Next Article in Journal
Conjugate Addition of Nucleophiles to the Vinyl Function of 2-Chloro-4-vinylpyrimidine Derivatives
Previous Article in Journal
New Cytotoxic Azaphilones from Monascus purpureus-Fermented Rice (Red Yeast Rice)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Preparation of the Pyridinium Salts Differing in the Length of the N-Alkyl Substituent

by
Jan Marek
1,2,
Petr Stodulka
1,
Jiri Cabal
1,
Ondrej Soukup
1,
Miroslav Pohanka
1,
Jan Korabecny
3,
Kamil Musilek
1 and
Kamil Kuca
1,*
1
Faculty of Military Health Science, University of Defence, Hradec Kralove, 50001, Czech Republic
2
Vakos XT, Prague, 18600, Czech Republic
3
Faculty of Pharmacy in Hradec Kralove, Charles University in Prague, Hradec Kralove 50001, Czech Republic
*
Author to whom correspondence should be addressed.
Molecules 2010, 15(3), 1967-1972; https://doi.org/10.3390/molecules15031967
Submission received: 14 December 2009 / Revised: 4 March 2010 / Accepted: 10 March 2010 / Published: 19 March 2010

Abstract

:
Quaternary pyridinium salts with chains ranging from C8 to C20 belong in the large group of cationic surfactants. In this paper, the preparation of such cationic surface active agents based on the pyridinium moiety and differing in the length of the N-alkyl chain is described. Additionally, HPLC technique was established to distinguish each prepared pyridinium analogue. This study represents universal method for preparation and identification of quaternary pyridinium detergents.

1. Introduction

Cationic surfactants offer a huge amount of applications for industry. They are frequently used in analytical and physical chemistry. Besides, some pyridinium salts are used as disinfectants in various medical applications (e.g., eye drops, solutions, foams). In this field, quaternary pyridinium salts are compounds of high interest, on account of their similarity to benzalkonium salts [1]. Their surfactant characteristics (e.g., critical micellar concentration, surface tension) have already been determined and published everal times [2]. Some pyridinium salts (C12 and C16) were also used as solubilizers for water insoluble compounds in Analytical Chemistry. Differently, such surfactants were employed in the complexes with dyes to determine the type and concentration of electrolytes by a spectrophotometric method [3]. A method describing spray application of a compound with a C16 side chain on the pyridinium ring (cetylpyridinium, CP) for protection of the poultry against bacterial contamination has also been published [4]. Another application of CP consisted in its addition to chewing gums as an anti-plaque agent [5]. The same application may be possible in the dentistry [6]. The surface activity of CP predicted its ability to form the micelles (Figure 1). This property was used by method called micellar-enhanced ultrafiltration that was applied for the removal of the pollutants (e.g., arsenic, perchlorate) from ground waters and also for the removal of the heavy metals from solutions [7,8]. The usage of pyridinium detergents as carriers improving adsorption of the oligonucleotides at the phospholipid membrane was also described [9,10]. The pyridinium analogues with the alkyl chains longer than C10 are known as nicotinic receptor antagonists [11]. Some pyridinium salts bearing shorter alkyl chains were prepared and tested as cholinesterase inhibitors [12].
Preparation of the different pyridinium salts was already described before [13,14]. However, there has been no description of the synthesis of the whole series of such salts differing in the alkyl chains (C8 to C20). Formerly, a similar method for preparation of the benzalkonium salts was reported [15]. Consequently, a universal method for preparation of monoquaternary pyridinium salts with chain substituents was developed. In this study, this method was successfully applied for the preparation of the described pyridinium salts.

2. Results and Discussion

The results regarding the synthesis and analysis of the newly prepared compounds are summarized in Table 1. Yield, melting point and HPLC retention time were determined. The preparation of salts with C12-C20 alkyl chains was found to be easier than in the case of C8-C10 ones and the final yields were higher. White crystaline products were obtained. Satisfactory purity was obtained after one (C12-C16) or several crystallizations (C8-C10, C18-C20) from ether suspensions. The compounds with C18-C20 alkyl chains were contaminated with starting material that was detected by TLC. Hence, they were crystallized from acetone. There were also some difficulties in the preparation of C8-C10 pyridinium salts. These compounds were crystallized several times from ether before the crystals were formed. After all these crystallizations, the yield was dramatically decreased.
Additionally, the HPLC analysis was carried out for individual compounds. A newly developed HPLC method was able to distinguish among all prepared quaternary pyridinium salts (Figure 2). The shortest retention time was found for the C8 pyridinium salt. This novel HPLC method could be easily used for characterization of mixtures of pyridinium detergents.

3. Experimental

3.1. General

The reactions were monitored by TLC (Kieselgel Merck; mobile phase chloroform/methanol 100/1; detection UV 254, Dragendorff reagent). The yields (%) and melting points (Boetius, m.p. were uncorrected) are summarized in Table 1. 1H-NMR spectra characterizing individual compounds were recorded in DMSO-d6 on a Varian Gemini 300 instrument.

3.2. Synthesis

A universal method for the preparation of monoquaternary pyridinium salts 3-9 was developed (Scheme 1): pure pyridine (1; 1 eq) in dry ethanol was mixed with 1-bromoalkane (2; 1.4 eq). The mixture was refluxed for 40 hours. The solution was evaporated under reduced pressure and the crude oily product was crystallized from ether, filtered under reduced pressure, washed with ether and allowed to dry at r.t..
N-Octylpyridinium bromide (3): 1H-NMR ppm 9.12 (d, J = 5.7 Hz, 2H), 8.61 (t, J = 7.6 Hz, 1H), 8.22-8.11 (m, 2H), 4.58 (t, J = 7.5 Hz, 2H), 1.98-1.80 (m, 2H), 1.22 (m, 12H), 0.82 (t, J = 6.6 Hz, 3H).
N-Decylpyridinium bromide (4): 1H-NMR ppm 9.14 (d, J = 5.63 Hz. 2H), 8.61 (t, J = 7.58 Hz, 1H), 8.22-8.11 (m, 2H), 4.60 (t, J = 7.46 Hz, 2H), 2.01-1.80 (m, 2H), 1.16 (m, 16H), 0.83 (t, J = 6.64 Hz, 3H).
N-Dodecylpyridinium bromide (5): 1H-NMR ppm 9.14 (d, J = 5.6 Hz. 2H), 8.61 (t, J = 7.8 Hz, 1H), 8.22-8.11 (m, 2H), 4.61 (t, J = 7.5 Hz, 2H), 1.98-1.80 (m, 2H), 1.23 (m, 20H), 0.79 (t, J = 6.3 Hz, 3H).
N-Tetradecylpyridinium bromide (6): 1H-NMR ppm 9.14 (d, J = 5.6 Hz, 2H), 8.60 (t, J = 7.8 Hz, 1H), 8.22-8.11 (m, 2H), 4.61 (t, J = 7. 5 Hz, 2H), 1.98-1.80 (m, 2H), 1.23 (m, 24H), 0.83 (t, J = 6.6 Hz, 3H).
N-Hexadecylpyridinium bromide (7): 1H-NMR ppm 9.13 (d, J = 5.7 Hz, 2H), 8.61 (s, 1H), 8.22-8.11 (m, 2H), 4.61 (t, J = 7.4 Hz, 2H), 2.03-1.80 (m, 2H), 1.21 (m, 28H), 0.83 (t, J = 6.5 Hz, 3H).
N-Octadecylpyridinium bromide (8): 1H-NMR ppm 9.12 (d, J = 5.4 Hz, 2H), 8.61 (s, 1H), 8.22-8.11 (m,, 2H), 4.60 (t, J = 7.3 Hz, 2H) 1.89-1.80 (m, 2H), 1.22 (m, 32H) 0.82 (d, J = 6.3 Hz. 3H).
N-Eicosylpyridinium bromide (9): 1H-NMR ppm 9.14 (d, J = 5.63 Hz, 2H), 8.60 (t, J = 7.8 Hz, 1H), 8.23-8.11 (m, 2H), 4.61 (t, J = 7.5 Hz, 2H), 1.98-1.80 (m, 2H), 1.23 (m. 36H), 0.83 (t. J = 6.6 Hz, 3H).

3.3. HPLC Analysis

A suitable method for the characterization of the newly prepared compounds in the mixture was developed using HPLC. The HPLC system consisted of a P200 gradient pump (Spectra-Physics Analytical, Fremont, CA, USA), a 7125 injection valve – 10 µL loop (Rheodyne, Cotati, CA, USA), an UV1000 detector (Spectra-Physics Analytical) and a CSW Chromatography Station 1.5 software (DataApex, Praha, Czech Republic). A 250 × 4.6 mm I.D. Waters Spherisorb cyano (5 µm) column was used for analysis (Supelco Inc., Bellefonte, WA, USA). The mobile phase consists of 45% acetonitrile and 55% water and was prepared as 0.1 M sodium acetate solution. Finally the pH was adjusted via acetic acid to 5.0. The column was rinsed isocratically with a flow-rate of 1 mL/min. The absorbance was measured at 257 nm.

4. Conclusions

The preparation of pyridinium salts with C8-C20 N-alkyl chains was described. Such preparation might be used not only under laboratory conditions, but also for semi-industrial purposes. It is anticipated that the synthesis wil prove very useful for the next effort of preparing micelle forming compounds. The developed HPLC experimental protocol was found valuable for the estimation of the purity of the newly prepared compounds.

Acknowledgements

This work was supported by the project of Ministry of Defence of the Czech Republic No. OVUOFVZ200803.

References and Notes

  1. Kuca, K.; Dohnal, V.; Bielavska, M.; Cabal, J. Determination of benzalkonium bromide homologues in disinfection products using high-performance liquid chromatography. Anal. Lett. 2005, 38, 673–682. [Google Scholar] [CrossRef]
  2. Semmler, A.; Kohler, H.H. Surface Properties of Alkylpyridinium Chlorides and the Applicability of the Pendant Drop Technique. J. Coll. Interf. Sci. 1999, 218, 137–144. [Google Scholar] [CrossRef] [PubMed]
  3. Akbas, H.; Kartal, C.I. Reactive Orange 16-dodecylpyridinium chloride interactions in electrolytic solutions. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2006, 65, 95–99. [Google Scholar] [CrossRef] [PubMed]
  4. Rodriguez-Morales, S.; Compadre, R.L.; Castillo, R.; Breen, P.J.; Compadre, C.M. 3D-QSAR, synthesis, and antimicrobial activity of 1-alkylpyridinium compounds as potential agents to improve food safety. Eur. J. Med. Chem. 2005, 40, 840–849. [Google Scholar] [CrossRef] [PubMed]
  5. Faraj, J.A.; Dorati, R.; Schoubben, A.; Worthen, D.; Selmin, F.; Capan, Y.; Leung, K.; DeLuca, P.P. Development of a peptide-containing chewing gum as a sustained release antiplaque antimicrobial delivery system. AAPS Pharm. Sci. Tech. 2007, 8, 26–29. [Google Scholar] [CrossRef] [PubMed]
  6. Tanzer, J.M.; Slee, A.M.; Kamay, B.; Scheer, E.R. In vitro evaluation of seven cationic detergents as antiplaque agents. Antimicrob. Agents Chemother. 1979, 15, 408–414. [Google Scholar] [CrossRef] [PubMed]
  7. Iqbal, J.; Kim, H.J.; Yang, J.S.; Baek, K.; Yang, J.W. Removal of arsenic from groundwater by micellar-enhanced ultrafiltration (MEUF). Chemosphere 2007, 66, 970–976. [Google Scholar] [CrossRef] [PubMed]
  8. Parette, R.; Cannon, F.S. The removal of perchlorate from groundwater by activated carbon tailored with cationic surfactants. Water Res. 2005, 39, 4020–4028. [Google Scholar] [CrossRef] [PubMed]
  9. Carlsson, S.; Kontturi, A.K.; Kontturi, K. Improving membrane activity of oligonucleotides by cetylpyridinium chloride: an electrochemical study. Eur. J. Pharm. Sci. 2006, 29, 451–459. [Google Scholar] [CrossRef] [PubMed]
  10. Bordbar, A.K.; Taheri-Kafrani, A. Binding of cetylpyridinum chloride to glucose oxidase. Coll. Surf. B, Biointer. 2007, 55, 84–89. [Google Scholar] [CrossRef] [PubMed]
  11. Grinevich, V.P.; Crooks, P.A.; Sumithran, S.P.; Haubner, A.J.; Ayers, J.T.; Dwoskin, L.P. N-n-alkylpyridinium analogs, a novel class of nicotinic receptor antagonists: selective inhibition of nicotine-evoked [3H] dopamine overflow from superfused rat striatal slices. J. Pharmacol. Exp. Ther. 2003, 306, 1011–1020. [Google Scholar] [CrossRef] [PubMed]
  12. Kuca, K.; Cabal, J.; Patocka, J.; Dohnal, V. Quaternary Heteroarenium Salts as the Competitive Inhibitors of the Brain Acetylcholinesterase. Lett. Drug Des. Discov. 2004, 4, 97–100. [Google Scholar] [CrossRef]
  13. Lang, J. Surfactant Aggregation Number and Polydispersity of Sds+1-Pentanol Mixed Micelles in Brine Determined by Time-Resolved Fluorescence Quenching. J. Phys. Chem. 1990, 94, 3734–3739. [Google Scholar] [CrossRef]
  14. Hashimoto, S.; Thomas, J.K. Fluorescence Study of Pyrene and Naphthalene in Cyclodextrin Amphiphile Complex-Systems. J. Am. Chem. Soc. 1985, 107, 4655–4662. [Google Scholar] [CrossRef]
  15. Nakamura, T.; Thomas, J.K. The Interaction of Alkylammonium Salts with synthetic clays. A fluorescence and laser excitation study. J. Phys. Chem. 1986, 90, 641–644. [Google Scholar] [CrossRef]
Sample Availability: Samples of all prepared compounds are available from the authors.
Figure 1. Forming of the micelles.
Figure 1. Forming of the micelles.
Molecules 15 01967 g001
Figure 2. HPLC chromatogram of the prepared pyridinium series mixture.
Figure 2. HPLC chromatogram of the prepared pyridinium series mixture.
Molecules 15 01967 g002
Scheme 1. Preparation of pyridinium salts.
Scheme 1. Preparation of pyridinium salts.
Molecules 15 01967 sch001
Table 1. Yield, melting point and retention time of prepared pyridinium salts.
Table 1. Yield, melting point and retention time of prepared pyridinium salts.
CompoundSide alkylating chain (R)Yield (%)m.p. (°C)HPLC Rt (min)
3C81328–303.98
4C104330–324.58
5C128273–755.27
6C147658–606.06
7C1699636.98
8C189565–678.05
9C207973–759.31

Share and Cite

MDPI and ACS Style

Marek, J.; Stodulka, P.; Cabal, J.; Soukup, O.; Pohanka, M.; Korabecny, J.; Musilek, K.; Kuca, K. Preparation of the Pyridinium Salts Differing in the Length of the N-Alkyl Substituent. Molecules 2010, 15, 1967-1972. https://doi.org/10.3390/molecules15031967

AMA Style

Marek J, Stodulka P, Cabal J, Soukup O, Pohanka M, Korabecny J, Musilek K, Kuca K. Preparation of the Pyridinium Salts Differing in the Length of the N-Alkyl Substituent. Molecules. 2010; 15(3):1967-1972. https://doi.org/10.3390/molecules15031967

Chicago/Turabian Style

Marek, Jan, Petr Stodulka, Jiri Cabal, Ondrej Soukup, Miroslav Pohanka, Jan Korabecny, Kamil Musilek, and Kamil Kuca. 2010. "Preparation of the Pyridinium Salts Differing in the Length of the N-Alkyl Substituent" Molecules 15, no. 3: 1967-1972. https://doi.org/10.3390/molecules15031967

Article Metrics

Back to TopTop