Sie sind auf Seite 1von 10

J. Chem.

Thermodynamics 135 (2019) 35–44

Contents lists available at ScienceDirect

J. Chem. Thermodynamics
journal homepage: www.elsevier.com/locate/jct

Solubility and vapor pressure data of bioactive 6-(acetylamino)-


N-(5-ethyl-1,3,4-thiadiazol-2-yl) hexanamide
Angelica Sharapova ⇑, Marina Ol’khovich, Svetlana Blokhina, German Perlovich
Institute of Solution Chemistry, Russian Academy of Sciences, 1 Akademicheskaya Street, 153045 Ivanovo, Russia

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

Article history: This paper presents new experimental data on some key physicochemical properties of original bioactive
Received 28 June 2018 6-(acetylamino)-N-(5-ethyl-1,3,4-thiadiazol-2-yl) hexanamide (AETH). Thermal analysis of the com-
Received in revised form 19 February 2019 pound under study has been carried out using differential scanning calorimetry and thermogravimetric
Accepted 11 March 2019
techniques. AETH solubility data in five pharmaceutically relevant solvents in the temperature range
Available online 13 March 2019
from 288.15 to 318.15 K have been obtained by the classic saturation shake-flask method. The bioactive
compound has poor solubility in hexane and buffer solutions, while it is slightly soluble in selected alco-
Keywords:
hols. The experimental solubility results have been correlated by means of modified Apelblat and van’t
Bioactive 1,3,4-thiadiazole derivative
Melting parameters
Hoff equations. The selected thermodynamic models produced acceptable results. The Hansen solubility
Solubility temperature dependences parameters and their components for AETH and solvents have been evaluated by using the van Krevelen–
Hansen solubility parameter Hoftyzer atomic group contribution method. The ideal solubility and the solution activity coefficient at
Sublimation enthalpy equilibrium in the selected solvents have been quantified based on experimental values of melting
enthalpy and temperature. All the investigated solutions exhibit a positive deviation from ideality with
the maximal solubility in alcohols. The equilibrium vapor pressure of the compound studied has been
determined as a function of temperature in the range of 433.15–454.15 K by the transpiration method.
The standard thermodynamic parameters of AETH sublimation have been calculated. It has been con-
cluded that the high crystal lattice energy value equal to 131.5 kJ/mol is the dominant factor determining
the poor solubility of the bioactive compound.
Ó 2019 Elsevier Ltd.

1. Introduction mino)-N-(5-ethyl-1,3,4-thiadiazol-2-yl) hexanamide (AETH)


containing a thiadiazole fragment as the pharmacophore (Fig. 1).
Thiadiazole is a prevalent and important five-membered hete- This compound has anaesthetic, anti-inflammatory, antiallergic
rocyclic system containing two nitrogen atoms and a sulfur atom. and analgesic activity, and in its preventive enteroprotective
Among several isomers of thiadiazole, 1,3,4-thiadiazole has been effect it excels the commercially available drug deanol
investigated more than the others [1]. Thiadiazole derivatives have aceglumate [6].
a wide spectrum of biological activity due to the strong aromaticity Studying physicochemical properties of new bioactive com-
of the cyclic system, which ensures their high in vivo stability [2]. pounds as potential drugs is one of the main problems of pharma-
The sulfur atom makes these compounds lipophilic and, thus, ceutical chemistry. Knowing these properties allows not only
enables them to easily permeate through biological membranes evaluating the characteristics that affect bioavailability of drug-
[3]. Besides, the derivatives of this series produce a low toxic effect like compounds at the in vitro stage but also predicting the ways
on the human body. Bonding different functional groups, capable and methods of improving their pharmacological profile. In the
of reacting with various receptors, to the thiadiazole ring allows present study, we provide a circumstantial summary of some
obtaining drug compounds with unique pharmacological proper- physicochemical properties of original bioactive 6-(acetylamino)-
ties [4]. This fact stimulates scientists to search for new chemical N-(5-ethyl-1,3,4-thiadiazol-2-yl) hexanamide focusing primarily
substances with a high anti-inflammatory, analgesic activities on an overview of the solubility data especially in pharmaceuti-
and relative safety of application among the new thiadiazole cally relevant solvents as well as thermophysical and sublimation
derivatives [5]. The object of the present study was 6-(acetyla parameters.

⇑ Corresponding author.
E-mail address: avs@isc-ras.ru (A. Sharapova).

https://doi.org/10.1016/j.jct.2019.03.015
0021-9614/Ó 2019 Elsevier Ltd.
36 A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44

water-methanol (42/58, v/v). The flow-rate was 0.4 ml/min. The


detector was operated at 257 nm. The injection volume was 10 ll.

Fig. 1. Chemical structure of 6-(acetylamino)-N-(5-ethyl-1,3,4-thiadiazol-2-yl) 2.2. Materials


hexanamide (AETH).
Detailed information about all the chemicals used in this study
is listed in Table 1. Buffer solutions were prepared by using bidis-
2. Experimental
tilled water with the electrical conductivity of 2.1 lS cm1. The
phosphate buffer pH 7.4 (I = 0.15 mol/dm3) was made by mixing
2.1. Synthesis and identification
KH2PO4 (9.1 g in 1 l) and Na2HPO412H2O (23.6 g in 1 l) salts. In
order to prepare the buffer solution pH 2.0 (I = 0.10 mol/dm3)
We placed 100 ml of acetonitrile, 12.91 g (0.1 mol/dm3) of 2-
6.57 g of KCl was dissolved in water, 119.0 ml of 0.1 mol/dm3
amino-5-ethyl-1,3,4-thiadiazole into a three-neck flask equipped
hydrochloric acid was added and the volume of the solution was
with an agitator, a thermometer and a fridge; stirred the mass
adjusted to 1 l with water. The pH values were measured by using
and gradually added 16.65 g (0.1 mol/dm3) of N-acetylamino hex-
a pH meter FG2-Kit (Mettler Toledo, Switzerland) standardized
anoic acid chloride. Then we heated the obtained mixture to the
with pH 1.68, 6.86 and 9.22 solutions.
boiling point, 3 h later removed the acetonitrile, added 150 ml of
distilled water and mixed carefully. The obtained solution was
heated to 80 °C, and 15–20 min later the residue was filtered and 2.3. Apparatus and procedure
washed on a filter with distilled water. The washing continued till
the backwash water reached pH 6–7. Then the residue was dried 2.3.1. Differential scanning calorimetry (DSC)
till it reached a constant weight at 100–105 °C. As a result, we The melting temperatures and enthalpies of the compounds
obtained 26.48 g of white crystalline powder representing 6-(acet under investigation have been determined using a Perkin-Elmer
ylamino)-N-(5-ethyl-1,3,4-thiadiazol-2-yl) hexanamide. Pyris 1 DSC differential scanning calorimeter (Perkin-Elmer Analyt-
The elemental analysis of the compound obtained was per- ical Instruments, Norwalk, Connecticut, USA) with Pyris software
formed on a CHN analyzer Carbo Erba (Czech Republic). Anal. for Windows NT. The DSC runs were performed in an atmosphere
Calcd. for C12 H20 N4 O2 S: C 50.68%, H 7.09%, N 19.71%, O of 20 cm3min1 of flowing dry helium gas of high purity (0.99996
11.25%, S 11.27%. Found: C 50.91%, H 7.23%, N 19.56%, S 11.23%. mass fraction) using standard aluminum sample pans at the heat-
The FT-IR spectrum of the compound synthesized was recorded ing rate of 2 K∙min1 Sealed and holed aluminum pans were used
by using a Bruker Vertex 80v spectrometer (Bruker Optik GmbH, for the experiment. An empty pan sealed in the same manner
Ettlingen, Germany) with the spectral resolution of 2 cm1 in the was used as the reference. The accuracy of weight measurements
spectral range 4000 – 400 cm1 at room temperature. The mea- was 0.005 mg.
sured vibrations (m, cm1) were 3310 and 3169 cm1(NH); 2988, The DSC was calibrated using a two-point calibration, measur-
2959, 2940 and 2860 (CH); 1669 (CO), 1560 (CONH), 1030 (C@S). ing the onset temperatures of indium and bismuth standards.
In addition, nuclear magnetic resonance spectra of the compound The onset of melting was used for calibration because it is almost
isolated were recorded in a Bruker Advance III HD 400 (400 MHz independent of the scan rate. The melting temperatures of indium
to 1H and 100 MHz to 13C analysis) (Bruker, Germany) in deuter- and bismuth were 429.7 K and 544.5 K, respectively (determined
ated chloroform (CDCl3, 99.9%). Graphical 1H NMR spectra of AETH by at least five measurements). The enthalpy scale was calibrated
have been provided in Fig. S1 (Supporting Information). The mass using the indium fusion heat. The measured fusion enthalpy value
fraction purity of AETH after recrystallization was determined by equaled 28.69 J∙g1 (the reference value 28.66 J∙g1) [7]. The
high performance liquid chromatography on an Agilent 1100 series expanded uncertainty (0.95 level of confidence) on the melting
apparatus with a Kinetex C18, 2.6 lm, 3  100 mm (Phenomenex, temperature was determined as twice the standard deviation of
USA) (Table 1). The mobile phase consisted of a mixture of five independent measurements.

Table 1
Description of the materials used in the experiments.

Chemical name CAS Formula M/g Source Initial mass Method of Final mass Analysis
register mol1 fraction purity purification fraction purity method
No.
6-(acetylamino)-N-(5-ethyl-1,3,4- 2184830- C12 H20 N4 O2 S 284.38 synthesis 0.95 recrystallization 0.98 HPLCa
thiadiazol-2-yl) hexanamide 56-4
Ethanol 64-17-5 C2H6O 46.07 Sigma- 0.99b – – –
Aldrich
1-Octanol 111-87-5 C8H18O 130.2 Sigma- 0.99b – – –
Aldrich
b
n-Hexane 110-54-3 C6H14 86.18 Sigma- 0.97 – – –
Aldrich
b
Potassium dihydrogen phosphate 7778-77- KH2PO4 136.08 Merck 0.99 – – –
0
Disodium hydrogen phosphate 10039- Na2HPO412H2O 358.14 Merck 0.99b – – –
dodecahydrate 32-4
Potassium chloride 7447-40- KCl 74.55 Sigma- 0.99b – – –
7 Aldrich
b
Benzoic acid 65-85-0 C7H6O2 122.12 Sigma- 0.995 – – –
Aldrich
a
High performance liquid chromatography.
b
As stated by the supplier.
A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44 37

2.3.2. Thermogravimetry (TG)


Thermogravimetric analysis was performed on a TG 209 F1 0.32
thermo-microbalance (Netzsch Gerätebau GmbH, Germany) using
0.28

-1
platinum crucibles in an atmosphere of dry argon at a flow rate of

m/mol kg
30 cm3∙min1 and a heating rate of 10 Kmin1. The sample weight
0.24
was 3–5 mg. The accuracy of the mass measurement was 1107 g.
The thermal curve reproducibility was monitored by performing 0.20
up to three replicated experiments.
0.16

2.3.3. Solubility determination 0.12


The saturated equilibrium solubility of AETH was determined 0.08
by the classical shake flask method at atmospheric pressure and
in the temperature range from 288.15 K to 318.15 K. The essence 0.04
of the above mentioned method consists in determination of the
compound concentration in the saturated solution. The indepen- 280 290 300 310 320 330 340 350 360
dent experiments were carried out in parallel in triplicate. The
solid sample was added carefully using a spatula to 5–10 ml of T/K
the solvents in a glass vial, while stirring until a heterogeneous sys-
Fig. 2. Plot of solubility (m) of benzoic acid in water as a function of temperature
tem (solid sample and liquid) was obtained. The concentrated sus-
(T). j – Reference [8]; ▲ – reference [9]; d – reference [10]; r – this work.
pensions of the compound in each solvent were shaken
continuously in an air thermostat containing a stirring device.
The point of the solution thermodynamic equilibrium was deter- benzoic acid solubility measured by us is in good agreement with
mined based on the solubility kinetic dependences and averaged literature data. The deviation of the measured solubilities from the
72 h. The sedimentation time of solid phase after stirring was literature values did not exceed 2%, so it was perform that this
4 h. After saturation was achieved, the solution aliquot was taken experimental technique was reliable.
and centrifuged in a centrifuge Biofuge stratus (Germany) for The chemical nature of the AETH solid phase in equilibrium
5 min at a fixed temperature. Then, the supernatant solutions were with the selected solvents after the solubility experiment was
filtered through a 0.20 lm filter MILLEXÒHA (Ireland). The satu- characterized by the DSC technique. No variation has been found
rated solution was diluted with the corresponding solvent to the in the DSC profile of the samples, indicating that neither salts nor
required concentration. solvates were formed.
The AETH concentration in molarity was determined by mea-
suring the absorbance at the wavelength of 253 nm using a UV–
2.3.4. Vapor pressure measurements
vis spectrophotometer (Cary-50, USA) at room temperature. The
The temperature dependence of AETH vapor pressure was mea-
experimental results are reported as an average value of at least
sured by the gas saturation method, also known as the transpira-
three replicated experiments and presented in Table S1 (Support-
tion method. A detailed description of the apparatus had been
ing Information).
presented previously [11]. The equipment was tested using ben-
The calibration was made at room temperature using the solu-
zoic acid. The verification experiment details and the comparison
tions with known concentrations of each substance in each inves-
of the obtained results with the literature data had been described
tigated solvent. The solutions were prepared by adding a substance
previously [12]. The standard value of the sublimation enthalpy at
of an appropriate mass and a certain volume of solvent (buffer pH
T = 298.15 K obtained in our experiments was
2.0, buffer pH 7.4, 1-octanol, ethanol, n-hexane) to the flask and
Dgcr Hom = 90.5 ± 0.3 kJ∙mol1. This was in good agreement with the
mixing until the substance was totally dissolved. The absorbance
value recommended by IUPAC (Dgcr Hom =89.7 ± 0.5 kJ∙mol1) [13].
of the solutions was measured and calibration curves were con-
The procedure in brief was as follows: approximately 0.5 g of
structed. The calibration curve was found to be linear in the con-
the investigated compound was blended with pyrex balls and
centration range of 0.2–1.0 lgml1 with a correlation coefficient
placed into a thermostated tube. At a constant temperature, the
of 0.9995.
flow of inert gas (N2) transported the saturated vapor of the sample
The conversion of molarity to the mole fraction concentration
under investigation through the tube until it was completely con-
scale was produced by the following equation:
densed at some point downstream. The vapor pressure of the sam-
M2 S ple at this temperature was calculated from the amount of the
x¼ ; ð1Þ sublimated sample and the volume of the inert gas used. The sta-
SðM2  M 1 Þ þ 1000q
bility of the gas flow was supported by a mass flow controller
where S is the molarity (mol/dm3), q is the density of solution MKS type 2179A. It is a prerequisite of the method that within
(g/cm3), M1 and M2 are the molar masses of AETH and solvent, the temperature interval used and during the whole experiment
respectively. The mole fractions for buffer solutions were calculated the compound does not get chemically decomposed in the measur-
with molar mass of water (without account of the buffer composi- ing cell and in the sublimated products, and the stripping gas vapor
tions). The molarity values of compound studied have been pre- pressure does not depend on the temperature. From the experi-
sented in Table S1 (Supporting Information). The densities of the mentally determined pressure – flow rate relationship, the optimal
solutions were measured at 288.15–318.15 K and atmospheric pres- flow rate of 1.69 dm3∙h1 was found. At this flow rate, the satu-
sure using an oscillation U-tube density meter DMA 4500 (Anton rated vapor pressure was independent of the flow rate and, thus,
Paar, Austria). The experimental values of the densities for the inves- the thermodynamic equilibrium was reached.
tigated solutions are given in Table S2 (Supporting Information). The sublimed substance amount is determined by the following
Standard uncertainty for density values was 0.004 g∙cm3. procedure. The condensed substance is dissolved in a definite vol-
The experimental setup and its accuracy were validated by ume of solvent Vsol. The mass of the substance is determined based
comparing the solubility data of benzoic acid in water with those on the measuring of absorbance A of its solution with a spec-
in literature [8–10] (Fig. 2 and Table S3). As shown in Fig. 2, the trophotometer (Cary, USA). Knowing the value of the extinction
38 A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44

coefficient e (dm3mol1cm1) of the studied compound dissolved


in the solvent one can express the concentration of the solution c
(moldm3) according to the Lambert-Beer law, by the following
relation:
A ¼ ecl ð2Þ
whereas the mass of the sublimed substance is calculated from:
m ¼ cV sol M ð3Þ
where l is the absorbing path length; M is the molar mass of the
studied substance.
The absolute vapor pressure p at each temperature T was calcu-
lated from the amount of the sublimed product within a definite
period. Considering that the vapor pressure of the substance was
very low, we were able to apply the ideal gas rule to the nitrogen
stream saturated with the substance, the values of p were calcu-
lated with Eq. (4): Fig. 3. TG (a) and DSC (b) curves of AETH.
p ¼ m  R  T a =V  M ð4Þ
where R is the universal gas constant, M is the molar mass of the tively. The subsequent exothermic peak at the temperature of
compound; Ta is the ambient temperature, V = VN2 + Vi (VN2 >> Vi) 275 °C corresponds to AETH decomposition, which agrees with
is the total volume of the gas phase consisting of the carrier gas the thermogravimetric data. The melting enthalpy was determined
and the gaseous compound under study at the atmospheric pres- by integrating the melting curve with the straight baseline and
sure. The gas volume VN2 (dm3) is calculated by the equation: equaled 35.2 ± 0.5 kJmol1. Uncertainties for melting parameters
correspond to expanded uncertainties of the mean (0.95 confi-
V N2 ¼ v  t ð5Þ
dence level). The melting entropy of AETH was calculated by equa-
where m (dm /h) is the gas flow velocity; t (h) is the sublimation
3
tion DSm = DHm/Tm. The value of DSm was found to be equal to
time period. 127.9 ± 4.2 J∙mol1K1.
The saturated vapor pressure values were measured five times
at each temperature with the standard deviation of no more than
3.2. Solubility experimental data
5%. Because the saturated vapor pressure of the investigated com-
pounds was low, we could assume that the change in the vapor
For the study, we selected five pharmaceutically relevant sol-
heat capacity with temperature was so small that it could be
vents: buffer pH 2.0 simulating the gastric fluid, buffer pH 7.4 –
neglected.
the blood system plasma, 1-octanol as the model of the properties
The data of vapor pressure were obtained as a function of tem-
of biological membranes phospholipids; hexane a component of
perature and were fitted using the following equation:
the blood–brain barrier media. Ethanol is a typical medium used
lnðp=PaÞ ¼ A þ B=T ð6Þ for delivering drugs in the pharmaceutical industry and is also
employed as a preservative, an extraction agent and a co-solvent.
The value of the sublimation enthalpy at the mean temperature
The measured mole fraction solubilities of AETH in five selected
T was calculated by the Clausius–Clapeyron equation:
solvents in the temperature ranging from 288.15 to 318.15 K are
 
@ðlnpÞ listed in Table 2, and graphically presented in Fig. 4A. As these data
Dcrg Hom ðTÞ ¼ R ð7Þ
@ð1=TÞ show, the solubility of the compound studied monotonically
increases with the rising temperature for all the studied solvents,
whereas the sublimation entropy at the given temperature T was which is characteristic of most organic substances. Ethanol has
calculated from the following relation: the highest temperature gradient of solubility facilitating the
ðDcrg Hom ðTÞ  Dcrg Gom ðTÞÞ solute–solvent interaction and enhancing dissolution power.
Dcrg Som ðTÞ ¼ ð8Þ Fig. 4B further shows that the mole fraction solubilities
T
decrease in the following order in different solvents: ethanol >
with Dcrg Gom ðTÞ = -RT ln(p/p0), where p0 = 105 Pa. 1-octanol > buffer pH 7.4 > buffer pH 2.0 > n-hexane at the temper-
ature below about 298.15 K, it is smaller in ethanol than in
3. Results and discussion 1-octanol. In order to qualitatively evaluate the solubility of the
bioactive compound studied in accordance with United States
3.1. Thermal analysis Pharmacopeia, we expressed the molarity of AETH (Table S1) at
T = 298.15 K in mg∙ml1 units and obtained the following values:
Fig. 3 shows the results of the thermal analysis of the studied buffer pH 2.0–0.54, buffer pH 7.4–0.46 n- hexane  0.04, ethanol
compound as TG (a) and DSC (b) curves. As it can be seen, AETH – 4.6, 1-octanol – 1.7. As can be seen from the data, AETH was con-
shows good thermal stability in the temperature range 30– sidered essentially insoluble n-hexane (S < 0.1 mg∙ml1), very
230 °C. No AETH mass loss is observed in this range, which indi- slightly soluble in buffer solutions (0.1 < S < 1 mg∙ml1) and
cates that this compound does not contain low-molecular volatile slightly soluble in 1-octanol and ethanol (S > 1 mg∙ml1). At a fixed
components. The main decomposition occurs between the temper- temperature, the solubilities are highest in ethanol, and lowest in
atures of approximately 250 and 300 °C. n-hexane. The solubility of AETH in ethanol at the temperatures
The DSC curve of sample heating at the temperature 190 °C has above 298.15 R is better than in 1-octanol. This is consistent with
a sharp exothermic peak that is evidently caused by compound the observation that lower molecular weight solvents are often
melting. The onset and maximum temperatures of the melting ‘‘stronger” than higher molecular weight solvents of the same
peak are Tonset = 205.2 ± 0.2 °C and Tmax = 208.5 ± 0.2 °C, respec- class. The drug solubility of the compound studied in muriatic buf-
A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44 39

Table 2
Experimental (xexp) and correlated (xcal) AETH solubility values (in mole fractions) in the organic solvents at different temperatures and pressure p = 0.1 MPa.

T/R xexp Modified Apelblat equation van’t Hoff equation


a
xcal RD xcal RD
Buffer pH 2.0b
288.15 2.49105 2.49105 0.0003 2.30105 0.0752
293.15 2.95105 2.93105 0.0067 2.71105 0.0824
298.15 3.40105 3.43105 0.0085 3.16105 0.0692
303.15 3.94105 3.99105 0.0124 3.68105 0.0657
308.15 4.67105 4.62105 0.0117 4.26105 0.0877
313.15 5.37105 5.31105 0.0107 4.91105 0.0860
318.15 6.03105 6.08105 0.0091 5.63105 0.0664
Buffer pH 7.4c
288.15 2.08105 2.08105 0.0043 2.04105 0.0185
293.15 2.46105 2.46105 0.0002 2.46105 0.0004
298.15 2.90105 2.90105 0.0081 2.95105 0.0166
303.15 3.41105 3.41105 0.0189 3.51105 0.0296
308.15 4.23105 4.23105 0.0243 4.16105 0.0171
313.15 4.91105 4.93105 0.0057 4.90105 0.0068
318.15 5.76105 5.76105 0.0106 5.74105 0.0039
n-Hexane
288.15 1.54105 1.53105 0.0028 1.51105 0.0162
293.15 1.81105 1.81105 0.002 1.81105 0.0026
298.15 2.15105 2.14105 0.0037 2.16105 0.0048
303.15 2.47105 2.53105 0.0243 2.56105 0.0353
308.15 3.05105 2.99105 0.0204 3.01105 0.013
313.15 3.55105 3.53105 0.0062 3.53105 0.0069
318.15 4.13105 4.16105 0.0084 4.11105 0.0053
Ethanol
288.15 5.78104 5.74104 0.0067 5.62104 0.0281
293.15 7.33104 7.40104 0.0097 7.41104 0.0104
298.15 9.65104 9.55104 0.0108 9.67104 0.0025
303.15 1.18103 1.23103 0.0350 1.25103 0.0526
308.15 1.62103 1.59103 0.0188 1.60103 0.0072
313.15 2.10103 2.06103 0.0232 2.05103 0.0244
318.15 2.60103 2.65103 0.0182 2.59103 0.0039
1-Octanol
288.15 6.85104 6.82104 0.0047 6.73104 0.0172
293.15 8.14104 8.13104 0.0011 8.15104 0.0018
298.15 9.61104 9.70104 0.0093 9.82104 0.0214
303.15 1.13103 1.16103 0.0244 1.17103 0.0392
308.15 1.42103 1.38103 0.0269 1.40103 0.0164
313.15 1.67103 1.65103 0.0123 1.65103 0.0108
318.15 1.94103 1.97103 0.0151 1.94103 0.0019

Standard uncertainties: u(T) = 0.15 K and u(p) = 3 kPa.


Relative standard uncertainties for solubility: ur(x) = 0.045 for buffer solutions and ur(x) = 0.04 for n-hexane, ethanol and 1-octanol.
a
RD is the relative deviation: RD = (xexp – xcal)/xexp.
b
Composition of aqueous buffer pH 2.0: KCl (6.57 g in 1 l) and 0.1 mol/dm3 hydrochloric acid (119.0 ml in 1 l).
c
Composition of aqueous buffer pH 7.4: KH2PO4 (9.1 g in 1 l) and Na2HPO412H2O (23.6 g in 1 l).

fer (pH 2.0) and phosphate buffer (pH 7.4) is the same, which B
ln x ¼ A þ þ ClnðT=K Þ ð9Þ
allows us to make a conclusion that the hydrogen index of aqueous T=K
solutions does not affect the dissolution process. The dissociation
where x is the experimental saturated mole fraction solubility of the
constant values pKa = 9.52 ± 0.50 calculated using Advanced
solute, T is the absolute temperature, A, B and C are the empirical
Chemistry Development (ACD/Labs) Software V11.02 show that
model parameters. Values A and B represent the variation in the
the studied compound is a weak acid, which is caused by the pres-
solution behavior resulting from the non-ideality of the solute sol-
ence of two secondary groups in the molecule structure. So, we can
ubility, whereas the value of C represents the association between
conclude that AETH is in an almost totally nonionized form in the
the temperatures and the enthalpy of fusion [16].
selected buffer solutions.
According to the van’t Hoff model, the mole fraction solubility
of a compound in the solvents studied is calculated using equation
3.3. Modeling of solubility data [17]:
B
In order to more quantitatively describe the solid–liquid equi- ln x ¼ A þ ð10Þ
ðT=K Þ
librium, the relationship between the experimental solubility and
the temperature was modeled by the modified Apelblat and van’t where x is the mole fraction solubility of the solute, A and B are the
Hoff equations. The modified Apelblat equation [14,15] has been model constants calculated using the least square analysis.
extensively used in correlation and determination of solubility The relative average deviation (RAD) and root-mean-square
data of different substances because of its simplicity and is deviation (RMSD) were used to evaluate the applicability and accu-
expressed as: racy of the models:
40 A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44

-6
buffer pH 2.0
buffer pH 7.4
2.0x10
-3 n-hexane
ethanol
-8

slightly soluble
soluble
1-octanol

lnx
-3
1.0x10

slightly
x

-5 -10
6.0x10

very slightly soluble

oc .4 very slightly soluble


insoluble
-5
4.0x10
-12

-5
2.0x10

l
bu r p e
r p .0

no

l
7
ffe ah

no
290 295 300 305 310 315 320

ffe H 2

ta
H
bu hex

ha
et
T/K

1-
n -
A B
Fig. 4. Solubility data of AETH in selected solvents: A  mole fraction solubility (x) of AETH at different temperatures, B – solubility histogram at 308.15 K.

N  
1X 
xexp  xcal 
organic substances is played by three types of interactions (disper-
RAD ¼  ð11Þ
N i¼1 xexp  sion interaction caused by interatomic forces of attraction; inter-
molecular dipole–dipole interaction; intermolecular hydrogen
 1=2 interaction that is seen as electrostatic interaction enhanced by
1 XN
 2 
 small hydrogen size) Hansen divided the solubility parameter into
RMSD ¼  xexp  xcal  ð12Þ
N i¼1  three components – dispersion (dd), polar (dp) and hydrogen bond-
ing interaction (dh). The total solubility value is then determined by
where N represents the number of experimental points, xexp and xcal the following equation:
are the experimental and calculated mole fraction solubility values
of the compound, respectively. d2t ¼ d2d þ d2p þ d2h ð13Þ
The mole fraction solubilities of the studied compound calcu-
lated by Eqs. (9), (10) are reported in Table 2. The calculated model The partial solubility parameters for AETH and selected solvents
parameters with the values of deviations are presented in Table 3. have been estimated using the combined group contribution meth-
As can be seen from the presented data, the modeling results ods of Van Krevelen–Hoftyzer and Fedors [21,22]. The group con-
obtained by the modified Apelblat and van’t Hoff equations show tribution parameters for the respective molecular forces and the
good consistency with the experimental solubility with RAD values associated molar volumes with the number of the corresponding
less than 2% and these models are all suitable for correlating the group for the compound studied are listed in Table 4. The HSP
solubility of AETH in the selected solvents. Obviously, the modified parameters of AETH and selected solvents as well as the calculated
Apelblat equation has the smallest values of deviations for the molar volume at 298.15 K are summarized in Table 5.
investigated systems. It is known that the similarity of HSP values of two substances
determines the degree of their interaction: the closer are the solute
3.4. Hansen solubility parameter and solvent parameters, the more effective is the dissolution pro-
cess. The calculated HSP values agree well with the experimental
The solubility behavior of AETH has been illustrated by using solubility values obtained by us. According to the data from Table 5,
the Hansen solubility parameter (HSP). HSP help to quantify the the best solvents for AETH are alcohols as their common dt param-
statements ‘like dissolves like’, and has been widely used to predict eters are closest to each other. It should be said that higher values
materials’ compatibilities, including in pharmaceutical drug devel- of all the HSP components in ethanol in comparison with those in
opment [18–20]. Taking into account the fact that the main role in 1-octanol lead to a three-time difference in the Dd value, which

Table 3
Parameters of modified Apelblat and van’t Hoff equations for AETH in the selected solvents.

Solvents A B C 105RMSD RAD


Modified Apelblat equation
Buffer pH 2.0 12.77 3357.8 20.69 0.043 0.0085
Buffer pH 7.4 137.65 3055.1 20.53 0.053 0.0103
n-Hexane 133.61 2951.9 19.83 0.035 0.0097
Ethanol 203.79 4910.5 31.66 1.309 0.0175
1-Octanol 144.90 3469.1 22.17 1.265 0.0134
van’t Hoff equation
Buffer pH 2.0 1.12 2731.6 – 0.033 0.0761
Buffer pH 7.4 0.16 3157.8 – 0.055 0.0133
n-Hexane 0.52 3048.3 – 0.149 0.0120
Ethanol 8.72 4670.3 – 0.971 0.0184
1-Octanol 3.94 3240.1 – 2.204 0.0155
A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44 41

Table 4
Group contribution parameters and associated molar volumes of AETH.

Individual functional group Frequency Fdi, (Jm3)0.5∙mol1 Fpi, (Jm3)0.5∙mol1 Ehi, J/mol Vi, cm3/mol
ACH3 2 336.6 0 0 33.5
ACH2A 6 234.6 0 0 16.1
>C< 2 214.2 0 0 19.2
ACOA 2 105.0 600 9500 10.8
NHA 2 122.4 700.7 1500 4.5
ASA 1 815.9 196.0 297.5 12.0
AN@ 2 380.0 100 250 5.0
Ring(5) 1 142.8 0 0 16.0

Table 5
Molar volumes and Hansen solubility parameters for AETH and selected solvents.

Compound V, cm3∙mol1 a
dd, MPa0.5 b
dp, MPa0.5 c
dh, MPa0.5 dt, MPa0.5 Dd
d
dv, MPa0.5
AETH 193.8 19.7 4.9 10.9 23.0 – 19.7
Buffer solutions 18.0 15.5 16.0 42.3 47.8 29.1 4.0
n-Hexane 131.6 14.9 0.0 0.0 14.9 12.9 14.9
Ethanol 58.5 15.8 8.8 19.4 26.5 10.1 13.1
1-Octanol 157.7 17.0 3.3 11.9 21.0 3.3 16.7
a
dd = RFdi/RVi.
b
dp = (RF2pi)0.5/RVi.
c
dh = (RFhi/RVi)0.5.
0:5
d
Dd ¼ ððdd1  dd2 Þ2 þ ðdp1  dp2 Þ2 þ ðdh1  dh2 Þ2 Þ .

should correspond to a significant difference in AETH solubility in The relation of dv versus dh, enables a projection of the three-
these solvents. However, the solubility of the considered com- dimensional solubility parameter into a two dimensional plot.
pound in alcohols is almost equal. The situation may be due not The bioactive AETH is placed in the center of the circle with coor-
only to solute–solvent interactions, but solvent–solvent interac- dinates dv = 19.7 MPa0.5 and dh = 10.9 MPa0.5. Earlier, using the poor
tions and molecular shapes and sizes. Hexane, only capable of dis- soluble drugs data set, Albers [24] has shown that two substances
persion interactions (dt = dd = 14.9), behaves as a very weak are miscible if their interaction radius in the Bagley
solvent, which results in very low solubility values (x = 105 mol. plot  5.6 MPa0.5. It is shown that only the point corresponding
frac.).. The poor solubility of the studied compound in buffer solu- to ethanol lies inside the solubility sphere, indicating that this alco-
tions is due to its highest dt value in water, with the maximum dif- hol can be a good solvent for AETH. As the presented diagram indi-
ference in hydrogen bonding parameters, which confirms the cates, the values of the other studied solvents lie beyond the area
hydrophobic character of AETH. of optimal solubility parameters, with buffer solutions and hexane
The results of the HSP calculations have been illustrated by being the farthest from the center. Thus, the Bagley’s plot provides
using the Bagley diagram (Fig. 5). Based on the thermodynamic a good approximation of the AETH behavior in the investigated sol-
considerations that dd and dp show similar values, whereas the vents with the measured solubility data.
effect of dh is of quite different nature, Bagley et al. [23] introduced The approach based on solubility parameters has been used as
a combined solubility parameter dv, which is defined as: an effective method to estimate the absorption behavior of new
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi substances in drug research and development [20]. By analysing
dv ¼ d2d þ d2p ð14Þ 32 known drug substances of different chemical nature, the author
found a region for optimal absorption with the centre coordinates
dv = 20.3.MPa0.5 and dh = 11.3 MPa0.5.. The bioactive compound that
we have studied has almost the same solubility parameters, which
allows us to make a conclusion that AETH can be absorbed along
the whole gastrointenstinal tract with maximal absorption
duration.
20
1-Octanol 3.5. Ideal solubility and activity coefficients of AETH
0.5

16 n-Hexane
, MPa

Ideal solubility is defined as the solubility of a solute in the per-


Ethanol fect solvent, for which there is no energy contribution associated
12 R with the dissolution process. Ideal solubility depends only on the
v

AETH
energy needed to break the crystalline structure of the compound
8 and does not take into account the solvent properties. If the ther-
Buffer mophysical parameters of the solute are known, the ideal solubility
solutions can be accurately calculated by the equation:
4 
 DH m T m
lnxid ¼ ð15Þ
0 10 20 30 40 RT m T
0.5
h
, MPa where xid is the ideal mole fraction solubility, Tm and DHm are the
temperature and enthalpy change of melting for the pure solute.
Fig. 5. Bagley diagram of AETH solubility (R = 5.6 – interaction radius). This approach has been especially frequently used by many
42 A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44

Table 6
Activity coefficients (c) of AETH in selected solvents at different temperatures and pressure p = 0.1 MPa.

Solvent c
288.15 K 293.15 K 298.15 K 303.15 K 308.15 K 313.15 K 318.15 K
Buffer pH 2.0 467.27 464.47 462.81 461.18 459.67 457.17 455.50
Buffer pH 7.4 555.24 545.66 537.03 529.38 521.07 516.92 505.68
n-Hexane 749.94 742.48 735.09 727.78 720.54 713.37 706.27
Ethanol 20.00 18.30 16.14 15.17 13.52 12.02 11.14
1-Octanol 16.85 16.50 16.21 15.97 15.53 15.20 15.03

The standard uncertainties: u(T) = 0.15 K, u(p) = 3 kPa.


Relative standard uncertainty: ur(c) = 0.04.

Table 7
Transpiration experiment parameters, vapor pressure of AETH at different temperatures and sublimation thermodynamic functions.

T/K m/mg V(N2)/dm3 Ta/K m/dm3h1 p/Pa


433.15 0.0144 9.700 299.15 1.69 0.0130
435.15 0.0193 10.917 297.15 1.69 0.0153
437.15 0.0215 10.478 297.15 1.69 0.0179
439.15 0.0207 8.653 298.15 1.69 0.0208
441.15 0.0282 10.275 298.15 1.69 0.0239
443.15 0.0235 7.520 299.15 1.69 0.0273
445.15 0.0336 8.957 297.65 1.69 0.0327
446.65 0.0396 9.413 299.15 1.69 0.0368
446.65 0.0428 9.413 299.15 1.69 0.0398
450.15 0.0454 8.196 298.15 1.69 0.0483
452.15 0.0549 8.653 299.15 1.69 0.0554
454.15 0.0269 3.634 298.15 1.69 0.0644
p(298.15 K)/Pa 2.21109
Dcrg Gom ð298:15 KÞ/kJ∙mol1 77.9 ± 1.6
Dcrg Hom ðT Þ/kJmol1 124.0 ± 1.2
C p (298.15 K)/J∙mol1K1 342.8
Dcrg Hom ð298:15 KÞ/kJ∙mol1 131.5 ± 1.2
T Dcrg Som ð298:15 KÞ/kJ∙mol1 53.6
Dcrg Som ð298:15 KÞ/J∙mol1∙K1 179.8 ± 5.1

The standard uncertainties: u(T) = 0.15 K, u(Ta) = 0.15 K, u(m) = 0.00001 g.


Relative standard uncertainty: ur(p) = 0.05.

researchers in the pharmaceutical field [25–27] and has been these organic solvents; d) the activity coefficients decrease with
shown by Grant [28] to provide a better fit with respect to available the temperature growth, which indicates a strengthening
experimental solubility data. solute–solvent molecular interaction; e) the c values of AETH
The ideal solubilities of AETH calculated by means of Eq. (15) corresponded to its solubility data: the lower was the activity coef-
from the Hm and Tm values obtained by us from the DSC experi- ficient, the higher was the compound solubility in this solvent.
ment are presented graphically in Fig. S2 (Supporting Information).
As this data shows, the ideal solubility of the studied substance
increases with temperature growth. In this temperature range, 3.6. Crystal lattice energy
the xid values vary from 1.15102 to 2.92102, which is three
orders of magnitude greater than the experimental solubility in According to the thermodynamic cycle, sublimation and solva-
buffer solutions and hexane and one order of magnitude higher tion parameters are the key properties connected with the under-
than in the selected alcohols. This phenomenon is typical, when standing of the fundamental aspects of dissolution process [29].
the solubility predicted based on the model of an ideal solution The enthalpy of sublimation characterizing lattice energy is an
is normally much higher than the solubility that is actually mea- important property of the solid state as it represents a quantitative
sured for a non-ideal solution. assessment of the intermolecular interactions in solid phase. It can
The activity coefficients (c) for AETH in the studied solvents be directly measured by vapor–pressure measurements. In our
were determined using the following equation: study, we used the transpiration method that is especially success-
fully applied at low vapor pressure values. Among its merits are:
xid sufficiently high accuracy, low sensitivity to impurities and short
c¼ ð16Þ duration to reach equilibrium [30].
xexp
The vapor pressure experimental values of AETH in the temper-
The calculated c values in each solvent in the studied tempera- ature range 433.15–454.15 K are summarized in Table 7 and
ture interval are listed in Table 6. graphically presented in Fig. 6. The obtained data are well approx-
By analysing the obtained data, we have found the following imated by the linear dependence: ln(p/Pa) = (30.08 ± 0.32) –
regularities: a) all the investigated solutions have a positive devia- (1491.0 ± 14.1)/T (R = 0.9996). The sublimation parameters of
tion from the ideality (c > 1); b) the biggest deviation from the ide- AETH calculated by Eqs. (7)–(8) are also represented in Table 7
ality is observed for buffer solutions and hexane with the along with their standard deviation. The sublimation enthalpy
maximum values of the activity coefficients; c) the minimum c obtained by the integrated form of the Clausius-Clapeyron equa-
value for AETH was obtained in ethanol and 1-octanol and is due tion corresponds to the average value over the experimental tem-
to stronger interactions between the bioactive compound and perature range studied.
A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44 43

The solutions in all the solvents exhibit a positive deviation from


-2.5 the ideality, with the highest solubility in alcohols. Moreover, the
temperature dependence of vapor pressure for the studied com-
-3.0
pound has been measured by the transpiration method. The crystal
ln(p/Pa)

lattice energy of AETH equal to 131.5 kJmol1 has been evaluated


from the experimental vapor pressure values. We have concluded
-3.5 that the thermodynamic reason for the low solubility of AETH is
the high value of energy required to break the crystal lattice.

-4.0
Appendix A. Supplementary data

-4.5
Supplementary data to this article can be found online at
https://doi.org/10.1016/j.jct.2019.03.015.

2.18 2.20 2.22 2.24 2.26 2.28 2.30 2.32


3 -1 -1 References
10 T /K
[1] Y. Hu, C.-Y. Li, X.-M. Wang, Y.-H. Yang, H.-L. Zhu, 1,3,4-Thiadiazole: synthesis,
Fig. 6. Plot of vapor pressure ln(p/Pa) against reciprocal temperature for AETH. reactions, and applications in medicinal, agricultural, and materials chemistry,
Chem. Rev. 12 (2014) 5572–5610.
[2] A. Castro, T. Castano, A. Encinas, W. Porcal, C. Gil, Advances in the synthesis and
recent therapeutic applications of 1,2,4-thiadiazole heterocycles, Bioorg. Med.
The standard molar enthalpies of sublimation Dcrg Hom ð298:15 KÞ Chem. 14 (2006) 1644–1652.
have been determined from the experimental data Dcrg Hom ðT Þ using [3] Y. Li, J. Geng, Y. Liu, S. Yu, G. Zhao, Thiadiazole  a Promising Structure in
Medicinal Chemistry, ChemMedChem 8 (2013) 27–41.
the heat capacity differences of both solid (C pc ) and gaseous (C pg ) [4] A.K. Jain, S. Sharma, A. Vaidya, V. Ravichandran, R.K. Agrawal, 1,3,4-Thiadiazole
phases at constant pressure: and Its derivatives: a review on recent progress in biological activities, Chem.
Biol. Drug Design 81 (2013) 557–576.


[5] N. Kushwaha, K.S. Swatantra, A.K. Kushwaha, Biological activities of
Dcrg Hom ð298:15 KÞ ¼ Dcrg Hom ðT Þ  C pcr  C pg ð298:15  T Þ thiadiazole derivatives: a review, Int. J. Chemtech. Res. 4 (2012) 517–531.
[6] A.A. Usanova, M.V. Zorkin, Influence of derivatives of acetylgluyamine and
¼ Dcrg Hom ðT Þ  Dcrg C p ð298:15  T Þ ð17Þ aminohexane acids on ulcerogenic processes and peroxidation of lipids in
experiment, Sci. Bull. Med. Pharm. 201 (2015) 85–91.
1  
where Dcrg C p ð298:15Þ=J  K1 mol ¼ 0:75 þ 0:15C pcr was derived [7] D.G. Archer, P. Steffen, J. Rudtsch, Enthalpy of fusion of indium: a certified
reference material for differential scanning calorimetry, J. Chem. Eng. Data 48
by Chickos et al. [31] on the basis of statistical results and C pcr (2003) 1157–1163.
was estimated using a group additivity approach (Table 7). The [8] H.L. Ward, S.S. Cooper, The system, benzoic acid, ortho phthalic acid, water, J.
uncertainty associated with the value of C pcr was comparable to that Phys. Chem. 24 (1930) 1484–1493.
[9] L.E. Strong, R.M. Neff, I. Whitesel, Thermodynamics of dissolving and solvation
assigned by Chickos [32]. processes for benzoic acid and the toluic acids in aqueous solution, J. Sol.
As the represented data show, AETH has a high value of subli- Chem. 18 (1989) 101–114.
mation enthalpy Dcrg Hom ð298:15KÞ = 131.5 ± 1.2 kJmol1. We can [10] A. Apelblat, E. Manzurola, N.A. Babal, The solubilities of benzene
polycarboxylic acids in water, J. Chem. Thermodyn. 38 (2006) 565–571.
assume that the presence of 2 donor (secondary amine [11] W. Zielenkiewicz, G.L. Perlovich, M. Wszelaka-Rylik, The vapour pressure and
groups > NH) and 4 acceptor (2 nitrogen atoms in the thiadiazole the enthalpy of sublimation: determination by inert gas flow method, J.
ring and 2 carbonyl COH groups) centers in the molecule structure Therm. Anal. Calorim. 57 (1999) 225–234.
[12] S. Blokhina, A. Sharapova, M. Ol’khovich, G. Perlovich, Sublimation
will lead to the formation of a diverse system of hydrogen bonding thermodynamics of four fluoroquinolone antimicrobial compounds, J. Chem.
in the crystal lattice of the compound and, as a result, to a rather Thermodyn. 105 (2017) 37–43.
strong molecular packing. That is why we can quite reasonably [13] J.D. Cox, G. Pilcher, Thermochemistry of Organic and Organometallic
Compounds, Academic Press, London, 1970.
conclude that the thermodynamic cause of the low solubility of [14] F. Yuan, Y. Wang, L. Xiao, Q. Huang, H. Hao, Solubility of cefoxitin acid in
AETH is the high standard sublimation enthalpy, i.e. the energy different solvent systems, J. Chem. Thermodyn. 103 (2016) 125–133.
contribution of the solvation process cannot compensate for the [15] S. Zong, J. Wang, Y. Xiao, H. Wu, H. Hao, Solubility and dissolution
thermodynamic properties of lansoprazole in pure solvents, J. Mol. Liq. 241
energy required to break the crystal lattice sufficiently. (2017) 399–406.
[16] K.D. Bhesaniya, K. Nandha, S. Baluja, Measurement, correlation and dissolution
thermodynamics of biological active chalcone in organic solvents at different
4. Conclusions temperatures, J. Chem. Thermodyn. 74 (2014) 32–38.
[17] P. Atkins, J. De Paula, Physical Chemistry, eighth ed., W.H. Freeman and
The data on the original bioactive 6-(acetylamino)-N-(5-ethyl- Company, New York, 2006.
[18] B.C. Hancock, P. York, R.C. Rowe, The use of solubility parameters in
1,3,4-thiadiazol-2-yl) hexanamide solubility in five pharmaceuti- pharmaceutical dosage form design, Int. J. Pharm. 148 (1) (1997) 1–21.
cally relevant solvents has been determined as a function of tem- [19] M.A. Mohammad, A. Alhalaweh, S.P. Velaga, Hansen solubility parameter as a
perature by the shake-flask method. The mole fraction tool to predict cocrystal formation, Int. J. Pharm. 407 (2011) 63–71.
[20] J. Breitkreutz, Prediction of intestinal drug absorption properties by three
solubilities of AETH at 298.15 decreased in the order: 1-octanol dimensional solubility parameters, Pharm. Res. 15 (1998) 1370–1375.
(9.61104) > ethanol (9.55104) > buffer pH 2.0 (3.43105) > buf- [21] R.F. Fedors, A method for estimating both the solubility parameters and molar
fer pH 7.4 (2.90105) > hexane (2.14105). The experimental sol- volumes of liquids, Polym. Eng. Sci. 14 (1974) 147–154.
[22] D.W. van Krevelen, K. Nijenhuis, Properties of Polymers: Their Correlation with
ubilities of AETH in mole fraction were correlated well with the Chemical Structure; Their Numerical Estimation and Prediction from Additive
van’t Hoff and modified Apelblat models. To assess the solubilizing Group Contributions, Elsevier Sciences Publishers, Amsterdam, 1990.
properties of the selected solvents a Hansen solubility parameter [23] E.B. Bagley, T.P. Nelson, J.M. Scigliano, Three-dimensional solubility
parameters and their relationship to internal pressure measurements in
based on van Krevelen–Hoftyzer contributions was used. The
polar and hydrogen bonding solvents, J. Paint Technol. 43 (1971) 35–42.
results of predicting the behavior of solvents in this approach are [24] J. Albers, Hot-Melt Extrusion with Poorly Soluble Drugs, Cuvillier Verlag,
in a reasonable agreement with the experimental AETH solubility Goettingen, Germany, 2008.
values. The ideal solubility has been obtained from melting param- [25] D. Mealeya, M. Svärda, A.C. Rasmuson, Thermodynamics of risperidone and
solubility in pure organic solvents, Fluid Phase Equil. 375 (2014) 73–79.
eters, measured by the DSC technique. The solution activity coeffi- [26] L.C. Garzon, F. Martinez, Temperature dependence of solubility for ibuprofen
cients at equilibrium in the selected solvents have been quantified. in some organic and aqueous solvents, J. Sol. Chem. 33 (2004) 1379–1395.
44 A. Sharapova et al. / J. Chem. Thermodynamics 135 (2019) 35–44

[27] P. Zhu, Y. Chen, J. Fang, Z. Wang, F. Xu, Solubility and solution thermodynamics [31] J.S. Chickos, D.G. Hesse, J.F. Liebman, A group additivity approach for the
of thymol in six pure organic solvents, J. Chem. Thermodyn. 92 (2016) 198– estimation of heat capacities of organic liquids and solids at 298 K, Struct.
206. Chem. 4 (1993) 261–269.
[28] D.J.W. Grant, T. Higuchi, Solubility Behavior of Organic Compounds [32] J.S. Chickos, W.E. Acree, Enthalpies of Sublimation of Organic and
(Techniques of Chemistry), Wiley-Interscience, New York, 1990. Organometallic Compounds, J. Phys. Chem. Ref. Data 2 (2002) 537–698.
[29] J. Westergren, In silico prediction of drug solubility: 1. Free energy of
hydration, J. Phys. Chem. B 111 (2007) 1872–1882.
[30] S.P. Verevkin, V.N. Emel’yanenko, Transpiration method: vapor pressures and
enthalpies of vaporization of some low-boiling esters, Fluid Phase Equil. 266
(2008) 64–75. JCT 2018-549

Das könnte Ihnen auch gefallen