Next Article in Journal
(E)-7-(4-(Diphenylamino)styryl)benzo[c][1,2,5]thiadiazole-4-carbonitrile
Previous Article in Journal
2-(4-Chlorophenyl)-4-(3,4-dimethoxy-phenyl)-6-methoxy-3-methylquinoline
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Direct Method for the Efficient Synthesis of Hydroxyalkyl-Containing Azoxybenzenes

by
Ekaterina S. Spesivaya
,
Ida A. Lupanova
,
Dzhamilya N. Konshina
and
Valery V. Konshin
*
Department of Chemistry and High Technology, Kuban State University, Stavropolskayast 149, 350040 Krasnodar, Russia
*
Author to whom correspondence should be addressed.
Molbank 2022, 2022(2), M1384; https://doi.org/10.3390/M1384
Submission received: 12 May 2022 / Revised: 7 June 2022 / Accepted: 8 June 2022 / Published: 14 June 2022
(This article belongs to the Section Organic Synthesis)

Abstract

:
Reaction of nitrobenzyl alcohol with glucose (200 mol%) in the presence of NaOH in water-ethanol medium gave 1,2-bis(4-(hydroxymethyl)phenyl)diazene oxide, 1,2-bis(2-(hydroxymethyl)phenyl)diazene oxide and 1,2-bis(4-(1-hydroxyethyl)phenyl)diazene oxide in 76%, 76% and 72% yields, respectively.

1. Introduction

The azoxybenzene moiety is a part of many compounds with practically valuable properties: liquid crystalline compounds [1,2,3] and bioactive substances [4,5,6]. Classical approaches to the synthesis of azoxybenzenes are based on numerous versions of the reduction of nitroaromatic substrates [7,8,9] or the oxidation of aminoarenes [10,11,12,13,14] and other methods [15]. However, such a route is not optimal where substrates contain functional groups susceptible to reduction or oxidation conditions. For this reason, it seems interesting to increase the molecular complexity of azoxybenzenes bearing reactive functions, such as carboxy- [16], formyl- [17], amino- [18,19], alkyne- [12,20,21], hydroxyl- [22], and other groups. Earlier, we demonstrated an example of the azoxybenzene preparation through reduction with glucose [23] in the presence of sodium hydroxide in the water–ethanol medium to synthesize 1,3-dioxolanyl-containing azoxybenzen [24]. In the present work, this method was extended to the preparation of hydroxyalkyl-containing azoxybenzenes 3, some of which were synthesized earlier using different methods [22,25,26,27,28,29,30] (Scheme 1).

2. Results and Discussion

Various nitroaryl-containing benzyl alcohols 2 (commercially available or easy synthesizable by known methods in high yields [31,32,33]) can serve as convenient starting compounds for the synthesis of hydroxyalkyl-containing azoxybenzenes.
Similar to 2-(4-nitrophenyl)-1,3-dioxolane [24], the reduction of nitrobenzyl alcohols 2 readily proceeds on exposure to glucose (200 mol%) in the presence of sodium hydroxide in the water–ethanol medium. The TLC or GC/MS control of the reaction course shows that complete conversion of the substrate occurs within 30 min of keeping the reaction mixture at 50 °C and vigorous stirring (Scheme 2).
The target products 3 precipitate as a crystalline solid after dilution of the reaction mass with water. Products were filtered off, washed with water, and further purified by recrystallization or flash chromatography.
The structure of 3 was confirmed by 1H and 13C-NMR spectroscopy, IR spectroscopy, and high-resolution mass spectrometry.
In this case, successful reduction to azoxy derivatives was due to a combination of experimental conditions, namely the reaction time, reaction medium and lack of aeration. It is also known that the reduction of nitroarenes by glucose can be carried out to obtain azo compounds [34,35,36] and aromatic amines [37].

3. Materials and Methods

The reactions were monitored by thin-layer chromatography (Sorbfil, Imid Ltd., Krasnodar, Russia). The 1H-NMR, 13C-NMR spectra were acquired on ECA400 (JEOL) (400 and 100 MHz, respectively) spectrometers in (CD3)2SO at room temperature. The chemical shifts δ were measured in ppm with reference to the residual solvent resonances (1H: (CD3)2SO, δ = 2.49 ppm; 13C: (CD3)2SO, δ = 39.5 ppm). The splitting patterns are referred to as s—singlet; d—doublet; t—triplet; m—multiplet. Coupling constants (J) are given in hertz. IR spectra were recorded on an IR Prestige (Shimadzu), using tablets of samples with KBr. High-resolution and accurate mass measurements were carried out using a Bruker MaXis Impact (electrospray ionization/time of flight). The melting points were determined on Stuart SMP30 apparatus and left uncorrected. The commercial reagents employed in the synthesis were 2-Nitrobenzaldehyde (98%, Aldrich, St. Louis, MS, USA), 4-Nitrobenzaldehyde (98%, Aldrich, St. Louis, MS, USA), 4′-Nitroacetophenone (98%, Aldrich, St. Louis, MS, USA), Sodium borohydride (powder, 98%, ABCR) and D-(+)-Glucose monohydrate (≥99%, Vekton, Russia).

General Procedure for the Preparation of Hydroxyalkyl-Containing Azoxybenzenes (3)

Nitrobenzyl alcohol 2 (3.26 mmol) in ethanol (6 mL) was added to 30% aq. NaOH (7.5 mL). At 50 °C, a solution of glucose monohydrate (1.29 g, 6.5 mmol) in water (1 mL) was added. The mixture was stirred for 30 min at 50 °C. The reaction was cooled, diluted with water, and the formed precipitate was filtered and washed with distilled water. The resulting residue was purified via recrystallization from EtOH.
1,2-bis(4-(hydroxymethyl)phenyl)diazene oxide 3a. Yield 0.32 g (76%); light beige crystals; mp 163–164 °C (EtOH). IR (KBr): ν = 3284 (OH), 3055 (Csp2-H), 2974, 2933, 2899, 2872 (Csp3-H), 1600, 1498 (Csp2-Csp2), 1465, 1408, 1327, 1278, 1199, 1159, 1035, 1010, 825 cm−1 (Supplementary Materials, Figure S1). 1H-NMR ((CD3)2SO, 399.78 MHz): δ = 4.57 (s., 2H, CH2), 4.62 (s., 2H, CH2), 5.33 (br. s., 1H, OH), 5.43 (br. s., 1H, OH), 7.45–7.50 (m., 2H, CH), 7.52–7.56 (m., 2H, CH), 8.07–8.11 (m., 2H, CH), 8.18–8.23 (m., 2H, CH) (Supplementary Materials, Figure S2). 13C-NMR ((CD3)2SO, 100.5 MHz): δ = 62.1 (CH2), 62.5 (CH2), 121.8 (CH), 125.0 (CH), 126.6 (CH), 126.7 (CH), 142.2 (C), 144.6 (C), 146.3 (C), 147.0 (C) (Supplementary Materials, Figure S3). HRMS ESI TOF: m/z = 259.1082 [M+H]+ (259,1077 calc. for C14H14N2O3) (Supplementary Materials, Figure S4).
1,2-bis(2-(hydroxymethyl)phenyl)diazene oxide 3b. Yield 0.32 g (76%); light beige crystals; mp 118–120 °C (EtOH). IR (KBr): ν = 3230 (OH), 2922, 2864 (Csp3-H), 1483, 1458, 1440, 1365, 1330, 1184, 1039, 1012, 921, 750 cm−1 (Supplementary Materials, Figure S5). 1H-NMR ((CD3)2SO, 399.78 MHz): δ = 4.58 (d., 5.7 Hz, 2H, CH2), 4.74 (d., 5.7 Hz, 2H, CH2), 5.20 (t., 5.7 Hz, 1H, OH), 5.41 (t., 5.7 Hz, 1H, OH), 7.35–7.42 (m., 2H), 7.45–7.51 (m., 1H), 7.58–7.62 (m., 2H), 7.70–7.74 (m., 1H), 7.77–7.82 (m., 1H), 7.88–7.92 (m., 1H) (Supplementary Materials, Figure S6). 13C-NMR ((CD3)2SO, 100.5 MHz): δ = 59.22 (CH2), 59.35 (CH2), 121.0 (CH), 123.4 (CH), 126.8 (CH), 127.2 (CH), 127.6 (CH), 128.1 (CH), 128.4 (CH), 130.5 (CH), 135.6 (C), 137.6 (C), 140.8 (C), 147.0 (C) (Supplementary Materials, Figure S7). HRMS ESI TOF: m/z = 241.0978 [(M-H2O)+H]+ (calc. 241,0971 for C14H12N2O2 [3b-H2O]) (Supplementary Materials, Figure S8).
1,2-bis(4-(1-hydroxyethyl)phenyl)diazene oxide 3c. Yield 0.34 g (72%); light beige crystals; mp 105–106 °C (EtOH). IR (KBr): ν = 3327 (OH), 3074, 3053, 3034 (Csp2-H), 2970, 2924 (Csp3-H), 1602, 1498 (Csp2-Csp2), 1462, 1409, 1367, 1327, 1286, 1203, 1083, 1010, 896, 840 cm−1 (Supplementary Materials, Figure S9). 1H-NMR ((CD3)2SO, 399.78 MHz): δ = 1.34 (d., 2.3 Hz, 3H, CH3), 1.35 (d., 2.5 Hz, 3H, CH3), 4.57–4.87 (m, 2H, CH), 5.30 (d., 4.3 Hz, 1H, OH), 5.40 (d., 4.4 Hz, 1H, OH), 7.48–7.52 (m, 2H, CH), 7.54–7.58 (m, 2H, CH), 8.06–8.09 (m, 2H, CH), 8.17–8.20 (m, 2H, CH) (Supplementary Materials, Figure S10). 13C-NMR ((CD3)2SO, 100.5 MHz): δ = 25.78 (CH3), 25.82 (CH3), 67.47 (CH), 67.51 (CH), 67.77 (CH), 67.80 (CH), 121.8 (CH), 125.0 (CH), 125.7 (CH), 126.0 (CH), 142.1 (C), 146.2 (C), 149.4 (C), 151.7 (C) (Supplementary Materials, Figure S11). HRMS ESI TOF: m/z = 287.1397 [M+H]+ (287,1390 calc. for C16H18N2O3) (Supplementary Materials, Figure S12).

Supplementary Materials

The following can be downloaded online. Figure S1: IR-spectrum of 3a; Figure S2: 1H-NMR of 3a; Figure S3: 13C-NMR of 3a; Figure S4: HRMS of 3a; Figure S5: IR-spectrum of 3b; Figure S6: 1H-NMR of 3b; Figure S7: 13C-NMR of 3b; Figure S8: HRMS of 3b; Figure S9: IR-spectrum of 3c; Figure S10: 1H-NMR of 3c; Figure S11: 13C-NMR of 3c; Figure S12: HRMS of 3c.

Author Contributions

Conceptualization, V.V.K.; methodology, V.V.K.; software, D.N.K.; validation, V.V.K. and D.N.K.; formal analysis, D.N.K.; investigation, I.A.L. and E.S.S.; resources, I.A.L. and E.S.S.; data curation, V.V.K.; writing—original draft preparation, V.V.K. and D.N.K.; writing—review and editing, V.V.K. and D.N.K.; supervision, V.V.K.; project administration, V.V.K.; funding acquisition, V.V.K., D.N.K. and I.A.L. All authors have read and agreed to the published version of the manuscript.

Funding

The research was carried out with the financial support of the Kuban Science Foundation in the framework of the scientific project N° MFI-20.1-28/20.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The HRMS study was accomplished with the use of scientific equipment of the Collective Employment Centre “Ecoanalytical Centre”, Kuban State University (A. Z. Temerdashev).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Pintre, I.C.; Gimeno, N.; Serrano, J.L.; Ros, M.B.; Alonso, I.; Folcia, C.L.; Ortega, J.; Etxebarria, J. Liquid crystalline and nonlinear optical properties of bent-shaped compounds derived from 3,4′-biphenylene. J. Mater. Chem. 2007, 17, 2219–2227. [Google Scholar] [CrossRef] [Green Version]
  2. Mulani, K.B.; Ganjave, N.V.; Chavan, N.N. Synthesis and characterization of azoxy based mesogenic diols. Indian J. Chem.-Sect. B Org. Med. Chem. 2014, 53, 359–362. [Google Scholar]
  3. Giricheva, N.I.; Lebedev, I.S.; Fedorov, M.S. Influence of structural features of azo-, azoxy-, azodioxy-benzenes and pyridines on mesomorphic properties of systems on their basis. Liq. Cryst. Their Appl. 2021, 21, 37–46. (In Russian) [Google Scholar] [CrossRef]
  4. Guo, Y.-Y.; Li, H.; Zhou, Z.-X.; Mao, X.-M.; Tang, Y.; Chen, X.; Jiang, X.-H.; Liu, Y.; Jiang, H.; Li, Y.-Q. Identification and biosynthetic characterization of natural aromatic azoxy products from Streptomyces Chattanoogensis L10. Org. Lett. 2015, 17, 6114–6117. [Google Scholar] [CrossRef] [PubMed]
  5. Takahashi, H.; Ishioka, T.; Koiso, Y.; Sodeoka, M.; Hashimoto, Y. Anti-androgenic activity of substituted azo- and azoxy-benzene derivatives. Biol. Pharm. Bull. 2000, 23, 1387–1390. [Google Scholar] [CrossRef] [Green Version]
  6. Kotova, V.A.; Rubanova, E.V.; Jatsynin, V.G. Winter Wheat and Barley Roots Stimulant Fertilizer. Patent RU2368140, 7 April 2008. [Google Scholar]
  7. Arnab, G.; Limaye, A.S.; Manjunatha, K.N.; Patil, S.A.; Dateer, R.B. Zn-Mediated Selective Reduction of Nitroarenes: A Sustainable Approach for Azoxybenzenes Synthesis. Org. Prep. Proced. Int. 2022, 54, 284–293. [Google Scholar] [CrossRef]
  8. Nishiyama, Y.; Fujii, A.; Mori, H. Photoreduction synthesis of various azoxybenzenes by visible-light irradiation under continuous flow conditions. J. Flow Chem. 2022, 12, 71–77. [Google Scholar] [CrossRef]
  9. Jin, M.; Liu, Y.; Zhang, X.; Wang, J.; Zhang, S.; Wang, G.; Zhang, Y.; Yin, H.; Zhang, H.; Zhao, H. Selective electrocatalytic hydrogenation of nitrobenzene over copper-platinum alloying catalysts: Experimental and theoretical studies. Appl. Catal. B 2021, 298, 120545. [Google Scholar] [CrossRef]
  10. Rezaeifard, A.; Jafarpour, M.; Naseri, M.A.; Shariati, R. A rapid and easy method for the synthesis of azoxy arenes using tetrabutylammonium peroxymonosulfate. Dyes Pigments 2008, 76, 840–843. [Google Scholar] [CrossRef]
  11. Singh, B.; Mandelli, D.; Pescarmona, P.P. Efficient and selective oxidation of aromatic amines to azoxy derivatives over aluminium and gallium oxide catalysts with nanorod morphology. ChemCatChem 2020, 12, 593–601. [Google Scholar] [CrossRef] [Green Version]
  12. Tanini, D.; Dalia, C.; Capperucci, A. The polyhedral nature of selenium-catalysed reactions: Se(IV) species instead of se(VI) species make the difference in the on water selenium-mediated oxidation of arylamines. Green Chem. 2021, 23, 5680–5686. [Google Scholar] [CrossRef]
  13. De Carvalho, G.S.G.; Chagas, L.H.; Fonseca, C.G.; De Castro, P.P.; Sant’Ana, A.C.; Leitão, A.A.; Amarante, G.W. Nb2O5 supported on mixed oxides catalyzed oxidative and photochemical conversion of anilines to azoxybenzenes. New J. Chem. 2019, 43, 5863–5871. [Google Scholar] [CrossRef]
  14. Qin, J.; Long, Y.; Sun, F.; Zhou, P.-P.; Wang, W.D.; Luo, N.; Ma, J. Zr(OH)4-catalyzed controllable selective oxidation of anilines to azoxybenzenes, azobenzenes and nitrosobenzenes. Angew. Chem. Int. Ed. 2022, 61, e202112907. [Google Scholar] [CrossRef] [PubMed]
  15. Zlotin, S.G.; Luk’yanov, O.A. Regioselective methods of synthesis of asymmetrically substituted diazene oxides. Russ. Chem. Rev. 1993, 62, 143–168. [Google Scholar] [CrossRef]
  16. Liu, Y.; Liu, B.; Guo, A.; Dong, Z.; Jin, S.; Lu, Y. Reduction of Nitroarenes to Azoxybenzenes by Potassium Borohydride in Water. Molecules 2011, 16, 3563–3568. [Google Scholar] [CrossRef]
  17. Boduszek, B.; Halama, A. Nitrobenzyl (α-amino)phosphonates. part 2[1]. cleavage of 4-nitrobenzyl(α-amino)phosphonic acids in aqueous sodium hydroxide solution. Phosphorus Sulfur Silicon Relat. Elem. 1998, 141, 239–250. [Google Scholar] [CrossRef]
  18. Yang, F.; Wang, Z.; Zhang, X.; Jiang, L.; Li, Y.; Wang, L. A Green Chemoenzymatic Process for the Synthesis of Azoxybenzenes. ChemCatChem 2015, 7, 3450–3453. [Google Scholar] [CrossRef]
  19. Pahalagedara, M.N.; Pahalagedara, L.R.; He, J.; Miao, R.; Gottlieb, B.; Rathnayake, D.; Suib, S.L. Room temperature selective reduction of nitrobenzene to azoxybenzene over magnetically separable urchin-like Ni/Graphene nanocomposites. J. Catal. 2016, 336, 41–48. [Google Scholar] [CrossRef] [Green Version]
  20. Kim, J.H.; Park, J.H.; Chung, Y.K.; Park, K.H. Ruthenium Nanoparticle-Catalyzed, Controlled and Chemoselective Hydrogenation of Nitroarenes using Ethanol as a Hydrogen Source. Adv. Synth. Catal. 2012, 354, 2412–2418. [Google Scholar] [CrossRef]
  21. Tani, H.; Tanaka, S.; Toda, F. Oligomers and Polymers Containing Triple Bonds. II. Derivatives of Ethynylazobenzene. Bull. Chem. Soc. Jpn. 1963, 36, 1267–1271. [Google Scholar] [CrossRef] [Green Version]
  22. Lakshminarayana, B.; Manna, A.K.; Satyanarayana, G.; Subrahmanyam, C. Palladium nanoparticles on silica nanospheres for switchable reductive coupling of nitroarenes. Catal. Lett. 2020, 150, 2309–2321. [Google Scholar] [CrossRef]
  23. Galbraith, H.W.; Degering, E.F.; Hitch, E.F. The alkaline reduction of aromatic nitro compounds with glucose. J. Am. Chem. Soc. 1951, 73, 1323–1324. [Google Scholar] [CrossRef]
  24. Spesivaya, E.S.; Lupanova, I.A.; Konshina, D.N.; Konshin, V.V. 1,2-Bis(4-(1,3-dioxolan-2-yl)phenyl)diazene Oxide. Molbank 2021, 2021, M1224. [Google Scholar] [CrossRef]
  25. Shine, H.J.; Mallory, H.E. The reduction of aromatic nitro compounds by potassium borohydride. J. Org. Chem. 1962, 27, 2390–2391. [Google Scholar] [CrossRef]
  26. Ohe, K.; Takahashi, H.; Uemura, S.; Sugita, N. Sodium benzenetellurolate-catalysed selective reduction of aromatic nitro compounds to azoxy compounds. J. Chem. Soc. Chem. Commun. 1988, 9, 591–592. [Google Scholar] [CrossRef]
  27. Ohe, K.; Uemura, S.; Sugita, N. Sodium arenetellurolate catalysed selective conversion of nitroaromatics to aromatic azoxy or azo compounds and its application for facile preparation of 3,3′- and 4,4′-bis[β-(aryltelluro)vinyl]azobenzenes from (3- and 4-nitrophenyl)acetylenes. J. Org. Chem. 1989, 54, 4169–4174. [Google Scholar] [CrossRef]
  28. Zeynizadeh, B.; Gilanizadeh, M. Green and highly efficient approach for the reductive coupling of nitroarenes to azoxyarenes using the new mesoporous Fe3O4@SiO2@Co–Zr–Sb catalyst. Res. Chem. Intermed. 2020, 46, 2969–2984. [Google Scholar] [CrossRef]
  29. Bhosale, S.M.; Momin, A.A.; Kunjir, S.; Rajamohanan, P.R.; Kusurkar, R.S. Unexpected observations during the total synthesis of calothrixin B–sodium methoxide as a source of hydride. Tetrahedron Lett. 2014, 55, 155–162. [Google Scholar] [CrossRef]
  30. Yan, Z.; Xie, X.; Song, Q.; Ma, F.; Sui, X.; Huo, Z.; Ma, M. Tandem selective reduction of nitroarenes catalyzed by palladium nanoclusters. Green Chem. 2020, 22, 1301–1307. [Google Scholar] [CrossRef]
  31. Desroches, J.; Champagne, P.A.; Benhassine, Y.; Paquin, J.-F. In situ activation of benzyl alcohols with XtalFluor-E: Formation of 1,1-diarylmethanes and 1,1,1-triarylmethanes through Friedel–Crafts benzylation. Org. Biomol. Chem. 2015, 13, 2243–2246. [Google Scholar] [CrossRef]
  32. Storz, M.P.; Maurer, C.K.; Zimmer, C.; Wagner, N.; Brengel, C.; de Jong, J.C.; Lucas, S.; Müsken, M.; Häussler, S.; Steinbach, A.; et al. Validation of PqsD as an Anti-biofilm Target in Pseudomonas aeruginosa by Development of Small-Molecule Inhibitors. J. Am. Chem. Soc. 2012, 134, 16143–16146. [Google Scholar] [CrossRef] [PubMed]
  33. Eppacher, S.; Giester, G.; Bats, J.W.; Noe, C.R. Enantiomerically Pure Poly(oxymethylene) Helices: Correlating Helicity with Centrochirality. Helv. Chim. Acta 2008, 91, 581–597. [Google Scholar] [CrossRef]
  34. Ðorđević, L.; Casimiro, L.; Demitri, N.; Baroncini, M.; Silvi, S.; Arcudi, F.; Credi, A.; Prato, M. Light-Controlled Regioselective Synthesis of Fullerene Bis-Adducts. Angew. Chem. Int. Ed. 2021, 60, 313. [Google Scholar] [CrossRef] [PubMed]
  35. Hossain, M.S.; Rahaman, S.A.; Hatai, J.; Saha, M.; Bandyopadhyay, S. Switching the recognition ability of a photoswitchable receptor towards phosphorylated anions. Chem. Commun. 2020, 56, 4172–4175. [Google Scholar] [CrossRef]
  36. Mutlu, H.; Barner-Kowollik, C. Green chain-shattering polymers based on a self-immolative azobenzene motif. Polym. Chem. 2016, 7, 2272–2279. [Google Scholar] [CrossRef] [Green Version]
  37. Kumar, M.; Sharma, U.; Sharma, S.; Kumar, V.; Singh, B.; Kumar, N. Catalyst-free water mediated reduction of nitroarenes using glucose as a hydrogen source. RSC Adv. 2013, 3, 4894–4898. [Google Scholar] [CrossRef]
Scheme 1. Examples of preparation hydroxyalkyl−containing azoxybenzenes.
Scheme 1. Examples of preparation hydroxyalkyl−containing azoxybenzenes.
Molbank 2022 m1384 sch001
Scheme 2. Reaction of nitrobenzyl alcohol 2ac with glucose.
Scheme 2. Reaction of nitrobenzyl alcohol 2ac with glucose.
Molbank 2022 m1384 sch002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Spesivaya, E.S.; Lupanova, I.A.; Konshina, D.N.; Konshin, V.V. A Direct Method for the Efficient Synthesis of Hydroxyalkyl-Containing Azoxybenzenes. Molbank 2022, 2022, M1384. https://doi.org/10.3390/M1384

AMA Style

Spesivaya ES, Lupanova IA, Konshina DN, Konshin VV. A Direct Method for the Efficient Synthesis of Hydroxyalkyl-Containing Azoxybenzenes. Molbank. 2022; 2022(2):M1384. https://doi.org/10.3390/M1384

Chicago/Turabian Style

Spesivaya, Ekaterina S., Ida A. Lupanova, Dzhamilya N. Konshina, and Valery V. Konshin. 2022. "A Direct Method for the Efficient Synthesis of Hydroxyalkyl-Containing Azoxybenzenes" Molbank 2022, no. 2: M1384. https://doi.org/10.3390/M1384

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