7

Click here to load reader

Determination of the self-association and inter-association equilibrium constants of a carboxylic acid and its mixtures with pyridine derivates

Embed Size (px)

Citation preview

Page 1: Determination of the self-association and inter-association equilibrium constants of a carboxylic acid and its mixtures with pyridine derivates

Determination of the self-association and inter-association equilibrium

constants of a carboxylic acid and its mixtures with pyridine derivates

A. Gonzalez a, L. Irusta a, M.J. Fernandez-Berridi a,*, J.J. Iruin a, T. Sierra b, L. Oriol b

a Polymer Science and Technology Department and Institute for Polymer Materials (POLYMAT),

University of the Basque Country, P.O. Box 1072, 20080 San Sebastian, Spainb Polymer and Liquid Crystal Group, Quımica Organica, Instituto de Ciencia de Materiales de Aragon,

Universidad de Zaragoza-CSIC, Pedro Cerbuna 12, 50009 Zaragoza, Spain

Received 11 November 2005; received in revised form 7 December 2005; accepted 8 December 2005

Available online 18 January 2006

This is paper is dedicated to Prof. Cecilia Sarasola

Abstract

Infrared spectroscopy has been employed to determine absorptivity coefficients and the self-association equilibrium constant of 3,4,5-

tris(dodecyloxy)benzoic acid in cyclohexane. The obtained value is consistent with data reported in bibliography for other similar compounds. The

equilibrium constant describing ethylpyridine/carbonyl hydrogen bonds has been determined for mixtures of the above-mentioned acid and two

ethylpyridine isomers. The effect of steric hindrance is discussed.

# 2005 Elsevier B.V. All rights reserved.

Keywords: H-bonding; Supramolecular chemistry; Self-association; Inter-association equilibrium constants; Polycatenar benzoic acid

www.elsevier.com/locate/vibspec

Vibrational Spectroscopy 41 (2006) 21–27

1. Introduction

Intermolecular hydrogen bonding is considered an important

driving force in the formation of supramolecular liquid crystals

[1–3]. Many supramolecules as well as supramolecular

polymers, prepared from components that are appropriately

substituted with functional H-donor and H-acceptor groups, are

able to self-organize into mesophases. In many cases, the

components are not necessarily mesogenic in their own right.

These complexes present the same types of mesophases as

covalently built mesogenic molecules. However, the resulting

organizations may rely on some additional features derived

from the dynamic character of the H-bond.

Most of the mesogenic supramolecular structures described

so far show mesophases of the calamitic type. Several disc-like

supramolecules built through hydrogen bonding have also been

reported to display columnar mesomorphism. This type of

liquid crystalline architecture deserves special attention due to

its applications such as (semi)conducting and photoconducting

materials, chemical sensors, etc. [4–7]. In this work, we have

* Corresponding author. Tel.: +34 943 018194; fax: +34 943 015270.

E-mail address: [email protected] (M.J. Fernandez-Berridi).

0924-2031/$ – see front matter # 2005 Elsevier B.V. All rights reserved.

doi:10.1016/j.vibspec.2005.12.003

studied the hydrogen bonding capability of 3,4,5-tris(dodecy-

loxy)benzoic acid, coded as 1. This acid has a polycatenar

structure and has been widely used on the preparation of

columnar liquid crystals both of low molecular weight and

polymers [8–10]. The study of the self-association and inter-

association of this type of carboxylic acids is the starting point

to understand the complex hydrogen bonding formation that

takes place in the supramolecular liquid crystals [11,12].

Infrared spectroscopy has been an important technique in the

study of both the nature and extent of hydrogen bonding [13] in

organic compounds, in particular on carboxylic acids [14–18].

It is well established in carboxylic acids that the self-

association takes place with the formation of cyclic dimers

(Scheme 1). This self-association equilibrium can be described

by a single equilibrium constant, KB [13,19,20].

The main objective of this work is to obtain the values of the

equilibrium constants describing the formation of 3,4,5-

tris(dodecyloxy)benzoic acid dimers (KB) and the inter-

association 1/pyridine complexes (KA), shown in Scheme 1,

by means of IR spectroscopy.

With the aim in mind, liquid sampling cells of known

thickness were used to obtain infrared spectra of 1 as a function

of concentration in a non-hydrogen bonding hydrocarbon

Page 2: Determination of the self-association and inter-association equilibrium constants of a carboxylic acid and its mixtures with pyridine derivates

A. Gonzalez et al. / Vibrational Spectroscopy 41 (2006) 21–2722

Scheme 1. Self-association of the carboxylic acid.

Fig. 1. Infrared spectra in the carbonyl-stretching vibration region for 1/

cyclohexane solutions at different concentrations.

solvent such as cyclohexane. Although most compounds

containing carboxylic acid groups are insoluble in non-polar

solvents, 1 presents a high solubility in cyclohexane, due to the

presence of large non-polar aliphatic chains. This fact offers the

possibility of studying the self-association of a carboxylic acid

in a non-polar inert solvent, which has not been reported on

bibliography so far. For the calculation of the acid/pyridine

inter-association equilibrium constant, mixtures of 1 and

ethylpyridine were prepared at different ratios in cyclohexane.

Infrared spectra of these mixtures were recorded in liquid

sampling cells of known thickness.

2. Experimental

Cyclohexane (Aldrich, HPLC grade), 2-ethylpyridine

(MERCK) and 4-ethylpyridine (Aldrich 98% (GC)) were used

as received. Compound 1 was synthesized according to a

previously reported method [8]. Infrared spectroscopic

measurements were recorded at room temperature in a Nicolet,

model Magna 560 FT-IR spectrometer at a resolution of 2 cm�1

and a minimum of 10 scans. Infrared spectra of 1 solutions

between 0.0003 and 0.14 mol/L were carried out using two

standard liquid cells of 0.1 and 1 mm path length, respectively.

Infrared spectra of 1/ethylpyridine mixtures (1/5, 1/10, 1/20 and

1/40 mol ratio) in cyclohexane were also obtained and in

solutions the concentration of 1 was 2.5 � 10�3 mol/L. Infrared

spectra of these mixture solutions were performed in a cell of

2 mm. All solution spectra were sustracted from the

cyclohexane spectrum obtained in the same liquid cell.

Infrared spectra of a thin film of 1 at different temperatures

were obtained using a variable temperature cell (Harrick) from

150 8C down to room temperature. The thin film of 1 was

prepared by casting from dichloromethane solutions onto a KBr

window.

3. Results and discussion

3.1. Self-association of 3,4,5-tris(dodecyloxy)benzoic acid

3.1.1. Infrared spectra

Fig. 1 shows FT-IR spectra in the carbonyl-stretching

vibration region of 1 in cyclohexane solutions at different

concentrations, recorded at room temperature. In the spectrum

of the most concentrated solution, a single band centred at

1690 cm�1 is observed. This band can be assigned to hydrogen-

bonded carbonyl-stretching vibration in the closed dimeric

form [21]. As the concentration of the solution is decreased a

new band centred at 1740 cm�1 appears. Furthermore, the

lower the solution concentration, the higher the relative

contribution of this band. Therefore, it can be assigned to

the free carbonyl-stretching vibration. The difference in

wavenumbers (about 50 cm�1) between the free and associated

carbonyl bands is very large, confirming that the self-associated

interaction is very strong [22].

The corresponding hydroxyl-stretching vibration band

appears in the region of 3400–3600 cm�1. However, the

resolution is so poor that it is not possible to discern between

associated and free bands, a prerequisite for performing

quantitative analysis. Therefore, only the carbonyl-stretching

region is used for such type of analysis.

3.1.2. Determination of the absorptivity coefficients

In order to quantify the self-association of 1, the relative

contribution of the free and associated carbonyl bands must be

calculated. However, the absorptivity coefficients of these two

vibrations are not the same and, in order to calculate the relative

concentration of free species, at least one of the two coefficients

must be known [14]. In order to quantify the absorptivity

coefficient of the hydrogen-bonded carbonyl vibration, the area

of the associated carbonyl band has been plotted as a function of

concentration. In the more concentrated solutions, where no

free band is observed, a straight line must be obtained. From the

slope of the line the absorptivity coefficient of the associated

band can be readily calculated, according to the Lambert–Beer

Page 3: Determination of the self-association and inter-association equilibrium constants of a carboxylic acid and its mixtures with pyridine derivates

A. Gonzalez et al. / Vibrational Spectroscopy 41 (2006) 21–27 23

Fig. 2. Determination of the absorptivity coefficient of bonded carbonyl groups

of 1.

Fig. 3. Determination of the absorptivity coefficient ratio of the carbonyl-

stretching vibration of 1/cyclohexane solutions.

equation. Fig. 2 shows the result of this calculation and a value

of eb = 159 L /mol mm has been obtained.

The fraction of bonded carbonyl groups at different

concentrations can be then determined using Eq. (1) [23,24]:

fB ¼AB

eBlC(1)

where l is the path length of the liquid cell, AB the area of the H-

bonded band and C is the solution concentration. The results of

these calculations are summarized in Table 1. As can be seen,

the fraction of self-associated carbonyl groups increases when

the concentration of [1] increases.

From the values of the associated carbonyl fraction as a

function of the concentration, the ratio between the two

absorption coefficients (r = eB/eF) can be calculated by means

of Eq. (2):

1

fB¼ 1þ r

AF

AB

(2)

where AF and AB are the areas of the free and associated bands,

respectively. Fig. 3 shows the result of this calculation. Only the

Table 1

Fraction of bonded carbonyl groups of 1 obtained at different concentrations

C (mol/L) Area of carbonyl

bonded band, AB

Area of carbonyl

free band, Af

Path length l,

(mm)

fB

0.140 22.38 0.106 0.1 1.00

0.112 16.52 0.070 0.1 0.93

0.088 12.16 0.080 0.1 0.87

0.072 10.70 0.083 0.1 0.93

0.056 8.33 0.064 0.1 0.94

0.046 7.21 0.086 0.1 0.98

0.028 4.19 0.071 0.1 0.94

0.017 2.60 0.062 0.1 0.96

0.010 1.53 0.049 0.1 0.96

0.006 0.84 0.030 0.1 0.88

0.004 3.61 0.213 1.0 0.57

0.002 2.64 0.194 1.0 0.83

0.0008 0.91 0.119 1.0 0.71

0.0005 0.53 0.084 1.0 0.67

0.0003 0.30 0.069 1.0 0.62

most diluted concentrations, where the free carbonyl band is

clearly observed, have been used. From the slope of the straight

line a value of r = 1.9 has been obtained which is consistent

with values previously reported in bibliography (r = 1.6) [13].

3.1.3. Determination of the self-association equilibrium

constant from the carbonyl-stretching region

Knowing the free carbonyl band fraction ( fF) and the

stoichiometry of the acid self-association, the equilibrium

constant can be calculated. The procedure is the following: the

volume fraction (fB) of each concentration can be obtained by

means of Eq. (3):

fB ¼ ciVM (3)

where VM is the molar volume of 1 (708.9 cm3/mol) calculated

using the method of group contributions [13]. The direct

calculation of VM from its density value was not possible

due to the lack of enough sample quantity to perform this

calculation.

From the value of fB and selecting a starting value for KB

and fB1(volume fraction of carboxylic acid monomer) can be

calculated using Eq. (4). From Eq. (5), fF can be calculated for

the whole composition range. The value of KB is systematically

varied and a least square method is employed in order to

determine the best fit of the experimental fF versus fB data

(Fig. 4):

fB ¼ fB1 þ 2KBf2B1

(4)

fF ¼1

1þ 2KBfB1

(5)

The commercial software (Fit K) developed by Painter and

Coleman has been used to perform this calculation.

The least square fit to the equations describing the system

results in a fine match of the experimental data over the entire

composition range considered. The best fit of the data has

Page 4: Determination of the self-association and inter-association equilibrium constants of a carboxylic acid and its mixtures with pyridine derivates

A. Gonzalez et al. / Vibrational Spectroscopy 41 (2006) 21–2724

Fig. 5. Infrared spectra in the carbonyl-stretching region of 1/4-ethylpyridine

mixtures.

Fig. 4. Experimental and simulated fractions of free carbonyl groups vs.

volume fraction of 1/cyclohexane solutions.

permitted to obtain a value of KB = 8900. This is a dimentionless

quantity equal to the definition of an equilibrium constant in

terms of molar concentration divided by the molar volume of the

structural unit (708 cm3/mol). It is interesting to note that the

equilibrium constant value determined from 1 is slighly higher to

that determined by Barela et al. [25] for benzoic acid in benzene

(6175). The same authors have reported [26] that the extent of

dimerization of the benzoic acic is lowered by substitution of the

benzene ring and that the effect is more pronounced by ortho

substituents. Taking into account these results a lower value of KB

should be obtained in 1. However, as Marcus [27] pointed out,

there is considerable evidence that aromatic hydrocarbons (as

benzene) interact with molecules containing OH groups,

possibly through the formation of very weak hydrogen bonds,

lowering the value of KB.

3.2. Determination of the inter-association equilibrium

constant in solution

After studying the self-association of 1, we accomplished

the study of the H-bonding equilibrium of the polycatenar acid

[1] and pyridines as H-acceptor. In particular, the inter-

association equilibrium constant, KA, can be calculated from

the infrared spectra of 1/ethylpyridine mixtures. In this study,

two pyridine derivatives, having an ethyl group in position 4 or

2, have been evaluated.

Scheme 2. Inter-association of the carboxylic acid with ethylpyridine.

3.2.1. [1]/4-ethylpyridine mixtures

Fig. 5 shows scale expanded room temperature infrared

spectra of 1/4-ethylpyridine mixtures recorded at different

compositions in the carbonyl-stretching region using cyclo-

hexane solutions having a fixed concentration of 1 (0.0025 M).

As can be seen, IR spectra show the two bands previously

observed for pure 1 and assigned to free and associated dimeric

species [28] (4-ethylpyridine does not present any band in the

carbonyl-stretching vibration region). In addition, the spectra of

the mixtures show a new band at 1710 cm�1, whose intensity

varies systematically as a function of composition [29–31].

This band can be assigned to free carbonyl groups whose OH is

hydrogen-bonded to the pyridine nitrogen, according to

Scheme 2 [32]. The band at 1690 cm�1, assigned to associated

carbonyl groups, decreases as increasing ethylpyridine con-

centration, meanwhile the band at 1740 cm�1 corresponding to

free carbonyl groups is still present but shows a low intensity.

This fact can be explained because the total carbonyl

concentration remains constant in all samples and, therefore,

the increase of the intensity of the 1710 cm�1 band must lead to

a decrease of the other band.

In addition to the three mentioned bands, the spectra of the

mixtures also exhibit a shoulder centered at about 1720 cm�1.

The assignment of this band to a specific interaction is not

straightforward and could be due to opened dimers of the

carboxylic acid. In order to check the origin of this band, a

Page 5: Determination of the self-association and inter-association equilibrium constants of a carboxylic acid and its mixtures with pyridine derivates

A. Gonzalez et al. / Vibrational Spectroscopy 41 (2006) 21–27 25

Fig. 6. Infrared spectra in the carbonyl-stretching region of 1 at different

temperatures.

Table 2

Fraction of bonded and free carbonyl groups for different 1/4-ethylpyridine

mixtures

1/4-Ethylpyridine fF fB

1/0 0.13 0.87

1/5 0.54 0.46

1/10 0.70 0.30

1/20 0.82 0.18

1/40 0.86 0.14

Fig. 7. Infrared spectra in the carbonyl-stretching vibration region of [1]/2-

ethylpyridine mixtures.

study of the dependence of the IR spectrum of 1 on the

temperature was accomplished on a thin film sample.

Fig. 6 displays the corresponding spectra in the carbonyl

region of a thin film of 1 at different temperatures. As can be

seen, three variable intensity bands as a function of temperature

appear in all spectra. Thus, at the highest temperature (150 8C)

a band centred at 1690 cm�1 dominates the region. This

position is characteristic of carbonyl groups associated in cyclic

dimers. Moreover, a band at 1738 cm�1 with a shoulder at

1720 cm�1 can be also detected. The band at higher

wavenumbers is due to totally free carbonyl groups. As

temperature decreases, the relative intensity of this shoulder

(1720 cm�1) increases in such a way that, at the lowest

temperature, its intensity predominates over that of 1738 cm�1.

Therefore, this shoulder can respond to carbonyl groups

participating in open dimeric associations.

Taking into account these results, the band at the same

position observed in 1/pyridine mixtures can be assigned to this

type of association (open dimer). It must be stressed out that

this band is centred at higher wavenumbers than the

corresponding to carbonyl groups whose OH is associated

with the nitrogen of pyridine. This fact indicates that the

pyridine-acid inter-association is stronger than the acid/acid

self-association, in accordance with previous results [28,33].

The calculation of KA implies the determination of free

carbonyl fractions for mixtures at different compositions. Due

to the proximity of the frequencies of the bands involved in the

calculation, the determination of their relative contributions

must be done by means of curve resolution. The software

Spectra Fit, developed by Painter and Coleman has been

employed for this purpose. The fractions of free and bonded

carbonyl groups have been determined by Eqs. (7) and (8).

fB ¼AB�

eB

eF1

�AF1þ�

eB

eF2

�AF2þ�

eB

eF3

�AF3þ AB

(7)

fF ¼ 1� fB (8)

where AB, AF1, AF2

and AF3are the areas of associated cyclic

closed dimers, totally free, open dimers and N–H bonded

bands, respectively. eFi (i = 1, 2 or 3) are the corresponding

infrared absorption coefficients.

To calculate the fraction of free and bonded groups the

absorption coefficients must be known. However, although

absorption coefficients of bonded (eB) and totally free carbonyl

groups (eF1) are known, absorption coefficients eF2

and eF3are

unfortunately unknown. Therefore, a linear dependence of the

absorption coefficients with wavenumbers was supposed, in

order to obtain the rest of the coefficients values. Infrared

absorption coefficients of 113 and 135 were obtained for the

open dimer and pyridine-acid carbonyl band, respectively.

Table 2 shows the free carbonyl group fractions at different

compositions. The fraction of bonded carbonyl groups

decreases as increasing pyridine concentration. It is worthwhile

to remember that as the total concentration of carbonyl groups

remains constant, carbonyl groups that are self-associated at

low pyridine concentrations become ‘‘free’’ when increasing

pyridine content.

The equilibrium constant KA can be obtained from free

carbonyl fraction data. Once again, the commercial software Fit

K was applied for this task. Using the self-association KB

previously determined (8900) and stating an initial value of KA,

a least square fit procedure is applied. The value of the free

fraction for the whole composition can be calculated using the

appropriate stoichiometric equations. The value of KA is

systematically varied to fit the experimental curve using a least

Page 6: Determination of the self-association and inter-association equilibrium constants of a carboxylic acid and its mixtures with pyridine derivates

A. Gonzalez et al. / Vibrational Spectroscopy 41 (2006) 21–2726

Table 3

Fraction of bonded and free carbonyl groups for 1/2-ethylpyridine mixtures at

different compositions

1/Ethylpyridine fF fB

1/0 0.07 0.93

1/5 0.34 0.66

1/10 0.49 0.51

1/20 0.67 0.33

1/40 0.82 0.18

squares procedure. Other parameters such as molar volume of

components have been employed to fit: V[1] = 708.9; VCHX = 99

and V4-ETPY = 114.8 cm3/mol (calculated by group contribu-

tions as stated previously [13,31]). A value of 335 has been

obtained.

3.2.2. [1]/2-ethylpyridine mixtures

The methodology for the calculation of the equilibrium

constant for 1/2-ethylpyridine system is the same as the

previously employed for 1/4-ethylpyridine mixtures in cyclo-

hexane.

Fig. 7 shows infrared spectra in the carbonyl-stretching

vibration region for 1/2-ethylpyridine mixtures. As can be seen,

the infrared spectra of these mixtures show the same absorption

bands and composition dependence as the observed for

mixtures with 4-ethylpyridine.

Nevertheless, for the same composition, the relative

absorbance of the band assigned to cyclic closed dimer is

higher in 2-ethylpyridine mixtures than the obtained for 4-

ethylpyridine, while the band corresponding to the association

between pyridine and 1 is lower in the former. This behaviour

can be numerically demonstrated when the free carbonyl

fraction is calculated, as can be seen in Table 3.

Applying the previously defined methodology, a value of

standarized equilibrium inter-association constant KA = 152

has been obtained. As expected, this value is lower than the

obtained for 1/4-ethylpyridine system presumably due to a

higher steric hindrance of the nitrogen atom in 2-ethylpyridine

to interact with the hydroxyl group of the carboxylic

compound.

4. Conclusions

The analysis of the infrared carbonyl-stretching vibration

region of 1 has shown different bands due to the presence of two

distinctive absorbing carbonyl groups. Thus, two bands

corresponding to free and hydrogen-bonded carbonyl groups

have been identified.

The constant of the self-association of 1 has been determined

by quantitative analysis of these bands. A value of 8900 has

been obtained, which is in accordance with data reported on in

bibliography.

The analysis of the carbonyl-stretching vibration region of 1/

ethylpyridine mixtures has shown the existence of a new band

assigned to free carbonyl groups generated as a consequence of

OH hydrogen bonding with pyridine nitrogen. This band is

located at lower frequencies than the corresponding to the

‘‘open dimer’’, indicative of a higher strength of the hydrogen

bonding between OH and nitrogen. The position of this band

does not show any dependence on the type of pyridine isomer,

what means that the hydrogen bonding force is about the same

for both isomers. Nevertheless, lower equilibrium constant

values have been obtained for 2-ethylpyridine because the

steric hindrance of the nitrogen in this component.

It is important to recognize that even though the strength of

the interaction between the carboxylic acid and pyridine groups

exceeds that of the carboxylic cyclic closed dimer, the

magnitude of the carboxylic acid dimer equilibrium constant

is greater.

FT-IR results show that the hydrogen bonding strength in

these systems is high enough to be considered the driving force

to achieve supramolecular liquid crystal polymers having

columnar mensomorphism based on this kind of polycatenar

acids.

Acknowledgements

We express our thanks to the University of the Basque

Country and Aragon Government for their continous support

through their respective consolidated research group programs.

References

[1] C.M. Paleos, D. Tsiourvas, Angew. Chem. Int. Ed. Eng. 34 (1995)

1696.

[2] T. Kato, N. Mizoshita, K. Kanie, Macromol. Rapid Commun. 22 (2001)

797.

[3] U. Beginn, Prog. Polym. Sci. 28 (2003) 1049.

[4] N. Boden, B. Movaghar, in: D. Demus, J. Goodby, G.W. Gray, H.-W.

Spiess, V. Vill (Eds.), Handbook of Liquid Crystals, vol. 2, Wiley–VCH,

Weinheim, 1998, p. 781.

[5] J. Simmerer, B. Glusen, W. Paulus, A. Kettnwe, P. Schuhmacher, A.

Adam, K.H. Etzbach, K. Siemensmeyer, J.H. Wendorff, H. Ringsdorf, D.

Haarer, Adv. Mater. 8 (1996) 815.

[6] J. Wright, P. Roisin, G.P. Rigby, R. Nolte, M.J. Cook, S.C. Thorpe, Sens.

Actuators B 13–14 (1993) 276.

[7] A. Adam, P. Schuhmacher, J. Simmerer, L. Haussling, K. Siemensmeyer,

K.H. Etzbach, H. Ringsdorf, D. Haarer, Nature 371 (1994) 141.

[8] C. Mertesdorf, H. Ringsdorf, Liq. Cryst. 5 (1989) 1757.

[9] C. Lin, H. Ringsdorf, M. Ebert, R. Kleppinger, J.H. Wendorff, Liq. Cryst. 5

(1989) 1841.

[10] G. Staufer, G. Latterman, Makromol. Chem. 192 (1991) 2421.

[11] T. Kato, M. Frechet, Macromolecules 23 (1990) 360.

[12] P.C. Painter, C. Cleveland, M.M. Coleman, Mol. Cryst. Liq. Cryst. Sci.

Technol., Sect. A 348 (2000) 269.

[13] M.M. Coleman, P.C. Painter, J.F. Graf, Specific Interactions and the

Miscibility of Polymer Blends, Technomic Publishing Ing, Lancaster,

PA, 1991.

[14] C. Cleveland, S. Fearneley, Y. Hu, M. Wagman, P. Painter, M. Coleman, J.

Macromol. Sci. Phys. 2 (2000) 197.

[15] I. Nahringbauer, Acta Chem. Scand. 24 (1970) 453.

[16] B. Brozoski, M. Coleman, P. Painter, Macromolecules 17 (1984)

230.

[17] B. Brozoski, M. Coleman, P. Painter, Macromolecules 17 (1984) 1591.

[18] M. Coleman, J. Lee, P. Painter, Macromolecules 23 (1990) 2339.

[19] M. Coleman, K. Guigley, P. Painter, Macromol. Chem. Phys. 200 (1999)

1167.

[20] M. Coleman, P. Painter, Macromol. Chem. Phys. 199 (1998) 1307.

Page 7: Determination of the self-association and inter-association equilibrium constants of a carboxylic acid and its mixtures with pyridine derivates

A. Gonzalez et al. / Vibrational Spectroscopy 41 (2006) 21–27 27

[21] M. Coleman, J. Lee, C. Serman, Z. Wang, P. Painter, Polymer 30 (1989)

1298.

[22] L. Zhang, S. Goh, S. Lee, J. Appl. Polym. Sci. 74 (1999) 383.

[23] L. Irusta, J.J. Iruin, M.J. Fernandez-Berridi, M. Sobkowiak, P. Painter, M.

Coleman, Vibr. Spectrosc. 23 (2000) 187.

[24] L. Irusta, M. Abee, J.J. Iruin, M.J. Fernandez-Berridi, Vibr. Spectrosc. 27

(2001) 183.

[25] R. Barela, G. Liwski, H. Szatylowicz, Fluid Phase Equilib. 105 (1995) 119.

[26] R. Barela, H. Buchowski, H. Szatylowicz, Fluid Phase Equilib. 92 (1994)

303.

[27] Y. Marcus, Introduction to Liquid State Chemistry, John Wiley and Sons,

New York, 1977;

J. Lee, P. Painter, M. Coleman, Macromolecules 21 (1988) 954.

[28] S. Johnson, K.J. Rumon, Phys. Chem. 69 (1965) 74.

[29] C. Bauzin, F. Brandys, Chem. Mater. 4 (1992) 970.

[30] L. Wilson, Macromolecules 27 (1994) 6683.

[31] M.J. Fernandez-Berridi, J.J. Iruin, L. Irusta, J. Mercero, J. Ugalde, J. Phys.

Chem. A 106 (2002) 4187.

[32] M. Coleman, P. Painter, Prog. Polym. Sci. 20 (1995) 1.

[33] J. Lee, P. Painter, M. Coleman, Macromolecules 21 (1988) 346.