Next Article in Journal
A Facile Synthesis of 3-Substituted 9H-Pyrido[3,4-b]indol-1(2H)-one Derivatives from 3-Substituted β-Carbolines
Previous Article in Journal
Emerging Biological Importance of Central Nervous System Lanthionines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of Cyclic Quaternary Ammonium Bromides by B3LYP Calculations, NMR and FTIR Spectroscopies

Laboratory of Microbiocides Chemistry, Faculty of Chemistry, A. Mickiewicz University, Grunwaldzka 6, 60-780 Poznań, Poland
*
Author to whom correspondence should be addressed.
Molecules 2010, 15(8), 5644-5657; https://doi.org/10.3390/molecules15085644
Submission received: 26 July 2010 / Revised: 11 August 2010 / Accepted: 13 August 2010 / Published: 16 August 2010
(This article belongs to the Section Organic Chemistry)

Abstract

:
N,N-dioctyl-azepanium, -piperidinium and -pyrrolidinium bromides 1-3, have been obtained and characterized by FTIR and NMR spectroscopy. DFT calculations have also been carried out. The optimized bond lengths, bond angles and torsion angles calculated by B3LYP/6-31G(d,p) approach have been presented. Both FTIR and Raman spectra of 1-3 are consistent with the calculated structures in the gas phase. The screening constants for 13C and 1H atoms have been calculated by the GIAO/B3LYP/6-31G(d,p) approach and analyzed. Linear correlations between the experimental 1H and 13C chemical shifts and the computed screening constants confirm the optimized geometry.

1. Introduction

Quaternary ammonium compounds (QACs) were introduced as antimicrobial agents by Domagk over seventy years ago [1]. The first generation of QACs were standard benzalkonium chlorides, i.e. alkylbenzyldimethylammonium chloride, with specific alkyl distributions, i.e., C12, 40%; C14, 50% and C16, 10% [2]. The second generation of QACs was obtained by substitution of the aromatic ring in alkylbenzyldimethylammonium chlorides by chlorine or alkyl groups to get products like alkyldimethylethylbenzylammonium chloride with C12, 50%; C14, 30%; C16, 17% and C18, 3% alkyl distribution. Dual quaternary ammonium salts are the third generation of QACs. These products are a mixture of equal proportions of alkyldimethylbenzylammonium chloride with alkyl distribution C12, 68%; C14, 32% and alkyldimethylethylbenzylammonium chloride with alkyl distribution C12, 50%; C14, 30%; C16, 17% and C18, 3%. The twin chain quaternary ammonium salts, like didecyldimethyl-ammonium chloride are the fourth generation of QACs. The concept of synergistic combinations of dual QACs has been applied to twin chain quaternary ammonium salts. The mixture of dialkyldimethylamoonium chloride (dioctyl, 25%; didecyl, 25%, octyldecyl, 50%) with benzalkonium chloride (C12, 40%; C14, 50%; C16, 10%) is the newest blend of quaternary ammonium salts which represents the fifth generation of QACs [2]. Because of the increasing resistance of microorganisms to commonly used disinfectants, the synthesis of new types of microbiocides is very important. One of the new groups with good antimicrobial activity are the cyclic quaternary ammonium salts [3]. The aim of this work was the synthesis of cyclic N,N-dioctyl quaternary ammonium salts, i.e. N,N-dioctyl-azepanium, N,N-dioctylpiperidinium and N,N-dioctylpyrrolidinium bromides, with potential antimicrobial activity. Some cyclic quaternary ammonium salts have previously been obtained by intramolecular cyclisation of amine derivatives [4,5,6,7,8,9]. Another way, i.e. reaction of alkyl halides with cyclic amines, can lead to chiral cyclic quaternary ammonium salts [10].
In recent years numbers of applications of the quaternary ammonium salts has been continuously increasing. They are used as biocides [11,12,13,14,15], and phase-transfer catalysts, especially in enantioselective reactions [16,17,18,19,20,21]. Pyrrolidinium salts are analogues of oxotremorine and are used as muscarinic agonists [5]. Some quaternary ammonium salts exist as ionic liquids, which can be used as “green solvents” [22,23,24,25,26] and electrolytes for liquid batteries [27,28].
The molecular structures of N,N-dioctyl-azepanium (1), -piperidinium (2) and -pyrrolidinium (3) bromides analyzed by FTIR and NMR spectroscopy and B3LYP calculations are presented in this paper. The above compounds belong to the cyclic quaternary ammonium bromide family investigated in our laboratory in order to better understand the mechanism of their biological activity.

2. Results and Discusion

2.1. Synthesis

N,N-dioctyl-azepanium, -piperidinium and -pyrrolidinium bromides 1-3 were obtained by reaction of N,N-dioctylamine with dibromohexane, dibromopentane and dibromobutane, respectively. The reaction of secondary amines with 1,5-dichloropentane and 1,4-dichlorobutane to produce dialkylpiperidinium and dialkylpyrrolidinium salts has previously been described by Ericsson and Keps [4]. In our work, using dibromoalkanes instead of dichloroalkanes, we formed five-, six- and seven-membered ammonium compounds in much higher yields and after shorter reaction times. In the first step of reaction of dioctylamine with α,ω-dibromoalkane, the halogenated tertiary amine is formed, which shows a strong tendency to form cyclic quaternary ammonium salts.

2.2. B3LYP Calculations

The structures and numbering for 1-3 are given in Figure 1. The structures optimized at the B3LYP/6-31G(d,p) level of theory are shown in Figure 2.
Figure 1. The structure and numbering for N,N-dioctylazepaniumbromide (1), N,N-dioctyl-piperidinium bromide (2) and N,N-dioctylpyrrolidinium bromide (3).
Figure 1. The structure and numbering for N,N-dioctylazepaniumbromide (1), N,N-dioctyl-piperidinium bromide (2) and N,N-dioctylpyrrolidinium bromide (3).
Molecules 15 05644 g001
Figure 2. Structures of (a) N,N-dioctylazepanium (1), (b) N,N-dioctylpiperidinium (2), (c) N,N-diocylpyrrolidinium (3) bromides optimized by the B3LYP/6-31G(d,p) method.
Figure 2. Structures of (a) N,N-dioctylazepanium (1), (b) N,N-dioctylpiperidinium (2), (c) N,N-diocylpyrrolidinium (3) bromides optimized by the B3LYP/6-31G(d,p) method.
Molecules 15 05644 g002
The computed B3LYP geometry parameters, energy and dipole moments are given in Table 1. The calculated energy for N,N-dioctylazepanium bromide (1) is about 1.2% lower than for N,N-dioctyl-piperidinium bromide (2) and 2.4% lower in comparison to N,N-dioctylpyrrolidinium bromide (3). The bromide anions in 1-3 are engaged in three non-linear weak intramolecular interactions with carbon atoms. Bromide anions additionally interact via Coulombic attractions with positively charged nitrogen atom. The N+(…)···Br- distances are 3.888 Å, 3.709 Å 3.674 Å, for 1, 2 and 3, respectively.
Table 1. Selected parameters of investigated molecules 1-3 estimated by B3LYP/6-31G(d,p) calculations.
Table 1. Selected parameters of investigated molecules 1-3 estimated by B3LYP/6-31G(d,p) calculations.
Parameters123
Energy (a.u)-3495.20808-3453.27811-3413.96044
Dipol moment (Debye)13.495111.409711.4657
Bond length (Å)
N+…Br-3.8883.7093.674
C(1)-H…Br-3.6363.5363.486
C(1’)-H…Br-3.686
C(4)-H…Br-3.5513.5703.616
C(4’)-H…Br- 3.360 3.346
N-C(1)1.535 1.538 1.532
N-C(1’)1.533 1.5141.513
N-C(4)1.548 1.542 1.529
N-C(4’)1.5311.551 1.543
Bond angle (o)
N-C(1)-C(2)119.5115.3106.2
N-C(1’)-C(2’)116.9114.2106.2
N-C(4)-C(5)117.9116.3115.6
N-C(4’)-C(5’)120.2119.9118.6
Dihedral angle (o)
N-C(1)-C(2)-C(3)-70.3-49.5
N-C(1’)-C(2’)-C(3’)88.6
N-C(1’)-C(2’)-C(3) 57.8
N-C(1)-C(2)-C(2’) -18.2
N-C(1’)-C(2’)-C(2) 25.2
N-C(4)-C(5)-C(6)-176.8-177.4-176.9
N-C(4’)-C(5’)-C(6’)-176.5-172.3-170.0

2.3. FTIR and Raman Spectra Study

Room-temperature solid-state FTIR and Raman spectra as well as the calculated spectra of 1 are shown in Figure 3.
Figure 3. Spectra of N,N-dioctylazepanium bromide (1);(a) FTIR, (b) Raman and (c) calculated spectra.
Figure 3. Spectra of N,N-dioctylazepanium bromide (1);(a) FTIR, (b) Raman and (c) calculated spectra.
Molecules 15 05644 g003
The observed and calculated harmonic frequencies and their tentative assignments are listed in Table 2. In general, the calculated frequency values with B3LYP 6-31G(d,p) basis set are close to experimental values of vibrational frequency.
Table 2. FTIR and Raman frequencies of N,N-dioctylazepanium bromide (1).
Table 2. FTIR and Raman frequencies of N,N-dioctylazepanium bromide (1).
RamanIRIR(calc.)INTProposed assignment
3437w νOH
2973m2956s301643.7νCH2
301362.3νCH2
301164.9νCH2
299923.3νCH2
298763.2νCH2
297418.6νCH2
2943112νCH2
2926s2925s293461.4νCH2
29196.4νCH2
2864s2856s2914200νCH2
2781vw νCH2
2727vw νCH2
2709vw2696vw νCH2
2669vw2670vw νCH2
1490vw1485m150121.9νCC
14814.7
1448w1468m14678.0νCC
14567.9
14522.3
1392w13961.5νCN
1377w13763.5νCN
13721.6
1358vw1360w13544.9νCC, βCH2
1349vw1338w13441.4βCH2
13212.8
1313vw1310w13082.7βCH2
12951.6
1280vw1277w12813.4νCC
1263vw1251vw12642.8νCC
12450,81
1217vw1218vw12050.63νCC
11861.3
1141vw1141vw11693.3νCN
1115vw1115vw11152.4νCN
1087vw1088w107515.3γCH2
1069vw1068vw10551.6γCH2
1048vw1047vw10293.8βCH2
1014vw1007w10142.9βCCC
960vw962w9972.6βCCC
930vw930vw9442.0βCCC
9339.7
865vw875w87813.7βCCC
846vw847vw8534.6βCCC
831vw832w βCCC
803vw800vw7882.5βCCC
767vw765vw74219.5βCCC
741vw738w βCCC
723w7144.0βCCC
706vw βCCC
659vw651vw6161.8βCNC
580vw578vw βNCC
542vw538vw βCCC
498vw499vw4993.8γCCC
403vw403vw4391.5γCCC
375vw 3461.3Lattice mode
360vw Lattice mode
330vw Lattice mode
303vw Lattice mode
288vw 2241.2Lattice mode
201vw 1234.1Lattice mode
86vw 910.59Lattice mode
512.5
The abbreviations used are: s, strong; m, medium; w, weak; vw, very weak; ν, stretching; β, in plane bending; δ, deformation; γ, out of plane bending; and τ, twisting.

2.4. 1H-NMR and 13C-NMR Spectra

The proton chemical shift assignments (Table 3-Table 5) are based on 2D COSY experiments, in which the proton-proton connectivity is observed through the off-diagonal peaks in the counter plot. The relations between the experimental 1H and 13C chemical shifts (δexp) and the GIAO (Gauge-Independent Atomic Orbitals) isotropic magnetic shielding (σcalc) for 1 is shown in Figure 4. Both correlations are linear, described by the relationship: δexp = a + b·σcalc. The parameters a and b are given in Table 3-Table 5. The very good correlation coefficients (r2=0.9379) for 1H and (r2=0.9984) for 13C confirm the optimized geometry of 1-3.
Figure 4. Plots of the experimental chemical shifts (δexp) vs the magnetic isotropic shielding (σcalc) from the GIAO/B3LYP/6-31G(d,p); N,N-dioctylazepanium bromide (1) δpred = a + b· σcalc. (a) carbon-13; (b) proton.
Figure 4. Plots of the experimental chemical shifts (δexp) vs the magnetic isotropic shielding (σcalc) from the GIAO/B3LYP/6-31G(d,p); N,N-dioctylazepanium bromide (1) δpred = a + b· σcalc. (a) carbon-13; (b) proton.
Molecules 15 05644 g004
Table 3. Chemical shifts (δ, ppm) in CDCl3 and calculated GIAO nuclear magnetic shielding (σcal) for N,N-dioctylazepanium bromide (1). The predicted GIAO chemical shifts were computed from the linear equation δexp= a + b·σcalc with a and b determined from the fit the experimental data.
Table 3. Chemical shifts (δ, ppm) in CDCl3 and calculated GIAO nuclear magnetic shielding (σcal) for N,N-dioctylazepanium bromide (1). The predicted GIAO chemical shifts were computed from the linear equation δexp= a + b·σcalc with a and b determined from the fit the experimental data.
δ expδcalcσcalc δexpδcalcσcalc
C(1)63.157.4118.0H(1)3.703.8527.44
C(2)22.223.3155.4H(2)2.011.6230.27
C(3)27.321.7157.1H(3)1.792.0629.72
C(4)61.364.6110.1H(4)3.453.2528.20
C(5)22.625.7152.7H(5)1.711.5930.32
C(6)26.427.1151.2H(6)1.271.2330.77
C(7)29.130.3147.7H(7)1.271.3330.65
C(8)29.030.3147.7H(8)1.271.3130.67
C(9)31.632.2145.6H(9)1.271.2730.72
C(10)22.624.1154.5H(10)1.271.3430.63
C(11)14.012.4167.3H(11)0.881.0531.00
a -0.9113 a -0.7865
b 164.9046 b 25.4318
r2 0.9622 r2 0.9609
Table 4. Chemical shifts (δ, ppm) in CDCl3 and calculated GIAO nuclear magnetic shielding (σcal) for N,N-dioctylpiperidinium bromide (2). The predicted GIAO chemical shifts were computed from the linear equation δexp= a + b·σcalc with a and b determined from the fit the experimental data.
Table 4. Chemical shifts (δ, ppm) in CDCl3 and calculated GIAO nuclear magnetic shielding (σcal) for N,N-dioctylpiperidinium bromide (2). The predicted GIAO chemical shifts were computed from the linear equation δexp= a + b·σcalc with a and b determined from the fit the experimental data.
δ expδcalcσcalc δexpδcalcσcalc
C(1)58.954.8129.6H(1)3.783.2827.90
C(2)20.020.5167.8H(2)1.901.6729.95
C(3)26.420.5167.8H(3)1.901.4730.23
C(4)58.161.0122.7H(4)3.463.8827.12
C(5)21.723.2161.8H(5)1.651.7729.84
C(6)22.525.7162.1H(6)1.271.4230.30
C(7)29.029.2158.1H(7)1.271.3330.41
C(8)28.929.2158.1H(8)1.271.3730.36
C(9)31.630.7156.5H(9)1.271.3030.45
C(10)20.623.1164.9H(10)1.271.3930.33
C(11)14.013.6175.5H(11)0.881.0230.81
a 170.9303 a 24.0232
b -0.8962 b -0.7758
r2 0.9640 r2 0.9168
Table 5. Chemical shifts (δ, ppm) in CDCl3 and calculated GIAO nuclear magnetic shielding (σcal) for N,N-dioctylpyrrolidinium bromide (3). The predicted GIAO chemical shifts were computed from the linear equation δexp= a + b·σcalc with a and b determined from the fit the experimental data.
Table 5. Chemical shifts (δ, ppm) in CDCl3 and calculated GIAO nuclear magnetic shielding (σcal) for N,N-dioctylpyrrolidinium bromide (3). The predicted GIAO chemical shifts were computed from the linear equation δexp= a + b·σcalc with a and b determined from the fit the experimental data.
δ expδcalcσcalc δexpδcalcσcalc
C(1)62.961.9126.0H(1)3.853.6927.57
C(2)21.818.6169.5H(2)2.311.7729.81
C(4)59.459.7128.2H(4)3.433.3228.00
C(5)23.424.2163.9H(5)1.702.4329.04
C(6)26.327.5160.6H(6)1.271.3030.49
C(7)29.029.2158.9H(7)1.271.2630.24
C(8)28.930.2157.8H(8)1.271.2530.41
C(9)31.531.1156.9H(9)1.271.4030.40
C(10)22.523.3164.8H(10)1.271.1830.35
C(11)14.012.5175.6H(11)0.880.9130.80
a 187.2433 a 27.4355
b -0.9949 b -0.8611
r2 0.9920 r2 0.9049
The correlation between the experimental chemical shifts and calculated isotropic screening constants are better for 13C atoms than for protons. The protons are located on the periphery of the molecule and thus are supposed to be more efficient in intermolecular (solute-solvent) effects than carbons. The differences between the exact values of the calculated and experimental shifts for protons are probably due to the fact that the shifts are calculated for single molecules in gas phase. For this reason the agreement between the experimental and the calculated data for proton is worse than for 13C.

3. Conclusions

N,N-dioctyl-azepanium, -piperidinium, -pyrrolidinium bromides 1-3 have been obtained by reaction of N,N-dioctylamine with dibromohexane, dibromopentane and dibromobutane, respectively. The structure of the investigated compounds has been analyzed by FTIR and NMR spectroscopy and B3LYP calculations. Both the FTIR and Raman spectra of 1-3 are consistent with the observed structures in the gas phase. The good correlations between the experimental 13C and 1H chemical shifts in D2O solution and GIAO/B3LYP/6-31G(d,p) calculated isotropic shielding tensors (δexp= a + b·σcalc) have confirmed the optimized geometry of 1-3.

4. Experimental

4.1. General

The NMR spectra were measured with a Varian Gemini 300VT spectrometer, operating at 300.07 and 75.4614 MHz for 1H and 13C, respectively. Typical conditions for the proton spectra were: pulse width 32o, acquisition time 5s, FT size 32 K and digital resolution 0.3 Hz per point, and for the carbon spectra pulse width 60o, FT size 60 K and digital resolution 0.6 Hz per point, the number of scans varied from 1200 to 10,000 per spectrum. The 13C and 1H chemical shifts were measured in CDCl3 relative to an internal standard of TMS. All proton and carbon-13 resonances were assigned by 1H (COSY) and 13C (HETCOR). All 2D NMR spectra were recorded at 298 K on a Bruker Avance DRX 600 spectrometer operating at the frequencies 600.315 MHz (1H) and 150.963 MHz (13C), and equipped with a 5 mm triple-resonance inverse probehead [1H/31P/BB] with a self-shielded z gradient coil (90o1H pulse width 9.0 μs and 13C pulse width 13.3 μs). Infrared spectra were recorded in the KBr pellets using a FT-IR Bruker IFS 66 spectrometer. The Raman spectrum was recorded on a Bruker IFS 66 spectrometer. The ESI (electron spray ionization) mass spectra were recorded on a Waters/Micromass (Manchester, UK) ZQ mass spectrometer equipped with a Harvard Apparatus syringe pump. The sample solutions were prepared in methanol at the concentration of approximately 10-5M. The standard ESI – MS mass spectra were recorded at the cone voltage 30V.

4.2. Computational Details

The calculations were performed using the Gaussian 03 program package [29] at the B3LYP [30,31] levels of theory with the 6-31G(d,p) basis set [30]. The NMR isotopic shielding constants were calculated using the standard GIAO (Gauge-Independent Atomic Orbital) approach [29,30,31,32] of GAUSSIAN 03 program package [33].

4.3. General procedure for the synthesis of N,N-dioctylcycloalkylammonium salts 1-3

Dioctylamine (5 g, 0.02 mol) was mixed with the appropriate dibromoalkane (0.02 mol) in the presence of anhydrous sodium carbonate (4.14 g, 0.04mol). The reaction mixture was heated under reflux for 15 h. The solvent was evaporated under reduced pressure and the residue was dried over P4O10 and then recrystallized from a suitable solvent, as indicated.
N,N-dioctylazepanium bromide (1). Prepared from 1,6-dibromohexane (5 g) and recrystallized from acetone/acetonitrile; yield: 65%, m.p. 212-214oC. Elemental analysis for C22H46NBr·H2O found (calc.) %C 62.80 (62.53); %H 11.49 (11.45); %N 3.30 (3.31); ES+MS m/z 325 (C22H46N); 1H-NMR (CDCl3): δ 3.70 (4H, t, C(1)H2, C(1’)H2), 2.01 (4H, m,C(2)H2, C(2’)H2), 1.79 (4H, m, C(3)H2, C(3’)H2), 3.45 (4H, t, C(4)H2, C(4’)H2), 1.71 (20H, m, C(5)H2, C(5’)H2, C(6)H2, C(6’)H2, C(7)H2, C(7’)H2, C(8)H2, C(8’)H2, C(9)H2, C(9’)H2), 1.27 (4H, m, C(10)H2, C(10’)H2), 0.88 (6H, t, C(11)H3, C(11’)H3); 13C-NMR (CDCl3): δ 63.1 (C(1), C(1’)), 22.2 (C(2), C(2’)), 27.3 (C(3), C(3’)), 61.3 (C(4), C(4’)), 22.6 (C(5), C(5’)), 26.4 (C(6), C(6’)), 29.1 (C(7), C(7’)), 29.0 (C(8), C(8’)), 31.6 (C(9), C(9’)), 22.6 (C(10), C(10’)), 14.0 (C(11), C(11’)).
N,N-dioctylpiperidinium bromide (2). From 1,5-dibromopentane (4.76 g, 0.02 mol ). Recrystallized from acetone; yield: 90%, m.p. 144-146oC. Elemental analysis for C21H44NBr found (calc) for %C 64.13 (64.59); %H 12.00 (11.36); %N 3.56 (3.59); ES+MS m/z 310 (C21H44N); 1H-NMR (CDCl3): δ 3.78 (4H, t, C(1)H2, C(1’)H2), 1.90 (6H, m, C(2)H2, C(2’)H2 C(3)H2), 3.46 (4H, t, C(4)H2, C(4’)H2), 1.65 (20H, m, C(5)H2, C(5’)H2, C(6)H2, C(6’)H2, C(7)H2, C(7’)H2, C(8)H2, C(8’)H2, C(9)H2, C(9’)H2), 1.27 (4H, m, C(10)H2, C(10’)H2), 0.88 (6H, t, C(11)H3, C(11’)H3); 13C-NMR (CDCl3): δ 58.9 (C(1), C(1’)), 20.0 (C(2), C(2’)), 26.4 (C(3)), 58.1 (C(4), C(4’)), 21.7 (C(5), C(5’)), 22.5 (C(6), C(6’)), 29.0 (C(7), C(7’)), 28.9 (C(8), C(8’)), 31.6 (C(9), C(9’)), 20.6 (C(10), C(10’)), 14.0 (C(11), C(11’)).
N,N-dioctylpyrrolidinium bromide (3). From 1,4-dibromobutane (4.2g, 0.02 mol). Recrystallized from ethyl acetate;yield: 98%, m.p. 120-124oC; Elemental analysis for C20H42NBr found (calc) %C 63.47 (63.81); %H 11.76 (11.24); %N 3.78 (3.72); ES+MS m/z 296(C20H42N); 1H-NMR (CDCl3): δ 3.85 (4H, t, C(1)H2, C(1’)H2 ), 2.31 (4H, m, C(2)H2, C(2’)H2), 3.43 (4H, t, C(4)H2, C(4’)H2), 1.70 (20H, m, C(5)H2, C(5’)H2, C(6)H2, C(6’)H2, C(7)H2, C(7’)H2, C(8)H2, C(8’)H2, C(9)H2, C(9’)H2), 1.27 (4H, m, C(10)H2, C(10’)H2), 0.88 (6H, t, C(11)H3, C(11’)H3 ); 13C-NMR (CDCl3): δ 62.9 (C(1), C(1’)), 21.8 (C(2), C(2’)), 59.4 (C(4), C(4’)), 23.4 (C(5), C(5’)), 26.3 (C(6), C(6’)), 29.0 (C(7), C(7’)), 28.9 (C(8), C(8’)), 31.5 (C(9), C(9’)), 22.5 (C(10), C(10’)), 14.0 (C(11), C(11’)).

Acknowledgments

This work was supported by the funds from Adam Mickiewicz University, Faculty of Chemistry.

References

  1. Domagk, G. A new class of disinfectants. Dtsch. Med. Wochenscher 1935, 61, 829–832. [Google Scholar]
  2. Block, S.S. Disinfection, Sterilization, and Preservation, 5th ed; Lippincott Williams & Wilkins: Philadelphia, PA, USA, 2001. [Google Scholar]
  3. Dega-Szafran, Z.; Dulewicz, E.; Brycki, B. Synthesis and characterization of 1-carbalkoxymethyl-4-hydroxy-1-methylpiperidinium chlorides. ARKIVOC 2007, VI, 90–102. [Google Scholar]
  4. Erickson, J.G.; Keps, J.S. Reactions of Long-chain Amines. IV. Preparation of N-Alkylpyrrolidines, N,N-Dialkylpyrrolidinium Chlorides and N,N-Dialkylpiperidinium Chlorides. J. Am. Chem. Soc. 1955, 77, 485–486. [Google Scholar] [CrossRef]
  5. Ringdahl, B.; Roch, M.; Jenden, D.J. Tertiary 3- and 4-Haloalkylamine Analogues of Oxotremorine as Prodrugs of Potent Muscarinic Agonists. J. Med. Chem. 1988, 31, 160–164. [Google Scholar]
  6. Chukhadzhyan, É.O.; Chukhadzhyan, Él.O.; Shakhatuni, K.G.; Gevorkyan, N.T. Synthesis of Chlorine-Substituted Naphto- and Benzisoindolinium Salts by Base-catalyzed Intramolecular Cyclization of Ammonium Salts. Chem. Heterocyclic Comp. 1998, 34, 912–915. [Google Scholar] [CrossRef]
  7. Schmitt, K.D. Surfactant-Mediated Phase Transfer as an Alternative to Propanesultone Alkylation. Formation of a New Class of Zwitterionic Surfactants. J. Org. Chem. 1995, 60, 5474–5479. [Google Scholar] [CrossRef]
  8. Venkataramu, S.D.; Macdonell, G.D.; Purdum, W.R.; Dilbeck, G.A.; Berlin, K.D. Polyphosphoric Acid Catalyzed Cyclization of Aralkenyl-Substituted Quaternary Ammonium Salts. J. Org. Chem. 1977, 42, 2195–2200. [Google Scholar]
  9. Walker, M.A. An Unusual Tandem Cyclization-Stevens Rearrangement Mediated by Ph3P/DEAD or Bu3P/ADDP. Tetrahedron Lett. 1996, 37, 8133–8136. [Google Scholar]
  10. Glaeske, K.W.; West, F.G. Chirality Transfer from Carbon to Nitrogen to Carbon via Cyclic Ammonium Ylides. Org. Lett. 1999, 1, 31–33. [Google Scholar] [CrossRef]
  11. Manivannan, G. Disinfection and Decontamination; Principles, Applications and Related Issues; CRC Press, Taylor & Francis Group: Boca Raton, FL, USA, 2008. [Google Scholar]
  12. Menger, F.M.; Keiper, J.S. Gemini surfactants. Angew. Chem. Int. Ed. 2000, 39, 1906–1920. [Google Scholar] [CrossRef]
  13. Paulus, W. Directory of Microbiocides for the Protection of Materials; A. Handbook, Springer: Dordrecht, The Netherlands, 2005. [Google Scholar]
  14. Zana, R.; Xia, J. Gemini Surfactants. Synthesis, Interfacial and Solution PhaseBehavior, and Applications; Marcel Dekker: New York, NY, USA, 2004. [Google Scholar]
  15. Fraise, A.P.; Lambert, P.A.; Maillard, J-Y. Russell, Hugo & Ayliffe’s Principles and Practice of Disinfection, Preservation and Sterilization, 4th ed; Blackwell Publishing: Malden, MA, USA, 2004. [Google Scholar]
  16. Park, E.J.; Kim, M.H.; Kim, D.Y. Enantioselective Alkylation of â-Keto Esters by Phase-Transfer Catalysis Using Chiral Quaternary Ammonium Salts. J. Org. Chem. 2004, 69, 6897–6899. [Google Scholar] [CrossRef]
  17. Kim, D.Y.; Huh, S.Ch.; Kim, S.M. Enantioselective Michael reaction of malonates and chalcones by phase-transfer catalysis using chiral quaternary ammonium salt. Tetrahedron Lett. 2001, 42, 6299–6301. [Google Scholar]
  18. Kim, D.Y.; Park, E.J. Catalytic Enantioselective Fluorination of α-Keto Esters by Phase-Transfer Catalysis Using Chiral Quaternary Ammonium Salts. Org. Lett. 2002, 4, 545–547. [Google Scholar] [CrossRef]
  19. Niess, B.; Jorgensen, K.A. The asymmetric vinylogous Mannich reaction of dicyanoalkylidenes with α-amido sulfones under phase-transfer conditions. Chem. Commun. 2007, 1620–1622. [Google Scholar]
  20. Orglmeister, E.; Mallat, T.; Baiker, A. Quaternary ammonium derivatives of cinchonidine as new chiral modifiers for platinum. J. Catal. 2005, 233, 333–341. [Google Scholar] [CrossRef]
  21. Breistein, P.; Karlsson, S.; Hendenström, E. Chiral pyrrolidinium salts as organocatalysts in the stereoselective 1,4-conjugate addition of N-methylpyrrole to cyclopent-1-ene carbaldehyde. Tetrahedron: Asymmetry 2006, 17, 107–111. [Google Scholar] [CrossRef]
  22. Fuijmori, T.; Fujii, K.; Kanzaki, R.; Chiba, K.; Yamamoto, H.; Umebayashi, Y.; Ishiguro, S. Conformational structure of room temperature ionic liquid N-butyl-N-methyl-pyrrolidinium bis(trifluoromethanesulfonyl) imide—Raman spectroscopic study and DFT calculations. J. Mol. Liquids 2007, 131-132, 216–224. [Google Scholar] [CrossRef]
  23. Yoshinawa-Fujita, M.; Johansson, K.; Newman, P.; MacFarlane, D.R.; Forsyth, M. Novel Lewis-base ionic liquids replacing typical anions. Tetrahedron Lett. 20 0647, 2755–2758. [Google Scholar]
  24. Henderson, W.A.; Passerini, S. Phase Behavior of Ionic Liquid-LiX Mixtures: Pyrrolidinium Cations and TFSI- Anions. Chem. Matter. 2004, 16, 2881–2885. [Google Scholar] [CrossRef]
  25. Sun, J.; MacFarlane, D.R.; Forsyth, M. A new family of ionic liquids based on the 1-alkyl-2-methyl pyrrolinium cation. Electrochim. Acta 2003, 48, 1707–1711. [Google Scholar] [CrossRef]
  26. MacFarlane, D.R.; Forsyth, S.A.; Golding, J.; Deacon, G.B. Ionic liquids based on imidazolium, ammonium and pyrrolidinium salts of the dicyanamide anion. Green Chem. 2002, 4, 444–448. [Google Scholar] [CrossRef]
  27. Sakaaebe, H.; Matsumoto, H. N-Methyl-N-propylpiperidinium bis(trifluoromethanesulfonyl) imide (PP13–TFSI) – novel electrolyte base for Li battery. Electrochem. Communi. 2003, 5, 594–598. [Google Scholar] [CrossRef]
  28. Tigelaar, D.M.; Meador, M.A.B.; Bennett, W.R. Composite Electrolytes for Lithium Batteries: Ionic Liquids in APTES Cross-Linked Polymers. Macromolecules 2007, 40, 4159–4164. [Google Scholar] [CrossRef]
  29. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A.; Vreven, T., Jr.; Kudin, K.N.; Burant, J.C.; Millam, J.M.; Iyengar, S.S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G.A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.X.Li.; Knox, J.E.; Hratchian, H.P.; Cross, J.B.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R.E.; Yazyev, O.; Austin, A.J.; Cammi, R.; Pomelli, C.; Ochterski, J.W.; Ayala, P.Y.; Morokuma, K.; Voth, G.A.; Salvador, P.; Dannenberg, J.J.; Zakrzewski, V.G.; Dapprich, S.; Daniels, A.D.; Strain, M.C.; Farkas, O.; Malick, D.K.; Rabuck, A.D.; Raghavachari, K.; Foresman, J.B.; Ortiz, J.V.; Cui, Q.; Baboul, A.G.; Clifford, S.; Cioslowski, J.; Stefanov, B.B.; Liu, G., Liashenko; Piskorz, P.; Komaromi, I.; Martin, R.L.; Fox, D.J.; Keith, T.; Al-Laham, M.A.; Peng, C.Y.; Nanayakkara, A.; Challacombe, M.; Gill, P.M.W.; Johnson, B.; Chen, W.; Wong, M.W.; Gonzalez, C.; Pople, J.A. Gaussian 03, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2004. [Google Scholar]
  30. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  31. Becke, A.D. Density-functional thermochemistry. V. Systematic optimization of exchange-correlation functionals. J. Chem. Phys. 1997, 107, 8554–8560. [Google Scholar] [CrossRef]
  32. Hehre, W.J.; Random, L.; Schleyer, P.v.R.; Pople, J.A. Ab Initio Molecular Orbital Theory; Wiley: New York, NY, USA, 1986. [Google Scholar]
  33. Wolinski, K.; Hilton, J.F.; Pulay, P. Efficient implementation of the gauge-independent atomic orbital method for NMR chemical shift calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds are available from the authors.

Share and Cite

MDPI and ACS Style

Brycki, B.; Szulc, A.; Kowalczyk, I. Study of Cyclic Quaternary Ammonium Bromides by B3LYP Calculations, NMR and FTIR Spectroscopies. Molecules 2010, 15, 5644-5657. https://doi.org/10.3390/molecules15085644

AMA Style

Brycki B, Szulc A, Kowalczyk I. Study of Cyclic Quaternary Ammonium Bromides by B3LYP Calculations, NMR and FTIR Spectroscopies. Molecules. 2010; 15(8):5644-5657. https://doi.org/10.3390/molecules15085644

Chicago/Turabian Style

Brycki, Bogumił, Adrianna Szulc, and Iwona Kowalczyk. 2010. "Study of Cyclic Quaternary Ammonium Bromides by B3LYP Calculations, NMR and FTIR Spectroscopies" Molecules 15, no. 8: 5644-5657. https://doi.org/10.3390/molecules15085644

Article Metrics

Back to TopTop