Sei sulla pagina 1di 21

See

discussions, stats, and author profiles for this publication at:


https://www.researchgate.net/publication/257209860

The urea–phenol(s) systems1

Article in Fluid Phase Equilibria · October 1998


DOI: 10.1016/S0378-3812(98)90206-0

CITATIONS READS

3 218

8 authors, including:

Małgorzata E. Jamróz Michał Henryk Jamróz


Instytut Chemii Przemysłowej Institute of Nuclear Chemistry and Tec…
15 PUBLICATIONS 117 CITATIONS 73 PUBLICATIONS 1,140 CITATIONS

SEE PROFILE SEE PROFILE

Available from: Michał Henryk Jamróz


Retrieved on: 11 October 2016
Fluid Phase Equilibria 152 Ž1998. 307–326

1
The urea–phenol Žs . systems
Małgorzata E. Jamroz ´ a, ) , Marcela Palczewska-Tulinska
´ a,
a a
Danuta Wyrzykowska-Stankiewicz , Andrzej M. Szafranski ´ , Jerzy Polaczek a ,
Jan Cz. Dobrowolski a,b,2 , Michał H. Jamroz ´ a , Aleksander P. Mazurek b
a
Industrial Chemistry Research Institute, 8, Rydygiera-Street, 01-793 Warsaw, Poland
b
Drug Institute, 30 r34, Chełmska-Street, 00-725 Warsaw, Poland
Received 18 June 1997; accepted 19 November 1997

Abstract

Solid–liquid equilibrium has been studied in the binary systems formed by urea with phenol, o-, m-,
p-cresol, 2,3-, 2,4-, 2,5-, 3,4-, 3,5-xylenol, and p-cumylphenol by the method involving crystal disappearance
temperature measurements and by differential scanning calorimetry. The congruently melting Ž1:2. complex
exists only in the urea–phenol system. The miscibility gap was found in three of the systems viz., urea–2,5-
xylenol, –2,6-xylenol, and –p-cumylphenol. An incongruently melting complex andror plateau-like section of
the liquidus curve characterize the SLE diagrams of the other systems. The enthalpy of formation of the
hydrogen bonded 1:1 complexes in the binary urea–phenol, o-, m-, p-cresol, and p-cumylphenol mixtures
dissolved in 1,2-dichloroethane was evaluated via temperature IR measurements. The enthalpy was found to be
about y30 kJrmol, except for the urea–o-cresol Ž1:1. complex Žy25 kJrmol.. The ab initio HFr6-31G))
calculations of stabilization energies for urea–phenol complexes bonded through a linear hydrogen bond show
very good agreement with the IR derived D H-values. q 1998 Elsevier Science B.V. All rights reserved.

Keywords: Ab initio; Enthalpy; Heat capacity; Hydrogen bond; Spectroscopy; Solid–liquid equilibria

1. Introduction

Although urea–phenolŽs. systems have been studied for more than a century, hydrogen bonding in
those systems has been very little looked into. Crystal structure has been established by the X-ray
diffraction method in four systems only, viz., urea–1,4-benzene diol Ž1:1. w1x, urea–1,3-benzene diol
Ž1:1. w2x, urea–phenol Ž1:2. w3x, and urea–4-nitrophenol Ž1:2. Žcited in Ref. w4x..

)
Corresponding author.
1
Paper presented at the International Conference on Applied Physical Chemistry, Warsaw, 13–15 November 1996.
2
Also corresponding author.

0378-3812r98r$ - see front matter q 1998 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 8 - 3 8 1 2 Ž 9 8 . 0 0 2 0 6 - 4
308 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

Solid–liquid equilibrium ŽSLE. has been studied most extensively in the urea–phenol system
w5–22x. The liquidus curves in the urea–methyl phenol systems have been studied less extensively
w23–26x, and those in the urea–dimethyl phenol systems have not been reported so far. SLE in ternary
systems involving urea and phenol have been extensively reviewed w27x.
The urea–1,3,5-trinitrophenol complex as well as a number of substituted urea–substituted phenol
complexes have been obtained and characterized w28x.
The IR spectrum of the urea–phenol system in the solid phase has been discussed w20x and IR
studies on the urea–phenol and urea–isomeric cresol systems in an inert solvent have been reported
ŽRefs. w22,29x, revision Ref. w30x.. The urea–p-nitrophenol system has been UV studied in water w31x.
A theoretical ab initio study on possible urea–phenol complexes Ž 1:1 and 1:2. has been carried out
w30,32x and the roles of the urea and the phenol molecules in the hydrogen bond formation, the
geometry of the complexŽ es., and the strength of the hydrogen bondŽs. in the urea–phenolŽs. systems
have been elucidated.
IR studies alone provide a deep insight but only rarely yield definitive answers to those questions.
Most bands that should be helpful in the analysis of H-bond formation are overlapped by other bands.
Actually, IR studies describe the H-bond geometry only as a bond type, i.e., single, double, cyclic,
etc. To be determined by IR spectroscopy, the H-bond strength Ž in terms of the enthalpy of complex
formation. requires the study to be carried out at several temperatures and in a solvent medium, and
the solvent does always perturb the complex formation equilibrium.
In keeping with previous results w20x, our preliminary IR studies on the urea–phenol solid phase
w27x showed urea’s C 5 O group and phenol’s OH group to be involved in the H-bond formation. The
IR spectrum of the molten urea–phenol Ž1:2. system was found to be a superposition of the spectra
recorded for molten urea and molten phenol, thereby bearing out our SLE evidence w22x for the
compound having a congruent melting point. However, it remained disputable whether or not the IR
revealed the presence of the N–H PPP OŽ H. hydrogen bonds existing in the crystal structure w3x. To
resolve this problem, a deuterated urea–phenol Ž 1:2. complex was prepared and its IR spectrum
recorded. Unfortunately, an intense band of phenol O–D stretching vibrations overlapped the N–H
stretching vibration bands and thus rendered interpretation of the spectra inconclusive. Two processes
were seen responsible for this situation: Ž i. deuterium could have been transferred from the urea-d 4 to
the phenol molecules, catalyzed by traces of water that is always present in such a hygroscopic
system; Žii. deuterium could have been directly transferred via the N–D PPP OŽH. hydrogen bond.
Thus, the problem of IR observation of the N–H PPP OŽH. hydrogen bond in the solid urea–phenol
Ž1:2. system has remained unresolved.
Therefore, it was deemed advisable to study the urea–phenol, urea–methyl phenol Ž o-, m-, and
p-. , and urea–4-Ž2-phenylisopropyl. phenol Ž i.e., p-cumylphenol. systems in the solvated state. The
purpose of this investigation was twofold: Ž i. to observe the N–H vibration bands and to establish the
type of the complex; Žii. to determine the enthalpy of formation of the complexes involved.

2. Experimental
2.1. Materials
Purest-grade urea Ž POCh, Gliwice, PL. , repeatedly recrystallized from distilled water and EtOH,
99.9 mol%, m.p. 132.48C ŽDSC.. Purest-grade phenol ŽMZRiP, Płock, PL. , dehydrated through a
´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 309

25-plate laboratory column, purity 99.9 mol% ŽGC. , b.p. 180.08Cr745–746 mm Hg. Pure-grade
¨
o-cresol ŽRutgerswerke, Germany. , distilled as above, 99.8 mol% Ž GC. , b.p. 189.5–189.98Cr745–746
mm Hg. Pure-grade m-cresol Ž Soyuzkhimexport, SU., distilled as above, 99.9 mol% Ž GC. , b.p.
201.3–201.48Cr760 mm Hg. Pure-grade p-cresol ŽSoyuzkhimexport, SU. , distilled as above, 99.9
mol% ŽGC., b.p. 200.08Cr754 mm Hg. Pure-grade p-cumylphenol ŽMZRiP, Płock, PL., fractionally
crystallized three times and then recrystallized from n-heptane, m.p. 73.08C ŽDSC. . Pure-grade
2,3-xylenol ŽUCB, Belgium., fractionally crystallized three times and distilled, b.p. 218.18Cr758 mm
Hg. Pure-grade 2,4-xylenol ŽFerak-Berlin, Germany. , crystallized from benzene, m.p. 24.18C ŽDSC..
Pure-grade 2,5-xylenol Ž Merck-Schuchardt, Germany. , crystallized from EtOH–Et 2 O, m.p. 74.98C
ŽDSC. . Pure-grade 2,6-xylenol Ž BDH, Great Britain. , distilled, b.p. 204.08Cr753 mm Hg. Pure-grade
3,4-xylenol ŽAldrich, USA. , fractionally crystallized three times and vacuum distilled, b.p. 140.08Cr25
mm Hg. Pure-grade 3,6-xylenol Ž Ferak-Berlin, Germany. , crystallized from benzene, m.p. 63.18C
ŽDSC. .
Pure-grade tetramethylurea Ž Merck, Germany. , distilled and dried over 3 A ˚ molecular sieves.
Spectroscopical reagent-grade 1,2-dichloroethane Merck, Germany , shaken with concentrated H 2 SO4
Ž .
and NaOH aq. solution, distilled and dried over CaH 2 or 3 A ˚ molecular sieves. Purest-grade carbon
tetrachloride Ž POCh, Gliwice, PL. , distilled and dried over CaH 2 .

2.2. Sample preparation

2.2.1. For SLE measurements


Synthetic mixtures of various compositions were prepared gravimetrically in small glass ampoules
and the ampoules were cooled to ca. 215 K, pumped down and sealed. The mixtures were melted,
cooled rapidly and heated progressively Žat a rate of 0.05 Krmin just before melting. , until the
crystals dissolved completely in the melt. Crystal disappearance temperatures were repeatedly read.
The average of three or more measurements was taken as the crystal-disappearance point of a given
mixture. The points were accurate to within "0.1 K. In this way, the liquidus boundaries were
ascertained for the systems examined over the whole concentration range.

2.2.2. For calorimetric measurements


Pure components of the system to be studied were weighed on a Cahn electric microbalance into a
DSC aluminum cup in desired proportions and then the cup was sealed by using a DSC sealing
assembly. The resulting synthetic mixture was melted for the first time in the DSC instrument and,
after its melting temperature had thus been established, the cup and its contents were cooled to prompt
crystallization of the specimen. Then the specimen was conditioned by being thermostated for a long
period of time Ž up to 120 h. at a temperature slightly lower than the melting point of the mixture
measured during the virgin run in the DSC instrument. A second run was carried out and the
specimen was cooled and reconditioned before a subsequent run was performed. The same condition-
ing procedure was applied before the determinations of the melting point and the enthalpies of melting
of pure substances.

2.2.3. For IR measurements


Complex-containing solutions in 1,2-dichloroethane were prepared by fusing well-known amounts
of an appropriate phenol and urea, and then dissolving the binary mixture in the solvent.
310 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

In the studies on 1:1 equilibria, the total concentrations of urea and phenol varied from 0.003 to
0.006 and from 0.07 to 0.10 molrdm3, respectively.
In the study of 1:2 equilibria, the total concentration of urea was around 0.05 molrdm3 and the
total concentration of phenol varied from 0.8 to 1.0 molrdm3.
For the study of the tetramethylurea–m-cresol system in CCl 4 , the total concentrations of
tetramethylurea and m-cresol were 0.011 and 0.0205 molrdm3, respectively.

3. Method and apparatus

3.1. The SLE measurements

Solubilities were determined by Alekseev’s method w33x. An original apparatus w34x was designed
and constructed. It consists of two major systems: Ži. the temperature controlled heatingrcooling
system and Žii. the shaking–mixing system. The first was employed in two versions according to the
experimental temperature range involved. Below 908C, the glass container was filled with a silicone
oil, surrounded by a water or an ethanol jacket connected to the thermo- or cryostat. Temperature was
controlled to within 0.01 K and measured with a Hewlett-Packard quartz thermometer installed inside
the container. Above 908C, the silicone oil was heated with a resistance wire wound around the
container. Temperature was controlled and measured with a PT-100 sensing element installed inside
the container.

3.2. Calorimetric measurements

The enthalpy of melting was determined by using a Perkin-Elmer DSC 1B scanning microcalorime-
ter w35x. For pure compounds, the classical procedure was used. For the complexes with incongruent
melting points, a special procedure was developed. Measurements were carried out for several
mixtures of compositions lying between that of the eutectic point and that of the end point of the
liquidus boundary produced by a given complex. In this case, the final result was obtained by
extrapolating the D H s f Ž x . relationship to the values corresponding to the presumed complex
composition. This procedure allowed also to estimate the hypothetical incongruent melting point of
the complex compound.
For pure compounds, DCp was evaluated as the difference of the liquid and solid heat capacity data
calculated at the compound’s melting point from the correlation equations adjusted to experimental
liquid and solid compound’s heat capacity data Ž in few cases, pure compound’s data were partially
aided by generalized correlations. . For the urea–phenol Ž 1:2. complex, DCp was established
experimentally. The value is higher Ž Table 2. than the sum of the DCp-values for pure phenol and
pure urea. Since DCp values are nonadditive nor are there any correlations of this property data with
other molecular property data, the DCp-values for urea–methyl phenol complexes, that could not be
determined experimentally, were assumed to be the same as the experimental value for the Ž 1:2.
urea–phenol complex. The same guiding rule was adopted to assume the DCp values for the
urea–dimethyl phenol complexes with incongruent melting points, except that due consideration was
given to the DCp-value of the pure dimethyl phenol that was the constituent of a given complex
compound and the actual DCp-values for pure urea was ignored as an additive term.
´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 311

3.3. IR spectra measurements

IR spectra were taken Ž"1 cmy1 . over the 3550–3150 cmy1 and 1800–1600 cmy1 ranges by
using a Perkin Elmer 580 IR spectrometer operated in conjunction with a PC computer. The spectra
were recorded at ca. 8–14 different temperatures covering the 300–330 K range by using two Karl
Zeiss ŽJena, Germany. thermostatic cells, one in either of the beams. The spectra of the solvent and
the dissolved phenol were compensated for. The cuvettes path lengths were 0.25 and 2.3 mm.
Temperature was sensed with iron–constantan thermocouples connected to digital voltmeters Ž"1 K. .
The thermocouples were calibrated for the solution temperature inside the cuvettes.

3.4. The determination of IR band parameters

IR spectra of urea in solvated complexes with phenols were fitted as a sum of the Gaussian and
Lorentzian functions by using the PASMO program w36x based on the least-squares optimization
procedure w37x. The positions, widths, and intensities of the bands were determined with the aid of the
band fitting routine. To verify the results obtained by the band fitting procedure, the urea–phenol
system was reanalyzed by the intensian method w38x.

3.5. The determination of the enthalpy of complex formation

The least-squares fit was made to the equation for the standard free Gibbs energy change:
D H y TD S s yRT lnŽ K ., which was conveniently written as:
R ln Ž K . s yD HrT q D S,
where D H and D S are respectively the enthalpy ŽkJrmol. and the entropy ŽJ moly1 Ky1 . of complex
formation, T is the absolute temperature Ž K ., R is the universal gas constant Ž8.314472 J moly1
Ky1 . and K is the equilibrium constant. The D H and D S values, and their errors at the 95%
confidential level, were calculated as the weighted means of the values found in two or more
independent measurements.
The constant of the hydrogen bond equilibrium:
X–H q Y s X–H PPP Y,
assumed to remain unaffected by any other equilibria, was described in terms of the molar ratios, x i .
The following expression was used:
K x s x X – H PPP Y rx X – H x Y s X–H PPP Y Ž X–H q Y q X–H PPP Y . r Ž X–H Y . ,
where brackets denote equilibrium concentrations.
Since the equilibrium concentration of phenol could hardly be determined when phenol was used in
a ten-fold excess over the urea concentration, the equilibrium constant was applied in the form of a
one-variable function, i.e., either the function of the urea, or of the complex equilibrium, concentra-
tion, for example:
K x s Ž cU y U .Ž c Ph q U . r U Ž c Ph y c U q U . .
The equilibrium urea and complex concentrations were determined by a special procedure. This
was necessary because the spectrum of the saturated urea solution in 1,2-dichloroethane used as the
312 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

standard was quite different from the spectra obtained for the equilibrium systems, where compensa-
tion of the solvent produced the nonlinear baseline varying from one sample to another and the
observed bands overlapped each other. Moreover, with the hand-made cuvettes of ca. 2.3 mm, the
path length was impossible to measure precisely by the interference method. Therefore, the absorptivi-
ties of the analyzed IR bands were determined by using the following scheme.
Assume there is only one equilibrium observed in the system. Then, the sum of the equilibrium
concentrations of the two components is constant and independent from temperature:
c1Ž T . q c 2 Ž T . s c,
where c1Ž T . and c 2 Ž T . are respectively the appropriate equilibrium concentrations of free urea and
the complex and c is the analytical urea concentration.
Assume also the Lambert–Beer’s law, the law of additivity of absorption bands, to hold for the
measured IR bands and the absorptivity to be stable with temperature. Let the cuvette path length be
temperature-independent. Then,
A1Ž T . ra1 b q A 2 Ž T . ra2 b s c,
where A i Ž T . is the absorbance of the i-th component at temperature T, a i is the band absorptivity and
b is the path length, all interrelated with concentration by Beer–Lambert’s law. If rewritten as:
A1Ž T . s ca1 b q A 2 Ž T . a1 bra2 b,
the expression yields a straight line of the type y s Sx q I, where S is the slope and I the intercept,
whose coefficients can be adjusted by the least squares routine. Then:
a1 b s Irc and a 2 b s IrSc
The absorptivity errors are found from the standard deviations of the slope S and the intercept I at the
adopted confidence level.

3.6. Theoretical method

The ab initio Hartree–Fock Ž HF. w39x calculations were carried out by using the Gaussian 94
system of programs w40x and the split valence 6-31G)) basis set w41,42x. Berny’s optimization
routine was used for molecular optimizations. Two kinds of structures of the complexes were
calculated: Ži. cyclic, the most stable in vacuum; and Žii. linear, used to model experimental
arrangements in 1,2-dichloroethane solution.
In the calculations of the cyclic structures no constraints were used. In the calculations of the linear
structures, the urea molecule was constrained to remain in the plane perpendicular to the H–O–C
plane of the phenol molecule, otherwise, the resulting structure of the complex was invariably cyclic.
The HF stabilization energies were corrected for the basis set superposition error ŽBSSE. .

4. Results and discussion

4.1. Solid–liquid equilibria studies

The SLE curves of the systems studied are presented in Figs. 1–11.
´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 313

Fig. 1. Liquidus curve of the phenol–urea system.

4.2. The phenol–urea system


So far the reported investigations were inconclusive as regards the congruency of the phenol–urea
Ž2:1. complex. Our investigations allowed to establish unambiguously that the 2:1 complex melts
congruently at 60.458C. The liquidus curve ŽFig. 1. in the vicinity to the congruent melting point is
almost flat, thus, indicating the complex formation equilibrium to be only moderately shifted toward
the complex. This conclusion was confirmed by the temperature IR studies, i.e., at temperatures above
the melting point of the 2:1 complex, the IR spectrum of the system is a superposition of the spectra
of the pure melted components of the complex.
4.3. The o-cresol–urea, m-cresol–urea and p-cresol–urea systems
In each system, the liquidus curve ŽFigs. 2–4. exhibits the existence of a complex with the
incongruent melting point. The length of the liquidus branch corresponding to complex formation
decreases from the m-cresol–urea through o-cresol–urea to the p-cresol–urea system. In the
o-cresol–urea and the m-cresol–urea systems, the composition of the hypothetical complexes is
presumably 2:1; for the p-cresol–urea the composition of the complex is hardly possible to establish.
Interestingly, in the m-cresol–urea, and also in the o-cresol–urea system, the urea-rich branch of
the liquidus curve approaches the urea melting point in a way suggestive of demixing in the liquid
phase.
4.4. The p-cumylphenol–urea system
The system is a simple eutectic type Ž Fig. 5. with a large miscibility gap extending over the
composition range x u s 0.59–0.95, which most likely arises on account of a very significant sterical
314 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

Fig. 2. Liquidus curve of the o-cresol–urea system.

Fig. 3. Liquidus curve of the m-cresol–urea system.


´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 315

Fig. 4. Liquidus curve of the p-cresol–urea system.

Fig. 5. Liquidus curve of the p-cumylphenol–urea system.


316 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

Fig. 6. Liquidus curve of the 2,3-xylenol–urea system.

hindrance due to the large isopropylbenzene substituent in the para position that disables the
formation of a crystalline p-cumylphenol–urea complex.

4.5. The 2,3-xylenol–urea and 3,4-xylenol–urea systems

The liquidus curves of the systems ŽFigs. 6 and 7. are very similar with the eutectic point on the
xylenol-rich side and a local minimum-like point near x u s 0.95. The reasons for the shape of the
curve at compositions rich in urea can be two-fold: first, the existence of a complex with the
congruent melting point and a hypothetical composition somewhere between 1:3 and 1:6. However,
such urea–phenol complexes are unlikely to occur, because the phenol can act as a donor of one
proton and acceptor of a maximum of two protons on the oxygen atom and one on the phenyl ring.
Secondly, the presence of the miscibility gap in the liquid phase for this range of compositions and
temperatures is more likely to occur, because two other xylenol–urea systems exhibit miscibility gaps
over analogous composition ranges.
The molecules of the two phenols, 2,3-xylenol and 3,4-xylenol, have one feature in common, viz.,
their two methyl groups are attached in either of the molecules to adjacent carbon atoms at the ring.

4.6. The 2,5-xylenol–urea and 2,6-xylenol–urea systems

Both systems have eutectic points and a miscibility gap ŽFigs. 8 and 9.. In the first and the second
systems, the gaps extend over the urea mole fraction intervals 0.92–0.67 and 0.95–0.34, respectively.
´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 317

Fig. 7. Liquidus curve of the 3,4-xylenol–urea system.

Fig. 8. Liquidus curve of the 2,5-xylenol–urea system.


318 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

Fig. 9. Liquidus curve of the 2,6-xylenol–urea system.

The reason for demixing in the 2,6-xylenol–urea system is quite obvious. The xylenol’s two methyl
substituents situated in the closest vicinity to the OH group strongly perturb the formation of
hydrogen bond nets. However, in the 2,5-xylenol–urea system the reason is not so evident.

4.7. The 2,4-xylenol–urea and 3,5-xylenol–urea system

These are the systems with eutectic points; each has a liquidus branch corresponding to an
incongruently melting complex and a short branch suggesting the presence of a miscibility gap in the
liquid phase Ž Figs. 10 and 11.. One can also suppose that, especially for the 3,5-xylenol–urea system
at high urea concentrations, there exists another eutectic point and that the large nearly flat segment of
the curve corresponds to another complex formation equilibrium.
A common structural feature of the 2,4-xylenol and the 3,5-xylenol molecules is that the two
methyl groups are separated from each other by one C–H group at the ring.

4.8. Thermodynamic interpretation of SLE

The thermodynamic interpretation of SLE is similar to that of VLE phase diagrams, i.e., a
correspondence is adopted between the liquid-phase activity coefficient the VLE case and the g L of
the SLE case. The crystal structures of organic compounds, usually incompatible enough to allow
zero solid solubility to be assumed Ž actually, a few tenths of a mol% solid solubility cannot be ruled
out as a possibility, at least on one side of the phase diagram. , offer a considerable simplification in
the calculations Ž g S s 1.. Since the composition uniformity of the solid phase is always a matter of
´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 319

Fig. 10. Liquidus curve of the 2,4-xylenol–urea system.

Fig. 11. Liquidus curve of the 3,5-xylenol–urea system.


320 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

concern, and diffusion rates of molecules are dramatically low, resort was taken in this study in
prolonged heating of the binary samples at temperatures very close to their melting points to secure
the solid phase as uniform as possible.
So far organic SLE phase diagrams have been correlated in terms of g L model equations much less
frequently than VLE or LLE. One reason is the lack of a general consistency test to assay the validity
of experimental data. Few attempts w43–46x, even if successful, are only partial solutions of the
problem. Another point of difference from VLE is that, in SLE, negative deviations from the
ideal-solution behavior are extremely rare.
The analytical expressions w46x adopted in this study to correlate g L in the SLE systems included
the Redlich–Kister expansions with two and with three adjustable constants Ž RK-2; RK-3. , the van
Laar equation Žv-L. with two adjustable constants and the NRTL equations with the parameter a
varied from 0.1 to 0.35. Results w47x of preliminary calculations, summarized as goodness-of-fit data,
are listed in Table 1. The data show the Redlich–Kister expansions to be generally adequate and in
some systems even prevail over the more complex g L model equations.
Further thermodynamic analysis w47x of the systems investigated was based on the principles of
calculation of phase diagrams w48x from the thermodynamic properties of the phases.

4.9. Calorimetric studies

The melting point Žm.p.. , enthalpies of fusion Ž D Hf . and heat capacity Ž DCp . for pure, congru-
ently, and incongruently melting complexes are collected in Table 2.
The largest Ž absolute. value of the fusion enthalpy was found for the urea–phenol complex, the
only one with a congruent melting point. In the urea–cresol systems, the lowest fusion enthalpy was

Table 1
The goodness of fit to selected model equations
Liquidusa RK-2 RK-3 v-L NRTL a s 0.3
Phenol– ŽPh–U complex. 4.6 2.05 8.4 5.6 b
ŽPh–U Complex. –Urea 17.6 14.0 13.8 4.5
o-Cresol– Ž o-C–U complex. 3.6 3.7 3.5 4.2
Ž o-C–U complex. –Urea 15.9 16.1 23.7 15.0
m-Cresol– Ž m-C–U complex. 6.4 4.7 6.1 6.2
Ž m-C–U complex. –Urea 9.7 8.6 – 9.2
p-Cresol– Ž p-C–U complex. 5.9 4.2 10.1 11.7 b
Ž p-C–U complex. –Urea 9.5 8.9 – 13.8 b
2,3-Xylenol–Urea 17.8 12.9 – 10.50 b
3,4-Xylenol–Urea 11.3 11.5 - 16.4
2,5-Xylenol– Ž2,5-X–U complex. 25.6 22.2 - y
2,6-Xylenol–Urea 13.9 7.8 18.4 15.1
2,4-Xylenol– Ž2,4-X–U complex. 12.4 9.6 12.8 14.8
Ž2,4-X–U complex. –Urea 13.5 13.2 y 12.4 b
3,5-Xylenol– Ž3,5-X–U complex. 7.4 3.2 y 8.1
Ž3,5-X–U complex. –Urea 13.5 11.5 y y
a
The names of compounds, e.g., ‘complex Ph–U–Urea’ ‘Phenol–complex Ph–U’, delimit the liquidus segment correlated.
b
a s 0.1.
´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 321

Table 2
The values of melting point Žm.p.., enthalpy of fusion Ž D Hf . and heat capacity difference Ž DCp . for urea, series of phenolŽs.
and urea–phenolŽs. complexes
Substance or complex Melting point Ž8C. Enthalpy of fusion Žcalrmol. Heat capacity difference Žcalrmol K.
Phenol 41.00 y2750.0 15.5
Urea–phenol 61.05 y14,300.0 22.0
o-Cresol 30.90 y3800.0 18.0
Urea– o-cresol 61.50 y3550.0 20.0
m-Cresol 12.10 y2550.0 16.7
Urea– m-cresol 73.00 y4000.0 22.0
p-Cresol 34.40 y2050.0 17.3
Urea– p-cresol 31.00 y4050.0 22.0
2,3-Xylenol 72.60 y4850.0 13.5
2,4-Xylenol 24.10 y3050.0 13.7
Urea–2,4-xylenol 51.00 y3300.0 20.0
2,5-Xylenol 74.90 y3300.0 8.0
2,6-Xylenol 45.40 y4500.0 8.0
3,4-Xylenol 65.30 y4550.0 8.6
3,5-Xylenol 63.10 y4150.0 8.0
Urea–3,5-xylenol 66.00 y3750.0 20.0
p-Cumylphenol 73.00 y5450.0 28.0
Urea 132.00 y3250.0 1.75

found in the urea–o-cresol system, where sterical hindrance is the largest. In the remaining
urea–cresol systems, the values are about the same. For the urea–xylenol systems, the estimation of
the fusion enthalpy was possible for only two systems, urea–2,4-xylenol and urea–3,5-xylenol, and
the lower value characterized the system with the larger sterical hindrance.

4.10. IR and theoretical inÕestigations

IR studies of H-bonded urea complexes in the solution need a nonpolar or a slightly polar solvent
transparent for IR radiation in the region where the analyzed bands appear. The quest for such a
solvent developed into a separate study w30,49x. 1,2-Dichloroethane was found to be the most
convenient solvent for the study.
Only three bands of urea dissolved in 1,2-dichloroethane have been observed, viz., the ones
corresponding to the asymmetric NH 2 stretching vibrations n ŽNH 2 . as , to the symmetric NH 2
stretching vibrations n ŽNH 2 . s , and to the C5O stretching vibrations n ŽC5O., positioned at 3515,
3407 and 1700 cmy1, respectively. In the presence of phenolŽ s. , the positions of the NH 2 stretching
vibration bands remain unchanged, and a new band appears at 1680 cmy1. The intensity of the new
band increases with the increasing phenol Ž cresol. concentration and decreases with increasing
temperature, while the intensity of the NH 2 bands remains unaffected. The only change in the NH 2
stretching vibration region is the change of the background shape which increases with an increase in
phenol concentration and decreases with an increase in temperature. Therefore, the band at 1680
cmy1 was assigned to the n ŽC5O. stretching vibration band of the urea–phenol Žor urea–cresol,
urea–p-cumylphenol. Ž1:1. complex and the background in the NH 2 stretching vibration region was
assumed to be due to a broad n ŽO–H. stretching vibrations band of a given complex. Such a spectral
322 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

pattern indicates that the Ž1:1. complex in the 1,2-dichloroethane solution is formed via the
O–H PPP O5C hydrogen bond interactions, and rules out the participation of the hypothetical
N–H PPP OŽH. hydrogen bond interactions.
A similar picture is obtained when the concentration of phenol Ž or cresol, p-cumylphenol. is
increased to ca. 1.0 molrdm3. A second new n ŽC5O. stretching vibrations band appears at 1667
cmy1 and changes with the phenol Žor cresol, p-cumylphenol. concentration and with temperature
likewise the 1680 cmy1 band does. However, the band at 1667 cmy1 overlaps with the band at 1680
cmy1 and the two maxima cannot be distinguished. The NH 2 stretching vibrations bands vary neither
with temperature nor with the phenol concentration. The background in the NH 2 stretching vibrations
region is more band-shaped and increases with the increasing phenol concentration and with the
decreasing temperature. The band at 1667 cmy1 was assigned to the n ŽC5O. stretching vibration
band of the urea–phenol Žor urea–cresol, urea–p-cumylphenol. Ž 1:2. complex and the background in
the NH 2 stretching vibration region was assumed to be due to broad n ŽO–H. stretching vibrations
band of a given complex. Such a spectral pattern indicates that the urea–phenol Ž or urea–cresol,
urea–p-cumylphenol. Ž1:2. complex in the 1,2-dichloroethane solution is formed by the two O–
H PPP O5C hydrogen bond interactions whereby participation of the hypothetical N–H PPP OŽ H.
hydrogen bond interactions is ruled out as a possibility. At that concentration of phenols the
O–H PPP OŽH. hydrogen bond interactions do play a significant role but, because the spectrum of a
given phenol is compensated for, they are hardly distinguished Žby the IR method. from the
O–H PPP O5C hydrogen bond interactions.
Thus, the qualitative analysis of the IR spectra of urea–phenolŽ s. systems in the 1,2-dichloroethane
solution strongly suggests that only O–H PPP O5C hydrogen bonds are important in the complex
formation process. This conclusion seems to be inconsistent with the X-ray w3x and theoretical w32x
results. In the ab initio studies, the energetics of several types of Ž1:1. and of two types of Ž 1:2.
urea–phenol complexes has been considered. In both cases, the structures with cyclic hydrogen
bonds, where two nonlinear O–H PPP O5C and N–H PPP OŽH. hydrogen bonds are supporting the
formation of a six-membered ringŽs. , appeared to be the most stable. However, the most stable
complex with a single hydrogen bond turned out to be that with the linear O–H PPP O5C hydrogen
bond.
Therefore, the enthalpies Ž D H . and entropies Ž D S . of hydrogen bond formation have been
reevaluated for the urea–phenol, urea–cresol Ž o-, m- and p-. and urea–p-cumylphenol complexes in
1,2-dichloroethane solutions by analyzing the changes of the C5O stretching vibration bands
intensities. The resulting D H and D S values are summarized in Table 3. The enthalpy of ca. 30
kJrmol is typical for a single O–H PPP O5C hydrogen bond interaction in phenol–amide or
phenol–urea derivative interactions w50,51x. The urea–phenol systems have identical D H values Žat
the 0.95 confidence level. . This is not surprising in view of the similarities in the molecular structures
and the values of some physico-chemical properties, e.g., p K a. The urea–o-cresol system is an
exception, however. Reinvestigation of this system gave a revised value of D H w30x higher by 5
kJrmol, the former value having been significantly perturbed by the formation of the 1:2 complexes.
In the case of the o-cresol–urea system, the lower stabilization of the complex is probably due to
sterical hindrance by the ortho-methyl substituent.
In order to confirm definitively the conclusion about the single O–H PPP O5C hydrogen bond
interaction in the systems, we observed also the variation of the C5O stretching vibrations band
region of the tetramethylurea–m-cresol system dissolved in CCl 4 . In this case, there are no N–H
´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 323

Table 3
The enthalpy and entropy of 1:1 urea–phenolŽs. hydrogen-bonded complex formation
Complex IR method in 1,2-dichloroethane solution RHFr6-31G)) calculations ŽBSSE included.
y1 y1 .
y D H ŽkJrmol. y D S ŽJrmol K y D H ŽkJrmol. linear y D H ŽkJrmol. cyclic
Urea–phenol 30.0"3.0 85.0"8.0 31.44 37.52
29.0"2.5a 85.0"8.0 a
Urea– o-cresol 25.0"3.0 70.0"10.0 26.43 b 35.77 b
30.94 c 36.58 c
Urea– m-cresol 30.0"3.5 85.0"10.0 30.66 b 36.75 b
30.71c 36.98 c
Urea– p-cresol 30.0"3.0 85.0"9.0 30.54 36.73
Urea– p-cumylphenol 32.0"3.0 160.0"15.0 y y
a
The value obtained using the intensian method w38x.
b
The methyl substituent is positioned in the vicinity of the urea molecule.
c
The methyl substituent is positioned far from the urea molecule.

groups to interact with the phenol’s oxygen-atom. Moreover, the C5O group of tetramethylurea is a
fair model of the C5O group in the unsubstituted urea. Technical reasons made the change of solvent
necessary. The C5O stretching vibration band of tetramethylurea in 1,2-dichloroethane solution
occurs at 1635 cmy1 and complex formation make it shifted toward lower wavenumbers where it
overlaps with the intense phenol ring vibration band. The enthalpy of formation of the tetramethy-
lurea–m-cresol hydrogen bond complex in CCl 4 solution is equal to y37 " 5 kJrmol and is even
lower than that for the urea–phenol systems. This fact means that in each case, urea–phenolŽ s. and
tetramethylurea–m-cresol systems, the same type of single O–H PPP O5C hydrogen bond interaction
is involved and the main differences are due to solute–solvent interactions.
The ab initio HFr6-31G)) stabilization energies ŽTable 3. calculated for the linear arrangements
of the complexes show very good agreement with the IR derived D H-values. They also confirm that,
in the ortho position, the methyl group does, whereas in other positions it does not perturb the
equilibrium. The stabilization energies calculated for the cyclic complexes show the NH PPP OŽH.
interactions to stabilize the complex by an additional amount of about 6 kJrmol. Thus, the overall
solvent–solute interaction energy should exceed that value.
The final conclusion about the type of the Ž1:1. complex is that the urea–phenol, urea–cresol and
urea–p-cumylphenol Ž 1:1. complexes are formed on account of the single O–H PPP O5C hydrogen
bond interaction and that the N–H urea groups do not stabilize the system in the 1,2-dichloroethane
medium.
Let us finally comment on the situation in more concentrated urea–phenols solutions where 1:2
complexes can be observed. Although the variation of the C5O stretching vibration band region with
temperature was measured, no reliable D H values were obtained for the systems, mainly because the
measured bands overlapped too much to allow maxima to be distinguished, and the band fitting
procedure gave no unique solution w37x. As a result, conflicting values of D H were obtained.
Nevertheless, qualitative conclusions can still be drawn. The 1:2 complexes are formed also on
account of the O–H PPP O5C hydrogen bond interactions and the N–H PPP OŽH. interaction are
rather unlikely to stabilize the systems.
324 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

5. Conclusions

The SLE diagrams have been measured for binary urea systems involving phenol, cresols,
p-cumylphenol and xylenols systems. The congruent melting point was found to occur certainly in the
urea–phenol system only.
The liquidus curves of all urea–cresol systems are similar in the sense that each system exhibits the
formation of an incongruent melting complex; in the –o and the –m-cresol systems the hypothetical
composition of the complex is 1:2.
The urea–p-cumylphenol system is similar to the urea–2,5- and urea–2,6-xylenol systems. In each
system, the liquidus curve is of the simple eutectic type with a more or less extensive miscibility gap
on the urea side of the diagram.
The urea–2,3- and urea–3,4-xylenol systems exhibit local minima and flattened humps on the
urea-rich side and form simple eutectics on the xylenol-rich side. The urea–2,4- and urea–3,5-xylenol
systems reveal the formation of an incongruent melting complex.
In the last-mentioned four systems and also in the urea–o- and urea–m-cresol systems the shape of
the curve at urea-rich compositions can be interpreted in two ways. One, suggesting the formation of
a congruent melting complex; and the other, supposing the existence of the miscibility gap in the
liquid phase at temperatures slightly exceeding that of hypothetical basis tie-line.
The urea–phenol, urea–cresols Ž o-, m- and p-. and urea–p-cumylphenol complexes in 1,2-dichlo-
roethane solution, both 1:1 and 1:2, are formed due to O–H PPP O5C hydrogen bonds.
In contrast to the crystalline urea–phenol 1:2 complex and to the results of quantum chemical
calculations, in 1,2-dichloroethane solution the NH 2 groups of urea do not participate in the formation
of a 1:1 or 1:2 complex with phenols.
The enthalpies of formation of 1:1 urea–phenol, urea–cresol, urea–p-cumylphenol complexes in
1,2-dichloroethane are y30 kJrmol, except for the urea–o-cresol complex for which it is equal to
y25 kJrmol.
The magnitude of the determined enthalpy of complex formation as well as the value of y37
kJrmol for the tetramethylurea–m-cresol 1:1 complex in CCl 4 , confirm the conclusion on the linear
O–H PPP O5C type of hydrogen bond present in the systems studied.
The ab initio HFr6-31G)) calculations of stabilization energies for urea–phenols complexes
bonded through a linear hydrogen bond show very good agreement with the IR derived D H-values.

6. List of symbols

A absorbance
a absorptivity
b path length
c analytical concentration
DCp heat capacity difference
DH enthalpy of formation
D Hf heat of fusion
I intercept of a straight line
K equilibrium constant
´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz 325

Kx equilibrium constant based on molar ratios


R universal gas constant
S slope of a straight line
DS entropy of formation
T absolute temperature
x molar ratio
wPx equilibrium concentration
Indices
1,2 components of the system
i current index

Acknowledgements

We wish to thank Mrs. Danuta Leciejewska for her able technical aid. This research was supported
by the Ministry of Chemistry Research Program CPBP 01.16 No. 03.05.01 Ž 1988. , by the Industrial
Chemistry Research Institute Ž 1990–1991. and by a grant from the State Committee for Scientific
Research Ž KBN. Ž 1991. .

References

w1x M.M. Mohamoud, S.C. Wallwork, Acta Crystallogr. Sect. B 31 Ž1975. 338.
w2x M. Pickering, R.W.H. Small, Acta Crystallogr. Sect. B 38 Ž1982. 3161.
w3x A.L. MacDonald, A. Murray, S. Townsley, P.R. Mallinson, Acta Crystallogr. Sect. C 43 Ž1987. 676–678.
w4x ´
V.G. Videnova-Adrabinska, The hydrogen bond as a design element of the crystal architecture. Crystal engineering.
From biology to materials. Scientific Papers of the Institute of Inorganic Chemistry and Technology of Rare Elements
of Technical University of Wrocław, No. 65, Monographs No. 32, Oficyna Wydawnicza Politechniki Wrocławskiej,
Wrocław, 1994.
w5x H. Eckenroth, Jahresber. Chem. Ž1886. 548.
w6x J.Ch. Philip, J. Chem. Soc. ŽLondon. 83 Ž1903. 814–834.
w7x R. Kremann, O. Rodinis, Monatsh. Chem. 27 Ž1906. 138.
w8x ¨
N.A. Pushin, D. Konig, Monatsh. Chem. 49 Ž1928. 75–82.
w9x N.A. Pushin, J.J. Rykovski, Z. Phys. Chem. A 161 Ž1932. 336–340.
w10x N.A. Pushin, J.J. Rykovski, Bull. Soc. Chim. Jugoslavie 3 Ž1932. 65.
w11x N.A. Pushin, J.J. Rykovski, Monatsh. Chem. Ž1932. 438.
w12x J.P. van der Hammen, Rec. Trav. Chim. 50 Ž1931. 347–350.
w13x C.A. Buchler, J.H. Wood, D.C. Hull, E.C. Erwin, J. Am. Chem. Soc. 54 Ž1932. 2398.
w14x K. Hrynkowski, F. Adamanis, Roczniki Chemii 14 Ž1934. 189.
w15x R. Cohen-Adad, Bull. Soc. Chem. Fr. 16 Ž1934. 824.
w16x D.E. Dionisev, N.Z. Rudenko, Zh. Obshch. Khim. 25 Ž1951. 990.
w17x A.G. Palobekov, I.I. Ilyasov, A.G. Bergman, Zh. Obshch. Khim. 34 Ž1964. 370.
w18x A.G. Palobekov, Zh. Obshch. Khim. 35 Ž1965. 939.
w19x A.G. Palobekov, A.G. Bergman, Zh. Obshch. Khim. 36 Ž1966. 1186.
w20x W.F. Chesnokov, I.M. Bokhovkin, Zh. Obshch. Khim. 36 Ž1966. 1549–1553.
w21x V.G. Vanyarkhoand, V.P. Zlomanov, Izv. Akad. Nauk. SSSR Neorg. Mater. 5 Ž1969. 1699.
326 ´ et al.r Fluid Phase Equilibria 152 (1998) 307–326
M.E. Jamroz

w22x J. Dobrowolski, M.H. Jamroz,´ M. Jamroz, ´ D. Wyrzykowska-Stankiewicz, A.M. Szafranski,


´ J. Polaczek, J. Mol. Struct.
175 Ž1988. 227–232.
w23x R. Kremann, Sitzungsber. Akad. Eiss. Wien. Math.-Naturwiss KL 116 ŽIIb. Ž1907. 1031.
w24x R. Kremann, Monatsh. Chem. 28 Ž1907. 1125.
w25x N.A. Pushin, L. Sladovic, J. Chem. Soc. ŽLondon. Ž1928. 2474–2481.
w26x I.M. Bokhovkin, Yu.I. Bokhovkina, E.O. Vitman, Zh. Obshch. Khim. 37 Ž1967. 31.
w27x J. Polaczek, T. Kreczmer, T. Tecza, M. Jamroz, ´ A. Krzeslak,
´ ´
A. Szafranski, D. Wyrzykowska-Stankiewicz, M.H.
´ J. Dobrowolski, Z. Lisicki, ICRI Internal Rep. 27 Ž1986. 1–63.
Jamroz,
w28x J.E. Barry, M. Finkelstein, G.A. Hutchins, S.D. Ross, Tetrahedron 39 Ž1983. 2151–2156.
w29x J.Cz. Dobrowolski, J. Mol. Struct. 219 Ž1990. 233–238.
w30x J.Cz. Dobrowolski, Ph.D. Thesis, Industrial Chemistry Research Institute, Warsaw, 1992.
w31x P.K. Das Gupta, S.P. Moulik, A.R. Das, Bull. Chem. Soc. Jpn. 64 Ž1991. 3156–3159.
w32x J.Cz. Dobrowolski, G. Karpinska,
´ A.P. Mazurek, Pol. J. Chem. 69 Ž1995. 1066–1070.
w33x V.F. Alekseev, Ann. Phys. Chim. 28 Ž1886. 305.
w34x A. Tulinski,
´ ´ Pol. Pat. Ž1984. 133 084.
J. Polaczek, B. Olejniczak, M.E. Jamroz,
w35x D. Wyrzykowska-Stankiewicz, A.M. Szafranski, ´ Wiss. Z. TH Leuna-Merseburg 17 Ž1975. 265–271.
w36x M.H. Jamroz,´ PASMO—the program fitting the spectrum in terms of superposition of different kind of functions,
Warsaw, 1990.
w37x W.F. Maddams, Appl. Spectrosc. 34 Ž1980. 245.
w38x J.Cz. Dobrowolski, G.J. Strzemecki, M.H. Jamroz, ´ Chemometr. Intell. Lab. Syst. 15 Ž1992. 39–50.
w39x C.C.J. Roothan, Rev. Mod. Phys. 23 Ž1951. 69.
w40x M.J. Frisch, G.W. Trucks, H.B. Schlegel, P.M.W. Gill, B.G. Johnson, M.A. Robb, J.R. Cheeseman, T.A. Keith, G.A.
Petersson, J.A. Montgomery, K. Raghavachari, M.A. Al-Laham, V.G. Zakrzewski, J.V. Ortiz, J.B. Foresman, J.
Cioslowski, B.B. Stefanov, A. Nanayakkara, M. Challacombe, C.Y. Peng, P.Y. Ayala, W. Chen, M.W. Wong, J.L.
Andres, E.S. Replogle, R. Gomperts, R.L. Martin, D.J. Fox, J.S. Binkley, D.J. Defrees, J. Baker, J.P. Stewart, M.
Head-Gordon, C. Gonzales, J.A. Pople, Gaussian 94 ŽRevision D.4.. Gaussian, Pittsburgh, PA, 1995.
w41x G.A. Petersson, A. Benett, T.G. Tensfeldt, M.A. Al-Laham, W.A. Shirley, J. Mantzaris, J. Chem. Phys. 89 Ž1988.
2193.
w42x G.A. Petersson, M.A. Al-Laham, J. Chem. Phys. 94 Ž1991. 6081.
w43x H.R. Null, AIChE J. 11 Ž1965. 780.
w44x H.R. Null, Chem. Eng. Prog. Symp. 81 Ž63. Ž1967. 52.
w45x A.M. Szafranski,
´ paper presented at the Polish Academy of Sciences Colloquium, Warsaw, 1972.
w46x R.C. Reid, J.M. Prausnitz, T.K. Sherwood, The Properties of Gases and Liquids, 3rd edn., McGraw-Hill, 1977.
w47x M.E. Jamroz,
´ Activity Coefficients in the Urea–PhenolŽs. Systems, to be published.
w48x J. Sangster, P.K. Talley, C.W. Bale, A.D. Pelton, Can. J. Chem. Eng. 66 Ž1988. 881.
w49x J.Cz. Dobrowolski, M.H. Jamroz, ´ A.P. Mazurek, Vib. Spectrosc. 8 Ž1994. 53–60.
w50x C. Dorval, Th. Zeegers-Hyskens, Tetrahedron Lett. 43 Ž1972. 4457.
w51x T. Gramstad, W.J. Fuglevik, Acta Chem. Scand. 16 Ž1962. 1369.

Potrebbero piacerti anche