Next Article in Journal
(2-[4-Methylpyrazol-1-yl]phenyl)platinum(II) (1,3-bis[1-methyl-1H-pyrazol-4-yl]propane-1,3-dione)
Previous Article in Journal
5-(Benzoyloxymethyl)isoxazole-3-carboxylic Acid Ethyl Ester
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

1-(2-(3,5-Di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) Quinolin-1-ium Bromide

1
Institute of Mineralogy and Crystallography “Acad. Ivan Kostov”, Bulgarian Academy of Sciences, Acad. G. Bonchev Street, bl. 107, 1113 Sofia, Bulgaria
2
Institute of Organic Chemistry with Centre of Phytochemistry, Bulgarian Academy of Sciences, Acad. G. Bonchev Street, bl. 9, 1113 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Molbank 2024, 2024(1), M1763; https://doi.org/10.3390/M1763
Submission received: 5 January 2024 / Revised: 19 January 2024 / Accepted: 22 January 2024 / Published: 24 January 2024

Abstract

:
The title compound 1-(2-(3,5-di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) quinolin-1-ium bromide was obtained in good yield by a facile one-pot, one-step synthetic procedure involving quinoline and an aromatic α-bromo ketone. The product was isolated using hot recrystallization from acetone/acetonitrile solution and characterized using powder and single-crystal XRD, 1D and 2D NMR, DSC, FT-IR, and HRMS analyses.

1. Introduction

The title compound 1-(2-(3,5-di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) quinolin-1-ium bromide is a quaternary ammonium salt obtained with a simple one-step quaternization reaction between quinoline and an aromatic α-bromo ketone. In chemistry, the conversion of tertiary aromatic/aliphatic amine to quaternary ammonium compound by reaction with alkyl/aryl halide is referred to as the Menshutkin reaction [1]. The reaction can be described as a second-order nucleophilic substitution (SN2), whereas both reagents are involved in the rate-determining step. Beneficial factors affecting the reaction rate are the use of polar aprotic solvents as reaction media, elevated reaction temperatures, and alkylating agents possessing a labile leaving group [2,3,4].
The molecular characteristics of the novel compound feature quinoline and 2,6-di-tert-butylphenol moieties connected by a 2-oxoethyl bridge. The quinoline is a versatile aromatic heterocyclic compound valued as an attractive molecular scaffold for a variety of scientific and industrial applications [5,6]. The unique structure of quinoline contributes to its interaction with biological systems, leading to its use in medicinal chemistry [7,8]. Quinoline and its derivatives have been employed in the synthesis of drugs with antimalarial [9,10], antitubercular [11,12], antitumor [13,14], antibacterial, and antifungal activities [15,16]. The nitrogen atom in the quinoline ring can serve as a coordination center for metal ions, a property exploited in many aspects of coordination chemistry [17,18]. Furthermore, quinoline derivatives often exhibit fluorescence, making them a valued building block for the development of fluorescent dyes and biomolecular labels for applications in imaging and clinical diagnostics [19,20,21]. The electronic properties of quinoline (the ability to donate and accept electrons) are fundamental in the design of materials for electronic devices, including OLEDs and other optoelectronic applications [22,23]. In recent years, great scientific attention has been given to quinoline derivatives acting as corrosion inhibitors [24,25]. Examples of quinoline derivatives found in the literature are given in Figure 1.
On the other hand, the research on 2,6-tert-butyl phenol (2,6-DTBP) derivatives is limited and is mainly focused on its antioxidative properties. For example, in a study by Milaeva et al., the 2,6-DTBP fragment was included in a series of organotin complexes showing antioxidative, luminescent, and cytotoxic properties [26]. Moreover, Berberova et al. demonstrated the antioxidative and cytotoxic activity of aromatic oligosulfides bearing the same fragment [27]. Primerova et al. reported on a series of pyrimidine, benzotriazole, and cyclic amide derivatives containing 2,6-DTBP that showed excellent redox properties and antibacterial activity against S. aureus [28]. The antioxidative potential of the 2,6-DTBP fragment was further explored by combining it with 1,2,4-Triazole-5-thione [29], 1,3,4-Thiadiazole [30], 1,3,4-oxadiazole [31], and ferrocene [32] moieties. Examples of 2,6-DTBP derivatives found in the literature are given in Figure 2.
A search of the scientific literature returned only a few molecular structures containing both quinoline and 2,6-DTBP fragments [33,34]. However, none of them possessed a positive nitrogen atom in their structure (i.e., quaternary ammonium compound). This provides an opportunity to combine the quinoline and 2,6-DTBP fragments into a quaternary ammonium molecule with the goal of unlocking a synergistic effect toward enhanced biological properties. Herein, we report on the facile one-step synthesis of the novel—1-(2-(3,5-di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) quinolin-1-ium bromide obtained in good yield (75%). The compound was characterized using 1D and 2D NMR techniques, powder and single-crystal X-ray diffraction analysis (PXRD and SCXRD), differential scanning calorimetry (DSC), Fourier-transform infrared spectroscopy (FT-IR) and high-resolution mass spectrometry (HR-MS).

2. Results and Discussion

2.1. Synthesis and Characterisation

The title compound was obtained by reacting the aromatic N-heterocyclic amine—quinoline with the aromatic α-bromoketone—2-bromo-1-(3,5-di-tert-butyl-4-hydroxyphenyl) ethan-1-one with slight excess (10%) of the latter. The synthesis was performed in the polar aprotic acetonitrile at 80 °C for 10 h with stirring. Furthermore, the reaction was conducted in an argon atmosphere due to the possibility of hydrolysis of the aromatic α-bromoketone. The reaction ended with the formation of a colorless precipitate. The precipitate was filtered and washed with cold acetonitrile. The product was isolated in 75% yield after recrystallization from a hot acetone/acetonitrile solution. The general procedure for synthesis of the title compound is given in Scheme 1.
The molecular structure of the studied compound in solution was confirmed with 1D and 2D NMR spectroscopy (Figures S1–S6). The 1H spectrum in DMSO-d6 and the numbering scheme used for the assignments are presented in Figure 3. Some of the protons were assigned unambiguously from the 1H NMR spectra by comparing the chemical shift value and the integral of the signals. Those protons are the singlets (s) for the CH3 groups of the tert-butyl fragment (1.457 ppm, 18H) and the hydroxyl group (8.240 ppm, 1H). Other signals required the observation of the coupling between the proton atoms positioned next to each other (2D COSY) or through space (2D NOESY). A close inspection of 1H NMR spectra revealed that all characteristic signals for quinoline protons are shifted towards a weak magnetic field (from 8.0 to 9.5 ppm). The signals for CH-8 (8.419 ppm) and CH-5 (8.542 ppm) are present as a doublet of doublets (dd), while for CH-7 (8.061 ppm), CH-6 (8.309 ppm), and CH-3 (8.217 ppm), signals are present as a doublet of doublets of doublets (ddd). The signals for the last two protons from the quinoline fragment—CH-4 and CH-2—overlap into a multiplet (9.403–9.473 ppm). The protons for the methylene group of the 2-oxoethyl bridge and protons for CH-2′ and CH-6′ give singlets (s) shifted at 7.007 ppm and 7.907 ppm, respectively. The information about the chemical shift of the carbon atoms in the 13C NMR spectra was extracted from HSQC and HMBC spectra. The FT-IR spectrum of the title compound (Figure S7) shows a characteristic C = O stretching for conjugated ketone at 1674 cm−1. The HR-MS (ESI+) experiment detected a m/z peak at 78.9176 corresponding to a Br and a m/z peak of 376.2271 corresponding to the quaternary ammonium product (Figure S8).
The thermal stability of the reported compound was studied with DSC analysis. The DSC thermogram (Figure 4) reveals a broad endothermic effect spanning the 48 °C–145 °C temperature region (H = 54.24 J/g), attributed to the prolonged release of both physisorbed acetonitrile and the one present in the crystal structure (see Section 2.2.). The acetonitrile-free structure is stable up to 170 °C when an exothermic effect starts to develop. This exothermic effect ending at ~200 °C (H = 19.37 J/g) can be associated with the monotropic solid–solid transition often observed in organic compounds [35,36]. The peak temperature of this exothermic effect at 188 °C could be defined as the transition point between the low-temperature, metastable and the high-temperature, stable form of the compound. Shortly after its formation, the high-temperature form melts at 223 ± 1 °C with a sharp and intensive endothermic effect (H = 75.38 J/g).

2.2. Crystallography

The title compound appears as colorless block-shaped crystals obtained by slow evaporation from a hot acetone/acetonitrile solution. The compound crystallizes in the triclinic P-1 space group (a = 12.402 Å, b = 12.458 Å, c = 17.978 Å, α = 90.79 °, β = 107.90°, and γ = 91.96°) with four molecules in the unit cell (Z = 4) and two independent molecules in the asymmetric unit (ASU, Z’ = 2, Table S1). Furthermore, the title compound is a crystal solvate having an acetonitrile molecule trapped in the crystal structure (Figure 5). At the molecular level, the studied compound can be described as constructed by quinoline and 2,6-di-tert-butylphenol fragments connected by a 2-oxoethyl (–CH2–C(=O)–) bridge. The quinoline moiety is planar with calculated RSMD values of 0.006 Å or 0.013 Å for both independent molecules. If only the phenol part of the 2,6-di-tert-butylphenol fragment is considered, the RMSD values are 0.001 Å and 0.009 Å, otherwise the RMSD values increase to 0.648 Å and 0.651 Å. Furthermore, the overall conformational flexibility of the molecule is enhanced by the presence of the 2-oxoethyl bridge. Thus, the quinoline and 2,6-di-tert-butylphenol fragments are distorted with respect to each other with a plane-to-plane twist angle of 87.84°/83.55° and plane-to-plane fold angle of 135.05°/137.45 for each independent molecule.
The molecule of the title compound possesses one carbonyl group (C = O) that is a typical proton acceptor and one OH group that can be both proton acceptor and donor. However, instead of the anticipated O–H…O = C interaction, the crystal structure is stabilized by two halogen bonding interactions (O–H…Br) with different D–H…X distances (Figure 6a, Table S4). For the molecule colored in orange (Figure 6a), the O-H…Br interaction is stronger (D-H…X = 2.532 Å) than for the one colored in blue (D-H…X = 2.631 Å).
The powder XRD pattern of the bulk of the studied compound was compared with the one generated from the single-crystal XRD (Figure 7). The comparison discloses full overlapping of the positions of the diffraction peaks, confirming the monophasicity of the sample.

3. Materials and Methods

3.1. General

All reagents were purchased from Sigma Aldrich and Alfa Aesar and were used without additional purification. PXRD analysis was performed on an Empyrean (Malvern Panalytical, Almelo, The Netherlands) diffractometer equipped with a PIXcel3D area detector and Cu X-ray source (λ = 1.5406 Å). Diffracting patterns were collected in the 2–50° 2Theta range using a low-background silicon single-crystal sample holder at operating conditions of 40 kV/30 mA and a step size of 0.013°. The NMR spectra were recorded on a Bruker Avance II+ 600 spectrometer (Rheinstetten, Germany) in DMSO-d6 and processed using the Topspin v2.1 program. Chemical shifts are given in ppm in δ-values against the solvent peak at 2.5 ppm, and the spin–spin coupling constants were calculated in Hz. Signals were assigned using a combination of two-dimensional NOESY, HSQC, and HMBC techniques. DSC analysis was performed on a Discovery DSC250 (TA Instruments, New Castle, DE, USA). A sample weighing around 2 mg was heated in a closed aluminum pan from 30 to 230 °C (heating rate 10 °C min –1) in nitrogen (flow rate 20 mL min –1). The reported melting point was determined from the DSC thermogram. The FT-IR spectrum was collected on a Tensor 37 (Bruker, Berlin, Germany) spectrometer using KBr pellets. The mass spectra were recorded using a Q Exactive Plus Hybrid Quadrupole-Orbitrap Mass Spectrometer from Thermo Scientific (ESI HRMS) in both positive and negative modes. The spectra were processed using the Xcalibur FreeStyle 1.8 SP1 program.

3.2. Synthesis of 1-(2-(3,5-Di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) Quinolin-1-ium Bromide

A solution of quinoline (2 mmol) and 2-bromo-1-(3,5-di-tert-butyl-4-hydroxyphenyl) ethan-1-one (2.2 mmol) in 10 mL acetonitrile was stirred for 10 h at 80 °C in an argon atmosphere. The precipitate was filtered and washed with acetonitrile and recrystallized from hot acetone/acetonitrile. Yield: 75%, colorless block-shaped crystals (acetone/acetonitrile); m.p. 222–224 °C; 1H NMR (DMSO-d6, 600 MHz,) δ 1.457 (s, 18H, CH3-tert-butyl), 7.007 (s, 2H, CH2-C = O), 7.907 (s, 2H, CH-2′ and CH-6′), 8.061 (ddd, 1H, J = 8.1 Hz, 6.9 Hz, 0.8 Hz, CH-6), 8.218 (ddd, 1H, J = 8.8 Hz, 7.0 Hz, 1.5 Hz, CH-7), 8.240 (s, 1H, OH) 8.309 (dd, 1H, J = 8.4 Hz, 5.8 Hz, CH-3), 8.419 (dd, 1H, J = 9.1 Hz, 0.9 Hz, CH-8), 8.540 (dd, 1H, J = 8.2 Hz, 1.5 Hz, CH-5), 9.40–9.47 (m, 2H, CH-4 and CH-2); 13C NMR (DMSO-d6, 151 MHz) δ 29.91 (C-tert-butyl), 34.64 (Cq-7′ and Cq-8′), 63.04 (CH2-C = O), 119.39 (C-8), 122.18 (C-3), 124.95 (Cq-1′), 126.03 (C-2′ and C-6′), 129.39 (Cq-4a), 129.98 (C-6), 130.57 (C-5), 135.83 (C-7), 138.43 (Cq-3′ and Cq-5′), 138.80 (Cq-8a), 148.41 (C-4), 150.80 (C-2), 160.28 (Cq-4′), 189.33 (C = O); FT-IR 3612, 3418, 3220, 3086, 2963, 2904, 2875, 1674, 1627, 1591, 1530, 1423, 1364, 1345, 1290, 1215, 1116, 1047, 875, 798, 767, 685, 582, 508, 461; HR-MS (ESI+) m/z calcd. for C25H30NO2+ 376.2271, found 376.2271, Δ = 0.0 mDa; HR-MS (ESI+) m/z calcd. for 79Br 78.9178, found 78.9176, Δ = −0.2 mDa.

3.3. Crystallography

Colorless block-shaped crystals of the title compound were obtained with slow evaporation from a hot acetone/acetonitrile solution. A suitable single crystal (dimensions 0.20 × 0.20 × 0.10 mm3) was mounted on a nylon loop using an oil-based cryoprotectant (Paratone ® N). Diffraction data were collected at 290 K on a Bruker D8 Venture diffractometer equipped with an IµS micro-focus sealed Mo X-ray source (λ = 071073 Å) and PHOTON II CPAD detector. Data reduction was performed with Bruker SAINT ver. 8.40B [38] software and a multi-scan absorption correction was performed with SADABS ver. 2016/2 [38] software, both included in the APEX4 [39] graphical interface. The structure was solved with the intrinsic phasing method (ShelxT ver. 2018/2 [40]), and the model was refined with the full-matrix least-squares method on F2 (ShelxL ver. 2019/1 [41]) integrated with Olex2 ver. 1.5 [42] graphical interface. All non-hydrogen atoms were located from electron density maps and were refined anisotropically. Hydrogen atoms were placed on calculated positions and refined using the riding model: Ueq = 1.2 for C-Haromatic = 0.93 Å and C-Hmethylenic = 0.97 Å and Ueq = 1.5 for C-Hmethyl = 0.96 Å. ORTEP-3 ver. 2020.1 software [37] was used to depict the molecules in the asymmetric unit (ASU). CCDC Mercury ver. 4.0 [43] was used for the illustration of the three-dimensional packing of the molecules. The most important data collection and crystallographic refinement parameters are given in Table S1. Bond lengths and bond angles for the crystal structure of the title compound are given in Tables S2 and S3. Complete crystallographic data for the reported structure were deposited in the CIF format with the Cambridge Crystallographic Data Centre as 2322651 (30 December 2023). These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html, or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +441223336033; E-mail: [email protected].

4. Conclusions

The title compound 1-(2-(3,5-di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) quinolin-1-ium bromide was obtained in good yield by a facile one-pot, one-step quaternization reaction between quinoline and an α-bromo ketone resulting in a colorless precipitate. The product was purified with hot recrystallization from an acetone/acetonitrile solution and characterized using powder and single-crystal XRD, NMR, DSC, FT-IR, and HRMS analyses. The single-crystal XRD analysis revealed that the compound crystallizes in the triclinic P-1 space group with acetonitrile molecule trapped in the crystal structure (crystal solvate). The crystal structure of the title compound is stabilized by halogen bonding interactions of the O–H…Br type.

Supplementary Materials

Figure S1: 1H NMR spectra of the title compound (1.0–9.5 ppm); Figure S2: 13C NMR of the title compound; Figure S3: NOESY spectra of the title compound (0.5–10 ppm); Figure S4: NOESY spectra of the title compound (7.0–9.5 ppm); Figure S5: HSQC spectra of the title compound; Figure S6: HMBC spectra of the title compound, Figure S7: FT-IR spectra of the title compound; Figure S8. HRMS spectrum in (a) positive mode and (b) negative mode of the title compound; Table S1: Most important crystallographic parameters obtained from the SCXRD of the title compound; Table S2: Bond lengths for the crystal structure of the title compound; Table S3: Bond angles for the crystal structure of the title compound; Table S4: Detected halogen bonding interactions for the crystal structure of the title compound.

Author Contributions

Conceptualization, R.R., K.I. and B.S.; formal analysis, R.R., V.K. and B.S.; visualization, R.R. and K.I.; investigation, R.R. and K.I.; methodology, R.R., V.K. and B.S.; writing—original draft, R.R.; funding acquisition, R.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Bulgarian National Science Fund (BNSF), grant number KP-06-M59/7.

Data Availability Statement

The data of the current study are available from the corresponding authors upon reasonable request.

Acknowledgments

The authors acknowledge the technical support from the project BG05M2OP001-1.002-0005, Perimed (2018–2023).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Abboud, J.-L.M.; Notario, R.; Bertrán, J.; Solà, M. One century of physical organic chemistry: The Menshutkin reaction. In Progress in Physical Organic Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 1993; Volume 19. [Google Scholar]
  2. Arnett, E.M.; Reich, R. Electronic effects on the Menshutkin reaction. A complete kinetic and thermodynamic dissection of alkyl transfer to 3-and 4-substituted pyridines. J. Am. Chem. Soc. 1980, 102, 5892–5902. [Google Scholar] [CrossRef]
  3. Sola, M.; Lledos, A.; Duran, M.; Bertran, J.; Abboud, J.L.M. Analysis of solvent effects on the Menshutkin reaction. J. Am. Chem. Soc. 1991, 113, 2873–2879. [Google Scholar] [CrossRef]
  4. Castejon, H.; Wiberg, K.B. Solvent effects on methyl transfer reactions. 1. The Menshutkin reaction. J. Am. Chem. Soc. 1999, 121, 2139–2146. [Google Scholar] [CrossRef]
  5. Lewinska, G.; Sanetra, J.; Marszalek, K.W. Application of quinoline derivatives in third-generation photovoltaics. J. Mater. Sci. Mater. Electron. 2021, 32, 18451–18465. [Google Scholar] [CrossRef]
  6. Sivaraman, G.; Iniya, M.; Anand, T.; Kotla, N.G.; Sunnapu, O.; Singaravadivel, S.; Gulyani, A.; Chellappa, D. Chemically diverse small molecule fluorescent chemosensors for copper ion. Coord. Chem. Rev. 2018, 357, 50–104. [Google Scholar] [CrossRef]
  7. Matada, B.S.; Pattanashettar, R.; Yernale, N.G. A comprehensive review on the biological interest of quinoline and its derivatives. Bioorg. Med. Chem. 2021, 32, 115973. [Google Scholar] [CrossRef]
  8. Yadav, P.; Shah, K. Quinolines, a perpetual, multipurpose scaffold in medicinal chemistry. Bioorg. Chem. 2021, 109, 104639. [Google Scholar] [CrossRef] [PubMed]
  9. El Shehry, M.F.; Ghorab, M.M.; Abbas, S.Y.; Fayed, E.A.; Shedid, S.A.; Ammar, Y.A. Quinoline derivatives bearing pyrazole moiety: Synthesis and biological evaluation as possible antibacterial and antifungal agents. Eur. J. Med. Chem. 2018, 143, 1463–1473. [Google Scholar] [CrossRef]
  10. Yeo, S.-J.; Liu, D.-X.; Kim, H.S.; Park, H. Anti-malarial effect of novel chloroquine derivatives as agents for the treatment of malaria. Malar. J. 2017, 16, 80. [Google Scholar] [CrossRef]
  11. Quimque, M.T.J.; Go, A.D.; Lim, J.A.K.; Vidar, W.S.; Macabeo, A.P.G. Mycobacterium tuberculosis inhibitors based on arylated quinoline carboxylic acid backbones with anti-mtb gyrase activity. Int. J. Mol. Sci. 2023, 24, 11632. [Google Scholar] [CrossRef]
  12. Abdelrahman, M.A.; Almahli, H.; Al-Warhi, T.; Majrashi, T.A.; Abdel-Aziz, M.M.; Eldehna, W.M.; Said, M.A. Development of novel isatin-tethered quinolines as anti-tubercular agents against multi and extensively drug-resistant mycobacterium tuberculosis. Molecules 2022, 27, 8807. [Google Scholar] [CrossRef] [PubMed]
  13. Mantu, D.; Antoci, V.; Moldoveanu, C.; Zbancioc, G.; Mangalagiu, I.I. Hybrid imidazole (benzimidazole)/pyridine (quinoline) derivatives and evaluation of their anticancer and antimycobacterial activity. J. Enzym. Inhib. Med. Chem. 2016, 31, 96–103. [Google Scholar] [CrossRef] [PubMed]
  14. Bingul, M.; Tan, O.; Gardner, C.R.; Sutton, S.K.; Arndt, G.M.; Marshall, G.M.; Cheung, B.B.; Kumar, N.; Black, D.S. Synthesis, characterization and anti-cancer activity of hydrazide derivatives incorporating a quinoline moiety. Molecules 2016, 21, 916. [Google Scholar] [CrossRef] [PubMed]
  15. Eissa, S.I.; Farrag, A.M.; Abbas, S.Y.; El Shehry, M.F.; Ragab, A.; Fayed, E.A.; Ammar, Y.A. Novel structural hybrids of quinoline and thiazole moieties: Synthesis and evaluation of antibacterial and antifungal activities with molecular modeling studies. Bioorg. Chem. 2021, 110, 104803. [Google Scholar] [CrossRef] [PubMed]
  16. Fu, H.-G.; Li, Z.-W.; Hu, X.-X.; Si, S.-Y.; You, X.-F.; Tang, S.; Wang, Y.-X.; Song, D.-Q. Synthesis and biological evaluation of quinoline derivatives as a novel class of broad-spectrum antibacterial agents. Molecules 2019, 24, 548. [Google Scholar] [CrossRef] [PubMed]
  17. Banerjee, A.; Dash, S.P.; Mohanty, M.; Sanna, D.; Sciortino, G.; Ugone, V.; Garribba, E.; Reuter, H.; Kaminsky, W.; Dinda, R. Chemistry of mixed-ligand oxidovanadium (iv) complexes of aroylhydrazones incorporating quinoline derivatives: Study of solution behavior, theoretical evaluation and protein/DNA interaction. J. Inorg. Biochem. 2019, 199, 110786. [Google Scholar] [CrossRef] [PubMed]
  18. El-Sonbati, A.Z.; Diab, M.A.; Morgan, S.M.; Abou-Dobara, M.I.; El-Ghettany, A.A. Synthesis, characterization, theoretical and molecular docking studies of mixed-ligand complexes of cu(ii), ni(ii), co(ii), mn(ii), cr(iii), uo2(ii) and cd(ii). J. Mol. Struct. 2020, 1200, 127065. [Google Scholar] [CrossRef]
  19. Elbatrawy, A.A.; Hyeon, S.J.; Yue, N.; Osman, E.E.A.; Choi, S.H.; Lim, S.; Kim, Y.K.; Ryu, H.; Cui, M.; Nam, G. “Turn-on” quinoline-based fluorescent probe for selective imaging of tau aggregates in alzheimer’s disease: Rational design, synthesis, and molecular docking. ACS Sens. 2021, 6, 2281–2289. [Google Scholar] [CrossRef]
  20. Pradhan, A.B.; Mandal, S.K.; Banerjee, S.; Mukherjee, A.; Das, S.; Bukhsh, A.R.K.; Saha, A. A highly selective fluorescent sensor for zinc ion based on quinoline platform with potential applications for cell imaging studies. Polyhedron 2015, 94, 75–82. [Google Scholar] [CrossRef]
  21. Yao, Y.; Ding, P.; Yan, C.; Tao, Y.; Peng, B.; Liu, W.; Wang, J.; Stuart, M.A.C.; Guo, Z. Fluorescent probes based on aiegen-mediated polyelectrolyte assemblies for manipulating intramolecular motion and magnetic relaxivity. Angew. Chem. Int. Ed. 2023, 62, e202218983. [Google Scholar] [CrossRef]
  22. Tselekidou, D.; Panagiotidis, L.; Papadopoulos, K.; Kyriazopoulos, V.; Gioti, M. A comparative study of ir (dmpq) 2 (acac) doped cbp, mcp, tapc and tcta for phosphorescent oleds. Photonics 2022, 9, 800. [Google Scholar] [CrossRef]
  23. Ali, B.; Khalid, M.; Asim, S.; Khan, M.U.; Iqbal, Z.; Hussain, A.; Hussain, R.; Ahmed, S.; Ali, A.; Hussain, A. Key electronic, linear and nonlinear optical properties of designed disubstituted quinoline with carbazole compounds. Molecules 2021, 26, 2760. [Google Scholar] [CrossRef] [PubMed]
  24. Daoud, D.; Hamani, H.; Douadi, T. Novel heterocyclic quinoline derivatives as green environmental corrosion inhibitors for carbon steel in hcl solution: Experimental and theoretical investigation. J. Adhes. Sci. Technol. 2021, 35, 2319–2345. [Google Scholar] [CrossRef]
  25. Abdelshafi, N.; Ibrahim, M.A.; Badran, A.-S.; Halim, S.A. Experimental and theoretical evaluation of a newly synthesized quinoline derivative as corrosion inhibitor for iron in 1.0 m hydrochloric acid solution. J. Mol. Struct. 2022, 1250, 131750. [Google Scholar] [CrossRef]
  26. Nikitin, E.; Mironova, E.; Shpakovsky, D.; Gracheva, Y.; Koshelev, D.; Utochnikova, V.; Lyssenko, K.; Oprunenko, Y.; Yakovlev, D.; Litvinov, R. Cytotoxic and luminescent properties of novel organotin complexes with chelating antioxidant ligand. Molecules 2022, 27, 8359. [Google Scholar] [CrossRef] [PubMed]
  27. Osipova, V.; Gracheva, Y.; Polovinkina, M.; Burmistrova, D.; Berberova, N. Antioxidant activity and cytotoxicity of aromatic oligosulfides. Molecules 2022, 27, 3961. [Google Scholar] [CrossRef] [PubMed]
  28. Koshelev, V.N.; Primerova, O.V.; Vorobyev, S.V.; Ivanova, L.V. Synthesis, redox properties and antibacterial activity of hindered phenols linked to heterocycles. Molecules 2020, 25, 2370. [Google Scholar] [CrossRef]
  29. Ariffin, A.; Rahman, N.A.; Yehye, W.A.; Alhadi, A.A.; Kadir, F.A. Pass-assisted design, synthesis and antioxidant evaluation of new butylated hydroxytoluene derivatives. Eur. J. Med. Chem. 2014, 87, 564–577. [Google Scholar] [CrossRef]
  30. Kong, C.; Yehye, W.A.; Rahman, N.A.; Tan, M.-W.; Nathan, S. Discovery of potential anti-infectives against staphylococcus aureus using a caenorhabditis elegans infection model. BMC Complement Altern. Med. 2014, 14, 4. [Google Scholar] [CrossRef]
  31. Shakir, R.M.; Ariffin, A.; Abdulla, M.A. Synthesis of new 2, 5-di-substituted 1, 3, 4-oxadiazoles bearing 2, 6-di-tert-butylphenol moieties and evaluation of their antioxidant activity. Molecules 2014, 19, 3436–3449. [Google Scholar] [CrossRef]
  32. Milaeva, E.; Filimonova, S.; Meleshonkova, N.; Dubova, L.; Shevtsova, E.; Bachurin, S.; Zefirov, N. Antioxidative activity of ferrocenes bearing 2, 6-di-tert-butylphenol moieties. Bioinorg. Chem. Appl. 2010, 2010, 165482. [Google Scholar] [CrossRef] [PubMed]
  33. Gibadullina, E.; Thu, N.T.; Bukharov, S.; Burilov, A. Synthesis of sterically hindered phenols based on 7-amino-2, 4-dimethylqunoline and 5, 7-dimethyl-1, 8-naphthyridin-2-amine. Russ. J. Gen. Chem. 2017, 87, 1305–1309. [Google Scholar] [CrossRef]
  34. Mikhalev, O.; Shpakovsky, D.; Gracheva, Y.A.; Antonenko, T.; Albov, D.; Aslanov, L.; Milaeva, E. Synthesis and study of new phenolic antioxidants with nitroaromatic and heterocyclic substituents. Russ. Chem. Bull. 2018, 67, 712–720. [Google Scholar] [CrossRef]
  35. Dudognon, E.; Correia, N.T.; Danède, F.; Descamps, M. Solid-solid transformation in racemic ibuprofen. Pharm. Res. 2013, 30, 81–89. [Google Scholar] [CrossRef] [PubMed]
  36. Rossi, P.; Macedi, E.; Paoli, P.; Bernazzani, L.; Carignani, E.; Borsacchi, S.; Geppi, M. Solid–solid transition between hydrated racemic compound and anhydrous conglomerate in na-ibuprofen: A combined X-ray diffraction, solid-state NMR, calorimetric, and computational study. Cryst. Growth Des 2014, 14, 2441–2452. [Google Scholar] [CrossRef]
  37. Farrugia, L.J. Wingx and ortep for windows: An update. J. Appl. Crystallogr. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  38. Bruker, A. Saint and Sadabs; Bruker AXS Inc.: Madison, WI, USA, 2009. [Google Scholar]
  39. Bruker, A. Apex 4; Bruker AXS Inc.: Madison, WI, USA, 2022. [Google Scholar]
  40. Sheldrick, G.M. Shelxt–integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A 2015, A71, 3–8. [Google Scholar] [CrossRef]
  41. Sheldrick, G.M. Crystal structure refinement with shelxl. Acta Crystallogr. Sect. C 2015, C71, 3–8. [Google Scholar] [CrossRef]
  42. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.; Puschmann, H. Olex2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  43. Macrae, C.F.; Sovago, I.; Cottrell, S.J.; Galek, P.T.; McCabe, P.; Pidcock, E.; Platings, M.; Shields, G.P.; Stevens, J.S.; Towler, M. Mercury 4.0: From visualization to analysis, design and prediction. J. Appl. Crystallogr. 2020, 53, 226–235. [Google Scholar] [CrossRef]
Figure 1. Representatives compounds bearing quinoline fragment [13,14,15,16,19,20,21,22,24].
Figure 1. Representatives compounds bearing quinoline fragment [13,14,15,16,19,20,21,22,24].
Molbank 2024 m1763 g001
Figure 2. Representatives compounds bearing 2,6-di-tert-butylphenol fragment [26,27,28,29,30,31,32].
Figure 2. Representatives compounds bearing 2,6-di-tert-butylphenol fragment [26,27,28,29,30,31,32].
Molbank 2024 m1763 g002
Scheme 1. Procedure for synthesis of 1-(2-(3,5-di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) quinolin-1-ium bromide.
Scheme 1. Procedure for synthesis of 1-(2-(3,5-di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) quinolin-1-ium bromide.
Molbank 2024 m1763 sch001
Figure 3. 1H NMR spectra of the title compound.
Figure 3. 1H NMR spectra of the title compound.
Molbank 2024 m1763 g003
Figure 4. DSC thermogram of the title compound.
Figure 4. DSC thermogram of the title compound.
Molbank 2024 m1763 g004
Figure 5. ORTEP [37] view of the molecules of the novel quaternary ammonium compound present in the asymmetric unit (ASU) with appropriate labeling scheme. Atomic displacement parameters (ADPs) are at 50%, and the hydrogen atoms are given as spheres with arbitrary radii.
Figure 5. ORTEP [37] view of the molecules of the novel quaternary ammonium compound present in the asymmetric unit (ASU) with appropriate labeling scheme. Atomic displacement parameters (ADPs) are at 50%, and the hydrogen atoms are given as spheres with arbitrary radii.
Molbank 2024 m1763 g005
Figure 6. Representation of (a) unit cell content and (b) three-dimensional packing with view perpendicular to b axis for the crystal structure of the title compound. For clarity, the two independent molecules in the ASU are given in blue and orange, while the bromine anions and the atoms of the acetonitrile molecules are given as spheres with arbitrary radii.
Figure 6. Representation of (a) unit cell content and (b) three-dimensional packing with view perpendicular to b axis for the crystal structure of the title compound. For clarity, the two independent molecules in the ASU are given in blue and orange, while the bromine anions and the atoms of the acetonitrile molecules are given as spheres with arbitrary radii.
Molbank 2024 m1763 g006
Figure 7. Comparison between the powder XRD pattern of the bulk sample (in red) and the one generated from the single-crystal XRD experiment (in blue) confirming the monophasicity of the studied compound.
Figure 7. Comparison between the powder XRD pattern of the bulk sample (in red) and the one generated from the single-crystal XRD experiment (in blue) confirming the monophasicity of the studied compound.
Molbank 2024 m1763 g007
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

Rusew, R.; Iliev, K.; Kurteva, V.; Shivachev, B. 1-(2-(3,5-Di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) Quinolin-1-ium Bromide. Molbank 2024, 2024, M1763. https://doi.org/10.3390/M1763

AMA Style

Rusew R, Iliev K, Kurteva V, Shivachev B. 1-(2-(3,5-Di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) Quinolin-1-ium Bromide. Molbank. 2024; 2024(1):M1763. https://doi.org/10.3390/M1763

Chicago/Turabian Style

Rusew, Rusi, Kostadin Iliev, Vanya Kurteva, and Boris Shivachev. 2024. "1-(2-(3,5-Di-tert-butyl-4-hydroxyphenyl)-2-oxoethyl) Quinolin-1-ium Bromide" Molbank 2024, no. 1: M1763. https://doi.org/10.3390/M1763

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