Next Article in Journal
Engineering Novel Amphiphilic Platinum(IV) Complexes to Co-Deliver Cisplatin and Doxorubicin
Previous Article in Journal
Ubiquitination and De-Ubiquitination in the Synthesis of Cow Milk Fat: Reality and Prospects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Preparation and Crystal Structures of Octaoxoketocalix[8]arene Derivatives: The Ketocalixarene Counterparts of the Largest “Major” Calixarene

Institute of Chemistry, The Hebrew University of Jerusalem, Jerusalem 9190401, Israel
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(17), 4094; https://doi.org/10.3390/molecules29174094
Submission received: 4 July 2024 / Revised: 18 August 2024 / Accepted: 26 August 2024 / Published: 29 August 2024
(This article belongs to the Section Organic Chemistry)

Abstract

:
The purpose of this study was to synthesize and structurally characterize ketocalixarenes (i.e., calixarenes where the bridging methylene bridges are replaced by carbonyl groups) derived from the largest “major” calixarene, namely p-tert-butylcalix[8]arene 3a. Ketocalix[8]arenes were synthesized by the oxidation of protected p-tert-butylcalix[8]arene derivatives. Octamethoxy-p-tert-butylketocalix[8]arene 6b was prepared by the photochemical reaction of the calixarene 3b with NBS in a CHCl3/H2O mixture. The oxidation of the methylene groups of octaacetoxy-p-tert-butylcalix[8]arene 3c was conducted by a reaction with CrO3 in Ac2O/AcOH. The basic hydrolysis of the acetate groups of the oxidation product yielded octahydroxy-p-tert-butylketocalix[8]arene 6a. In the crystal, the molecule adopts a saddle-like conformation of crystallographic C2 and idealized S4 symmetry. Strikingly, the array of OH/OH intramolecular hydrogen bonds present in the parent 3a is completely disrupted in 6a.

1. Introduction

Calix[n]arenes are synthetic macrocycles consisting of n phenol rings interconnected by methylene units [1,2,3,4,5,6]. These systems possess a well-defined cavity capable of hosting smaller molecules or ions. During the last few decades, calixarenes have received considerable attention because of their multitude of potential applications [6]. Their preparation usually involves the base-catalyzed condensation of p-tert-butylphenol and formaldehyde. Only three calixarenes were initially synthetically accessible in both multigram scale and high yields: p-tert-butylcalix[4]arene (1a), p-tert-butylcalix[6]arene (2a) and p-tert-butylcalix[8]arene (3a) [7,8,9,10]. These systems were dubbed by Gutsche as “major calixarenes”(Scheme 1) [11]. Substituents are usually introduced at the scaffold via the electrophilic substitution of aromatic rings and/or via the alkylation or acylation of phenolic hydroxyl groups [12]. These modifications usually aim to alter the chemical properties, three-dimensional shape and rigidity of the macrocycle. A structural modification which is relatively unexplored involves the oxidation of methylene bridges to carbonyl groups.
Calixarenes where phenol rings are interconnected by carbonyl groups (“ketocalixarenes”) are of both stereochemical and synthetic interest [13]. The formal replacement of the bridging methylene groups by carbonyls changes the conformational preferences of the macrocycle. Remarkably, whereas the parent p-tert-butylcalix[4]arene (1a) adopts a cone conformation stabilized by a circular array of hydrogen bonds [14], tetrahydroxyketocalix[4]arene 4a adopts in the crystal and in solution a 1,3-alternate conformation (Figure 1) [15,16]. The carbonyl groups are not involved in intramolecular hydrogen bonding, and it was conjectured that the conformational change is due to dipole–dipole repulsion in the cone form and the increased conjugation of aryl rings with carbonyl groups in the conformation adopted [15].
Tetrahydroxyketocalix[4]arene 4a (Scheme 2) was first synthesized in 1990 by Görmar and coworkers [17]. The reaction sequence involved the acetylation of the hydroxyl groups of 1a (to protect the phenol moieties during the subsequent oxidation step), CrO3 oxidation of the methylene groups of 1c and hydrolysis of the acetoxy groups [17,18]. Starting from hexahydroxycalix[6]arene 2a, a similar synthetic sequence was utilized for the preparation of hexahydroxyketocalix[6]arene 5a [19]. There is only a single report on the attempted oxidation of the methylenes of a derivative of calix[8]arene 3a which is the largest “classical” calixarene. In a pioneering study, Ninagawa and coworkers reported in 1985 on the CrO3 oxidation of the methylene groups of 3c in an Ac2O/AcOH mixture at 45 °C. A calix[8]arene derivative was obtained with three methylene bridges oxidized to carbonyls, but the identity of the structural isomer isolated was not determined [20]. The goal of this work was to synthesize and structurally characterize the derivatives of p-tert-butylcalix[8]arene, where all methylene bridges were oxidized to carbonyl groups.

2. Results

2.1. General Considerations

In all cases reported where the methylene groups of a calix scaffold were oxidized to carbonyl groups, the calixarene phenolic OH groups were protected as methyl ether (i.e., 1b, 2b) or acetoxy (1c, 2c) groups. This protection is necessary to avoid the oxidation of the phenol rings (e.g., to yield calixquinones [21] or 5,5’-bicalix[n]arenes [22]). In general, methyl ethers functionalities are considered more robust protecting groups of phenolic hydroxyls. This is of importance if, after oxidation, a further chemical modification of carbonyl bridges is planned on the protected derivative (e.g., by reaction with an organometallic reagent). For example, the reaction of tetramethoxyketocalix[4]arene 4b with PhLi [23] or MeLi [24] proceeds, readily affording the corresponding tetra-addition product. However, in our previous experience, we found that the deprotection of all the OH groups by the cleavage of the O-Me bonds in a methylene functionalized system sometimes fails, even under strenuous reaction conditions. Labile acetoxy protecting groups, although incompatible with organometallic reagents, can be readily cleaved under basic hydrolysis conditions and unmask protected OH groups.

2.2. Oxidation of Octamethylether of Calix[8]arene

In 2013, Fischer and coworkers reported on a facile method for the oxidation of the methylene groups of tetramethoxycalix[4]arene 1b. The photochemical reaction of 1b with eight equivalents of N-bromosuccinimide (NBS) in a mixture of CHCl3/H2O afforded ketocalix[4]arenes 4b in relatively high yields [25,26]. Presumably, the reaction involved the radical dibromination of methylenes, followed by the hydrolysis of the brominated bridges. In the case of the larger 2b, the initial experiments showed that the reaction afforded the target 5b, but a pentamethoxy ketocalix[6]arene derivative (7, Figure 2) was also obtained as a side product, presumably formed via the cleavage of an O-Me bond by the HBr generated in the reaction. The optimization of the reaction time yielded the ketocalixarene 5b in high yield [27].
The oxidation of 3b (NBS, irradiation with a spot lamp, CHCl3/H2O) was conducted using the optimized reaction conditions found previously for 2b (Equation (1)). An examination of the crude product by NMR indicated the formation of a single major product. After recrystallization from CHCl3/EtOH, 6b was isolated in 36% yield.
Molecules 29 04094 i001
The 1H NMR spectrum of 6b (500 MHz, CDCl3, rt) displayed a symmetrical pattern, with one singlet each for the aromatic, methoxy and t-Bu protons (Supplementary Materials). The observed downfield shift in the aromatic protons (δ = 7.60 ppm) is consistent with the presence of carbonyl groups at the bridges, which results in the deshielding of these protons. A symmetrical pattern was also observed in the 13C NMR spectrum. Both NMR spectra are consistent with a molecule possessing averaged eightfold symmetry.

2.3. Crystal Structure of Ketocalix[8]arene 6b

Single crystals of 6b were grown from a mixture of CHCl3/EtOH. The molecule crystallized with CHCl3 molecules (Figure 3) in a conformation of Ci symmetry. A pair of opposite methoxy groups (C(21a)-O(4a) and C(21b)-O(4b)) were disordered in two orientations: “in” (pointing towards the cavity) and “out” at a 2:1 population ratio. In their “out” arrangement, a disordered chloroform molecule was located within the calix cavity. Assuming that the preferred conformation in solution is similar to the crystal conformation, the observed eightfold symmetry observed in the NMR spectra can be ascribed to a fast dynamic process involving rotation about the carbonyl–aryl and methoxy–aryl bonds.

2.4. Demethylation of 6b with Iodocyclohexane/DMF

An iodocyclohexane/DMF mixture was introduced by Wang, Duan and coworkers as a useful reagent for the demethylation of aryl methylethers [28]. This reagent has been successfully used in the exhaustive demethylation of calixarene methyl ether derivatives [29,30]. An attractive feature of the reagent is the slow release of HI generated by an elimination reaction. We therefore attempted the demethylation of 6b by reaction with iodocyclohexane/DMF for 24, 48 and 96 h. Unfortunately, in all cases, the 1H NMR spectrum of the crude reaction mixture displayed a complex pattern of broad signals in both the t-Bu and aromatic regions, suggesting that the demethylation did not proceed to completion (Equation (2)).
Molecules 29 04094 i002
As an alternative pathway for the preparation of 6a, we also examined the CrO3 oxidation of 3c since if the oxidation of methylenes proceeds as intended, the readily hydrolyzed acetoxy group should provide synthetic access into the octahydroxy derivative.

2.5. Oxidation of Octaacetoxy p-tert-Butyl-calix[8]arene 3c

We have previously shown that the oxidation of the 1,3-alternate form of tetracetoxycalix[4]arene 1c to the corresponding tetraoxoketocalix[4]arene can be conducted by reaction with NBS in a CHCl3/H2O mixture under irradiation with a 100 W lamp. Although HBr is generated in the reaction, the acetate groups were not cleaved during the oxidation step [31]. However, preliminary experiments indicated that the attempted oxidation of 3c with NBS under similar reaction conditions resulted only in decomposition products. We therefore resorted to Görmar’s reaction conditions (CrO3 in refluxing Ac2O/AcOH) [17]. Satisfactorily, the reaction afforded the desired octaacetoxyketocalix[8]arene 6c, albeit in low yield (12%, Scheme 3).
The 1H NMR spectrum of 6c displayed a simple spectrum with sharp singlets for the aromatic and t-Bu groups (Figure 4). As observed for 6b, the aromatic signal was downfield shifted resonating at 7.69 ppm. Notably, the acetoxy group appeared as a broad signal at 1.65 ppm. The relative upfield shift of the acetoxy groups suggest that at least some of the acetoxy groups are under the shielding effects of the aromatic rings. The increased broadening of the acetoxy signal may be indicative of a dynamic exchange between conformers resulting in an observed time averaged signal. A similar broadening of the acetoxy signal in the 1H NMR spectrum was previously observed for the smaller analog 5c [19].

2.6. Crystal Structure of Octaacetoxy Ketocalix[8]arene 6c

A single crystal of 6c was grown from chlorobenzene and submitted to X-ray crystallography. The solvent molecule that cocrystallized was disordered, and its hydrogen atoms were not located nor calculated. The molecule adopts in the crystal a conformation of Ci symmetry (Figure 5). The eight carbonyl groups can be divided into two types: “corner” and “edge”. The four “corner” carbonyls radiate from the center of the macrocycle. The acetoxy groups attached to four non-vicinal rings are oriented towards the cavity center, while the rest of the acetoxy groups are oriented “out”.

2.7. Preparation and Crystal Structure of Octahydroxyketocalix[8]arene 6a

The removal of the protecting acetyl groups of 6c was conducted by reaction with aq. NaOH in MeOH yielding 6a (Scheme 3). As with 6b and 6c, the 1H NMR spectrum of 6a (500 MHz, CDCl3, rt) showed single singlets for all groups. The OH groups resonated at low field (δ 11.01 ppm), suggesting that these groups are involved in hydrogen bonds. A single crystal of 6a was grown by the slow evaporation of an ethanolic solution and submitted to X-ray diffraction.
In contrast to 6b and 6c, 6a adopts in the crystal a saddle-like conformation of crystallographic C2 symmetry and idealized S4 symmetry. In general, the parent 3a displays a closed loop of hydrogen bonds where each OH serves both as a donor and acceptor of a hydrogen bond. A detailed analysis of the crystal structure of 3a cocrystallized with different solvents showed that although in some cases, a solvent molecule can disrupt the closed loop, most of the intramolecular hydrogen bonds between neighboring phenol rings remain unaffected [32]. Remarkably, the hydrogen bond pattern in 6a displays four pairs of convergent intramolecular hydrogen bonds involving carbonyl groups and a pair of geminal phenol rings (Figure 6). To the best of our knowledge, such a pattern is unprecedented for a calix[8]arene analog. The non-hydrogen-bonded carbonyls are located at the corners of the somewhat square-shaped geometry and are pointing away from the cavity center. An analysis of the crystal structures of substituted benzophenones in the literature has shown that in non-hydrogen-bonded derivatives, the C=O bond length is on the average 1.22 Å, whereas when hydrogen-bonded, this bond length increases to 1.24–1.25 Å [33]. In 6a, the C=O bond lengths of the two nonequivalent carbonyls at the corners are O(8)C(44):1.221(3) and O(4)C(22):1.224(3) Å, whereas for the hydrogen-bonded carbonyls, they are O(6)C(33):1.247(3) and O(2)C(11):1.246(3) Å, in agreement with the previous analysis.
Pairs of convergent hydrogen bonds involving carbonyl oxygens as acceptors were not observed in the crystal structure of the smaller ketocalix[4]arene 4a. It seems likely that the larger (and thus more flexible) macrocyclic scaffold in 6a allows for conformations where pairs of OH groups on geminal rings are in steric proximity to a central carbonyl. These conformations are sterically inaccessible for the lower homolog 4a.
The structure of 6a can be viewed as comprising four non-overlapping diarylketone subunits. Diarylketones prefer a helical “propeller” conformation where both rings are twisted in the same direction relative to the C-C(=O)C central plane [33,34,35]. The propeller conformation is a compromise between two opposite effects: (i) to maximize the conjugation of the aryl rings with the carbonyls by attaining a coplanar arrangement and (ii) to minimize the steric repulsion between ortho hydrogens (or substituents if present) on both rings. This propeller conformation is chiral; thus, two enantiomeric arrangements are possible. These two arrangements are depicted in Figure 7. An inspection of the crystal structure of 6a indicates that the two possible helicities alternate along the macrocycle, as expected for a conformation of idealized S4 symmetry (Figure 7).

2.8. Low-Temperature 1H NMR Spectra of 6a

In the crystal conformation, the unique C2 axis is perpendicular to the main macrocyclic plane and exchanges pairs of opposite rings. However, this chiral conformation may be due to a minor distortion (due to crystal packing forces) of the more symmetric achiral conformation of S4 symmetry. In this conformation, alternating rings (i.e., in “1,3,5,7” or “2,4,6,8” positions) are exchanged by the S4 axis, but pairs of geminal rings are symmetry nonequivalent. Thus, it should be expected that if the conformation is “frozen” on the NMR timescale, precluding accidental isochrony, two OH signals should be observed. Upon lowering the temperature of an NMR sample in CD2Cl2, the OH signals broadened and decoalesced into two relatively sharp signals in a 1:1 ratio and three broad signals in a 1:1:2 ratio (Figure 8). If the two sharp singlets are assigned to a conformation of S4 symmetry, it can be concluded that an additional conformation with C2 or Ci symmetry is also present in solution.

3. Experimental Methods

3.1. General Information

All common reagents and solvents were used from commercial suppliers without further purification. 1H NMR and 13C NMR spectra were recorded at room temperature on DRX400 (Bruker, Karlsruhe, Germany) and AVANCE II 500 (Bruker, Karlsruhe, Germany) instruments. Proton chemical shifts (δ in ppm) were referenced to the residual protium resonance of the solvent (CDCl3, δ = 7.26 ppm) or to TMS. Carbon chemical shifts (δ in ppm) were referenced to the carbon resonances of the solvent (CDCl3, δ = 77.2 ppm). High-resolution mass spectra (HRMS) were recorded at the Analytical Chemistry Lab (The Hebrew University) using an X500R QTOF (SCIEX, Framingham, MA, USA) instrument. Melting points were measured using a Fisher-Johns (Fischer Scientific Co., Hampton, VA, USA) melting point apparatus and were uncorrected.

3.2. Crystal Data

CCDC (Deposition Number 2363603-2363605) contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/? (accessed on 28 August 2024) (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44 1223 336033; E-mail: [email protected]).
  • Crystal data for 6a: empirical formula: C98.5H108O21.25, formula weight: 1631.84, temperature: 150.0(1) K, crystal system: monoclinic, space group: I2/a, a: 18.4709(5) Å, b: 22.0612(6) Å, c: 23.7204(7) Å, β: 90.484(3)°, V: 9665.5(5) Å3, Z: 4, ρcalc: 1.121 g/cm3, μ: 0.078 mm−1, F(000): 3476.0, crystal size: 0.24 × 0.09 × 0.08 mm3, radiation: Mo Kα (λ = 0.71073), 2θ range for data collection: 4.41 to 64.572°, index ranges: −23 ≤ h ≤ 27, −32 ≤ k ≤ 28, −33 ≤ l ≤ 35, reflections collected: 44,192, independent reflections: 14328 [Rint =0.0435, Rsigma =0.0584], data/restraints/parameters: 14,328/0/606, goodness-of-fit on F2: 1.028, final R indexes [I ≥ 2σ (I)]: R1: 0.0916, wR2: 0. 0.2518, final R indexes [all data]: R1: 0.1404, wR2: 0.2827, largest diff. peak/hole/e Å−3 0.80/−0.34.
  • Crystal data for 6b: empirical formula: C97.67H113.33Cl5O16, formula weight: 1720.50, temperature: 200.0(1) K, crystal system: triclinic, space group: P-1, a: 12.4386(1)Å, b: 13.4767(2) Å, c: 14.6784(1) Å, α: 97.887(1)°, β: 93.365(1)°, γ: 96.941(1)°, V: 2412.32(4) Å3, Z: 1, ρcalc: 1.184 g/cm3, μ: 0.212 mm−1, F(000): 912.0, crystal size: 0.43 × 0.11 × 0.07 mm3, radiation: Mo Kα (λ = 0.71073), 2θ range for data collection: 4.174 to 64.604°, index ranges: −17 ≤ h ≤ 18, −19 ≤ k ≤ 19, −21 ≤ l ≤ 21, reflections collected: 104,881, independent reflections: 15145 [Rint = 0.0309, Rsigma = 0.0221], data/restraints/parameters: 15,145/0/631, goodness-of-fit on F2: 1.031, final R indexes [I ≥ 2σ (I)]: R1: 0.0695, wR2: 0.2071, final R indexes [all data]: R1: 0.0871, wR2: 0.2205, largest diff. peak/hole/e Å−3 0.84/−0.66.
  • Crystal data for 6c: empirical formula: C116H112Cl2O24, formula weight: 1960.95, temperature: 150.0(1) K, crystal system: triclinic, space group: P-1, a: 10.0905(1) Å, b: 15.8706(2) Å, c: 17.1177(2) Å, α: 84.812(1)°, β: 76.830(1)°, γ: 82.254(1)°, V: 2639.74(5) Å3, Z: 1, ρcalc: 1.234 g/cm3, μ: 0.134 mm−1, F(000): 1034.0, crystal size: 0.26 × 0.19 × 0.09 mm3, radiation: Mo Kα (λ = 0.71073), 2θ range for data collection: 4.174 to 64.798°, index ranges: −14 ≤ h ≤ 14, −22 ≤ k ≤ 23, −23 ≤ l ≤ 24, reflections collected: 94,717, independent reflections: 16,346 [Rint = 0.0412, Rsigma = 0.0312], data/restraints/parameters: 16,346/0/778, goodness-of-fit on F2: 1.027, final R indexes [I ≥ 2σ (I)]: R1: 0.0737, wR2: 0.1987, final R indexes [all data]: R1: 0.0918, wR2: 0.2101, largest diff. peak/hole/e Å−3 0.57/−0.68.

3.3. Synthesis of Compounds 6a6c

3.3.1. Preparation of 5,11,17,23,29,35,41,47-Octa-tert-butyl-49,50,51,52,53,54,55,56-octahydroxy-2,8,14,20,26,32,38,44-octaoxo-calix[8]arene (6a)

A mixture of 6c (0.5 g, 0.29 mmol), aq. NaOH (75 mL, 2 M) and methanol (50 mL) was heated to reflux for 4 h. After cooling to rt, the mixture was acidified with conc. HCl until pH = 1, and the solid was filtered by vacuum yielding 0.48 g of a yellow solid. Trituration with cold methanol afforded 0.38 g 6a (94%), mp 305–307 °C.
IR ν = 3527 (OH), 1669 (C=O) cm−1. 1H NMR (CDCl3, 500 MHz) δ 11.01 (s, 8H, OH), 7.86 (s, 16H, ArH), 1.35 (s, 72H, t-Bu) ppm. 13C NMR (CDCl3, 125 MHz) δ 198.4, 158.9, 141.6, 133.1, 124.8, 34.5, 31.3 ppm. HRMS (ESI-QTOF) m/z calcd. for C88H96O16 + H+: 1409.6771. Found: 1409.6730.

3.3.2. Preparation of 5,11,17,23,29,35,41,47-Octa-tert-butyl-49,50,51,52,53,54,55,56-octamethoxy-2,8,14,20,26,32,38,44-octaoxocalix[8]arene (6b)

A mixture of octamethoxycalix[8]arene 3b (4.7 g, 4.36 mmol), NBS (19.7 g, 110.0 mmol), chloroform (90 mL) and water (2 mL) was refluxed during 3.5 h while irradiated with a spotlight (100 W). After cooling, the solution was washed successively with aq Na2S2O5, water, aq NaHCO3, water, dried over MgSO4 and evaporated. A 1H NMR analysis of the crude product indicated a nearly complete oxidation of the methylene groups. The crude product was recrystallized from CHCl3/EtOH, yielding the product (1.82 g, 36% yield), mp 312 °C.
1H NMR (CDCl3, 400 MHz) δ 7.60 (s, 16H, ArH), 3.23 (s, 24H, OMe), 1.25 (s, 72H, t-Bu) ppm. 13C NMR (CDCl3, 100 MHz) δ 195.3, 155.9, 145.9, 133.3, 130.6, 63.1, 34.5, 31.1 ppm. HRMS (ESI-QTOF) m/z calcd. for C96H112O16 + H+: 1521.8023. Found: 1521.7952.

3.3.3. Preparation of 5,11,17,23,29,35,41,47-Octa-tert-butyl-49,50,51,52,53,54,55,56-octaacetoxy-2,8,14,20,26,32,38,44-octaoxo-calix[8]arene (6c)

To a mixture of octaacetoxycalix[8]arene 3c (3.03 g, 1.85 mmol), 200 mL acetic anhydride and 4 mL acetic acid was added CrO3 (6.5 g, 0.065 mmol). The mixture was heated to reflux for 3h. During the heating period, the color of the reaction changed from dark orange to green. After cooling to rt, the mixture was filtered through a sintered glass filter, 200 mL of chloroform was added and the mixture was washed with water (200 mL) several times until the water phase was no longer colored. The organic phase was dried (MgSO4) and evaporated. The residue was recrystallized from CHCl3/MeOH, yielding 0.4 g (12%) 6c, mp 310–312 °C.
1H NMR (CDCl3, 500 MHz) δ 7.69 (s, 16H, ArH), 1.65 (br s, 24H, CH3C=O), 1.28 (s, 72H, t-Bu) ppm. 13C NMR (CDCl3, 125 MHz) δ 192.3, 168.5, 149.1, 144.9, 132.4, 130.8, 34.9, 31.0, 20.0 ppm. Anal. calcd. for C104H112O24: C, 71.54, H, 6.47. Found: C, 71.54, H, 6.46. HRMS (ESI-QTOF) m/z calcd. for C104H112O24 + H+: 1745.7616. Found: 1745.7603.

4. Conclusions

Octamethoxy and octaacetoxyketocalix[8]arenes can be prepared by the oxidation of the corresponding calix[8]arenes with NBS/light and CrO3, respectively. An X-ray analysis of a crystal of octahydroxyketocalix[8]arene 6a grown from ethanol indicates that the molecule adopts a conformation of idealized S4 symmetry where the helicities of the dihydroxybenzophenone subunits alternate along the macrocycle. The carbonyl groups disrupt all the intramolecular hydrogen bonds between phenol rings.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules29174094/s1, Figure S1: 1H NMR spectrum (500 MHz, CDCl3, rt) of 6a, Figure S2: 13C NMR spectrum (125 MHz, CDCl3, rt) of 6a, Figure S3: HRMS of 6a, Figure S4: 1H NMR spectrum (400 MHz, CDCl3, rt) of 6b, Figure S5: 13C NMR spectrum (100 MHz, CDCl3, rt) of 6b, Figure S6: HRMS of 6b, Figure S7: 1H NMR spectrum (500 MHz, CDCl3, rt) of 6c, Figure S8: 13C NMR spectrum (125 MHz, CDCl3, rt) of 6c, Figure S9: HRMS of 6c.

Author Contributions

Conceptualization, K.K., S.O. and S.E.B.; methodology, K.K., S.O. and S.E.B.; investigation, K.K. and S.O.; crystal structure determinations, B.B.; writing—original draft preparation, S.E.B.; writing—review and editing, S.E.B.; supervision, S.E.B.; project administration, S.E.B.; funding acquisition, S.E.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Israel Science Foundation (ISF) Grant No. 262/20.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in this article/Supplementary Material; further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Böhmer, V. Calixarenes, Macrocycles with (Almost) Unlimited Possibilities. Angew. Chem. Int. Ed. Engl. 1995, 34, 713–745. [Google Scholar] [CrossRef]
  2. Gutsche, C.D. Calixarenes Revisited; Royal Society of Chemistry: Cambridge, UK, 1998. [Google Scholar] [CrossRef]
  3. Asfari, Z.; Böhmer, V.; Harrowfield, J.; Vicens, J.; Saadioui, M. (Eds.) Calixarenes 2001; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2001. [Google Scholar] [CrossRef]
  4. Böhmer, V. The Chemistry of Phenols; Rappoport, Z., Ed.; Wiley: Chichester, UK, 2003; pp. 1369–1454. [Google Scholar] [CrossRef]
  5. Gutsche, C.D. Calixarenes: An Introduction, 2nd ed.; Royal Society of Chemistry: Cambridge, UK, 2008. [Google Scholar] [CrossRef]
  6. Neri, P.; Sessler, J.L.; Wang, M.-X. (Eds.) Calixarenes and Beyond; Springer: Cham, Switzerland, 2016. [Google Scholar] [CrossRef]
  7. Gutsche, C.D.; Iqbal, M. p-tert-butylcalix[4]arene. Org. Synth. 1990, 68, 234–237. [Google Scholar] [CrossRef]
  8. Gutsche, C.D.; Dhawan, B.; Leonis, M.; Stewart, D. p-tert-butylcalix[6]arene . Org. Synth. 1990, 68, 238–242. [Google Scholar] [CrossRef]
  9. Munch, J.H.; Gutsche, C.D. p-tert-butylcalix[8]arene. Org. Synth. 1990, 68, 243–244. [Google Scholar] [CrossRef]
  10. For the preparation of “giant” calix[n]arenes derivatives containing up to 90 rings see. Guerineau, V.; Rollet, M.; Viel, S.; Lepoittevin, B.; Costa, L.; Saint-Aguet, P.; Laurent, R.; Roger, P.; Gigmes, D.; Martini, C.; et al. The synthesis and characterization of giant Calixarenes. Nat. Commun. 2019; 10, 113. [Google Scholar] [CrossRef]
  11. Stewart, D.R.; Gutsche, C.D. Isolation, Characterization, and Conformational Characteristics of p-tert-Butylcalix[9-20]arenes. J. Am. Chem. Soc. 1999, 121, 4136–4146. [Google Scholar] [CrossRef]
  12. For a recent review on the functionalization at the meta positions of the aromatic rings see. Lhotak, P. Direct meta substitution of calix[4]arenes. Org. Biomol. Chem. 2022; 20, 7377–7390. [Google Scholar] [CrossRef]
  13. For a review on calixarenes modified at the bridges see. Sliwa, W.; Deska, M. Calixarenes containing modified meso bridges. Arkivoc, 2015; i, 29–47. [Google Scholar] [CrossRef]
  14. For a review on the intramolecular hydrogen bonding in calixarenes see. Rudkevich, D.M. Intramolecular Hydrogen Bonding in Calixarenes. Chem. Eur. J. 2000; 6, 2679–2686. [Google Scholar] [CrossRef]
  15. Seri, N.; Simaan, S.; Botoshansky, M.; Kaftory, M.; Biali, S.E. A Conformationally Flexible Tetrahydroxycalix[4]arene Adopting the Unusual 1,3-Alternate Conformation. J. Org. Chem. 2003, 68, 7140–7142. [Google Scholar] [CrossRef]
  16. A 1,3-alternate conformation was also observed in a tetrahydroxy tetrasulfinylcalix[4]arene. Mislin, G.; Graf, E.; Hosseini, M.W.; De Cian, A.; Fischer, J. Tetrasulfinylcalix[4]arenes: Synthesis and solid state structural analysis. Tetrahedron Lett. 1999; 40, 1129–1132. [Google Scholar] [CrossRef]
  17. Görmar, G.; Seiffarth, K.; Schulz, M.; Zimmermann, J.; Flämig, G. Synthese und Reduktion des Tetra-tert-butyltetraoxocalix[4]arens. Makromol. Chem. 1990, 191, 81–87. [Google Scholar] [CrossRef]
  18. Seri, N.; Thondorf, I.; Biali, S.E. Preparation and Conformation of Dioxocalix[4]arene Derivatives. J. Org. Chem. 2004, 69, 4774–4780. [Google Scholar] [CrossRef]
  19. Kogan, K.; Biali, S.E. Intramolecular SNAr Reactions in a Large-Ring Ketocalix[6]arene. Org. Lett. 2007, 9, 2393–2396. [Google Scholar] [CrossRef]
  20. Ninagawa, A.; Cho, K.; Matsuda, H. Preparation of 4-tert-butyloxocalix[n]arenes and their properties as UV-absorbers. Makromol. Chem. 1985, 186, 1379–1385. [Google Scholar] [CrossRef]
  21. Reddy, P.A.; Kashyap, R.P.; Watson, W.H.; Gutsche, C.D. Calixarenes 30. Calixquinones. Isr. J. Chem. 1992, 32, 89–96. [Google Scholar] [CrossRef]
  22. Neri, P.; Bottino, A.; Cunsolo, F.; Piattelli, M.; Gavuzzo, E. 5,5′-Bicalix[4]arene: The Bridgeless Prototype of Double Calix[4]arenes of the “Head-to-Head” Type. Angew. Chem. Int. Ed. 1998, 37, 166–169. [Google Scholar] [CrossRef]
  23. Kuno, L.; Seri, N.; Biali, S.E. Tetraaddition of PhLi to a Ketocalixarene Derivative. Org. Lett. 2007, 9, 1577–1580. [Google Scholar] [CrossRef]
  24. Poms, D.; Itzhak, N.; Kuno, L.; Biali, S.E. Calixradialenes: Calixarene Derivatives with Exocyclic Double Bonds. J. Org. Chem. 2014, 79, 538–545. [Google Scholar] [CrossRef] [PubMed]
  25. Fischer, C.; Lin, G.; Seichter, W.; Weber, E. Simple two step reaction towards a 2,7,14,21-tetraoxocalix[4]arene via bridge brominated intermediate. Tetrahedron Lett. 2013, 54, 2187–2189. [Google Scholar] [CrossRef]
  26. Saikia, I.; Borah, A.J.; Phukan, P. Use of Bromine and Bromo-Organic Compounds in Organic Synthesis. Chem. Rev. 2016, 116, 6837–7042. [Google Scholar] [CrossRef]
  27. Itzhak, N.; Biali, S.E. An Expedient Synthesis of Ketocalix[6]arene Hexamethyl Ether. Synthesis 2018, 50, 3897–3901. [Google Scholar] [CrossRef]
  28. Zuo, L.; Yao, S.; Wang, W.; Duan, W. An efficient method for demethylation of aryl methyl ethers. Tetrahedron Lett. 2008, 49, 4054–4056. [Google Scholar] [CrossRef]
  29. Carroll, L.T.; Hill, P.A.; Ngo, C.Q.; Klatt, K.P.; Fantini, J.L. Synthesis and reactions of a 2-chlorocalix[4]arene and a 2,2′-coupled dicalixarene. Tetrahedron 2013, 69, 5002–5007. [Google Scholar] [CrossRef]
  30. Fong, A.; McCormick, L.; Teat, S.J.; Brechin, E.K.; Dalgarno, S.J. Exploratory studies into 3d/4f cluster formation with fully bridge-substituted calix[4]arenes. Supramol. Chem. 2018, 30, 504–509. [Google Scholar] [CrossRef]
  31. Itzhak, N.; Biali, S.E. Preparation and Crystal Structure of the 1,3-alternate Atropisomer of a Calix[4]radialene. Eur. J. Org. Chem. 2015, 2015, 3221–3225. [Google Scholar] [CrossRef]
  32. For a recent study see. Kieliszek, A.; Malinska, M. Conformations of p-tert-Butylcalix[8]arene in Solvated Crystal Structures. Cryst. Growth. Des. 2021; 21, 6862–6871. [Google Scholar] [CrossRef]
  33. Rappoport, Z.; Biali, S.E.; Kaftory, M. Application of the structural correlation method to ring-flip processes in benzophenones. J. Am. Chem. Soc. 1990, 112, 7742–7748. [Google Scholar] [CrossRef]
  34. Abraham, R.J.; Haworth, I.S. Conformational analysis. Part 12. A theoretical and lanthanide-induced shift (LIS) investigation of the conformation of benzophenone and 3,4′-dichlorobenzophenone. J. Chem. Soc. Perkin Trans. 1988, 8, 1429–1437. [Google Scholar] [CrossRef]
  35. Grilli, S.; Lunazzi, L.; Mazzanti, A.; Casarini, D.; Femoni, C. Conformational Studies by Dynamic NMR. 78. Stereomutation of the Helical Enantiomers of Trigonal Carbon Diaryl-Substituted Compounds:  Dimesitylketone, Dimesitylthioketone, and Dimesitylethylene. J. Org. Chem. 2001, 66, 488–495. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. “Major” calixarenes 1a3a and their protected derivatives.
Scheme 1. “Major” calixarenes 1a3a and their protected derivatives.
Molecules 29 04094 sch001
Figure 1. Tetrahydroxy-p-tert-butyl-calix[4]arene 1a (left) adopts a cone conformation, while tetrahydroxyketocalix[4]arene 4a (right) adopts a 1,3-alternate conformation.
Figure 1. Tetrahydroxy-p-tert-butyl-calix[4]arene 1a (left) adopts a cone conformation, while tetrahydroxyketocalix[4]arene 4a (right) adopts a 1,3-alternate conformation.
Molecules 29 04094 g001
Scheme 2. Ketocalixarene derivatives of “major” calixarenes.
Scheme 2. Ketocalixarene derivatives of “major” calixarenes.
Molecules 29 04094 sch002
Figure 2. Side product 7 obtained in the oxidation of 2b.
Figure 2. Side product 7 obtained in the oxidation of 2b.
Molecules 29 04094 g002
Figure 3. The crystal structure of octamethoxyketocalix[8]arene 6b. Chloroform molecules that cocrystallized are omitted. A pair of opposite methoxy groups and t-Bu groups are disordered. For clarity, only one orientation of the disordered groups is shown.
Figure 3. The crystal structure of octamethoxyketocalix[8]arene 6b. Chloroform molecules that cocrystallized are omitted. A pair of opposite methoxy groups and t-Bu groups are disordered. For clarity, only one orientation of the disordered groups is shown.
Molecules 29 04094 g003
Scheme 3. Preparation of octaacetoxyketocalix[8]arene 6c and octahydroxyketocalix[8]arene 6a.
Scheme 3. Preparation of octaacetoxyketocalix[8]arene 6c and octahydroxyketocalix[8]arene 6a.
Molecules 29 04094 sch003
Figure 4. 1H NMR spectrum (CDCl3, rt) of octaacetoxy ketocalix[8]arene 6c.
Figure 4. 1H NMR spectrum (CDCl3, rt) of octaacetoxy ketocalix[8]arene 6c.
Molecules 29 04094 g004
Figure 5. Crystal structure of octaacetoxyketocalix[8]arene 6c. Disordered t-Bu groups and chlorobenzene molecule that cocrystallized are omitted for clarity.
Figure 5. Crystal structure of octaacetoxyketocalix[8]arene 6c. Disordered t-Bu groups and chlorobenzene molecule that cocrystallized are omitted for clarity.
Molecules 29 04094 g005
Figure 6. An ORTEP (left) and CPK representation (right) of the crystal structure of octahydroxy ketocalix[8]arene 6a. Disordered t-Bu and ethanol molecules that cocrystallized are omitted. The molecule adopts a saddle-like conformation of idealized S4 symmetry. Four pairs or congruent hydrogen bonds are present, involving OH and carbonyl groups. Selected O…O distances (Å): O(1)O(2): 2.611(3), O(2)O(3): 2.584(3), O(5)O(6): 2.548(2), O(6)O(7): 2.696(2).
Figure 6. An ORTEP (left) and CPK representation (right) of the crystal structure of octahydroxy ketocalix[8]arene 6a. Disordered t-Bu and ethanol molecules that cocrystallized are omitted. The molecule adopts a saddle-like conformation of idealized S4 symmetry. Four pairs or congruent hydrogen bonds are present, involving OH and carbonyl groups. Selected O…O distances (Å): O(1)O(2): 2.611(3), O(2)O(3): 2.584(3), O(5)O(6): 2.548(2), O(6)O(7): 2.696(2).
Molecules 29 04094 g006
Figure 7. (left): The two possible helicities of a 2,2-dihydroxybenzophenone subunit possessing convergent hydrogen bonds. The two helicities are colored red and blue. (right): The helicities of the intramolecularly hydrogen bonded diarylketone subunits according to the color convention depicted on the left. The helicities alternate along the macrocycle resulting in a structure of idealized S4 symmetry.
Figure 7. (left): The two possible helicities of a 2,2-dihydroxybenzophenone subunit possessing convergent hydrogen bonds. The two helicities are colored red and blue. (right): The helicities of the intramolecularly hydrogen bonded diarylketone subunits according to the color convention depicted on the left. The helicities alternate along the macrocycle resulting in a structure of idealized S4 symmetry.
Molecules 29 04094 g007
Figure 8. Temperature-dependent 1H NMR spectrum (400 MHz, CD2Cl2) of OH region of octahydroxy ketocalix[8]arene 6a. From top to bottom: at 288, 217 and 188 K.
Figure 8. Temperature-dependent 1H NMR spectrum (400 MHz, CD2Cl2) of OH region of octahydroxy ketocalix[8]arene 6a. From top to bottom: at 288, 217 and 188 K.
Molecules 29 04094 g008
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

Kogan, K.; Omar, S.; Bogoslavsky, B.; Biali, S.E. The Preparation and Crystal Structures of Octaoxoketocalix[8]arene Derivatives: The Ketocalixarene Counterparts of the Largest “Major” Calixarene. Molecules 2024, 29, 4094. https://doi.org/10.3390/molecules29174094

AMA Style

Kogan K, Omar S, Bogoslavsky B, Biali SE. The Preparation and Crystal Structures of Octaoxoketocalix[8]arene Derivatives: The Ketocalixarene Counterparts of the Largest “Major” Calixarene. Molecules. 2024; 29(17):4094. https://doi.org/10.3390/molecules29174094

Chicago/Turabian Style

Kogan, Katerina, Suheir Omar, Benny Bogoslavsky, and Silvio E. Biali. 2024. "The Preparation and Crystal Structures of Octaoxoketocalix[8]arene Derivatives: The Ketocalixarene Counterparts of the Largest “Major” Calixarene" Molecules 29, no. 17: 4094. https://doi.org/10.3390/molecules29174094

Article Metrics

Back to TopTop