands planes to be closer in order to maximize the π-π interactions [59,60], the alkyl chains present inside the cavity of BuRu, EtRu and BuRh strongly limits structural deviation from right prismatic geometry and maintain both truxene moieties at distances close to the intermetallic one. These are however slightly shorter than the M-M distances. As a result, the trigonal prisms are only slightly distorted with average Bailar angles ranging from 7.9° to 16° (Figure S20). Moreover, in each case, the pyridyl rings are tilted by ca. 20° out of the plane of the truxene moiety. All these parameters generate two types of chirality, a double rosette (M/P) and a propeller isomerism (Δ/Λ). Both are dependent one other in these structures since only the enantiomers (AA, M, Δ) and (CC, P, Λ) are observed [60,61].

3. Materials and Methods

3.1. Chemicals

Compounds 1b [56], 2b and LBu [53], as well as compounds 1a and 2a [62], and complexes Ru [63] and Rh [64] were synthesized using procedures described in the literature. All reagents were of commercial reagent grade and were used without further purification. Silica gel chromatography was performed with a SIGMA Aldrich Chemistry SiO2 (pore size 60 Ã, 40–63 µm technical grades).

3.2. Instrumentation

Characterizations and NMR experiments were carried out on a NMR Bruker Avance III 300 spectrometer at 298 K, using perdeuterated solvents. 1H DOSY NMR spectra were analyzed with MestReNova software. MALDI-TOF-MS spectra were recorded on a MALDI-TOF Bruker Biflex III instrument using a positive-ion mode. ESI-FTICR mass spectra at very high resolution were performed in positive detection mode on a 7T Solarix 2xR (Bruker Daltonics, Marne la Vallée, France).

3.3. Experimental Procedure and Characterization Data

3.3.1. Ligand LEt

To a stirred solution of 2a (0.50 g, 0.669 mmol) in toluene (25 mL) and EtOH (12 mL) was added 4-pyridine boronic acid (0.33 g, 2.68 mmol, 4 equivalents) at room temperature. Then K2CO3 (1.94 g, 14.05 mmol, 21 equivalents) in water (8 mL) was added to the solution at room temperature. The solution was degassed with argon for 40 min at room temperature. Pd(PPh3)4 (0.23 g, 30%) was then added to the solution. The solution was stirred at 90 °C. After 64 h, the mixture was cooled to room temperature and extracted with dichloromethane. The organic extracts were washed with water, dried over magnesium sulfate and the solvent were evaporated. The residue was purified by chromatography on silica gel using ethyl acetate/dichloromethane/methanol/triethylamine (from 75/25/0/0 to 42/42/15/1) as an eluant to give ligand LEt as a yellow powder (369 mg, 74%). 1H NMR (300 MHz, 298 K, CDCl3): δ 8.74–8.72 (m, 6H), 8.50–8.47 (d, 3J = 8.9 Hz, 3H), 7.76 (s, 3H), 7.75 (d, 3J = 7.1 Hz, 3H), 7.68–7.65 (m, 6H), 3.12–3.05 (m, 6H), 2.31–2.17 (m, 6H), 0.29 (t, 3J = 7.1 Hz, 18H). 13C NMR (76 MHz, CDCl3): δ 153.85, 150.43, 148.23, 145.27, 141.45, 138.50, 136.52, 125.45, 125.20, 121.56, 120.67, 57.18, 29.65, 8.65. HRMS (MALDI-TOF): found: 741.4110; Calculated: 741.4083.

3.3.2. Self-Assembly EtRu

A mixture of LEt (10.00 mg, 13.5 μmol, 2 equiv.) and complex Ru (19.35 mg, 20.0 µmol, 3 equiv.) in methanol (2 mL) was stirred 20 h at 50 °C. Then, diethyl ether (5 mL) was added and the resulting suspension was centrifuged. The resulting solid was washed twice with diethyl ether to give EtRu (24.02 mg, 4.5 μmol, 83%) as a dark solid. 1H NMR (300 MHz, MeOD): δ 8.50–8.46 (m, 12H), 8.32–8.26 (m, 6H), 7.90–7.77 (m, 24H), 7.32–7.31 (m, 12H), 5.92–5.88 (m, 12H), 5.69–5.65 (m, 12H), 2.92–2.88 (m, 12H), 2.70–2.50 (m, 6H), 2.17–2.16 (m, 24H), 2.02–1.95 (m, 6H), 1.37 (d, 3J = 6.9 Hz, 36H), 0.10–(−0.02) (m, 18H), −0.32–(−0.49) (m, 18H). 1H DOSY NMR (300 MHz, MeOD) D = 3.03 × 10−10 m2·s−1. FTICR-HRMS (m/z), [EtRu − 3TfO]3+: found: 1302.59965, calculated 1302.59947, [EtRu − 4TfO]4+: found: 939.71154, calculated 939.71148, [EtRu − 5TfO]5+: found: 721.97871, calculated 72197870.

3.3.3. Self-Assembly BuRh

A mixture of LBu (10.00 mg, 10.6 μmol, 2 equiv.) and complex Rh (23.98 mg, 15.9 µmol, 3 equiv.) in methanol (2.5 mL) was stirred 1 h at 50 °C. Then, diethyl ether (5 mL) was added and the resulting suspension was centrifuged. The resulting solid was washed twice with diethyl ether to give BuRh (25.14 mg, 3.92 μmol, 74%) as a yellow solid. 1H NMR (300 MHz, 298 K, MeOD): δ 10.38–10.35 (m, 12H), 9.93 (d, 3J = 7.3 Hz, 12H), 8.72–8.68 (m, 12H), 8.47–8.43 (m, 12H), 8.33–8.30 (m 6H), 7.95–7.75 (m, 24H), 2.90–2.86 (m, 6H), 2.71–2.67 (m 6H), 2.13–2.08 (m, 6H), 1.95–1.82 (m, 96H), 0.95–0.87 (12H), 0.50–0.05 (m, 42H), −0.06–(−0.20) (m, 12H), −0.36 (t, 3J = 7.2Hz), −0.44 (t, 3J = 7.3 Hz, 9H). 1H DOSY NMR (300 MHz, MeOD): D = 2.34 × 10−10 m2·s−1. FTICR-HRMS (m/z): [BuRh − 4OTf]4+: found: 1398.57256; calculated 1398.57199, [BuRh − 5Otf]5+: found: 1089.06707; calculated 1089.06712, [BuRh − 6Otf]6+: found: 882.72998; calculated: 882.73054, [BuRh − 7Otf]7+: found: 735.20397; calculated: 735.20408.

3.4. Molecular Modelling

Molecular modelling was performed by using the molecular mechanics force field MM+ method from the HyperChem 8.0.3 program (Hypercube, Inc., Waterloo, ON, Canada,) configured in vacuo, with a RMS of 10−5 kcal/mole, a number of maximum cycles of 32,500, and a Polak-Ribiere algorithm. Counter anions were omitted to simplify the calculation.

3.5. X-ray Crystallographic Analysis

X-ray single-crystal diffraction data were collected at low temperature on a Rigaku Oxford Diffraction SuperNova diffractometer equipped with Atlas CCD detector and micro-focus Cu-Kα radiation (λ = 1.54184 Å). The structures were solved by dual-space algorithm, expanded and refined on F2 by full matrix least-squares techniques using SHELX programs (G. M. Sheldrick, SHELXT 2018/2 and SHELXL 2018/3). All non-H atoms were refined anisotropically and multiscan empirical absorption was corrected using CrysAlisPro program (CrysAlisPro 1.171.40.45a and 1.171.41.118a, Rigaku Oxford Diffraction, 2019–2021). The H atoms were placed at calculated positions and refined using a riding model.
The structure refinement of BuRh showed disordered electron density which could not be reliably modeled. The program PLATON/SQUEEZE was used to add the corresponding scattering contribution to the calculated structure factors. This electron density can be attributed to solvent molecules (methyl tert-butyl ether (MTBE)) and missing triflate molecules (48 CF3SO3 anions). The assumed solvent composition and missing anions were included in the calculation of the empirical formula, formula weight, density, linear absorption coefficient, and F(000).
Crystallographic data for LEt: C54H51N3, M = 741.97, T = 200K, colorless prism, 0.243 × 0.137 × 0.108 mm3, monoclinic, space group Cc, a = 16.8429(4) Å, b = 29.2254(7) Å, c = 8.8287(3) Å, β = 104.094(3)°, V = 4215.0(2) Å3, Z = 4, ρcalc = 1.169 g/cm3, μ = 0.513 mm−1, F(000) = 1584, θmin = 3.024°, θmax = 72.419°, 15675 reflections collected, 6363 unique (Rint = 0.0661), parameters/restraints = 520/2, R1 = 0.0631 and wR2 = 0.1739 using 5960 reflections with I > 2σ(I), R1 = 0.0657 and wR2 = 0.1786 using all data, GOF = 1.047, −0.300 < Δρ < 0.237 e.Å-3. CCDC 2183860.
Crystallographic data for BuRh: C376H516F36N24O56Rh6S12, M = 7954.28, T = 150K, yellow prism, 0.13 × 0.076 × 0.027 mm3, monoclinic, space group C2/c, a = 39.414(3) Å, b = 24.8816(17) Å, c = 40.900(6) Å, β = 98.08(1)°, V = 39711(7) Å3, Z = 4, ρcalc = 1.330 g/cm3, μ = 3.291 mm−1, F(000) = 16696, θmin = 2.916°, θmax = 73.304°, 85378 reflections collected, 37770 unique (Rint = 0.1374), parameters/restraints = 1315/28, R1 = 0.1173 and wR2 = 0.3152 using 10726 reflections with I > 2σ(I), R1 = 0.2119 and wR2 = 0.3999 using all data, GOF = 0.927, −0.739 < Δρ < 0.594 e.Å-3. CCDC 2183861.

4. Conclusions

In summary, we synthetized a series of three chiral M6L2 metalla-cages from prochiral hexa-alkylated truxene ligands and achiral dinuclear metallic complexes. Remarkably, the diastereomeric ratio between enantiomers AA/CC and the meso AC form can be tuned thanks to the non-covalent interactions occurring between the internal alkyl chains upheld by two opposite facing ligands. While only the AA/CC enantiomers couple is observed in solution in the case of BuRu, the metalla-cage BuRh exists as the statistical mixture of the three possible stereoisomers. The latter are also observed for EtRu through preferentially producing the enantioenriched AA/CC forms. These results unambiguously confirm the crucial role of inter-ligand communication through the alkyl chains in the self-sorting process. Work is under progress to evaluate the binding abilities of these M6L2 chiral cages.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics10070103/s1, NMR spectra, cif files and check cif files of LEt, EtRu and BuRh.

Author Contributions

Conceptualization, M.S. and S.G.; methodology, S.G.; synthesis and NMR experiments, A.B., S.S., A.L., E.C. and L.M.; HRMS-FTICR analyses, V.C. and F.A.; X-ray analyses, M.A.; writing—original draft preparation, S.G.; writing—review and editing, M.S., A.B. and S.G.; supervision, S.G.; funding acquisition, S.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CNRS grant EMERGENCE@INC2019 and RFI LUMOMAT grant 2019 (PHOTOCAGE).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge the University of Angers for a PhD grant (S.S.). They also acknowledge the ASTRAL platform (SFRMATRIX, Univ. Angers) for their assistance in spectroscopic analyses. Finally, the financial support from the National FT-ICR network (FR3624 CNRS) for conducting the research is gratefully acknowledged.

Conflicts of Interest

Authors declare no conflicts of interest.

References

  1. Sun, Y.; Chen, C.; Liu, J.; Stang, P.J. Recent developments in the construction and applications of platinum-based metallacycles and metallacages via coordination. Chem. Soc. Rev. 2020, 49, 3889–3919. [Google Scholar] [CrossRef]
  2. Debata, N.B.; Tripathy, D.; Sahoo, H.S. Development of coordination driven self-assembled discrete spherical ensembles. Coord. Chem. Rev. 2019, 387, 273–298. [Google Scholar] [CrossRef]
  3. Cook, T.R.; Stang, P.J. Recent Developments in the Preparation and Chemistry of Metallacycles and Metallacages via Coordination. Chem. Rev. 2015, 115, 7001–7045. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, D.; Ronson, T.K.; Zou, Y.-Q.; Nitschke, J.R. Metal–organic cages for molecular separations—Nature Reviews Chemistry. Nat. Rev. Chem. 2021, 5, 168–182. [Google Scholar] [CrossRef]
  5. Grommet, A.B.; Feller, M.; Klajn, R. Chemical reactivity under nanoconfinement. Nat. Nanotechnol. 2020, 15, 256–271. [Google Scholar] [CrossRef]
  6. Percástegui, E.G.; Ronson, T.K.; Nitschke, J.R. Design and Applications of Water-Soluble Coordination Cages. Chem. Rev. 2020, 120, 13480–13544. [Google Scholar] [CrossRef] [PubMed]
  7. Yadav, S.; Kannan, P.; Qiu, G. Cavity-based applications of metallo-supramolecular coordination cages (MSCCs). Org. Chem. Front. 2020, 7, 2842–2872. [Google Scholar] [CrossRef]
  8. Fang, Y.; Powell, J.A.; Li, E.; Wang, Q.; Perry, Z.; Kirchon, A.; Yang, X.; Xiao, Z.; Zhu, C.; Zhang, L.; et al. Catalytic reactions within the cavity of coordination cages. Chem. Soc. Rev. 2019, 48, 4707–4730. [Google Scholar] [CrossRef]
  9. Gao, W.-X.; Zhang, H.-N.; Jin, G.-X. Supramolecular catalysis based on discrete heterometallic coordination-driven metallacycles and metallacages. Coord. Chem. Rev. 2019, 386, 69–84. [Google Scholar] [CrossRef]
  10. Rizzuto, F.J.; von Krbek, L.K.S.; Nitschke, J.R. Strategies for binding multiple guests in metal–organic cages. Nat. Rev. Chem. 2019, 3, 204–222. [Google Scholar] [CrossRef]
  11. Zhao, L.; Jing, X.; Li, X.; Guo, X.; Zeng, L.; He, C.; Duan, C. Catalytic properties of chemical transformation within the confined pockets of Werner-type capsules. Coord. Chem. Rev. 2019, 378, 151–187. [Google Scholar] [CrossRef]
  12. Casini, A.; Woods, B.; Wenzel, M. The Promise of Self-Assembled 3D Supramolecular Coordination Complexes for Biomedical Applications. Inorg. Chem. 2017, 56, 14715–14729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zarra, S.; Wood, D.M.; Roberts, D.A.; Nitschke, J.R. Molecular containers in complex chemical systems. Chem. Soc. Rev. 2015, 44, 419–432. [Google Scholar] [CrossRef] [Green Version]
  14. Amouri, H.; Desmarets, C.; Moussa, J. Confined Nanospaces in Metallocages: Guest Molecules, Weakly Encapsulated Anions, and Catalyst Sequestration. Chem. Rev. 2012, 112, 2015–2041. [Google Scholar] [CrossRef]
  15. Cai, K.; Zhang, L.; Astumian, R.D.; Stoddart, J.F. Radical-pairing-induced molecular assembly and motion. Nat. Rev. Chem. 2021, 5, 447–465. [Google Scholar] [CrossRef]
  16. Goeb, S.; Sallé, M. Electron-rich Coordination Receptors Based on Tetrathiafulvalene Derivatives: Controlling the Host-Guest Binding. Acc. Chem. Res. 2021, 54, 1043–1055. [Google Scholar] [CrossRef]
  17. Wezenberg, S.J. Light-switchable Metal-Organic Cages. Chem. Lett. 2020, 49, 609–615. [Google Scholar] [CrossRef]
  18. Chen, L.-J.; Yang, H.-B. Construction of Stimuli-Responsive Functional Materials via Hierarchical Self-Assembly Involving Coordination Interactions. Acc. Chem. Res. 2018, 51, 2699–2710. [Google Scholar] [CrossRef]
  19. Kim, T.Y.; Vasdev, R.A.S.; Preston, D.; Crowley, J.D. Strategies for Reversible Guest Uptake and Release from Metallosupramolecular Architectures. Chem. Eur. J. 2018, 24, 14878–14890. [Google Scholar] [CrossRef]
  20. Zhang, D.; Ronson, T.K.; Nitschke, J.R. Functional Capsules via Subcomponent Self-Assembly. Acc. Chem. Res. 2018, 51, 2423–2436. [Google Scholar] [CrossRef]
  21. Diaz-Moscoso, A.; Ballester, P. Light-responsive molecular containers. Chem. Commun. 2017, 53, 4635–4652. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, W.; Wang, Y.-X.; Yang, H.-B. Supramolecular transformations within discrete coordination-driven supramolecular architectures. Chem. Soc. Rev. 2016, 45, 2656–2693. [Google Scholar] [CrossRef] [PubMed]
  23. Qu, D.-H.; Wang, Q.-C.; Zhang, Q.-W.; Ma, X.; Tian, H. Photoresponsive Host–Guest Functional Systems. Chem. Rev. 2015, 115, 7543–7588. [Google Scholar] [CrossRef]
  24. Trapp, O. Self-amplification of Enantioselectivity in Asymmetric Catalysis by Supramolecular Recognition and Stereodynamics. In Supramolecular Catalysis: New Directions and Developments; van Leeuwen, P.W.N.M., Raynal, M., Eds.; WILEY-VCH: Weinheim, Germany, 2022. [Google Scholar]
  25. Zhang, L.; Liu, H.; Yuan, G.; Han, Y.-F. Chiral Coordination Metallacycles/Metallacages for Enantioselective Recognition and Separation. Chin. J. Chem. 2021, 39, 2273–2286. [Google Scholar] [CrossRef]
  26. Zhang, J.H.; Xie, S.M.; Zi, M.; Yuan, L.M. Recent advances of application of porous molecular cages for enantioselective recognition and separation. J. Sep. Sci. 2020, 43, 134–149. [Google Scholar] [CrossRef]
  27. Pan, M.; Wu, K.; Zhang, J.-H.; Su, C.-Y. Chiral metal–organic cages/containers (MOCs): From structural and stereochemical design to applications. Coord. Chem. Rev. 2019, 378, 333–349. [Google Scholar] [CrossRef]
  28. Tan, C.; Chu, D.; Tang, X.; Liu, Y.; Xuan, W.; Cui, Y. Supramolecular Coordination Cages for Asymmetric Catalysis. Chem. Eur. J. 2019, 25, 662–672. [Google Scholar] [CrossRef]
  29. Li, X.; Wu, J.; He, C.; Meng, Q.; Duan, C. Asymmetric Catalysis within the Chiral Confined Space of Metal-Organic Architectures. Small 2019, 15, e1804770. [Google Scholar] [CrossRef]
  30. Chen, L.-J.; Yang, H.-B.; Shionoya, M. Chiral metallosupramolecular architectures. Chem. Soc. Rev. 2017, 46, 2555–2576. [Google Scholar] [CrossRef]
  31. Liu, M.; Zhang, L.; Wang, T. Supramolecular Chirality in Self-Assembled Systems. Chem. Rev. 2015, 115, 7304–7397. [Google Scholar] [CrossRef]
  32. Li, C.L.; Zuo, Y.; Zhao, Y.Q.; Zhang, S.D. Chiral Self-sorting in Cage-like Compounds. Chem. Lett. 2020, 49, 1356–1366. [Google Scholar] [CrossRef]
  33. Jędrzejewska, H.; Szumna, A. Making a Right or Left Choice: Chiral Self-Sorting as a Tool for the Formation of Discrete Complex Structures. Chem. Rev. 2017, 117, 4863–4899. [Google Scholar] [CrossRef] [PubMed]
  34. Schulte, T.R.; Holstein, J.J.; Clever, G.H. Chiral Self-Discrimination and Guest Recognition in Helicene-Based Coordination Cages. Angew. Chem. Int. Ed. 2019, 58, 5562–5566. [Google Scholar] [CrossRef]
  35. Beaudoin, D.; Rominger, F.; Mastalerz, M. Chiral Self-Sorting of [2+3] Salicylimine Cage Compounds. Angew. Chem. Int. Ed. 2017, 56, 1244–1248. [Google Scholar] [CrossRef]
  36. Rota Martir, D.; Escudero, D.; Jacquemin, D.; Cordes, D.B.; Slawin, A.M.Z.; Fruchtl, H.A.; Warriner, S.L.; Zysman-Colman, E. Homochiral Emissive Λ8- and Δ8-[Ir8Pd4]16+ Supramolecular Cages. Chem. Eur. J. 2017, 23, 14358–14366. [Google Scholar] [CrossRef] [Green Version]
  37. Boer, S.A.; Turner, D.R. Self-selecting homochiral quadruple-stranded helicates and control of supramolecular chirality. Chem. Commun. 2015, 51, 17375–17378. [Google Scholar] [CrossRef]
  38. Maeda, C.; Kamada, T.; Aratani, N.; Osuka, A. Chiral self-discriminative self-assembling of meso–meso linked diporphyrins. Coord. Chem. Rev. 2007, 251, 2743–2752. [Google Scholar] [CrossRef]
  39. Lützen, A.; Hapke, M.; Griep-Raming, J.; Haase, D.; Saak, W. Synthesis and Stereoselective Self-Assembly of Double- and Triple-Stranded Helicates. Angew. Chem. Int. Ed. 2002, 41, 2086–2089. [Google Scholar] [CrossRef]
  40. Xu, C.; Lin, Q.; Shan, C.; Han, X.; Wang, H.; Wang, H.; Zhang, W.; Chen, Z.; Guo, C.; Xie, Y.; et al. Metallo-Supramolecular Octahedral Cages with Three Types of Chirality towards Spontaneous Resolution. Angew. Chem. 2022, 134, e202203099. [Google Scholar] [CrossRef]
  41. Masood, M.A.; Enemark, E.J.; Stack, T.D.P. Ligand Self-Recognition in the Self-Assembly of a [{Cu(L)}2]2+ Complex: The Role of Chirality. Angew. Chem. Int. Ed. 1998, 37, 928–932. [Google Scholar] [CrossRef]
  42. Arribas, C.S.; Wendt, O.F.; Sundin, A.P.; Carling, C.-J.; Wang, R.; Lemieux, R.P.; Wärnmark, K. Formation of an heterochiral supramolecular cage by diastereomer self-discrimination: Fluorescence enhancement and C60 sensing. Chem. Commun. 2010, 46, 4381–4383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Weilandt, T.; Kiehne, U.; Schnakenburg, G.; Lützen, A. Diastereoselective self-assembly of dinuclear heterochiral metallosupramolecular rhombs in a self-discriminating process. Chem. Commun. 2009, 17, 2320–2322. [Google Scholar] [CrossRef] [PubMed]
  44. Qiu, G.; Nava, P.; Colomban, C.; Martinez, A. Control and Transfer of Chirality Within Well-Defined Tripodal Supramolecular Cages. Front. Chem. 2020, 8, 599893. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, D.; Ronson, T.K.; Greenfield, J.L.; Brotin, T.; Berthault, P.; Léonce, E.; Zhu, J.-L.; Xu, L.; Nitschke, J.R. Enantiopure [Cs+/Xe⊂Cryptophane]⊂FeII4L4 Hierarchical Superstructures. J. Am. Chem. Soc. 2019, 141, 8339–8345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Zhang, P.; Wang, X.; Xuan, W.; Peng, P.; Li, Z.; Lu, R.; Wu, S.; Tian, Z.; Cao, X. Chiral separation and characterization of triazatruxene-based face-rotating polyhedra: The role of non-covalent facial interactions. Chem. Commun. 2018, 54, 4685–4688. [Google Scholar] [CrossRef]
  47. Qu, H.; Tang, X.; Wang, X.; Li, Z.; Huang, Z.; Zhang, H.; Tian, Z.; Cao, X. Chiral molecular face-rotating sandwich structures constructed through restricting the phenyl flipping of tetraphenylethylene. Chem. Sci. 2018, 9, 8814–8818. [Google Scholar] [CrossRef] [Green Version]
  48. Wang, Y.; Fang, H.; Tranca, I.; Qu, H.; Wang, X.; Markvoort, A.J.; Tian, Z.; Cao, X. Elucidation of the origin of chiral amplification in discrete molecular polyhedra. Nat. Commun. 2018, 9, 488. [Google Scholar] [CrossRef] [Green Version]
  49. Wang, X.; Peng, P.; Xuan, W.; Wang, Y.; Zhuang, Y.; Tian, Z.; Cao, X. Narcissistic chiral self-sorting of molecular face-rotating polyhedra. Org. Biomol. Chem. 2018, 16, 34–37. [Google Scholar] [CrossRef]
  50. Wang, Y.; Fang, H.; Zhang, W.; Zhuang, Y.; Tian, Z.; Cao, X. Interconversion of molecular face-rotating polyhedra through turning inside out. Chem. Commun. (Camb.) 2017, 53, 8956–8959. [Google Scholar] [CrossRef]
  51. Wang, X.; Wang, Y.; Yang, H.; Fang, H.; Chen, R.; Sun, Y.; Zheng, N.; Tan, K.; Lu, X.; Tian, Z.; et al. Assembled molecular face-rotating polyhedra to transfer chirality from two to three dimensions. Nat. Commun. 2016, 7, 12469. [Google Scholar] [CrossRef] [Green Version]
  52. Zhu, J.-L.; Zhang, D.; Ronson, T.K.; Wang, W.; Xu, L.; Yang, H.-B.; Nitschke, J.R. A Cavity-Tailored Metal-Organic Cage Entraps Gases Selectively in Solution and the Amorphous Solid State. Angew. Chem. Int. Ed. 2021, 60, 11789–11792. [Google Scholar] [CrossRef] [PubMed]
  53. Séjourné, S.; Labrunie, A.; Dalinot, C.; Benchohra, A.; Carré, V.; Aubriet, F.; Allain, M.; Sallé, M.; Goeb, S. Chiral Self-Sorting in Truxene-Based Metallacages. Inorganics 2020, 8, 1. [Google Scholar] [CrossRef] [Green Version]
  54. Bols, P.S.; Anderson, H.L. Shadow Mask Templates for Site-Selective Metal Exchange in Magnesium Porphyrin Nanorings. Angew. Chem. Int. Ed. 2018, 57, 7874–7877. [Google Scholar] [CrossRef]
  55. Tehfe, M.-A.; Lalevée, J.; Telitel, S.; Contal, E.; Dumur, F.; Gigmes, D.; Bertin, D.; Nechab, M.; Graff, B.; Morlet-Savary, F.; et al. Polyaromatic Structures as Organo-Photoinitiator Catalysts for Efficient Visible Light Induced Dual Radical/Cationic Photopolymerization and Interpenetrated Polymer Networks Synthesis. Macromolecules 2012, 45, 4454–4460. [Google Scholar] [CrossRef]
  56. Kanibolotsky, A.L.; Berridge, R.; Skabara, P.J.; Perepichka, I.F.; Bradley, D.D.C.; Koeberg, M. Synthesis and Properties of Monodisperse Oligofluorene-Functionalized Truxenes:  Highly Fluorescent Star-Shaped Architectures. J. Am. Chem. Soc. 2004, 126, 13695–13702. [Google Scholar] [CrossRef]
  57. Guan, J.; Xu, F.; Tian, C.; Pu, L.; Yuan, M.-S.; Wang, J. Tricolor Luminescence Switching by Thermal and Mechanical Stimuli in the Crystal Polymorphs of Pyridyl-substituted Fluorene. Chem. Asian J. 2019, 14, 216–222. [Google Scholar] [CrossRef] [PubMed]
  58. Cohen, Y.; Avram, L.; Frish, L. Diffusion NMR Spectroscopy in Supramolecular and Combinatorial Chemistry: An Old Parameter-New Insights. Angew. Chem. Int. Ed. 2005, 44, 520–554. [Google Scholar] [CrossRef]
  59. 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. [Google Scholar] [CrossRef]
  60. Govindaswamy, P.; Linder, D.; Lacour, J.; Süss-Fink, G.; Therrien, B. Self-assembled hexanuclear arene ruthenium metallo-prisms with unexpected double helical chirality. Chem. Commun. 2006, 45, 4691–4693. [Google Scholar] [CrossRef] [Green Version]
  61. Barry, N.P.E.; Austeri, M.; Lacour, J.; Therrien, B. Highly Efficient NMR Enantiodiscrimination of Chiral Octanuclear Metalla-Boxes in Polar Solvent. Organometallics 2009, 28, 4894–4897. [Google Scholar] [CrossRef] [Green Version]
  62. Yuan, M.-S.; Fang, Q.; Liu, Z.-Q.; Guo, J.-P.; Chen, H.-Y.; Yu, W.-T.; Xue, G.; Liu, D.-S. Acceptor or Donor (Diaryl B or N) Substituted Octupolar Truxene:  Synthesis, Structure, and Charge-Transfer-Enhanced Fluorescence. J. Org. Chem. 2006, 71, 7858–7861. [Google Scholar] [CrossRef] [PubMed]
  63. Barry, N.P.E.; Furrer, J.; Therrien, B. In and Out of Cavity Interactions by Modulating the Size of Ruthenium Metallarectangles. Helv. Chim. Acta 2010, 93, 1313–1328. [Google Scholar] [CrossRef] [Green Version]
  64. Zhang, H.-N.; Gao, W.-X.; Deng, Y.-X.; Lin, Y.-J.; Jin, G.-X. Stacking-interaction-induced host–guest chemistry and Borromean rings based on a polypyridyl ligand. Chem. Commun. 2018, 54, 1559–1562. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Hexa-alkylated truxene-based ligand (L) showing two rotational faces (Clockwise (C), green and Anticlockwise (A), blue); and (b) the three possible metalla-cage structures (M6L2) that can be obtained upon self-assembling with bis-metallic complexes.
Figure 1. (a) Hexa-alkylated truxene-based ligand (L) showing two rotational faces (Clockwise (C), green and Anticlockwise (A), blue); and (b) the three possible metalla-cage structures (M6L2) that can be obtained upon self-assembling with bis-metallic complexes.
Inorganics 10 00103 g001
Figure 2. X-Ray crystal structures of LEt (left) and LBu (right): (a,c) top views (clockwise face in green); and (b,d) lateral views showing both rotating faces (clockwise in green, anticlockwise in blue).
Figure 2. X-Ray crystal structures of LEt (left) and LBu (right): (a,c) top views (clockwise face in green); and (b,d) lateral views showing both rotating faces (clockwise in green, anticlockwise in blue).
Inorganics 10 00103 g002
Scheme 1. Synthesis of ligands LEt and LBu, and metalla-cages BuRu, EtRu and BuRh, as well as their respective stereoisomeric ratio.
Scheme 1. Synthesis of ligands LEt and LBu, and metalla-cages BuRu, EtRu and BuRh, as well as their respective stereoisomeric ratio.
Inorganics 10 00103 sch001
Figure 3. ESI-FTICR mass spectra recorded in methanol (C = 10−4 M) of: (a) BuRu; (b) EtRu; and (c) BuRh.
Figure 3. ESI-FTICR mass spectra recorded in methanol (C = 10−4 M) of: (a) BuRu; (b) EtRu; and (c) BuRh.
Inorganics 10 00103 g003
Figure 4. 1H NMR (298 K, C = 10−3 M, high-field region) of: (a) ligand LBu in CDCl3; (b) ligand LEt in CDCl3; (c) BuRu in methanol-d4 after 12 h; (d) EtRu in methanol-d4 after 4 h; and (e) BuRh in methanol-d4 after 12 h and (d’) EtRu in methanol-d4 after 12 h. The following nomenclature has been used: protons 4 correspond to terminal CH3 from butyl chains and protons 2 correspond to terminal CH3 from ethyl chains; grey and black assignments correspond to inner and outer cavity protons respectively; i and ii denote the proper signatures of the AA/CC enantiomers and the AC meso derivative respectively; 1 and 2 are used to differentiate the diastereotopic aliphatic protons (SI).
Figure 4. 1H NMR (298 K, C = 10−3 M, high-field region) of: (a) ligand LBu in CDCl3; (b) ligand LEt in CDCl3; (c) BuRu in methanol-d4 after 12 h; (d) EtRu in methanol-d4 after 4 h; and (e) BuRh in methanol-d4 after 12 h and (d’) EtRu in methanol-d4 after 12 h. The following nomenclature has been used: protons 4 correspond to terminal CH3 from butyl chains and protons 2 correspond to terminal CH3 from ethyl chains; grey and black assignments correspond to inner and outer cavity protons respectively; i and ii denote the proper signatures of the AA/CC enantiomers and the AC meso derivative respectively; 1 and 2 are used to differentiate the diastereotopic aliphatic protons (SI).
Inorganics 10 00103 g004
Figure 5. X-ray crystal structures of: (a) BuRu; (c) BuRh and MM+ simulation of (b) EtRu (only CC forms are represented). Schematic representation of chirality elements (d).
Figure 5. X-ray crystal structures of: (a) BuRu; (c) BuRh and MM+ simulation of (b) EtRu (only CC forms are represented). Schematic representation of chirality elements (d).
Inorganics 10 00103 g005
Table 1. a, b and c values within the self-assembled structures BuRu, EtRu and BuRh.
Table 1. a, b and c values within the self-assembled structures BuRu, EtRu and BuRh.
Alkyl Chain
Length a (Å)
Metal-Metal
Distance b (Å)
c = b − 2a
(Å)
BuRu5.08.4−1.6
EtRu2.68.42.2
BuRh5.012.92.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Benchohra, A.; Séjourné, S.; Labrunie, A.; Miller, L.; Charbonneau, E.; Carré, V.; Aubriet, F.; Allain, M.; Sallé, M.; Goeb, S. Controlling Chiral Self-Sorting in Truxene-Based Self-Assembled Cages. Inorganics 2022, 10, 103. https://doi.org/10.3390/inorganics10070103

AMA Style

Benchohra A, Séjourné S, Labrunie A, Miller L, Charbonneau E, Carré V, Aubriet F, Allain M, Sallé M, Goeb S. Controlling Chiral Self-Sorting in Truxene-Based Self-Assembled Cages. Inorganics. 2022; 10(7):103. https://doi.org/10.3390/inorganics10070103

Chicago/Turabian Style

Benchohra, Amina, Simon Séjourné, Antoine Labrunie, Liam Miller, Enzo Charbonneau, Vincent Carré, Frédéric Aubriet, Magali Allain, Marc Sallé, and Sébastien Goeb. 2022. "Controlling Chiral Self-Sorting in Truxene-Based Self-Assembled Cages" Inorganics 10, no. 7: 103. https://doi.org/10.3390/inorganics10070103

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop