Next Article in Journal
Rheological Investigation of Highly Filled Copper(II) Oxide Nanosuspensions to Optimize Precursor Particle Content in Reductive Laser-Sintering
Previous Article in Journal
Frustrated-Laser-Induced Thermal Starting Plumes in Fresh and Salt Water
Previous Article in Special Issue
Conventional and Green Rubber Plasticizers Classified through Nile Red [E(NR)] and Reichardt’s Polarity Scale [ET(30)]
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Solvent Replacement Strategies for Processing Pharmaceuticals and Bio-Related Compounds—A Review

by
Jia Lin Lee
1,
Gun Hean Chong
1,*,
Masaki Ota
2,3,
Haixin Guo
4 and
Richard Lee Smith, Jr.
2,*
1
Faculty of Food Science and Technology, Universiti Putra Malaysia, Serdang 43400, Selangor, Malaysia
2
Graduate School of Environmental Studies, Tohoku University, Aramaki Aza Aoba, 468-1, Aoba-ku, Sendai 980-8572, Japan
3
Graduate School of Engineering, Tohoku University, Aramaki Aza Aoba, 6-6-11-403, Aoba-ku, Sendai 980-8579, Japan
4
Agro-Environmental Protection Institute, Ministry of Agriculture and Rural Affairs, No. 31 Fukang Road, Nankai District, Tianjin 300191, China
*
Authors to whom correspondence should be addressed.
Liquids 2024, 4(2), 352-381; https://doi.org/10.3390/liquids4020018
Submission received: 31 December 2023 / Revised: 22 February 2024 / Accepted: 21 March 2024 / Published: 9 April 2024

Abstract

:
An overview of solvent replacement strategies shows that there is great progress in green chemistry for replacing hazardous di-polar aprotic solvents, such as N,N-dimethylformamide (DMF), 1-methyl-2-pyrrolidinone (NMP), and 1,4-dioxane (DI), used in processing active industrial ingredients (APIs). In synthetic chemistry, alcohols, carbonates, ethers, eucalyptol, glycols, furans, ketones, cycloalkanones, lactones, pyrrolidinone or solvent mixtures, 2-methyl tetrahydrofuran in methanol, HCl in cyclopentyl methyl ether, or trifluoroacetic acid in propylene carbonate or surfactant water (no organic solvents) are suggested replacement solvents. For the replacement of dichloromethane (DCM) used in chromatography, ethyl acetate ethanol or 2-propanol in heptanes, with or without acetic acid or ammonium hydroxide additives, are suggested, along with methanol acetic acid in ethyl acetate or methyl tert-butyl ether, ethyl acetate in ethanol in cyclohexane, CO2-ethyl acetate, CO2-methanol, CO2-acetone, and CO2-isopropanol. Supercritical CO2 (scCO2) can be used to replace many organic solvents used in processing materials from natural sources. Vegetable, drupe, legume, and seed oils used as co-extractants (mixed with substrate before extraction) can be used to replace the typical organic co-solvents (ethanol, acetone) used in scCO2 extraction. Mixed solvents consisting of a hydrogen bond donor (HBD) solvent and a hydrogen bond acceptor (HBA) are not addressed in GSK or CHEM21 solvent replacement guides. Published data for 100 water-soluble and water-insoluble APIs in mono-solvents show polarity ranges appropriate for the processing of APIs with mixed solvents. When water is used, possible HBA candidate solvents are acetone, acetic acid, acetonitrile, ethanol, methanol, 2-methyl tetrahydrofuran, 2,2,5,5-tetramethyloxolane, dimethylisosorbide, Cyrene, Cygnet 0.0, or diformylxylose. When alcohol is used, possible HBA candidates are cyclopentanone, esters, lactone, eucalytol, MeSesamol, or diformylxylose. HBA—HBA mixed solvents, such as Cyrene—Cygnet 0.0, could provide interesting new combinations. Solubility parameters, Reichardt polarity, Kamlet—Taft parameters, and linear solvation energy relationships provide practical ways for identifying mixed solvents applicable to API systems.

1. Introduction

Solvents are commonly viewed as being polar or nonpolar, depending on whether their molecular structure contains highly electronegative (N, O, S, Cl, Br, I) elements or only (C, H) elements. However, for a molecule to be polar, it must contain a polar bond and have asymmetry in its structure that causes an imbalance in charge separation between two (+ and −) poles referred to as dipoles. The presence of an asymmetrically arranged polar bond, such as C-Cl in chloromethane (CH3C1), causes the molecule to be polar, whereas the presence of four symmetrically arranged C-Cl bonds in carbon tetrachloride (CCl4) cause the molecule to be nonpolar. For two solvents to be miscible, similarity in molecular polarity is required, as given by the well-known adage, “like dissolves like”, which in other words means that, for the solvation of polar molecules to occur, dipole—dipole interactions must exist, and conversely, for the solvation of nonpolar molecules to occur, dipole—dipole interactions must be absent. There are many exceptions to this adage, and certainly, system conditions (temperature, pressure) and van der Waals-London forces (dispersion) play important roles in solvation processes. Moreover, for solvent mixtures as discussed in this review, composition and interactions between hydrogen bond donor (HBD) and hydrogen bond acceptor (HBA) molecules are important.
Physical properties such as dipole moment (μD), dielectric constant (ε), octanol-water partition coefficient (logKow or logP), normal boiling point (Tb), melting temperature (Tm), entropy of fusion (∆fusS), Hildebrand solubility parameter, and Hansen solubility parameter help to characterize the macroscopic polarity of a molecule. On the other hand, empirical polarity scales based on solvatochromic probes (dyes), such as Reichardt ET(30) [1] and normalized ETN values [2], Kamlet—Taft (KT) acidity (α), basicity (β) and dipolar/polarizability (π*) values [3,4], and Catalán parameters [5], help to characterize the microscopic polarity of a solvent [6]. In solvent selection guides developed by the industry [7,8] and chemical societies [9,10,11,12], pure component solvent properties are analyzed in detail for developing solvent replacement strategies; however, as a focus of this review, considerable opportunities exist if mixtures of two kinds of polar solvents are used to create environments of microscopic polarity. For example, mixing an HBD solvent with an HBA solvent causes complex molecules (e.g., HBD—HBA pairs) to form, such that heterogeneity (local composition) is observed for simple alcohol—water mixtures [13,14] or ethylene glycol-water mixtures [15]. In this review, the emphasis is placed on taking advantage of the local composition and microscopic polarity of a solvent mixture as opposed to the bulk properties of a pure solvent, even though temperature and pressure can also be used to vary the properties of a pure solvent.
Solutes, in the context of this review, are active pharmaceutical ingredients (APIs) and bio-related molecules that can have multiple functional groups and can contain both polar (hydrophilic) and nonpolar (hydrophobic) regions in their structure. Functional groups in the solute can interact within the molecule (intramolecular) or between neighboring molecules (intermolecular) to form associated, cyclic, complex, network, or tertiary structures, and thus, the dissolution of an API into a solvent can be the result of many different molecular interactions. The composition of a solvent mixture can be used to fine-tune dipole—dipole interactions that sometimes lead to the solubility enhancement of the API in solution that is higher than that in either of the pure mono-solvents, which is known as synergistic behavior.

2. Substances of Very High Concern (SVHC)

In the synthesis and processing of APIs, polar protic (water, alcohols, carboxylic acids), dipolar aprotic (ketones, lactones, esters, ethers), or nonpolar aprotic (hydrocarbons) solvents are used. Notably, hazardous and unsafe dipolar aprotic chemicals (e.g., N,N-dimethylformamide (DMF), 1-methyl-2-pyrrolidinone (NMP), 1,4-dioxane (DI)) account for over 40% of total solvents used in synthetic, medicine-related, and process chemistry [16], and these solvents and more than 480 others are on the candidate list of substances of very high concern (SVHC), as designated under the European Chemicals Agency (ECHA), as the European Union Registration, Evaluation Authorization and Restriction of Chemicals (REACH) guidelines limit or prohibit the use of chemicals, especially those having reproductive toxicity, carcinogenicity, or explosive decomposition properties (Table 1). Thus, the key motivation of employing mixed solvents instead of mono-solvents, new solvents, or newly developed solvents is based on environmental health and safety (EHS) guidelines for compounds with known chemical properties and conformity with the “International Council for Harmonisation of Technical Requirements for Pharmaceuticals for Human Use” (ICH). Namely, EHS and ICH should be primary factors in solvent replacement, rather than apparent greenness or economic or sustainability factors, because few newly developed solvents or solvent systems have had sufficient time for scrutiny in all areas highlighted by governmental agencies and in the solvent guides discussed below. In this review, solvent replacement strategies are analyzed with the aim of highlighting a method for identifying safe solvent mixtures for the research development and chemical processing of organic compounds.

3. Solvent Guides

To address the issue of the overuse of hazardous dipolar aprotic chemicals in API synthesis and processing and to improve the awareness of chemical professionals who perform solvent selection on a day-to-day basis, pharmaceutical industries have developed solvent guides with ranking systems. Chemical agencies have developed lists for solvents evaluated as hazardous that require formal authorization for use in chemical processes.
The GlaxoSmithKline (GSK) solvent guide [7,18,19] contains detailed analyses of a total of 154 small molecules (e.g., alcohols, aromatics, carbonates, chlorinated, dipolar aprotics, esters, ethers, hydrocarbons, ketones, organic acids, water) commonly used in pharmaceutical industries. The GSK solvent guide has the following categories: (i) waste (incineration, recycling, biotreatment, VOC emissions), (ii) environment (aquatic impact, air impact), (iii) human health (health hazard, exposure potential), and (iv) safety (flammability and explosion, reactivity). The GSK solvent guide allows for the quick evaluation and qualitative comparison of replacement solvents based on four primary categories that include life-cycle assessment (LCA), and it ranks solvents in their categories on a scale from 1 (major issues) to 10 (few known issues).
The European consortium and Innovative Medicines Initiative (IMI) produced CHEM21 [8], which contains guidelines and metrics for solvent usage. Byrne et al. reported environmental, health, and safety (EHS) tools and guidelines for solvents and highlighted key points in available guidelines [11]. The CHEM21 solvent guide ranks solvents in EHS categories on a scale from 1 (recommended) to 10 (hazardous), which is contrary (and opposite in order) to the scale of the GSK solvent guide. Both solvent guides provide extremely useful evaluations of solvent risks and issues and provide solvent replacement recommendations.
The American Chemical Society (ACS) Green Chemistry Institute (CGI) and pharmaceutical roundtable produced a solvent selection website (Figure 1) [10,12] dedicated to solvent usage in pharmaceutical and chemical industries and a solvent guideline [9]. Figure 1 shows a sample screen of a solvent selection tool developed for the ACS GCI Pharmaceutical Roundtable (GCIPR) that uses principle component analysis (PCA) to identify potential solvent replacements. PCA combines many physical properties, characteristics (presence of functional groups), and environmental data to generate correlations and scores according to user constraints. The solvent selection tool (Figure 1) was described by Diorazio et al. [20] and was originally designed by AstraZeneca in Spotfire, and a version was donated to GCIPR. The GCIPR solvent selection tool is useful for identifying replacement solvents based on both quantitative and qualitative characteristics (Figure 1).

4. Replacement Solvents in Synthetic Chemistry

Syntheses of APIs are commonly performed in multistep batch processes that use hazardous or unsafe dipolar aprotic solvents in some of the key steps. Replacement strategies for non-green dipolar aprotic solvents used in reactions were suggested by Gao et al. [21].
Table 2 summarizes replacement solvents for 15 classes of synthetic reactions identified by Jordan et al. [16]. Possible replacement solvents for dipolar aprotics (Table 2) include novel water-surfactant (PS-750-M) systems that eliminate organic solvents [22], dipolar aprotic solvents with improved safety and sustainability, namely N-butyl-2-pyrrolidinone (NBP), propylene carbonate (PC), dimethylisosorbide (DMI) [23], dihydrolevoglucosenone (Cyrene) [24], eucalyptol [25], or dimethylcarbonate (DMC), or the use of mixed solvents, such as 2-methyltetrahydrofuran (2-MeTHF) with methanol (Table 2). Besides THF or DMF in Sonogashira cross-coupling reactions (Table 2), eucalyptol can possibly replace solvents such as anisole, bromobenzene, chlorobenzene, chloroform, diethyl ether (DE), N,N-dimethylacetamide (DMA), dimethyl ether (DME), DI, ethyl acetate, ethyl benzoate, and toluene [25].
In the synthesis of APIs with solvents, the type of process employed is an important point that deserves attention. A less obvious way to lower risks associated with solvent usage in API synthesis is through continuous manufacturing (CM) [26], as opposed to batch processing. In a CM process, systems can be automated, quality can be improved, waste can be reduced, and, most importantly, solvent volumes can be greatly lowered over those quantities used in batch systems by lowering the total system volumes and by eliminating the storage of API reaction intermediates, such that overall safety of the synthesis can be improved. The number of papers published on the continuous manufacturing of APIs has roughly tripled in the past 5 years, making it a highly active research area. In CM processes, solvent selection and solvent additives play key roles in flow chemistry, product quality, system operability, economics, and sustainability. Furthermore, there are some recent new approaches for CM processes; amidation by reactive extrusion has been developed as a solventless synthesis method and has been used for the preparation of teriflunomide and moclobemide APIs [27].

5. Solubility Parameters

Solubility parameters (SP) are used to characterize substances in solvent replacement strategies. The Hildebrand SP (δ) has the basis of regular solution theory [28], and its development in solubility theory relates the cohesive energy density defined by Equation (1) to the activity coefficient [29].
δ ( Δ U ¯ vap / V ¯ ) 1 / 2
In Equation (1), U and V are the molar internal energy of vaporization and molar volume of the substance in its liquid state, respectively. The definition of the Hildebrand SP is typically simplified by replacing U with (HPV) and assuming ideal gas behavior:
δ = ( ( Δ H ¯ vap R T ) / V ¯ ) 1 / 2
Hansen [30] divided the total cohesive energy given in Equation (1) into three parts: (i) dispersion (van der Waals (London) forces) interactions (δd), hydrogen bonding interactions (δh), and polar (or dipole-dipole) interactions (δp). Hansen solubility parameters (HSPs) are used to determine a solubility parameter distance (Ra) between two substances “1” and “2” as follows:
( Ra ) 2 = 4 ( δ d 1 δ d 2 ) 2 + ( δ h 1 δ h 2 ) 2 + ( δ p 1 δ p 2 ) 2
where the sphere provides a region of favorable solvation for a solute “1” and solvent “2”, i.e., as values of Ra become closer to zero according to a chosen solvent with given HSP values, affinity becomes higher, and the solubility of the solute in the solvent should increase. The factor of four in Equation (3) is empirical and adds statistical weighting to dispersion interactions as being most important in solvation. By taking a substance such as a polymer or biomolecule and seeing whether it dissolves into solvents with known HSP values, the radius of interaction (Ro) can be determined for that compound. Then, a relative energy difference can be defined as follows:
RED = Ra/Ro
and solvents or solvent mixtures that have RED < 1 are candidates that dissolve the compound. It is possible, for example, for two solvents outside of the solvation sphere to be mixed, such that they form a good solvent as mixture for a polymer. HSP theory has been used to estimate the solubilities of anti-inflammatory drugs in pure and mixed solvents [31]. Fractional HSP values, which can be plotted on ternary diagrams to facilitate the assessment of interactions, have been used to identify green extraction solvents for alkaloids [32] and to screen solvent mixtures for pharmaceutical cocrystal formation [33]. HSP is a powerful tool used for solvent screening and is especially useful for large molecules, such as polymers or biomolecules, as highlighted by Abbott [34].
In comparing the Hildebrand solubility parameter theory with that of the Hansen solubility theory, the Hildebrand solubility parameter theory has some notable failures in predicting miscibility between materials [30]. However, in a critical comparison of solvent selection for 75 polymers, both theories gave similar results in predicting polymer—solvent miscibility [35]. Namely, Hildebrand SP had a prediction accuracy of 60% for solvents and 76% for non-solvents, whereas HSP had a prediction accuracy of 67% for solvents and 76% for non-solvents [35]. On the other hand, for polar polymers, the Hildebrand SP theory gave a prediction accuracy of only 57% [35]. Both Hildebrand solubility parameters and Hansen solubility parameters are useful screening tools for solvent replacement. Hildebrand SP theory is simple and provides qualitative estimation of solvent interactions for nonpolar molecules or slightly polar molecules; Hansen SP theory accounts for detailed molecular interactions and is applicable to both nonpolar and polar molecules. HSP can be applied to complex molecules, such as lignin [36] or phytochemicals [37]; however, HSP is qualitative when hydrogen bond donor (HBD) and hydrogen bond acceptor (HBA) molecular systems are considered [38].

6. Empirical Polarity Scales

Reichardt ET(30) parameters are based on the solvatochromic properties of Betaine 30 dye and provide the sensitive characterization of solvent polarity. Reichardt ETN values are normalized based on the ET(30) values of water and tetramethylsilane. Reichardt parameters are firmly established in the chemical literature and form the basis of a widely used polarity scale for organic chemicals [6].
Kamlet—Taft (KT) parameters are based on the solvatochromism of dyes specific to Lewis acidity (α), Lewis basicity (β), and dipolarity/polarizability (π*) and have independent scales that depend on reference solvents [39]. The Kamlet—Taft polarity scales are meant to have values of α, β, and π* that are between zero and one; however, when a solvent has a Lewis acidity, Lewis basicity, or dipolarity/polarizability that is outside of the range of reference compounds, (π* = 0 (cyclohexane) and π* = 1 (dimethylsulfoxide)) values of KT parameters can be greater than unity or less than zero.
Catalán parameters improved the KT parameter approach by using specific dyes for solvent polarizability (SP), solvent dipolarity (SdP), solvent acidity (SA), and solvent basicity (SB) parameters rather than by average values, as in the KT approach. Catalán parameters separate the polarizability (SP) and dipolarity (SdP) contributions of the KT parameter approach. All three scales have wide use in the chemical literature, although there are issues in data reduction methods and parameter values, as pointed out by Spange et al. [40], who reanalyzed polarity scales considering molar concentrations of the solvent (N), and Spange and Weiß [41], who proposed a method to unify the acid—based (pKa) and density effects of hydrogen bond donor solvents.
According to Reichardt and Welton [6], common molecular solvents (Figure 2) can be roughly divided into three groupings: (i) dipolar protic (HBD), ETN > 0.5; (ii) dipolar aprotic (HBA), 0.3 < ETN < 0.5; and (iii) apolar (non-HBD or nonpolar), ETN < 0.3. Examination of the KT dipolarity/polarizability parameters (Figure 2) shows that longer chain hydrocarbons have π* values less than zero, and water has a π* greater than unity, which is due to the choice of reference solvents in the KT method. Most solvent replacement strategies consider Reichardt, Kamlet—Taft or Catalán parameters in their analysis. For example, dipolar aprotic solvents generally have high KT basicity and low KT acidity (Figure 3). Direct replacement solvents for dipolar aprotics could be N-butyl-2-pyrrolidinone (NBP), CyreneTM (Cyr), γ-valerolactone (GVL), γ-butyrolactone (GBL), eucalyptol (Eupt), tetramethyloxolane (TMO), dimethyl isosorbide (DMI), or cyclopentyl methyl ether (CPME). However, many solvents have ETN polarity values that are much lower than that of dipolar aprotics (Figure 2) and KT acidities that are either too high or KT basicities that are too low (Figure 3) to allow direct replacement of dipolar aprotics. Nevertheless, the range of Kamlet—Taft parameters of dipolar aprotics provide valuable information for considering mixed solvents and mixed solvent composition.

7. Opportunities with Mixed Solvents

Mixtures of solvents (mixed solvents) allow one to vary the chemical properties of the solution in a unique number of ways. For example, when an HBD solvent is mixed with an HBA solvent, KT parameters vary continuously with composition (Figure 4). KT parameters of mixed solvents can show synergistic behavior, which means that their β or π* values can be higher than the KT parameters of the pure solvents (Figure 4), especially when water is the HBD solvent. Duereh et al. [42] showed that there is a clear relationship between microscopic (local) polarity, complex molecule (HBD—HBA solvent pairs) interactions, and synergistic behavior in thermodynamic properties (Figure 5).
Thus, solvent composition of mixed solvents allows one to vary microscopic polarity (local composition) and the concentration of HBD—HBA complex molecules that can be used advantageously in solvent replacement schemes.
In this section, strategies for using mixed solvents to replace hazardous chemicals are highlighted for chromatography solvents, CO2 expanded liquids, supercritical fluids, low-transition temperature mixtures, switchable solvents, and HBD—HBA mixtures of molecular solvents.

7.1. Chromatography Solvents

In chromatographic methods, great progress has been made with the introduction of mixed solvents, such as ethyl acetate (EtAc)-ethanol (EtOH) in heptanes being demonstrated as a superior replacement for dichloromethane (DCM) [43]. Mixed solvent stock solutions are marketed by leading chemical suppliers for HPLC, TLC, and flash chromatography (FC) methods [44], confirming the success of the EtAc—ethanol mixtures.
The reason why EtAc—EtOH in a heptane mixed solvent system can replace DCM can be understood by examining the variation in KT parameters of the mixture compared with the KT parameters of the DCM—MeOH system. In this case, EtOH is the HBD solvent, EtAc is the HBA solvent, and the heptanes have low overall KT acidity for the mobile phase. Composition variation of EtAc–EtOH mixtures allows for the fine control of the basicity and dipolarity/polarizability that transverse methanol KT parameters (Figure 4).
To replace hexane, CO2–EtAc has been suggested to be applicable to thin-layer chromatography (Table 3), and CO2–MeOH has been demonstrated to be applicable to flash chromatography [45]. The entire corporate chemistry division of Syngenta (Table 3) reduced the overall volume of seven hazardous dipolar aprotic solvents (DCM, CHCl3, DCE, DI, DME, DMF, DE) by 75% over a period of two years by using solvent replacement (e.g., EtAc–EtOH mixtures for DCM) and by emphasizing reverse phase chromatography for the separation of polar compounds [46] (Table 3); however, DMF usage increased during that period. Solvent pairs, such as cyclohexanone–MeOH, cyclohexanone–EtOH, cyclopentanone–MeOH, cyclopentanone–EtOH, GBL–MeOH, GBL–EtOH, GBL–water, GVL–MeOH, GVL–EtOH, and GVL–water, have been demonstrated as replacements for NMP or DMF in polyamide synthesis and, thus, have possibilities as solvent replacements in analytical method development [47].
Improvements in high-pressure liquid chromatography (HPLC) have been made with the introduction of ultra-high-pressure liquid chromatography (UHPLC), supercritical fluid chromatography (SFC), and ultra-high-pressure supercritical fluid chromatography (UHPSFC), which reduce the amount of solvents necessary in analyses while improving resolution [48]. When UHPSFC—tandem mass spectroscopy is employed, the determination of plant hormones (cytokinins) can be analyzed in 9 min at detection limits close to 0.03 fmol [49]. ACS has introduced the analytical method greenness score (AMGS) calculator developed by Hicks et al. [48] that ranks chromatography methods according to instrument energy, solvent energy, and solvent EHS scores [10].
Table 3. Replacement solvents for dichloromethane (DCM) in high-performance liquid (HPLC), thin-layer chromatography (TLC) and flash chromatography (FC) methods. Analytes consist of neutral, basic, acidic, and polar API.
Table 3. Replacement solvents for dichloromethane (DCM) in high-performance liquid (HPLC), thin-layer chromatography (TLC) and flash chromatography (FC) methods. Analytes consist of neutral, basic, acidic, and polar API.
Mixed Solvent aAnalyte bSystemRef.
EtAc:EtOH (3:1) in heptanesNeutralLC[43]
EtAc:EtOH in heptanesNeutralLC[43]
iPrOH in heptanesNeutralLC[43]
EtAc:EtOH (3:1) in MTBENeutralLC[43]
MeOH in MTBENeutralLC[43]
EtAc:EtOH (3:1) (2% NH4OH) in heptanesBasicLC[43]
MeOH: NH4OH (10:1) in EtAcBasicLC[43]
MeOH: NH4OH (10:1) in MTBEBasicLC[43]
EtAc:EtOH (3:1) (2% AcOH) in heptanesAcidicLC[43]
MeOH:AcOH (10:1) in EtAcAcidicLC[43]
MeOH:AcOH (10:1) in MTBEAcidicLC[43]
EtAc:EtOH (3:1) in cyclohexanen.s.LC[46]
acetonitrile:waterPolarLC[46]
tert-butyl acetateAllLC[50]
sec-butyl acetateAllLC[50]
ethyl isobutyrateAllLC[50]
methyl pivalateAllLC[50]
CO2:EtAcn.s.TLC[51]
EtAc in heptanesn.s.TLC[51]
iPrOH in heptanesn.s.TLC[51]
Ace in heptanesn.s.TLC[51]
CO2:MeOHNeutralFC[45]
CO2:EtAcn.s.FC[51]
CO2:Acen.s.FC[51]
CO2:iPrOHn.s.FC[51]
a AcOH: acetic acid; EtAc: ethyl acetate; MTBE: methyl tert-butyl ether; b n.s. not specified.

7.2. Expanded Liquids and Supercritical Fluids

Chemists and chemical engineers have introduced many new types of solvents through major research initiatives. CO2-expanded bio-based liquids (CXL) have been demonstrated to be favorable for enantioselective biocatalysis [52], and supercritical fluids have been shown to be able to replace the hazardous solvents used in processing APIs [53]. Supercritical carbon dioxide (scCO2) has been shown to have a wide application in processing bioactive lipids [54] and bioactive-related food ingredients [55]. A comprehensive review is available on the supercritical extraction of bioactive molecules from plant matrices [56]. A less-studied methodology in the supercritical extraction of bioactive molecules from natural sources is to eliminate organic co-solvents, such as ethanol or acetone, by replacing them with co-extractants that are typically oils from plant materials (Table 4).
In co-extractant methodology (Table 4), natural source substrates (petals, pericarp, etc.) are mixed with a natural oil (co-extractant) from a vegetable, drupe, legume, or seed (or fruit) before extraction with pure scCO2. The co-extractant serves to increase the mass transfer of active components from the natural source to the supercritical phase by solubilization and polarity matching, and the co-extractant properties are enhanced due to scCO2 dissolution into the co-extractant phase that causes the reduction of both surface tension and viscosity while enhancing heat transfer and related properties. Thus, with co-extractant methodology (Table 4), organic co-solvents are completely eliminated in scCO2 extraction such that the contamination of extracts with organic compounds is not an issue. Furthermore, with co-extractant methodology, a final product is realized directly, the cultural processing of many types of food is possible, and food safety is strictly enhanced [70].
Related to developments in supercritical fluid theory, entropy based solubility parameters have been proposed that allow the extension of traditional solubility parameter theory to chemical systems containing supercritical fluids and ethanol [71] or systems at high temperatures or high pressures [72]. Experimental systems for measuring the KT parameters of methanol, ethanol, 2-propanol, and 1,1,1,2-tetrafluoroethane (HFC134a) co-solvents in CO2 have been developed for assessing the HBD alcohol interactions with the HBA Lewis acidity of CO2 in the supercritical state for quantifying polarity enhancements [73].

7.3. Low Transition Temperature Mixtures

Low transition temperature mixtures (LTTMs) are special combinations of mixed solvents made up of a hydrogen bond donor (HBD) molecule and a hydrogen bond acceptor (HBA) molecule for the purpose of liquefying the mixture [74]. Ionic liquids (ILs) are combinations of discrete organic moiety containing cations and anions that are in the liquid state at room temperature. Deep eutectic solvents (DESs) are mixtures of Lewis or Brønsted acids and bases that are in the liquid state at room temperature.
The possibility of using either ILs or DESs as solvent replacements or for processing APIs allows them to have many potential innovative applications due to their solvation and tailorable properties [75]. Issues with ILs are their cost, recyclability, and relatively higher viscosity compared to molecular solvents. While DESs are inexpensive, they share some of the same issues as ILs, and in addition, their separation from chemical products may be problematic due to the formation of strong HBD—HBA complexes with the API. One innovative approach that addresses some of these issues is to incorporate the IL chemical structure into the API to improve the bioavailability in drug delivery systems [76,77]. Reviews in the area of combining HBD- or HBA-containing APIs into the structure of ILs for drug delivery systems and other purposes show that there is much activity in this research area [78,79].

7.4. Switchable Solvents

Switchable polarity solvents (SPS) [80], switchable hydrophilicity solvents (SHS) [81], switchable water (SW) [82], solvent-assisted switchable water (SASW) [83], and high-pressure switchable water (HPSW) [84] are new types of mixed solvents that can change their polarity, hydrophilicity, or characteristics through the introduction or removal of CO2. Switchable solvent systems would seem to have many applications in processing API, and furthermore, it could be highly advantageous if APIs with an existing or added amidine group could have modified hydrophilicity with CO2 [85] for the purposes of separation, purification, or analysis.

7.5. HBD—HBA Mixtures of Molecular Solvents

The attractiveness of using molecular solvents to form HBD—HBA mixtures is that their EHS data are available, making it possible to assess their safety. With the EHS safety of the solvents assessed, it becomes possible to focus on the technical issue of solubilizing the API in the mixed solvent for processing operations.
Duereh et al. [86] developed a methodology for replacing dipolar aprotic solvents with safe HBD—HBA solvent pairs based on solubility and Kamlet—Taft windows (Figure 6). In the methodology [86], solvent pairs are evaluated from a database with user-defined solubility parameter and KT parameter windows of an API to determine working compositions and a prioritized list of mixed solvents according to a composite GSK score. The open-access software given in ref. [86] can be extended with activity coefficient models or quantum chemistry methods to broaden the scope of the methodology.
In developing solvent replacement methodologies, physical properties can be important attributes for solvent selection. Jouyban and Acree [87] developed a single functional form for the correlation of viscosity, density, dielectric constant, surface tension, speed of sound, Reichardt ETN, molar volume, and isentropic compressibility of binary mixed solvents. Nazemieh et al. [88] reported data for a new set of mixed solvents, namely p-cymene with α-pinene, limonene, and citral correlated with the Jouyban—Acree model for physico-chemical properties (PCPs). Lee et al. [89] developed a local composition regular solution theory model for the correlation and prediction of API solubility in mixed solvents that had a single functional form for all compounds studied. The advantage of similar functional forms in correlative and predictive schemes for PCPs and activity coefficients is that machine learning techniques can be applied as the size of the database increases.

8. Kamlet—Taft Parameter Windows for APIs

APIs are commonly designated as being water soluble or non-water soluble. When the Reichardt ETN and KT parameters are plotted for mono-solvents that solvate 45 water-soluble APIs (Figure 7) and 47 water-insoluble APIs (Figure 8), the range of ETN and KT parameter values becomes visible, which characterizes the apparent polarity of the API. Although many water-soluble APIs are solvated by dipolar protic solvents (ETN > 0.5) and dipolar aprotic solvents (0.3 < ETN < 0.5) over a wide range of KT acidities (α), there are minimum values of π* and β required for solvation (Figure 7). On the other hand, water-insoluble APIs are solvated by a relatively narrow range of solvent polarities (0.2 < ETN < 0.75) and KT acidities (α), in which there are maximum values of π* and minimum values of β required for solvation (Figure 8).
When an API is dipolar protic, it interacts with basic dipolar aprotic solvents by forming hydrogen bonds with the solvent that must be stronger than those in the solid phase for solvation to occur. If the dipolar aprotic solvent is not basic or if it has insufficient basicity, then the dipolar protic API will have low solubility in the solvent, because dipole-dipole interactions generally do not have sufficient strength to break hydrogen bonds in the solid phase. On the other hand, when an API is dipolar aprotic, it interacts with dipolar aprotic solvents through dipole—dipole interactions that must be stronger than those in the solid phase for solvation to occur.
Conversely, if a solvent is dipolar protic, then the API must have sufficient basicity to accept hydrogen bonds that must be stronger than those in the solvent phase. Thus, scales for molecular basicity are extremely important in identifying potential new solvents and solvent systems. However, note that all KT α, β, and π* parameters influence API solubility in a mixed solvent and that they depend on the mixed solvent local composition, frequently in a non-linear or synergistic way [90,91].
Figure 7. Reichardt ETN and Kamlet—Taft parameters of mono-solvents that solvate water-soluble APIs at ca. 25 °C. Data from refs. [1,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302]. Water (Blue). Less-hazardous solvents (Green). Hazardous solvents (Red). Detailed information in Supplementary Materials.
Figure 7. Reichardt ETN and Kamlet—Taft parameters of mono-solvents that solvate water-soluble APIs at ca. 25 °C. Data from refs. [1,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302]. Water (Blue). Less-hazardous solvents (Green). Hazardous solvents (Red). Detailed information in Supplementary Materials.
Liquids 04 00018 g007
Figure 8. Reichardt ETN and Kamlet—Taft parameters of mono-solvents that solvate water-insoluble APIs at ca. 25 °C. Data from refs. [1,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302]. Less-hazardous solvents (Green). Hazardous solvents (Red). Detailed information in Supplementary Materials.
Figure 8. Reichardt ETN and Kamlet—Taft parameters of mono-solvents that solvate water-insoluble APIs at ca. 25 °C. Data from refs. [1,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302]. Less-hazardous solvents (Green). Hazardous solvents (Red). Detailed information in Supplementary Materials.
Liquids 04 00018 g008
Consider the HBD—HBA mixed-solvent systems shown in Figure 9. When water is used as the HBD solvent (Figure 9a–c), the HBA solvent addition (increasing x2) lowers the KT α and generally lowers π* values depending on the HBA polarity and causes KT β values to initially sharply increase, during which the microscopic polarity changes greatly due to the formation of complex molecules [91]. For example, water—lactone mixed solvents have been shown to exhibit synergy in KT basicity [91]. For an alcohol as the HBD solvent, the addition of an HBA solvent lowers the KT α, and it causes the KT π* and KT β values to linearly increase or decrease with bulk composition depending on whether the pure alcohol KT π* or KT β values are less than, equal to, or greater than those of the HBA solvent alcohol KT π* or KT β values (Figure 9d–i). Duereh et al. showed examples of the case of ethanol (HBD) –cyclopentanone (HBA), in which mixed solvent composition can be used to favorably solvate an API (paracetamol), and they also showed a case of methanol (HBD) –cyclopentanone (HBA), in which mixed solvent composition failed to provide any solvation benefit, along with examples of 12 APIs [90].
There are a number of HBD—HBA solvent combinations that could be replacements for hazardous solvents (Figure 9). For possible HBD solvents, water, methanol, and ethanol are good candidates. When water is the HBD solvent, possible candidate HBA solvents are acetone, acetic acid, acetonitrile, EtOH, MeOH, 2-MeTHF, water-2,2,5,5-tetramethyloxolane (TMO), DMI, Cyrene, Cygnet 0.0, or possibly diformylxylose. Safety and conditions must be considered carefully. For example, 2-MeTHF forms peroxides more rapidly than IPE, THF, or CPME when inhibitors are not present; ethereal solvents form peroxides [7]. Cygnet 0.0 is solid at room temperature [303], and 2-MeTHF in water has inverse temperature behavior up until temperatures of 340 K [304], meaning that its solubility in water decreases with increasing temperature.
When alcohols are used as the HBD solvent, cyclohexanone (CHN), cyclopentanone (CPN), many kinds of esters, GBL, GVL, eucalytol (water insoluble), or possibly MeSesamol (water insoluble) [305] or diformylxylose [306] are candidates. Furthermore, interesting HBA—HBA combinations, such as Cyrene—Cygnet 0.0, are being suggested for polymer syntheses [303] to replace hazardous dipolar aprotic solvents, and these types of HBA—HBA mixed solvents could have advantages in processing APIs.

9. Linear Solvation Energy Relationships (LSER)

Polarity parameters originally reported by Kamlet, Abboud, Abraham, and Taft were intended for use in linear solvation energy relationships (LSER) [307], expressed as follows:
X Y Z = X Y Z 0 + s ( π + d δ ) + a α + b β + h δ H + e ξ
where XYZ is a chemical phenomenon, XYZ0 is a reference phenomenon, and s, a, b, h, and e are descriptors that are used to correlate polarity parameters to XYZ. Many adaptations have been made of Equation (5), and a well-known one is due to Abraham [308], which expressed water—octanol partition coefficients (log P) and gas—solvent partition coefficient (log K) as follows [309]:
log (P) = c + eE + sS+ aA + bB + vV
log (K) = c + eE + sS+ aA + bB + vL
Where the bold symbols are properties of the solute related to excess molar refraction (E), dipolarity/polarizability (S), hydrogen bond acidity (A), Lewis basicity (B), McGowan’s molecular volume (V), and gas-to-hexadecane partition coefficient (L). LSER models are directly applicable to predict the solubility of APIs in solvents [309]. LSER models are widely used in the field of chromatography for characterizing columns and estimating retention times [310,311] or in the analysis of petroleum distillate conditions with group contribution activity coefficient models such as UNIFAC [312], but they do not appear to have been used more broadly (in reverse) in mixed solvent replacement schemes, although environmentally related partition coefficients are incorporated into life-cycle assessment tools, such as EPA’s CompTox chemical dashboard system [313] or machine learning studies for solvent characterization factors [314].

10. Conclusions

In this work, several strategies were highlighted for the replacement of hazardous dipolar aprotic solvents related to pharmaceutical and bio-related compounds. Solvent guides form the basis of solvent replacement and consider categories of safety, human health, environment, waste, and sustainability. Linking online solvent selection sites with GSK, CHEM21, ECHA, and other guidelines would allow for the efficient dissemination of solvent replacements.
An example of drop-in replacement solvents and several mixed solvent combinations for synthesizing APIs is one strategy that shows it is possible for academia and the industry to replace hazardous dipolar aprotic solvents by adopting new chemical systems that are both efficient and safe. The universal guide for the replacement of hazardous dipolar aprotic solvents in synthetic chemistry is one of the key strategies.
Mixed solvents can be used in many ways to replace hazardous solvents, often with a performance benefit. Dichloromethane can be replaced by ethanol (HBD) and ethyl acetate (HBA) mixed solvents, as is evident from marketed stock solutions by chemical companies. The use of CO2 with esters or alcohols instead of hexane or chlorinated hydrocarbons is seen to be effective for thin-layer, flash, and supercritical chromatography, and with the introduction of marketed industrial analytical equipment, it is clear that the new technology will become established.
Expanded liquids, supercritical fluids, low-transition temperature (HBD—HBA) mixtures, and switchable solvents all offer safer chemical systems that have low energy, performance, and sustainability benefits. Chemical systems based on HBD—HBA mixtures of molecular solvents for processing APIs offer a simple way to replace hazardous solvents by considering the range of solubility parameters, Reichardt polarity, and Kamlet—Taft parameters of the pure components. Reichardt polarity and Kamlet—Taft parameters of pure components are necessary physical properties for the development of solvent replacement strategies. By using the available solubility data of APIs in mono-solvents, new mixed solvent combinations can be seen.

11. Future Outlook

Presently, there are many measurements of Reichardt polarity and Kamlet—Taft parameters of pure compounds, but far fewer measurements have been made for mixed solvent systems that can potentially replace hazardous dipolar aprotic solvents. Many new measurements are needed of Reichardt polarity and Kamlet—Taft parameters of HBD—HBA and HBA—HBA mixed solvents, especially those systems such as ethanol—ethyl acetate, to understand fundamental interactions of complex molecules with APIs.
Theoretical methods applied to HBD—HBA systems could greatly accelerate the identification of new chemical systems for processing APIs. COSMO-RS is able to quantitatively predict Kamlet—Taft parameters for both molecular solvents and deep eutectic solvents [315]. COSMO-RS gives qualitative predictions of Hansen solubility parameters [316], which is encouraging because the values of APIs could lead to a great reduction in experimental effort.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/liquids4020018/s1, Table S1. Water-soluble APIs solvated by monosolvents and their solvent polarity (ETN), Kamlet-Taft acidity (α), basicity (β) window, dipolarity/polarizability (π*) and corresponding literature. Solvents listed as hazardous in GSK solvent guide are highlighted in red. Table S2. Water-insoluble APIs solvated by monosolvents and their solvent polarity (ETN), Kamlet-Taft acidity (α), basicity (β) window, dipolarity/polarizability (π*) and corresponding literature. Solvents listed as hazardous in GSK solvent guide are highlighted in red.

Author Contributions

J.L.L.: data curation, investigation, formal analysis, visualization, software. G.H.C.: supervision, resources, project administration, funding acquisition, writing—reviewing and editing. M.O.: writing-reviewing and editing. H.G.: writing—reviewing and editing. R.L.S.J.: conceptualization, supervision, methodology, formal analysis, project administration, funding acquisition, writing—original draft, reviewing and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge Trumer Medicare Sdn Bhd (Vot number: 6300827-11801) for the funding of the subscription of software.

Acknowledgments

Partial support of this project in the form of facilities and resources of the Research Center of Supercritical Fluid Technology, Tohoku University, is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Reichardt, C. Empirical Parameters of Solvent Polarity as Linear Free-Energy Relationships. Angew. Chem. Int. Ed. Engl. 1979, 18, 98–110. [Google Scholar] [CrossRef]
  2. Krygowski, T.M.; Wrona, P.K.; Zielkowska, U.; Reichardt, C. Empirical parameters of lewis acidity and basicity for aqueous binary solvent mixtures. Tetrahedron 1985, 41, 4519–4527. [Google Scholar] [CrossRef]
  3. Reichardt, C. Solvation Effects in Organic Chemistry: A Short Historical Overview. J. Org. Chem. 2022, 87, 1616–1629. [Google Scholar] [CrossRef] [PubMed]
  4. Kamlet, M.J.; Taft, R.W. The Solvatochromic Comparison Method. I. The β-Scale Of Solvent Hydrogen-Bond Acceptor (HBA) Basicities. J. Am. Chem. Soc. 1976, 98, 377–383. [Google Scholar] [CrossRef]
  5. Catalán, J. On the empirical scales of organic solvents established using probe/homomorph pairs. J. Phys. Org. Chem. 2021, 34, e4206. [Google Scholar] [CrossRef]
  6. Reichardt, C.; Welton, T. Solvents and Solvent Effects in Organic Chemistry; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2011. [Google Scholar] [CrossRef]
  7. Alder, C.M.; Hayler, J.D.; Henderson, R.K.; Redman, A.M.; Shukla, L.; Shuster, L.E.; Sneddon, H.F. Updating and further expanding GSK’s solvent sustainability guide. Green Chem. 2016, 18, 3879–3890. [Google Scholar] [CrossRef]
  8. Prat, D.; Wells, A.; Hayler, J.; Sneddon, H.; McElroy, C.R.; Abou-Shehada, S.; Dunn, P.J. CHEM21 selection guide of classical- and less classical-solvents. Green Chem. 2016, 18, 288–296. [Google Scholar] [CrossRef]
  9. ACS. ACS GCI Pharmaceutical Roundtable. Collaboration to Deliver a Solvent Selection Guide for the Pharmaceutical Industry. 2018. Available online: https://www.acs.org/content/dam/acsorg/greenchemistry/industriainnovation/roundtable/solvent-selection-guide.pdf (accessed on 16 December 2023).
  10. ACS. Tools for Innovation in Chemistry. 2023. Available online: https://www.acsgcipr.org/tools-for-innovation-in-chemistry/ (accessed on 16 December 2023).
  11. Byrne, F.P.; Jin, S.; Paggiola, G.; Petchey, T.H.M.; Clark, J.H.; Farmer, T.J.; Hunt, A.J.; Robert McElroy, C.; Sherwood, J. Tools and techniques for solvent selection: Green solvent selection guides. Sustain. Chem. Process. 2016, 4, 7. [Google Scholar] [CrossRef]
  12. Diorazio, L.J.; Richardson, P.; Sneddon, H.F.; Moores, A.; Briddell, C.; Martinez, I. Making Sustainability Assessment Accessible: Tools Developed by the ACS Green Chemistry Institute Pharmaceutical Roundtable. ACS Sustain. Chem. Eng. 2021, 9, 16862–16864. [Google Scholar] [CrossRef]
  13. Dixit, S.; Crain, J.; Poon, W.C.K.; Finney, J.L.; Soper, A.K. Molecular segregation observed in a concentrated alcohol–water solution. Nature 2002, 416, 829–832. [Google Scholar] [CrossRef]
  14. Ono, T.; Horikawa, K.; Ota, M.; Sato, Y.; Inomata, H. Insight into the local composition of the Wilson equation at high temperatures and pressures through molecular simulations of methanol-water mixtures. J. Chem. Eng. Data 2014, 59, 1024–1030. [Google Scholar] [CrossRef]
  15. Ono, T.; Ito, Y.; Ota, M.; Takebayashi, Y.; Furuya, T.; Inomata, H. Difference in aqueous solution structure at 293.2 and 473.2 K between ethanol and ethylene glycol via molecular dynamics. J. Mol. Liq. 2022, 368, 120764. [Google Scholar] [CrossRef]
  16. Jordan, A.; Hall, C.G.J.; Thorp, L.R.; Sneddon, H.F. Replacement of Less-Preferred Dipolar Aprotic and Ethereal Solvents in Synthetic Organic Chemistry with More Sustainable Alternatives. Chem. Rev. 2022, 122, 6749–6794. [Google Scholar] [CrossRef] [PubMed]
  17. ECHA. Candidate List of Substances of Very High Concern for Authorisation. 2023. Available online: https://echa.europa.eu/candidate-list-table (accessed on 24 December 2023).
  18. Henderson, R.K.; Jiménez-González, C.; Constable, D.J.C.; Alston, S.R.; Inglis, G.G.A.; Fisher, G.; Sherwood, J.; Binks, S.P.; Curzons, A.D. Expanding GSK’s solvent selection guide—Embedding sustainability into solvent selection starting at medicinal chemistry. Green Chem. 2011, 13, 854–862. [Google Scholar] [CrossRef]
  19. Jiménez-González, C.; Curzons, A.D.; Constable, D.J.C.; Cunningham, V.L. Expanding GSK’s Solvent Selection Guide—Application of life cycle assessment to enhance solvent selections. Clean Technol. Environ. Policy 2004, 7, 42–50. [Google Scholar] [CrossRef]
  20. Diorazio, L.J.; Hose, D.R.J.; Adlington, N.K. Toward a More Holistic Framework for Solvent Selection. Org. Process Res. Dev. 2016, 20, 760–773. [Google Scholar] [CrossRef]
  21. Gao, F.; Bai, R.; Ferlin, F.; Vaccaro, L.; Li, M.; Gu, Y. Replacement strategies for non-green dipolar aprotic solvents. Green Chem. 2020, 22, 6240–6257. [Google Scholar] [CrossRef]
  22. Gabriel, C.M.; Keener, M.; Gallou, F.; Lipshutz, B.H. Amide and Peptide Bond Formation in Water at Room Temperature. Org. Lett. 2015, 17, 3968–3971. [Google Scholar] [CrossRef]
  23. Dalla Torre, D.; Annatelli, M.; Aricò, F. Acid catalyzed synthesis of dimethyl isosorbide via dimethyl carbonate chemistry. Catal. Today 2023, 423, 113892. [Google Scholar] [CrossRef]
  24. Sherwood, J.; Constantinou, A.; Moity, L.; McElroy, C.R.; Farmer, T.J.; Duncan, T.; Raverty, W.; Hunt, A.J.; Clark, J.H. Dihydrolevoglucosenone (Cyrene) as a bio-based alternative for dipolar aprotic solvents. Chem. Commun. 2014, 50, 9650–9652. [Google Scholar] [CrossRef]
  25. Campos, J.F.; Scherrmann, M.-C.; Berteina-Raboin, S. Eucalyptol: A new solvent for the synthesis of heterocycles containing oxygen, sulfur and nitrogen. Green Chem. 2019, 21, 1531–1539. [Google Scholar] [CrossRef]
  26. Domokos, A.; Nagy, B.; Szilágyi, B.; Marosi, G.; Nagy, Z.K. Integrated Continuous Pharmaceutical Technologies—A Review. Org. Process Res. Dev. 2021, 25, 721–739. [Google Scholar] [CrossRef]
  27. Lavayssiere, M.; Lamaty, F. Amidation by reactive extrusion for the synthesis of active pharmaceutical ingredients teriflunomide and moclobemide. Chem. Commun. 2023, 59, 3439–3442. [Google Scholar] [CrossRef] [PubMed]
  28. Hildebrand, J.H. solubility. xii. Regular solutions1. J. Am. Chem. Soc. 1929, 51, 66–80. [Google Scholar] [CrossRef]
  29. Prausnitz, J.M.; Lichtenthaler, R.N.; Azevedo, E.G.D. Molecular Thermodynamics of Fluid-Phase Equilibria, 3rd ed.; Prentice Hall PTR: London, UK, 1999. [Google Scholar]
  30. Hansen, C.M. Hansen Solubility Parameters. In A User’s Handbook, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2007. [Google Scholar] [CrossRef]
  31. Takebayashi, Y.; Sue, K.; Furuya, T.; Yoda, S. Solubilities of Organic Semiconductors and Nonsteroidal Anti-inflammatory Drugs in Pure and Mixed Organic Solvents: Measurement and Modeling with Hansen Solubility Parameter. J. Chem. Eng. Data 2018, 63, 3889–3901. [Google Scholar] [CrossRef]
  32. Wu, Y.; Li, W.; Vovers, J.; Thuan Lu, H.; Stevens, G.W.; Mumford, K.A. Investigation of green solvents for the extraction of phenol and natural alkaloids: Solvent and extractant selection. Chem. Eng. J. 2022, 442, 136054. [Google Scholar] [CrossRef]
  33. Kumar, A.; Nanda, A. In-silico methods of cocrystal screening: A review on tools for rational design of pharmaceutical cocrystals. J. Drug Deliv. Sci. Technol. 2021, 63, 102527. [Google Scholar] [CrossRef]
  34. Abbott, S. Solubility, similarity, and compatibility: A general-purpose theory for the formulator. Curr. Opin. Colloid Interface Sci. 2020, 48, 65–76. [Google Scholar] [CrossRef]
  35. Venkatram, S.; Kim, C.; Chandrasekaran, A.; Ramprasad, R. Critical Assessment of the Hildebrand and Hansen Solubility Parameters for Polymers. J. Chem. Inf. Model. 2019, 59, 4188–4194. [Google Scholar] [CrossRef]
  36. Ma, Q.; Yu, C.; Zhou, Y.; Hu, D.; Chen, J.; Zhang, X. A review on the calculation and application of lignin Hansen solubility parameters. Int. J. Biol. Macromol. 2024, 256, 128506. [Google Scholar] [CrossRef]
  37. Novaes, F.J.M.; de Faria, D.C.; Ferraz, F.Z.; de Aquino Neto, F.R. Hansen Solubility Parameters Applied to the Extraction of Phytochemicals. Plants 2023, 12, 3008. [Google Scholar] [CrossRef] [PubMed]
  38. Otárola-Sepúlveda, J.; Cea-Klapp, E.; Aravena, P.; Ormazábal-Latorre, S.; Canales, R.I.; Garrido, J.M.; Valerio, O. Assessment of Hansen solubility parameters in deep eutectic solvents for solubility predictions. J. Mol. Liq. 2023, 388, 122669. [Google Scholar] [CrossRef]
  39. Kamlet, M.J.; Abboud, J.L.; Taft, R.W. The solvatochromic comparison method. 6. The .pi.* scale of solvent polarities. J. Am. Chem. Soc. 1977, 99, 6027–6038. [Google Scholar] [CrossRef]
  40. Spange, S.; Weiß, N.; Schmidt, C.H.; Schreiter, K. Reappraisal of Empirical Solvent Polarity Scales for Organic Solvents. Chem. Methods 2021, 1, 42–60. [Google Scholar] [CrossRef]
  41. Spange, S.; Weiß, N. Empirical Hydrogen Bonding Donor (HBD) Parameters of Organic Solvents Using Solvatochromic Probes—A Critical Evaluation. ChemPhysChem 2023, 24, e202200780. [Google Scholar] [CrossRef] [PubMed]
  42. Duereh, A.; Sato, Y.; Smith, R.L.; Inomata, H.; Pichierri, F. Does Synergism in Microscopic Polarity Correlate with Extrema in Macroscopic Properties for Aqueous Mixtures of Dipolar Aprotic Solvents? J. Phys. Chem. B 2017, 121, 6033–6041. [Google Scholar] [CrossRef]
  43. Taygerly, J.P.; Miller, L.M.; Yee, A.; Peterson, E.A. A convenient guide to help select replacement solvents for dichloromethane in chromatography. Green Chem. 2012, 14, 3020–3025. [Google Scholar] [CrossRef]
  44. Sigma-Aldrich. Greener Chromatography Solvents. 2015. Available online: https://www.sigmaaldrich.com/content/dam/sigma-aldrich/docs/Sigma-Aldrich/General_Information/1/greener-chromatography-solvents-82207.pdf (accessed on 27 December 2023).
  45. McClain, R.; Rada, V.; Nomland, A.; Przybyciel, M.; Kohler, D.; Schlake, R.; Nantermet, P.; Welch, C.J. Greening Flash Chromatography. ACS Sustain. Chem. Eng. 2016, 4, 4905–4912. [Google Scholar] [CrossRef]
  46. Dutta, P.; McGranaghan, A.; Keller, I.; Patil, Y.; Mulholland, N.; Murudi, V.; Prescher, H.; Smith, A.; Carson, N.; Martin, C.; et al. A case study in green chemistry: The reduction of hazardous solvents in an industrial R&D environment. Green Chem. 2022, 24, 3943–3956. [Google Scholar] [CrossRef]
  47. Duereh, A.; Sato, Y.; Smith, R.L.; Inomata, H. Replacement of Hazardous Chemicals Used in Engineering Plastics with Safe and Renewable Hydrogen-Bond Donor and Acceptor Solvent-Pair Mixtures. ACS Sustain. Chem. Eng. 2015, 3, 1881–1889. [Google Scholar] [CrossRef]
  48. Hicks, M.B.; Farrell, W.; Aurigemma, C.; Lehmann, L.; Weisel, L.; Nadeau, K.; Lee, H.; Moraff, C.; Wong, M.; Huang, Y.; et al. Making the move towards modernized greener separations: Introduction of the analytical method greenness score (AMGS) calculator. Green Chem. 2019, 21, 1816–1826. [Google Scholar] [CrossRef]
  49. Petřík, I.; Pěnčík, A.; Stýskala, J.; Tranová, L.; Amakorová, P.; Strnad, M.; Novák, O. Rapid profiling of cytokinins using supercritical fluid chromatography coupled with tandem mass spectrometry. Anal. Chim. Acta 2024, 1285, 342010. [Google Scholar] [CrossRef]
  50. Lynch, J.; Sherwood, J.; McElroy, C.R.; Murray, J.; Shimizu, S. Dichloromethane replacement: Towards greener chromatography via Kirkwood–Buff integrals. Anal. Methods 2023, 15, 596–605. [Google Scholar] [CrossRef] [PubMed]
  51. Duereh, A.; Smith, R.L. Strategies for using hydrogen-bond donor/acceptor solvent pairs in developing green chemical processes with supercritical fluids. J. Supercrit. Fluids 2018, 141, 182–197. [Google Scholar] [CrossRef]
  52. Hoang, H.N.; Nagashima, Y.; Mori, S.; Kagechika, H.; Matsuda, T. CO2-expanded bio-based liquids as novel solvents for enantioselective biocatalysis. Tetrahedron 2017, 73, 2984–2989. [Google Scholar] [CrossRef]
  53. Kankala, R.K.; Zhang, Y.S.; Wang, S.-B.; Lee, C.-H.; Chen, A.-Z. Supercritical Fluid Technology: An Emphasis on Drug Delivery and Related Biomedical Applications. Adv. Healthc. Mater. 2017, 6, 1700433. [Google Scholar] [CrossRef] [PubMed]
  54. Temelli, F. Perspectives on supercritical fluid processing of fats and oils. J. Supercrit. Fluids 2009, 47, 583–590. [Google Scholar] [CrossRef]
  55. Temelli, F. Perspectives on the use of supercritical particle formation technologies for food ingredients. J. Supercrit. Fluids 2018, 134, 244–251. [Google Scholar] [CrossRef]
  56. De Melo, M.M.R.; Silvestre, A.J.D.; Silva, C.M. Supercritical fluid extraction of vegetable matrices: Applications, trends and future perspectives of a convincing green technology. J. Supercrit. Fluids 2014, 92, 115–176. [Google Scholar] [CrossRef]
  57. Krichnavaruk, S.; Shotipruk, A.; Goto, M.; Pavasant, P. Supercritical carbon dioxide extraction of astaxanthin from Haematococcus pluvialis with vegetable oils as co-solvent. Bioresour. Technol. 2008, 99, 5556–5560. [Google Scholar] [CrossRef]
  58. Saravana, P.S.; Getachew, A.T.; Cho, Y.-J.; Choi, J.H.; Park, Y.B.; Woo, H.C.; Chun, B.S. Influence of co-solvents on fucoxanthin and phlorotannin recovery from brown seaweed using supercritical CO2. J. Supercrit. Fluids 2017, 120, 295–303. [Google Scholar] [CrossRef]
  59. Sun, M.; Temelli, F. Supercritical carbon dioxide extraction of carotenoids from carrot using canola oil as a continuous co-solvent. J. Supercrit. Fluids 2006, 37, 397–408. [Google Scholar] [CrossRef]
  60. Kok, S.L.; Lee, W.J.; Smith, R.L.; Suleiman, N.; Jom, K.N.; Vangnai, K.; Bin Sharaai, A.H.; Chong, G.H. Role of virgin coconut oil (VCO) as co-extractant for obtaining xanthones from mangosteen (Garcinia mangostana) pericarp with supercritical carbon dioxide extraction. J. Supercrit. Fluids 2021, 176, 105305. [Google Scholar] [CrossRef]
  61. Lee, W.J.; Ng, C.C.; Ng, J.S.; Smith, R.L.; Kok, S.L.; Hee, Y.Y.; Lee, S.Y.; Tan, W.K.; Zainal Abidin, N.H.; Halim Lim, S.A.; et al. Supercritical carbon dioxide extraction of α-mangostin from mangosteen pericarp with virgin coconut oil as co-extractant and in-vitro bio-accessibility measurement. Process Biochem. 2019, 87, 213–220. [Google Scholar] [CrossRef]
  62. Gao, Y.; Liu, X.; Xu, H.; Zhao, J.; Wang, Q.; Liu, G.; Hao, Q. Optimization of supercritical carbon dioxide extraction of lutein esters from marigold (Tagetes erecta L.) with vegetable oils as continuous co-solvents. Sep. Purif. Technol. 2010, 71, 214–219. [Google Scholar] [CrossRef]
  63. Ma, Q.; Xu, X.; Gao, Y.; Wang, Q.; Zhao, J. Optimisation of supercritical carbon dioxide extraction of lutein esters from marigold (Tagetes erect L.) with soybean oil as a co-solvent. Int. J. Food Sci. Technol. 2008, 43, 1763–1769. [Google Scholar] [CrossRef]
  64. Pattiram, P.D.; Abas, F.; Suleiman, N.; Mohamad Azman, E.; Chong, G.H. Edible oils as a co-extractant for the supercritical carbon dioxide extraction of flavonoids from propolis. PLoS ONE 2022, 17, e0266673. [Google Scholar] [CrossRef]
  65. Shi, X.; Wu, H.; Shi, J.; Xue, S.J.; Wang, D.; Wang, W.; Cheng, A.; Gong, Z.; Chen, X.; Wang, C. Effect of modifier on the composition and antioxidant activity of carotenoid extracts from pumpkin (Cucurbita maxima) by supercritical CO2. LWT-Food Sci. Technol. 2013, 51, 433–440. [Google Scholar] [CrossRef]
  66. Fikri, I.; Yulianah, Y.; Lin, T.-C.; Lin, R.-W.; Chen, U.-C.; Lay, H.-L. Optimization of supercritical fluid extraction of dihydrotanshinone, cryptotanshinone, tanshinone I, and tanshinone IIA from Salvia miltiorrhiza with a peanut oil modifier. Chem. Eng. Res. Des. 2022, 180, 220–231. [Google Scholar] [CrossRef]
  67. Saldaña, M.D.A.; Temelli, F.; Guigard, S.E.; Tomberli, B.; Gray, C.G. Apparent solubility of lycopene and β-carotene in supercritical CO2, CO2+ethanol and CO2+canola oil using dynamic extraction of tomatoes. J. Food Eng. 2010, 99, 1–8. [Google Scholar] [CrossRef]
  68. Vasapollo, G.; Longo, L.; Rescio, L.; Ciurlia, L. Innovative supercritical CO2 extraction of lycopene from tomato in the presence of vegetable oil as co-solvent. J. Supercrit. Fluids 2004, 29, 87–96. [Google Scholar] [CrossRef]
  69. Shi, J.; Yi, C.; Xue, S.J.; Jiang, Y.; Ma, Y.; Li, D. Effects of modifiers on the profile of lycopene extracted from tomato skins by supercritical CO2. J. Food Eng. 2009, 93, 431–436. [Google Scholar] [CrossRef]
  70. Yara-Varón, E.; Li, Y.; Balcells, M.; Canela-Garayoa, R.; Fabiano-Tixier, A.S.; Chemat, F. Vegetable oils as alternative solvents for green oleo-extraction, purification and formulation of food and natural products. Molecules 2017, 22, 1474. [Google Scholar] [CrossRef] [PubMed]
  71. Ota, M.; Hashimoto, Y.; Sato, M.; Sato, Y.; Smith, R.L.; Inomata, H. Solubility of flavone, 6-methoxyflavone and anthracene in supercritical CO2 with/without a co-solvent of ethanol correlated by using a newly proposed entropy-based solubility parameter. Fluid Phase Equilibria 2016, 425, 65–71. [Google Scholar] [CrossRef]
  72. Ota, M.; Sugahara, S.; Sato, Y.; Smith, R.L.; Inomata, H. Vapor-liquid distribution coefficients of hops extract in high pressure CO2 and ethanol mixtures and data correlation with entropy-based solubility parameters. Fluid Phase Equilibria 2017, 434, 44–48. [Google Scholar] [CrossRef]
  73. Duereh, A.; Sugimoto, Y.; Ota, M.; Sato, Y.; Inomata, H. Kamlet-Taft Dipolarity/Polarizability of Binary Mixtures of Supercritical Carbon Dioxide with Cosolvents: Measurement, Prediction, and Applications in Separation Processes. Ind. Eng. Chem. Res. 2020, 59, 12319–12330. [Google Scholar] [CrossRef]
  74. Francisco, M.; van den Bruinhorst, A.; Kroon, M.C. Low-Transition-Temperature Mixtures (LTTMs): A New Generation of Designer Solvents. Angew. Chem. Int. Ed. 2013, 52, 3074–3085. [Google Scholar] [CrossRef] [PubMed]
  75. Smith, E.L.; Abbott, A.P.; Ryder, K.S. Deep Eutectic Solvents (DESs) and Their Applications. Chem. Rev. 2014, 114, 11060–11082. [Google Scholar] [CrossRef]
  76. Moshikur, R.M.; Goto, M. Ionic Liquids as Active Pharmaceutical Ingredients (APIs). In Application of Ionic Liquids in Drug Delivery; Goto, M., Moniruzzaman, M., Eds.; Springer Singapore: Singapore, 2021; pp. 13–33. [Google Scholar] [CrossRef]
  77. Md Moshikur, R.; Goto, M. Pharmaceutical Applications of Ionic Liquids: A Personal Account. Chem. Rec. 2023, 23, e202300026. [Google Scholar] [CrossRef]
  78. Wu, X.; Zhu, Q.; Chen, Z.; Wu, W.; Lu, Y.; Qi, J. Ionic liquids as a useful tool for tailoring active pharmaceutical ingredients. J. Control. Release 2021, 338, 268–283. [Google Scholar] [CrossRef]
  79. Moshikur, R.M.; Carrier, R.L.; Moniruzzaman, M.; Goto, M. Recent Advances in Biocompatible Ionic Liquids in Drug Formulation and Delivery. Pharmaceutics 2023, 15, 1179. [Google Scholar] [CrossRef] [PubMed]
  80. Jessop, P.G.; Mercer, S.M.; Heldebrant, D.J. CO2-triggered switchable solvents, surfactants, and other materials. Energy Environ. Sci. 2012, 5, 7240–7253. [Google Scholar] [CrossRef]
  81. Cunha, I.T.; McKeeman, M.; Ramezani, M.; Hayashi-Mehedy, K.; Lloyd-Smith, A.; Bravi, M.; Jessop, P.G. Amine-free CO2-switchable hydrophilicity solvents and their application in extractions and polymer recycling. Green Chem. 2022, 24, 3704–3716. [Google Scholar] [CrossRef]
  82. Mercer, S.M.; Jessop, P.G. “Switchable water”: Aqueous solutions of switchable ionic strength. ChemSusChem 2010, 3, 467–470. [Google Scholar] [CrossRef] [PubMed]
  83. Liberato, V.S.; Ferreira, T.F.; MacDonald, A.R.; Dias Ribeiro, B.; Zarur Coelho, M.A.; Jessop, P.G. A CO2-responsive method for separating hydrophilic organic molecules from aqueous solutions: Solvent-assisted switchable water. Green Chem. 2023, 25, 4705–4712. [Google Scholar] [CrossRef]
  84. Cunha, I.T.; Yang, H.; Jessop, P.G. High pressure switchable water: An alternative method for separating organic products from water. Green Chem. 2021, 23, 3996–4007. [Google Scholar] [CrossRef]
  85. Phan, L.; Jessop, P.G. Switching the hydrophilicity of a solute. Green Chem. 2009, 11, 307–330. [Google Scholar] [CrossRef]
  86. Duereh, A.; Sato, Y.; Smith, R.L.; Inomata, H. Methodology for replacing dipolar aprotic solvents used in API processing with safe hydrogen-bond donor and acceptor solvent-pair mixtures. Org. Process Res. Dev. 2017, 21, 114–124. [Google Scholar] [CrossRef]
  87. Jouyban, A.; Acree, W.E. A single model to represent physico-chemical properties of liquid mixtures at various temperatures. J. Mol. Liq. 2021, 323, 115054. [Google Scholar] [CrossRef]
  88. Nazemieh, A.; Acree, W.E.; Jouyban, A. Further computations on physico-chemical properties of binary mixtures of p-cymene with α-pinene, limonene and citral. J. Mol. Liq. 2022, 350, 118211. [Google Scholar] [CrossRef]
  89. Lee, J.L.; Chong, G.H.; Kanno, A.; Ota, M.; Guo, H.; Smith, R.L. Local composition-regular solution theory for analysis of pharmaceutical solubility in mixed-solvents. J. Mol. Liq. 2024, 397, 124012. [Google Scholar] [CrossRef]
  90. Duereh, A.; Guo, H.; Honma, T.; Hiraga, Y.; Sato, Y.; Lee Smith, R.; Inomata, H. Solvent Polarity of Cyclic Ketone (Cyclopentanone, Cyclohexanone): Alcohol (Methanol, Ethanol) Renewable Mixed-Solvent Systems for Applications in Pharmaceutical and Chemical Processing. Ind. Eng. Chem. Res. 2018, 57, 7331–7344. [Google Scholar] [CrossRef]
  91. Duereh, A.; Sato, Y.; Smith, R.L.; Inomata, H. Analysis of the Cybotactic Region of Two Renewable Lactone-Water Mixed-Solvent Systems that Exhibit Synergistic Kamlet-Taft Basicity. J. Phys. Chem. B 2016, 120, 4467–4481. [Google Scholar] [CrossRef] [PubMed]
  92. Abbasi, M.; Vaez-Gharamaleki, J.; Fazeli-Bakhtiyari, R.; Martinez, F.; Jouyban, A. Prediction of deferiprone solubility in some non-aqueous binary solvent mixtures at various temperatures. J. Mol. Liq. 2015, 203, 16–19. [Google Scholar] [CrossRef]
  93. Acree, W.E. IUPAC-NIST Solubility Data Series. 102. Solubility of Nonsteroidal Anti-inflammatory Drugs (NSAIDs) in Neat Organic Solvents and Organic Solvent Mixtures. J. Phys. Chem. Ref. Data 2014, 43, 023102. [Google Scholar] [CrossRef]
  94. Akay, S.; Kayan, B.; Martínez, F. Solubility of fluconazole in (ethanol + water) mixtures: Determination, correlation, dissolution thermodynamics and preferential solvation. J. Mol. Liq. 2021, 333, 115987. [Google Scholar] [CrossRef]
  95. Akay, S.; Kayan, B.; Peña, M.Á.; Jouyban, A.; Martínez, F. Solubility of Salicylic Acid in Some (Ethanol + Water) Mixtures at Different Temperatures: Determination, Correlation, Thermodynamics and Preferential Solvation. Int. J. Thermophys. 2023, 44, 121. [Google Scholar] [CrossRef]
  96. Ali, H.S.M.; York, P.; Blagden, N.; Soltanpour, S.; Acree, W.E.; Jouyban, A. Solubility of Budesonide, Hydrocortisone, and Prednisolone in Ethanol + Water Mixtures at 298.2 K. J. Chem. Eng. Data 2009, 55, 578–582. [Google Scholar] [CrossRef]
  97. Almandoz, M.C.; Sancho, M.I.; Blanco, S.E. Spectroscopic and DFT study of solvent effects on the electronic absorption spectra of sulfamethoxazole in neat and binary solvent mixtures. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 118, 112–119. [Google Scholar] [CrossRef]
  98. Alsubaie, M.; Aljohani, M.; Erxleben, A.; McArdle, P. Cocrystal Forms of the BCS Class IV Drug Sulfamethoxazole. Cryst. Growth Des. 2018, 18, 3902–3912. [Google Scholar] [CrossRef]
  99. Alvani-Alamdari, S.; Rezaei, H.; Rahimpour, E.; Hemmati, S.; Martinez, F.; Barzegar-Jalali, M.; Jouyban, A. Mesalazine solubility in the binary mixtures of ethanol and water at various temperatures. Phys. Chem. Liq. 2019, 59, 12–25. [Google Scholar] [CrossRef]
  100. Aniya, V.; De, D.; Mohammed, A.M.; Thella, P.K.; Satyavathi, B. Measurement and modeling of solubility of para-tert-butylbenzoic acid in pure and mixed organic solvents at different temperatures. J. Chem. Eng. Data 2017, 62, 1411–1421. [Google Scholar] [CrossRef]
  101. Anwer, M.K.; Mohammad, M.; Fatima, F.; Alshahrani, S.M.; Aldawsari, M.F.; Alalaiwe, A.; Al-Shdefat, R.; Shakeel, F. Solubility, solution thermodynamics and molecular interactions of osimertinib in some pharmaceutically useful solvents. J. Mol. Liq. 2019, 284, 53–58. [Google Scholar] [CrossRef]
  102. Assis, G.P.; Garcia, R.H.L.; Derenzo, S.; Bernardo, A. Solid-liquid equilibrium of paracetamol in water-ethanol and water-propylene glycol mixtures. J. Mol. Liq. 2021, 323, 114617. [Google Scholar] [CrossRef]
  103. Aydi, A.; Claumann, C.A.; Wüst Zibetti, A.; Abderrabba, M. Differential Scanning Calorimetry Data and Solubility of Rosmarinic Acid in Different Pure Solvents and in Binary Mixtures (Methyl Acetate + Water) and (Ethyl Acetate + Water) from 293.2 to 313.2 K. J. Chem. Eng. Data 2016, 61, 3718–3723. [Google Scholar] [CrossRef]
  104. Banerjee, D.; Laha, A.K.; Bagchi, S. Preferential solvation in mixed binary solvent. J. Chem. Soc. Faraday Trans. 1995, 91, 631. [Google Scholar] [CrossRef]
  105. Barzegar-Jalali, M.; Mirheydari, S.N.; Rahimpour, E.; Shekaari, H.; Martinez, F.; Jouyban, A. Experimental determination and correlation of bosentan solubility in (PEG 200+ water) mixtures at T = (293.15–313.15) K. Phys. Chem. Liq. 2018, 57, 504–515. [Google Scholar] [CrossRef]
  106. Bernal-García, J.M.; Guzmán-López, A.; Cabrales-Torres, A.; Estrada-Baltazar, A.; Iglesias-Silva, G.A. Densities and viscosities of (N,N-dimethylformamide+ water) at atmospheric pressure from (283.15 to 353.15) K. J. Chem. Eng. Data 2008, 53, 1024–1027. [Google Scholar] [CrossRef]
  107. Bhesaniya, K.; Nandha, K.; Baluja, S. Thermodynamics of Fluconazole Solubility in Various Solvents at Different Temperatures. J. Chem. Eng. Data 2014, 59, 649–652. [Google Scholar] [CrossRef]
  108. Blanco-Márquez, J.H.; Ortiz, C.P.; Cerquera, N.E.; Martínez, F.; Jouyban, A.; Delgado, D.R. Thermodynamic analysis of the solubility and preferential solvation of sulfamerazine in (acetonitrile + water) cosolvent mixtures at different temperatures. J. Mol. Liq. 2019, 293, 111507. [Google Scholar] [CrossRef]
  109. Blokhina, S.V.; Sharapova, A.V.; Ol’khovich, M.V.; Levshin, I.B.; Perlovich, G.L. Solid–liquid phase equilibrium and thermodynamic analysis of novel thiazolidine-2,4-dione derivative in different solvents. J. Mol. Liq. 2021, 326, 115273. [Google Scholar] [CrossRef]
  110. Bosch, E.; Rived, F.; Rosés, M. Solute–solvent and solvent–solvent interactions in binary solvent mixtures. Part 4. Preferential solvation of solvatochromic indicators in mixtures of 2-methylpropan-2-ol with hexane, benzene, propan-2-ol, ethanol and methanol. J. Chem. Soc. Perkin Trans. 1996, 2, 2177–2184. [Google Scholar] [CrossRef]
  111. Bosch, E.; Rosés, M.; Herodes, K.; Koppel, I.; Leito, I.; Koppel, I.; Taal, V. Solute-solvent and solvent-solvent interactions in binary solvent mixtures. 2. Effect of temperature on the ET(30) polarity parameter of dipolar hydrogen bond acceptor-hydrogen bond donor mixtures. J. Phys. Org. Chem. 1996, 9, 403–410. [Google Scholar] [CrossRef]
  112. Calvo, B.; Cepeda, E.A. Solubilities of Stearic Acid in Organic Solvents and in Azeotropic Solvent Mixtures. J. Chem. Eng. Data 2008, 53, 628–633. [Google Scholar] [CrossRef]
  113. Calvo, B.; Collado, I.; Cepeda, E.A. Solubilities of Palmitic Acid in Pure Solvents and Its Mixtures. J. Chem. Eng. Data 2008, 54, 64–68. [Google Scholar] [CrossRef]
  114. Cañadas, R.; González-Miquel, M.; González, E.J.; Díaz, I.; Rodríguez, M. Evaluation of bio-based solvents for phenolic acids extraction from aqueous matrices. J. Mol. Liq. 2021, 338, 116930. [Google Scholar] [CrossRef]
  115. Carmen Grande, M.d.; Juliá, J.A.; García, M.; Marschoff, C.M. On the density and viscosity of (water+dimethylsulphoxide) binary mixtures. J. Chem. Thermodyn. 2007, 39, 1049–1056. [Google Scholar] [CrossRef]
  116. Castro, G.T.; Filippa, M.A.; Peralta, C.M.; Davin, M.V.; Almandoz, M.C.; Gasull, E.I. Solubility and Preferential Solvation of Piroxicam in Neat Solvents and Binary Systems. Z. Für Phys. Chem. 2017, 232, 257–280. [Google Scholar] [CrossRef]
  117. Chen, F.; Zhao, M.; Feng, L.; Ren, B. Measurement and Correlation for Solubility of Diosgenin in Some Mixed Solvents. Chin. J. Chem. Eng. 2014, 22, 170–176. [Google Scholar] [CrossRef]
  118. Chen, F.-X.; Zhao, M.-R.; Ren, B.-Z.; Zhou, C.-R.; Peng, F.-F. Solubility of diosgenin in different solvents. J. Chem. Thermodyn. 2012, 47, 341–346. [Google Scholar] [CrossRef]
  119. Chen, S.; Liu, Q.; Dou, H.; Zhang, L.; Pei, L.; Huang, R.; Shu, G.; Yuan, Z.; Lin, J.; Zhang, W.; et al. Solubility and dissolution thermodynamic properties of Mequindox in binary solvent mixtures. J. Mol. Liq. 2020, 303, 112619. [Google Scholar] [CrossRef]
  120. Chen, Y.; Xu, X.; Xie, L. Thermodynamic parameters on corresponding solid-liquid equilibrium of hydroxyapatite in pure and mixture organic solvents. J. Mol. Liq. 2017, 229, 189–197. [Google Scholar] [CrossRef]
  121. Chen, Z.; Zhai, J.; Liu, X.; Mao, S.; Zhang, L.; Rohani, S.; Lu, J. Solubility measurement and correlation of the form A of ibrutinib in organic solvents from 278.15 to 323.15 K. J. Chem. Thermodyn. 2016, 103, 342–348. [Google Scholar] [CrossRef]
  122. Cui, Z.; Yao, L.; Ye, J.; Wang, Z.; Hu, Y. Solubility measurement and thermodynamic modelling of curcumin in twelve pure solvents and three binary solvents at different temperature (T = 278.15–323.15 K). J. Mol. Liq. 2021, 338, 116795. [Google Scholar] [CrossRef]
  123. Cysewski, P.; Jeliński, T.; Przybyłek, M.; Nowak, W.; Olczak, M. Solubility Characteristics of Acetaminophen and Phenacetin in Binary Mixtures of Aqueous Organic Solvents: Experimental and Deep Machine Learning Screening of Green Dissolution Media. Pharmaceutics 2022, 14, 2828. [Google Scholar] [CrossRef] [PubMed]
  124. De la Rosa, M.V.G.; Santiago, R.; Romero, J.M.; Duconge, J.; Monbaliu, J.-C.; López-Mejías, V.; Stelzer, T. Solubility Determination and Correlation of Warfarin Sodium 2-Propanol Solvate in Pure, Binary, and Ternary Solvent Mixtures. J. Chem. Eng. Data 2019, 64, 1399–1413. [Google Scholar] [CrossRef] [PubMed]
  125. del Mar Muñoz, M.; Delgado, D.R.; Peña, M.Á.; Jouyban, A.; Martínez, F. Solubility and preferential solvation of sulfadiazine, sulfamerazine and sulfamethazine in propylene glycol + water mixtures at 298.15K. J. Mol. Liq. 2015, 204, 132–136. [Google Scholar] [CrossRef]
  126. Delgado, D.R.; Martínez, F. Solubility and solution thermodynamics of sulfamerazine and sulfamethazine in some ethanol+water mixtures. Fluid Phase Equilibria 2013, 360, 88–96. [Google Scholar] [CrossRef]
  127. Delgado, D.R.; Martínez, F. Preferential solvation of sulfadiazine, sulfamerazine and sulfamethazine in ethanol+water solvent mixtures according to the IKBI method. J. Mol. Liq. 2014, 193, 152–159. [Google Scholar] [CrossRef]
  128. Dizechi, M.; Marschall, E. Viscosity of some binary and ternary liquid mixtures. J. Chem. Eng. Data 1982, 27, 358–363. [Google Scholar] [CrossRef]
  129. Domańska, U.; Pobudkowska, A.; Pelczarska, A.; Winiarska-Tusznio, M.; Gierycz, P. Solubility and pKa of select pharmaceuticals in water, ethanol, and 1-octanol. J. Chem. Thermodyn. 2010, 42, 1465–1472. [Google Scholar] [CrossRef]
  130. Dong, Q.; Yu, S.; Wang, X.; Ding, S.; Li, E.; Cai, Y.; Xue, F. Solubility Measurement and Correlation of Itraconazole Hydroxy Isobutyltriazolone in Four Kinds of Binary Solvent Mixtures with Temperature from 283.15 to 323.15 K. ACS Omega 2023, 8, 39390–39400. [Google Scholar] [CrossRef] [PubMed]
  131. Elizalde-Solis, O.; Arenas-Quevedo, M.G.; Verónico-Sánchez, F.J.; García-Morales, R.; Zúñiga-Moreno, A. Solubilities of Binary Systems α-Tocopherol + Capsaicin and α-Tocopherol + Palmitic Acid in Supercritical Carbon Dioxide. J. Chem. Eng. Data 2019, 64, 1948–1955. [Google Scholar] [CrossRef]
  132. Fakhree, M.A.A.; Ahmadian, S.; Panahi-Azar, V.; Acree, W.E.; Jouyban, A. Solubility of 2-Hydroxybenzoic Acid in Water, 1-Propanol, 2-Propanol, and 2-Propanone at (298.2 to 338.2) K and Their Aqueous Binary Mixtures at 298.2 K. J. Chem. Eng. Data 2012, 57, 3303–3307. [Google Scholar] [CrossRef]
  133. Filippa, M.A.; Gasull, E.I. Ibuprofen solubility in pure organic solvents and aqueous mixtures of cosolvents: Interactions and thermodynamic parameters relating to the solvation process. Fluid Phase Equilibria 2013, 354, 185–190. [Google Scholar] [CrossRef]
  134. Gheitasi, N.; Nazari, A.H.; Haghtalab, A. Thermodynamic Modeling and Solubility Measurement of Cetirizine Hydrochloride and Deferiprone in Pure Solvents of Acetonitrile, Ethanol, Acetic Acid, Sulfolane, and Ethyl Acetate and Their Mixtures. J. Chem. Eng. Data 2019, 64, 5486–5496. [Google Scholar] [CrossRef]
  135. Gonçalves Bonassoli, A.B.; Oliveira, G.; Bordón Sosa, F.H.; Rolemberg, M.P.; Mota, M.A.; Basso, R.C.; Igarashi-Mafra, L.; Mafra, M.R. Solubility measurement of lauric, palmitic, and stearic acids in ethanol, n-propanol, and 2-propanol using differential scanning calorimetry. J. Chem. Eng. Data 2019, 64, 2084–2092. [Google Scholar] [CrossRef]
  136. Guo, S.; Yang, W.; Hu, Y.; Wang, K.; Luan, Y. Measurement and Correlation of the Solubility of N-Acetylglycine in Different Solvents at Temperatures from 278.15 to 319.15 K. J. Solut. Chem. 2013, 42, 1879–1887. [Google Scholar] [CrossRef]
  137. Guo, Y.; He, H.; Huang, H.; Qiu, J.; Han, J.; Hu, S.; Liu, H.; Zhao, Y.; Wang, P. Solubility determination and thermodynamic modeling of n-acetylglycine in different solvent systems. J. Chem. Eng. Data 2021, 66, 1344–1355. [Google Scholar] [CrossRef]
  138. Gusain, K.; Garg, S.; Kumar, R. Solubility Prediction of Pharmaceutical Compounds in Pure Solvent by Different Correlations and Thermodynamic Models. SSRN Electron. J. 2020. [Google Scholar] [CrossRef]
  139. Ha, E.-S.; Lee, S.-K.; Jeong, J.-S.; Sim, W.-Y.; Yang, J.-I.; Kim, J.-S.; Kim, M.-S. Solvent effect and solubility modeling of rebamipide in twelve solvents at different temperatures. J. Mol. Liq. 2019, 288, 111041. [Google Scholar] [CrossRef]
  140. Hatefi, A.; Jouyban, A.; Mohammadian, E.; Acree, W.E.; Rahimpour, E. Prediction of paracetamol solubility in cosolvency systems at different temperatures. J. Mol. Liq. 2019, 273, 282–291. [Google Scholar] [CrossRef]
  141. He, Q.; Zheng, M.; Zhao, H. Baicalin solubility in aqueous co-solvent mixtures of methanol, ethanol, isopropanol and n-propanol revisited: Solvent–solvent and solvent–solute interactions and IKBI preferential solvation analysis. Phys. Chem. Liq. 2019, 58, 820–832. [Google Scholar] [CrossRef]
  142. Hellstén, S.; Qu, H.; Louhi-Kultanen, M. Screening of Binary Solvent Mixtures and Solvate Formation of Indomethacin. Chem. Eng. Technol. 2011, 34, 1667–1674. [Google Scholar] [CrossRef]
  143. Heryanto, R.; Hasan, M.; Abdullah, E.C.; Kumoro, A.C. Solubility of Stearic Acid in Various Organic Solvents and Its Prediction using Non-ideal Solution Models. ScienceAsia 2007, 33, 469–472. Available online: https://www.scienceasia.org/2007.33.n4/v33_469_472.pdf (accessed on 30 December 2023).
  144. Hu, W.; Shang, Z.; Wei, N.; Hou, B.; Gong, J.; Wang, Y. Solubility of benorilate in twelve monosolvents: Determination, correlation and COSMO-RS analysis. J. Chem. Thermodyn. 2021, 152, 106272. [Google Scholar] [CrossRef]
  145. Hu, X.; Gong, Y.; Cao, Z.; Huang, Z.; Sha, J.; Li, Y.; Li, T.; Ren, B. Solubility, Hansen solubility parameter and thermodynamic properties of etodolac in twelve organic pure solvents at different temperatures. J. Mol. Liq. 2020, 316, 113779. [Google Scholar] [CrossRef]
  146. Hu, X.; Tian, Y.; Cao, Z.; Sha, J.; Huang, Z.; Li, Y.; Li, T.; Ren, B. Solubility measurement, Hansen solubility parameter and thermodynamic modeling of etodolac in four binary solvents from 278.15 K to 323.15 K. J. Mol. Liq. 2020, 318, 114155. [Google Scholar] [CrossRef]
  147. Imran, S.; Hossain, A.; Mahali, K.; Guin, P.S.; Datta, A.; Roy, S. Solubility and peculiar thermodynamical behaviour of 2-aminobenzoic acid in aqueous binary solvent mixtures at 288.15 to 308.15 K. J. Mol. Liq. 2020, 302, 112566. [Google Scholar] [CrossRef]
  148. Ivanov, E.V.; Batov, D.V. Enthalpy-related parameters of interaction of simplest α-amino acids with the pharmaceutical mebicar (N-tetramethylglycoluril) in water at 298.15 K. J. Chem. Thermodyn. 2019, 128, 159–163. [Google Scholar] [CrossRef]
  149. Jia, L.; Yang, J.; Cui, P.; Wu, D.; Wang, S.; Hou, B.; Zhou, L.; Yin, Q. Uncovering solubility behavior of Prednisolone form II in eleven pure solvents by thermodynamic analysis and molecular simulation. J. Mol. Liq. 2021, 342, 117376. [Google Scholar] [CrossRef]
  150. Jiménez Cruz, J.M.; Vlaar, C.P.; López-Mejías, V.; Stelzer, T. Solubility Measurements and Correlation of MBQ-167 in Neat and Binary Solvent Mixtures. J. Chem. Eng. Data 2021, 66, 832–839. [Google Scholar] [CrossRef]
  151. Jiménez, D.M.; Cárdenas, Z.J.; Martínez, F. Solubility and solution thermodynamics of sulfadiazine in polyethylene glycol 400 + water mixtures. J. Mol. Liq. 2016, 216, 239–245. [Google Scholar] [CrossRef]
  152. Jouyban, A.; Acree, W.E.; Martínez, F. Dissolution thermodynamics and preferential solvation of ketoconazole in some {ethanol (1) + water (2)} mixtures. J. Mol. Liq. 2020, 313, 113579. [Google Scholar] [CrossRef]
  153. Jouyban, A.; Mazaher Haji Agha, E.; Rahimpour, E.; Acree, W.E., Jr. Further computation and some comments on “Stearic acid solubility in mixed solvents of (water + ethanol) and (ethanol + ethyl acetate): Experimental data and comparison among different thermodynamic models”. J. Mol. Liq. 2020, 310, 113228. [Google Scholar] [CrossRef]
  154. Jouyban, K.; Mazaher Haji Agha, E.; Hemmati, S.; Martinez, F.; Kuentz, M.; Jouyban, A. Solubility of 5-aminosalicylic acid in N-methyl-2-pyrrolidone + water mixtures at various temperatures. J. Mol. Liq. 2020, 310, 113143. [Google Scholar] [CrossRef]
  155. Jouyban-Gharamaleki, V.; Jouyban, A.; Kuentz, M.; Hemmati, S.; Martinez, F.; Rahimpour, E. A laser monitoring technique for determination of mesalazine solubility in propylene glycol and ethanol mixtures at various temperatures. J. Mol. Liq. 2020, 304, 112714. [Google Scholar] [CrossRef]
  156. Jouyban-Gharamaleki, V.; Jouyban, A.; Martinez, F.; Zhao, H.; Rahimpour, E. A laser monitoring technique for solubility study of ketoconazole in propylene glycol and 2-propanol mixtures at various temperatures. J. Mol. Liq. 2020, 320, 114444. [Google Scholar] [CrossRef]
  157. Kalam, M.A.; Alshehri, S.; Alshamsan, A.; Haque, A.; Shakeel, F. Solid liquid equilibrium of an antifungal drug itraconazole in different neat solvents: Determination and correlation. J. Mol. Liq. 2017, 234, 81–87. [Google Scholar] [CrossRef]
  158. Kandi, S.; Charles, A.L. Measurement, correlation, and thermodynamic properties for solubilities of bioactive compound (−)-epicatechin in different pure solvents at 298.15 K to 338.15 K. J. Mol. Liq. 2018, 264, 269–274. [Google Scholar] [CrossRef]
  159. Karpiuk, I.; Wilczura-Wachnik, H.; Myśliński, A. α-Tocopherol/AOT/alkane/water system. J. Therm. Anal. Calorim. 2017, 131, 2885–2892. [Google Scholar] [CrossRef]
  160. Khajir, S.; Shayanfar, A.; Acree, W.E.; Jouyban, A. Effects of N-methylpyrrolidone and temperature on phenytoin solubility. J. Mol. Liq. 2019, 285, 58–61. [Google Scholar] [CrossRef]
  161. Kuhs, M.; Svärd, M.; Rasmuson, Å.C. Thermodynamics of fenoxycarb in solution. J. Chem. Thermodyn. 2013, 66, 50–58. [Google Scholar] [CrossRef]
  162. Kumari, A.; Kadakanchi, S.; Tangirala, R.; Thella, P.K.; Satyavathi, B. Measurement and modeling of solid–liquid equilibrium of para-tert-butylbenzoic acid in acetic acid/methanol+ water and acetic acid+ para-tert-butyltoluene binary systems at various temperatures. J. Chem. Eng. Data 2016, 62, 87–95. [Google Scholar] [CrossRef]
  163. Lange, L.; Heisel, S.; Sadowski, G. Predicting the Solubility of Pharmaceutical Cocrystals in Solvent/Anti-Solvent Mixtures. Molecules 2016, 21, 593. [Google Scholar] [CrossRef] [PubMed]
  164. Lee, S.-K.; Sim, W.-Y.; Ha, E.-S.; Park, H.; Kim, J.-S.; Jeong, J.-S.; Kim, M.-S. Solubility of bisacodyl in fourteen mono solvents and N-methyl-2-pyrrolidone + water mixed solvents at different temperatures, and its application for nanosuspension formation using liquid antisolvent precipitation. J. Mol. Liq. 2020, 310, 113264. [Google Scholar] [CrossRef]
  165. Li, A.; Si, Z.; Yan, Y.; Zhang, X. Solubility and thermodynamic properties of hydrate lenalidomide in phosphoric acid solution. J. Mol. Liq. 2021, 330, 115446. [Google Scholar] [CrossRef]
  166. Li, M.; Liu, S.; Li, S.; Yang, Y.; Cui, Y.; Gong, J. Determination and Correlation of Dipyridamole p-Toluene Sulfonate Solubility in Seven Alcohol Solvents and Three Binary Solvents. J. Chem. Eng. Data 2017, 63, 208–216. [Google Scholar] [CrossRef]
  167. Li, R.; Fu, L.; Zhang, J.; Wang, W.; Chen, X.; Zhao, J.; Han, D. Solid-liquid equilibrium and thermodynamic properties of dipyridamole form II in pure solvents and mixture of (N-methyl pyrrolidone + isopropanol). J. Chem. Thermodyn. 2020, 142, 105981. [Google Scholar] [CrossRef]
  168. Li, R.; Jin, Y.; Yu, B.; Xu, Q.; Chen, X.; Han, D. Solubility determination and thermodynamic properties calculation of macitentan in mixtures of ethyl acetate and alcohols. J. Chem. Thermodyn. 2021, 156, 106344. [Google Scholar] [CrossRef]
  169. Li, R.; Liu, L.; Khan, A.; Li, C.; He, Z.; Zhao, J.; Han, D. Effect of Cosolvents on the Solubility of Lenalidomide and Thermodynamic Model Correlation of Data. J. Chem. Eng. Data 2019, 64, 4272–4279. [Google Scholar] [CrossRef]
  170. Li, R.; Yan, H.; Wang, Z.; Gong, J. Correlation of Solubility and Prediction of the Mixing Properties of Ginsenoside Compound K in Various Solvents. Ind. Eng. Chem. Res. 2012, 51, 8141–8148. [Google Scholar] [CrossRef]
  171. Li, R.; Yin, X.; Jin, Y.; Chen, X.; Zhao, B.; Wang, W.; Zhong, S.; Han, D. The solubility profile and dissolution thermodynamic properties of minocycline hydrochloride in some pure and mixed solvents at several temperatures. J. Chem. Thermodyn. 2021, 157, 106399. [Google Scholar] [CrossRef]
  172. Li, R.; Zhao, B.; Chen, X.; Zhang, J.; Liu, Z.; Zhu, X.; Han, D. Solubility and apparent thermodynamic analysis of pomalidomide in (acetone + ethanol/isopropanol) and (ethyl acetate + ethanol/isopropanol) and its correlation with thermodynamic model. J. Chem. Thermodyn. 2021, 154, 106345. [Google Scholar] [CrossRef]
  173. Li, W.; Yuan, J.; Wang, X.; Shi, W.; Zhao, H.; Xing, R.; Jouyban, A.; Acree, W.E. Bifonazole dissolved in numerous aqueous alcohol mixtures: Solvent effect, enthalpy–entropy compensation, extended Hildebrand solubility parameter approach and preferential solvation. J. Mol. Liq. 2021, 338, 116671. [Google Scholar] [CrossRef]
  174. Li, X.; Du, C.; Cong, Y.; Zhao, H. Solubility determination and thermodynamic modelling of 3-amino-1,2,4-triazole in ten organic solvents from T = 283.15 K to T = 318.15 K and mixing properties of solutions. J. Chem. Thermodyn. 2017, 104, 189–200. [Google Scholar] [CrossRef]
  175. Li, X.; Ma, M.; Du, C.; Zhao, H. Solubility of cetilistat in neat solvents and preferential solvation in (acetone, isopropanol or acetonitrile) + water co-solvent mixtures. J. Mol. Liq. 2017, 242, 618–624. [Google Scholar] [CrossRef]
  176. Li, X.; Wang, M.; Du, C.; Cong, Y.; Zhao, H. Preferential solvation of rosmarinic acid in binary solvent mixtures of ethanol + water and methanol + water according to the inverse Kirkwood–Buff integrals method. J. Mol. Liq. 2017, 240, 56–64. [Google Scholar] [CrossRef]
  177. Li, Y.; Wang, Y.; Ning, Z.; Cui, J.; Wu, Q.; Wang, X. Solubilities of Adipic Acid and Succinic Acid in a Glutaric Acid + Acetone or n-Butanol Mixture. J. Chem. Eng. Data 2014, 59, 4062–4069. [Google Scholar] [CrossRef]
  178. Lin, L.; Zhao, K.; Yu, B.; Wang, H.; Chen, M.; Gong, J. Measurement and Correlation of Solubility of Cefathiamidine in Water + (Acetone, Ethanol, or 2-Propanol) from (278.15 to 308.15) K. J. Chem. Eng. Data 2015, 61, 412–419. [Google Scholar] [CrossRef]
  179. Liu, J.-Q.; Wang, Y.; Tang, H.; Wu, S.; Li, Y.-Y.; Zhang, L.-Y.; Bai, Q.-Y.; Liu, X. Experimental Measurements and Modeling of the Solubility of Aceclofenac in Six Pure Solvents from (293.35 to 338.25) K. J. Chem. Eng. Data 2014, 59, 1588–1592. [Google Scholar] [CrossRef]
  180. Liu, W.; Bao, Z.; Shen, Y.; Yao, T.; Bai, H.; Jin, X. Solubility measurement and thermodynamic modeling of carbamazepine (form III) in five pure solvents at various temperatures. Chin. J. Chem. Eng. 2021, 33, 231–235. [Google Scholar] [CrossRef]
  181. Liu, Y.; Wang, Y.; Liu, Y.; Xu, S.; Chen, M.; Du, S.; Gong, J. Solubility of L-histidine in different aqueous binary solvent mixtures from 283.15 K to 318.15 K with experimental measurement and thermodynamic modelling. J. Chem. Thermodyn. 2017, 105, 1–14. [Google Scholar] [CrossRef]
  182. Lou, Y.; Wang, Y.; Li, Y.; He, M.; Su, N.; Xu, R.; Meng, X.; Hou, B.; Xie, C. Thermodynamic equilibrium and cosolvency of florfenicol in binary solvent system. J. Mol. Liq. 2018, 251, 83–91. [Google Scholar] [CrossRef]
  183. Mabhoot, A.; Jouyban, A. Solubility of Sodium Phenytoin in Propylene Glycol + Water Mixtures in the Presence of B-Cyclodextrin. Pharm. Sci. 2015, 21, 152–156. [Google Scholar] [CrossRef]
  184. Mahali, K.; Guin, P.S.; Roy, S.; Dolui, B.K. Solubility and solute–solvent interaction phenomenon of succinic acid in aqueous ethanol mixtures. J. Mol. Liq. 2017, 229, 172–177. [Google Scholar] [CrossRef]
  185. Marcus, Y. The use of chemical probes for the characterization of solvent mixtures. Part 2. Aqueous mixtures. J. Chem. Soc. Perkin Trans. 1994, 2, 1751. [Google Scholar] [CrossRef]
  186. Marcus, Y. Use of chemical probes for the characterization of solvent mixtures. Part 1. Completely non-aqueous mixtures. J. Chem. Soc. Perkin Trans. 1994, 2, 1015. [Google Scholar] [CrossRef]
  187. Matsuda, H.; Kaburagi, K.; Matsumoto, S.; Kurihara, K.; Tochigi, K.; Tomono, K. Solubilities of Salicylic Acid in Pure Solvents and Binary Mixtures Containing Cosolvent. J. Chem. Eng. Data 2008, 54, 480–484. [Google Scholar] [CrossRef]
  188. McHedlov-Petrosyan, N.O. Book review: Christian Reichardt and Thomas Welton, Solvents and Solvent Effects in Organic Chemistry (Fourth Edition, Updated and Enlarged; Wiley-VCH Verlag & Co. KGaA, Weinheim, 2011; 718 p. Hardcover). Russ. J. Phys. Chem. A 2011, 85, 1482. [Google Scholar] [CrossRef]
  189. Mealey, D.; Svärd, M.; Rasmuson, Å.C. Thermodynamics of risperidone and solubility in pure organic solvents. Fluid Phase Equilibria 2014, 375, 73–79. [Google Scholar] [CrossRef]
  190. Mirmehrabi, M.; Rohani, S. Measurement and Prediction of the Solubility of Stearic Acid Polymorphs by the UNIQUAC Equation. Can. J. Chem. Eng. 2004, 82, 335–342. [Google Scholar] [CrossRef]
  191. Mo, F.; Ma, J.; Zhang, P.; Zhang, D.; Fan, H.; Yang, X.; Zhi, L.; Zhang, J. Solubility and thermodynamic properties of baicalein in water and ethanol mixtures from 283.15 to 328.15 K. Chem. Eng. Commun. 2019, 208, 183–196. [Google Scholar] [CrossRef]
  192. Mohamadian, E.; Hamidi, S.; Martínez, F.; Jouyban, A. Solubility prediction of deferiprone in N-methyl-2-pyrrolidone+ ethanol mixtures at various temperatures using a minimum number of experimental data. Phys. Chem. Liq. 2017, 55, 805–816. [Google Scholar] [CrossRef]
  193. Mohammadian, E.; Jouyban, A.; Barzegar-Jalali, M.; Acree, W.E.; Rahimpour, E. Solubilization of naproxen: Experimental data and computational tools. J. Mol. Liq. 2019, 288, 110985. [Google Scholar] [CrossRef]
  194. Mohammadzade, M.; Barzegar-Jalali, M.; Jouyban, A. Solubility of naproxen in 2-propanol+water mixtures at various temperatures. J. Mol. Liq. 2015, 206, 110–113. [Google Scholar] [CrossRef]
  195. Moodley, K.; Rarey, J.; Ramjugernath, D. Experimental solubility of diosgenin and estriol in various solvents between T = (293.2–328.2)K. J. Chem. Thermodyn. 2017, 106, 199–207. [Google Scholar] [CrossRef]
  196. Moodley, K.; Rarey, J.; Ramjugernath, D. Experimental solubility data for prednisolone and hydrocortisone in various solvents between (293.2 and 328.2) K by employing combined DTA/TGA. J. Mol. Liq. 2017, 240, 303–312. [Google Scholar] [CrossRef]
  197. Mora, C.P.; Martínez, F. Solubility of naproxen in several organic solvents at different temperatures. Fluid Phase Equilibria 2007, 255, 70–77. [Google Scholar] [CrossRef]
  198. Moradi, M.; Mazaher Haji Agha, E.; Hemmati, S.; Martinez, F.; Kuentz, M.; Jouyban, A. Solubility of 5-aminosalicylic acid in {N-methyl-2-pyrrolidone + ethanol} mixtures at T = (293.2 to 313.2) K. J. Mol. Liq. 2020, 306, 112774. [Google Scholar] [CrossRef]
  199. Ning, L.; Gong, X.; Li, P.; Chen, X.; Wang, H.; Xu, J. Measurement and correlation of the solubility of estradiol and estradiol-urea co-crystal in fourteen pure solvents at temperatures from 273.15 K to 318.15 K. J. Mol. Liq. 2020, 304, 112599. [Google Scholar] [CrossRef]
  200. Noda, K.; Aono, Y.; Ishida, K. Viscosity and Density of Ethanol-Acetic Acid-Water Mixtures. Kagaku Kogaku Ronbunshu 1983, 9, 237–240. [Google Scholar] [CrossRef]
  201. Noda, K.; Ohashi, M.; Ishida, K. Viscosities and densities at 298.15 K for mixtures of methanol, acetone, and water. J. Chem. Eng. Data 1982, 27, 326–328. [Google Scholar] [CrossRef]
  202. Noubigh, A. Stearic acid solubility in mixed solvents of (water + ethanol) and (ethanol + ethyl acetate): Experimental data and comparison among different thermodynamic models. J. Mol. Liq. 2019, 296, 112101. [Google Scholar] [CrossRef]
  203. Oliveira, G.; Bonassoli, A.B.G.; Rolemberg, M.P.; Mota, M.A.; Basso, R.C.; Soares, R.d.P.; Igarashi-Mafra, L.; Mafra, M.R. Water Effect on Solubilities of Lauric and Palmitic Acids in Ethanol and 2-Propanol Determined by Differential Scanning Calorimetry. J. Chem. Eng. Data 2021, 66, 2366–2373. [Google Scholar] [CrossRef]
  204. Ortiz, C.P.; Cardenas-Torres, R.E.; Martínez, F.; Delgado, D.R. Solubility of Sulfamethazine in the Binary Mixture of Acetonitrile + Methanol from 278.15 to 318.15 K: Measurement, Dissolution Thermodynamics, Preferential Solvation, and Correlation. Molecules 2021, 26, 7588. [Google Scholar] [CrossRef]
  205. Osorio, I.P.; Martínez, F.; Peña, M.A.; Jouyban, A.; Acree, W.E. Solubility, dissolution thermodynamics and preferential solvation of sulfadiazine in (N-methyl-2-pyrrolidone + water) mixtures. J. Mol. Liq. 2021, 330, 115693. [Google Scholar] [CrossRef]
  206. Pabba, S.; Kumari, A.; Ravuri, M.G.; Thella, P.K.; Satyavathi, B.; Shah, K.; Kundu, S.; Bhargava, S.K. Experimental determination and modelling of the co-solvent and antisolvent behaviour of binary systems on the dissolution of pharma drug; L-aspartic acid and thermodynamic correlations. J. Mol. Liq. 2020, 314, 113657. [Google Scholar] [CrossRef]
  207. Pacheco, D.P.; Martínez, F. Thermodynamic analysis of the solubility of naproxen in ethanol + water cosolvent mixtures. Phys. Chem. Liq. 2007, 45, 581–595. [Google Scholar] [CrossRef]
  208. Padervand, M.; Naseri, S.; Boroujeni, H.C. Preferential solvation of pomalidomide, an anticancer compound, in some binary mixed solvents at 298.15 K. Chin. J. Chem. Eng. 2020, 28, 2626–2633. [Google Scholar] [CrossRef]
  209. Pasham, F.; Jabbari, M.; Farajtabar, A. Solvatochromic Measurement of KAT Parameters and Modeling Preferential Solvation in Green Potential Binary Mixtures of N-Formylmorpholine with Water, Alcohols, and Ethyl Acetate. J. Chem. Eng. Data 2020, 65, 5458–5466. [Google Scholar] [CrossRef]
  210. Patel, A.; Vaghasiya, A.; Gajera, R.; Baluja, S. Solubility of 5-Amino Salicylic Acid in Different Solvents at Various Temperatures. J. Chem. Eng. Data 2010, 55, 1453–1455. [Google Scholar] [CrossRef]
  211. Przybyłek, M.; Miernicka, A.; Nowak, M.; Cysewski, P. New Screening Protocol for Effective Green Solvents Selection of Benzamide, Salicylamide and Ethenzamide. Molecules 2022, 27, 3323. [Google Scholar] [CrossRef]
  212. Qiu, J.; Huang, H.; He, H.; Liu, H.; Hu, S.; Han, J.; Guo, Y.; Wang, P. Measurement and Correlation of trans-4-Hydroxyl-proline Solubility in Sixteen Individual Solvents and a Water + Acetonitrile Binary Solvent System. J. Chem. Eng. Data 2020, 66, 575–587. [Google Scholar] [CrossRef]
  213. Radmand, S.; Rezaei, H.; Zhao, H.; Rahimpour, E.; Jouyban, A. Solubility and thermodynamic study of deferiprone in propylene glycol and ethanol mixture. BMC Chem 2023, 17, 37. [Google Scholar] [CrossRef]
  214. Ràfols, C.; Rosés, M.; Bosch, E. Solute–solvent and solvent–solvent interactions in binary solvent mixtures. Part 5. Preferential solvation of solvatochromic indicators in mixtures of propan-2-ol with hexane, benzene, ethanol and methanol. J. Chem. Soc. Perkin Trans. 1997, 2, 243–248. [Google Scholar] [CrossRef]
  215. Rani, R.S.; Rao, G.N. Stability of binary complexes of L-aspartic acid in dioxan–water mixtures. Bull. Chem. Soc. Ethiop. 2013, 27, 367–376. [Google Scholar] [CrossRef]
  216. Rashid, A.; White, E.T.; Howes, T.; Litster, J.D.; Marziano, I. Effect of Solvent Composition and Temperature on the Solubility of Ibuprofen in Aqueous Ethanol. J. Chem. Eng. Data 2014, 59, 2699–2703. [Google Scholar] [CrossRef]
  217. Rathi, P.B.; Kale, M.; Soleymani, J.; Jouyban, A. Solubility of Etoricoxib in Aqueous Solutions of Glycerin, Methanol, Polyethylene Glycols 200, 400, 600, and Propylene Glycol at 298.2 K. J. Chem. Eng. Data 2018, 63, 321–330. [Google Scholar] [CrossRef]
  218. Ren, J.; Chen, D.; Yu, Y.; Li, H. Solubility of dicarbohydrazide bis[3-(5-nitroimino-1,2,4-triazole)] in common pure solvents and binary solvents at different temperatures. R. Soc. Open Sci. 2019, 6, 190728. [Google Scholar] [CrossRef]
  219. Rezaei, H.; Rahimpour, E.; Martinez, F.; Jouyban, A. Measurement and correlation of solubility data for deferiprone in propylene glycol and 2-propanol at different temperatures. Heliyon 2023, 9, e17402. [Google Scholar] [CrossRef]
  220. Rezaei, H.; Rezaei, H.; Rahimpour, E.; Martinez, F.; Jouyban, A. Solubility profile of phenytoin in the mixture of 1-propanol and water at different temperatures. J. Mol. Liq. 2021, 334, 115936. [Google Scholar] [CrossRef]
  221. Rodríguez, A.; Trigo, M.; Aubourg, S.P.; Medina, I. Optimisation of Low-Toxicity Solvent Employment for Total Lipid and Tocopherol Compound Extraction from Patagonian Squid By-Products. Foods 2023, 12, 504. [Google Scholar] [CrossRef] [PubMed]
  222. Rosales-García, T.; Rosete-Barreto, J.M.; Pimentel-Rodas, A.; Davila-Ortiz, G.; Galicia-Luna, L.A. Solubility of Squalene and Fatty Acids in Carbon Dioxide at Supercritical Conditions: Binary and Ternary Systems. J. Chem. Eng. Data 2017, 63, 69–76. [Google Scholar] [CrossRef]
  223. Roses, M.; Ortega, J.; Bosch, E. Variation ofE T(30) polarity and the Kamlet-Taft solvatochromic parameters with composition in alcohol-alcohol mixtures. J. Solut. Chem. 1995, 24, 51–63. [Google Scholar] [CrossRef]
  224. Ruidiaz, M.A.; Delgado, D.R.; Martínez, F. Indomethacin solubility estimation in 1,4-dioxane + water mixtures by the extended hildebrand solubility approach. Química Nova 2011, 34, 1569–1574. [Google Scholar] [CrossRef]
  225. Sajedi-Amin, S.; Barzegar-Jalali, M.; Fathi-Azarbayjani, A.; Kebriaeezadeh, A.; Martínez, F.; Jouyban, A. Solubilization of bosentan using ethanol as a pharmaceutical cosolvent. J. Mol. Liq. 2017, 232, 152–158. [Google Scholar] [CrossRef]
  226. Serna-Carrizales, J.C.; Zárate-Guzmán, A.I.; Aguilar-Aguilar, A.; Forgionny, A.; Bailón-García, E.; Flórez, E.; Gómez-Durán, C.F.A.; Ocampo-Pérez, R. Optimization of Binary Adsorption of Metronidazole and Sulfamethoxazole in Aqueous Solution Supported with DFT Calculations. Processes 2023, 11, 1009. [Google Scholar] [CrossRef]
  227. Sha, J.; Ma, T.; Huang, Z.; Hu, X.; Zhang, R.; Cao, Z.; Wan, Y.; Sun, R.; He, H.; Jiang, G.; et al. Corrigendum to “Solubility determination, model evaluation, Hansen solubility parameter and thermodynamic properties of benorilate in six pure solvents and two binary solvent mixtures”. J. Chem. Thermodyn. 2021, 158, 106365. [Google Scholar] [CrossRef]
  228. Sha, J.; Yang, X.; Hu, X.; Huang, Z.; Cao, Z.; Wan, Y.; Sun, R.; Jiang, G.; He, H.; Li, Y.; et al. Solubility determination, model evaluation, Hansen solubility parameter and thermodynamic properties of benflumetol in pure alcohol and ester solvents. J. Chem. Thermodyn. 2021, 154, 106323. [Google Scholar] [CrossRef]
  229. Sha, J.; Yang, X.; Ji, L.; Cao, Z.; Niu, H.; Wan, Y.; Sun, R.; He, H.; Jiang, G.; Li, Y.; et al. Solubility determination, model evaluation, Hansen solubility parameter, molecular simulation and thermodynamic properties of benflumetol in four binary solvent mixtures from 278.15 K to 323.15 K. J. Mol. Liq. 2021, 333, 115867. [Google Scholar] [CrossRef]
  230. Shakeel, F.; Iqbal, M.; Ezzeldin, E.; Haq, N. Thermodynamics of solubility of ibrutinib in ethanol+water cosolvent mixtures at different temperatures. J. Mol. Liq. 2015, 209, 461–464. [Google Scholar] [CrossRef]
  231. Shao, D.; Yang, Z.; Zhou, G.; Chen, J.; Zheng, S.; Lv, X.; Li, R. Improving the solubility of acipimox by cosolvents and the study of thermodynamic properties on solvation process. J. Mol. Liq. 2018, 262, 389–395. [Google Scholar] [CrossRef]
  232. Sharapova, A.; Ol’khovich, M.; Blokhina, S.; Perlovich, G. Solubility and vapor pressure data of bioactive 6-(acetylamino)-N-(5-ethyl-1,3,4-thiadiazol-2-yl) hexanamide. J. Chem. Thermodyn. 2019, 135, 35–44. [Google Scholar] [CrossRef]
  233. Shen, B.; Wang, Q.; Wang, Y.; Ye, X.; Lei, F.; Gong, X. Solubilities of Adipic Acid in Acetic Acid + Water Mixtures and Acetic Acid + Cyclohexane Mixtures. J. Chem. Eng. Data 2013, 58, 938–942. [Google Scholar] [CrossRef]
  234. Sheng, X.; Luo, W.; Wang, Q. Determination and Correlation for the Solubilities of Succinic Acid in Cyclohexanol + Cyclohexanone + Cyclohexane Solvent Mixtures. J. Chem. Eng. Data 2018, 63, 801–811. [Google Scholar] [CrossRef]
  235. Shi, S.; Yan, M.; Tao, B.; Luo, W. Measurement and correlation for solubilities of succinic acid, glutaric acid and adipic acid in five organic solvents. J. Mol. Liq. 2020, 297, 111735. [Google Scholar] [CrossRef]
  236. Singh, S. Studies on the Interactions of Paracetamol in Water and Binary Solvent Mixtures at T = (298.15–313.15) K: Viscometric and Surface Tension Approach. Biointerface Res. Appl. Chem. 2021, 12, 2776–2786. [Google Scholar] [CrossRef]
  237. Smirnov, V.I. Thermochemical investigation of L-glutamine dissolution processes in aqueous co-solvent mixtures of acetonitrile, dioxane, acetone and dimethyl sulfoxide at T = 298.15 K. J. Chem. Thermodyn. 2020, 150, 106227. [Google Scholar] [CrossRef]
  238. Smirnov, V.I.; Badelin, V.G. Similarity and differences of the thermochemical characteristics of l-glutamine dissolution in aqueous solutions of some acetamides and formamides at T = 298.15 K. J. Mol. Liq. 2019, 285, 84–88. [Google Scholar] [CrossRef]
  239. Soltanpour, S.; Gharagozlu, A. Piroxicam Solubility in Binary and Ternary Solvents of Polyethylene Glycols 200 or 400 with Ethanol and Water at 298.2 K: Experimental Data Report and Modeling. J. Solut. Chem. 2015, 44, 1407–1423. [Google Scholar] [CrossRef]
  240. Soltanpour, S.; Jouyban, A. Solubility of Acetaminophen and Ibuprofen in Binary and Ternary Mixtures of Polyethylene Glycol 600, Ethanol and Water. Chem. Pharm. Bull. 2010, 58, 219–224. [Google Scholar] [CrossRef] [PubMed]
  241. Soltanpour, S.; Nazemi, V. Solubility of Ketoconazole in Binary and Ternary Solvents of Polyethylene Glycols 200, 400 or 600 with Ethanol and Water at 298.2 K. Data Report and Analysis. J. Solut. Chem. 2018, 47, 65–79. [Google Scholar] [CrossRef]
  242. Soltanpour, S.; Shekarriz, A.-H. Naproxen solubility in binary and ternary solvents of polyethylene glycols 200, 400 or 600 with ethanol and water at 298.2 K—Experimental data report and modelling. Phys. Chem. Liq. 2015, 53, 748–762. [Google Scholar] [CrossRef]
  243. Solymosi, T.; Tóth, F.; Orosz, J.; Basa-Dénes, O.; Angi, R.; Jordán, T.; Ötvös, Z.; Glavinas, H. Solubility Measurements at 296 and 310 K and Physicochemical Characterization of Abiraterone and Abiraterone Acetate. J. Chem. Eng. Data 2018, 12, 4453–4458. [Google Scholar] [CrossRef]
  244. Sun, H.; Liu, B.; Liu, P.; Zhang, J.; Wang, Y. Solubility of Fenofibrate in Different Binary Solvents: Experimental Data and Results of Thermodynamic Modeling. J. Chem. Eng. Data 2016, 61, 3177–3183. [Google Scholar] [CrossRef]
  245. Sun, J.; Liu, X.; Fang, Z.; Mao, S.; Zhang, L.; Rohani, S.; Lu, J. Solubility Measurement and Simulation of Rivaroxaban (Form I) in Solvent Mixtures from 273.15 to 323.15 K. J. Chem. Eng. Data 2015, 61, 495–503. [Google Scholar] [CrossRef]
  246. Swinerd, G.G. Orbital Mechanics: Theory and Applications; Logsdon, T., Ed.; John Wiley and Sons Limited: Chichester, UK, 1998. [Google Scholar] [CrossRef]
  247. Tang, W.; Wang, Z.; Feng, Y.; Xie, C.; Wang, J.; Yang, C.; Gong, J. Experimental Determination and Computational Prediction of Androstenedione Solubility in Alcohol + Water Mixtures. Ind. Eng. Chem. Res. 2014, 53, 11538–11549. [Google Scholar] [CrossRef]
  248. Tang, W.; Xie, C.; Wang, Z.; Wu, S.; Feng, Y.; Wang, X.; Wang, J.; Gong, J. Solubility of androstenedione in lower alcohols. Fluid Phase Equilibria 2014, 363, 86–96. [Google Scholar] [CrossRef]
  249. Teutenberg, T.; Wiese, S.; Wagner, P.; Gmehling, J. High-temperature liquid chromatography. Part II: Determination of the viscosities of binary solvent mixtures—Implications for liquid chromatographic separations. J. Chromatogr. A 2009, 1216, 8470–8479. [Google Scholar] [CrossRef]
  250. Thati, J.; Nordström, F.L.; Rasmuson, Å.C. Solubility of Benzoic Acid in Pure Solvents and Binary Mixtures. J. Chem. Eng. Data 2010, 55, 5124–5127. [Google Scholar] [CrossRef]
  251. Torres, N.; Escalera, B.; Martínez, F.; Peña, M.Á. Thermodynamic Analysis of Etoricoxib in Amphiprotic and Amphiprotic: Aprotic Solvent Mixtures at Several Temperatures. J. Solut. Chem. 2020, 49, 272–288. [Google Scholar] [CrossRef]
  252. Valavi, M.; Ukrainczyk, M.; Dehghani, M.R. Prediction of solubility of active pharmaceutical ingredients by semi- predictive Flory Huggins/Hansen model. J. Mol. Liq. 2017, 246, 166–172. [Google Scholar] [CrossRef]
  253. Vargas-Santana, M.S.; Cruz-González, A.M.; Ortiz, C.P.; Delgado, D.R.; Martínez, F.; Peña, M.Á.; Acree, W.E.; Jouyban, A. Solubility of sulfamerazine in (ethylene glycol + water) mixtures: Measurement, correlation, dissolution thermodynamics and preferential solvation. J. Mol. Liq. 2021, 337, 116330. [Google Scholar] [CrossRef]
  254. Vieira, A.W.; Molina, G.; Mageste, A.B.; Rodrigues, G.D.; de Lemos, L.R. Partitioning of salicylic and acetylsalicylic acids by aqueous two-phase systems: Mechanism aspects and optimization study. J. Mol. Liq. 2019, 296, 111775. [Google Scholar] [CrossRef]
  255. Volkova, T.V.; Levshin, I.B.; Perlovich, G.L. New antifungal compound: Solubility thermodynamics and partitioning processes in biologically relevant solvents. J. Mol. Liq. 2020, 310, 113148. [Google Scholar] [CrossRef]
  256. Wang, H.; Yao, G.; Zhang, H. Measurement and Correlation of the Solubility of Baicalin in Several Mixed Solvents. J. Chem. Eng. Data 2019, 64, 1281–1287. [Google Scholar] [CrossRef]
  257. Wang, S.; Chen, N.; Qu, Y. Solubility of Florfenicol in Different Solvents at Temperatures from (278 to 318) K. J. Chem. Eng. Data 2011, 56, 638–641. [Google Scholar] [CrossRef]
  258. Wang, S.; Chen, Y.; Gong, T.; Dong, W.; Wang, G.; Li, H.; Wu, S. Solid-liquid equilibrium behavior and thermodynamic analysis of dipyridamole in pure and binary solvents from 293.15 K to 328.15 K. J. Mol. Liq. 2019, 275, 8–17. [Google Scholar] [CrossRef]
  259. Wang, S.; Li, Q.-S.; Su, M.-G. Solubility of 1H-1,2,4-Triazole in Ethanol, 1-Propanol, 2-Propanol, 1,2-Propanediol, Ethyl Formate, Methyl Acetate, Ethyl Acetate, and Butyl Acetate at (283 to 363) K. J. Chem. Eng. Data 2007, 52, 856–858. [Google Scholar] [CrossRef]
  260. Wang, S.; Qin, L.; Zhou, Z.; Wang, J. Solubility and Solution Thermodynamics of Betaine in Different Pure Solvents and Binary Mixtures. J. Chem. Eng. Data 2012, 57, 2128–2135. [Google Scholar] [CrossRef]
  261. Wang, S.; Song, Z.; Wang, J.; Dong, Y.; Wu, M. Solubilities of Ibuprofen in Different Pure Solvents. J. Chem. Eng. Data 2010, 55, 5283–5285. [Google Scholar] [CrossRef]
  262. Wang, S.; Zhang, Y.; Wang, J. Solubility Measurement and Modeling for Betaine in Different Pure Solvents. J. Chem. Eng. Data 2014, 59, 2511–2516. [Google Scholar] [CrossRef]
  263. Wang, X.; Zhang, D.; Liu, S.; Chen, Y.; Jia, L.; Wu, S. Thermodynamic Study of Solubility for Imatinib Mesylate in Nine Monosolvents and Two Binary Solvent Mixtures from 278.15 to 318.15 K. J. Chem. Eng. Data 2018, 63, 4114–4127. [Google Scholar] [CrossRef]
  264. Wang, Z.; Xu, Z.; Xu, X.; Yang, A.; Luo, W.; Luo, Y. Solubility of benzoic acid in twelve organic solvents: Experimental measurement and thermodynamic modeling. J. Chem. Thermodyn. 2020, 150, 106234. [Google Scholar] [CrossRef]
  265. Watterson, S.; Hudson, S.; Svärd, M.; Rasmuson, Å.C. Thermodynamics of fenofibrate and solubility in pure organic solvents. Fluid Phase Equilibria 2014, 367, 143–150. [Google Scholar] [CrossRef]
  266. Wei, H.; Gao, N.; Dang, L. Solubility and Thermodynamic Properties of Sulfamethazine–Saccharin Cocrystal in Pure and Binary (Acetonitrile + 2-Propanol) Solvents. Trans. Tianjin Univ. 2020, 27, 460–472. [Google Scholar] [CrossRef]
  267. Wu, J.; Gu, L.; Wang, H.; Tao, L.; Wang, X. Solubility of Baicalein in Different Solvents from (287 to 323) K. Int. J. Thermophys. 2014, 35, 1465–1475. [Google Scholar] [CrossRef]
  268. Wu, K.; Li, Y. Solubility and solution thermodynamics of isobutyramide in 15 pure solvents at temperatures from 273.15 to 324.75 K. J. Mol. Liq. 2020, 311, 113294. [Google Scholar] [CrossRef]
  269. Wu, S.; Shi, Y.; Zhang, H. Solubility Measurement and Correlation for Amrinone in Four Binary Solvent Systems at 278.15–323.15 K. J. Chem. Eng. Data 2020, 65, 4108–4115. [Google Scholar] [CrossRef]
  270. Wu, Y.; Ren, M.; Zhang, X. Solubility Determination and Model Correlation of Benorilate between T = 278.18 and 318.15 K. J. Chem. Eng. Data 2020, 65, 3690–3695. [Google Scholar] [CrossRef]
  271. Wu, Y.; Wu, C.; Yan, S.; Hu, B. Solubility of Bisacodyl in Pure Solvent at Various Temperatures: Data Correlation and Thermodynamic Property Analysis. J. Chem. Eng. Data 2019, 65, 43–48. [Google Scholar] [CrossRef]
  272. Wu, Y.; Wu, J.; Wang, J.; Gao, J. Effect of Solvent Properties and Composition on the Solubility of Ganciclovir Form I. J. Chem. Eng. Data 2019, 64, 1501–1507. [Google Scholar] [CrossRef]
  273. Wüst Zibetti, A.; Aydi, A.; Claumann, C.A.; Eladeb, A.; Adberraba, M. Correlation of solubility and prediction of the mixing properties of rosmarinic acid in different pure solvents and in binary solvent mixtures of ethanol + water and methanol + water from (293.2 to 318.2) K. J. Mol. Liq. 2016, 216, 370–376. [Google Scholar] [CrossRef]
  274. Xia, Q.; Chen, S.-N.; Chen, Y.-S.; Zhang, M.-S.; Zhang, F.-B.; Zhang, G.-L. Solubility of decanedioic acid in binary solvent mixtures. Fluid Phase Equilibria 2011, 304, 105–109. [Google Scholar] [CrossRef]
  275. Xu, R.; Han, T.; Shen, L.; Zhao, J.; Lu, X.a. Solubility Determination and Modeling for Artesunate in Binary Solvent Mixtures of Methanol, Ethanol, Isopropanol, and Propylene Glycol + Water. J. Chem. Eng. Data 2019, 64, 755–762. [Google Scholar] [CrossRef]
  276. Marcus, Y. Preferential Solvation of Drugs in Binary Solvent Mixtures. Pharm. Anal. Acta 2017, 10, 4172. [Google Scholar] [CrossRef]
  277. Yan, M.; Li, X.; Tao, B.; Yang, L.; Luo, W. Solubility of succinic acid, glutaric acid and adipic acid in propionic acid + ε-caprolactone mixtures and propionic acid + cyclohexanone mixtures: Experimental measurement and thermodynamic modeling. J. Mol. Liq. 2018, 272, 106–119. [Google Scholar] [CrossRef]
  278. Yang, H.; Rasmuson, Å.C. Solubility of Butyl Paraben in Methanol, Ethanol, Propanol, Ethyl Acetate, Acetone, and Acetonitrile. J. Chem. Eng. Data 2010, 55, 5091–5093. [Google Scholar] [CrossRef]
  279. Yang, H.; Zhang, T.; Xu, S.; Han, D.; Liu, S.; Yang, Y.; Du, S.; Li, M.; Gong, J. Measurement and Correlation of the Solubility of Azoxystrobin in Seven Monosolvents and Two Different Binary Mixed Solvents. J. Chem. Eng. Data 2017, 62, 3967–3980. [Google Scholar] [CrossRef]
  280. Yang, L.; Zhang, Y.; Cheng, J.; Yang, C. Solubility and thermodynamics of polymorphic indomethacin in binary solvent mixtures. J. Mol. Liq. 2019, 295, 111717. [Google Scholar] [CrossRef]
  281. Yang, Z.; Shao, D.; Zhou, G. Analysis of solubility parameters of fenbendazole in pure and mixed solvents and evaluation of thermodynamic model. J. Chem. Thermodyn. 2020, 140, 105876. [Google Scholar] [CrossRef]
  282. Yang, Z.; Shao, D.; Zhou, G. Solubility profile of imatinib in pure and mixed solvents and calculation of thermodynamic properties. J. Chem. Thermodyn. 2020, 144, 106031. [Google Scholar] [CrossRef]
  283. Yang, Z.; Shao, D.; Zhou, G. Improvement of solubility and analysis thermodynamic properties of β tegafur in pure and mixed organic solvents. J. Chem. Thermodyn. 2020, 146, 106090. [Google Scholar] [CrossRef]
  284. Yaws, C.L. Physical Properties—Inorganic Compounds. In The Yaws Handbook of Physical Properties for Hydrocarbons and Chemicals; Elsevier: Amsterdam, The Netherlands, 2015; pp. 684–810. [Google Scholar]
  285. Yu, Q.; Ma, X.; Xu, L. Solubility, dissolution enthalpy and entropy of l-glutamine in mixed solvents of ethanol+water and acetone+water. Thermochim. Acta 2013, 558, 6–9. [Google Scholar] [CrossRef]
  286. Zadaliasghar, S.; Jouyban, A.; Martinez, F.; Barzegar-Jalali, M.; Rahimpour, E. Solubility of ketoconazole in the binary mixtures of 2-propanol and water at different temperatures. J. Mol. Liq. 2020, 300, 112259. [Google Scholar] [CrossRef]
  287. Zhang, C.-L.; Li, B.-Y.; Wang, Y. Solubilities of Sulfadiazine in Methanol, Ethanol, 1-Propanol, 2-Propanol, Acetone, and Chloroform from (294.15 to 318.15) K. J. Chem. Eng. Data 2010, 55, 2338–2339. [Google Scholar] [CrossRef]
  288. Zhang, F. Commentary on the “Solubility of l-histidine in different aqueous binary solvent mixtures from 283.15 K to 318.15 K with experimental measurement and thermodynamic modelling”. J. Chem. Thermodyn. 2018, 124, 98–100. [Google Scholar] [CrossRef]
  289. Zhang, H.; Yin, Q.; Liu, Z.; Gong, J.; Bao, Y.; Zhang, M.; Hao, H.; Hou, B.; Xie, C. Measurement and correlation of solubility of dodecanedioic acid in different pure solvents from T = (288.15 to 323.15)K. J. Chem. Thermodyn. 2014, 68, 270–274. [Google Scholar] [CrossRef]
  290. Zhang, J.; Huang, C.; Chen, J.; Xu, R. Equilibrium Solubility Determination and Modeling of Fenbendazole in Cosolvent Mixtures at (283.15–328.15) K. J. Chem. Eng. Data 2019, 64, 4095–4102. [Google Scholar] [CrossRef]
  291. Zhang, J.; Huang, C.; Xu, R. Solubility of Bifonazole in Four Binary Solvent Mixtures: Experimental Measurement and Thermodynamic Modeling. J. Chem. Eng. Data 2019, 64, 2641–2648. [Google Scholar] [CrossRef]
  292. Zhang, J.; Huang, C.; Xu, R. Solubility Determination and Mathematical Modeling of Nicorandil in Several Aqueous Cosolvent Systems at Temperature Ranges of 278.15–323.15 K. J. Chem. Eng. Data 2020, 65, 4063–4070. [Google Scholar] [CrossRef]
  293. Zhang, J.; Song, X.; Xu, R. Solubility Determination and Modeling for Milrinone in Binary Solvent Mixtures of Ethanol, Isopropanol, Ethylene Glycol, and N,N-Dimethylformamide + Water. J. Chem. Eng. Data 2020, 65, 4100–4107. [Google Scholar] [CrossRef]
  294. Zhang, J.; Zhang, H.; Xu, R. Solubility Determination and Modeling for Tirofiban in Several Mixed Solvents at 278.15–323.15 K. J. Chem. Eng. Data 2020, 65, 4071–4078. [Google Scholar] [CrossRef]
  295. Zhang, N.; Li, S.; Yang, H.; Li, M.; Yang, Y.; Tang, W. Measurement and Correlation of the Solubility of Tetramethylpyrazine in Nine Monosolvents and Two Binary Solvent Systems. J. Chem. Eng. Data 2019, 64, 995–1006. [Google Scholar] [CrossRef]
  296. Zhang, P.; Sha, J.; Wan, Y.; Zhang, C.; Li, T.; Ren, B. Apparent thermodynamic analysis and the dissolution behavior of levamisole hydrochloride in three binary solvent mixtures. Thermochim. Acta 2019, 681, 178375. [Google Scholar] [CrossRef]
  297. Zhang, P.; Wan, Y.; Zhang, C.; Zhao, R.; Sha, J.; Li, Y.; Li, T.; Ren, B. Solubility and mixing thermodynamic properties of levamisole hydrochloride in twelve pure solvents at various temperatures. J. Chem. Thermodyn. 2019, 139, 105882. [Google Scholar] [CrossRef]
  298. Zhang, P.; Zhang, C.; Zhao, R.; Wan, Y.; Yang, Z.; He, R.; Chen, Q.; Li, T.; Ren, B. Measurement and Correlation of the Solubility of Florfenicol Form A in Several Pure and Binary Solvents. J. Chem. Eng. Data 2018, 63, 2046–2055. [Google Scholar] [CrossRef]
  299. Zhang, X.; Chen, J.; Hu, J.; Liu, M.; Cai, Z.; Xu, Y.; Sun, B. The solubilities of benzoic acid and its nitro-derivatives, 3-nitro and 3,5-dinitrobenzoic acids. J. Chem. Res. 2021, 45, 1100–1106. [Google Scholar] [CrossRef]
  300. Zhang, X.; Cui, P.; Yin, Q.; Zhou, L. Measurement and Correlation of the Solubility of Florfenicol in Four Binary Solvent Mixtures from T = (278.15 to 318.15) K. Crystals 2022, 12, 1176. [Google Scholar] [CrossRef]
  301. Zhu, Y.; Yang, H.; Si, Z.; Zhang, X. Solubility and thermodynamics of l-hydroxyproline in water and (methanol, ethanol, n-propanol) binary solvent mixtures. J. Mol. Liq. 2020, 298, 112043. [Google Scholar] [CrossRef]
  302. Zorrilla-Veloz, R.I.; Stelzer, T.; López-Mejías, V. Measurement and Correlation of the Solubility of 5-Fluorouracil in Pure and Binary Solvents. J. Chem. Eng. Data 2018, 63, 3809–3817. [Google Scholar] [CrossRef]
  303. Milescu, R.A.; Zhenova, A.; Vastano, M.; Gammons, R.; Lin, S.; Lau, C.H.; Clark, J.H.; McElroy, C.R.; Pellis, A. Polymer Chemistry Applications of Cyrene and its Derivative Cygnet 0.0 as Safer Replacements for Polar Aprotic Solvents. ChemSusChem 2021, 14, 3367–3381. [Google Scholar] [CrossRef]
  304. Glass, M.; Aigner, M.; Viell, J.; Jupke, A.; Mitsos, A. Liquid-liquid equilibrium of 2-methyltetrahydrofuran/water over wide temperature range: Measurements and rigorous regression. Fluid Phase Equilibria 2017, 433, 212–225. [Google Scholar] [CrossRef]
  305. Dargo, G.; Kis, D.; Gede, M.; Kumar, S.; Kupai, J.; Szekely, G. MeSesamol, a bio-based and versatile polar aprotic solvent for organic synthesis and depolymerization. Chem. Eng. J. 2023, 471, 144365. [Google Scholar] [CrossRef]
  306. Komarova, A.O.; Dick, G.R.; Luterbacher, J.S. Diformylxylose as a new polar aprotic solvent produced from renewable biomass. Green Chem. 2021, 23, 4790–4799. [Google Scholar] [CrossRef]
  307. Kamlet, M.J.; Abboud, J.L.M.; Abraham, M.H.; Taft, R.W. Linear solvation energy relationships. 23. A comprehensive collection of the solvatochromic parameters,. pi.*,. alpha., and. beta., and some methods for simplifying the generalized solvatochromic equation. J. Org. Chem. 1983, 48, 2877–2887. [Google Scholar] [CrossRef]
  308. Abraham, M.H. Scales of solute hydrogen-bonding: Their construction and application to physicochemical and biochemical processes. Chem. Soc. Rev. 1993, 22, 73–83. [Google Scholar] [CrossRef]
  309. Liu, X.; Acree, W.E.; Abraham, M.H. Descriptors for some compounds with pharmacological activity; calculation of properties. Int. J. Pharm. 2022, 617, 121597. [Google Scholar] [CrossRef]
  310. Vitha, M.; Carr, P.W. The chemical interpretation and practice of linear solvation energy relationships in chromatography. J. Chromatogr. A 2006, 1126, 143–194. [Google Scholar] [CrossRef]
  311. West, C.; Lesellier, E. Characterisation of stationary phases in subcritical fluid chromatography with the solvation parameter model: III. Polar stationary phases. J. Chromatogr. A 2006, 1110, 200–213. [Google Scholar] [CrossRef]
  312. Efimov, I.; Povarov, V.G.; Rudko, V.A. Comparison of UNIFAC and LSER Models for Calculating Partition Coefficients in the Hexane–Acetonitrile System Using Middle Distillate Petroleum Products as an Example. Ind. Eng. Chem. Res. 2022, 61, 9575–9585. [Google Scholar] [CrossRef]
  313. EPA. CompTox Chemicals Dashboard v2.3.0. 2023. Available online: https://comptox.epa.gov/dashboard/ (accessed on 31 December 2023).
  314. Hou, P.; Jolliet, O.; Zhu, J.; Xu, M. Estimate ecotoxicity characterization factors for chemicals in life cycle assessment using machine learning models. Environ. Int. 2020, 135, 105393. [Google Scholar] [CrossRef]
  315. Wojeicchowski, J.P.; Abranches, D.O.; Ferreira, A.M.; Mafra, M.R.; Coutinho, J.A.P. Using COSMO-RS to Predict Solvatochromic Parameters for Deep Eutectic Solvents. ACS Sustain. Chem. Eng. 2021, 9, 10240–10249. [Google Scholar] [CrossRef]
  316. Wojeicchowski, J.P.; Ferreira, A.M.; Okura, T.; Pinheiro Rolemberg, M.; Mafra, M.R.; Coutinho, J.A.P. Using COSMO-RS to Predict Hansen Solubility Parameters. Ind. Eng. Chem. Res. 2022, 61, 15631–15638. [Google Scholar] [CrossRef]
Figure 1. ACS Green Chemistry Institute and Pharmaceutical Roundtable (GCIPR) solvent selection tool https://www.acsgcipr.org/tools-for-innovation-in-chemistry/solvent-tool/ (accessed on 1 April 2024) described by Diorazio et al. [20]. Copyright ACS, 2023.
Figure 1. ACS Green Chemistry Institute and Pharmaceutical Roundtable (GCIPR) solvent selection tool https://www.acsgcipr.org/tools-for-innovation-in-chemistry/solvent-tool/ (accessed on 1 April 2024) described by Diorazio et al. [20]. Copyright ACS, 2023.
Liquids 04 00018 g001
Figure 2. Reichardt ETN parameters plotted against Kamlet—Taft dipolarity/polarizability parameters for selected molecular solvents.
Figure 2. Reichardt ETN parameters plotted against Kamlet—Taft dipolarity/polarizability parameters for selected molecular solvents.
Liquids 04 00018 g002
Figure 3. Kamlet—Taft basicity parameter plotted against acidity parameter for selected molecular solvents.
Figure 3. Kamlet—Taft basicity parameter plotted against acidity parameter for selected molecular solvents.
Liquids 04 00018 g003
Figure 4. Kamlet—Taft acidity (α) and basicity (β) versus dipolar/polarizability (π*) for aqueous and non-aqueous mixed solvents and pure solvents. Dashed lines show approximate behavior of mixed solvent KT parameters with composition.
Figure 4. Kamlet—Taft acidity (α) and basicity (β) versus dipolar/polarizability (π*) for aqueous and non-aqueous mixed solvents and pure solvents. Dashed lines show approximate behavior of mixed solvent KT parameters with composition.
Liquids 04 00018 g004
Figure 5. Dynamic viscosity (η) of water (HBD) –hydrogen bond acceptor (HBA) mixed solvent systems as a function of mole fraction of HBA solvent (x2) at 25 °C. HBA solvents are ordered in terms of Hunter basicity ( β 2 H ) values (low to high): acetonitrile (Liquids 04 00018 i001 ACN), γ-valerolactone (Liquids 04 00018 i002 GVL), γ-butyrolactone (Liquids 04 00018 i003 GBL), tetrahydrofuran (Liquids 04 00018 i004 THF), 1,4-dioxane (Liquids 04 00018 i005 DI), acetone (Liquids 04 00018 i006 Ace), pyridine (Liquids 04 00018 i007 PYR), N-methyl-2-pyrrolidone (Liquids 04 00018 i008 NMP), N,N-dimethylformamide (Liquids 04 00018 i009 DMF), N,N-dimethylacetamide (Liquids 04 00018 i010 DMA), and dimethyl sulfoxide (Liquids 04 00018 i011 DMSO). Reprinted with permission from [42]. Copyright American Chemical Society, 2017.
Figure 5. Dynamic viscosity (η) of water (HBD) –hydrogen bond acceptor (HBA) mixed solvent systems as a function of mole fraction of HBA solvent (x2) at 25 °C. HBA solvents are ordered in terms of Hunter basicity ( β 2 H ) values (low to high): acetonitrile (Liquids 04 00018 i001 ACN), γ-valerolactone (Liquids 04 00018 i002 GVL), γ-butyrolactone (Liquids 04 00018 i003 GBL), tetrahydrofuran (Liquids 04 00018 i004 THF), 1,4-dioxane (Liquids 04 00018 i005 DI), acetone (Liquids 04 00018 i006 Ace), pyridine (Liquids 04 00018 i007 PYR), N-methyl-2-pyrrolidone (Liquids 04 00018 i008 NMP), N,N-dimethylformamide (Liquids 04 00018 i009 DMF), N,N-dimethylacetamide (Liquids 04 00018 i010 DMA), and dimethyl sulfoxide (Liquids 04 00018 i011 DMSO). Reprinted with permission from [42]. Copyright American Chemical Society, 2017.
Liquids 04 00018 g005
Figure 6. Concept of solubility parameter and Kamlet—Taft windows for identifying replacement solvents of an API (paracetamol): (a) window for solubility parameter, (b) window for API acidity, (c) window for API basicity, (d) window for API dipolarity/polarizability. (left): Range of solubility and Kamlet—Taft parameters for dissolution of API in known solvents, including hazardous ones. (right): Range of solubility and Kamlet—Taft parameters superimposed onto theoretical calculations and available literature data to determine working composition ranges for a given mixed solvent pair (acetone—water). Reprinted with permission from ref. [86]. Copyright American Chemical Society, 2016.
Figure 6. Concept of solubility parameter and Kamlet—Taft windows for identifying replacement solvents of an API (paracetamol): (a) window for solubility parameter, (b) window for API acidity, (c) window for API basicity, (d) window for API dipolarity/polarizability. (left): Range of solubility and Kamlet—Taft parameters for dissolution of API in known solvents, including hazardous ones. (right): Range of solubility and Kamlet—Taft parameters superimposed onto theoretical calculations and available literature data to determine working composition ranges for a given mixed solvent pair (acetone—water). Reprinted with permission from ref. [86]. Copyright American Chemical Society, 2016.
Liquids 04 00018 g006
Figure 9. Kamlet—Taft acidity, basicity, and polarity for selected mixed solvents versus HBA solvent mole fraction: (a,d,g) water—HBA; (b,e,h) methanol—HBA; (c,f,i) ethanol—HBA. Trends shown are based on estimations (dashed lines) and actual data (solid lines) [51,86,90,91].
Figure 9. Kamlet—Taft acidity, basicity, and polarity for selected mixed solvents versus HBA solvent mole fraction: (a,d,g) water—HBA; (b,e,h) methanol—HBA; (c,f,i) ethanol—HBA. Trends shown are based on estimations (dashed lines) and actual data (solid lines) [51,86,90,91].
Liquids 04 00018 g009
Table 1. Selected chemicals from candidate list of substances of very high concern (SVHC) for authorization by the European Chemicals Agency (ECHA) as of 2023. Chemicals shown in various categories are for educational purposes only. Specific hazards, detailed information, case decisions, or discussion should be accessed from ECHA website [17]. LD50 values from PubChem or online sources based on rat/mouse oral or dermal (d) studies.
Table 1. Selected chemicals from candidate list of substances of very high concern (SVHC) for authorization by the European Chemicals Agency (ECHA) as of 2023. Chemicals shown in various categories are for educational purposes only. Specific hazards, detailed information, case decisions, or discussion should be accessed from ECHA website [17]. LD50 values from PubChem or online sources based on rat/mouse oral or dermal (d) studies.
Chemical (CAS No.)LD50 (mg/kg)Chemical (CAS No.)LD50 (mg/kg)
Carcinogenic Respiratory Sensitizing
1,2,3-trichloropropane (96-18-4)120cis-cyclohexane-1,2-dicarboxylic anhydride (13149-00-3)-
1,2-dichloroethane (107-06-2)670Cyclohexane-1,2-dicarboxylic anhydride (85-42-7)958
1,4-dioxane (123-91-1) (DI)1550Glutaral (111-30-8)134
2,4-dinitrotoluene (121-14-2)268Toxic to Reproduction
4,4′-Diaminodiphenylmethane (101-77-9)1201-Methyl-2-pyrrolidone (NMP) (872-50-4)3914
4-aminoazobenzene (60-09-3)2001-vinylimidazole (1072-63-5)180
Acrylamide (79-06-1)1702-ethoxyethanol (110-80-5)2125
Anthracene oil (90640-80-5)2000 d2-ethoxyethyl acetate (111-15-9)2700
Biphenyl-4-ylamine (92-67-1)2052-methoxyethanol (109-86-4)2370
Chrysene (218-01-9)3202-methoxyethyl acetate (110-49-6)2900
Furan (110-00-9)5.22-methylimidazole (693-98-1)1400
Propylene oxide (75-56-9)1245 d4,4′-sulphonyldiphenol (80-09-1)4556
N-(hydroxymethyl)acrylamide (924-42-5)474Dibutyl phthalate (84-74-2) (DBP) 7499
o-aminoazotoluene (97-56-3)300 (dog)Dicyclohexyl phthalate (84-61-7)30
o-toluidine (95-53-4)670Dihexyl phthalate (84-75-3)29,600
Phenolphthalein (77-09-8)>1Diisobutyl phthalate (84-69-5)15
Potassium dichromate (7778-50-9)25Diisohexyl phthalate (71850-09-4)-
Trichloroethylene (79-01-6)1282Diisopentyl phthalate (605-50-5)2000
Endocrine disruptor Dioctyltin dilaurate (3648-18-8)6450
2-(isononylphenoxy)ethanol (85005-55-6)-Formamide (75-12-7)5577
4-(1-ethyl-1-methylhexyl)phenol (52427-13-1)-Methoxyacetic acid (625-45-6)1000
4,4′-(1-methylpropylidene)bisphenol (77-40-7)500.1N,N-dimethylformamide (68-12-2) (DMF) 2800
4-tert-butylphenol (98-54-4)2951Nitrobenzene (98-95-3)349
Isobutyl 4-hydroxybenzoate (4247-02-3)2600N-methylacetamide (79-16-3)5
Nonylphenol (25154-52-3)1200 n-pentyl-isopentyl phthalate (776297-69-9)-
Nonylphenol, ethoxylated (9016-45-9) 1300 Perfluoroheptanoic acid (375-85-9)500
Human health effects Phenol, 4-dodecyl, branched (210555-94-5)2000
Melamine (108-78-1)3161Phenol, tetrapropylene- (57427-55-1)2000
Persistent, Bioaccumulative and Toxic (PBT)Very Persistent, Very Bioaccumulative (vPvB)
Alkanes, C14-16, chloro (1372804-76-6)23 Phenanthrene (85-01-8)700
Anthracene (120-12-7)>17Terphenyl, hydrogenated (61788-32-7)17,500
Dodecamethylcyclohexasiloxane (540-97-6)>50
Octamethylcyclotetrasiloxane (556-67-2)1540
Pyrene (129-00-0)2700
Table 2. Possible replacement solvents for dipolar aprotic solvents used in synthetic chemistry transformations. Content was summarized and adapted from Unified solvent selection guide for replacement of common dipolar aprotic solvents in synthetically useful transformations contained in ref. [16]. Copyright ACS, 2022.
Table 2. Possible replacement solvents for dipolar aprotic solvents used in synthetic chemistry transformations. Content was summarized and adapted from Unified solvent selection guide for replacement of common dipolar aprotic solvents in synthetically useful transformations contained in ref. [16]. Copyright ACS, 2022.
ReactionUnsafe Dipolar AproticsReplacement Solvents
Amide formationDCM; DMFCyrene; surfactant-water
Boc deprotectionDIHCl in CPME; TFA in PC
Borylation chemistryDI2-MeTHF:MeOH (1:1); CPME; MTBE; CH
Buchwald—Hartwig aminationDI2-MeTHF; tBuOH
CarbonylationTHF; DEDMC
CarboxylationTHF; DE2-MeTHF; DMI; DMC
C-H activationTHF; DMF; DI2-MeTHF; CH
Mizoroki—Heck cross-couplingDI; THF; DMFNBP; DMI; PC
Nucleophilic aromatic substitutionTHF; DMF; DI2-MeTHF; PEG-400
Organometallic reactionR-MgX; R-Li; hydrides2-MeTHF; CPME
Solid-phase peptide synthesisDMF; DMAc; NMPNBP; GVL
Sonogashira cross-couplingTHF; DMFCyrene; NBP; DMI; Eucalyptol
Steglich EsterificationDMFDMC
Suzuki-Miyaura cross-couplingDI; THF; DMFCyrene; NBP; DMI; 2-MeTHF
Urea synthesisDMF; THFCyrene
Table 4. Co-extractant methodology for obtaining bio-products from supercritical CO2 extraction of natural sources. Co-extractants: vegetable, drupe, legume, or seed oils or triacylglycerols (TAGs, triglycerides). Bio-product yields shown are maximum values normalized to 100%.
Table 4. Co-extractant methodology for obtaining bio-products from supercritical CO2 extraction of natural sources. Co-extractants: vegetable, drupe, legume, or seed oils or triacylglycerols (TAGs, triglycerides). Bio-product yields shown are maximum values normalized to 100%.
Natural SourceCo-ExtractantBio-ProductT (°C)P (MPa)%YieldRef.
AlgaeSoybean oilastaxanthin704036[57]
Brown seaweedSunflower oilcarotenoids503099[58]
CarrotsCanola oilcarotenoids705592[59]
MangosteenVirgin coconut oilxanthonoids704331[60]
MangosteenVirgin coconut oilα-mangostin603576[61]
MarigoldMedium-chain TAGslutein esters654398[62]
MarigoldSoybean oillutein esters533093[63]
PropolisVirgin coconut oilflavonoids501525[64]
PumpkinOlive oil carotenoid502541[65]
Red sagePeanut oil diterpenoids503890[66]
TomatoCanola oillycopene 404086[67]
TomatoHazelnut oillycopene664540[68]
Tomato skinOlive oil lycopene753558[69]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lee, J.L.; Chong, G.H.; Ota, M.; Guo, H.; Smith, R.L., Jr. Solvent Replacement Strategies for Processing Pharmaceuticals and Bio-Related Compounds—A Review. Liquids 2024, 4, 352-381. https://doi.org/10.3390/liquids4020018

AMA Style

Lee JL, Chong GH, Ota M, Guo H, Smith RL Jr. Solvent Replacement Strategies for Processing Pharmaceuticals and Bio-Related Compounds—A Review. Liquids. 2024; 4(2):352-381. https://doi.org/10.3390/liquids4020018

Chicago/Turabian Style

Lee, Jia Lin, Gun Hean Chong, Masaki Ota, Haixin Guo, and Richard Lee Smith, Jr. 2024. "Solvent Replacement Strategies for Processing Pharmaceuticals and Bio-Related Compounds—A Review" Liquids 4, no. 2: 352-381. https://doi.org/10.3390/liquids4020018

Article Metrics

Back to TopTop