ArticlePDF Available

N,N'-(1,2-Phenylene)-bis[4-(azidomethyl)benzamide]

Authors:

Abstract

The synthesis of N,N''-(1,2-phenylene)-bis[4-(azidomethyl)benzamide] (2) by direct nucleophilic disubstitution of the suitable dihalogen precursor 1 with NaN3 is reported. The structure of the title compound was fully characterized by FT-IR, 1H NMR, 13C NMR, EI-MS, elemental analysis and melting point determination.
Molbank 2010, M707; doi:10.3390/M707
molbank
ISSN 1422-8599
www.mdpi.com/journal/molbank
Short Note
N,N'-(1,2-Phenylene)-bis[4-(azidomethyl)benzamide]
rgen Bachl 1 and David D. Díaz 1,2,*
1 Institut für Organische Chemie, Universität Regensburg, Universitätsstr. 31, 93040 Regensburg,
Germany
2 ICMA, CSIC-Universidad de Zaragoza, Pedro Cerbuna 12, 50009 Zaragoza, Spain
* Author to whom correspondence should be addressed;
E-Mail: David.Diaz@chemie.uni-regensburg.de
Received: 8 October 2010 / Accepted: 16 November 2010 / Published: 18 November 2010
Abstract: The synthesis of N,N''-(1,2-phenylene)-bis[4-(azidomethyl)benzamide] (2) by
direct nucleophilic disubstitution of the suitable dihalogen precursor 1 with NaN3 is
reported. The structure of the title compound was fully characterized by FT-IR, 1H NMR,
13C NMR, EI-MS, elemental analysis and melting point determination.
Keywords: o-phenylenediamine; azide; benzamide; nucleophilic substitution
Organic azides (R-N3) are an important class of energy-rich and versatile intermediates for organic
synthesis [1] that have drawn great interest since the late 19th century [2]. They currently engage a key
interdisciplinary position at the interfaces between chemistry, biomedicine and materials science [3].
The industrial interest in these compounds derives from their use as precursors for the synthesis of
amines [3,4], nitrenes [5], heterocycles such as triazoles and tetrazoles [3,6], as latent amino groups in
the synthesis of natural products [3], as detonators [7], blowing agents for polymeric foams [3], and
functional groups in pharmaceuticals as illustrated by azidonucleosides in the AIDS treatment [8] and
their bioconjugation via Staudinger ligation [9]. Moreover, in the last decade organic azides have
emerged as a key component of the powerful ‘click’ chemistry [10]. In particular, polyvalent azides are
valued cross-linking agents and monomers in materials science [11,12].
Herein, we report the synthesis of N,N'-(1,2-phenylene)-bis[4-(azidomethyl)benzamide] (2). This
compound represents a versatile building block for the synthesis of a library of 3,1,5-
benzoxadiazepines with potential bioactivities that could lead to potential new drug candidates [13].
The preparation of 2 was carried out by double nucleophilic substitution of the suitable dihalogen
OPEN ACCESS
Molbank 2010
M707 (Page 2)
precursor 1 with NaN3 in DMSO at 100 °C for 24 h (Scheme 1). Dihalogen precursor 1 was prepared
by double nucleophilic addition of o-phenylenediamine on 4-(chloromethyl)benzoyl chloride as
reported previously [14].
Scheme 1.
Experimental Section
General
1H and 13C NMR spectra were recorded at 25 °C on a Bruker Avance 300 spectrometer in CDCl3 as
solvent, and chemical shifts are reported relative to Me4Si (
= 0) [15]. The low-resolution mass
spectrum was obtained by using a Varian MAT 311A spectrometer. Elemental analysis was performed
on a Heraeus Mikro-Rapid analyzer. The infrared spectrum was recorded using a Diamond ATR
(attenuated total reflection) accessory (Golden Gate) on a Bio-Rad Excalibur FTS 3000 MX
spectrophotometer. The melting point (mp) was measured in a chi 510 and is uncorrected. Thin-
layer chromatography was carried out on Merck aluminium sheets coated with silica gel
60 F254. Compounds were visualized by use of 254 nm UV light and/or iodine as staining reagent. All
solvents were of p.a. grade or purified by standard techniques [16]. Anhydrous sodium sulfate was
used for drying solutions.
Safety Note
Although the title compound was found to be a stable solid (C/N = 2.75) at room temperature,
highly energy-rich organic azides (C/N < 1) should never be isolated and they should be stored away
from sources of heat, light, pressure, and shock. Azide-containing reaction mixtures should not be
concentrated through rotary evaporation or distillation.
Synthesis of N,N'-(1,2-phenylene)-bis[4-(azidomethyl)benzamide] (2)
To a 50 mL round-bottomed flask equipped with a stirring bar and condenser were added N,N'-(1,2-
phenylene)-bis[4-(chloromethyl)benzamide] (1) (500 mg, 1.21 mmol), sodium azide (236 mg,
3.63 mmol), and DMSO (10 mL). The mixture was heated at 100 °C for 24 h, cooled, and water
(30 mL) and brine (10 mL) were added. The mixture was extracted five times with EtOAc, and the
combined organic phases were washed with brine, dried (Na2SO4), filtered, and concentrated. The final
Molbank 2010
M707 (Page 3)
traces of solvent were removed under vacuum to yield 2 (417 mg, 81%) as a white solid: TLC Rf
(AcOEt/hexane 1:1) 0.44; m.p. = 116118 °C; 1H NMR (300 MHz, CDCl3) /ppm = 4.46 (s, 2H),
6.846.47 (m, 1H), 7.347.26 (m, 1H), 7.48 (d, J = 8.4 Hz, 2H), 8.04 (d, J = 8.3 Hz, 2H), 9.61 (s, 1H);
13C NMR (75 MHz, CDCl3) /ppm = 54.28, 125.88, 126.17, 128.29, 128.32, 130.58, 133.30, 139.68,
166.06; FT-IR (ATR) max (cm-1) 3225 (N-H stretching), 2086 (N3 asymmetric stretching), 1646 (C=O
stretching, amide I band), 1505 (N-H bending, amide II band), 1252 (N3 symmetric stretching); MS
(ESI) m/z 427 [MH+]. Elemental analysis calculated for C22H18N8O2: C, 61.96; H, 4.25; N, 26.28;
found: C, 62.09; H, 4.69; N, 26.89.
Acknowledgements
D.D.D. thanks the Alexander von Humboldt Foundation for a research fellowship for experienced
researchers.
References and Notes
1. Scriven, E.F.V.; Tumbull, K. Azides: Their preparation and synthetic uses. Chem. Rev. 1988, 88,
297-368.
2. Curtius, T. Ueber Stickstoffwasserstoffsäure (Azoimid) N3H. Ber. Dtsch. Chem. Ges. 1890, 23,
3023-3033.
3. Bräse, S.; Gil, C.; Knepper, K.; Zimmermann, V. Organic azides: An exploding diversity of a
unique class of compounds. Angew. Chem. Int. Ed. 2005, 44, 5188-5240.
4. Chandrasekhar, S.; Prakash, S.Y.; Rao, C.L. Poly(ethylene glycol) (400) as superior solvent
medium against ionic liquids for catalytic hydrogenations with PtO2. J. Org. Chem. 2006, 71,
2196-2199.
5. Rao, H.S.P.; Siva, P. Facile reduction of azides with sodium borohydride/copper(II) sulfate
system. Synth. Commun. 1994, 24, 549-555.
6. Kolb, H.C.; Sharpless, K.B. The growing impact of click chemistry on drug discovery. Drug
Discov. Today 2003, 8, 1128-1137.
7. Agrawal, J.P.; Hodgson, R. Organic Chemistry of Explosives, 1st ed.; John Wiley & Sons, Inc.:
Chichester, UK, 2006.
8. Lin, T.S.; Prusoff, W.H. Synthesis and biological activity of several amino analogues of
thymidine. J. Med. Chem. 1978, 21, 109-112.
9. Kohn, M.; Breinbauer, R. The Staudinger ligation- A gift to chemical biology. Angew. Chem. Int.
Ed. 2004, 43, 3106-3116.
10. Kolb, H.C.; Finn, M.G.; Sharpless, K.B. Click chemistry: Diverse chemical function from a few
good reactions. Angew. Chem. Int. Ed. 2001, 40, 2004-2021.
11. Binder, W.H.; Sachsenhofer, R. ‘Click’ chemistry in polymer and materials science. Macromol.
Rapid Commun. 2007, 28, 15-54.
12. az, D.D.; Punna, S.; Holzer, P.; McPherson, A.K.; Sharpless, K.B.; Fokin, V.V.; Finn, M.G.
Click chemistry in materials synthesis. 1. Adhesive polymers from copper-catalyzed azide-alkyne
cycloaddition. J. Polym. Sci. Polym. Chem. 2004, 42, 4392-4403.
Molbank 2010
M707 (Page 4)
13. Mazurkiewicz, R. Novel synthesis and rearrangement of 3,l,5-benzoxadiazepines. Monatsh.
Chem. 1988, 119, 1279-1287.
14. Bachl, J.; az, D.D. N,N'-1,2-Phenylenebis[4-(chloromethyl)benzamide]. Molbank 2010, 2010,
M705.
15. See Supplementary Files.
16. Armarego, W.L.F.; Perrin, D.D. Purification of Laboratory Chemicals, 4th ed.; Butterworth-
Heinemann: Oxford, UK, 1996.
© 2010 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article
distributed under the terms and conditions of the Creative Commons Attribution license
(http://creativecommons.org/licenses/by/3.0/).
ResearchGate has not been able to resolve any citations for this publication.
Book
routes used for the synthesis of organic explosives in a single volume. Assuming no prior knowledge, the book discusses everything from the simplest mixed acid nitration of toluene, to the complex synthesis of highly energetic caged nitro compounds. Reviews laboratory and industrial methods, which can be used to introduce aliphatic C-nitro, aromatic C-nitro, N-nitro, and nitrate ester functionality into organic compounds. Discusses the advantages and disadvantages of each synthetic method or route, with scope, limitations, substrate compatibility and other important considerations. Features numerous examples in the form of text, reaction diagrams, and tables.
Article
Sodium borohydride/copper (II) sulphate reduces alkyl and aryl azides to primary amines and aroyl azides to amides under mild conditions.
Article
Aus dem 2′-Desoxy-trityl-lyxosyl-thymin (I) wird über 4 Zwischenstufen das Diazid (III) hergestellt und zum Diamin (IV) hydriert.
Article
Examination of nature's favorite molecules reveals a striking preference for making carbon–heteroatom bonds over carbon–carbon bonds—surely no surprise given that carbon dioxide is nature's starting material and that most reactions are performed in water. Nucleic acids, proteins, and polysaccharides are condensation polymers of small subunits stitched together by carbon–heteroatom bonds. Even the 35 or so building blocks from which these crucial molecules are made each contain, at most, six contiguous C−C bonds, except for the three aromatic amino acids. Taking our cue from nature's approach, we address here the development of a set of powerful, highly reliable, and selective reactions for the rapid synthesis of useful new compounds and combinatorial libraries through heteroatom links (C−X−C), an approach we call “click chemistry”. Click chemistry is at once defined, enabled, and constrained by a handful of nearly perfect “spring-loaded” reactions. The stringent criteria for a process to earn click chemistry status are described along with examples of the molecular frameworks that are easily made using this spartan, but powerful, synthetic strategy.
Article
The copper(I)-catalyzed cycloaddition reaction between azides and alkynes has been employed to make metal-adhesive materials. Copper and brass surfaces supply the necessary catalytic Cu ions, and thus the polymerization process occurs selectively on these metals in the absence of added catalysts. Alternatively, copper compounds can be added to monomer mixtures and then introduced to reducing metal surfaces such as zinc to initiate polymerization. The resulting materials were found to possess comparable or superior adhesive strength to standard commercial glues, and structure-activity correlations have identified several important properties of the monomers. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 4392- 4403, 2004
Article
The modification of polymers after the successful achievement of a polymerization process represents an important task in macromolecular science. Cycloaddition reactions, among them the metal catalyzed azide/alkyne ‘click’ reaction (a variation of the Huisgen 1,3‐dipolar cycloaddition reaction between terminal acetylenes and azides) represents an important contribution towards this endeavor. They combine high efficiency (usually above 95%) with a high tolerance of functional groups and solvents, as well as moderate reaction temperatures (25–70 °C). The present review assembles recent literature for applications of this reaction in the field of polymer science (linear polymers, dendrimers, gels) as well as the use of this and related reactions for surface modification on carbon nanotubes, fullerenes, and on solid substrates, and includes the authors own publications in this field. A number of references (>100) are included. magnified image