Apparent molal volume and viscosity values for a new synthesized diazoted resorcin[4]arene in DMSO at several temperatures

Apparent molal volume and viscosity values for a new synthesized diazoted resorcin[4]arene in DMSO at several temperatures

Journal of Molecular Liquids 231 (2017) 142–148 Contents lists available at ScienceDirect Journal of Molecular Liquids journal homepage: www.elsevie...

720KB Sizes 0 Downloads 29 Views

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 Departamento de Química, Universidad de los Andes, Cr. 1 No. 18A 10, Bogotá 111711, Colombia 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 b c

a r t i c l e

i n f o

Article history: Received 30 November 2016 Received in revised form 25 January 2017 Accepted 27 January 2017 Available online 31 January 2017 Keywords: Diazoted resorcin[4]arenes p-(3-carboxyphenylazo)propylresorcin[4]arene (APRA) DMSO Apparent molal volumes Viscosities

a b s t r a c t A new diazoted resorcin[4]arene was synthesized and characterized by spectral techniques. The densities and viscosities of their solutions in dimethylsulfoxide (DMSO) were measured at temperatures between 293.15 and 313.15 K, over a range of concentrations from (0.0058 to 0.023) (mol kg−1) at atmospheric pressure. By using these experimental data, some thermodynamic functions were derived to contribute to the understanding of the intermolecular interactions taking place in solution. The results obtained from the volumes study indicate the presence of strong solute-solute interactions. The study of viscosities suggests the existence of strong solutesolvent interactions. The above result was confirmed by calculating the Gibbs free energies of activation, per mole, of viscous flow of both the solvent and the solute. The flow parameter values found indicate an activated state less organized compared with the ground state and a process driven by the enthalpy. © 2017 Published by Elsevier B.V.

1. Introduction Azo dyes have an important role in the chemical industry and are of interest for different fields of study, such as organic chemistry, analytical chemistry and physical chemistry. Their applications included dyeing of fibers [1], sensors for molecules and ions [2–4], metallochromic indicators [5], solar cells [6], photochromic materials [7], food additives [8], therapeutics agents [9], among other applications. The azo dyes derived from resorcin[4]arenes constitute an interesting type of compounds since they have several conformations in solution and solid state [10– 12]. In the literature, few studies are found about the synthesis of azoresorcin[4]arene [13,14]. The results of these studies indicate that the best synthetic route is the diazotization in the free position of the annular ring between hydroxyl groups. The synthesis and characterization of new azo dyes provides guidance for further studies. In this work, the synthesis of a new azoresorcin[4]arene was carried out by diazotization of 3-aminobenzoic acid and subsequent coupling with tetrapropylresorcin[4]arene. The compound obtained was purified and characterized and its solution properties were studied in dimethylsulfoxide (DMSO) at several temperatures in the range (293.15 to 313.15) K in order to obtain information about these macrocyclic systems covering both the standard and the physiological ⁎ Corresponding authors. E-mail addresses: [email protected] (M. Maldonado), [email protected] (M.Á. Esteso).

http://dx.doi.org/10.1016/j.molliq.2017.01.093 0167-7322/© 2017 Published by Elsevier B.V.

temperatures. The results are interpreted in terms of solute-solute, solvent-solvent and solute-solvent interactions. 2. Experimental 2.1. Materials In Table 1, source, purity, quantification method and CAS number of the chemicals used are shown. They were used as supplied, without further purification. Tetrapropylresorcin[4]arene (PRA) and p-(3carboxyphenylazo)propylresorcin[4]arene (APRA) were prepared according to procedures reported in the literature [15]; their purity was determined by HPLC as better than 98%. The resorcin[4]arenes were kept in dark bottles, to protect them from sunlight, over freshly activated silica gel. Their yields are referred to isolated product after purification. DMSO was stored over 3 Å molecular sieve and was used without further purification. 2.2. Analytical characterization studies The characterization of the synthesized compounds was done by FTIR, 1H‐NMR, 13C‐NMR, thermogravimetric analysis and mass spectroscopy. FT-IR spectra were recorded on KBr disc, using a Thermo Nicolet IS10 spectrophotometer. The 1H‐NMR and 13C‐NMR spectra were recorded on a Varian spectrometer Unity 300 apparatus. The chemical

M. Maldonado et al. / Journal of Molecular Liquids 231 (2017) 142–148 Table 1 Source, purity and CAS number of the chemicals used. Chemical name

CAS number

DMSO

67-68-5

Source

Fluka Purum Resorcinol 108-46-3 Merck Butanal 123-72-8 Merck Hydrochloric acid 7647-01-0 Merck Sodium nitrite 7632-00-0 Merck 3-Aminobenzoic acid 99-05-8 Acros Sodium hydroxide 1310-72-2 Merck Tetrapropylcalix[4]resorcinarene – Synthesized Azacalix[4]resorcinarene – Synthesized

Mass fraction purity

Method

≥0.99 0.99 0.99 0.37 0.99 0.99 0.99 ≥0.99 0.98

143

400 MHz, DMSO-d6): 0.91 (t, J = 8.0 Hz, 12H), 1.20 (m, 8H), 1.85 (m, 8H), 4.45 (t, J = 7.6 Hz, 4H), 7.35 (s, 4H), 7.56–8.33 (m, 16H), 9.01 (s, 8H), 13.2 (s, 4H); 13C-NMR (δ, 100 MHz, DMSO-d6): 13.7, 20.5, 29.3, 55.1, 115.6, 116.0, 117.5, 119.6, 122.3, 125.1, 127.2, 129.6, 132.0, 141.7, 166.3; MALDI-TOF-MS (4-nitroaniline): calcd. for C68H64N8O16: m/ z = 1249.280 [M]+; found: m/z = 1272.564 [M + Na]+. The values measured for 1H-NMR in DMSO show that only one conformational isomer was obtained, the crown isomer. 2.5. Determination of density

HPLC HPLC

shifts (δ), in ppm, were obtained in DMSO-d6 using TMS as an internal standard. The thermogravimetric analysis (TG) was performed by using a Netzsch STA 409 thermobalance with a sample weight of 28 mg, over a temperature range of 20–500 °C and a heating rate of 10 °C/min; the measurements were carried out in a nitrogen atmosphere (flow rate: 16.66 mL/min) by using an alumina crucible. Molar mass was determined on a MALDI-TOF spectrometer (Bruker Daltonics) using 4-nitroaniline as matrix for the desorption/ionization process.

Solutions were prepared by direct weighting of both the solute and 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 concentration range studied was (0.0058 to 0.023) (mol kg−1). Densities were measured with an Anton Paar DMA 5000 densimeter, which has a sensitivity of 1 ∙ 10−6 (g cm−3) and accuracy of 5 ∙ 10−6 (g cm−3) in the ranges of (0–90) °C of temperature and (0–1.0) MPa of pressure. This instrument uses the vibrating U-tube measuring principle and is equipped with a Peltier type thermostating unit, that ensures a temperature control of ±0.005 K. The densimeter was calibrated with dry air and purified water (milli-Q® quality) at each studied temperatures, according to the recommendation given by the manufacturer.

2.3. Synthesis and characterization of PRA PRA was synthesized by condensation in acid medium of resorcinol and butanaldehyde in a 1:1 water-ethanol mixture, following the literature described method [15], with a yield of 86%. The values found for its characterization were: IR (KBr, υ cm−1): 3314 (O\\H), 2957 (C\\H), 1616 (C=C), 1194 (C\\O); 1H‐NMR (δ, 400 MHz, DMSO-d6): 0.90 (t, J = 8.0 Hz, 12H), 1.21 (m, 8H), 2.09 (q, J = 8.0 Hz, 8H), 4.23 (t, J = 8.0 Hz, 8H), 6.15 (s, 4H), 7.24 (s, 4H), 8.94 (s, 8H); 13C-NMR (δ, 100 MHz, DMSO-d6): 16.1, 25.3, 32.7, 49.2, 112.1, 127.2, 129.0, 151.1

2.6. Determination of viscosity Viscosities were measured by using an Ostwal type viscometer, model Cannon-Fenske, calibrated with purified water (milli-Q® quality) at each temperature studied, with uncertainty of ±0.25% and efflux times in the range (374–670) s in order to make irrelevant the kinetic energy correction (Hagenbach correction). The temperature control was ±0.02 K. The viscometer constants were calculated with the help of the Poiseuille equation:

2.4. Synthesis and characterization of APRA Diazotation and coupling reaction were done by following the literature 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 (1 mL at 37%). The resulting mixture was slowly added over a solution of PRA (0.25 mmol) in 10 mL of NaOH 1 M solution. The reaction mixture was stirred for 2 h at a temperature below 5 °C. The red solid obtained was filtered, washed with water and then dried at 50 °C for 24 h. The yield was of 91%. The characterization values obtained were: IR (KBr, υ cm−1): 3314 (O\\H), 2957 (C\\H), 1616 (C=C), 1194 (C\\O); 1H-NMR (δ,

η β ¼ α− 2 ρt t

ð1Þ

where η is the viscosity, α and β are the viscometer constants, and ρ and t are the density and the flow time, respectively, of a liquid of known viscosity (in this case water). These constants were obtained from the plot of η/ρt against 1/t2 concerning literature data of pure water [16]. The viscometer constants were used to determine the absolute viscosity values of solutions of APRA in DMSO. The viscosity values reported in this work are the mean ones of at least three sets of measurements (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 3.1. Spectral data Synthesis of PRA has been reported in the literature, while similar information about APRA has not been previously reported. Consequently, in the case of PRA the results obtained from the spectral studies could be compared with values reported in the literature for this compound, whereas in the case of APRA such comparison had to be done by recurring to values for similar compounds [17–20]. In this way, the results for 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 molar mass, determined by high resolution MALDI-TOF-MS using 4nitroaniline as matrix, revealed the expected [M + Na]. The FT-IR spectrum showed absorptions from azo group (1044 cm−1 and 773 cm−1), aromatic ring (1609 cm−1), alkyl chains (2929 cm−1) and hydroxyl groups (3421 cm−1). The 1H‐NMR was recorded in DMSO-d6 and displayed characteristic signals of propyl chains (1.09, 1.60, and 2.35 ppm), a methylene bridge fragment between the aromatic rings (4.60 ppm), and the aromatic hydrogen of penta substituted resorcinol units (7.39 ppm). 3.2. Thermal properties To examine the thermal stability of APRA and the possible formation of solvates, TG analysis was performed under the conditions given in section 2.2. The TG curve related to the thermal decomposition of APRA is shown in Fig. 1. This indicates that the decomposition occurs near 250 °C; that is, this compound is stable below 250 °C. On the other hand, in the solvents used for the crystallization processes, formation of solvates is not observed. This result contrast with previous studies for other resorcin[4]arene systems, which show that molecules, such as water, alcohol, pyridine and dimethylformamide may join the compound to generate solvates [21] 3.3. Density studies Density values for pure DMSO, to be used in the present work, were experimentally measured and they are collected in Table 2 together

with those available from the literature. A satisfactory agreement between them is observed at each temperature. Experimental densities and apparent molal volumes together with their respective standard deviations of the measurements are given in Table 3 for the solutions of APRA in DMSO at all temperatures studied. In all cases, the measurement uncertainty value (better than 0.7 cm3 mol−1) was calculated according to the law of propagating the variances [31]. Apparent molal volumes,Vφ, were calculated from density values by using the equation Vφ ¼

M 2 1000 ðρ−ρ0 Þ − ρ mρρ0

ð2Þ

where M2 is the molar mass of APRA, ρ is the density of its solution in DMSO, ρo is the density of the pure solvent and m denotes the molality of the solution. These apparent molal volume values were fitted against the molal concentration, m, by means of a weighted linear regression to the second order polynomial equation [32]: V φ ¼ V 0φ þ Sv m þ Bv m2

ð3Þ

where V0φ is the apparent molal volume of the solute at infinite dilution (equal to its partial molar volume at infinite dilution,V02), Sv is the experimental slope (which has been related to the effect of the solute on the solvent structure [33]) and Bv is an empirical parameter. The fittings for the different temperatures studied are shown in Fig. 2, and the Vφ∘, Sv, Bv values are listed in Table 4 at the interest temperatures. As it can be seen, Bv b 0 and Sv N 0, independently of the temperature. By taking into account that a positive value of SV is associated with strong solute-solute interactions [33], an overlapping of the solvation spheres when increase the APRA concentration would be suggested. The variation of V02 with the temperature was adjusted to the following equation V 02 ¼ a þ bT

Fig. 1. TG curve for APRA.

ð4Þ

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

2.200

298.15

1.095555

303.15

1.090534

308.15

1.085517

313.15

1.080499

1.100865 [22] 1.100730 [23] 1.100530 [24] 1.09537 [25] 1.09537 [26] 1.09574 [27] 1.090812 [22] 1.09074 [23] 1.09050 [24] 1.08573 [28] 1.0852 [29] 1.08607 [30] 1.080770 [22] 1.08075 [23] 1.08046 [24]

2.184[38] 2.213[39] 2.202[40] 1.99 [41] 1.975 [42] 1.975[43] 1.79 [41] 1.788 [43] 1.810 [38] 1.65 [41] 1.630 [43] 1.605 [44] 1.516 [38] 1.513 [45] 1.4484 [46]

where T is the temperature (in Kelvin) and a and b are empirical constants, which were determined, by a least squares method, as equal to a = (817.3. ± 1.5) (cm3 mol−1) and b = (0.130 ± 0.005) (cm3·mol−1·K−1). Eq. (4) can be differentiated with respect to the temperature to obtain the 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 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 sum of four major contributions [34]: V 02 ¼ V 0int þ V 0T þ V 0I þ κ T0 RT

ð5Þ

where V0int is the intrinsic volume of the solute; V0T is related to the packing effects; V0I is the contribution from the solute-solvent interactions and κT0 is the isothermal compressibility of the solvent. The last addend term on the right side of the equation accounts the volume effect related to the kinetic contribution to the pressure of the solute molecule, due to translational degrees of freedom [34]; this contribution is small and usually can be neglected. The intrinsic volume of the solute is understood as the volume actually occupied by the solute molecules which 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 on the solute. On the contrary, V0T and V0I depend on the solute-solvent interactions and they can be considered as an interaction volume V0inter = V0T + V0I . Therefore the Eq. (5) reduces to: V 02 ¼ V 0int þ V 0inter

ð6Þ

On the other hand, since the APRA was synthetized under homogeneous conditions, it would be expected that only the crown conformer was obtained [35]. In fact, as it was previously indicated, only the crown isomer was experimentally observed. Additionally, due to the functional groups introduced in the coupling reaction (carboxyphenylazo) are very bulky, the mobility of the resorcin[4]arene APRA is expected to be reduced. The last assertion was confirmed by 1H-NMR in DMSO-d6. As a consequence, the intrinsic volume for APRA, V0int, was estimated, by using the Winmostar software [36], after an optimization of the geometry of the ions with MOPAC 2012, by means of a semi-empirical PM3 method [37], as equal to 642.6 cm3·mol−1. The interactions volume, V0inter, was then calculated from eq. 6. In Table 4 the values of these V0inter at the studied temperatures are summarized. The positive value for the partial molar expansibility at infinite dilution, E0φ, can be understood in terms of the thermal motion increase with the temperature. DMSO is a solvent loosely structured having compressing weak forces which are less effective at high temperatures. As a result, when the temperature increases the solvent structure would

1.977

1.791

1.631

1.498

decrease and consequently the volume of the solute increases. This assertion is supported by the values obtained for V0inter. In fact, as the temperature increases the values of V0inter also increase, indicating stronger solute-solvent interactions at low temperatures. 3.4. Viscosity studies The viscosity values of pure DMSO used in this work were experimentally obtained by us. They are shown in Table 2 as well as the available ones from the literature. It can be observed that there is a good agreement among them at all the temperatures studied. Experimental absolute and relative viscosities, η and ηr, respectively, as well as their standard deviations [31] are also given in Table 3 for APRA in DMSO solutions at the temperatures studied. The dependence of ηr with the molal concentration was analyzed using the Tsangaris-Martin equation [47–49]:

ηr ¼

η ¼ 1 þ Bm þ Dm2 η0

ð7Þ

where the viscosity B-coefficient is related to solvation (solute-solvent interactions), structure, size and shape or hydrodynamic effects of the solute in the solution and D is an empirical constant related with contributions due to both the higher terms of the hydrodynamic effects and the change with concentration of the solute-solute interactions [47– 49]. D coefficient is important at high solute concentration (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:

ηr ¼

η ¼ 1 þ Bm η0

ð8Þ

The dependence of ηr on m at the temperatures studied for APRA is shown in Fig. 3. The values of the B-coefficient and their uncertainties are given in Table 5 and its dependence on the temperature, shown in Fig. 4. As it can be observed, the viscosity increases with the solute concentration and decreases with the increasing of the temperature, being B N 0 at all studied temperatures (Table 5). It is accepted that positive B values indicate the presence of strong solute-solvent interactions [51] (structure-making capacity [52]). Therefore, dipole-dipole interactions taking place between the carboxyl, the azo and the hydroxyl groups of APRA with the DMSO molecules could be the responsible of 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 0.00585 0.00697 0.00789 0.00907 0.01093 0.01293 0.01438 0.01731 0.01911 0.02097 0.02294

1.100580 1.102542 1.102908 1.103212 1.103595 1.104198 1.104840 1.105303 1.106230 1.106797 1.107385 1.107999

– 856.9 857.4 857.7 858.3 859.1 859.8 860.3 861.1 861.6 861.9 862.3

± ± ± ± ± ± ± ± ± ± ±

0.7 0.6 0.5 0.4 0.4 0.3 0.3 0.2 0.2 0.2 0.2

2.200 2.314 2.345 2.369 2.391 2.425 2.471 2.520 2.556 2.633 2.678 2.724

± ± ± ± ± ± ± ± ± ± ± ±

0.001 0.002 0.002 0.001 0.002 0.003 0.004 0.002 0.003 0.001 0.003 0.001

1.000 1.051 1.066 1.076 1.087 1.102 1.123 1.145 1.162 1.197 1.217 1.238

298.15 K 0.00000 0.00585 0.00697 0.00789 0.00907 0.01093 0.01293 0.01438 0.01731 0.01911 0.02097 0.02294

1.095555 1.097531 1.097899 1.098204 1.098589 1.099195 1.099843 1.100309 1.101242 1.101817 1.102407 1.103024

– 857.5 858.1 858.6 859.2 860.1 860.7 861.2 862.0 862.3 862.7 863.1

± ± ± ± ± ± ± ± ± ± ±

0.7 0.6 0.5 0.4 0.4 0.3 0.3 0.2 0.2 0.2 0.2

1.977 2.072 2.100 2.124 2.143 2.167 2.209 2.250 2.284 2.346 2.380 2.430

± ± ± ± ± ± ± ± ± ± ± ±

0.001 0.003 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.002 0.003 0.004

1.000 1.048 1.062 1.074 1.084 1.096 1.117 1.138 1.155 1.187 1.204 1.229

303.15 K 0.00000 0.00585 0.00697 0.00789 0.00907 0.01093 0.01293 0.01438 0.01731 0.01911 0.02097 0.02294

1.090534 1.092523 1.092894 1.093201 1.093590 1.094202 1.094851 1.095322 1.096262 1.096839 1.097431 1.098055

– 858.3 858.9 859.3 859.8 860.6 861.5 861.9 862.7 863.1 863.6 863.9

± ± ± ± ± ± ± ± ± ± ±

0.7 0.6 0.5 0.5 0.4 0.3 0.3 0.2 0.2 0.2 0.2

1.791 1.870 1.893 1.907 1.931 1.952 1.991 2.024 2.045 2.104 2.138 2.176

± ± ± ± ± ± ± ± ± ± ± ±

0.000 0.002 0.004 0.005 0.001 0.001 0.001 0.002 0.001 0.003 0.001 0.005

1.000 1.044 1.057 1.065 1.078 1.090 1.112 1.130 1.142 1.175 1.194 1.215

308.15 K 0.00000 0.00585 0.00697 0.00789 0.00907 0.01093 0.01293 0.01438 0.01731 0.01911 0.02097 0.02294

1.085517 1.087520 1.087893 1.088202 1.088593 1.089209 1.089865 1.090339 1.091281 1.091863 1.092456 1.093085

– 858.9 859.5 860.0 860.6 861.3 862.1 862.5 863.6 863.9 864.6 864.8

± ± ± ± ± ± ± ± ± ± ±

0.7 0.6 0.5 0.5 0.4 0.3 0.3 0.2 0.2 0.2 0.2

1.631 1.698 1.724 1.737 1.756 1.779 1.812 1.838 1.858 1.905 1.934 1.967

± ± ± ± ± ± ± ± ± ± ± ±

0.001 0.002 0.002 0.000 0.001 0.006 0.002 0.000 0.000 0.001 0.002 0.000

1.000 1.041 1.057 1.065 1.077 1.091 1.111 1.127 1.139 1.168 1.186 1.206

313.15 K 0.00000 0.00585 0.00697 0.00789 0.00907 0.01093 0.01293 0.01438 0.01731 0.01911 0.02097 0.02294

1.080499 1.082515 1.082891 1.083202 1.083596 1.084216 1.084876 1.085354 1.086302 1.086888 1.087484 1.088118

– 859.6 860.2 860.7 861.2 862.0 862.8 863.2 864.3 864.7 865.3 865.6

Fig. 2. Apparent molal volume values for solutions of APRA in DMSO at 293.15 (◊), 298.15 (ᴏ), 303.15 (Δ), 308.15 (x) and 313.15 (□) K.

expressions suggested by Feakins et al. [53] after the transition state theory of Eyring: Δμ ∘1# ¼ RT  ln

  η0 V ∘1 hNA

ð9Þ

  RT  Δμ ∘2# ¼ Δμ ∘1# þ ∘ B− V ∘1 −V ∘2 V1

ð10Þ

where h is the Planck constant; NA is the Avogradro number; η0 and V01 are the viscosity and the molar volume of the solvent, respectively, at each temperature; V02, the limiting partial molar volume of the solute and B is the viscosity coefficient. The values of Δμ∘1# and Δμ∘# 2 at all temperatures are summarized in Table 5. ∘# As it can be ascertained, Δμ∘# 2 N Δμ1 what according to Feakins et al. means that the solute-solvent interactions are strong (as it was already concluded from the viscosity B coefficient analysis) and the formation of the activated state is less favorable compared with the ground state [51]. In addition, from the slope of the plot of the Gibbs free energy of activation per mole of solute against temperature, the entropy of activation per mole of viscous flow of solution (ΔS∘# 2 ) was obtained according to: ΔS2°# ¼ −d

Δμ ∘2# dT

! ð11Þ

and the enthalpy of activation per mole of viscous flow of solution (ΔH∘# 2 ) was determined by using the equation: ΔH 2°# ¼ Δμ ∘2# þ T  ΔS∘2#

± ± ± ± ± ± ± ± ± ± ±

0.7 0.6 0.5 0.5 0.4 0.3 0.3 0.2 0.2 0.2 0.2

1.498 1.555 1.579 1.597 1.607 1.625 1.651 1.680 1.695 1.739 1.764 1.790

± ± ± ± ± ± ± ± ± ± ± ±

0.000 0.001 0.000 0.001 0.000 0.003 0.002 0.001 0.001 0.001 0.000 0.001

1.000 1.039 1.054 1.066 1.073 1.085 1.102 1.121 1.132 1.161 1.178 1.195

Table 4 Vφ0, Sv, Bv and V0inter values for APRA at (293.15, 298.15, 303.15, 308.15 and 313.15) K. T/K

V0φ (uv)a /(cm3 ∙mol−1)

Sv (us)b /(cm3∙ kg1/2 ∙mol−3/2)

Bv(uB)c /(cm3 kg mol−2)

V0interd /(cm3 mol−1)

293.15 298.15 303.15 308.15 313.15

853.5 854.3 855.0 855.6 856.1

631 635 628 623 664

−10850 ± 405 −11039 ± 1072 −10461 ± 742 −9667 ± 1132 −11069 ± 757

210.9 211.8 212.4 213.1 213.5

a

3.5. Thermodynamic activation parameters The Gibbs free energies of activation per mole of viscous flow of both ∘# the solvent (Δμ∘# 1 ) and the solute (Δμ2 ) were calculated by using the

ð12Þ

± ± ± ± ±

0.1 0.3 0.2 0.3 0.1

± ± ± ± ±

13 34 23 36 16

V0φ is the apparent molal volume of the solute at infinite dilution. Sv is the experimental slope (related to the effect of the solute on the solvent structure). c Bv is an empirical parameter. d V0inter is an interaction volume of the solute (depending on the solute-solvent interactions). b

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.

The values found for this thermodynamics parameters at 298.15 K −1 −1 and ΔH∘# . The positive enwere ΔS∘# 2 = 2246 J mol 2 =1054 kJ mol 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. Besides that, the relative contributions to Δμ∘# 2 of both the enthalpy (ζH) and the entropy (ζTS) were calculated using the following equations:

ζH

ζ TS

¼

¼

   ∘#  ΔH2       ∘#   ∘#  ΔH 2  þ T ΔS2 

ð13Þ

    T ΔS∘2#       ∘#   ∘#  ΔH2  þ T ΔS2 

ð14Þ

The values obtained for these ζH and ζTS were 0.61 and 0.39, respectively. 4. Conclusions Resorcinarenes PRA and APRA were synthesized in good yields. The new synthesized Azaresorcinarene (APRA) was characterized by mass spectrometry, IR, 1H and 13C NMR studies. Also, absorption spectra were recorded. Such characterization concluded that only the crown isomer of APRA was obtained. Moreover, a TG analysis was performed from which no formation of solvates of APRA in this solvent (DMSO) was concluded. Using experimental density and viscosity data, values for both the apparent partial molal volume and the apparent molal expansibility at infinite dilution, as well as the viscosity B-coefficient, were determined. The positive values found for B suggest the presence of strong APRA-DMSO interactions, probably of dipole-dipole type, Table 5 Values of the viscosity B coefficient and the activation Gibbs energies of the solvent,Δμ∘1#, and of the solute, Δμ∘2#, at (293.15, 298.15, 303.15, 308.15 and 313.15) K for APRA in DMSO. T/K

B/(kg∙mol−1)

Δμ∘1#/(kJ·mol−1)

−1 Δμ∘# ) 2 /(kJ·mol

293.15 298.15 303.15 308.15 313.15

10.51 ± 0.01 9.87 ± 0.03 9.28 ± 0.04 9.00 ± 0.02 8.73 ± 0.02

14.55 14.54 14.55 14.56 14.59

402.22 384.83 368.56 362.66 357.17

between the hydroxyl groups of the resorcinol rings, aza groups and carboxylic groups. This result was confirmed by the values calculated for the Gibbs free energy of activation per mol of viscous flow of both the solvent and the solute. References [1] M.S. Deshmukh, N. Sekar, A combined experimental and TD-DFT investigation of three disperse azo dyes having the nitroterephthalate skeleton, Dyes Pigments 103 (2014) 25–33. [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. [3] Y. Egawa, R. Miki, T. Seki, Colorimetric sugar sensing using boronic acid-substituted azobenzenes, Materials 7 (2014) 1201–1220. [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-soluble reagents for metals by planar electrophoresis, Analyst 111 (1986) 351–354. [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 (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. [8] L. Abramsson-Zetterberg, N. Ilbäck, The synthetic food colouring agent Allura Red 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 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. Chem. 6 (2009) 398–411. [11] J. Han, C.-G. Yan, Synthesis, crystal structure and configuration of resorcinarene amides, J. Incl. Phenom. Macrocycl. Chem. 61 (2008) 119–126. [12] J. Sun, L.-L. Zhang, Y. Yao, C.-G. Yan, Synthesis, crystal structures and complexing properties of tetramethoxyresorcinarene functionalized tetraacylhydrazones, J. Incl. Phenom. Macrocycl. Chem. 79 (2014) 485–494. [13] V.K. Jain, P.H. Kanaiya, N. Bhojak, Synthesis, spectral characterization of azo dyes derived from calix[4]resorcinarenes and their application in dyeing of fibers, Fiber. Polym. 9 (2008) 720–726. [14] S. Elcin, M.M. Ilhan, H. Deligoz, Synthesis and spectral characterization of azo dyes derived from calix[4]arene and their application in dyeing of fibers, J. Incl. Phenom. Macrocycl. Chem. 77 (2013) 259–267. [15] A.G.S. Hoegberg, Two stereoisomeric macrocyclic resorcinol-acetaldehyde condensation 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. [17] A. Wzorek, J. Mattay, W. Iwanek, Synthesis and structural investigation of the cyclochiral boron resorcinarenes obtained from L-amino acids and boronic acid, Tetrahedron Asymmetry 23 (2012) 271–277. [18] M. Urbaniak, J. Mattay, W. Iwanek, Synthesis of resorcinarene derivatives via the catalyzed mannich reaction, part 3: glycol derivatives of resorcinarene, Synth. Commun. 41 (2011) 670–676. [19] A. Galán, E.C. Escudero-Adán, A. Frontera, P. Ballester, Synthesis, structure, and binding properties of lipophilic cavitands base on calix[4]pyrrole-resorcinarene hybrid scaffold, J. Organomet. Chem. 79 (2012) 5545–5557. [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 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 investigation of some selected solvates of calix[4]resorcinarene, J. Therm. Anal. 54 (1998) 249–255. [22] X. Wang, F. Yang, Y. Gao, Z. Liu, Volumetric properties of binary mixtures of dimethylsulfoxide with amines from (293.15 to 363.15) K, J. Chem. Thermodyn. 57 (2013) 145–151. [23] O. Ciocirlan, O. Lulian, Vapor pressure, density, viscosity and refractive index of dimethyl sulfoxide + 1,4-dimethylbenzene system, J. Serb. Chem. Soc. 73 (2008) 73–85. [24] N.G. Tsierkesos, A.E. Kelarakis, M.M. Palaiologou, Densities, viscosities, refractive indices, and surface tension of dimethyl sulfoxide + butyl acetate mixtures at (293.15, 303.15 and 313.15) K, J. Chem. Eng. Data 45 (2000) 395–398. [25] V. Govinda, P.M. Reddy, P. Attri, P. Venkatesu, V. Venkateswarlu, Influence of anion on thermophysical properties of ionic liquids with polar solvent, J. Chem. Thermodyn. 58 (2013) 269–278. [26] R. Radhamma, P. Venkatesu, T. Hoffman, Vapor-liquid equilibrium for the binary mixtures of dimethyl sulfoxide with substituted benzenes, Fluid Phase Equilib. 262 (2007) 32–36. [27] R. Ghosh, S. Banerjee, S. Chakrabarty, B. Bagchi, Anomalous behavior of linear hydrocarbon chains in water-DMSO binary mixture at low DMSO concentration, J. Phys. Chem. B 115 (2011) 7612–7620. [28] A. Ali, S. Ansari, A.K. Nain, Densities, refractive indices and excess properties of binary mixtures of dimethylsulphoxide with some poly(ethylene glycol)s at different temperatures, J. Mol. Liq. 178 (2013) 178–184. [29] P.K. Thakur, S. Patre, R. Pande, Thermophysical and excess properties of hydroxamic acids in DMSO, J. Chem. Thermodyn. 58 (2013) 226–236. [30] V. Govinda, P. Attri, P. Venkatesu, V. Venkateswarlu, Thermophysical properties of dimethylsulfoxide with ionic liquids at various temperatures, Fluid Phase Equilib. 304 (2011) 35–43. [31] NIST, Guidelines for evaluating and expressing the uncertainty of NIST measurement results, adapted from NIST technical note 1297, 1994 (edn). [32] L. Bernazani, V. Mollica, M.R. Tiné, Partial molar volumes of organic compounds in C8 solvents at 298.15 K, Fluid Phase Equilib. 203 (2002) 15–29. [33] D. Warminska, J. Wawer, W. Grzybkowski, Apparent molar volumes and compressibilities of alkaline earth metal ions in methanol and dimethylsulfoxide, J. Chem. Thermodyn. 42 (2010) 1116–1125. [34] O. Likhodi, T.V. Chalikian, Partial molar volumes and adiabatic compressibilities of a series of aliphatic amino acids and oligoglycines in D2O, J. Am. Chem. Soc. 121 (1999) 1156–1163. [35] P. Timmerman, W. Verboom, D.N. Reinhoudt, Resorcinarenes, Tetrahedron 52 (1996) 2663–2704. [36] N. Senda, Idemitsu Technical Report, 49, 2006 106–111. [37] J.P. Stewart, MOPAC2012, Stewart Computational Chemistry, Colorado Springs, CO, USA(HTTP://OpenMOPAC.net) 2012. [38] A. Nissema, A. Kuvaja, Thermodynamic properties of the binary system dimethyl sulphoxide-hexafluorobezene, Suomen Kemistilehti, B 45 (1972) 206–273. [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.

[40] A. Nissema, L. Koskenniska, Thermodynamic properties of the binary system dimethyl sulphoxide-p-Xileno, Suomen Kemistilehti B 45 (1972) 203–205. [41] V. Govinda, P. Attri, P. Venkatesu, V. Venkateswarlu, Temperature effect on the molecular interactions between two ammonium ionic liquids and dimethylsulfoxide, J. Mol. Liq. 164 (2011) 218–225. [42] L. Börnstein, Group IV physical chemistry. Pure organometallic and organononmetallic liquids, binary liquid mixtures, viscosity of pure organic liquids and binary liquid mixtures, in: Ch. Wohlfarth, B. Wohlfarth, M.D. Lechner (Eds.), Mixtures of Organic Compounds. Part 2, Springer-Verlag, Berlin Heidenberg 2001, p. 692. [43] M.I. Aralaguppi, T.M. Aminabhavi, S.B. Harogoppad, R.H. Balundgi, Thermodynamic interactions in binary mixtures of dimethyl sulfoxide with benzene, toluene, 1,3dimethylbenzene, 1,3,5-trimethylbenzene and methoxybenzene from 298.15 to 308.15 K, J. Chem. Eng. Data 37 (1992) 298–303. [44] G.S. Gokavi, J.R. Raju, T.M. Aminabhavi, R.H. Balundgi, M.V. Muddapur, Viscosities and densities of binary liquid mistures of dimethyl sulfoxide with chlorobenzene, pyridine, and methyl ethyl ketone at 25, 35, 45, and 55. degree. C, J. Chem. Eng. Data 31 (1986) 15–18. [45] P.G. Sears, T.M. Stoeckinger, L.R. Dawson, Dielectric constants, viscosities, fusion point curves, and other properties of three nonaqueous binary systems, J. Chem. Eng. Data 16 (1971) 220–222. [46] P. Umadevi, K. Rambabu, M.N. Rao, K.S. Rao, C. Rambabu, Densities, adiabatic compressibility, free-length, viscosities and excess volumes of P-cresol(L) + dimethyl sulfoxide (2), + dimethyl formamide (2), and 1,4-dioxane at 303.15–318.15 K, Phys. Chem. Liq. 30 (1995) 29–46. [47] J.M. Tsangaris, R.B. Martin, Viscosities of aqueous solutions of dipolar ions, Biochem. Biophys. 112 (1965) 267–272. [48] C.I.A.V. Santos, C. Teijeiro, A.C.F. Ribeiro, D.F.S.L. Rodrigues, C.M. Romero, M.A. Esteso, Drug delivery systems: Study of inclusion complex formation for ternary caffeine-βcyclodextrin-water mixtures from apparent molar volume values at 298.15 K and 310.15 K, J. Mol. Liq. 223 (2016) 209–216. [49] C.M. Romero, D.M. Rodriguez, A.C.F. Ribeiro, M.A. Esteso, Effect of temperature on the partial molar volume, isentropic compressibiblity and viscosity of DL-2aminobutyric acid in water and in aqueous sodium chloride solutions, J. Chem. Thermodyn. 104 (2017) 274–280. [50] H. Donald, B. Jenkins, Y. Marcus, Viscosity B-coefficients of ions in solution, Chem. Rev. 95 (1995) 2695–2724. [51] S.S. Dhondge, P.N. Dahasahasra, L.J. Paliwal, D.W. Deshmukh, Density and viscosity study of nicotinic acid and nicotinamide in dilute aqueous solutions and around the temperature of the maximum density of water, J. Chem. Thermodyn. 76 (2014) 16–23. [52] Y. Marcus, Effect of ions on the structure of water: structure making and breaking, Chem. Rev. 109 (2009) 1346–1370. [53] D. Feakins, D. Freemantle, K. Lawrence, Transition state treatment of the relative viscosity of electrolytic solutions. Applications to aqueous, non-aqueous and methanol + water systems, J. Chem. Soc. Faraday Trans. I 70 (1974) 795–806.