Solubility Modeling and Solvent Effects of Allopurinol in 15 Neat

Aug 28, 2018 - Experimental allopurinol solubility in water and 14 organic solvents, namely, N,N-dimethylformamide (DMF), isopropanol, ethanol, methan...
0 downloads 0 Views 941KB Size
Article Cite This: J. Chem. Eng. Data XXXX, XXX, XXX−XXX

pubs.acs.org/jced

Solubility Modeling and Solvent Effects of Allopurinol in 15 Neat Solvents Min Zheng,† Jiao Chen,† Gaoquan Chen,† Renjie Xu,‡ and Hongkun Zhao*,†,‡ †

College of Chemistry & Chemical Engineering, YangZhou University, YangZhou, Jiangsu 225002, People’s Republic of China Guangling College, YangZhou University, YangZhou, Jiangsu 225009, People’s Republic of China

J. Chem. Eng. Data Downloaded from pubs.acs.org by UNIV OF SOUTH DAKOTA on 08/30/18. For personal use only.



ABSTRACT: Experimental allopurinol solubility in water and 14 organic solvents, namely, N,N-dimethylformamide (DMF), isopropanol, ethanol, methanol, acetone, acetonitrile, n-propanol, n-heptanol, 1-hexanol, nbutanol, 2-butanone, N-methyl-2-pyrrolidone (NMP), 1,4-dioxane, and cyclohexane, has been studied by using the shake-flask method. The experiments were carried out over the temperature range from 278.15 to 333.15 K under p = 101.2 kPa. The solubility of allopurinol in selected solvents rose with the rise in temperature. The solubility values of allopurinol were recorded as the maximum in NMP and the minimum in cyclohexane. The obtained solubility data were mathematically described with the λh equation, Apelblat equation, and NRTL equation. The maximum percentage of the average relative deviation was 1.73 × 10−2, and the maximum value of the root-mean-square deviation was 2.416 × 10−4. Furthermore, the solvent effect was analyzed through the method of SERL (linear solvation energy relationship). The extent and types of solvent−solute and solvent−solvent interactions were recognized.



INTRODUCTION Solubility is of critical importance in the pharmaceutical industry for the design of products due to its impact on the efficacy of a drug and its forthcoming improvement in the formulation and pharmacokinetics including transport, release, absorption, and so forth.1−3 In the initial developing stages, solubility evaluation is important in the drug development process. The low solubility in water of drugs commonly results in insufficient absorption and low bioavailability. On the other hand, crystallization operation is very significant in the pharmaceutical field as a purification method and is usually used as the final production step of APIs (active pharmaceutical ingredients).4,5 The solid solubility offers the important thermodynamic knowledge needed for the crystallization process design and for controlling the distribution of crystal size and polymorphism in solutions.1−7 Allopurinol (CAS reg. no. 315-30-0, structure shown in Figure 1) is chemically described as 4-hydroxy-pyrazolo[3,4d]pyrimidine. It is generally used as a medication to reduce the levels of uric acid. Allopurinol is especially employed in preventing gout and particular types of kidney stones.8−10 The

key problem relating to the design of the dosage form of allopurinol is its very low water solubility,6,11,12 which decreases its bioavailability. Because it is highly permeable and very soluble in water,11 allopurinol is a class II drug. To improve its solubility in water, bioavailability, and therapeutic activity, various approaches, namely, solid dispersion,13−16 crystallization,17−20 varying pH values,12 surfactants,21 cyclodextrin inclusion complexation,22,23 cosolubilization,24,25 and hydrotropes,26−29 have been investigated. These techniques require precise solubility data in several neat solvents and solvent mixtures. Therefore, it is of crucial importance to know the solubility of allopurinol. However, the physicochemical properties of allopurinol in pure solvents and solvent mixtures have not yet been systematically investigated. Only the allopurinol solubility in water,11,12 different pH solutions,12 and in aqueous solutions of some compounds25−29 is available in the literature. In addition, the allopurinol solubilities in water, carbon tetrachloride, acetone, n-hexane, ethanol, and ethyl acetate at three temperatures (298, 310, and 313 K) have been reported elsewhere.6 According to the above considerations, the purpose of the work is to expand the allopurinol solubility data in different solvents, namely, water, N,N-dimethylformamide (DMF), isopropanol, methanol, acetone, ethanol, acetonitrile, nheptanol, n-butanol, n-propanol, 1-hexanol, 2-butanone, Nmethyl-2-pyrrolidone (NMP), cyclohexane, and 1,4-dioxane. Received: May 24, 2018 Accepted: August 17, 2018

Figure 1. Chemical structure of allopurinol. © XXXX American Chemical Society

A

DOI: 10.1021/acs.jced.8b00430 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Detailed Information on the Materials Used in This Work chemicals

molar mass, g·mol−1 melting point, K

allopurinol methanol ethanol isopropanol n-propanol n-butanol acetone acetonitrile DMF 2-butanone 1,4-dioxane NMP n-heptanol 1-hexanol cyclohexane water

136.11 32.04 46.07 60.10 60.10 74.12 58.08 41.05 73.09 72.10 88.11 99.13 116.2 102.18 84.16 18.02

melting enthalpy, kJ·mol−1

source

mass fraction purity

analytical method

38.5a

Sigma Chemical Co., Ltd. Merck

0.992 0.997 0.994 0.995 0.995 0.994 0.995 0.996 0.995 0.994 0.995 0.994 0.994 0.996 0.995 conductivity DMF > 2-butanone > acetone > n-heptanol > acetonitrile >1-hexanol > n-butanol > n-propanol > isopropanol > ethanol > methanol > 1,4-dioxane > water > cyclohexane. For the allopurinol and alcohol solutions, the sequence of allopurinol solubility is approximately consistent with the solvent polarity. The methanol polarity is the largest among the selected alcohols,43 therefore the allopurinol solubility in methanol is lowest in these alcohols. The same tendency may be observed for 2-butanone and acetone. Although the cyclohexane polarity is the lowest for these solvents, the solubility of allopurinol in cyclohexane is lower than in the other solvents. In addition, the dipole moments of allopurinol are very large. An allopurinol molecule may give large dipole−dipole interactions with solvents owing to the >HH group.44 It can form hydrogen bonds with solvents. The formed hydrogen bonds with the molecules of solvents have a significant effect on solute solubility. The solubility of allopurinol in NMP and DMF is larger than that in the other neat solvents. Obviously the hydrogen bonds are formed between the >HH group of

Co. Ltd., China. The solution was shaken through the shaker kept at 100 rpm. So as to obtain an equilibration time, 0.5 mL of liquor was withdrawn with a 2 mL syringe at intervals of 1 h and then analyzed using high-performance liquid chromatography (HPLC, Agilent 1260). Results indicated that 23 h was enough for all solutions to be in equilibrium. Then each mixture was removed from the shaker, and the allopurinol particles were allowed to settle for 4 h. The upper liquid was withdrawn carefully, diluted, and analyzed by HPLC. Analytical Analysis of Allopurinol Using HPLC. The content of allopurinol in the equilibrium liquor was tested by using HPLC (Agilent 1260). The HPLC system included a quaternary pump and a UV detector (G1314F). The separation was conducted by using a Waters C18 reversephase column (250 mm × 4.6 mm, 5 μm) (Waters Inc., Bedford, MA, USA). The flow rate of mobile phase was 1 mL· min−1, which comprised a solution of 50 mmol·L−1 KH2PO4 (pH 4.63) and methanol (97:3, v/v). The temperature of the column was about 303 K, and the detection wavelength of allopurinol was 254 nm.42 Prior to the test, the HPLC system was calibrated by using standard solutions. The calibration curve was built between the HPLC area and the allopurinol concentration. Each determination was carried in triplicate, and three samples were taken for every mixture. The obtained solubility data was the mean value of three measurements. The relative standard uncertainty was estimated to be 0.0224 for the obtained solubility data in mole fraction. Solid-State Characterization of Allopurinol. So as to evaluate the possible transformation and crystallinity of allopurinol during the experiment, we characterized the excess solid allopurinol in solution. The excess solid was gathered and analyzed by using XPRD (X-ray powder diffraction). The XPRD spectra of equilibrated and pure allopurinol were obtained by using a HaoYuan DX-2700B instrument over the 2θ range of 10−90°. The anode was Cu with Kα radiation (λ = 1.54184 nm). The spectra of equilibrated and pure allopurinol were attained at a tube current and voltage of 40 mA and 40 kV, respectively, in a scan step size of 0.02°.



RESULTS AND DISCUSSION XPRD Analysis. The XPRD scan of equilibrated and pure allopurinol is given in Figure 2. The XPRD spectra of neat and equilibrated allopurinol present sharp characteristic peaks at different values of 2θ (Figure 2). The XPRD spectra of neat allopurinol is very similar to that of the crystalline allopurinol D

DOI: 10.1021/acs.jced.8b00430 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 4. Solubility of allopurinol in water determined in this work and reported in previous publications. (a) ■, this work; •, ref. 6; ▲, ref. 11; ▼, ref. 12. (b) ◆, Ethanol, this work; ◇, ethanol, ref. 6; ★, acetone, this work; ☆, acetone, ref. 6.

allopurinol and the oxygen atoms of NMP and DMF. Consequently, the x values of allopurinol are much higher in the two solvents than in the other pure solvents studied. Generally, it is very difficult to illustrate the allopurinol solubility presented in Table 2 for one reason. This case is perhaps due to numerous factors including molecule polarity, solvent−solute interactions, and molecular shapes, sizes, and interactions between solvent molecules. It should be noted that the solubility (expressed as g·L−1) of allopurinol in ethanol and acetone at T = 298, 310, and 313 K and in water at T = 298, 303, 310, 313, and 315 K has been published by Mota and co-workers.6 In addition, the solubility (expressed as g·L−1) of allopurinol in H2O at T = 298 K is also available in the literature.11,12 It is regretful that the densities of the corresponding solutions are not reported in the published papers. So as to compare the reported solubility values with those determined in this work, they are converted to the mole fraction solubility by assuming that the solution density is equal to that of the corresponding pure solvent due to lower solubility values. The comparison of the solubility values is made in Figure 4. For the allopurinol and water system (Figure 4a), the allopurinol solubility in mole fraction at 298.15 K obtained in the present work is 0.2398 × 10−4, which is much lower than 0.7533 × 10−4 reported by Dave11 and 0.5296 × 10−4 reported by Hendriksen.12 However, they are very close to those reported by Mota6 at different temperatures, which are 0.2648 × 10−4 at 298 K, 0.3707 × 10−4 at 303 K, 0.5560 × 10−4 at 310 K, 0.7282 × 10−4 at 313 K, and 0.7679 × 10−4 at 315 K. Nevertheless, for ethanol and acetone solvents (Figure 4b), a few large differences are observed between the reported solubility data6 and the ones determined in this work. The difference perhaps results from many factors, e.g., the experimental technique, solute purity, equilibrium time, analysis method, sampling, and so on. Solvent Effect upon Solubility. In order to obtain deep information on the solvent effect on solubility, the KAT-LSER model is employed to examine the allopurinol solubility in the selected neat solvents at a temperature of 298.15 K. The required parameters including α, β, π*, and δH are obtained from the literature45−47 and tabulated in Table 3. The molar volume of allopurinol is evaluated with the method suggested by Fedors,48 which is Vs = 71.2 cm3·mol−1 (Table 4). The acquired solubility values are correlated with some parameters of the solvents by using MLRA. The regressed results are described as eq 11 for the selected monosolvents at a temperature of 298.15 K.

Table 3. Hildebrand Solubility Parameters (δH) and Solvatochromic Parameters α, β, and π* for Neat Solventsa

a

solvent

α

β

π*

Vsδ2H/(100RT)

methanol ethanol isopropanol n-propanol n-butanol acetone acetonitrile DMF water NMP n-heptanol 1-hexanol cyclohexane 1,4-dioxane 2-butanone

0.98 0.86 0.76 0.84 0.84 0.08 0.19 0.00 1.17 0.03 0.64 0.67 0 0 0.06

0.66 0.75 0.84 0.90 0.84 0.43 0.40 0.77 0.47 0.75 0.96 0.94 0 0.37 0.48

0.60 0.54 0.48 0.52 0.47 0.71 0.75 0.89 1.09 0.97 0.39 0.40 0 0.55 0.67

0.451 0.361 0.286 0.311 0.277 0.204 0.306 0.178 1.175 0.151 0.127 0.130 0.145 0.216 0.187

Taken from refs. 45−47.

Table 4. Application of the Fedors’ Method to Estimate Molar Volume of Allopurinol group

group number

V (cm3·mol−1)

>CO −N= >NH >C= −CH= ring closure with five or more atoms conjugation in ring for each double bond total

1 2 2 2 2 2 3

10.8 5.0 4.5 −5.5 13.5 16 −2.2 71.2

ln(x) = −12.846(0.599) + 0.431(1.592)α + 1.322(1.856)β ij V δ 2 yz + 9.291(1.630)π * − 7.758(2.710)jjjj s H zzzz k 100RT {

(11)

n = 15, R2 = 0.932, RSS = 3.774, and F = 49.02. Here, R2 denotes the correlation coefficient, RSS is the residual sum of squares, and F is the F test. The standard deviation for every coefficient is given in square brackets. Equation 11 shows that the KAT-LSER model comprising all variables provides a satisfactory result for the determined solubility over all of the neat solvents. The equation coefficients’ values show that the contributions to allopurinol solubility are 2.29, 7.08, 49.42, and 41.28% for α, β, π*, and δH, respectively. Therefore, the specific interactions involving hydrogen bonding play a small E

DOI: 10.1021/acs.jced.8b00430 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 5. Parameters in the Modified Apelblat Equation and λh Equation for Allopurinol in Different Solvents λh equation

Apelblat equation solvent

A

B

C

100 RD

106 RMSD

λ

h

100 RD

106 RMSD

methanol ethanol isopropanol n-propanol n-butanol acetone acetonitrile DMF water NMP n-heptanol 1-hexanol 1,4-dioxane cyclohexane 2-butanone

12.57 36.63 31.2 1.58 32.89 1.50 18.12 −49.23 55.88 −3.81 −18.88 32.49 41.12 77.71 −0.02

−4782.5 −5720.4 −5077.1 −3526.0 −4794.1 −3028.4 −3809.9 −368.59 −7294.69 −1785.5 −1776.5 −4702.4 −5443.4 −7816.0 −2908.5

−1.10 −4.74 −4.12 0.21 −4.49 0.37 −2.43 7.90 −7.35 1.07 2.90 −4.45 −5.73 −11.46 0.59

0.67 0.70 0.75 1.19 0.47 0.02 0.44 1.14 1.18 0.68 0.31 0.99 0.74 2.14 0.02

0.92 1.04 1.38 2.65 1.04 0.33 2.26 118 0.42 235 1.47 2.99 1.15 0.04 0.39

0.1913 0.1631 0.0931 0.0764 0.0744 0.4187 0.0523 0.7089 0.2694 1.2190 0.0289 0.0749 0.0450 0.002233 0.4518

23 150.6 26 097.4 40 822.0 46 700.2 45 782.0 7480.69 58 297.0 3940.64 18 624.0 1745.89 91 230.3 44 307.4 81 336.3 1 892 103 6818.96

0.57 0.77 0.62 1.12 0.53 0.01 0.56 0.70 1.00 0.70 0.30 0.80 0.81 1.72 0.04

0.93 1.36 1.63 2.71 1.91 0.30 2.62 132.7 0.73 241.6 1.74 3.44 1.34 0.05 1.52

Table 6. Parameters of the NRTL Model in Different Solvents solvent

a12

b12/K

a21

b21/K

α

100 RAD

106 RMSD

methanol ethanol isopropanol n-propanol n-butanol acetone acetonitrile DMF water 1-hexanol n-heptanol 2-butanone NMP 1,4-dioxane cyclohexane

−2.2256 4.3776 4.1837 −2.691 4.5911 −1.7243 4.5186 −2.7657 4.0548 36.081 −2.584 −1.987 28.193 3.708 32.556

−5.4266 0.0572 −0.4912 −73.4555 0.3233 −118.81 0.2206 −11.947 −0.7477 −8825.7 5.8573 −76.0472 −17118.5 0.057 −7923.147

6.3702 1.1909 2.4314 10.4751 3.276 5.8541 4.1399 9.4419 −0.4855 7.3724 12.2914 6.6197 −38.368 3.326 9.157

−203.27 −369.12 −824.24 −815.97 −1214.3 −1107.4 −1561.9 −1783.1 387.26 −2415.1 −2041.6 −1273.8 15 556.1 −956.72 −1424.7

0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3

0.59 0.77 0.61 1.11 0.52 0.01 0.52 0.68 1.00 0.55 0.25 0.03 0.69 0.81 1.73

0.97 1.39 1.62 2.92 1.89 0.20 2.56 114.7 0.74 2.73 1.44 1.15 222.0 1.35 0.04

deviation), and RMSD (root-mean-square deviation) are employed

role in the solubility of allopurinol. For the specific interactions, β of the solvent makes a larger contribution to solubility than does α. The influence of the interactions of dipolarity/polarizability on allopurinol solubility almost has the same degree as that of the cavity term, which is accounted for by Vsδ2H/(100RT). The α, β, and π* coefficients are all positive, which shows that the solubility of allopurinol rises with the increase in solvent dipolarity/polarizability and the formation ability of a hydrogen bond. On the contrary, the solubility of allopurinol decreases with the increase in the cohesive energy density of the studied pure solvents. Data Correlation and Calculation. The allopurinol solubility is correlated in terms of the following objective function for the NRTL model F=

∑ (ln γie − ln γic)2 i=1

RAD =

∑ (xie − xic)2 i=1

N



xie − xic xie

ÄÅ N É ÅÅ ∑ (x c − x e)2 ÑÑÑ1/2 ÅÅ i = 1 i ÑÑ i ÑÑ RMSD = ÅÅÅ ÑÑ ÅÅ N ÑÑÖ ÅÇ i=1

(14)

(15)

Here, γei is the activity coefficient obtained with eq 3, and γci is the activity coefficient evaluated with the corresponding models; xei and xci represent, respectively, the determined and evaluated allopurinol solubility, and N refers to the number of experimental points. The melting properties of allopurinol are cited in the literature,6 which are 653.5 K for the melting temperature (Tm) and 38.5 kJ·mol−1 for the melting enthalpy (ΔfusH). The curvefit parameters λ and h in the Buchowski−Ksiazaczak λh equation; A, B, and C in the Apelblat equation; aij and bij in the NRTL equation; and along with the values of RMSD and RAD are tabulated in Tables 5 and 6. Additionally, the backcalculated solubility values with the Apelblat model are plotted in Figure 3.

(12)

whereas the objective function for the λh and modified Apelblat equations is represented as eq 13 F=

1 N

(13)

Furthermore, in order to estimate the three solubility models, the RD (relative deviation), RAD (relative average F

DOI: 10.1021/acs.jced.8b00430 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Metastable Polymorph of Carbamazepine. Org. Process Res. Dev. 2013, 17, 512−518. (8) Metzner, J. E.; Buchberger, D.; Lauter, J.; Pech, R. Investigation into the Bioequivalence of a New Allopurinol Tablet Formulation Compared with a Standard Preparation. Arzneim.-Forsch. 1997, 47, 1236−1241. (9) Changdeo, J. S.; Vinod, M.; Shankar, K. B.; Rajaram, C. A. Physicochemical Characterization and Solubility Enhancement Studies of Allopurinol Solid Dispersions. Braz. J. Pharm. Sci. 2011, 47, 513−523. (10) Pacher, P.; Nivorozhkin, A.; Szabó, C. Therapeutic Effects of Xanthine Oxidase Inhibitors: Renaissance Half a Century after the Discovery of Allopurinol. Pharmacol. Rev. 2006, 58, 87−114. (11) Dave, R. A.; Morris, M. E. Novel high/low solubility classification methods for new molecular entities. Int. J. Pharm. 2016, 511, 111−126. (12) Hendriksen, B. A.; Sanchez Felix, M. V.; Bolger, M. B. The Composite Solubility versus pH Profile and its Role in Intestinal Absorption Prediction. AAPS PharmSci 2003, 5, 35−49. (13) Fayjunessa, R.; Sarkar, M. R.; Sultana, R.; Jahan, K.; Labu, Z. K. Study on Dissolution Improvement of Allopurinol by Co-grinding and Fusion Method using Solid Dispersion Technique. J. Biomed. Pharmaceut. Res. 2013, 2, 1−7. (14) Narasaki, M.; Okamura, N.; Fujii, T. Solubilization Enhancement of Water-Insoluble Drugs by Coating with Dispersants and Solubilization Aids. JP Patent 11,246,393, Sept 14, 1999. (15) Hamza, Y. E.; Kata, M. Improvement of Dissolution Characteristics and Solubility of Allopurinol via Solid Dispersion and Hydrotropy with Gentisic Acid Ethanolamide. Pharm. Ind. 1990, 52, 363−368. (16) Samy, A. M.; Marzouk, M. A.; Ammar, A. A.; Ahmed, M. K. Enhancement of the Dissolution Profile of Allopurinol by a Solid Dispersion Technique. Drug Discov. Ther. 2010, 4, 77−84. (17) Hanna, M.; Shan, N.; Cheney, M. L.; Weyna, R. D.; Houck, R. Crystallization Method for Improvement of Oral Bioavailability of Drugs Including Bisphosphonates. U.S. Patent 20,130,035,315, Feb 7, 2013. (18) Hanna, M.; Shan, N.; Cheney, M. L.; Weyna, D. R.; Houck, R. Improvement of Solubility and Permeability of Drugs through Generating Crystalline Forms. WO Patent 2,011,097,269, Aug 11, 2011. (19) Docherty, R.; Pencheva, K.; Abramov, Y. A. Low Solubility in Drug Development: Deconvoluting the Relative Importance of Solvation and Crystal Packing. J. Pharm. Pharmacol. 2015, 67, 847− 856. (20) Mortada, S. A. M.; Boraie, N. A. Enhancement of Dissolution Rate of Allopurinol by Crystallization in Aqueous Surfactant and Hydrophilic Polymer Solutions. Alexandria J. Pharm. Sci. 1989, 3, 45− 50. (21) Hamza, Y. E.; Kata, M. Influence of Certain Nonionic Surfactants on the Solubilization and inVitro Availability of Allopurinol. Pharm. Ind. 1989, 51, 1441−1444. (22) Hamza, Y. E.; Kata, M. Influence of Certain Hydrotropic and Complexing Agents on the Solubilization of Allopurinol. Pharm. Ind. 1989, 51, 1159−1162. (23) Ammar, H. O.; El-Nahhas, S. A. Improvement of Some Pharmaceutical Properties of Drugs by Cyclodextrin Complexation. 1. Allopurinol. Pharmazie 1995, 50, 49−51. (24) Hamza, Y. E.; Kata, M. Cosolubilization and in vitro Availability of Solubilized Allopurinol. Pharm. Ind. 1990, 52, 241−244. (25) Collett, J. H.; Kesteven, G. The solubility of Allopurinol in Aqueous Solutions of Polyvinylpyrrolidone. Drug Dev. Ind. Pharm. 1978, 4, 555−568. (26) Ammar, H. O.; El-Nahhas, S. A.; Khalil, R. M.; Omar, S. M. Solubilization of Allopurinol with Methylxanthines. Pharmazie 1994, 49, 839−842. (27) Ammar, H. O.; El-Nahhas, S. A. Effect of Aromatic Hydrotropes on the Solubility of Allopurinol. Part 1: Effect of

Tables 5 and 6 show that the evaluated solubility of allopurinol in 15 solvents coincides well with the determined ones. The largest value of the RMSD is 2.416 × 10−4, which is attained using the λh model for NMP. All of the RAD values are no greater than 1.73%. Therefore, the computed solubility values by using the three solubility models show good agreement with the determined values. That is to say, the selected three solubility models are suitable for describing the solubility of allopurinol in all 15 solvents in the temperature range from 278.15 to 333.15 K.



CONCLUSIONS The allopurinol solubility was determined in 15 neat solvents at temperatures ranging from 278.15 to 333.15 K. At a certain temperature, the solubility of allopurinol in the 15 solvents ranked as NMP > DMF > 2-butanone > acetone > n-heptanol > acetonitrile > 1-hexanol > n-butanol > n-propanol > isopropanol alcohol > ethanol > methanol > 1,4-dioxane > water > cyclohexane. The Apelblat equation, NRTL model, and Buchowski−Ksiazaczak λh equation were used in the solubility correlation. The largest RMSD value obtained was 2.416 × 10−4 and that of RAD was 1.73 × 10−2. The concept of LSER was employed in describing the change in solubility according to the solvent effect. The solubility parameter and dipolarity/polarizability made a large contribution to solubility.



AUTHOR INFORMATION

Corresponding Author

*Tel: +86 514 87975244. E-mail: [email protected]. ORCID

Renjie Xu: 0000-0003-1888-8864 Hongkun Zhao: 0000-0001-5972-8352 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are thankful for the financial support of the Innovation and Entrepreneurship Training Project for Undergraduate of Jiangsu Higher Education Institutions (grant no. 201813987006Y) and the Natural Science Foundation of Guangling College, Yangzhou University (grant no. ZKYB17007).



REFERENCES

(1) Avdeef, A. Absorption and Drug Development, Solubility. Permeability and Charge State; Wiley-Interscience: Hoboken, NJ, 2003. (2) Aulton, M. E. Pharmaceutics: The Science of Dosage Forms Design, 2nd ed.; Churchill Livingstone: London, 2002. (3) Jouyban, A. Handbook of Solubility Data for Pharmaceuticals; CRC Press: Boca Raton, FL, 2010. (4) Chen, J.; Sarma, B.; Evans, J. M. B.; Myerson, A. S. Pharmaceutical Crystallization. Cryst. Growth Des. 2011, 11, 887−895. (5) Pinho, S. P.; Macedo, E. A. Solubility in Food, Pharmaceutical and Cosmetic Industries. In Developments and Applications in Solubility; Letcher, T. M., Ed.; Royal Society of Chemistry: Cambridge, 2007; pp 309−326. (6) Mota, F. L.; Carneiro, A. P.; Queimada, A. J.; Pinho, S. P.; Macedoa, E. A. Temperature and Solvent Effects in the Solubility of Some Pharmaceutical Compounds: Measurements and Modeling. Eur. J. Pharm. Sci. 2009, 37, 499−507. (7) O’Mahony, M. A.; Croker, D. M.; Rasmuson, Å. C.; Veesler, S.; Hodnett, B. K. Measuring the Solubility of a Quickly Transforming G

DOI: 10.1021/acs.jced.8b00430 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Sodium Salts of Hydroxy and Amino Derivatives of Benzoic Acid. Pharmazie 1993, 48, 842−845. (28) Ammar, H. O.; El-Nahhas, S. A. Effect of Aromatic Hydrotropes on the Solubility of Allopurinol. Part 2: Effect of Nicotinamide and Sodium Salts of Benzoic, Naphthoic and Nicotinic Acids. Pharmazie 1993, 48, 534−536. (29) Ammar, H. O.; El-Nahhas, S. A. Effect of Aromatic Hydrotropes on the Solubility of Allopurinol. Part 3: Sodium Salts of Toluic Acids. Pharmazie 1993, 48, 751−754. (30) Manzurola, E.; Apelblat, A. Solubilities of L-Glutamic acid, 3Nitrobenzoic Acid, P-Toluic Acid, Calcium-L-Lactate, Calcium Gluconate, Magnesium-DL-Aspartate, and Magnesium-L-Lactate in Water. J. Chem. Thermodyn. 2002, 34, 1127−1136. (31) Apelblat, A.; Manzurola, E. Solubilities of o-Acetylsalicylic, 4Aminosalicylic, 3,5-Dinitrosalicylic, and P-Toluic Acid, and Magnesium-DL-Aspartate in Water from T = (278 to 348) K. J. Chem. Thermodyn. 1999, 31, 85−91. (32) Buchowski, H.; Ksiazczak, A.; Pietrzyk, S. Solvent Activity along a Saturation Line and Solubility of Hydrogen-Bonding Solids. J. Phys. Chem. 1980, 84, 975−979. (33) Renon, H.; Prausnitz, J. M. Estimation of Parameters for the NRTL Equation for Excess Gibbs Energies of Strongly Nonideal Liquid Mixtures. Ind. Eng. Chem. Process Des. Dev. 1969, 8, 413−419. (34) Li, W. X.; Chen, J. T.; Han, S.; Du, C. B.; Zhao, H. K. Thermodynamic Solubility of Tetraethyl Ranelate in Ten Organic Solvents at Different Temperatures. J. Mol. Liq. 2016, 216, 771−780. (35) Kamlet, M. J.; Abboud, J. L. M.; Taft, R. W. An Examination of Linear Solvation Energy Relationships. Prog. Phys. Org. Chem. 1980, 13, 485−630. (36) Marcus, Y. Solubility and Solvation in Mixed Solvent Systems. Pure Appl. Chem. 1990, 62, 2069−2076. (37) Carr, P. W. Solvatochromism, Linear Solvation Energy Relationships, and Chromatography. Microchem. J. 1993, 48, 4−28. (38) Jouyban, A.; Fakhree, M. A. A. Experimental, Computational Methods Pertaining to Drug Solubility. In Toxicity and Drug Testing; Acree, W.E., Jr., Ed.; InTech: Rijeka, Croatia, 2012; pp 187−218. (39) Jouyban, A.; Martinez, F.; Panahi-Azar, V. Solubility of Fluphenazine Decanoate in Aqueous Mixtures of Polyethylene Glycols 400 and 600 at Various Temperatures. Fluid Phase Equilib. 2014, 368, 58−64. (40) Shakeel, F.; Haq, N.; Siddiqui, N. A.; Alanazi, F. K.; Alsarra, I. A. Correlation of Solubility of Bioactive Compound Reserpine in Eight Green Solvents at (298.15 to 338.15) K. J. Chem. Eng. Data 2015, 60, 775−780. (41) Shakeel, F.; Haq, N.; Siddiqui, N. A. Solubility and Thermodynamics Function of Vanillin in Ten Different Environmentally Benign Solvents. Food Chem. 2015, 180, 244−248. (42) Su, D.; Wu, W. C.; Xin, H. W.; Xiong, L. Determination the Concentration of Allopurinol and its Metabolite-Oxypurinol in Human Plasma by HPLC. China Pharmacist 2009, 12, 1033−1034. (43) Smallwood, I. M. Handbook of Organic Solvent Properties; Amoled: London, 1996. (44) Huyskens, F.; Morissen, H.; Huyskens, P. Solubilities of pNitroanilines in Various Classes of Solvents. Specific Solute-Solvent Interactions. J. Mol. Struct. 1998, 441, 17−25. (45) Marcus, Y. The Properties of Organic Liquids that are Relevant to Their Use as Solvating Solvents. Chem. Soc. Rev. 1993, 22, 409− 416. (46) Zakerhamidi, M. S.; Golghasemi Sorkhabi, Sh. Solvent Effects on the Molecular Resonance Structures and Photo-Physical Properties of a Group of Oxazine Dyes. J. Lumin. 2015, 157, 220−228. (47) Hansen, C. M. Hansen Solubility Parameters: A User’s Handbook; CRC Press: 2007. (48) Fedors, R. F. A method for Estimating both the Solubility Parameters and Molar Volumes of Liquids. Polym. Eng. Sci. 1974, 14, 147−154.

H

DOI: 10.1021/acs.jced.8b00430 J. Chem. Eng. Data XXXX, XXX, XXX−XXX