Next Article in Journal
Review of the Role of Nanotechnology in Overcoming the Challenges Faced in Oral Cancer Diagnosis and Treatment
Next Article in Special Issue
Synthesis, Reactivity and Coordination Chemistry of Group 9 PBP Boryl Pincer Complexes: [(PBP)M(PMe3)n] (M = Co, Rh, Ir; n = 1, 2)
Previous Article in Journal
Medicinal Mushrooms: Their Bioactive Components, Nutritional Value and Application in Functional Food Production—A Review
Previous Article in Special Issue
Recent Progress in Crystalline Borates with Edge-Sharing BO4 Tetrahedra
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

1,8-Dihydroxy Naphthalene—A New Building Block for the Self-Assembly with Boronic Acids and 4,4′-Bipyridine to Stable Host–Guest Complexes with Aromatic Hydrocarbons

Department of Chemistry & Biochemistry, Texas Tech University, P.O. Box 41061, Lubbock, TX 79409-1061, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(14), 5394; https://doi.org/10.3390/molecules28145394
Submission received: 30 April 2023 / Revised: 22 June 2023 / Accepted: 7 July 2023 / Published: 14 July 2023

Abstract

:
The new Lewis acid–base adducts of general formula X(nad)B←NC5H4-C5H4N→B(nad)X [nad = 1,8-O2C10H6, X = C6H5 (2c), 3,4,5-F3-C6H2 (2d)] were synthesized in high yields via reactions of 1,8-dihydroxy naphthalene [nadH2] and 4,4′-bipyridine with the aryl boronic acids C6H5B(OH)2 and 3,4,5-F3-C6H2B(OH)2, respectively, and structurally characterized by multi-nuclear NMR spectroscopy and SCXRD. Self-assembled H-shaped Lewis acid–base adduct 2d proved to be effective in forming thermally stable host–guest complexes, 2d × solvent, with aromatic hydrocarbon solvents such as benzene, toluene, mesitylene, aniline, and m-, p-, and o-xylene. Crystallographic analysis of these solvent adducts revealed host–guest interactions to primarily occur via π···π contacts between the 4,4′-bipyridyl linker and the aromatic solvents, resulting in the formation of 1:1 and 1:2 host–guest complexes. Thermogravimetric analysis of the isolated complexes 2d × solvent revealed their high thermal stability with peak temperatures associated with the loss of solvent ranging from 122 to 147 °C. 2d, when self-assembled in an equimolar mixture of m-, p-, and o-xylene (1:1:1), preferentially binds to o-xylene. Collectively, these results demonstrate the ability of 1,8-dihydroxy naphthalene to serve as an effective building block in the selective self-assembly to supramolecular aggregates through dative covalent N→B bonds.

Graphical Abstract

1. Introduction

The last decades have witnessed increased interest in the rational design of supramolecular architectures based on the self-assembly of various suitable building blocks [1,2,3,4,5,6,7,8,9,10,11,12]. Amongst the various strategies, the utilization of intermolecular dative covalent N→B bond interactions in the design and self-assembly of well-defined supramolecular structures has become a promising synthetic approach with potential in host–guest chemistry, gas storage, and crystal engineering [13,14,15,16]. Examples of this recent development are trigonal planar catechol-based aryl boronic esters [17]. These strongly Lewis acidic building blocks, in combination with N-containing strongly Lewis basic ligands, have been demonstrated to facilitate self-assembly to give—through coordinate covalent N→B bonds—boronic ester-based supramolecular structures, such as coordination polymers, cage- and nanostructures, and gels [18,19,20,21,22,23,24,25,26,27]. Recently, the groups of Hőpfl and Morales-Rojas [28] have shown that reactions of catechol-based aryl boronic esters with bipyridine linkers in the presence of various aromatic hydrocarbon solvents readily form self-assembled Lewis-type N→B adducts with aromatic hydrocarbons incorporated. It was hypothesized that upon N→B coordination, the aromatic bipyridine linker becomes increasingly electron-deficient, enabling the recognition and selective isolation of host–guest complexes with aromatic hydrocarbons [29]. However, catecholate-based boronic esters, despite their high Lewis acidity, readily hydrolyze, particularly when used in “wet” solvents. Our group has recently shown that 1,8-dihydroxy naphthalene-based aryl boronic esters are as strongly Lewis acidic as their catechol-based counterparts but significantly more hydrolytically stable (Scheme 1) [30]. Based on our recent findings, we envisioned that 1,8-dihydroxy naphthalene should be an effective building block in the self-assembly with aryl boronic acids and a suitable bipyridine linker to form thermally stable host–guest complexes with aromatic hydrocarbons.

2. Results and Discussion

At the outset, we investigated the ability of our recently prepared boronic esters of general formula R-Bnad (1ae), where nad = 1,8-naphthtalenediolate (a: R = mesityl; b: R = 2,6-dichlorophenyl; c: R = phenyl; d: R = 3,4,5-trifluorophenyl; e: R = pentafluorophenyl) to form stable Lewis acid–base adducts with 4,4′-bipyridine (Scheme 2). The reactions were performed in acetone-D6 as the solvent and monitored by 11B and 1H NMR spectroscopy. However, the 1H NMR spectra of all five reactions exclusively gave one set of signals, even when an excess of 4,4′-bipyridine or one of the respective boronic esters was used, indicating rapid equilibrium to occur at room temperature. The 11B NMR spectra, on the other hand, showed very different spectral features (Figure 1). Thus, a 1:2 mixture of 4,4′-bipyridine and the bulkiest and least Lewis acidic ester 1a (R = mesityl) showed a signal at around 29 ppm, which is close to the 11B NMR chemical shift of 1a (δ = 27 ppm in CDCl3) [30], showing that the equilibrium strongly favors the individual acid and base components over adduct 2a. For the system 4,4′-bipyridine/1b (1:2), two distinct signals at ca. 27 ppm and 7 ppm were observed; the former can be assigned to ester 1b, while the latter is due to the desired Lewis acid–base adduct 2b. In contrast, mixtures of 4,4′-bipyridine with 1c, 1d, and 1e, respectively, showed up-field shifted signals at ca. 12 ppm (1c), 6 ppm (1d), and 4 ppm (1e), indicating the formation of the respective adducts 2c, 2d, and 2e.
To confirm our structural assignments, efforts were undertaken to grow single crystals suitable for X-ray analysis from concentrated acetone solutions of 4,4′-bipyridine and 2ae. The results for 2a, 2b, and 2d are shown in Figure 2 and reveal that two boronic ester units bind to a 4,4′-bipyridine moiety linked via dative N→B bonds. Compounds 2a and 2b can structurally be described as double-tweezers. 2a exhibits an almost perfect double-tweezer-shaped structure with nearly planar bipyridine units, while 2b is markedly twisted with a twisting angle of the bipyridine moiety of about 35°. Adduct 2d, on the other hand, is composed of a more H-shaped structure with a nearly planar bipyridine unit (twisting angle ~10°). The bond parameters for all three adducts are similar, with B-N and B-O distances ranging from 1.61 to 1.67 Å and 1.45–1.47 Å, respectively. The central boron atoms in 2a, 2b, and 2d have slightly distorted tetrahedral coordination environments, with O-B-O and N-B-O angles ranging from 111–116° to 102–105°. Unfortunately, it was not possible to obtain suitable single crystals of 2c and 2e from acetone as the solvent. Not only was 2e highly soluble in acetone, but it also slowly degraded to a second product exhibiting a relatively sharp signal in the 11B NMR with a chemical shift of ca. 0 ppm (see also Figure 1). This new product crystallized from acetone after a week and was identified by NMR spectroscopy and SCXRD as a spirocyclic 4,4′-bipyridinium borate (see supporting information). However, switching the solvent from acetone to benzene resulted in the formation of single crystalline material, which by X-ray analysis was identified as the phenyl derivative 2c × benzene (Figure 2). Similar to 2d, adduct 2c × benzene has an H-shaped structure with a nearly planar bipyridine unit (twisting angle ~ 5°) and average B-O and B-N distances of 1.45 and 1.67 Å, respectively. Notably, 2c × benzene co-crystallizes with two molecules of benzene, both being entrapped via π–π stacking with the 4,4′-bipyridine moiety and C-Hπ contacts with both the phenyl and the naphthalene substituents.
It was also possible to synthesize and isolate adducts 2c and 2d via a one-pot procedure in excellent yields starting from the respective boronic acids, without the need of preparing the respective esters 1c and 1d (Scheme 3). For example, when two equivalents of 1,8-dihydroxy naphthalene and one equivalent of 4,4′-bipyridine were reacted with two equivalents of the respective aryl boronic acid in acetone as a solvent, 2c and 2d were isolated in yields of 80% and 90%, respectively, as bright orange crystalline materials. Both compounds were fully characterized by 1H, 13C, 19F, and 11B NMR spectroscopy as well as by combustion analysis.
Encouraged by these results, the self-assembly with diboronic acids was investigated. Thus, upon reacting 1,8-dihydroxy naphthalene with 4,4′-bipyridine and 1,4-phenylene diboronic acid in a 2:1:1 ratio in acetone as the solvent, an orange crystalline material formed. However, once precipitated, the isolated material proved to be insoluble in common organic solvents, rendering its characterization by NMR spectroscopy impossible. Attempts to grow single crystals suitable for X-ray analysis failed as well. On the other hand, replacing 4,4′-bipyridine with trans-1,2-di(4-pyridyl)ethylene gave— upon reaction with 1,8-dihydroxy naphthalene and 1,4-phenylene diboronic acid in acetone—an orange microcrystalline precipitate, from which single crystals suitable for X-ray analysis could be obtained. The results revealed the compound to be polymeric Lewis acid–base adduct 3. Polymer 3 is composed of diboronic ester units coordinated with the 4,4′-bipyridine moieties in a zig-zag fashion (Scheme 3, Figure 3a). Due to poor diffraction and multiple twinning of the single crystals measured, only the connectivity of the polymer could be confirmed.
In contrast, the reaction of 1,3-phenylene diboronic acid under otherwise identical conditions did not lead to the formation of a polymeric structure; instead, the rectangular dimer 4 (yield 94%) was formed as confirmed by single-crystal X-ray analysis (Scheme 3, Figure 3b). Both bipyridine units in 4 are markedly twisted, with twisting angles of about 40°. The average B-N and B-O distances are in the expected range of 1.675 and 1.455 Å, respectively.
Encouraged by the ability of 2c × benzene to form a stable host–guest complex with two molecules of benzene, a detailed study of the inclusion properties of H-shaped 2d was undertaken (Scheme 4). 2d was selected because of its enhanced hydrolytic stability in solution and the ability of the fluorine substituents to significantly decrease the electron density at boron, thereby increasing the stability of the Lewis acid–base adducts to be formed. Suspensions of 1,8-dihydroxy naphthalene, 4,4′-bipyridine, and 3,4,5-trifluorophenyl boronic acid in the respective aromatic hydrocarbon solvent (i.e., benzene, toluene, o-xylene, m-xylene, p-xylene, mesitylene, aniline) were heated until clear solutions were obtained. Upon slowly cooling to room temperature, crystalline solids were obtained in isolated yields ranging from 34–81%. In addition to being characterized by 1H NMR spectroscopy and combustion analysis, all compounds were analyzed by X-ray crystallography.
The results of the X-ray analysis are shown in Figure 4 and reveal that 2d is capable of forming various stable host–guest complexes with all aromatic hydrocarbon solvents used in this study via π–π stacking between the 4,4′-bipyridyl unit of 2d and the respective aromatic solvent. Thus, in benzene as a solvent, 2d × benzene is formed (Figure 4a), which is composed of three benzene molecules per two molecules of 2d. Identical results were obtained with aniline as the solvent, generating host–guest complex 2d × aniline (Figure 4b) with a 2d/aniline ratio of 2:3. In contrast, carrying out the self-assembly reactions in toluene and p-xylene gave the host–guest complexes 2d × toluene and 2d × p-xylene (Figure 4c,d), respectively, which consist of one aromatic guest per one molecule of 2d (1:1 complex). Interestingly, crystallization experiments in meta-xylene gave two crystallographically distinct single crystals of 2d × m-xylene; the results for one revealed a classical 1:1 host–guest complex (Figure 4e), while X-ray analysis of the second one revealed a 1:2 complex (Figure 4f) with two additional unbound m-xylene molecules per three equivalents of the 1:2 complex. On the other hand, X-ray analysis of crystalline 2d × o-xylene obtained from o-xylene as a solvent confirmed a 1:2 complex (Figure 4g). The X-ray analysis for 2d × mesitylene (Figure 4h) also confirms a 1:2 host–guest complex but with two molecules of 2d and two molecules of unbound mesitylene in the unit cell.
Table 1 compares the host–guest ratios of 2d × solvent determined from the X-ray analysis of single crystals with those obtained from the 1H NMR data of the bulk materials; the latter were dried in air for circa 24 h. It was found that crystals of 2d × o-xylene and 2d × m-xylene, upon drying in air, lose guest molecules to form host–guest complexes with a formal 1:1 host–guest ratio. While in 2d × aniline and 2d × benzene, the 2:3 host–guest ratio remains unchanged, and 2d × mesitylene contains significantly more mesitylene than what was found via X-ray analysis of single crystalline material. This seems to imply that 2d × mesitylene may crystallize in various forms with different host–guest ratios.
To study the thermal behavior of the host–guest complexes 2d × solvent, a thermogravimetric analysis of the bulk materials was performed; the results are shown in Figure 5, Figures S30 and S31, and Table S1. Figure 5 (bottom) shows the TG curve of 2d, revealing complete weight loss to occur at about 260 °C. This indicates that the acid–base adduct 2d either sublimes or that the decomposition and evaporation of its individual acid and base components occur within the same temperature range. The TG graphs of all host–guest compounds exhibit a two-step weight loss. This is exemplarily shown for the solvent adducts 2d × benzene, 2d × toluene, 2d × o-xylene, 2d × m-xylene, and 2d × p-xylene (Figure 5), for which the observed weight losses are consistent with the stoichiometry of the bulk materials established by 1H NMR spectroscopy (Table 1). After the loss of the guest, the residual material likely consists of the solid phase established for solvent-free 2d. In this series, for 2d × benzene and 2d × toluene, the peak temperatures (Tpeak) associated with the liberation of the guest were found to be 131 °C and 147 °C, respectively, suggesting stronger host–guest interactions in 2d × toluene, and consistent with the results of DSC analysis (Figures S32 and S33). For the xylene series, the peak temperatures for the loss of solvent are 122 °C (2d × m-xylene), 133 °C (2d × o-xylene), and 136 °C (2d × p-xylene). The slightly higher peak temperatures for 2d × o-xylene and 2d × p-xylene suggest somewhat stronger host–guest interactions compared to 2d × m-xylene, which is supported by DSC measurements (Figures S34–S36).
In a related study, Hőpfl and Morales-Rojas recently disclosed the thermal properties of self-assembled host–guest complexes derived from reactions of 1,2-catechol and 3,4,5-trifluorphenylboronic acid with 4,4′-bipyridine in various aromatic hydrocarbon solvents [28]. Thermogravimetric analysis of these catechol-based host–guest complexes revealed the peak temperatures associated with the loss of solvent to be 65 °C (m-xylene), 71 °C (o-xylene), 76 °C (p-xylene), 121 °C (toluene), 139 °C (benzene), and 100 °C (mesitylene). Except for benzene, these values are significantly lower than what was measured for our 1,8-dihydroxy naphthalene-based host–guest complexes 2d × solvent (Figure 5). The replacement of catechol with the 1,8-dihydroxy naphthalene ligand at the central boron appears to result in stronger π-type host–guest interactions between the 4,4′-bipyridine unit and the aromatic solvent. Unfortunately, Hőpfl and Morales-Rojas did not disclose thermodynamic data from DSC measurements for comparison [28].
Based on the results from the host–guest chemistry with m-, o-, and p-xylene, we wondered whether self-assembled 2d would selectively incorporate one of the three xylene isomers. Thus, 1,8-dihydroxy naphthalene, 4,4′-bipyridine, and 3,4,5-trifluoroboronic acid were reacted in an equimolar solvent mixture of m-, p-, and o-xylene (1:1:1). The obtained crystalline material was collected and analyzed by 1H NMR spectroscopy using DMSO-D6 as the solvent (see supporting information for more details). The results revealed a higher selectivity of 2d for o-xylene (ca. 57%), with an m-, p-, to o-xylene ratio of about 2:1:4, respectively, which is similar to those of the respective catechol-based host–guest complexes reported by Hőpfl and Morales-Rojas [28]. Considering the results from the TGA and DSC measurements, the fairly low uptake of p-xylene appears to be surprising but perhaps can be attributed to kinetic effects during the crystallization process.
Finally, to further broaden the scope of the host–guest approach, the modified bipyridine linker 1,2-bis(4-pyridyl)ethane was employed in reactions with 1,8-dihydroxy naphthalene and various boronic acids. Unfortunately, in none of the cases could host–guest complexes with aromatic hydrocarbon solvents be crystallized and structurally determined. Nonetheless, single crystals of the solvent-free double-tweezers 5 and 6 could be obtained from acetone solutions and were characterized by single-crystal X-ray crystallography (Figure 6a,b).

3. Conclusions

We have demonstrated the ability of 1,8-dihydroxy naphthalene to self-assemble with various aryl boronic acids and 4,4′-bipyridine as a linker to give double-tweezer and H-shaped Lewis acid–base adducts in very good isolated yields by means of N→B bond formation. In addition, a series of host–guest complexes, 2d × solvent, was prepared via reactions of 4,4′-bipyridine, 1,8-dihydroxy naphthalene, and 3,4,5-trifluorophenyl boronic acid with various aromatic hydrocarbons. SCXRD analysis of these solvent adducts revealed host–guest interactions to primarily occur via π···π contacts between the 4,4′-bipyridyl linker and the aromatic solvents to give 1:1 and 1:2 host–guest complexes, which was confirmed by 1H NMR spectroscopic analysis of the bulk samples. TGA measurements of the bulk samples showed well-defined weight losses at peak temperatures ranging from 122 to 147 °C, which in every case could be attributed to the complete loss of solvent (guest). Self-assembled Lewis acid–base adduct 2d, when recrystallized from an equimolar mixture of m-, p-, and o-xylene (1:1:1), preferentially binds to o-xylene, generating 2d × o-xylene as the major component, similar to what was reported for structurally related catechol-based host–guest complexes [28]. Collectively, the present contribution demonstrates the effectiveness of 1,8-dihydroxy naphthalene in constructing supramolecular structures via dative N-B bond formation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28145394/s1. Figure S1: 1H NMR spectrum of 2c (acetone-d6); Figure S2: 13C NMR spectrum of 2c (acetone-d6); Figure S3: 11B NMR spectrum of 2c (acetone-d6); Figure S4: 11B NMR spectrum of 2c (DMSO-d6); Figure S5: 1H NMR spectrum of 2c × benzene in DMSO-d6; Figure S6: 1H NMR spectrum 2d (acetone-d6); Figure S7: 13C NMR spectrum of 2d (acetone-d6); Figure S8: 19F NMR spectrum of 2d (acetone-d6); Figure S9: 11B NMR spectrum of 2d (acetone-d6); Figure S10: 1H NMR spectrum of 4,4-bipyridinium-bis(1,8-naphthalenediolato)borate (DMSO-d6); Figure S11: 11B NMR spectrum of 4,4-bipyridinium-bis(1,8-naphthalenediolato)borate (DMSO-d6); Figure S12: 13C NMR spectrum of 4,4-bipyridinium-bis(1,8-naphthalenediolato)borate (DMSO-d6); Figure S13: 13C NMR-DEPT spectrum of 4,4-bipyridinium-bis(1,8-naphthalenediolato) borate (DMSO-d6); Figure S14: 1H NMR spectrum of 1,4-benzene-diboronic-acid-1,8-naphthalenediolate ester in DMSO-d6; Figure S15: 13C NMR spectrum of 1,4-benzene-diboronic-acid-1,8-naphthalenediolate ester in DMSO-d6; Figure S16: 11B NMR spectrum of 1,4-benzene-diboronic-acid-1,8-naphthalenediolate ester in DMSO-d6; Figure S17: 11B NMR spectrum of 1,4-benzene-diboronic-acid-1,8-naphthalenediolate ester in acetone-d6; Figure S18: 1H NMR spectrum of 4 in DMSO-d6; Figure S19: 1H NMR spectrum of 2d × benzene in acetone-d6; Figure S20: 1H NMR spectrum of 2d × toluene in DMSO-d6; Figure S21: 1H NMR spectrum of 2d × m-xylene in DMSO-d6; Figure S22: 1H NMR spectrum of 2d × p-xylene in DMSO-d6; Figure S23: 1H NMR spectrum of 2d × o-xylene in DMSO-d6; Figure S24: 1H NMR spectrum of 2d × mesitylene in DMSO-d6; Figure S25: 1H NMR spectrum of 2d × aniline in DMSO-d6; Figure S26: 1H NMR spectrum (DMSO-d6) of adduct 2d × xylene isolated from a 1:1:1 mixture of m-, p-, and o-xylene; Figure S27: Aliphatic region of the 1H NMR spectrum (DMSO-d6) of adduct 2d × xylene isolated from a 1:1:1 mixture of m-, p-, and o-xylene; Figure S28: 1H NMR spectrum (DMSO-d6) of 2d × m-xylene after soaking in o-xylene for 24 h; Figure S29: 1H NMR spectrum (DMSO-d6) of 2d × p-xylene after soaking in o-xylene for 24 h; Figure S30: Thermogravimetric analysis of 2d × mesitylene with weight loss [%] and peak temperature Tpeak [°C] for the loss of solvent; Figure S31: Thermogravimetric analysis of 2d × aniline with weight loss [%] and peak temperature Tpeak [°C] for the loss of solvent; Figure S32: DSC analysis of 2d × benzene for the loss of solvent; Figure S33: DSC analysis of 2d × toluene for the loss of solvent; Figure S34: DSC analysis of 2d × p-xylene for the loss of solvent; Figure S35: DSC analysis of 2d × o-xylene for the loss of solvent; Figure S36: DSC analysis of 2d × m-xylene for the loss of solvent; Table S1: Selected information from the TGA analysis for 2d × solvent; Table S2: Crystal data and structure refinement for 2d × toluene; Table S3: Crystal data and structure refinement for 6; Table S4: Crystal data and structure refinement for 2b; Table S5: Crystal data and structure refinement for 2d × benzene; Table S6: Crystal data and structure refinement for 2d × p-xylene; Table S7: Crystal data and structure refinement for 2d × m-xylene; Table S8: Crystal data and structure refinement for 2d × aniline; Table S9: Crystal data and structure refinement for 5; Table S10: Crystal data and structure refinement for 2d; Table S11: Crystal data and structure refinement for 3; Table S12: Crystal data and structure refinement for 2d × m-xylene; Table S13: Crystal data and structure refinement for 2c; Table S14: Crystal data and structure refinement for 2a; Table S15: Crystal data and structure refinement for 2d × mesitylene; Table S16: Crystal data and structure refinement for 2d × o-xylene; Table S17: Crystal data and structure refinement for 4. References [31,32,33,34,35,36,37,38] are cited in the supplementary materials.

Author Contributions

Conceptualization, C.K.; methodology, C.P.M. and R.P.; formal analysis, C.P.M., R.P. and D.K.U.; investigation, C.P.M. and R.P.; writing—review and editing, C.K.; supervision, C.K.; project administration, C.K.; funding acquisition, C.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Catalysis Science Program, under Award DE-SC0019094.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Desiraju, G.R. Supramolecular synthons in crystal engineering—A new organic synthesis. Angew. Chem. Int. Ed. Engl. 1995, 34, 2311–2327. [Google Scholar] [CrossRef]
  2. 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. [Google Scholar] [CrossRef]
  3. Hosseini, M.W. Molecular tectonics: From simple tectons to complex molecular networks. Acc. Chem. Res. 2005, 38, 313–323. [Google Scholar] [CrossRef] [PubMed]
  4. Mastalerz, M. Shape-Persistent Organic Cage Compounds by Dynamic Covalent Bond Formation. Angew. Chem. Int. Ed. 2010, 49, 5042–5053. [Google Scholar] [CrossRef]
  5. Chakrabarty, R.; Mukherjee, P.S.; Stang, P.J. Supramolecular coordination: Self-assembly of finite two- and three-dimensional ensembles. Chem. Rev. 2011, 111, 6810–6918. [Google Scholar]
  6. Smulders, M.M.J.; Riddell, I.A.; Browne, C.; Nitschke, J.R. Building on architectural principles for three-dimensional metallosupramolecular construction. Chem. Soc. Rev. 2013, 42, 1728–1754. [Google Scholar] [CrossRef]
  7. Goesten, M.G.; Kapteijn, F.; Gascon, J. Fascinating chemistry or frustrating unpredictability: Observations in crystal engineering of metal−organic frameworks. CrystEngComm 2013, 15, 9249–9257. [Google Scholar] [CrossRef] [Green Version]
  8. Harris, K.; Fujita, D.; Fujita, M. Giant hollow MnL2n spherical complexes: Structure, functionalisation and applications. Chem. Commun. 2013, 49, 6703–6712. [Google Scholar] [CrossRef]
  9. Song, M.; Sun, Z.; Han, C.; Tian, D.; Li, H.; Kim, J.S. Calixarene based chemosensors by means of click chemistry. Chem. Asian J. 2014, 9, 2344–2357. [Google Scholar] [CrossRef]
  10. Durola, F.; Heitz, V.; Reviriego, F.; Roche, C.; Sauvage, J.-P.; Sour, A.; Trolez, Y. Cyclic [4]rotaxanes containing two parallel porphyrinic plates: Toward switchable molecular receptors and Compressors. Acc. Chem. Res. 2014, 47, 633–645. [Google Scholar] [CrossRef] [PubMed]
  11. Zhang, G.; Mastalerz, M. Organic cage compounds—From shape-persistency to function. Chem. Soc. Rev. 2014, 43, 1934–1947. [Google Scholar] [CrossRef]
  12. Li, H.; Yao, Z.-J.; Liu, D.; Jin, G.-X. Multi-component coordination-driven self-assembly toward heterometallic macrocycles and cages. Coord. Chem. Rev. 2015, 293–294, 139–157. [Google Scholar]
  13. Torres-Huerta, A.; Velásquez-Hernández, M.; Martínez-Otero, D.; Höpfl, H.; Jancik, V. Structural induction via solvent variation in assemblies of triphenylboroxine and piperazine—Potential application as self-assembly molecular sponge. Cryst. Growth Des. 2017, 17, 2438–2452. [Google Scholar] [CrossRef]
  14. Wang, W.; Wang, L.; Du, F.; Wang, G.-D.; Hou, L.; Zhu, Z.; Liu, B.; Wang, Y.-Y. Dative B←N bonds based crystalline organic framework with permanent porosity for acetylene storage and separation. Chem. Sci. 2023, 14, 533–539. [Google Scholar]
  15. Stephens, A.J.; Scopelliti, R.; Tirani, F.F.; Solari, E.; Severin, K. Crystalline Polymers Based on Dative Boron-Nitrogen Bonds and the Quest for Porosity. ACS Mater. Lett. 2019, 1, 3–7. [Google Scholar] [CrossRef]
  16. Bhandary, S.; Shukla, R.; Van Hecke, K. Effect of chemical substitution on the construction of boroxine-based supramolecular crystalline polymers featuring B←N dative bonds. CrystEngComm 2022, 24, 1695–1699. [Google Scholar] [CrossRef]
  17. Madura, I.D.; Czerwinska, K.; Jakubczyk, M.; Pawelko, A.; Adamczyk-Wozniak, A.; Sporzynski, A. Weak C–H···O and Dipole–Dipole Interactions as Driving Forces in Crystals of Fluorosubstituted Phenylboronic Catechol Esters. Cryst. Growth Des. 2013, 13, 5344–5352. [Google Scholar] [CrossRef]
  18. Luisier, N.; Scopelliti, R.; Severin, K. Supramolecular gels based on boronate esters and imidazolyl donors. Soft Matter 2016, 12, 588–593. [Google Scholar] [CrossRef] [PubMed]
  19. Salazar-Mendoza, D.; Cruz-Huerta, J.; Höpfl, H.; Hernández-Ahuactzi, I.F.; Sanchez, M. Macrocycles and coordination polymers derived from self-complementary tectons based on N-containing boronic acids. Cryst. Growth Des. 2013, 13, 2441–2454. [Google Scholar] [CrossRef]
  20. Ray, K.K.; Campillo-Alvarado, G.; Morales-Rojas, H.; Hopfl, H.; MacGillivray, L.R.; Tivanski, A.V. Semiconductor Cocrystals Based on Boron: Generated Electrical Response with π-Rich Aromatic Molecules. Cryst. Growth Des. 2020, 20, 3–8. [Google Scholar] [CrossRef]
  21. Icli, B.; Solari, E.; Kilbas, B.; Scopelliti, R.; Severin, K. Multicomponent Assembly of Macrocycles and Polymers by Coordination of Pyridyl Ligands to 1,4-Bis(benzodioxaborole)benzene. Chem. Eur. J. 2012, 18, 14867–14874. [Google Scholar] [CrossRef]
  22. Christinat, N.; Scopelliti, R.; Severin, K. Boron-based rotaxanes by multicomponent self-assembly. Chem. Commun. 2008, 3660–3662. [Google Scholar] [CrossRef] [PubMed]
  23. Hartwick, C.J.; Yelgaonkar, S.P.; Reinheimer, E.W.; Campillo-Alvarado, G.; MacGillivray, L.R. Self-Assembly of Diboronic Esters with U-Shaped Bipyridines: “Plugin-Socket” Assemblies. Cryst. Growth Des. 2021, 21, 4482–4487. [Google Scholar] [CrossRef] [PubMed]
  24. Christinat, N.; Croisier, E.; Scopelliti, R.; Cascella, M.; Rothlisberger, U.; Severin, K. Formation of boronate ester polymers with efficient intrastrand charge-transfer transitions by three-component reactions. Eur. J. Inorg. Chem. 2007, 2007, 5177–5181. [Google Scholar] [CrossRef]
  25. Herrera-España, A.D.; Campillo-Alvarado, G.; Román-Bravo, P.; Herrera-Ruiz, D.; Höpfl, H.; Morales-Rojas, H. Selective Isolation of Polycyclic Aromatic Hydrocarbons by Self-Assembly of a Tunable N→B Clathrate. Cryst. Growth Des. 2015, 15, 1572–1576. [Google Scholar] [CrossRef]
  26. Dhara, A.; Beuerle, F. Reversible Assembly of a Supramolecular Cage Linked by Boron–Nitrogen Dative Bonds. Chem. Eur. J. 2015, 21, 17391–17396. [Google Scholar] [CrossRef]
  27. Icli, B.; Sheepwash, E.; Riis-Johannessen, T.; Schenk, K.; Filinchuk, Y.; Scopelliti, R.; Severin, K. Dative boron–nitrogen bonds in structural supramolecular chemistry: Multicomponent assembly of prismatic organic cages. Chem. Sci. 2011, 2, 1719–1721. [Google Scholar]
  28. Campillo-Alvarado, G.; Vargas-Olvera, E.C.; Hopfl, H.; Herrera-Espana, A.D.; Sanchez-Guadarrama, O.; Morales-Rojas, H.; MacGillivray, L.R.; Rodriguez-Molina, B.; Farfan, N. Self-Assembly of Fluorinated Boronic Esters and 4,4′-Bipyridine into 2:1 N→B Adducts and Inclusion of Aromatic Guest Molecules in the Solid State: Application for the Separation of o,m,p-Xylene. Cryst. Growth Des. 2018, 18, 2726–2743. [Google Scholar] [CrossRef]
  29. Herrera-Espana, A.D.; Hoepfl, H.; Morales-Rojas, H. Boron-Nitrogen Double Tweezers Comprising Arylboronic Esters and Diamines: Self-Assembly in Solution and Adaptability as Hosts for Aromatic Guests in the Solid State. ChemPlusChem 2020, 85, 548–560. [Google Scholar] [CrossRef] [PubMed]
  30. Manankandayalage, C.P.; Unruh, D.K.; Krempner, C. Boronic, diboronic and boric acid esters of 1,8-naphthalenediol–synthesis, structure and formation of boronium salts. Dalton Trans. 2020, 49, 4834–4842. [Google Scholar] [CrossRef]
  31. Bruker (2021) SADABS v2016/2, Bruker AXS Inc.: Madison, WI, USA, 2015.
  32. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, C71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Spek, A.L. Structure validation in chemical crystallography. Acta Crystallogr. Sect. D Biol. Crystallogr. 2009, D65, 148–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Spek, A.L. PLATON SQUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Crystallogr. Sect. C Struct. Chem. 2015, C71, 9–18. [Google Scholar] [CrossRef] [Green Version]
  35. CrysAlisPRO. Oxford Diffraction; Agilent Technologies UK Ltd.: Yarnton, UK, 2018. [Google Scholar]
  36. SCALE3 ABSPACK, Oxford Diffraction Ltd.: Abingdon, UK, 2005.
  37. Sheldrick, G.M. SHELXT–Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, A71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef]
Scheme 1. Formation of catechol- and 1,8-dihydroxy naphthalene-based aryl boronic esters.
Scheme 1. Formation of catechol- and 1,8-dihydroxy naphthalene-based aryl boronic esters.
Molecules 28 05394 sch001
Scheme 2. Equilibrium between 4,4′-bipyridine/1ae and the Lewis acid–base adducts 2ae in dilute acetone-d6 at room temperature.
Scheme 2. Equilibrium between 4,4′-bipyridine/1ae and the Lewis acid–base adducts 2ae in dilute acetone-d6 at room temperature.
Molecules 28 05394 sch002
Figure 1. 11B NMR spectroscopic study of the equilibrium reaction of the esters 1ae with 4,4′-bipyridine in acetone-D6 at room temperature.
Figure 1. 11B NMR spectroscopic study of the equilibrium reaction of the esters 1ae with 4,4′-bipyridine in acetone-D6 at room temperature.
Molecules 28 05394 g001
Figure 2. Solid-state structures of the 4,4′-bipyridine adducts 2a, 2b, 2d (co-crystallizing acetone molecules omitted for clarity), and 2c × benzene. Color codes: red = oxygen, orange = boron, green = fluorine, white = hydrogen, black = carbon, blue = nitrogen. Selected bond lengths [Å] and angles [°]: (a) 2a, O1 B1 1.476(2), O2 B1 1.462(2), N1 B1 1.627(2), O2 B1 O1 110.9(1), O2 B1 N1 104.7(1), O1 B1 N1 100.3(1); (b) 2b, O1 B1 1.448(2), O2 B1 1.449(2), O3 B2 1.453(2), O4 B2 1.442(2), N1 B1 1.611(2), N2 B2 1.650(2), O1 B1 O2 112.1(1), O1 B1 N1 104.9(1), O2 B1 N1 105.4(1), O4 B2 N2 106.2(1), O3 B2 N2 102.5(1), O4 B2 O3 116.2(1); (c) 2d, B1 O2 1.449(2), B1 O1 1.453(2), B1 N1 1.675(2), B2 O4 1.447(2), B2 O3 1.452(2), B2 N2 1.669(2), O2 B1 N1 105.9(1), O1 B1 N1 104.5(1), O2 B1 O1 115.0(1), O4 B2 N2 104.4(1), O3 B2 N2 105.9(1), O4 B2 O3 115.2(1); (d) 2c × benzene, O1 B1 1.455(2), O2 B1 1.456(2), O3 B2 1.458(2), O4 B2 1.453(2), N1 B1 1.673(2), N2 B2 1.670(2), O1 B1 O2 114.1(1), O1 B1 N1 104.8(1), O2 B1 N1 104.4(1), O4 B2 O3 114.1(1), O4 B2 N2 104.9(1), O3 B2 N2 104.5(1).
Figure 2. Solid-state structures of the 4,4′-bipyridine adducts 2a, 2b, 2d (co-crystallizing acetone molecules omitted for clarity), and 2c × benzene. Color codes: red = oxygen, orange = boron, green = fluorine, white = hydrogen, black = carbon, blue = nitrogen. Selected bond lengths [Å] and angles [°]: (a) 2a, O1 B1 1.476(2), O2 B1 1.462(2), N1 B1 1.627(2), O2 B1 O1 110.9(1), O2 B1 N1 104.7(1), O1 B1 N1 100.3(1); (b) 2b, O1 B1 1.448(2), O2 B1 1.449(2), O3 B2 1.453(2), O4 B2 1.442(2), N1 B1 1.611(2), N2 B2 1.650(2), O1 B1 O2 112.1(1), O1 B1 N1 104.9(1), O2 B1 N1 105.4(1), O4 B2 N2 106.2(1), O3 B2 N2 102.5(1), O4 B2 O3 116.2(1); (c) 2d, B1 O2 1.449(2), B1 O1 1.453(2), B1 N1 1.675(2), B2 O4 1.447(2), B2 O3 1.452(2), B2 N2 1.669(2), O2 B1 N1 105.9(1), O1 B1 N1 104.5(1), O2 B1 O1 115.0(1), O4 B2 N2 104.4(1), O3 B2 N2 105.9(1), O4 B2 O3 115.2(1); (d) 2c × benzene, O1 B1 1.455(2), O2 B1 1.456(2), O3 B2 1.458(2), O4 B2 1.453(2), N1 B1 1.673(2), N2 B2 1.670(2), O1 B1 O2 114.1(1), O1 B1 N1 104.8(1), O2 B1 N1 104.4(1), O4 B2 O3 114.1(1), O4 B2 N2 104.9(1), O3 B2 N2 104.5(1).
Molecules 28 05394 g002
Scheme 3. One-pot synthesis of 2c, 2d, 3, and 4.
Scheme 3. One-pot synthesis of 2c, 2d, 3, and 4.
Molecules 28 05394 sch003
Figure 3. Solid-state structures of 3 (a) (disordered DMSO molecules omitted for clarity) and 4 (b) (disordered acetone molecules omitted for clarity). Color codes: red = oxygen, orange = boron, white = hydrogen, black = carbon, blue = nitrogen. Selected bond lengths in [Å] and angles [°]: 4, O1 B1 1.454(2), O2 B1 1.458(2), O3 B2 1.451(2), O4 B2 1.457(2), N1 B1 1.672(2), N2 B2 1.678(2), O1 B1 O2 114.9(1), O1 B1 N1 105.6(1), O2 B1 N1 105.3(1), O3 B2 O4 114.9(1), O3 B2 N2 105.1(1), O4 B2 N2 105.9(1).
Figure 3. Solid-state structures of 3 (a) (disordered DMSO molecules omitted for clarity) and 4 (b) (disordered acetone molecules omitted for clarity). Color codes: red = oxygen, orange = boron, white = hydrogen, black = carbon, blue = nitrogen. Selected bond lengths in [Å] and angles [°]: 4, O1 B1 1.454(2), O2 B1 1.458(2), O3 B2 1.451(2), O4 B2 1.457(2), N1 B1 1.672(2), N2 B2 1.678(2), O1 B1 O2 114.9(1), O1 B1 N1 105.6(1), O2 B1 N1 105.3(1), O3 B2 O4 114.9(1), O3 B2 N2 105.1(1), O4 B2 N2 105.9(1).
Molecules 28 05394 g003
Scheme 4. Host–guest chemistry of 2c with various aromatic hydrocarbon solvents.
Scheme 4. Host–guest chemistry of 2c with various aromatic hydrocarbon solvents.
Molecules 28 05394 sch004
Figure 4. Solid-state structures of host–guest complexes of 2c with various aromatic solvents (color codes: red = oxygen, orange = boron, white = hydrogen, black = carbon, blue = nitrogen, green = fluorine): (a) 2d × benzene; (b) 2d × aniline, (c) 2d × toluene, (d) 2d × p-xylene; (e) 2d × m-xylene; (f) 2d × m-xylene (unbound m-xylene omitted for clarity); (g) 2d × o-xylene; (h) 2d × mesitylene (unbound mesitylene omitted for clarity).
Figure 4. Solid-state structures of host–guest complexes of 2c with various aromatic solvents (color codes: red = oxygen, orange = boron, white = hydrogen, black = carbon, blue = nitrogen, green = fluorine): (a) 2d × benzene; (b) 2d × aniline, (c) 2d × toluene, (d) 2d × p-xylene; (e) 2d × m-xylene; (f) 2d × m-xylene (unbound m-xylene omitted for clarity); (g) 2d × o-xylene; (h) 2d × mesitylene (unbound mesitylene omitted for clarity).
Molecules 28 05394 g004
Figure 5. Thermogravimetric analysis of 2d and its solvent adducts 2d × solvent with corresponding weight loss [%] and peak temperature Tpeak [°C] for the loss of solvent.
Figure 5. Thermogravimetric analysis of 2d and its solvent adducts 2d × solvent with corresponding weight loss [%] and peak temperature Tpeak [°C] for the loss of solvent.
Molecules 28 05394 g005
Figure 6. Solid-state structures of 5 and 6 (color codes: red = oxygen, orange = boron, white = hydrogen, black = carbon, blue = nitrogen). Selected bond lengths [Å] and angles [°]: (a) 5, O1 B1 1.447(4), O2 C3 1.367(3), O2 B1 1.466(4), N1 B1 1.656(4), N2 B2 1.682(4), O1 B1 O2 114.4(3), O1 B1 N1 105.4(2), O2 B1 N1 104.7(2); (b) 6, O1 B1 1.471(2), O2 B1 1.467(2), N1 B1 1.622(2), O2 B1 O1 112.3(1), O2 B1 N1 106.0(1), O1 B1 N1 99.7(1).
Figure 6. Solid-state structures of 5 and 6 (color codes: red = oxygen, orange = boron, white = hydrogen, black = carbon, blue = nitrogen). Selected bond lengths [Å] and angles [°]: (a) 5, O1 B1 1.447(4), O2 C3 1.367(3), O2 B1 1.466(4), N1 B1 1.656(4), N2 B2 1.682(4), O1 B1 O2 114.4(3), O1 B1 N1 105.4(2), O2 B1 N1 104.7(2); (b) 6, O1 B1 1.471(2), O2 B1 1.467(2), N1 B1 1.622(2), O2 B1 O1 112.3(1), O2 B1 N1 106.0(1), O1 B1 N1 99.7(1).
Molecules 28 05394 g006
Table 1. Molar 2d/solvent ratios of 2d × solvent from SCXRD and 1H NMR data.
Table 1. Molar 2d/solvent ratios of 2d × solvent from SCXRD and 1H NMR data.
2d × Solvent2d/Solvent Ratio
(from X-ray Data)
2d/Solvent Ratio
(from 1H NMR Data) 1
2d × benzene2:31:1.5
2d × toluene1:11:1
2d × o-xylene1:21:1
2d × m-xylene 23:6 3 and 1:11:1
2d × p-xylene1:11:1
2d × mesitylene1:2 41:2.66
2d × aniline2:31:1.5
1 The isolated crystalline materials were dried in air for 24 h; 2 Two crystallographically distinct crystals were found; 3 Crystal contains two additional unbound m-xylene molecules; 4 Crystal contains two molecules of 2d and two molecules of unbound mesitylene.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Manankandayalage, C.P.; Unruh, D.K.; Perry, R.; Krempner, C. 1,8-Dihydroxy Naphthalene—A New Building Block for the Self-Assembly with Boronic Acids and 4,4′-Bipyridine to Stable Host–Guest Complexes with Aromatic Hydrocarbons. Molecules 2023, 28, 5394. https://doi.org/10.3390/molecules28145394

AMA Style

Manankandayalage CP, Unruh DK, Perry R, Krempner C. 1,8-Dihydroxy Naphthalene—A New Building Block for the Self-Assembly with Boronic Acids and 4,4′-Bipyridine to Stable Host–Guest Complexes with Aromatic Hydrocarbons. Molecules. 2023; 28(14):5394. https://doi.org/10.3390/molecules28145394

Chicago/Turabian Style

Manankandayalage, Chamila P., Daniel K. Unruh, Ryan Perry, and Clemens Krempner. 2023. "1,8-Dihydroxy Naphthalene—A New Building Block for the Self-Assembly with Boronic Acids and 4,4′-Bipyridine to Stable Host–Guest Complexes with Aromatic Hydrocarbons" Molecules 28, no. 14: 5394. https://doi.org/10.3390/molecules28145394

Article Metrics

Back to TopTop