You are on page 1of 7

Journal of Molecular Liquids 231 (2017) 142–148

Contents lists available at ScienceDirect

Journal of Molecular Liquids

journal homepage: www.elsevier.com/locate/molliq

Apparent molal volume and viscosity values for a new synthesized


diazoted resorcin[4]arene in DMSO at several temperatures
Mauricio Maldonado a,⁎, Edilma Sanabria b, Belén Batanero c, Miguel Ángel Esteso d,⁎
a
Departamento de Química, Facultad de Ciencias, Universidad Nacional de Colombia, Sede Bogotá, Cr. 30 No. 45-03, Bogotá 111321, Colombia
b
Departamento de Química, Universidad de los Andes, Cr. 1 No. 18A 10, Bogotá 111711, Colombia
c
Department of Organic Chemistry, Universidad de Alcalá, Alcalá de Henares 28871, Spain
d
U.D. Química Física, Universidad de Alcalá, Alcalá de Henares 28871, Spain

a r t i c l e i n f o a b s t r a c t

Article history: A new diazoted resorcin[4]arene was synthesized and characterized by spectral techniques. The densities and
Received 30 November 2016 viscosities of their solutions in dimethylsulfoxide (DMSO) were measured at temperatures between 293.15
Received in revised form 25 January 2017 and 313.15 K, over a range of concentrations from (0.0058 to 0.023) (mol kg−1) at atmospheric pressure. By
Accepted 27 January 2017
using these experimental data, some thermodynamic functions were derived to contribute to the understanding
Available online 31 January 2017
of the intermolecular interactions taking place in solution. The results obtained from the volumes study indicate
Keywords:
the presence of strong solute-solute interactions. The study of viscosities suggests the existence of strong solute-
Diazoted resorcin[4]arenes solvent interactions. The above result was confirmed by calculating the Gibbs free energies of activation, per
p-(3-carboxyphenylazo)propylresorcin[4]arene mole, of viscous flow of both the solvent and the solute. The flow parameter values found indicate an activated
(APRA) state less organized compared with the ground state and a process driven by the enthalpy.
DMSO © 2017 Published by Elsevier B.V.
Apparent molal volumes
Viscosities

1. Introduction temperatures. The results are interpreted in terms of solute-solute,


solvent-solvent and solute-solvent interactions.
Azo dyes have an important role in the chemical industry and are of
interest for different fields of study, such as organic chemistry, analytical 2. Experimental
chemistry and physical chemistry. Their applications included dyeing of
fibers [1], sensors for molecules and ions [2–4], metallochromic indica- 2.1. Materials
tors [5], solar cells [6], photochromic materials [7], food additives [8],
therapeutics agents [9], among other applications. The azo dyes derived In Table 1, source, purity, quantification method and CAS number of
from resorcin[4]arenes constitute an interesting type of compounds the chemicals used are shown. They were used as supplied, without
since they have several conformations in solution and solid state [10– further purification.
12]. In the literature, few studies are found about the synthesis of Tetrapropylresorcin[4]arene (PRA) and p-(3-
azoresorcin[4]arene [13,14]. The results of these studies indicate that carboxyphenylazo)propylresorcin[4]arene (APRA) were prepared accord-
the best synthetic route is the diazotization in the free position of the ing to procedures reported in the literature [15]; their purity was deter-
annular ring between hydroxyl groups. The synthesis and character- mined by HPLC as better than 98%. The resorcin[4]arenes were kept in
ization of new azo dyes provides guidance for further studies. In this dark bottles, to protect them from sunlight, over freshly activated silica
work, the synthesis of a new azoresorcin[4]arene was carried out by gel. Their yields are referred to isolated product after purification. DMSO
diazotization of 3-aminobenzoic acid and subsequent coupling with was stored over 3 Å molecular sieve and was used without further
tetrapropylresorcin[4]arene. The compound obtained was purified purification.
and characterized and its solution properties were studied in
dimethylsulfoxide (DMSO) at several temperatures in the range 2.2. Analytical characterization studies
(293.15 to 313.15) K in order to obtain information about these
macrocyclic systems covering both the standard and the physiological The characterization of the synthesized compounds was done by FT-
IR, 1H‐NMR, 13C‐NMR, thermogravimetric analysis and mass spectros-
⁎ Corresponding authors.
copy. FT-IR spectra were recorded on KBr disc, using a Thermo Nicolet
E-mail addresses: mmaldonadov@unal.edu.co (M. Maldonado), miguel.esteso@uah.es IS10 spectrophotometer. The 1H‐NMR and 13C‐NMR spectra were re-
(M.Á. Esteso). corded on a Varian spectrometer Unity 300 apparatus. The chemical

http://dx.doi.org/10.1016/j.molliq.2017.01.093
0167-7322/© 2017 Published by Elsevier B.V.
M. Maldonado et al. / Journal of Molecular Liquids 231 (2017) 142–148 143

Table 1 400 MHz, DMSO-d6): 0.91 (t, J = 8.0 Hz, 12H), 1.20 (m, 8H), 1.85 (m,
Source, purity and CAS number of the chemicals used. 8H), 4.45 (t, J = 7.6 Hz, 4H), 7.35 (s, 4H), 7.56–8.33 (m, 16H), 9.01 (s,
Chemical name CAS Source Mass Method 8H), 13.2 (s, 4H); 13C-NMR (δ, 100 MHz, DMSO-d6): 13.7, 20.5, 29.3,
number fraction 55.1, 115.6, 116.0, 117.5, 119.6, 122.3, 125.1, 127.2, 129.6, 132.0, 141.7,
purity 166.3; MALDI-TOF-MS (4-nitroaniline): calcd. for C68H64N8O16: m/
DMSO 67-68-5 Fluka ≥0.99 z = 1249.280 [M]+; found: m/z = 1272.564 [M + Na]+.
Purum The values measured for 1H-NMR in DMSO show that only one con-
Resorcinol 108-46-3 Merck 0.99
formational isomer was obtained, the crown isomer.
Butanal 123-72-8 Merck 0.99
Hydrochloric acid 7647-01-0 Merck 0.37
Sodium nitrite 7632-00-0 Merck 0.99 2.5. Determination of density
3-Aminobenzoic acid 99-05-8 Acros 0.99
Sodium hydroxide 1310-72-2 Merck 0.99
Tetrapropylcalix[4]resorcinarene – Synthesized ≥0.99 HPLC
Solutions were prepared by direct weighting of both the solute and
Azacalix[4]resorcinarene – Synthesized 0.98 HPLC the solvent by using a Mettler AE 240 analytical balance (accuracy of
1 ∙ 10−5 g in the range of interest, 0–40 g). The estimated uncertainty
concerning the solutions concentration was less than ±0.1%. The con-
shifts (δ), in ppm, were obtained in DMSO-d6 using TMS as an internal centration range studied was (0.0058 to 0.023) (mol kg−1). Densities
standard. The thermogravimetric analysis (TG) was performed by were measured with an Anton Paar DMA 5000 densimeter, which has
using a Netzsch STA 409 thermobalance with a sample weight of a sensitivity of 1 ∙ 10−6 (g cm−3) and accuracy of 5 ∙ 10−6 (g cm−3) in
28 mg, over a temperature range of 20–500 °C and a heating rate of the ranges of (0–90) °C of temperature and (0–1.0) MPa of pressure.
10 °C/min; the measurements were carried out in a nitrogen atmo- This instrument uses the vibrating U-tube measuring principle and is
sphere (flow rate: 16.66 mL/min) by using an alumina crucible. Molar equipped with a Peltier type thermostating unit, that ensures a temper-
mass was determined on a MALDI-TOF spectrometer (Bruker Daltonics) ature control of ±0.005 K. The densimeter was calibrated with dry air
using 4-nitroaniline as matrix for the desorption/ionization process. and purified water (milli-Q® quality) at each studied temperatures, ac-
cording to the recommendation given by the manufacturer.
2.3. Synthesis and characterization of PRA
2.6. Determination of viscosity
PRA was synthesized by condensation in acid medium of resorcinol
and butanaldehyde in a 1:1 water-ethanol mixture, following the liter-
Viscosities were measured by using an Ostwal type viscometer,
ature described method [15], with a yield of 86%. The values found for its
model Cannon-Fenske, calibrated with purified water (milli-Q® quali-
characterization were: IR (KBr, υ cm−1): 3314 (O\\H), 2957 (C\\H),
ty) at each temperature studied, with uncertainty of ±0.25% and efflux
1616 (C=C), 1194 (C\\O); 1H‐NMR (δ, 400 MHz, DMSO-d6): 0.90 (t,
times in the range (374–670) s in order to make irrelevant the kinetic
J = 8.0 Hz, 12H), 1.21 (m, 8H), 2.09 (q, J = 8.0 Hz, 8H), 4.23 (t, J =
energy correction (Hagenbach correction). The temperature control
8.0 Hz, 8H), 6.15 (s, 4H), 7.24 (s, 4H), 8.94 (s, 8H); 13C-NMR (δ,
was ±0.02 K. The viscometer constants were calculated with the help
100 MHz, DMSO-d6): 16.1, 25.3, 32.7, 49.2, 112.1, 127.2, 129.0, 151.1
of the Poiseuille equation:

2.4. Synthesis and characterization of APRA


η β
Diazotation and coupling reaction were done by following the liter- ¼ α− 2 ð1Þ
ρt t
ature described method (Scheme 1) [13]. A solution of NaNO2 (1 mmol)
in water (10.0 mL) cooled to 0 °C was drop wise added over a solution of
3-aminobenzoic acid (1 mmol) in water (15 mL) and concentrated HCl where η is the viscosity, α and β are the viscometer constants, and ρ and
(1 mL at 37%). The resulting mixture was slowly added over a solution of t are the density and the flow time, respectively, of a liquid of known
PRA (0.25 mmol) in 10 mL of NaOH 1 M solution. The reaction mixture viscosity (in this case water). These constants were obtained from the
was stirred for 2 h at a temperature below 5 °C. The red solid obtained plot of η/ρt against 1/t2 concerning literature data of pure water [16].
was filtered, washed with water and then dried at 50 °C for 24 h. The The viscometer constants were used to determine the absolute viscosity
yield was of 91%. values of solutions of APRA in DMSO. The viscosity values reported in
The characterization values obtained were: IR (KBr, υ cm−1): 3314 this work are the mean ones of at least three sets of measurements
(O\\H), 2957 (C\\H), 1616 (C=C), 1194 (C\\O); 1H-NMR (δ, (standard uncertainty of 0.005 mPa s).

Scheme 1. Coupling reaction of PRA and diazonium salt.


144 M. Maldonado et al. / Journal of Molecular Liquids 231 (2017) 142–148

3. Results and discussion with those available from the literature. A satisfactory agreement
between them is observed at each temperature.
3.1. Spectral data Experimental densities and apparent molal volumes together with
their respective standard deviations of the measurements are given in
Synthesis of PRA has been reported in the literature, while similar in- Table 3 for the solutions of APRA in DMSO at all temperatures studied.
formation about APRA has not been previously reported. Consequently, In all cases, the measurement uncertainty value (better than
in the case of PRA the results obtained from the spectral studies could be 0.7 cm3 mol−1) was calculated according to the law of propagating
compared with values reported in the literature for this compound, the variances [31].
whereas in the case of APRA such comparison had to be done by recur- Apparent molal volumes,Vφ, were calculated from density values by
ring to values for similar compounds [17–20]. In this way, the results for using the equation
PRA are in good agreement with those already reported in the literature
[20]. In relation to APRA, it shown an UVλmax (H2O) = 495 nm and its M 2 1000 ðρ−ρ0 Þ
Vφ ¼ − ð2Þ
molar mass, determined by high resolution MALDI-TOF-MS using 4- ρ mρρ0
nitroaniline as matrix, revealed the expected [M + Na]. The FT-IR spec-
trum showed absorptions from azo group (1044 cm−1 and 773 cm−1), where M2 is the molar mass of APRA, ρ is the density of its solution in
aromatic ring (1609 cm−1), alkyl chains (2929 cm−1) and hydroxyl DMSO, ρo is the density of the pure solvent and m denotes the molality
groups (3421 cm−1). The 1H‐NMR was recorded in DMSO-d6 and of the solution.
displayed characteristic signals of propyl chains (1.09, 1.60, and These apparent molal volume values were fitted against the molal
2.35 ppm), a methylene bridge fragment between the aromatic rings concentration, m, by means of a weighted linear regression to the sec-
(4.60 ppm), and the aromatic hydrogen of penta substituted resorcinol ond order polynomial equation [32]:
units (7.39 ppm).
V φ ¼ V 0φ þ Sv m þ Bv m2 ð3Þ
3.2. Thermal properties
where V0φ is the apparent molal volume of the solute at infinite dilution
To examine the thermal stability of APRA and the possible formation (equal to its partial molar volume at infinite dilution,V02), Sv is the exper-
of solvates, TG analysis was performed under the conditions given in imental slope (which has been related to the effect of the solute on the
section 2.2. The TG curve related to the thermal decomposition of solvent structure [33]) and Bv is an empirical parameter.
APRA is shown in Fig. 1. This indicates that the decomposition occurs The fittings for the different temperatures studied are shown in Fig.
near 250 °C; that is, this compound is stable below 250 °C. On the 2, and the Vφ∘, Sv, Bv values are listed in Table 4 at the interest
other hand, in the solvents used for the crystallization processes, forma- temperatures.
tion of solvates is not observed. This result contrast with previous stud- As it can be seen, Bv b 0 and Sv N 0, independently of the tempera-
ies for other resorcin[4]arene systems, which show that molecules, such ture. By taking into account that a positive value of SV is associated
as water, alcohol, pyridine and dimethylformamide may join the com- with strong solute-solute interactions [33], an overlapping of the solva-
pound to generate solvates [21] tion spheres when increase the APRA concentration would be
suggested.
3.3. Density studies The variation of V02 with the temperature was adjusted to the follow-
ing equation
Density values for pure DMSO, to be used in the present work, were
experimentally measured and they are collected in Table 2 together V 02 ¼ a þ bT ð4Þ

Fig. 1. TG curve for APRA.


M. Maldonado et al. / Journal of Molecular Liquids 231 (2017) 142–148 145

Table 2
Experimental density (ρ) and viscosity (η) values obtained for pure DMSO in the temperature range (293.15 to 313.15) K and their comparison with the literature values.

T/K ρ 10−3/(kg.m−3) η/(mPa.s)

Experimental Literature Experimental Literature

293.15 1.100580 1.100865 [22] 2.200 2.184[38]


1.100730 [23] 2.213[39]
1.100530 [24] 2.202[40]
298.15 1.095555 1.09537 [25] 1.977 1.99 [41]
1.09537 [26] 1.975 [42]
1.09574 [27] 1.975[43]
303.15 1.090534 1.090812 [22] 1.791 1.79 [41]
1.09074 [23] 1.788 [43]
1.09050 [24] 1.810 [38]
308.15 1.085517 1.08573 [28] 1.631 1.65 [41]
1.0852 [29] 1.630 [43]
1.08607 [30] 1.605 [44]
313.15 1.080499 1.080770 [22] 1.498 1.516 [38]
1.08075 [23] 1.513 [45]
1.08046 [24] 1.4484 [46]

where T is the temperature (in Kelvin) and a and b are empirical constants, decrease and consequently the volume of the solute increases. This as-
which were determined, by a least squares method, as equal to a = sertion is supported by the values obtained for V0inter. In fact, as the tem-
(817.3. ± 1.5) (cm3 mol−1) and b = (0.130 ± 0.005) (cm3·mol−1·K−1). perature increases the values of V0inter also increase, indicating stronger
Eq. (4) can be differentiated with respect to the temperature to obtain the solute-solvent interactions at low temperatures.
partial molar expansibility at infinite dilution, E0φ =b= (0.130 ±0.005)
(cm3 mol-1 K-1). This property is positive under the experimental conditions 3.4. Viscosity studies
used in this work and does not show dependence on the temperature.
The partial molar volume of the solute, V02, can be expressed as the The viscosity values of pure DMSO used in this work were experi-
sum of four major contributions [34]: mentally obtained by us. They are shown in Table 2 as well as the avail-
able ones from the literature. It can be observed that there is a good
V 02 ¼ V 0int þ V 0T þ V 0I þ κ T0 RT ð5Þ agreement among them at all the temperatures studied.
Experimental absolute and relative viscosities, η and ηr, respectively,
where V0int is the intrinsic volume of the solute; V0T is related to the pack- as well as their standard deviations [31] are also given in Table 3 for
ing effects; V0I is the contribution from the solute-solvent interactions APRA in DMSO solutions at the temperatures studied.
and κT0 is the isothermal compressibility of the solvent. The last addend The dependence of ηr with the molal concentration was analyzed
term on the right side of the equation accounts the volume effect related using the Tsangaris-Martin equation [47–49]:
to the kinetic contribution to the pressure of the solute molecule, due to
translational degrees of freedom [34]; this contribution is small and
η
usually can be neglected. The intrinsic volume of the solute is under- ηr ¼ ¼ 1 þ Bm þ Dm2 ð7Þ
stood as the volume actually occupied by the solute molecules which η0
is impenetrable for the solvents molecules and it can be assumed to
be identical to the van der Waals volume, V0W [32], which only depends
where the viscosity B-coefficient is related to solvation (solute-solvent
on the solute. On the contrary, V0T and V0I depend on the solute-solvent
interactions), structure, size and shape or hydrodynamic effects of the
interactions and they can be considered as an interaction volume
solute in the solution and D is an empirical constant related with contri-
V0inter = V0T + V0I .
butions due to both the higher terms of the hydrodynamic effects and
Therefore the Eq. (5) reduces to:
the change with concentration of the solute-solute interactions [47–
49]. D coefficient is important at high solute concentration
V 02 ¼ V 0int þ V 0inter ð6Þ
(N0.5 mol·dm−3) [50], so that in this study (m b 0.023 mol kg−1) it
was disregarded and Eq. (7) reduced to:
On the other hand, since the APRA was synthetized under homoge-
neous conditions, it would be expected that only the crown conformer
was obtained [35]. In fact, as it was previously indicated, only the crown η
ηr ¼ ¼ 1 þ Bm ð8Þ
isomer was experimentally observed. Additionally, due to the functional η0
groups introduced in the coupling reaction (carboxyphenylazo) are very
bulky, the mobility of the resorcin[4]arene APRA is expected to be re-
duced. The last assertion was confirmed by 1H-NMR in DMSO-d6. As a The dependence of ηr on m at the temperatures studied for APRA is
consequence, the intrinsic volume for APRA, V0int, was estimated, by shown in Fig. 3.
using the Winmostar software [36], after an optimization of the The values of the B-coefficient and their uncertainties are given in
geometry of the ions with MOPAC 2012, by means of a semi-empirical Table 5 and its dependence on the temperature, shown in Fig. 4.
PM3 method [37], as equal to 642.6 cm3·mol−1. The interactions vol- As it can be observed, the viscosity increases with the solute concen-
ume, V0inter, was then calculated from eq. 6. In Table 4 the values of tration and decreases with the increasing of the temperature, being
these V0inter at the studied temperatures are summarized. B N 0 at all studied temperatures (Table 5). It is accepted that positive
The positive value for the partial molar expansibility at infinite B values indicate the presence of strong solute-solvent interactions
dilution, E0φ, can be understood in terms of the thermal motion increase [51] (structure-making capacity [52]). Therefore, dipole-dipole interac-
with the temperature. DMSO is a solvent loosely structured having tions taking place between the carboxyl, the azo and the hydroxyl
compressing weak forces which are less effective at high temperatures. groups of APRA with the DMSO molecules could be the responsible of
As a result, when the temperature increases the solvent structure would the positive B values found.
146 M. Maldonado et al. / Journal of Molecular Liquids 231 (2017) 142–148

Table 3
Density (ρ), apparent molal volume (Vφ), viscosity (η), and relative viscosity (ηr) values
for APRA in DMSO at (293.15, 298.15, 303.15, 308.15 and 313.15) K.

m/(mol kg−1) ρ 10−3/(kg·m−3) Vφ/(cm3 ∙mol−1) η/(mPa∙s) ηr

293.15 K
0.00000 1.100580 – 2.200 ± 0.001 1.000
0.00585 1.102542 856.9 ± 0.7 2.314 ± 0.002 1.051
0.00697 1.102908 857.4 ± 0.6 2.345 ± 0.002 1.066
0.00789 1.103212 857.7 ± 0.5 2.369 ± 0.001 1.076
0.00907 1.103595 858.3 ± 0.4 2.391 ± 0.002 1.087
0.01093 1.104198 859.1 ± 0.4 2.425 ± 0.003 1.102
0.01293 1.104840 859.8 ± 0.3 2.471 ± 0.004 1.123
0.01438 1.105303 860.3 ± 0.3 2.520 ± 0.002 1.145
0.01731 1.106230 861.1 ± 0.2 2.556 ± 0.003 1.162
0.01911 1.106797 861.6 ± 0.2 2.633 ± 0.001 1.197
0.02097 1.107385 861.9 ± 0.2 2.678 ± 0.003 1.217
0.02294 1.107999 862.3 ± 0.2 2.724 ± 0.001 1.238

298.15 K
0.00000 1.095555 – 1.977 ± 0.001 1.000 Fig. 2. Apparent molal volume values for solutions of APRA in DMSO at 293.15 (◊), 298.15
0.00585 1.097531 857.5 ± 0.7 2.072 ± 0.003 1.048 (ᴏ), 303.15 (Δ), 308.15 (x) and 313.15 (□) K.
0.00697 1.097899 858.1 ± 0.6 2.100 ± 0.001 1.062
0.00789 1.098204 858.6 ± 0.5 2.124 ± 0.001 1.074
0.00907 1.098589 859.2 ± 0.4 2.143 ± 0.001 1.084
0.01093 1.099195 860.1 ± 0.4 2.167 ± 0.001 1.096
expressions suggested by Feakins et al. [53] after the transition state
0.01293 1.099843 860.7 ± 0.3 2.209 ± 0.001 1.117 theory of Eyring:
0.01438 1.100309 861.2 ± 0.3 2.250 ± 0.001 1.138  
0.01731 1.101242 862.0 ± 0.2 2.284 ± 0.001 1.155 η0 V ∘1
0.01911 1.101817 862.3 ± 0.2 2.346 ± 0.002 1.187 Δμ ∘1# ¼ RT  ln ð9Þ
hNA
0.02097 1.102407 862.7 ± 0.2 2.380 ± 0.003 1.204
0.02294 1.103024 863.1 ± 0.2 2.430 ± 0.004 1.229
RT   
303.15 K
Δμ ∘2# ¼ Δμ ∘1# þ ∘ B− V ∘1 −V ∘2 ð10Þ
V1
0.00000 1.090534 – 1.791 ± 0.000 1.000
0.00585 1.092523 858.3 ± 0.7 1.870 ± 0.002 1.044
0.00697 1.092894 858.9 ± 0.6 1.893 ± 0.004 1.057
where h is the Planck constant; NA is the Avogradro number; η0 and V01
0.00789 1.093201 859.3 ± 0.5 1.907 ± 0.005 1.065 are the viscosity and the molar volume of the solvent, respectively, at
0.00907 1.093590 859.8 ± 0.5 1.931 ± 0.001 1.078 each temperature; V02, the limiting partial molar volume of the solute
0.01093 1.094202 860.6 ± 0.4 1.952 ± 0.001 1.090 and B is the viscosity coefficient. The values of Δμ∘1# and Δμ∘#2 at all tem-
0.01293 1.094851 861.5 ± 0.3 1.991 ± 0.001 1.112
peratures are summarized in Table 5.
0.01438 1.095322 861.9 ± 0.3 2.024 ± 0.002 1.130
0.01731 1.096262 862.7 ± 0.2 2.045 ± 0.001 1.142 As it can be ascertained, Δμ∘# ∘#
2 N Δμ1 what according to Feakins et al.
0.01911 1.096839 863.1 ± 0.2 2.104 ± 0.003 1.175 means that the solute-solvent interactions are strong (as it was already
0.02097 1.097431 863.6 ± 0.2 2.138 ± 0.001 1.194 concluded from the viscosity B coefficient analysis) and the formation of
0.02294 1.098055 863.9 ± 0.2 2.176 ± 0.005 1.215 the activated state is less favorable compared with the ground state
308.15 K [51].
0.00000 1.085517 – 1.631 ± 0.001 1.000 In addition, from the slope of the plot of the Gibbs free energy of acti-
0.00585 1.087520 858.9 ± 0.7 1.698 ± 0.002 1.041 vation per mole of solute against temperature, the entropy of activation
0.00697 1.087893 859.5 ± 0.6 1.724 ± 0.002 1.057
per mole of viscous flow of solution (ΔS∘# 2 ) was obtained according to:
0.00789 1.088202 860.0 ± 0.5 1.737 ± 0.000 1.065
0.00907 1.088593 860.6 ± 0.5 1.756 ± 0.001 1.077 !
0.01093 1.089209 861.3 ± 0.4 1.779 ± 0.006 1.091 Δμ ∘2#
0.01293 1.089865 862.1 ± 0.3 1.812 ± 0.002 1.111 ΔS2°# ¼ −d ð11Þ
0.01438 1.090339 862.5 ± 0.3 1.838 ± 0.000 1.127
dT
0.01731 1.091281 863.6 ± 0.2 1.858 ± 0.000 1.139
0.01911 1.091863 863.9 ± 0.2 1.905 ± 0.001 1.168 and the enthalpy of activation per mole of viscous flow of solution
0.02097 1.092456 864.6 ± 0.2 1.934 ± 0.002 1.186 (ΔH∘#
2 ) was determined by using the equation:
0.02294 1.093085 864.8 ± 0.2 1.967 ± 0.000 1.206

313.15 K ΔH 2°# ¼ Δμ ∘2# þ T  ΔS∘2# ð12Þ


0.00000 1.080499 – 1.498 ± 0.000 1.000
0.00585 1.082515 859.6 ± 0.7 1.555 ± 0.001 1.039
0.00697 1.082891 860.2 ± 0.6 1.579 ± 0.000 1.054
0.00789 1.083202 860.7 ± 0.5 1.597 ± 0.001 1.066 Table 4
0.00907 1.083596 861.2 ± 0.5 1.607 ± 0.000 1.073 Vφ0, Sv, Bv and V0inter values for APRA at (293.15, 298.15, 303.15, 308.15 and 313.15) K.
0.01093 1.084216 862.0 ± 0.4 1.625 ± 0.003 1.085
T/K V0φ (uv)a Sv (us)b Bv(uB)c V0interd
0.01293 1.084876 862.8 ± 0.3 1.651 ± 0.002 1.102
/(cm3 ∙mol−1) /(cm3∙ kg1/2 ∙mol−3/2) /(cm3 kg mol−2) /(cm3 mol−1)
0.01438 1.085354 863.2 ± 0.3 1.680 ± 0.001 1.121
0.01731 1.086302 864.3 ± 0.2 1.695 ± 0.001 1.132 293.15 853.5 ± 0.1 631 ± 13 −10850 ± 405 210.9
0.01911 1.086888 864.7 ± 0.2 1.739 ± 0.001 1.161 298.15 854.3 ± 0.3 635 ± 34 −11039 ± 1072 211.8
0.02097 1.087484 865.3 ± 0.2 1.764 ± 0.000 1.178 303.15 855.0 ± 0.2 628 ± 23 −10461 ± 742 212.4
0.02294 1.088118 865.6 ± 0.2 1.790 ± 0.001 1.195 308.15 855.6 ± 0.3 623 ± 36 −9667 ± 1132 213.1
313.15 856.1 ± 0.1 664 ± 16 −11069 ± 757 213.5
a
V0φ is the apparent molal volume of the solute at infinite dilution.
b
3.5. Thermodynamic activation parameters Sv is the experimental slope (related to the effect of the solute on the solvent
structure).
c
Bv is an empirical parameter.
The Gibbs free energies of activation per mole of viscous flow of both d
V0inter is an interaction volume of the solute (depending on the solute-solvent
the solvent (Δμ∘# ∘#
1 ) and the solute (Δμ2 ) were calculated by using the interactions).
M. Maldonado et al. / Journal of Molecular Liquids 231 (2017) 142–148 147

Fig. 4. Dependence of the viscosity B coefficient on the temperature for solutions of APRA
in DMSO.
Fig. 3. Dependence of the relative viscosity (ηr) on the molal concentration (m) for APRA in
solution at 293.15 (□), 298.15 (◊), 303.15 (Δ), 308.15 (x) and 313.15 (o) K.
between the hydroxyl groups of the resorcinol rings, aza groups and
carboxylic groups. This result was confirmed by the values calculated
The values found for this thermodynamics parameters at 298.15 K for the Gibbs free energy of activation per mol of viscous flow of both
were ΔS∘#2 = 2246 J mol
−1
and ΔH∘#2 =1054 kJ mol
−1
. The positive en- the solvent and the solute.
tropy value, which suggests a transition state less organized than the
ground state, together with the unfavorable enthalpy value (ΔH∘# 2 N 0)
point to the process is enthalpy driven. References
Besides that, the relative contributions to Δμ∘#
2 of both the enthalpy [1] M.S. Deshmukh, N. Sekar, A combined experimental and TD-DFT investigation of
(ζH) and the entropy (ζTS) were calculated using the following equa- three disperse azo dyes having the nitroterephthalate skeleton, Dyes Pigments
103 (2014) 25–33.
tions:
[2] O.A. Adegoke, T.E. Adesuji, O.E. Thomas, Novel colorimetric sensors for cyanide
  based on azo-hydrazone tautomeric skeletons, Spectrochim. Acta, Part A 128
 ∘#  (2014) 147–152.
ΔH2  [3] Y. Egawa, R. Miki, T. Seki, Colorimetric sugar sensing using boronic acid-substituted
ζH ¼     ð13Þ
 ∘#   ∘#  azobenzenes, Materials 7 (2014) 1201–1220.
ΔH 2  þ T ΔS2  [4] J. Isaad, A. Perwuelz, New color chemosensor for cyanide based on water soluble azo
dyes, Tetrahedron Lett. 51 (2010) 5810–5814.
  [5] M.M. Ferris, M.A. Leonard, Examination of metallochromic indicators and water-sol-
 
T ΔS∘2#  uble reagents for metals by planar electrophoresis, Analyst 111 (1986) 351–354.
ζ TS ¼     ð14Þ [6] M. Han, X. Zhang, X. Zhang, C. Liao, B. Zhu, Q. Li, Azo-coupled zinc phthalocyanines:
 ∘#   ∘#  towards broad absortion and application in dye-sensitized solar cells, Polyhedron 85
ΔH2  þ T ΔS2 
(2015) 864–873.
[7] E. Ortyl, J. Jaworowska, P.T. Seifi, R. Barille, S. Kucharski, Photochromic polymer and
hybrid materials containing azo methylsoxasole dye, Soft Matter 9 (2011) 335–346.
The values obtained for these ζH and ζTS were 0.61 and 0.39, [8] L. Abramsson-Zetterberg, N. Ilbäck, The synthetic food colouring agent Allura Red
respectively. AC (E129) is not genotoxic in a flow cytometry-based micronucleus assay in vivo,
Food Chem. Toxicol. 59 (2013) 86–89.
[9] L.T. May, S.J. Briddon, S.J. Hill, Antagonist selective modulation of adenosine A1 and
4. Conclusions A3 receptor pharmacology by the food dye brilliant black BN: evidence for allosteric
interactions, Mol. Pharmacol. 77 (2010) 678–686.
[10] W. Iwanek, A. Wzorek, Introduction to the chirality of resorcinarenes, Mini-Rev. Org.
Resorcinarenes PRA and APRA were synthesized in good yields. The
Chem. 6 (2009) 398–411.
new synthesized Azaresorcinarene (APRA) was characterized by mass [11] J. Han, C.-G. Yan, Synthesis, crystal structure and configuration of resorcinarene am-
spectrometry, IR, 1H and 13C NMR studies. Also, absorption spectra ides, J. Incl. Phenom. Macrocycl. Chem. 61 (2008) 119–126.
were recorded. Such characterization concluded that only the crown [12] J. Sun, L.-L. Zhang, Y. Yao, C.-G. Yan, Synthesis, crystal structures and complexing
properties of tetramethoxyresorcinarene functionalized tetraacylhydrazones, J.
isomer of APRA was obtained. Moreover, a TG analysis was performed Incl. Phenom. Macrocycl. Chem. 79 (2014) 485–494.
from which no formation of solvates of APRA in this solvent (DMSO) [13] V.K. Jain, P.H. Kanaiya, N. Bhojak, Synthesis, spectral characterization of azo dyes de-
was concluded. Using experimental density and viscosity data, values rived from calix[4]resorcinarenes and their application in dyeing of fibers, Fiber.
Polym. 9 (2008) 720–726.
for both the apparent partial molal volume and the apparent molal ex- [14] S. Elcin, M.M. Ilhan, H. Deligoz, Synthesis and spectral characterization of azo dyes
pansibility at infinite dilution, as well as the viscosity B-coefficient, derived from calix[4]arene and their application in dyeing of fibers, J. Incl. Phenom.
were determined. The positive values found for B suggest the presence Macrocycl. Chem. 77 (2013) 259–267.
[15] A.G.S. Hoegberg, Two stereoisomeric macrocyclic resorcinol-acetaldehyde conden-
of strong APRA-DMSO interactions, probably of dipole-dipole type, sation products, J. Organomet. Chem. 45 (1980) 4498–4500.
[16] J. Kestin, M. Sokolov, W.A. Wakeham, Viscosity of the liquid in the range water −8
to 150 °C, J. Phys. Chem. Ref. Data 7 (3) (1978) 941–948.
Table 5
[17] A. Wzorek, J. Mattay, W. Iwanek, Synthesis and structural investigation of the
Values of the viscosity B coefficient and the activation Gibbs energies of the solvent,Δμ∘1#,
cyclochiral boron resorcinarenes obtained from L-amino acids and boronic acid, Tet-
and of the solute, Δμ∘2#, at (293.15, 298.15, 303.15, 308.15 and 313.15) K for APRA in
rahedron Asymmetry 23 (2012) 271–277.
DMSO. [18] M. Urbaniak, J. Mattay, W. Iwanek, Synthesis of resorcinarene derivatives via the cat-
alyzed mannich reaction, part 3: glycol derivatives of resorcinarene, Synth.
T/K B/(kg∙mol−1) Δμ∘1#/(kJ·mol−1) Δμ∘#
2 /(kJ·mol
−1
)
Commun. 41 (2011) 670–676.
293.15 10.51 ± 0.01 14.55 402.22 [19] A. Galán, E.C. Escudero-Adán, A. Frontera, P. Ballester, Synthesis, structure, and bind-
298.15 9.87 ± 0.03 14.54 384.83 ing properties of lipophilic cavitands base on calix[4]pyrrole-resorcinarene hybrid
303.15 9.28 ± 0.04 14.55 368.56 scaffold, J. Organomet. Chem. 79 (2012) 5545–5557.
308.15 9.00 ± 0.02 14.56 362.66 [20] L.M. Tundstad, J.A. Tucker, E. Dalcanale, J. Weiser, J.A. Bryant, J.C. Sherman, R.C.
Helgeson, C.B. Knobler, D.J. Cram, Host-guest complexation. 48. Octol building
313.15 8.73 ± 0.02 14.59 357.17
blocks for cavitands and carcerands, J. Organomet. Chem. 54 (1989) 1305–1312.
148 M. Maldonado et al. / Journal of Molecular Liquids 231 (2017) 142–148

[21] O. Pietraszkiewicz, E. Utzig, W. Zielenkiewics, M. Pietraszkiewicz, Thermochemical [40] A. Nissema, L. Koskenniska, Thermodynamic properties of the binary system di-
investigation of some selected solvates of calix[4]resorcinarene, J. Therm. Anal. 54 methyl sulphoxide-p-Xileno, Suomen Kemistilehti B 45 (1972) 203–205.
(1998) 249–255. [41] V. Govinda, P. Attri, P. Venkatesu, V. Venkateswarlu, Temperature effect on the mo-
[22] X. Wang, F. Yang, Y. Gao, Z. Liu, Volumetric properties of binary mixtures of lecular interactions between two ammonium ionic liquids and dimethylsulfoxide, J.
dimethylsulfoxide with amines from (293.15 to 363.15) K, J. Chem. Thermodyn. Mol. Liq. 164 (2011) 218–225.
57 (2013) 145–151. [42] L. Börnstein, Group IV physical chemistry. Pure organometallic and
[23] O. Ciocirlan, O. Lulian, Vapor pressure, density, viscosity and refractive index of di- organononmetallic liquids, binary liquid mixtures, viscosity of pure organic liquids
methyl sulfoxide + 1,4-dimethylbenzene system, J. Serb. Chem. Soc. 73 (2008) and binary liquid mixtures, in: Ch. Wohlfarth, B. Wohlfarth, M.D. Lechner (Eds.),
73–85. Mixtures of Organic Compounds. Part 2, Springer-Verlag, Berlin Heidenberg 2001,
[24] N.G. Tsierkesos, A.E. Kelarakis, M.M. Palaiologou, Densities, viscosities, refractive in- p. 692.
dices, and surface tension of dimethyl sulfoxide + butyl acetate mixtures at (293.15, [43] M.I. Aralaguppi, T.M. Aminabhavi, S.B. Harogoppad, R.H. Balundgi, Thermodynamic
303.15 and 313.15) K, J. Chem. Eng. Data 45 (2000) 395–398. interactions in binary mixtures of dimethyl sulfoxide with benzene, toluene, 1,3-
[25] V. Govinda, P.M. Reddy, P. Attri, P. Venkatesu, V. Venkateswarlu, Influence of anion dimethylbenzene, 1,3,5-trimethylbenzene and methoxybenzene from 298.15 to
on thermophysical properties of ionic liquids with polar solvent, J. Chem. 308.15 K, J. Chem. Eng. Data 37 (1992) 298–303.
Thermodyn. 58 (2013) 269–278. [44] G.S. Gokavi, J.R. Raju, T.M. Aminabhavi, R.H. Balundgi, M.V. Muddapur, Viscosities
[26] R. Radhamma, P. Venkatesu, T. Hoffman, Vapor-liquid equilibrium for the binary and densities of binary liquid mistures of dimethyl sulfoxide with chlorobenzene,
mixtures of dimethyl sulfoxide with substituted benzenes, Fluid Phase Equilib. pyridine, and methyl ethyl ketone at 25, 35, 45, and 55. degree. C, J. Chem. Eng.
262 (2007) 32–36. Data 31 (1986) 15–18.
[27] R. Ghosh, S. Banerjee, S. Chakrabarty, B. Bagchi, Anomalous behavior of linear hydro- [45] P.G. Sears, T.M. Stoeckinger, L.R. Dawson, Dielectric constants, viscosities, fusion
carbon chains in water-DMSO binary mixture at low DMSO concentration, J. Phys. point curves, and other properties of three nonaqueous binary systems, J. Chem.
Chem. B 115 (2011) 7612–7620. Eng. Data 16 (1971) 220–222.
[28] A. Ali, S. Ansari, A.K. Nain, Densities, refractive indices and excess properties of bina- [46] P. Umadevi, K. Rambabu, M.N. Rao, K.S. Rao, C. Rambabu, Densities, adiabatic com-
ry mixtures of dimethylsulphoxide with some poly(ethylene glycol)s at different pressibility, free-length, viscosities and excess volumes of P-cresol(L) + dimethyl
temperatures, J. Mol. Liq. 178 (2013) 178–184. sulfoxide (2), + dimethyl formamide (2), and 1,4-dioxane at 303.15–318.15 K,
[29] P.K. Thakur, S. Patre, R. Pande, Thermophysical and excess properties of hydroxamic Phys. Chem. Liq. 30 (1995) 29–46.
acids in DMSO, J. Chem. Thermodyn. 58 (2013) 226–236. [47] J.M. Tsangaris, R.B. Martin, Viscosities of aqueous solutions of dipolar ions, Biochem.
[30] V. Govinda, P. Attri, P. Venkatesu, V. Venkateswarlu, Thermophysical properties of Biophys. 112 (1965) 267–272.
dimethylsulfoxide with ionic liquids at various temperatures, Fluid Phase Equilib. [48] C.I.A.V. Santos, C. Teijeiro, A.C.F. Ribeiro, D.F.S.L. Rodrigues, C.M. Romero, M.A. Esteso,
304 (2011) 35–43. Drug delivery systems: Study of inclusion complex formation for ternary caffeine-β-
[31] NIST, Guidelines for evaluating and expressing the uncertainty of NIST measure- cyclodextrin-water mixtures from apparent molar volume values at 298.15 K and
ment results, adapted from NIST technical note 1297, 1994 (edn). 310.15 K, J. Mol. Liq. 223 (2016) 209–216.
[32] L. Bernazani, V. Mollica, M.R. Tiné, Partial molar volumes of organic compounds in [49] C.M. Romero, D.M. Rodriguez, A.C.F. Ribeiro, M.A. Esteso, Effect of temperature on
C8 solvents at 298.15 K, Fluid Phase Equilib. 203 (2002) 15–29. the partial molar volume, isentropic compressibiblity and viscosity of DL-2-
[33] D. Warminska, J. Wawer, W. Grzybkowski, Apparent molar volumes and compress- aminobutyric acid in water and in aqueous sodium chloride solutions, J. Chem.
ibilities of alkaline earth metal ions in methanol and dimethylsulfoxide, J. Chem. Thermodyn. 104 (2017) 274–280.
Thermodyn. 42 (2010) 1116–1125. [50] H. Donald, B. Jenkins, Y. Marcus, Viscosity B-coefficients of ions in solution, Chem.
[34] O. Likhodi, T.V. Chalikian, Partial molar volumes and adiabatic compressibilities of a Rev. 95 (1995) 2695–2724.
series of aliphatic amino acids and oligoglycines in D2O, J. Am. Chem. Soc. 121 [51] S.S. Dhondge, P.N. Dahasahasra, L.J. Paliwal, D.W. Deshmukh, Density and viscosity
(1999) 1156–1163. study of nicotinic acid and nicotinamide in dilute aqueous solutions and around
[35] P. Timmerman, W. Verboom, D.N. Reinhoudt, Resorcinarenes, Tetrahedron 52 the temperature of the maximum density of water, J. Chem. Thermodyn. 76
(1996) 2663–2704. (2014) 16–23.
[36] N. Senda, Idemitsu Technical Report, 49, 2006 106–111. [52] Y. Marcus, Effect of ions on the structure of water: structure making and breaking,
[37] J.P. Stewart, MOPAC2012, Stewart Computational Chemistry, Colorado Springs, CO, Chem. Rev. 109 (2009) 1346–1370.
USA(HTTP://OpenMOPAC.net) 2012. [53] D. Feakins, D. Freemantle, K. Lawrence, Transition state treatment of the relative vis-
[38] A. Nissema, A. Kuvaja, Thermodynamic properties of the binary system dimethyl cosity of electrolytic solutions. Applications to aqueous, non-aqueous and
sulphoxide-hexafluorobezene, Suomen Kemistilehti, B 45 (1972) 206–273. methanol + water systems, J. Chem. Soc. Faraday Trans. I 70 (1974) 795–806.
[39] A. Nissema, H. Saeynaejaekangas, Thermodynamic properties of binary and ternary
systems. II. Excess molar volumes and excess viscosities of toluene + dimethyl
sulphoxide mixtures, Finn. Chem. Lett. 129 (1976) 40–44.

You might also like